首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
4.
5.
PTG and GL are hepatic protein phosphatase-1 (PP1) glycogen-targeting subunits, which direct PP1 activity against glycogen synthase (GS) and/or phosphorylase (GP). The C-terminal 16 amino residues of GL comprise a high affinity binding site for GP that regulates bound PP1 activity against GS. In this study, a truncated GL construct lacking the GP-binding site (GLtr) and a chimeric PTG molecule containing the C-terminal site (PTG-GL) were generated. As expected, GP binding to glutathione S-transferase (GST)-GLtr was reduced, whereas GP binding to GST-PTG-GL was increased 2- to 3-fold versus GST-PTG. In contrast, PP1 binding to all proteins was equivalent. Primary mouse hepatocytes were infected with adenoviral constructs for each subunit, and their effects on glycogen metabolism were investigated. GLtr expression was more effective at promoting GP inactivation, GS activation, and glycogen accumulation than GL. Removal of the regulatory GP-binding site from GLtr completely blocked the inactivation of GS seen in GL-expressing cells following a drop in extracellular glucose. As a result, GLtr expression prevented glycogen mobilization under 5 mm glucose conditions. In contrast, equivalent overexpression of PTG or PTG-GL caused a similar increase in glycogen-targeted PP1 levels and GS dephosphorylation. Surprisingly, GP dephosphorylation was significantly reduced in PTG-GL-overexpressing cells. As a result, PTG-GL expression permitted glycogenolysis under 5 mm glucose conditions that was prevented in PTG-expressing cells. Thus, expression of constructs that contained the high affinity GP-binding site (GL and PTG-GL) displayed reduced glycogen accumulation and enhanced glycogenolysis compared with their respective controls, albeit via different mechanisms.Hepatic glycogen metabolism plays a central role in the maintenance of circulating plasma glucose levels under various physiological conditions. The rate-controlling enzymes in glycogen metabolism, glycogen synthase (GS)2 and glycogen phosphorylase (GP), are subject to multiple levels of regulation, including allosteric binding of activators and inhibitors, protein phosphorylation, and changes in subcellular localization. GS is phosphorylated on up to 9 residues by a variety of kinases, although site 2 appears to be the most important regulator of hepatic GS (1). In contrast, GP is phosphorylated on a single N-terminal serine residue by phosphorylase kinase, which increases GP activity and its sensitivity to allosteric activators. Both GS and GP are in turn also regulated by protein phosphatases, most notably PP1. Although PP1 is a cytosolic protein, a family of five molecules has been reported that targets the enzyme to glycogen particles (27), whereas another two glycogen-targeting subunits have been putatively identified based on sequence homology (8). Published work has indicated that each targeting subunit confers differential regulation of PP1 activity by extracellular hormonal signals and/or intracellular changes in metabolites (911).Four PP1-glycogen-targeting proteins are expressed in rodent liver, although GL and PTG/R5 have been most extensively studied (9, 1215). GL is present at higher levels in rat liver than PTG (12), but the expression of both proteins is subject to coordinate regulation by fasting/refeeding and insulin (12, 13). Previous studies indicated that the PTG-PP1 complex is primarily responsible for GP dephosphorylation and regulation of glycogenolysis (13, 16), whereas the GL-PP1 complex preferentially mediates the activation of GS upon elevation of extracellular glucose (9, 13). However, the molecular mechanisms underlying these differential properties of PTG and GL have not been completely defined.Both PTG and GL directly bind to specific PP1 substrates involved in glycogen metabolism, albeit for different physiological reasons. The extreme C-terminal 16 amino acids of GL comprises a unique, high affinity binding site for phosphorylated GP (GPa (17)), which has been further delineated to two critical tyrosine residues (18, 37). Interaction of PP1 with GL reduces phosphatase activity against GPa (3). In turn, GPa binding to the GL-PP1 complex potently inhibits phosphatase activity against GS in vitro (3, 19) and regulates glycogen-targeted PP1 activity in liver cells and extracts (2022). PTG contains a single substrate-binding site that interacts with GS and GP (5, 23). In contrast to the regulatory role of the GPa binding to GL, interaction of substrates with PTG increases PP1 activity against these proteins (24). Indeed, disruption of the substrate-binding site by point mutagenesis abrogated the ability of mutant PTG expression to increase cellular glycogen levels (23), indicating an important role for substrate binding to the PTG-PP1 complex.Previous work has comprehensively compared the metabolic impact of PTG versus GL overexpression in hepatocytes and thus was not the goal of this study (9, 10). Instead, two novel PP1 targeting constructs were generated in which the high affinity GPa-binding site was removed from GL or added to the C terminus of PTG. The effects of expressing wild-type and mutant constructs on GS and GP activities and on the regulation of glycogen metabolism by extracellular glucose were investigated using primary mouse hepatocytes.  相似文献   

6.
7.
Cross-talk of BMP and Wnt signaling pathways has been implicated in many aspects of biological events during embryogenesis and in adulthood. A secreted protein Wise and its orthologs (Sostdc1, USAG-1, and Ectodin) have been shown to modulate Wnt signaling and also inhibit BMP signals. Modulation of Wnt signaling activity by Wise is brought about by an interaction with the Wnt co-receptor LRP6, whereas BMP inhibition is by binding to BMP ligands. Here we have investigated the mode of action of Wise on Wnt and BMP signals. It was found that Wise binds LRP6 through one of three loops formed by the cystine knot. The Wise deletion construct lacking the LRP6-interacting loop domain nevertheless binds BMP4 and inhibits BMP signals. Moreover, BMP4 does not interfere with Wise-LRP6 binding, suggesting separate domains for the physical interaction. Functional assays also show that the ability of Wise to block Wnt1 activity through LRP6 is not impeded by BMP4. In contrast, the ability of Wise to inhibit BMP4 is prevented by additional LRP6, implying a preference of Wise in binding LRP6 over BMP4. In addition to the interaction of Wise with BMP4 and LRP6, the molecular characteristics of Wise, such as glycosylation and association with heparan sulfate proteoglycans on the cell surface, are suggested. This study helps to understand the multiple functions of Wise at the molecular level and suggests a possible role for Wise in balancing Wnt and BMP signals.Wise is a secreted protein that was isolated from a functional screen of a chick cDNA library of embryonic tissues. It was identified as being able to alter the antero-posterior character of neuralized Xenopus animal caps by promoting activity of the Wnt pathway (1). Independently, the homologous protein was isolated from a functional screen to detect genes that are preferentially expressed in the rat endometrium, which had been maximally sensitized to implantation, and named USAG-1 (uterine sensitization-associated gene-1) (2). The protein was identified a third time from the GenBankTM sequence data base of mouse as a putative secreted protein, shown to be a BMP antagonist, and named Ectodin (3). The gene has also been called Sostdc1 (Sclerostin domain-containing 1) or Sostl (Sclerostin-like) due to the homology with Sclerostin-encoding gene Sost (4, 5). USAG-1/Wise/Ectodin/Sostdc1 is expressed in various tissues, such as the surface ectoderm of the posterior axis (1, 6), branchial arches (3, 6), the dermal papilla in hair follicles (7), vibrissae (3), mammalian tooth cusps (3, 8), rat endometrium (2), developing testis (911), interdigital tissues (12), and embryonic and adult kidneys (13, 14).Wise appears to have a dual role in modulating the Wnt pathway. Injection of Wnt8 RNA into a ventral vegetal blastomere of Xenopus embryos at the four-cell stage induces a full secondary axis to form, and this is blocked by the addition of Wise RNA as well as other Wnt inhibitors (1). Activation of the Wnt/β-catenin pathway in hair follicles triggers regeneration of hair growth, and expression of Wise appears to have a defined role to inhibit this (15). In this context, Wise expression is repressed by the nuclear receptor co-repressor, Hairless, which results in activation of the Wnt pathway; thus, a model of periodic regeneration of hair follicles has been proposed (15, 16). In addition, Wise and its homologue USAG-1 have been shown to block Wnt1, Wnt3a, and Wnt10b activities in reporter assays (14, 15, 17). Wise was found to bind to the Wnt co-receptor, LRP6, sharing the binding domain with Wnt ligands. Importantly, Wise was found to compete with Wnt8 for binding to LRP6, therefore suggesting a mechanism for inhibition of the Wnt pathway whereby Wise blocks the binding of ligand and receptor (1). Wise may also be retained in the endoplasmic reticulum and inhibit the trafficking of LRP6 to the cell surface (18). Wise also binds LRP4 (19), a member of the LRP family functioning inhibitory to Wnt signals (20). It is noteworthy that Wise was isolated from a screen designed to detect the activation of the Wnt/β-catenin pathway, not inhibition. The exact mechanism of how Wise exerts such a context-dependent modulation on the Wnt pathway is yet to be clarified.Osteoblast differentiation of MC3T3-E1 cells, as measured by alkaline phosphatase activity, can be induced by a wide range of BMP molecules. In this assay, Ectodin, the mouse ortholog of Wise, was shown to inhibit differentiation induced by BMP2, -4, -6, or -7 in a dose-dependent manner (3). Similarly, Ectodin (also known as USAG-1) was also found to inhibit the bone differentiation induced by BMP2, -4, or -7 in C2C12 cells (14). Ectodin also inhibits BMP2- or BMP7-induced Msx2 expression in dissected mouse tooth buds in organ culture (3). In tooth buds, Ectodin expression is detected in the dental ectoderm and mesenchymal cells excluding from the enamel knot (3). Ectodin/USAG-1-deficient mice created by targeted-disruption show altered tooth morphology and extra teeth, indicating that Ectodin and BMP tightly control tooth development and patterning in mammals (8, 2123). Furthermore, in mouse adult kidneys, the ability of BMP7 to repair established renal injury is blocked by USAG-1 (13). All of these findings indicate that USAG-1/Wise/Ectodin has a clear antagonistic effect on BMP signaling, where it binds BMP2, -4, -6, and -7 (3, 14) and presumably prevents BMP binding to its receptors.Analysis of the sequence of Wise reveals that it has the C1XnC2XGXC3XnC4XnC5XC6 motif of a six-membered cystine knot, where C1 forms a disulfide bond with C4, C2 with C5, and C3 with C6 (for a review of the cystine knot, see Refs. 2427). This arrangement results in a globular protein with three loops, “finger 1,” “heel,” and “finger 2,” held together with an eight-membered ring of C2XGXC3C6XC5C2 (Fig. 1). BMP antagonists represent a subfamily in the cystine knot superfamily, and this is further subdivided into three subfamilies based on the size of the cystine knot. These are the CAN family (eight-membered ring), Twisted Gastrulation (nine-membered ring), and Chordin and Noggin (10-membered ring) (27). There is generally little sequence homology between family members in the heel, finger 1, and finger 2 regions, yet Wise does show a moderate homology with Sclerostin (28). Sclerostin is involved in regulating bone mass (4, 5) and also appears to antagonize both Wnt (2932) and BMP (28, 33, 34) signals. This paper aims to analyze the dual role of Wise on Wnt and BMP pathways by probing the structural features of the protein and reconciling them to physiological properties. It also aims to reveal the molecular nature of the protein in view of possible glycosylation, secretion, and association with extracellular matrix.Open in a separate windowFIGURE 1.Structure of chick Wise protein. A, stereo ribbon representation of the chick Wise three-dimensional structural model (residues 68–186). Purple, β-strands; green, loop regions. Yellow, disulfide bonds in the cystine knot plus a further disulfide (cysteines 89 and 147) linking two fingers of the structure. N- and C-terminal ends are indicated. B, schematic drawing of the full-length chick Wise structure. Arrowhead, the predicted signal sequence cleavage site for secretion; black dot, asparagine at position 47 (N47), the glycosylated site revealed in this study. Six cysteine residues forming the “cystine knot” are shown in circles, and disulfide bonds for the knot formation are shown by dotted lines. Three loops (Finger 1, Heel, and Finger 2) are indicated. The scheme also shows the deleted parts of Wise constructs ΔN, Δheel, and ΔC.  相似文献   

8.
YVH1 was one of the first eukaryotic dual specificity phosphatases cloned, and orthologues posses a unique C-terminal zinc-coordinating domain in addition to a cysteine-based phosphatase domain. Our recent results revealed that human YVH1 (hYVH1) protects cells from oxidative stress. This function requires phosphatase activity and the zinc binding domain. This current study provides evidence that the thiol-rich zinc-coordinating domain may act as a redox sensor to impede the active site cysteine from inactivating oxidation. Furthermore, using differential thiol labeling and mass spectrometry, it was determined that hYVH1 forms intramolecular disulfide bonds at the catalytic cleft as well as within the zinc binding domain to avoid irreversible inactivation during severe oxidative stress. Importantly, zinc ejection is readily reversible and required for hYVH1 activity upon returning to favorable conditions. This inimitable mechanism provides a means for hYVH1 to remain functionally responsive for protecting cells during oxidative stimuli.Human YVH1 (hYVH12; also known as DUSP12) is a member of the dual specificity phosphatase (DUSP) subfamily of protein-tyrosine phosphatases (PTPs) (1, 2). It is constructed of an N-terminal DUSP catalytic domain and a unique C-terminal zinc coordinating domain (3). Poor characterization and lack of mitogen-activated protein kinase targeting motifs further classify this enzyme as an atypical DUSP (1). YVH1 orthologues exhibit high evolutionary conservation and similar domain organization (3). Deletion of the yvh1 gene in yeast disrupts normal growth processes (4), whereas insertion and expression of the hyvh1 gene is capable of restoring a normal yeast growth phenotype (3). Amplification of the dusp12/hyvh1 gene has been reported in multiple sarcomas, implicating a role for hYVH1 in human disease (57).Recently, deletion studies from our laboratory have shown that the C-terminal zinc binding domain of hYVH1 is not essential for intrinsic phosphatase activity in vitro; however, it is required for interaction with the ATPase domain of heat shock protein 70 (8). Similarly, overexpression of wild type hYVH1 but not catalytically dead or zinc coordinating domain deletion mutants prevents cell death induced by Fas receptor activation, heat shock, and hydrogen peroxide (H2O2) (8). Despite these findings, current information on hYVH1 enzymatic and physiological functions remains limited.PTPs and DUSPs share similar active site architecture and catalytic mechanism, characterized by the conserved HCX5R(S/T) motif (9, 10). The unique microenvironment within the HCX5R(S/T) motif reduces the pKa value of the active site cysteine, enhancing both nucleophilicity and oxidation susceptibility (11, 12). Stimulated or constituent generation of ROS can result in oxidative second messenger signaling responses capable of transient and reversible post-translational inactivation of both PTPs and DUSPs through oxidation of the catalytic cysteine (1315).This oxidative susceptibility and modification varies among PTPs and DUSPs, a likely consequence of slight variations in active site conformations or mediated through unique regulatory domains (1618). Accumulating evidence suggests that redox-mediated oxidation of PTPs is a dynamic modification that can differentially regulate PTPs (13, 19). Sulfenic acid, cyclic sulfenamide, and disulfide bond formation have all been shown to facilitate stable, reversible active site modifications among various PTPs and DUSPs (12, 14, 20). Furthermore, evidence suggests that oxidation predominantly and rapidly targets the active site cysteine, whereas other cysteinyl residues remain in the reduced state (15, 20).This study investigated the relationship between the zinc-coordinating C-terminal domain and the catalytic domain of hYVH1 during oxidative conditions. We provide data suggesting that the zinc binding domain can serve as a reducing agent during oxidative stress to impede the oxidation of the active site cysteine. Increased exposure to oxidative conditions readily induces disulfide bond formation within the zinc-coordinating and catalytic domains, resulting in concomitant zinc ejection and enzymatic inactivation. Zinc ejection is readily reversible and required for hYVH1 activity upon returning to reducing conditions. Thus, we propose a mechanism for phosphatase active site protection through the intrinsic redox buffering capacity of this unique zinc binding domain.  相似文献   

9.
10.
Sphingosine 1-phosphate (S1P) is a bioactive lipid signal transmitter present in blood. Blood plasma S1P is supplied from erythrocytes and plays an important role in lymphocyte egress from lymphoid organs. However, the S1P export mechanism from erythrocytes to blood plasma is not well defined. To elucidate the mechanism of S1P export from erythrocytes, we performed the enzymatic characterization of S1P transporter in rat erythrocytes. Rat erythrocytes constitutively released S1P without any stimulus. The S1P release was reduced by an ABCA1 transporter inhibitor, glyburide, but not by a multidrug resistance-associated protein inhibitor, MK571, or a multidrug resistance protein inhibitor, cyclosporine A. Furthermore, we measured S1P transport activity using rat erythrocyte inside-out membrane vesicles (IOVs). Although the effective S1P transport into IOVs was observed in the presence of ATP, this activity was also supported by dATP and adenosine 5′-(β,γ-imido)triphosphate. The rate of S1P transport increased depending on S1P concentration, with an apparent Km value of 21 μm. Two phosphorylated sphingolipids, dihydrosphingosine 1-phosphate and ceramide 1-phosphate, did not inhibit S1P transport. Similar to the intact erythrocytes, the uptake of S1P into IOVs was inhibited by glyburide and vanadate but not by the other ABC transporter inhibitors. These results suggest that S1P is exported from the erythrocytes by a novel ATP-dependent transporter.Sphingosine 1-phosphate (S1P),2 a bioactive lipid molecule present in the blood, plays an important role in diverse cellular responses, such as migration, proliferation, and differentiation (1, 2). These processes are triggered by the binding of S1P to its specific receptors (3), of which five subtypes (S1P1-S1P5) have been identified in endothelial and immune cells (4). Studies using S1P1 receptor-deficient mice showed abnormalities in lymphocyte egress from lymph nodes, spleen, and thymus (5, 6). Whereas blood plasma contains a basal level of S1P from the nanomolar to the micromolar range (712), lymphoid tissues maintain a low S1P environment through the activity of S1P lyase (13). It has been proposed that a higher concentration of S1P in the blood plasma than in the lymphoid organs establishes an essential gradient along which lymphocytes expressing the S1P1 receptor on cell surfaces migrate (2, 5, 6, 1315).The source of plasma S1P remains unclear despite its importance in the cellular responses of endothelial cells and lymphocytes. Unlike most cells, blood cells, astrocytes, and vascular endothelial cells are reported to release S1P (8, 1618). These cells contain sphingosine kinase, which synthesizes S1P through the phosphorylation of sphingosine (16, 18, 19). Whereas platelets and mast cells release S1P in a stimulus-dependent manner (17, 20), erythrocytes, neutrophils, and mononuclear cells release S1P in a stimulus-independent manner (16). The roles of S1P derived from erythrocytes, the most abundant of these blood cells, have not been elucidated. However, recent reports suggest that S1P released from erythrocytes is a major source of plasma S1P (7, 9) and promotes lymphocyte egress to blood (9).Previously, we showed that S1P is released from rat platelets upon stimulation by thrombin or Ca2+ (21). We proposed that an ATP-dependent transporter plays a key role in S1P release from platelets (21). However, the detailed mechanism of S1P release is unclear because there is no way to assay the transport of S1P across the membrane. In this study we compared the properties of S1P release from erythrocytes with that of platelets and showed that S1P release from erythrocytes does not require any stimuli. We then established an assay to measure the ATP-dependent S1P uptake into inside-out membrane vesicles (IOVs) prepared from rat erythrocytes and characterized S1P transport in erythrocytes.  相似文献   

11.
12.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

13.
Membrane trafficking is regulated in part by small GTP-binding proteins of the ADP-ribosylation factor (Arf) family. Arf function depends on the controlled exchange and hydrolysis of GTP. We have purified and cloned two variants of a 130-kDa phosphatidylinositol 4,5-biphosphate (PIP2)-dependent Arf1 GTPase-activating protein (GAP), which we call ASAP1a and ASAP1b. Both contain a pleckstrin homology (PH) domain, a zinc finger similar to that found in another Arf GAP, three ankyrin (ANK) repeats, a proline-rich region with alternative splicing and SH3 binding motifs, eight repeats of the sequence E/DLPPKP, and an SH3 domain. Together, the PH, zinc finger, and ANK repeat regions possess PIP2-dependent GAP activity on Arf1 and Arf5, less activity on Arf6, and no detectable activity on Arl2 in vitro. The cDNA for ASAP1 was independently identified in a screen for proteins that interact with the SH3 domain of the tyrosine kinase Src. ASAP1 associates in vitro with the SH3 domains of Src family members and with the Crk adapter protein. ASAP1 coprecipitates with Src from cell lysates and is phosphorylated on tyrosine residues in cells expressing activated Src. Both coimmunoprecipitation and tyrosine phosphorylation depend on the same proline-rich class II Src SH3 binding site required for in vitro association. By directly interacting with both Arfs and tyrosine kinases involved in regulating cell growth and cytoskeletal organization, ASAP1 could coordinate membrane remodeling events with these processes.Membrane traffic, the transfer of material between membrane-bound compartments, is needed for such diverse cellular processes as secretion, endocytosis, and changes in cell shape that accompany cell growth, division, and migration (reviewed in references 84, 85, and 87). It is mediated by transport vesicles that are formed by budding from a donor membrane. The process of budding is driven by the assembly of a proteinaceous coat. Once the vesicle is formed, the coat must dissociate to permit fusion with an acceptor membrane and the consequent delivery of the vesicle’s contents. These steps are regulated in part by the Arf family of small GTP-binding proteins (reviewed in references 8, 23, 61, and 63). Arfs are highly conserved and are found in eukaryotes ranging from yeast to humans. The mammalian Arf family is divided into several classes based largely on sequence similarity: class I (Arfs 1 through 3), class II (Arfs 4 and 5), class III (Arf6), and the more distantly related Arf-like (Arl) class. By linking GTP binding and hydrolysis to coat assembly and disassembly, Arfs regulate membrane trafficking at a number of sites. Arf1 has been implicated in endoplasmic reticulum-to-Golgi and intra-Golgi transport, endosome-to-endosome fusion, and synaptic vesicle formation (8, 23, 28, 61, 63, 66). Arf6 has been implicated in regulation of membrane traffic between the plasma membrane and a specialized endocytic compartment, and its function has been linked to cytoskeletal reorganization (25, 26, 71, 73, 74). The specific sites of action of the other Arf family members are not known.The hydrolysis of GTP on Arf requires a GTPase-activating protein (GAP) (19, 61). With multiple Arfs and multiple sites of action, the existence of several unique Arf GAPs had been anticipated. A number of activities have been purified or partially purified from mammalian sources, including rat liver (19, 57, 77), rat spleen (21), and bovine brain (79), and two Arf GAP activities from rat liver have been resolved (77). They have similar Arf specificities but differ in their lipid dependencies. One of the Arf GAPs (ArfGAP/ArfGAP1, hereafter referred to as ArfGAP1) which functions in the Golgi is activated by dioleoglycerols (3, 4, 19, 40). ArfGAP1, in common with a yeast Arf GAP, GCS1 (72), contains a zinc finger domain which is required for activity (19). The second Arf GAP (ArfGAP2) is specifically activated by phosphatidylinositol 4,5-bisphosphate (PIP2) and phosphatidic acid (PA). Based on lipid requirements, ArfGAP2 was speculated to function at the plasma membrane and be regulated independently of ArfGAP1 (77). ArfGAP1 and ArfGAP2 were antigenically distinct and, therefore, likely to be distinct gene products; however, prior to this study, only ArfGAP1 had been cloned (19).Src, a cytoplasmic tyrosine kinase with N-terminal Src homology 3 (SH3) and SH2 domains, transduces signals important for cell growth and cytoskeletal organization (12, 68, 91). A number of studies suggest that Src is also involved in regulating membrane traffic. Src associates primarily with endosomal membranes and in several cell types has been localized to specialized secretory vesicles, including synaptic vesicles (5, 20, 34, 46, 54, 69, 81). Overexpression of Src accelerates endocytosis (95). In addition, Src associates with or phosphorylates several proteins involved in membrane trafficking (5, 31, 43, 65).Here, we report the purification and cloning of a PIP2-dependent Arf GAP, ASAP1. ASAP1 contains a zinc finger domain similar to that required for GAP activity in ArfGAP1 and GCS1. ASAP1 also contains a number of domains that are likely to be involved in regulation and/or localization: a pleckstrin homology (PH) domain, three ankyrin (ANK) repeats, a proline-rich region with SH3 binding motifs, and an SH3 domain. In addition, ASAP1 was identified independently as a binding protein for Src and was found to be phosphorylated on tyrosine in cells that express activated Src. ASAP1 also associated with the adapter protein c-Crk in vitro. ASAP1 was localized to the cytoplasm and the cell edge likely associated with the plasma membrane. We propose that ASAP1, by binding both Src and PIP2, could coordinate membrane trafficking with cell growth or actin cytoskeleton remodeling.  相似文献   

14.
15.
16.
17.
18.
Maspin is a serpin that has multiple effects on cell behavior, including inhibition of migration. How maspin mediates these diverse effects remains unclear, as it is devoid of protease inhibitory activity. We have previously shown that maspin rapidly inhibits the migration of vascular smooth muscle cells (VSMC), suggesting the involvement of direct interactions with cell surface proteins. Here, using immunofluorescence microscopy, we demonstrate that maspin binds specifically to the surface of VSMC in the dedifferentiated, but not the differentiated, phenotype. Ligand blotting of VSMC lysates revealed the presence of several maspin-binding proteins, with a protein of 150 kDa differentially expressed between the two VSMC phenotypes. Western blotting suggested that this protein was the β1 integrin subunit, and subsequently both α3β1 and α5β1, but not αvβ3, were shown to associate with maspin by coimmunoprecipitation. Specific binding of these integrins was also observed using maspin-affinity chromatography, using HT1080 cell lysates. Direct binding of maspin to α5β1 was confirmed using a recombinant α5β1-Fc fusion protein. Using conformation-dependent anti-β1 antibodies, maspin binding to VSMC was found to lead to a decrease in the activation status of the integrin. The functional involvement of α5β1 in mediating the effect of maspin was established by the inhibition of migration of CHO cells overexpressing human α5 integrin, but not those lacking α5 expression. Our observations suggest that maspin engages in specific interactions with a limited number of integrins on VSMC, leading to their inactivation, and that these interactions are responsible for the effects of maspin in the pericellular environment.Maspin is a member of the serpin family of serine protease inhibitors (SERPINB5).2 It was originally identified as a gene down-regulated in invasive breast cancer and proposed as a class II tumor suppressor (1), and has since been shown to have many effects on cellular behavior that are consistent with this activity. It has been shown to decrease the proliferation, migration, and metastasis of tumor cells in vivo (1, 2) and their invasion in vitro (3, 4), and to increase apoptosis of endothelial cells (5) and inhibit angiogenesis (6). However, the cellular effects of maspin are not restricted to tumor cells, and we have demonstrated that maspin can inhibit the migration of vascular smooth muscle cells (7).VSMC migration is a key event in the development of atherosclerosis (8), and contributes significantly to restenosis after angioplasty (9) and transplant arteriosclerosis (10). VSMC are not terminally differentiated and acquire migratory capacity as part of a phenotypic switch from a contractile, quiescent state to a dedifferentiated phenotype, characterized by proliferation and increased extracellular matrix synthesis, in addition to motility (11). This allows VSMC to respond to environmental cues following vascular injury. The phenotypic plasticity of VSMC is regulated by an array of signals, among which integrin-mediated association with surrounding extracellular matrix and changes in the expression of matrix-degrading proteases are prominent (1214).How maspin mediates its various cellular effects is unclear. Maspin has been reported to be an inhibitor of plasminogen activation (3, 15, 16), but we have shown that maspin is unable to inhibit either uPA- or tPA-catalyzed plasminogen activation under conditions in which the serpin PAI-1 was completely inhibitory (7). The anti-proteolytic inhibitory mechanism of serpins is dependent on characteristics of the reactive center loop (RCL) allowing it to adopt the necessary canonical conformation and rearrangements subsequent to protease binding (17). The RCL of maspin does not have the required characteristics (7, 18), and the conclusion that maspin is a non-inhibitory serpin is fully supported by its crystal structure (19, 20).Another confounding factor in understanding the mechanisms underlying the cellular effects of maspin is that, in common with the serpin PAI-2, it lacks an authentic secretion signal sequence. Nevertheless it has been shown to enter secretory vesicles (21) and is found extracellularly, in the cytoplasm and also in the nucleus (21, 22). Cytoplasmic and nuclear binding proteins for maspin have been identified (2325), and may be responsible for its effects on proliferation and apoptosis. How secreted, extracellular maspin exerts its effects is unclear, but a function as a cell signaling ligand has been proposed (2628). However, the characteristics of the maspin inhibitory effect on VSMC migration point to a more direct effect of maspin.To determine the mechanism of the maspin effect on VSMC migration, we have now attempted to identify maspin-binding proteins on the surface of these cells. In this report we provide biochemical, cellular, and functional evidence that the effect of maspin on cell migration is mediated by specific binding to cell adhesion receptors of the integrin family. We find that maspin binds specifically to β1 integrins on the surface of dedifferentiated VSMC, which leads to a reduction in the activation status of the integrin, and that the binding of maspin to α5β1 is sufficient for its inhibitory effects on cell migration and may represent a more general mechanism underlying its diverse biological effects.  相似文献   

19.
Although exosites 1 and 2 regulate thrombin activity by binding substrates and cofactors and by allosterically modulating the active site, it is unclear whether there is direct allosteric linkage between the two exosites. To begin to address this, we first titrated a thrombin variant fluorescently labeled at exosite 1 with exosite 2 ligands, HD22 (a DNA aptamer), γ′-peptide (an analog of the COOH terminus of the γ′-chain of fibrinogen) or heparin. Concentration-dependent and saturable changes in fluorescence were elicited, supporting inter-exosite linkage. To explore the functional consequences of this phenomenon, we evaluated the capacity of exosite 2 ligands to inhibit thrombin binding to γAA-fibrin, an interaction mediated solely by exosite 1. When γAA-fibrinogen was clotted with thrombin in the presence of HD22, γ′-peptide, or prothrombin fragment 2 there was a dose-dependent and saturable decrease in thrombin binding to the resultant fibrin clots. Furthermore, HD22 reduced the affinity of thrombin for γAA-fibrin 6-fold and accelerated the dissociation of thrombin from preformed γAA-fibrin clots. Similar responses were obtained when surface plasmon resonance was used to monitor the interaction of thrombin with γAA-fibrinogen or fibrin. There is bidirectional communication between the exosites, because exosite 1 ligands, HD1 (a DNA aptamer) or hirudin-(54–65) (an analog of the COOH terminus of hirudin), inhibited the exosite 2-mediated interaction of thrombin with immobilized γ′-peptide. These findings provide evidence for long range allosteric linkage between exosites 1 and 2 on thrombin, revealing further complexity to the mechanisms of thrombin regulation.As the central effector of hemostasis, thrombin is engaged in procoagulant, anticoagulant, and fibrinolytic processes. These seemingly contrasting roles are regulated, at least in part, by thrombin''s interactions with other factors in the blood and vasculature. The binding of ligands to thrombin is promoted by exosites 1 and 2, which are positively charged domains that flank the active site. These exosites facilitate the binding of substrates or cofactors and align them for optimal interaction with the active site (1).Exosite 1 is predominantly used to gain access to the active site by substrates such as fibrinogen (2), factors V (3) and VIII (4), and the protease-activated receptors (PARs)2 on platelets (5). Effectors that modulate thrombin activity, including thrombomodulin (6), hirudin (7), and heparin cofactor II (8), also utilize exosite 1. Thrombomodulin alters the specificity of thrombin by hindering access of other substrates to exosite 1 (9) and by providing new binding sites for protein C and thrombin-activable fibrinolysis inhibitor, thereby promoting anticoagulant and antifibrinolytic pathways, respectively (10, 11). Fewer processes are mediated by exosite 2, which serves largely as a tether that anchors thrombin for participation in other reactions. Thus, heparin binds exosite 2 (12) and catalyzes thrombin inhibition by antithrombin and heparin cofactor II (13, 14). Exosite 2 also is used by glycoprotein 1bα on platelets to localize thrombin for activation of PARs (1517).Although the prevailing role of the exosites is to bring substrates and cofactors into proximity with thrombin, there is evidence that the exosites also serve as allosteric regulators of thrombin activity. Crystallographic studies reveal that, when peptides derived from PAR1 or PAR3 are bound to exosite 1 on thrombin, an obstructing surface loop moves out of the active site pocket, thereby providing access to substrates (18). The binding of a thrombomodulin fragment to exosite 1 was shown to alter the environment of an active site fluorescent probe (19), which accelerates the rate of protein C and thrombin-activable fibrinolysis inhibitor activation in an allosteric fashion. In contrast, exosite 1-binding peptides from heparin cofactor II or fibrinogen decrease the rate of protein C activation (20). Additionally, the binding of ligands to exosite 1 alters the rates of chromogenic substrate hydrolysis (21, 22). Allosteric effects are not limited to exosite 1, because prothrombin fragment 2 (F2), a cleavage product of prothrombin, binds exosite 2 and decreases the rate at which thrombin converts fibrinogen to fibrin (23, 24) and is inhibited by antithrombin (25, 26). In support of the concept that these alterations are allosteric in origin, fluorescent probes bound to the active site of thrombin undergo a change in fluorescence intensity when exosite 2 is occupied (24, 27).Although there is good evidence for allosteric regulation of the active site by the exosites, it remains unclear whether there is direct allosteric connection between the exosites. Reciprocal effects between exosites 1 and 2 have been observed by some investigators (2830), but not by others (25, 31). The aim of the current study was to use different techniques and additional ligands to resolve this controversy. First, we examined the effect of exosite 2-directed ligands on the fluorescence intensity of a thrombin variant that was labeled in exosite 1. Next, we examined the effect of these ligands on thrombin binding to fibrin. To exploit the observation that thrombin binds γAA-fibrinogen exclusively via exosite 1 (2, 32), leaving exosite 2 accessible, this subpopulation was isolated (32). We used intact fibrin clots and surface plasmon resonance (SPR) to examine the influence of exosite 2-directed ligands on thrombin binding to γAA-fibrin. In addition, diffusion studies were performed to examine the effect of exosite-directed ligands on the rate of thrombin dissociation from preformed fibrin clots. Finally, we explored whether exosite 1-directed ligands modulate the binding of thrombin to an exosite 2-directed ligand.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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