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 31 毫秒
1.
N-Glycosylation of integrin α5β1 plays a crucial role in cell spreading, cell migration, ligand binding, and dimer formation, but the detailed mechanisms by which N-glycosylation mediates these functions remain unclear. In a previous study, we showed that three potential N-glycosylation sites (α5S3–5) on the β-propeller of the α5 subunit are essential to the functional expression of the subunit. In particular, site 5 (α5S5) is the most important for its expression on the cell surface. In this study, the function of the N-glycans on the integrin β1 subunit was investigated using sequential site-directed mutagenesis to remove the combined putative N-glycosylation sites. Removal of the N-glycosylation sites on the I-like domain of the β1 subunit (i.e. the Δ4-6 mutant) decreased both the level of expression and heterodimeric formation, resulting in inhibition of cell spreading. Interestingly, cell spreading was observed only when the β1 subunit possessed these three N-glycosylation sites (i.e. the S4-6 mutant). Furthermore, the S4-6 mutant could form heterodimers with either α5S3-5 or α5S5 mutant of the α5 subunit. Taken together, the results of the present study reveal for the first time that N-glycosylation of the I-like domain of the β1 subunit is essential to both the heterodimer formation and biological function of the subunit. Moreover, because the α5S3-5/β1S4-6 mutant represents the minimal N-glycosylation required for functional expression of the β1 subunit, it might also be useful for the study of molecular structures.Integrin is a heterodimeric glycoprotein that consists of both an α and a β subunit (1). The interaction between integrin and the extracellular matrix is essential to both physiologic and pathologic events, such as cell migration, development, cell viability, immune homeostasis, and tumorigenesis (2, 3). Among the integrin superfamily, β1 integrin can combine with 12 distinct α subunits (α1–11, αv) to form heterodimers, thereby acquiring a wide variety of ligand specificity (1, 4). Integrins are thought to be regulated by inside-out signaling mechanisms that provoke conformational changes, which modulate the affinity of integrin for the ligand (5). However, an increasing body of evidence suggests that cell-surface carbohydrates mediate a variety of interactions between integrin and its extracellular environment, thereby affecting integrin activity and possibly tumor metastasis as well (68).Guo et al. (9) reported that an increase in β1–6-GlcNAc sugar chains on the integrin β1 subunit stimulated cell migration. In addition, elevated sialylation of the β1 subunit, because of Ras-induced STGal-I transferase activity, also induced cell migration (10, 11). Conversely, cell migration and spreading were reduced by the addition of a bisecting GlcNAc, which is a product of N-acetylglucosaminyltransferase III (GnT-III),2 to the α5β1 and α3β1 integrins (12, 13). Alterations of N-glycans on integrins might also regulate their cis interactions with membrane-associated proteins, including the epidermal growth factor receptor, the galectin family, and the tetraspanin family of proteins (1419).In addition to the positive and negative regulatory effects of N-glycan, several research groups have reported that N-glycans must be present on integrin α5β1 for the αβ heterodimer formation and proper integrin-matrix interactions. Consistent with this hypothesis, in the presence of the glycosylation inhibitor, tunicamycin, normal integrin-substrate binding and transport to the cell surface are inhibited (20). Moreover, treatment of purified integrin with N-glycosidase F blocked both the inherent association of the subunits and the interaction between integrin and fibronectin (FN) (21). These results suggest that N-glycosylation is essential to the functional expression of α5β1. However, because integrin α5β1 contains 26 potential N-linked glycosylation sites, 14 in the α subunit and 12 in the β subunit, identification of the sites that are essential to its biological functions is key to understanding the molecular mechanisms by which N-glycans alter integrin function. Recently, our group determined that N-glycosylation of the β-propeller domain on the α5 subunit is essential to both heterodimerization and biological functions of the subunit. Furthermore, we determined that sites 3–5 are the most important sites for α5 subunit-mediated cell spreading and migration on FN (22). The purpose of this study was to clarify the roles of N-glycosylation of the β1 subunit. Therefore, we performed combined substitutions in the putative N-glycosylation sites by replacement of asparagine residues with glutamine residues. We subsequently introduced these mutated genes into β1-deficient epithelial cells (GE11). The results of these mutation experiments revealed that the N-glycosylation sites on the I-like domain of the β1 subunit, sites number 4–6 (S4-6), are essential to both heterodimer formation and biological functions, such as cell spreading.  相似文献   

2.
3.
Integrin receptor plays key roles in mediating both inside-out and outside-in signaling between cells and the extracellular matrix. We have observed that the tissue-specific loss of the integrin β1 subunit in striated muscle results in a near complete loss of integrin β1 subunit protein expression concomitant with a loss of talin and to a lesser extent, a reduction in F-actin content. Muscle-specific integrin β1-deficient mice had no significant difference in food intake, weight gain, fasting glucose, and insulin levels with their littermate controls. However, dynamic analysis of glucose homeostasis using euglycemichyperinsulinemic clamps demonstrated a 44 and 48% reduction of insulin-stimulated glucose infusion rate and glucose clearance, respectively. The whole body insulin resistance resulted from a specific inhibition of skeletal muscle glucose uptake and glycogen synthesis without any significant effect on the insulin suppression of hepatic glucose output or insulin-stimulated glucose uptake in adipose tissue. The reduction in skeletal muscle insulin responsiveness occurred without any change in GLUT4 protein expression levels but was associated with an impairment of the insulin-stimulated protein kinase B/Akt serine 473 phosphorylation but not threonine 308. The inhibition of insulin-stimulated serine 473 phosphorylation occurred concomitantly with a decrease in integrin-linked kinase expression but with no change in the mTOR·Rictor·LST8 complex (mTORC2). These data demonstrate an in vivo crucial role of integrin β1 signaling events in mediating cross-talk to that of insulin action.Integrin receptors are a large family of integral membrane proteins composed of a single α and β subunit assembled into a heterodimeric complex. There are 19 α and 8 β mammalian subunit isoforms that combine to form 25 distinct α,β heterodimeric receptors (1-5). These receptors play multiple critical roles in conveying extracellular signals to intracellular responses (outside-in signaling) as well as altering extracellular matrix interactions based upon intracellular changes (inside-out signaling). Despite the large overall number of integrin receptor complexes, skeletal muscle integrin receptors are limited to seven α subunit subtypes (α1, α3, α4, α5, α6, α7, and αν subunits), all associated with the β1 integrin subunit (6, 7).Several studies have suggested an important cross-talk between extracellular matrix and insulin signaling. For example, engagement of β1 subunit containing integrin receptors was observed to increase insulin-stimulated insulin receptor substrate (IRS)2 phosphorylation, IRS-associated phosphatidylinositol 3-kinase, and activation of protein kinase B/Akt (8-11). Integrin receptor regulation of focal adhesion kinase was reported to modulate insulin stimulation of glycogen synthesis, glucose transport, and cytoskeleton organization in cultured hepatocytes and myoblasts (12, 13). Similarly, the integrin-linked kinase (ILK) was suggested to function as one of several potential upstream kinases that phosphorylate and activate Akt (14-18). In this regard small interfering RNA gene silencing of ILK in fibroblasts and conditional ILK gene knockouts in macrophages resulted in a near complete inhibition of insulin-stimulated Akt serine 473 (Ser-473) phosphorylation concomitant with an inhibition of Akt activity and phosphorylation of Akt downstream targets (19). However, a complex composed of mTOR·Rictor·LST8 (termed mTORC2) has been identified in several other studies as the Akt Ser-473 kinase (20, 21). In addition to Ser-473, Akt protein kinase activation also requires phosphorylation on threonine 308 Thr-30 by phosphoinositide-dependent protein kinase, PDK1 (22-24).In vivo, skeletal muscle is the primary tissue responsible for postprandial (insulin-stimulated) glucose disposal that results from the activation of signaling pathways leading to the translocation of the insulin-responsive glucose transporter, GLUT4, from intracellular sites to the cell surface membranes (25, 26). Dysregulation of any step of this process in skeletal muscle results in a state of insulin resistance, thereby predisposing an individual for the development of diabetes (27-33). Although studies described above have utilized a variety of tissue culture cell systems to address the potential involvement of integrin receptor signaling in insulin action, to date there has not been any investigation of integrin function on insulin action or glucose homeostasis in vivo. To address this issue, we have taken advantage of Cre-LoxP technology to inactivate the β1 integrin receptor subunit gene in striated muscle. We have observed that muscle creatine kinase-specific integrin β1 knock-out (MCKItgβ1 KO) mice display a reduction of insulin-stimulated glucose infusion rate and glucose clearance. The impairment of insulin-stimulated skeletal muscle glucose uptake and glycogen synthesis resulted from a decrease in Akt Ser-473 phosphorylation concomitant with a marked reduction in ILK expression. Together, these data demonstrate an important cross-talk between integrin receptor function and insulin action and suggests that ILK may function as an Akt Ser-473 kinase in skeletal muscle.  相似文献   

4.
Autophagy is a degradative process that recycles long-lived and faulty cellular components. It is linked to many diseases and is required for normal development. ULK1, a mammalian serine/threonine protein kinase, plays a key role in the initial stages of autophagy, though the exact molecular mechanism is unknown. Here we report identification of a novel protein complex containing ULK1 and two additional protein factors, FIP200 and ATG13, all of which are essential for starvation-induced autophagy. Both FIP200 and ATG13 are critical for correct localization of ULK1 to the pre-autophagosome and stability of ULK1 protein. Additionally, we demonstrate by using both cellular experiments and a de novo in vitro reconstituted reaction that FIP200 and ATG13 can enhance ULK1 kinase activity individually but both are required for maximal stimulation. Further, we show that ATG13 and ULK1 are phosphorylated by the mTOR pathway in a nutrient starvation-regulated manner, indicating that the ULK1·ATG13·FIP200 complex acts as a node for integrating incoming autophagy signals into autophagosome biogenesis.Macroautophagy (herein referred to as autophagy) is a catabolic process whereby long-lived proteins and damaged organelles are shuttled to lysosomes for degradation. This process is conserved in all eukaryotes. Under normal growth conditions a housekeeping level of autophagy exists. Under stress, such as nutrient starvation, autophagy is strongly induced resulting in the engulfment of cytosolic components and organelles in specialized double-membrane structures termed autophagosomes. Following fusion of the outer autophagosomal membrane with lysosomes, the inner membrane and its cytoplasmic cargo are degraded and recycled (13). Recent work has implicated autophagy in many disease pathologies, including cancer, neurodegeneration, as well as in eliminating intracellular pathogens (48).The morphology of autophagy was first described in mammalian cells over 50 years ago (9). However, it is only recently through yeast genetic screens, that multiple autophagy-related (ATG) genes have been identified (1012). The yeast ATG proteins have been classified into four major groups: the Atg1 protein kinase complex, the Vps34 phosphatidylinositol 3-phosphate kinase complex, the Atg8/Atg12 conjugation systems, and the Atg9 recycling complex (13). Even though many ATG genes are now known, most of which have functional homologs in mammalian cells (14, 15), the molecular mechanism by which they sense the initial triggers and subsequently dictate autophagy-specific intracellular membrane events is far from understood.In yeast, one of the earliest autophagy-specific events is believed to involve the Atg1 protein kinase complex. Atg1 is a serine/threonine protein kinase and a key autophagy-regulator (16). Atg1 is complexed to at least two other proteins during autophagy, Atg13 and Atg17, both of which are required for normal Atg1 function and autophagosome generation (1719). Classical signaling pathways such as the cAMP-dependent kinase (PKA) pathway or the Tor kinase pathway appear to converge upon this complex, placing Atg1 at an early stage during autophagosome biogenesis (2022). Atg1 phosphorylation by PKA blocks its association with the forming autophagosome (21), while the Tor pathway hyperphosphorylates Atg13 causing a reduced affinity of Atg13 for Atg1, resulting in repression of autophagy (17, 19). In contrast, nutrient starvation or inhibition of Tor leads to dephosphorylation of Atg13 thus increased Atg1 complex formation and kinase activity, resulting in stimulation of autophagy (19). Surprisingly, the physiological substrates of Atg1 kinase have not been identified; thus how Atg1 transduces upstream autophagic signaling is undefined. Recently, mammalian homologs of Atg1 have been identified as ULK1 and ULK2 (Unc-51-like kinase)2 (2325). ULK1 and ULK2 are ubiquitously expressed and localize to the isolation membrane, or forming autophagosome, upon nutrient starvation (25); RNAi-mediated depletion of ULK1 in HEK293 cells compromises autophagy (23, 24). The exact role of ULK1 versus ULK2 in autophagy is unclear, and it is possible some redundancy exists between the two isoforms (26).Given the conservation of autophagy from yeast to man, it is interesting to note that no mammalian counterpart to yeast Atg13 or Atg17 had been identified until very recently. The protein FIP200 (focal adhesion kinase family-interacting protein of 200 kDa) was identified as an autophagy-essential binding partner of both ULK1 and ULK2 (25), and it has been speculated that FIP200 might be the equivalent of yeast Atg17, despite low sequence similarity (25, 27).In this study, we delve deeper into the molecular regulation of ULK1 to gain a better insight into how mammalian signaling pathways affect autophagy initiation. We describe here the identification of a triple complex consisting of ULK1, FIP200, and the mammalian equivalent of Atg13. This complex is required not only for localization of ULK1 to the isolation membrane but also for maximal kinase activity. In addition, both ATG13 and ULK1 are kinase substrates in the mTOR pathway and thus might function to sense nutrient starvation. Therefore, this study defines the role of mammalian ULK1-ATG13-FIP200 complex in mediating the initial autophagic triggers and to transduce the signal to the core autophagic machinery.  相似文献   

5.
The paralyzed zebrafish strain relaxed carries a null mutation for the skeletal muscle dihydropyridine receptor (DHPR) β1a subunit. Lack of β1a results in (i) reduced membrane expression of the pore forming DHPR α1S subunit, (ii) elimination of α1S charge movement, and (iii) impediment of arrangement of the DHPRs in groups of four (tetrads) opposing the ryanodine receptor (RyR1), a structural prerequisite for skeletal muscle-type excitation-contraction (EC) coupling. In this study we used relaxed larvae and isolated myotubes as expression systems to discriminate specific functions of β1a from rather general functions of β isoforms. Zebrafish and mammalian β1a subunits quantitatively restored α1S triad targeting and charge movement as well as intracellular Ca2+ release, allowed arrangement of DHPRs in tetrads, and most strikingly recovered a fully motile phenotype in relaxed larvae. Interestingly, the cardiac/neuronal β2a as the phylogenetically closest, and the ancestral housefly βM as the most distant isoform to β1a also completely recovered α1S triad expression and charge movement. However, both revealed drastically impaired intracellular Ca2+ transients and very limited tetrad formation compared with β1a. Consequently, larval motility was either only partially restored (β2a-injected larvae) or not restored at all (βM). Thus, our results indicate that triad expression and facilitation of 1,4-dihydropyridine receptor (DHPR) charge movement are common features of all tested β subunits, whereas the efficient arrangement of DHPRs in tetrads and thus intact DHPR-RyR1 coupling is only promoted by the β1a isoform. Consequently, we postulate a model that presents β1a as an allosteric modifier of α1S conformation enabling skeletal muscle-type EC coupling.Excitation-contraction (EC)3 coupling in skeletal muscle is critically dependent on the close interaction of two distinct Ca2+ channels. Membrane depolarizations of the myotube are sensed by the voltage-dependent 1,4-dihydropyridine receptor (DHPR) in the sarcolemma, leading to a rearrangement of charged amino acids (charge movement) in the transmembrane segments S4 of the pore-forming DHPR α1S subunit (1, 2). This conformational change induces via protein-protein interaction (3, 4) the opening of the sarcoplasmic type-1 ryanodine receptor (RyR1) without need of Ca2+ influx through the DHPR (5). The release of Ca2+ from the sarcoplasmic reticulum via RyR1 consequently induces muscle contraction. The protein-protein interaction mechanism between DHPR and RyR1 requires correct ultrastructural targeting of both channels. In Ca2+ release units (triads and peripheral couplings) of the skeletal muscle, groups of four DHPRs (tetrads) are coupled to every other RyR1 and hence are geometrically arranged following the RyR-specific orthogonal arrays (6).The skeletal muscle DHPR is a heteromultimeric protein complex, composed of the voltage-sensing and pore-forming α1S subunit and auxiliary subunits β1a, α2δ-1, and γ1 (7). While gene knock-out of the DHPR γ1 subunit (8, 9) and small interfering RNA knockdown of the DHPR α2δ-1 subunit (10-12) have indicated that neither subunit is essential for coupling of the DHPR with RyR1, the lack of the α1S or of the intracellular β1a subunit is incompatible with EC coupling and accordingly null model mice die perinatally due to asphyxia (13, 14). β subunits of voltage-gated Ca2+ channels were repeatedly shown to be responsible for the facilitation of α1 membrane insertion and to be potent modulators of α1 current kinetics and voltage dependence (15, 16). Whether the loss of EC coupling in β1-null mice was caused by decreased DHPR membrane expression or by the lack of a putative specific contribution of the β subunit to the skeletal muscle EC coupling apparatus (17, 18) was not clearly resolved. Recently, other β-functions were identified in skeletal muscle using the β1-null mutant zebrafish relaxed (19, 20). Like the β1-knock-out mouse (14) zebrafish relaxed is characterized by complete paralysis of skeletal muscle (21, 22). While β1-knock-out mouse pups die immediately after birth due to respiratory paralysis (14), larvae of relaxed are able to survive for several days because of oxygen and metabolite diffusion via the skin (23). Using highly differentiated myotubes that are easy to isolate from these larvae, the lack of EC coupling could be described by quantitative immunocytochemistry as a moderate ∼50% reduction of α1S membrane expression although α1S charge movement was nearly absent, and, most strikingly, as the complete lack of the arrangement of DHPRs in tetrads (19). Thus, in skeletal muscle the β subunit enables EC coupling by (i) enhancing α1S membrane targeting, (ii) facilitating α1S charge movement, and (iii) enabling the ultrastructural arrangement of DHPRs in tetrads.The question arises, which of these functions are specific for the skeletal muscle β1a and which ones are rather general properties of Ca2+ channel β subunits. Previous reconstitution studies made in the β1-null mouse system (24, 25) using different β subunit constructs (26) did not allow differentiation between β-induced enhancement of non-functional α1S membrane expression and the facilitation of α1S charge movement, due to the lack of information on α1S triad expression levels. Furthermore, the β-induced arrangement of DHPRs in tetrads was not detected as no ultrastructural information was obtained.In the present study, we established zebrafish mutant relaxed as an expression system to test different β subunits for their ability to restore skeletal muscle EC coupling. Using isolated myotubes for in vitro experiments (19, 27) and complete larvae for in vivo expression studies (28-31) and freeze-fracture electron microscopy, a clear differentiation between the major functional roles of β subunits was feasible in the zebrafish system. The cloned zebrafish β1a and a mammalian (rabbit) β1a were shown to completely restore all parameters of EC coupling when expressed in relaxed myotubes and larvae. However, the phylogenetically closest β subunit to β1a, the cardiac/neuronal isoform β2a from rat, as well as the ancestral βM isoform from the housefly (Musca domestica), could recover functional α1S membrane insertion, but led to very restricted tetrad formation when compared with β1a, and thus to impaired DHPR-RyR1 coupling. This impairment caused drastic changes in skeletal muscle function.The present study shows that the enhancement of functional α1S membrane expression is a common function of all the tested β subunits, from β1a to even the most distant βM, whereas the effective formation of tetrads and thus proper skeletal muscle EC coupling is an exclusive function of the skeletal muscle β1a subunit. In context with previous studies, our results suggest a model according to which β1a acts as an allosteric modifier of α1S conformation. Only in the presence of β1a, the α1S subunit is properly folded to allow RyR1 anchoring and thus skeletal muscle-type EC coupling.  相似文献   

6.
Integrin α8β1 interacts with a variety of Arg-Gly-Asp (RGD)-containing ligands in the extracellular matrix. Here, we examined the binding activities of α8β1 integrin toward a panel of RGD-containing ligands. Integrin α8β1 bound specifically to nephronectin with an apparent dissociation constant of 0.28 ± 0.01 nm, but showed only marginal affinities for fibronectin and other RGD-containing ligands. The high-affinity binding to α8β1 integrin was fully reproduced with a recombinant nephronectin fragment derived from the RGD-containing central “linker” segment. A series of deletion mutants of the recombinant fragment identified the LFEIFEIER sequence on the C-terminal side of the RGD motif as an auxiliary site required for high-affinity binding to α8β1 integrin. Alanine scanning mutagenesis within the LFEIFEIER sequence defined the EIE sequence as a critical motif ensuring the high-affinity integrin-ligand interaction. Although a synthetic LFEIFEIER peptide failed to inhibit the binding of α8β1 integrin to nephronectin, a longer peptide containing both the RGD motif and the LFEIFEIER sequence was strongly inhibitory, and was ∼2,000-fold more potent than a peptide containing only the RGD motif. Furthermore, trans-complementation assays using recombinant fragments containing either the RGD motif or LFEIFEIER sequence revealed a clear synergism in the binding to α8β1 integrin. Taken together, these results indicate that the specific high-affinity binding of nephronectin to α8β1 integrin is achieved by bipartite interaction of the integrin with the RGD motif and LFEIFEIER sequence, with the latter serving as a synergy site that greatly potentiates the RGD-driven integrin-ligand interaction but has only marginal activity to secure the interaction by itself.Integrins are a family of adhesion receptors that interact with a variety of extracellular ligands, typically cell-adhesive proteins in the extracellular matrix (ECM).2 They play mandatory roles in embryonic development and the maintenance of tissue architectures by providing essential links between cells and the ECM (1). Integrins are composed of two non-covalently associated subunits, termed α and β. In mammals, 18 α and 8 β subunits have been identified, and combinations of these subunits give rise to at least 24 distinct integrin heterodimers. Based on their ligand-binding specificities, ECM-binding integrins are classified into three groups, namely laminin-, collagen- and RGD-binding integrins (2, 3), of which the RGD-binding integrins have been most extensively investigated. The RGD-binding integrins include α5β1, α8β1, αIIbβ3, and αV-containing integrins, and have been shown to interact with a variety of ECM ligands, such as fibronectin and vitronectin, with distinct binding specificities.The α8 integrin subunit was originally identified in chick nerves (4). Integrin α8β1 is expressed in the metanephric mesenchyme and plays a crucial role in epithelial-mesenchymal interactions during the early stages of kidney morphogenesis. Disruption of the α8 gene in mice was found to be associated with severe defects in kidney morphogenesis (5) and stereocilia development (6). To date, α8β1 integrin has been shown to bind to fibronectin, vitronectin, osteopontin, latency-associated peptide of transforming growth factor-β1, tenascin-W, and nephronectin (also named POEM) (713), among which nephronectin is believed to be an α8β1 integrin ligand involved in kidney development (10).Nephronectin is one of the basement membrane proteins whose expression and localization patterns are restricted in a tissue-specific and developmentally regulated manner (10, 11). Nephronectin consists of five epidermal growth factor-like repeats, a linker segment containing the RGD cell-adhesive motif (designated RGD-linker) and a meprin-A5 protein-receptor protein-tyrosine phosphatase μ (MAM) domain (see Fig. 3A). Although the physiological functions of nephronectin remain only poorly understood, it is thought to play a role in epithelial-mesenchymal interactions through binding to α8β1 integrin, thereby transmitting signals from the epithelium to the mesenchyme across the basement membrane (10). Recently, mice deficient in nephronectin expression were produced by homologous recombination (14). These nephronectin-deficient mice frequently displayed kidney agenesis, a phenotype reminiscent of α8 integrin knock-out mice (14), despite the fact that other RGD-containing ligands, including fibronectin and osteopontin, were expressed in the embryonic kidneys (9, 15). The failure of the other RGD-containing ligands to compensate for the deficiency of nephronectin in the developing kidneys suggests that nephronectin is an indispensable α8β1 ligand that plays a mandatory role in epithelial-mesenchymal interactions during kidney development.Open in a separate windowFIGURE 3.Binding activities of α8β1 integrin to nephronectin and its fragments. A, schematic diagrams of full-length nephronectin (NN) and its fragments. RGD-linker and RGD-linker (GST), the central RGD-containing linker segments expressed in mammalian and bacterial expression systems, respectively; PRGDV, a short RGD-containing peptide modeled after nephronectin and expressed as a GST fusion protein (see Fig. 4A for the peptide sequence). The arrowheads indicate the positions of the RGD motif. B, purified recombinant proteins were analyzed by SDS-PAGE in 7–15% gradient (left and center panels) and 12% (right panels) gels, followed by Coomassie Brilliant Blue (CBB) staining, immunoblotting with an anti-FLAG mAb, or lectin blotting with PNA. The quantities of proteins loaded were: 0.5 μg (for Coomassie Brilliant Blue staining) and 0.1 μg (for blotting with anti-FLAG and PNA) in the left and center panels;1 μg in the right panel. C, recombinant proteins (10 nm) were coated on microtiter plates and assessed for their binding activities toward α8β1 integrin (10 nm) in the presence of 1 mm Mn2+. The backgrounds were subtracted as described in the legend to Fig. 2. The results represent the mean ± S.D. of triplicate determinations. D, titration curves of α8β1 integrin bound to full-length nephronectin (NN, closed squares), the RGD-linker segments expressed in 293F cells (RGD-linker, closed triangles) and E. coli (RGD-linker (GST), open triangles), the MAM domain (MAM, closed diamonds), and the PRGDV peptide expressed as a GST fusion protein in E. coli (PRGDV (GST), open circles). The assays were performed as described in the legend to Fig. 2B. The results represent the means of duplicate determinations.Although ligand recognition by RGD-binding integrins is primarily determined by the RGD motif in the ligands, it is the residues outside the RGD motif that define the binding specificities and affinities toward individual integrins (16, 17). For example, α5β1 integrin specifically binds to fibronectin among the many RGD-containing ligands, and requires not only the RGD motif in the 10th type III repeat but also the so-called “synergy site” within the preceding 9th type III repeat for fibronectin recognition (18). Recently, DiCara et al. (19) demonstrated that the high-affinity binding of αVβ6 integrin to its natural ligands, e.g. foot-and-mouth disease virus, requires the RGD motif immediately followed by a Leu-Xaa-Xaa-Leu/Ile sequence, which forms a helix to align the two conserved hydrophobic residues along the length of the helix. Given the presence of many naturally occurring RGD-containing ligands, it is conceivable that the specificities of the RGD-binding integrins are dictated by the sequences flanking the RGD motif or those in neighboring domains that come into close proximity with the RGD motif in the intact ligand proteins. However, the preferences of α8β1 integrin for RGD-containing ligands and how it secures its high-affinity binding toward its preferred ligands remain unknown.In the present study, we investigated the binding specificities of α8β1 integrin toward a panel of RGD-containing cell-adhesive proteins. Our data reveal that nephronectin is a preferred ligand for α8β1 integrin, and that a LFEIFEIER sequence on the C-terminal side of its RGD motif serves as a synergy site to ensure the specific high-affinity binding of nephronectin to α8β1 integrin.  相似文献   

7.
Proteolytic processing of amyloid precursor protein (APP) by β- and γ-secretases generates β-amyloid (Aβ) peptides, which accumulate in the brains of individuals affected by Alzheimer disease. Detergent-resistant membrane microdomains (DRM) rich in cholesterol and sphingolipid, termed lipid rafts, have been implicated in Aβ production. Previously, we and others reported that the four integral subunits of the γ-secretase associate with DRM. In this study we investigated the mechanisms underlying DRM association of γ-secretase subunits. We report that in cultured cells and in brain the γ-secretase subunits nicastrin and APH-1 undergo S-palmitoylation, the post-translational covalent attachment of the long chain fatty acid palmitate common in lipid raft-associated proteins. By mutagenesis we show that nicastrin is S-palmitoylated at Cys689, and APH-1 is S-palmitoylated at Cys182 and Cys245. S-Palmitoylation-defective nicastrin and APH-1 form stable γ-secretase complexes when expressed in knock-out fibroblasts lacking wild type subunits, suggesting that S-palmitoylation is not essential for γ-secretase assembly. Nevertheless, fractionation studies show that S-palmitoylation contributes to DRM association of nicastrin and APH-1. Moreover, pulse-chase analyses reveal that S-palmitoylation is important for nascent polypeptide stability of both proteins. Co-expression of S-palmitoylation-deficient nicastrin and APH-1 in cultured cells neither affects Aβ40, Aβ42, and AICD production, nor intramembrane processing of Notch and N-cadherin. Our findings suggest that S-palmitoylation plays a role in stability and raft localization of nicastrin and APH-1, but does not directly modulate γ-secretase processing of APP and other substrates.Alzheimer disease is the most common among neurodegenerative diseases that cause dementia. This debilitating disorder is pathologically characterized by the cerebral deposition of 39–42 amino acid peptides termed Aβ, which are generated by proteolytic processing of amyloid precursor protein (APP)2 by β- and γ-secretases (1, 2). The β-site APP cleavage enzyme 1 cleaves full-length APP within its luminal domain to generate a secreted ectodomain leaving behind a C-terminal fragment (β-CTF). γ-Secretase cleaves β-CTF within the transmembrane domain to release Aβ and APP intracellular C-terminal domain (AICD). γ-Secretase is a multiprotein complex, comprising at least four subunits: presenilins (PS1 and PS2), nicastrin, APH-1, and PEN-2 for its activity (3). PS1 is synthesized as a 42–43-kDa polypeptide and undergoes highly regulated endoproteolytic processing within the large cytoplasmic loop domain connecting putative transmembrane segments 6 and 7 to generate stable N-terminal (NTF) and C-terminal fragments (CTF) by an uncharacterized proteolytic activity (4). This endoproteolytic event has been identified as the activation step in the process of PS1 maturation as it assembles with other γ-secretase subunits (3). Nicastrin is a heavily glycosylated type I membrane protein with a large ectodomain that has been proposed to function in substrate recognition and binding (5), but this putative function has not been confirmed by others (6). APH-1 is a seven-transmembrane protein encoded by two human or three rodent genes that are alternatively spliced (7). Although PS1 (or PS2), nicastrin, APH-1, and PEN-2 are sufficient for γ-secretase processing of APP, a type I membrane protein, termed p23 (also referred toTMP21), was recently identified as a γ-secretase component that modulates γ-secretase activity and regulates secretory trafficking of APP (8, 9).A growing number of type I integral membrane proteins has been identified as γ-secretase substrates within the last few years, including Notch1 homologues, Notch ligands, Delta and Jagged, cell adhesion receptors N- and E-cadherins, low density lipoprotein receptor-related protein, ErbB-4, netrin receptor DCC, and others (10). Mounting evidence suggests that APP processing occurs within cholesterol- and sphingolipid-enriched lipid rafts, which are biochemically defined as detergentresistant membrane microdomains (DRM) (11, 12). Previously we reported that each of the γ-secretase subunits localizes in lipid rafts in post-Golgi and endosome membranes enriched in syntaxin 6 (13). Moreover, loss of γ-secretase activity by gene deletion or exposure to γ-secretase inhibitors results in the accumulation of APP CTFs in lipid rafts indicating that cleavage of APP CTFs likely occurs in raft microdomains (14). In contrast, CTFs derived from Notch1, Jagged2, N-cadherin, and DCC are processed by γ-secretase in non-raft membranes (14). The mechanisms underlying association of γ-secretase subunits with lipid rafts need further clarification to elucidate spatial segregation of amyloidogenic processing of APP in membrane microdomains.Post-translational S-palmitoylation is increasingly recognized as a potential mechanism for regulating raft association, stability, intracellular trafficking, and function of several cytosolic and transmembrane proteins (1517). S-palmitoylation refers to the addition of 16-carbon palmitoyl moiety to certain cysteine residues through thioester linkage. Cysteines close to transmembrane domains or membrane-associated domains in non-integral membrane proteins are preferred S-palmitoylation sites, although no conserved motif has been identified (18). Palmitoylation modifies numerous neuronal proteins, including postsynaptic density protein PSD-95 (19), a-amino-3-hydroxyl-5-methyl-4-isoxazole propionic acid receptors (20), nicotinic α7 receptors (21), neuronal t-SNAREs SNAP-25, synaptobrevin 2 and synaptogagmin (22, 23), neuronal growth-associated protein GAP-43 (24), protein kinase CLICK-III (CL3)/CaMKIγ (25), β-secretase (26), and Huntingtin (27). Although palmitoylation can occur in vitro without the involvement of an enzyme, a family of palmitoyltransferases that specifically catalyze S-palmitoylation has been identified (28, 29).In this study, we have identified S-palmitoylation of γ-secretase subunits nicastrin and APH-1, and characterized its role on DRM association, protein stability, and γ-secretase enzyme activities. We show that nicastrin is S-palmitoylated at Cys689, and APH-1 at Cys182 and Cys245. Mutagenesis of palmitoylation sites results in increased degradation of nascent nicastrin and APH-1 polypeptides and reduced association with DRM. Nevertheless, in cultured cells overexpression of S-palmitoylation-deficient nicastrin and APH-1 does not modulate γ-secretase processing of APP or other substrates.  相似文献   

8.
Activation of protein kinase C (PKC) promotes the salvage pathway of ceramide formation, and acid sphingomyelinase has been implicated, in part, in providing substrate for this pathway (Zeidan, Y. H., and Hannun, Y. A. (2007) J. Biol. Chem. 282, 11549–11561). In the present study, we examined whether acid β-glucosidase 1 (GBA1), which hydrolyzes glucosylceramide to form lysosomal ceramide, was involved in PKC-regulated formation of ceramide from recycled sphingosine. Glucosylceramide levels declined after treatment of MCF-7 cells with a potent PKC activator, phorbol 12-myristate 13-acetate (PMA). Silencing GBA1 by small interfering RNAs significantly attenuated acid glucocerebrosidase activity and decreased PMA-induced formation of ceramide by 50%. Silencing GBA1 blocked PMA-induced degradation of glucosylceramide and generation of sphingosine, the source for ceramide biosynthesis. Reciprocally, forced expression of GBA1 increased ceramide levels. These observations indicate that GBA1 activation can generate the source (sphingosine) for PMA-induced formation of ceramide through the salvage pathway. Next, the role of PKCδ, a direct effector of PMA, in the formation of ceramide was determined. By attenuating expression of PKCδ, cells failed to trigger PMA-induced alterations in levels of ceramide, sphingomyelin, and glucosylceramide. Thus, PKCδ activation is suggested to stimulate the degradation of both sphingomyelin and glucosylceramide leading to the salvage pathway of ceramide formation. Collectively, GBA1 is identified as a novel source of regulated formation of ceramide, and PKCδ is an upstream regulator of this pathway.Sphingolipids are abundant components of cellular membranes, many of which are emerging as bioactive lipid mediators thought to play crucial roles in cellular responses (1, 2). Ceramide, a central sphingolipid, serves as the main precursor for various sphingolipids, including glycosphingolipids, gangliosides, and sphingomyelin. Regulation of formation of ceramide has been demonstrated through the action of three major pathways: the de novo pathway (3, 4), the sphingomyelinase pathway (5), and the salvage pathway (68). The latter plays an important role in constitutive sphingolipid turnover by salvaging long-chain sphingoid bases (sphingosine and dihydrosphingosine) that serve as sphingolipid backbones for ceramide and dihydroceramide as well as all complex sphingolipids (Fig. 1A).Open in a separate windowFIGURE 1.The scheme of the sphingosine salvage pathway of ceramide formation and inhibition of PMA induction of ceramide by fumonisin B1. A, the scheme of the sphingosine salvage pathway of ceramide formation. B, previously published data as to effects of fumonisin B1 on ceramide mass profiles (23) are re-plotted as a PMA induction of ceramide. In brief, MCF-7 cells were pretreated with or without 100 μm fumonisin B1 for 2 h followed by treatment with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. Results are expressed as sum of increased mass of ceramide species. Dotted or open columns represents C16-ceramide or sum of other ceramide species (C14-ceramide, C18-ceramide, C18:1-ceramide, C20-ceramide, C24-ceramide, and C24:1-ceramide), respectively. The data represent mean ± S.E. of three to five values.Metabolically, ceramide is also formed from degradation of glycosphingolipids (Fig. 1A) usually in acidic compartments, the lysosomes and/or late endosomes (9). The stepwise hydrolysis of complex glycosphingolipids eventually results in the formation of glucosylceramide, which in turn is converted to ceramide by the action of acid β-glucosidase 1 (GBA1)2 (9, 10). Severe defects in GBA1 activity cause Gaucher disease, which is associated with aberrant accumulation of the lipid substrates (1014). On the other hand, sphingomyelin is cleaved by acid sphingomyelinase to also form ceramide (15, 16). Either process results in the generation of lysosomal ceramide that can then be deacylated by acid ceramidase (17), releasing sphingosine that may escape the lysosome (18). The released sphingosine may become a substrate for either sphingosine kinases or ceramide synthases, forming sphingosine 1-phosphate or ceramide, respectively (3, 1921).In a related line of investigation, our studies (20, 22, 23) have begun to implicate protein kinase Cs (PKC) as upstream regulators of the sphingoid base salvage pathway resulting in ceramide synthesis. Activation of PKCs by the phorbol ester (PMA) was shown to stimulate the salvage pathway resulting in increases in ceramide. All the induced ceramide was inhibited by pretreatment with a ceramide synthase inhibitor, fumonisin B1, but not by myriocin, thus negating acute activation of the de novo pathway and establishing a role for ceramide synthesis (20, 23). Moreover, labeling studies also implicated the salvage pathway because PMA induced turnover of steady state-labeled sphingolipids but did not affect de novo labeled ceramide in pulse-chase experiments.Moreover, PKCδ, among PKC isoforms, was identified as an upstream molecule for the activation of acid sphingomyelinase in the salvage pathway (22). Interestingly, the PKCδ isoform induced the phosphorylation of acid sphingomyelinase at serine 508, leading to its activation and consequent formation of ceramide. The activation of acid sphingomyelinase appeared to contribute to ∼50% of the salvage pathway-induced increase in ceramide (28) (also, see Fig. 4C). This raised the possibility that distinct routes of ceramide metabolism may account for the remainder of ceramide generation. In this study, we investigated glucocerebrosidase GBA1 as a candidate for one of the other routes accounting for PKC-regulated salvage pathway of ceramide formation.Open in a separate windowFIGURE 4.Effects of knockdown of lysosomal enzymes on the generation of ceramide after PMA treatment. A, MCF-7 cells were transfected with 5 nm siRNAs of each of four individual sequences (SCR, GBA1-a, GBA1-b, and GBA1-c) for 48 h and then stimulated with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of the C16-ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. The data represent mean ± S.E. of three to nine values. B, MCF-7 cells were transfected with 5 nm siRNAs of SCR or GBA1-a (GBA1) for 48 h and then stimulated with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of individual ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. The data represent mean ± S.E. of three to five values. C14-Cer, C14-ceramide; C16-Cer, C16-ceramide; C18-Cer; C18-ceramide; C18:1-Cer, C18:1-ceramide; C20-Cer, C20-ceramide; C20-Cer, C24-ceramide; C24:1-Cer, C24:1-ceramide. C, MCF-7 cells were transfected with 5 nm siRNAs of SCR, acid sphingomyelinase (ASM), or GBA1-a (GBA1) for 48 h following stimulation with (PMA) or without (Control) 100 nm PMA for 1 h. Lipids were extracted, and then the levels of ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. Levels of C16-ceramide are shown. The data represent mean ± S.E. of four to five values. Significant changes from SCR-transfected cells treated with PMA are shown in A–C (*, p < 0.02; **, p < 0.05; ***, p < 0.01).  相似文献   

9.
10.
Vasopressin controls renal water excretion largely through actions to regulate the water channel aquaporin-2 in collecting duct principal cells. Our knowledge of the mechanisms involved has increased markedly in recent years with the advent of methods for large-scale systems-level profiling such as protein mass spectrometry, yeast two-hybrid analysis, and oligonucleotide microarrays. Here we review this progress.Regulation of water excretion by the kidney is one of the most visible aspects of everyday physiology. An outdoor tennis game on a hot summer day can result in substantial water losses by sweating, and the kidneys respond by reducing water excretion. In contrast, excessive intake of water, a frequent occurrence in everyday life, results in excretion of copious amounts of clear urine. These responses serve to exact tight control on the tonicity of body fluids, maintaining serum osmolality in the range of 290–294 mosmol/kg of H2O through the regulated return of water from the pro-urine in the renal collecting ducts to the bloodstream.The importance of this process is highlighted when the regulation fails. For example, polyuria (rapid uncontrolled excretion of water) is a sometimes devastating consequence of lithium therapy for bipolar disorder. On the other side of the coin are water balance disorders that result from excessive renal water retention causing systemic hypo-osmolality or hyponatremia. Hyponatremia due to excessive water retention can be seen with severe congestive heart failure, hepatic cirrhosis, and the syndrome of inappropriate antidiuresis.The chief regulator of water excretion is the peptide hormone AVP,2 whereas the chief molecular target for regulation is the water channel AQP2. In this minireview, we describe new progress in the understanding of the molecular mechanisms involved in regulation of AQP2 by AVP in collecting duct cells, with emphasis on new information derived from “systems-level” approaches involving large-scale profiling and screening techniques such as oligonucleotide arrays, protein mass spectrometry, and yeast two-hybrid analysis. Most of the progress with these techniques is in the identification of individual molecules involved in AVP signaling and binding interactions with AQP2. Additional related issues are addressed in several recent reviews (14).  相似文献   

11.
12.
Excessive accumulation of β-amyloid peptides in the brain is a major cause for the pathogenesis of Alzheimer disease. β-Amyloid is derived from β-amyloid precursor protein (APP) through sequential cleavages by β- and γ-secretases, whose enzymatic activities are tightly controlled by subcellular localization. Delineation of how intracellular trafficking of these secretases and APP is regulated is important for understanding Alzheimer disease pathogenesis. Although APP trafficking is regulated by multiple factors including presenilin 1 (PS1), a major component of the γ-secretase complex, and phospholipase D1 (PLD1), a phospholipid-modifying enzyme, regulation of intracellular trafficking of PS1/γ-secretase and β-secretase is less clear. Here we demonstrate that APP can reciprocally regulate PS1 trafficking; APP deficiency results in faster transport of PS1 from the trans-Golgi network to the cell surface and increased steady state levels of PS1 at the cell surface, which can be reversed by restoring APP levels. Restoration of APP in APP-deficient cells also reduces steady state levels of other γ-secretase components (nicastrin, APH-1, and PEN-2) and the cleavage of Notch by PS1/γ-secretase that is more highly correlated with cell surface levels of PS1 than with APP overexpression levels, supporting the notion that Notch is mainly cleaved at the cell surface. In contrast, intracellular trafficking of β-secretase (BACE1) is not regulated by APP. Moreover, we find that PLD1 also regulates PS1 trafficking and that PLD1 overexpression promotes cell surface accumulation of PS1 in an APP-independent manner. Our results clearly elucidate a physiological function of APP in regulating protein trafficking and suggest that intracellular trafficking of PS1/γ-secretase is regulated by multiple factors, including APP and PLD1.An important pathological hallmark of Alzheimer disease (AD)4 is the formation of senile plaques in the brains of patients. The major components of those plaques are β-amyloid peptides (Aβ), whose accumulation triggers a cascade of neurodegenerative steps ending in formation of senile plaques and intraneuronal fibrillary tangles with subsequent neuronal loss in susceptible brain regions (1, 2). Aβ is proteolytically derived from the β-amyloid precursor protein (APP) through sequential cleavages by β-secretase (BACE1), a novel membrane-bound aspartyl protease (3, 4), and by γ-secretase, a high molecular weight complex consisting of at least four components: presenilin (PS), nicastrin (NCT), anterior pharynx-defective-1 (APH-1), and presenilin enhancer-2 (PEN-2) (5, 6). APP is a type I transmembrane protein belonging to a protein family that includes APP-like protein 1 (APLP1) and 2 (APLP2) in mammals (7, 8). Full-length APP is synthesized in the endoplasmic reticulum (ER) and transported through the Golgi apparatus. Most secreted Aβ peptides are generated within the trans-Golgi network (TGN), also the major site of steady state APP in neurons (911). APP can be transported to the cell surface in TGN-derived secretory vesicles if not proteolyzed to Aβ or an intermediate metabolite. At the cell surface APP is either cleaved by α-secretase to produce soluble sAPPα (12) or reinternalized for endosomal/lysosomal degradation (13, 14). Aβ may also be generated in endosomal/lysosomal compartments (15, 16). In contrast to neurotoxic Aβ peptides, sAPPα possesses neuroprotective potential (17, 18). Thus, the subcellular distribution of APP and proteases that process it directly affect the ratio of sAPPα to Aβ, making delineation of the mechanisms responsible for regulating trafficking of all of these proteins relevant to AD pathogenesis.Presenilin (PS) is a critical component of the γ-secretase. Of the two mammalian PS gene homologues, PS1 and PS2, PS1 encodes the major form (PS1) in active γ-secretase (19, 20). Nascent PSs undergo endoproteolytic cleavage to generate an amino-terminal fragment (NTF) and a carboxyl-terminal fragment (CTF) to form a functional PS heterodimer (21). Based on observations that PSs possess two highly conserved aspartate residues indispensable for γ-secretase activity and that specific transition state analogue γ-secretase inhibitors bind to PS1 NTF/CTF heterodimers (5, 22), PSs are believed to be the catalytic component of the γ-secretase complex. PS assembles with three other components, NCT, APH-1, and PEN-2, to form the functional γ-secretase (5, 6). Strong evidence suggests that PS1/γ-secretase resides principally in the ER, early Golgi, TGN, endocytic and intermediate compartments, most of which (except the TGN) are not major subcellular sites for APP (23, 24). In addition to generating Aβ and cleaving APP to release the APP intracellular domain, PS1/γ-secretase cleaves other substrates such as Notch (25), cadherin (26), ErbB4 (27), and CD44 (28), releasing their respective intracellular domains. Interestingly, PS1/γ-secretase cleavage of different substrates seems to occur at different subcellular compartments; APP is mainly cleaved at the TGN and early endosome domains, whereas Notch is predominantly cleaved at the cell surface (9, 11, 29). Thus, perturbing intracellular trafficking of PS1/γ-secretase may alter interactions between PS1/γ-secretase and APP, contributing to either abnormal Aβ generation and AD pathogenesis or decreased access of PS1/γ-secretase to APP such that Aβ production is reduced. However, mechanisms regulating PS1/γ-secretase trafficking warrant further investigation.In addition to participating in γ-secretase activity, PS1 regulates intracellular trafficking of several membrane proteins, including other γ-secretase components (nicastrin, APH-1, and PEN-2) and the substrate APP (reviewed in Ref. 30). Intracellular APP trafficking is highly regulated and requires other factors such as mint family members and SorLA (2). Moreover, we recently found that phospholipase D1 (PLD1), a phospholipid-modifying enzyme that regulates membrane trafficking events, can interact with PS1, and can regulate budding of APP-containing vesicles from the TGN and delivery of APP to the cell surface (31, 32). Interestingly, Kamal et al. (33) identified an axonal membrane compartment that contains APP, BACE1, and PS1 and showed that fast anterograde axonal transport of this compartment is mediated by APP and kinesin-I, implying a traffic-regulating role for APP. Increased APP expression is also shown to decrease retrograde axonal transport of nerve growth factor (34). However, whether APP indeed regulates intracellular trafficking of proteins including BACE1 and PS1/γ-secretase requires further validation. In the present study we demonstrate that intracellular trafficking of PS1, as well as that of other γ-secretase components, but not BACE1, is regulated by APP. APP deficiency promotes cell surface delivery of PS1/γ-secretase complex and facilitates PS1/γ-secretase-mediated Notch cleavage. In addition, we find that PLD1 also regulates intracellular trafficking of PS1 through a different mechanism and more potently than APP.  相似文献   

13.
14.
15.
TANK-binding kinase 1 (TBK1) is a key component of Toll-like receptor-dependent and -independent signaling pathways. In response to microbial components, TBK1 activates interferon regulatory factor 3 (IRF3) and cytokine expression. Here we show that TBK1 is a novel target of the γ134.5 protein, a virulence factor whose expression is regulated in a temporal fashion. Remarkably, the γ134.5 protein is required to inhibit IRF3 phosphorylation, nuclear translocation, and the induction of antiviral genes in infected cells. When expressed in mammalian cells, the γ134.5 protein forms complexes with TBK1 and disrupts the interaction of TBK1 and IRF3, which prevents the induction of interferon and interferon-stimulated gene promoters. Down-regulation of TBK1 requires the amino-terminal domain. In addition, unlike wild type virus, a herpes simplex virus mutant lacking γ134.5 replicates efficiently in TBK1-/- cells but not in TBK1+/+ cells. Addition of exogenous interferon restores the antiviral activity in both TBK1-/- and TBK+/+ cells. Hence, control of TBK1-mediated cell signaling by the γ134.5 protein contributes to herpes simplex virus infection. These results reveal that TBK1 plays a pivotal role in limiting replication of a DNA virus.Herpes simplex virus 1 (HSV-1)3 is a large DNA virus that establishes latent or lytic infection, in which the virus triggers innate immune responses. In HSV-infected cells, a number of antiviral mechanisms operate in a cell type- and time-dependent manner (1). In response to double-stranded RNA (dsRNA), Toll-like receptor 3 (TLR3) recruits an adaptor TIR domain-containing adaptor inducing IFN-β and stimulates cytokine expression (2, 3). In the cytoplasm, RNA helicases, RIG-I (retinoid acid-inducible gene-I), and MDA5 (melanoma differentiation associated gene 5) recognize intracellular viral 5′-triphosphate RNA or dsRNA (2, 4). Furthermore, a DNA-dependent activator of IFN-regulatory factor (DAI) senses double-stranded DNA in the cytoplasm and induces cytokine expression (5). There is also evidence that viral entry induces antiviral programs independent of TLR and RIG-I pathways (6). While recognizing distinct viral components, these innate immune pathways relay signals to the two IKK-related kinases, TANK-binding kinase 1 (TBK1) and inducible IκB kinase (IKKi) (2).The IKK-related kinases function as essential components that phosphorylate IRF3 (interferon regulatory factor 3), as well as the closely related IRF7, which translocates to the nucleus and induces antiviral genes, such as interferon-α/β and ISG56 (interferon-stimulated gene 56) (7, 8). TBK1 is constitutively expressed, whereas IKKi is engaged as an inducible gene product of innate immune signaling (9, 10). IRF3 activation is attenuated in TBK1-deficient but not in IKKi-deficient cells (11, 12). Its activation is completely abolished in double-deficient cells (12), suggesting a partially redundant function of TBK1 and IKKi. Indeed, IKKi also negatively regulates the STAT-signaling pathway (13). TBK1/IKKi interacts with several proteins, such as TRAF family member-associated NF-κB activator (TANK), NAP1 (NAK-associated protein 1), similar to NAP1TBK1 adaptor (SINTBAD), DNA-dependent activator of IFN-regulatory factors (DAI), and secretory protein 5 (Sec5) in host cells (5, 1418). These interactions are thought to regulate TBK1/IKKi, which delineates innate as well as adaptive immune responses.Upon viral infection, expression of HSV proteins interferes with the induction of antiviral immunity. When treated with UV or cycloheximide, HSV induces an array of antiviral genes in human lung fibroblasts (19, 20). Furthermore, an HSV mutant, with deletion in immediate early protein ICP0, induces ISG56 expression (21). Accordingly, expression of ICP0 inhibits the induction of antiviral programs mediated by IRF3 or IRF7 (2123). However, although ICP0 negatively regulates IFN-β expression, it is not essential for this effect (24). In HSV-infected human macrophages or dendritic cells, an immediate early protein ICP27 is required to suppress cytokine induction involving IRF3 (25). In this context, it is notable that an HSV mutant, lacking a leaky late gene γ134.5, replicates efficiently in cells devoid of IFN-α/β genes (26). Additionally, the γ134.5 null mutant induces differential cytokine expression as compared with wild type virus (27). Thus, HSV modulation of cytokine expression is a complex process that involves multiple viral components. Currently, the molecular mechanism governing this event is unclear. In this study, we show that HSV γ134.5 targets TBK1 and inhibits antiviral signaling. The data herein reveal a previously unrecognized mechanism by which γ134.5 facilitates HSV replication.  相似文献   

16.
17.
Osteopontin (OPN) is a cytokine and ligand for multiple members of the integrin family. OPN undergoes the in vivo polymerization catalyzed by cross-linking enzyme transglutaminase 2, which consequently increases the bioactivity through enhanced interaction with integrins. The integrin α9β1, highly expressed on neutrophils, binds to the sequence SVVYGLR only after intact OPN is cleaved by thrombin. The SVVYGLR sequence appears to be cryptic in intact OPN because α9β1 does not recognize intact OPN. Because transglutaminase 2-catalyzed polymers change their physical and chemical properties, we hypothesized that the SVVYGLR site might also be exposed on polymeric OPN. As expected, α9β1 turned into a receptor for polymeric OPN, a result obtained by cell adhesion and migration assays with α9-transfected cells and by detection of direct binding of recombinant soluble α9β1 with colorimetry and surface plasmon resonance analysis. Because the N-terminal fragment of thrombin-cleaved OPN, a ligand for α9β1, has been reported to attract neutrophils, we next examined migration of neutrophils to polymeric OPN using time-lapse microscopy. Polymeric OPN showed potent neutrophil chemotactic activity, which was clearly inhibited by anti-α9β1 antibody. Unexpectedly, mutagenesis studies showed that α9β1 bound to polymeric OPN independently of the SVVYGLR sequence, and further, SVVYGLR sequence of polymeric OPN was cryptic because SVVYGLR-specific antibody did not recognize polymeric OPN. These results demonstrate that polymerization of OPN generates a novel α9β1-binding site and that the interaction of this site with the α9β1 integrin is critical to the neutrophil chemotaxis induced by polymeric OPN.Acidic phosphorylated secreted glycoprotein osteopontin (OPN),4 known as a cytokine, has multiple functions, including roles in tissue remodeling, fibrosis, mineralization, immunomodulation, inflammation, and tumor metastasis (13). OPN is also an integrin ligand. At least nine integrins can function as OPN receptors. α5β1, α8β1, αvβ1, αvβ3, αvβ5 (1), and αvβ6 (4) recognize the linear tripeptide RGD, and α9β1, α4β1, and α4β7 recognize the sequence, SVVYGLR (5), adjacent to RGD but only after OPN has been cleaved by the protease, thrombin (Fig. 1).Open in a separate windowFIGURE 1.Schematic diagram of OPN. Two integrin-binding sites (boxed), a thrombin cleavage site (arrow), and a putative transglutamination site (circled) are shown. The term thrombin-cleaved nOPN is defined as in the figure.The overlap of receptors for OPN does not necessarily mean that these integrins play redundant roles in cellular responses to OPN because the patterns of integrin expression and utilization vary widely among cell types. In addition, interactions of different integrins with a single ligand can exert distinct effects on cell behavior in a single cell type. For example, we have previously reported that signals by ligation of αvβ3, αvβ6, or α9β1 to a single ligand, tenascin-C, differently affected cell adhesion, spreading, and proliferation of the colon cancer cell line, SW480 (6). Furthermore, intact OPN or thrombin- or matrix metalloproteinase-cleaved OPN interact with distinct subsets of integrins and exhibit distinct effects on cell behavior (4, 7, 8). Collectively, some of the functional diversity of OPN could be attributed to this multiplicity of receptors and responses. We have recently shown that polymerization of OPN results in enhanced biological activity (9). We thus set out to determine whether polymerized OPN exerts its effects through unique interactions with integrins.OPN is polymerized by transglutaminase 2 (TG2, EC 2.3.2.13) (10) that catalyzes formation of isopeptide cross-links between glutamine and lysine residues in substrate proteins (11) including OPN. Polymeric OPN has been identified in vivo in bone (12) and calcified aorta (13). We have previously reported that upon polymerization, OPN displays increased integrin binding accompanied by enhanced cell adhesion, spreading, migration, and focal contact formation (9). However, very little is known about how polymeric OPN induces its biological effects.Integrin α9β1, highly expressed on neutrophils (14), does not act as a receptor for intact OPN but does bind to an N-terminal fragment of OPN (nOPN) that is generated by thrombin cleavage (15) through the new C-terminal sequence, SVVYGLR. Protein polymerization can expose otherwise cryptic domains (16), so we hypothesized that the SVVYGLR site might be exposed upon polymerization and serve as a binding site for α9β1. In the present study, we demonstrate that α9β1 is indeed a receptor for polymeric OPN and that neutrophil migration induced by polymeric OPN is largely mediated by this interaction. However, mutational analysis and antibody studies demonstrate that this interaction does not involve the SVVYGLR site, suggesting the presence of de novo binding site in polymeric OPN.  相似文献   

18.
Recently we reported that N-glycans on the β-propeller domain of the integrin α5 subunit (S-3,4,5) are essential for α5β1 heterodimerization, expression, and cell adhesion. Herein to further investigate which N-glycosylation site is the most important for the biological function and regulation, we characterized the S-3,4,5 mutants in detail. We found that site-4 is a key site that can be specifically modified by N-acetylglucosaminyltransferase III (GnT-III). The introduction of bisecting GlcNAc into the S-3,4,5 mutant catalyzed by GnT-III decreased cell adhesion and migration on fibronectin, whereas overexpression of N-acetylglucosaminyltransferase V (GnT-V) promoted cell migration. The phenomenon is similar to previous observations that the functions of the wild-type α5 subunit were positively and negatively regulated by GnT-V and GnT-III, respectively, suggesting that the α5 subunit could be duplicated by the S-3,4,5 mutant. Interestingly GnT-III specifically modified the S-4,5 mutant but not the S-3,5 mutant. This result was confirmed by erythroagglutinating phytohemagglutinin lectin blot analysis. The reduction in cell adhesion was consistently observed in the S-4,5 mutant but not in the S-3,5 mutant cells. Furthermore mutation of site-4 alone resulted in a substantial decrease in erythroagglutinating phytohemagglutinin lectin staining and suppression of cell spread induced by GnT-III compared with that of either the site-3 single mutant or wild-type α5. These results, taken together, strongly suggest that N-glycosylation of site-4 on the α5 subunit is the most important site for its biological functions. To our knowledge, this is the first demonstration that site-specific modification of N-glycans by a glycosyltransferase results in functional regulation.Glycosylation is a crucial post-translational modification of most secreted and cell surface proteins (1). Glycosylation is involved in a variety of physiological and pathological events, including cell growth, migration, differentiation, and tumor invasion. It is well known that glycans play important roles in cell-cell communication, intracellular signal transduction, protein folding, and stability (2, 3).Integrins comprise a family of receptors that are important for cell adhesion. The major function of integrins is to connect cells to the extracellular matrix, activate intracellular signaling pathways, and regulate cytoskeletal formation (4). Integrin α5β1 is well known as a fibronectin (FN)3 receptor. The interaction between integrin α5 and FN is essential for cell migration, cell survival, and development (58). In addition, integrins are N-glycan carrier proteins. For example, α5β1 integrin contains 14 and 12 putative N-glycosylation sites on the α5 and β1 subunits, respectively. Several studies suggest that N-glycosylation is essential for functional integrin α5β1. When human fibroblasts were cultured in the presence of 1-deoxymannojirimycin, which prevents N-linked oligosaccharide processing, immature α5β1 integrin appeared on the cell surface, and FN-dependent adhesion was greatly reduced (9). Treatment of purified integrin α5β1 with N-glycosidase F, which cleaves between the innermost N-acetylglucosamine (GlcNAc) and asparagine N-glycan residues of N-linked glycoproteins, prevented the inherent association between subunits and blocked α5β1 binding to FN (10).A growing body of evidence indicates that the presence of the appropriate oligosaccharide can modulate integrin activation. N-Acetylglucosaminyltransferase III (GnT-III) catalyzes the addition of GlcNAc to mannose that is β1,4-linked to an underlying N-acetylglucosamine, producing what is known as a “bisecting” GlcNAc linkage as shown in Fig. 1B. GnT-III is generally regarded as a key glycosyltransferase in N-glycan biosynthetic pathways and contributes to inhibition of metastasis. The introduction of a bisecting GlcNAc catalyzed by GnT-III suppresses additional processing and elongation of N-glycans. These reactions, which are catalyzed in vitro by other glycosyltransferases, such as N-acetylglucosaminyltransferase V (GnT-V), which catalyzes the formation of β1,6 GlcNAc branching structures (Fig. 1B) and plays important roles in tumor metastasis, do not proceed because the enzymes cannot utilize the bisected N-glycans as a substrate. Introduction of the bisecting GlcNAc to integrin α5 by overexpression of GnT-III resulted in decreased in ligand binding and down-regulation of cell adhesion and migration (1113). Contrary to the functions of GnT-III, overexpression of GnT-V promoted integrin α5β1-mediated cell migration on FN (14). These observations clearly demonstrate that the alteration of N-glycan structure affected the biological functions of integrin α5β1. Similarly characterization of the carbohydrate moieties in integrin α3β1 from non-metastatic and metastatic human melanoma cell lines showed that expression of β1,6 GlcNAc branched structures was higher in metastatic cells compared with non-metastatic cells, confirming the notion that the β1,6 GlcNAc branched structure confers invasive and metastatic properties to cancer cells. In fact, Partridge et al. (15) reported that GnT-V-modified N-glycans containing poly-N-acetyllactosamine, the preferred ligand for galectin-3, on surface receptors oppose their constitutive endocytosis, promoting intracellular signaling and consequently cell migration and tumor metastasis.Open in a separate windowFIGURE 1.Potential N-glycosylation sites on the α5 subunit and its modification by GnT-III and GnT-V. A, schematic diagram of potential N-glycosylation sites on the α5 subunit. Putative N-glycosylation sites are indicated by triangles, and point mutations are indicated by crosses (N84Q, N182Q, N297Q, N307Q, N316Q, N524Q, N530Q, N593Q, N609Q, N675Q, N712Q, N724Q, N773Q, and N868Q). B, illustration of the reaction catalyzed by GnT-III and GnT-V. Square, GlcNAc; circle, mannose. TM, transmembrane domain.In addition, sialylation on the non-reducing terminus of N-glycans of α5β1 integrin plays an important role in cell adhesion. Colon adenocarcinomas express elevated levels of α2,6 sialylation and increased activity of ST6GalI sialyltransferase. Elevated ST6GalI positively correlated with metastasis and poor survival. Therefore, ST6GalI-mediated hypersialylation likely plays a role in colorectal tumor invasion (16, 17). In fact, oncogenic ras up-regulated ST6GalI and, in turn, increased sialylation of β1 integrin adhesion receptors in colon epithelial cells (18). However, this is not always the case. The expression of hyposialylated integrin α5β1 was induced by phorbol esterstimulated differentiation in myeloid cells in which the expression of the ST6GalI was down-regulated by the treatment, increasing FN binding (19). A similar phenomenon was also observed in hematopoietic or other epithelial cells. In these cells, the increased sialylation of the β1 integrin subunit was correlated with reduced adhesiveness and metastatic potential (2022). In contrast, the enzymatic removal of α2,8-linked oligosialic acids from the α5 integrin subunit inhibited cell adhesion to FN (23). Collectively these findings suggest that the interaction of integrin α5β1 with FN is dependent on its N-glycosylation and the processing status of N-glycans.Because integrin α5β1 contains multipotential N-glycosylation sites, it is important to determine the sites that are crucial for its biological function and regulation. Recently we found that N-glycans on the β-propeller domain (sites 3, 4, and 5) of the integrin α5 subunit are essential for α5β1 heterodimerization, cell surface expression, and biological function (24). In this study, to further investigate the underlying molecular mechanism of GnT-III-regulated biological functions, we characterized the N-glycans on the α5 subunit in detail using genetic and biochemical approaches and found that site-4 is a key site that can be specifically modified by GnT-III.  相似文献   

19.
20.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  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