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 296 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
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   

6.
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   

7.
8.
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   

9.
10.
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   

11.
Three heterozygous mutations were identified in the genes encoding platelet integrin receptor αIIbβ3 in a patient with an ill defined platelet disorder: one in the β3 gene (S527F) and two in the αIIb gene (R512W and L841M). Five stable Chinese hamster ovary cell lines were constructed expressing recombinant αIIbβ3 receptors bearing the individual R512W, L841M, or S527F mutation; both the R512W and L841M mutations; or all three mutations. All receptors were expressed on the cell surface, and mutations R512W and L841M had no effect on integrin function. Interestingly, the β3 S527F mutation produced a constitutively active receptor. Indeed, both fibrinogen and the ligand-mimetic antibody PAC-1 bound to non-activated αIIbβ3 receptors carrying the S527F mutation, indicating that the conformation of this receptor was altered and corresponded to the high affinity ligand binding state. In addition, the conformational change induced by S527F was evident from basal anti-ligand-induced binding site antibody binding to the receptor. A molecular model bearing this mutation was constructed based on the crystal structure of αIIbβ3 and revealed that the S527F mutation, situated in the third integrin epidermal growth factor-like (I-EGF3) domain, hindered the αIIbβ3 receptor from adopting a wild type-like bent conformation. Movement of I-EGF3 into a cleft in the bent conformation may be hampered both by steric hindrance between Phe527 in β3 and the calf-1 domain in αIIb and by decreased flexibility between I-EGF2 and I-EGF3.The platelet receptor αIIbβ3 belongs to the family of integrin receptors that consist of noncovalently linked α/β-heterodimers. They are cell-surface receptors that play a role in cell-cell and cell-matrix interactions. Under resting conditions, integrin receptors adopt the low affinity conformation and do not interact with their ligands. Inside-out signaling turns the receptor into a high affinity conformation capable of ligand binding. Ligand binding itself induces additional conformational changes resulting in exposure of neoantigenic sites called ligand-induced binding sites (LIBS)3 and generates in turn outside-in signaling, which triggers a range of downstream signals (1, 2).Integrin αIIbβ3 is expressed on platelets and megakaryocytes. In flowing blood under resting conditions, αIIbβ3 does not interact with its ligand fibrinogen. When a blood vessel is damaged, platelets adhere at sites of vascular injury and become activated. As a consequence, αIIbβ3 adopts the high affinity conformation and binds fibrinogen. This results in platelet aggregation and thrombus formation, which eventually will stop the bleeding (3).The topology of integrins comprises an extracellular, globular, N-terminal ligand-binding head domain (the β-propeller domain in the αIIb chain and the βI domain in the β3 chain) standing on two long legs or stalks (consisting of thigh, calf-1, and calf-2 domains in the αIIb chain and hybrid, plexin/semaphorin/integrin (PSI), four integrin endothelial growth factor-like (I-EGF), and β-tail domains in the β3 chain), followed by transmembrane and cytoplasmic domains (1, 2). X-ray crystal structures of the extracellular domain of non-activated αVβ3 revealed that the legs are severely bent, putting the head domain next to the membrane-proximal portions of the legs (4, 5). The bending occurs between I-EGF1 and I-EGF2 in the β-subunit and between the thigh and calf-1 domains in the α-subunit. This bent conformation represents the low affinity state of the receptor. The high affinity state of the receptor is induced by activation and is associated with a large-scale conformational rearrangement in which the integrin extends with a switchblade-like motion (2). Recently, the crystal structure of the entire extracellular domain of αIIbβ3 in its low affinity conformation was resolved and revealed that this integrin also adopts the bent conformation under resting conditions (6). Structural rearrangements in αIIbβ3 between the bent and extended conformations are similar to what has been reported for other integrins (7).We report here that the S527F mutation in the I-EGF3 region of the β3 polypeptide chain of the αIIbβ3 receptor induces a constitutively active receptor adopting an extended high affinity conformation. This was evidenced by spontaneous PAC-1, fibrinogen, and anti-LIBS antibody binding. These data were further corroborated by modeling the replacement of Ser527 with Phe in the crystal structure of the extracellular domain of αIIbβ3. In this model, the S527F mutation decreases the flexibility of I-EGF3 and appears to prevent movement of the lower β-leg into the cleft between the upper β-leg and the lower α-leg. As a consequence, formation of the bent conformation of the non-activated receptor is hampered.  相似文献   

12.
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   

13.
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   

14.
The lack of β1 integrins on chondrocytes leads to severe chondrodysplasia associated with high mortality rate around birth. To assess the impact of β1 integrin-mediated cell-matrix interactions on the function of adult knee joints, we conditionally deleted the β1 integrin gene in early limb mesenchyme using the Prx1-cre transgene. Mutant mice developed short limbed dwarfism and had joint defects due to β1 integrin deficiency in articular regions. The articular cartilage (AC) was structurally disorganized, accompanied by accelerated terminal differentiation, altered shape, and disrupted actin cytoskeleton of the chondrocytes. Defects in chondrocyte proliferation, cytokinesis, and survival resulted in hypocellularity. However, no significant differences in cartilage erosion, in the expression of matrix-degrading proteases, or in the exposure of aggrecan and collagen II cleavage neoepitopes were observed between control and mutant AC. We found no evidence for disturbed activation of MAPKs (ERK1/2, p38, and JNK) in vivo. Furthermore, fibronectin fragment-stimulated ERK activation and MMP-13 expression were indistinguishable in control and mutant femoral head explants. The mutant synovium was hyperplastic and frequently underwent chondrogenic differentiation. β1-null synoviocytes showed increased proliferation and phospho-focal adhesion kinase expression. Taken together, deletion of β1 integrins in the limb bud results in multiple abnormalities of the knee joints; however, it does not accelerate AC destruction, perturb cartilage metabolism, or influence intracellular MAPK signaling pathways.Chondrocytes of the articular cartilage (AC)2 secrete a unique set of extracellular matrix (ECM) molecules that assemble into interactive associates composed of collagens, proteoglycans (PGs), and non-collagenous glycoproteins (1). The fibrillar collagen meshwork supplies cartilage with its tensile strength, whereas the hydrated glycosaminoglycan (GAG) chains of PGs (mainly aggrecan) generate an osmotic swelling pressure that resists compressive forces. In diarthrodial joints, the molecular composition and the physical properties of the cartilage are principal determinants for the shock-absorbing function of articular surfaces upon mechanical loading. During the development of osteoarthritis (OA), an imbalance between anabolic and catabolic processes increases the proteolysis of PGs and collagens (2, 3), which eventually leads to the mechanical weakening of the AC and culminates in its progressive destruction. Physiological and pathological remodeling of the AC ECM is primarily attributed to the activities of matrix metalloproteinases (MMPs) and a disintegrin and metalloproteinase with thrombospondin-like repeat (ADAMTS) proteases (4, 5) and is controlled by the communication between the cells and their environment.An increasing amount of evidence suggests that interactions between chondrocytes and the ECM through the integrin family of heterodimeric (αβ) transmembrane receptors play a central role in cartilage function (6). Integrins connect the pericellular matrix to cytoskeletal and intracellular signaling complexes and modulate various cellular functions, including survival, proliferation, differentiation, and matrix assembly and metabolism (7, 8). Chondrocytes express several integrin receptors for cartilage matrix ligands, such as α1β1, α2β1, and α10β1 (for collagen II); α5β1, αvβ3, and αvβ5 (for fibronectin); and α6β1 (for laminin) (6, 9). We have previously demonstrated that β1fl/fl-Col2a1cre+ mice, in which the floxed β1 integrin gene (β1fl/fl) was deleted using the chondrocyte-specific Col2a1cre transgene, display severe chondrodysplasia and a high mortality rate at birth (10). Homozygous mutant mice exhibit multiple growth plate abnormalities during endochondral bone formation, characterized by defects in chondrocyte adhesion, shape, proliferation, cytokinesis, and actin organization. In addition, the cartilage matrix shows a sparse, distorted collagen network. Similar, but milder abnormalities were found in mice lacking the collagen-binding integrin α10β1 or integrin-linked kinase in cartilage (11, 12).Although these works have identified β1 integrins as essential regulators of growth plate development, the role of integrins in joint morphogenesis, adult joint function, and pathology is incompletely understood. In the embryonic mouse limb culture system, administration of β1 and α5 blocking antibodies or RGD peptides induced ectopic joint formation between proliferating and hypertrophic chondrocytes of the growth plate, suggesting that α5β1 integrin controls the decision between cartilage differentiation and joint formation during development (13). In adult joints, increased immunostaining of β1 integrin was reported in osteoarthritic monkey cartilage compared with normal cartilage (14) and in human OA samples at minimally damaged locations compared with areas with more severe lesions (15). In another study, the neoexpression of α2, α4, and β2 integrins was observed in osteoarthritic human femoral head cartilage (16). In vitro experiments have suggested that signaling through the fibronectin (FN) receptor α5β1 integrin is pivotal to prevent cell death of normal and osteoarthritic human articular chondrocytes (17). FN fragments (FN-fs) present in synovial fluid and cartilage of OA patients have been implicated in cartilage breakdown (1821). Human AC chondrocytes treated with the central, 110–120-kDa cell-binding FN-f but not with intact FN were shown to increase MMP-13 synthesis through the stimulation of α5β1 integrin and the subsequent activation of the proline-rich tyrosine kinase-2 and mitogen-activated protein kinases (MAPKs) ERK-1/2, JNK, and p38 (22, 23). Similarly, an adhesion-blocking antibody against α2β1 integrin induced the phosphorylation of MAPKs in human AC chondrocytes (22). Treatment of cultured rabbit synovial fibroblasts with central FN-fs or activating antibodies against α5β1 integrin elevated MMP-1 and MMP-3 expression (24). Although these experiments suggest that blocking integrin signaling through α2β1/α5β1 in response to degradation fragments may attenuate OA, mice lacking α1β1 integrin are prone to osteoarthritis (25). Knee joints of α1-null mice display precocious PG loss, cartilage erosion associated with increased MMP-2 and MMP-3 expression, and synovial hyperplasia.To further explore the role of β1 integrins in joint biology, here we report the deletion of the floxed β1 integrin gene in embryonic limb bud mesenchymal cells using the Prx1cre transgene (26). β1fl/fl-Prx1cre+ mice were born alive with short limbs due to the lack of β1 integrin heterodimers on chondrocytes. We found that β1 integrin deficiency in knee joints leads to multiple abnormalities of the AC and the synovium, but it is not associated with accelerated AC destruction, perturbed AC metabolism, and MAPK signaling. Our data suggest that β1 integrins are required for the proper structural organization of the AC by anchoring chondrocytes to the ECM, but signaling through β1 integrins is less important for normal cartilage homeostasis.  相似文献   

15.
The novel α1D L-type Ca2+ channel is expressed in supraventricular tissue and has been implicated in the pacemaker activity of the heart and in atrial fibrillation. We recently demonstrated that PKA activation led to increased α1D Ca2+ channel activity in tsA201 cells by phosphorylation of the channel protein. Here we sought to identify the phosphorylated PKA consensus sites on the α1 subunit of the α1D Ca2+ channel by generating GST fusion proteins of the intracellular loops, N terminus, proximal and distal C termini of the α1 subunit of α1D Ca2+ channel. An in vitro PKA kinase assay was performed for the GST fusion proteins, and their phosphorylation was assessed by Western blotting using either anti-PKA substrate or anti-phosphoserine antibodies. Western blotting showed that the N terminus and C terminus were phosphorylated. Serines 1743 and 1816, two PKA consensus sites, were phosphorylated by PKA and identified by mass spectrometry. Site directed mutagenesis and patch clamp studies revealed that serines 1743 and 1816 were major functional PKA consensus sites. Altogether, biochemical and functional data revealed that serines 1743 and 1816 are major functional PKA consensus sites on the α1 subunit of α1D Ca2+ channel. These novel findings provide new insights into the autonomic regulation of the α1D Ca2+ channel in the heart.L-type Ca2+ channels are essential for the generation of normal cardiac rhythm, for induction of rhythm propagation through the atrioventricular node and for the contraction of the atrial and ventricular muscles (15). L-type Ca2+ channel is a multisubunit complex including α1, β and α2/δ subunits (57). The α1 subunit contains the voltage sensor, the selectivity filter, the ion conduction pore, and the binding sites for all known Ca2+ channel blockers (69). While α1C Ca2+ channel is expressed in the atria and ventricles of the heart (1013), expression of α1D Ca2+ channel is restricted to the sinoatrial (SA)2 and atrioventricular (AV) nodes, as well as in the atria, but not in the adult ventricles (2, 3, 10).Only recently it has been realized that α1D along with α1C Ca2+ channels contribute to L-type Ca2+ current (ICa-L) and they both play important but unique roles in the physiology/pathophysiology of the heart (69). Compared with α1C, α1D L-type Ca2+ channel activates at a more negative voltage range and shows slower current inactivation during depolarization (14, 15). These properties may allow α1D Ca2+ channel to play critical roles in SA and AV nodes function. Indeed, α1D Ca2+ channel knock-out mice exhibit significant SA dysfunction and various degrees of AV block (12, 1619).The modulation of α1C Ca2+ channel by cAMP-dependent PKA phosphorylation has been extensively studied, and the C terminus of α1 was identified as the site of the modulation (2022). Our group was the first to report that 8-bromo-cAMP (8-Br-cAMP), a membrane-permeable cAMP analog, increased α1D Ca2+ channel activity using patch clamp studies (2). However, very little is known about potential PKA phosphorylation consensus motifs on the α1D Ca2+ channel. We therefore hypothesized that the C terminus of the α1 subunit of the α1D Ca2+ channel mediates its modulation by cAMP-dependent PKA pathway.  相似文献   

16.
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   

17.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   

18.
ATP synthase uses a unique rotational mechanism to convert chemical energy into mechanical energy and back into chemical energy. The helix-turn-helix motif, termed “DELSEED-loop,” in the C-terminal domain of the β subunit was suggested to be involved in coupling between catalysis and rotation. Here, the role of the DELSEED-loop was investigated by functional analysis of mutants of Bacillus PS3 ATP synthase that had 3–7 amino acids within the loop deleted. All mutants were able to catalyze ATP hydrolysis, some at rates several times higher than the wild-type enzyme. In most cases ATP hydrolysis in membrane vesicles generated a transmembrane proton gradient, indicating that hydrolysis occurred via the normal rotational mechanism. Except for two mutants that showed low activity and low abundance in the membrane preparations, the deletion mutants were able to catalyze ATP synthesis. In general, the mutants seemed less well coupled than the wild-type enzyme, to a varying degree. Arrhenius analysis demonstrated that in the mutants fewer bonds had to be rearranged during the rate-limiting catalytic step; the extent of this effect was dependent on the size of the deletion. The results support the idea of a significant involvement of the DELSEED-loop in mechanochemical coupling in ATP synthase. In addition, for two deletion mutants it was possible to prepare an α3β3γ subcomplex and measure nucleotide binding to the catalytic sites. Interestingly, both mutants showed a severely reduced affinity for MgATP at the high affinity site.F1F0-ATP synthase catalyzes the final step of oxidative phosphorylation and photophosphorylation, the synthesis of ATP from ADP and inorganic phosphate. F1F0-ATP synthase consists of the membrane-embedded F0 subcomplex, with, in most bacteria, a subunit composition of ab2c10, and the peripheral F1 subcomplex, with a subunit composition of α3β3γδε. The energy necessary for ATP synthesis is derived from an electrochemical transmembrane proton (or, in some organisms, a sodium ion) gradient. Proton flow down the gradient through F0 is coupled to ATP synthesis on F1 by a unique rotary mechanism. The protons flow through (half) channels at the interface of the a and c subunits, which drives rotation of the ring of c subunits. The c10 ring, together with F1 subunits γ and ε, forms the rotor. Rotation of γ leads to conformational changes in the catalytic nucleotide binding sites on the β subunits, where ADP and Pi are bound. The conformational changes result in the formation and release of ATP. Thus, ATP synthase converts electrochemical energy, the proton gradient, into mechanical energy in the form of subunit rotation and back into chemical energy as ATP. In bacteria, under certain physiological conditions, the process runs in reverse. ATP is hydrolyzed to generate a transmembrane proton gradient, which the bacterium requires for such functions as nutrient import and locomotion (for reviews, see Refs. 16).F1 (or F1-ATPase) has three catalytic nucleotide binding sites located on the β subunits at the interface to the adjacent α subunit. The catalytic sites have pronounced differences in their nucleotide binding affinity. During rotational catalysis, the sites switch their affinities in a synchronized manner; the position of γ determines which catalytic site is the high affinity site (Kd1 in the nanomolar range), which site is the medium affinity site (Kd2 ≈ 1 μm), and which site is the low affinity site (Kd3 ≈ 30–100 μm; see Refs. 7 and 8). In the original crystal structure of bovine mitochondrial F1 (9), one of the three catalytic sites, was filled with the ATP analog AMP-PNP,2 a second was filled with ADP (plus azide) (see Ref. 10), and the third site was empty. Hence, the β subunits are referred to as βTP, βDP, and βE. The occupied β subunits, βTP and βDP, were in a closed conformation, and the empty βE subunit was in an open conformation. The main difference between these two conformations is found in the C-terminal domain. Here, the “DELSEED-loop,” a helix-turn-helix structure containing the conserved DELSEED motif, is in an “up” position when the catalytic site on the respective β subunit is filled with nucleotide and in a “down” position when the site is empty (Fig. 1A). When all three catalytic sites are occupied by nucleotide, the previously open βE subunit assumes an intermediate, half-closed (βHC) conformation. It cannot close completely because of steric clashes with γ (11).Open in a separate windowFIGURE 1.The βDELSEED-loop. A, interaction of the βTP and βE subunits with theγ subunit.β subunits are shown in yellow andγ in blue. The DELSEED-loop (shown in orange, with the DELSEED motif itself in green)of βTP interacts with the C-terminal helixγ and the short helix that runs nearly perpendicular to the rotation axis. The DELSEED-loop of βE makes contact with the convex portion of γ, formed mainly by the N-terminal helix. A nucleotide molecule (shown in stick representation) occupies the catalytic site of βTP, and the subunit is in the closed conformation. The catalytic site on βE is empty, and the subunit is in the open conformation. This figure is based on Protein Data Bank file 1e79 (32). B, deletions in the βDELSEED-loop. The loop was “mutated” in silico to represent the PS3 ATP synthase. The 3–4-residue segments that are removed in the deletion mutants are color-coded as follows: 380LQDI383, pink; 384IAIL387, green; 388GMDE391, yellow; 392LSD394, cyan; 395EDKL398, orange; 399VVHR402, blue. Residues that are the most involved in contacts with γ are labeled. All figures were generated using the program PyMOL (DeLano Scientific, San Carlos, CA).The DELSEED-loop of each of the three β subunits makes contact with the γ subunit. In some cases, these contacts consist of hydrogen bonds or salt bridges between the negatively charged residues of the DELSEED motif and positively charged residues on γ. The interactions of the DELSEED-loop with γ, its movement during catalysis, the conservation of the DELSEED motif (see 1214). Thus, the finding that an AALSAAA mutant in the α3β3γ complex of ATP synthase from the thermophilic Bacillus PS3, where several hydrogen bonds/salt bridges to γ are removed simultaneously, could drive rotation of γ with the same torque as the wild-type enzyme (14) came as a surprise. On the other hand, it seems possible that it is the bulk of the DELSEED-loop, more so than individual interactions, that drives rotation of γ. According to a model favored by several authors (6, 15, 16) (see also Refs. 1719), binding of ATP (or, more precisely, MgATP) to the low affinity catalytic site on βE and the subsequent closure of this site, accompanied by its conversion into the high affinity site, are responsible for driving the large (80–90°) rotation substep during ATP hydrolysis, with the DELSEED-loop acting as a “pushrod.” A recent molecular dynamics (20) study supports this model and implicates mainly the region around several hydrophobic residues upstream of the DELSEED motif (specifically βI386 and βL387)3 as being responsible for making contact with γ during the large rotation substep.

TABLE 1

Conservation of residues in the DELSEED-loop Amino acids found in selected species in the turn region of the DELSEED-loop. Listed are all positions subjected to deletions in the present study. Residue numbers refer to the PS3 enzyme. Consensus annotation: p, polar residue; s, small residue; h, hydrophobic residue; –, negatively charged residue; +, positively charged residue.Open in a separate windowIn the present study, we investigated the function of the DELSEED-loop using an approach less focused on individual residues, by deleting stretches of 3–7 amino acids between positions β380 and β402 of ATP synthase from the thermophilic Bacillus PS3. We analyzed the functional properties of the deletion mutants after expression in Escherichia coli. The mutants showed ATPase activities, which were in some cases surprisingly high, severalfold higher than the activity of the wild-type control. On the other hand, in all cases where ATP synthesis could be measured, the rates where below or equal to those of the wild-type enzyme. In Arrhenius plots, the hydrolysis rates of the mutants were less temperature-dependent than those of wild-type ATP synthase. In those cases where nucleotide binding to the catalytic sites could be tested, the deletion mutants had a much reduced affinity for MgATP at high affinity site 1. The functional role of the DELSEED-loop will be discussed in light of the new information.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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