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.
Focal adhesions are specialized sites of cell attachment to the extracellular matrix where integrin receptors link extracellular matrix to the actin cytoskeleton, and they are constantly remodeled during cell migration. Focal adhesion kinase (FAK) is an important regulator of focal adhesion remodeling. AGAP2 is an Arf GTPase-activating protein that regulates endosomal trafficking and is overexpressed in different human cancers. Here we examined the regulation of the FAK activity and the focal adhesion remodeling by AGAP2. Our results show that FAK binds the pleckstrin homology domain of AGAP2, and the binding is independent of FAK activation following epidermal growth factor receptor stimulation. Overexpression of AGAP2 augments the activity of FAK, and concordantly, the knockdown of AGAP2 expression with RNA interference attenuates the FAK activity stimulated by epidermal growth factor or platelet-derived growth factor receptors. AGAP2 is localized to the focal adhesions, and its overexpression results in dissolution of the focal adhesions, whereas knockdown of its expression stabilizes them. The AGAP2-induced dissolution of the focal adhesions is independent of its GTPase-activating protein activity but may involve its N-terminal G protein-like domain. Our results indicate that AGAP2 regulates the FAK activity and the focal adhesion disassembly during cell migration.Focal adhesions are macromolecular structures that connect actin cytoskeleton to the extracellular matrix and play an important role in cell migration (1). Components of focal adhesions include signaling proteins such as focal adhesion kinase (FAK),3 c-Src, and paxillin, as well as structural proteins such as talin and vinculin (2, 3). Focal adhesions are constantly formed and disassembled (i.e. remodeled) at the leading edge of migrating cells, and they are disassembled at the trailing edge during the cell migration (4, 5). Available evidence demonstrates that the remodeling of focal adhesions is regulated by FAK (6) and Arf-directed GTPase-activating proteins (Arf GAPs) (7).FAK is a member of the Src family nonreceptor tyrosine kinases whose activities are regulated by intra-molecular phosphorylation (8). Autophosphorylation of FAK on tyrosine residue 397 provides docking sites for Src homology 2 domain-containing proteins, including c-Src. After binding to FAK, c-Src phosphorylates FAK on Tyr-576 and Tyr-577 to activate fully the intrinsic kinase activity of FAK (9). Cellular functions of FAK are many and include cell migration, survival, and proliferation; and activation of FAK occurs upon integrin clustering or stimulation of cell surface receptors such as the epidermal growth factor (EGF) or platelet-derived growth factor (PDGF) receptors. FAK activation following integrin clustering results in recruitment of structural and signaling proteins that collectively contribute to the formation of the focal adhesions (10). In FAK null cells, focal adhesions are formed but cannot disassemble (11), suggesting that FAK is required for the focal adhesion disassembly.ADP-ribosylation factors (Arfs) are GTP-binding proteins that lack detectable intrinsic GTPase activities. Therefore, hydrolysis of GTP bound to Arf is mediated by Arf GAPs (12, 13). The AZAP family of Arf GAPs are multidomain proteins that contain a catalytic core of pleckstrin homology (PH), Arf GAP, and ankyrin repeat domains (14), and each subgroup possesses characteristic domain(s). For example, ASAPs have a BAR (Bin, Amphiphysin, Rvs) domain at their N termini and a Src homology 3 domain at their C termini; ARAPs have a Rho GAP domain and five PH domains; and AGAPs have a G protein-like domain (GLD) at their N termini and their PH domains are split, i.e. there is an insert of 80–100 amino acids between the β5 strand and β6 strand. The Arf GAPs regulate membrane trafficking and remodeling of the actin cytoskeleton (7, 15), but the molecular mechanisms underlying the contribution of individual Arf GAPs to membrane trafficking and actin remodeling are being defined. We have reported that AGAP2 binds the clathrin adaptor protein AP-1 and regulates the AP-1/Rab4-dependent endosomal trafficking (16). Studies from other groups have indicated that AGAP2 is overexpressed in different human cancers, including glioblastoma, and that AGAP2 enhances the invasion of glioblastoma cells (17, 18).In this study, we tested the hypothesis that AGAP2 regulates focal adhesion remodeling and cell migration. We find that AGAP2 forms a complex with FAK, increases the FAK activity, and provokes the focal adhesion disassembly that may lead to increased cell migration. Some Arf GAPs have been shown to regulate focal adhesions, and each Arf GAP seems to regulate the focal adhesions by a distinct mechanism. Our results introduce a new way to regulate the focal adhesions by the Arf GAP AGAP2, i.e. through the regulation of FAK activity. These observations support the idea that various Arf GAPs function coordinately to provide temporal and spatial regulation of the focal adhesions during cell migration.  相似文献   

2.
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   

3.
Cell migration is central to normal physiology in embryogenesis, the inflammatory response and wound healing. In addition, the acquisition of a motile and invasive phenotype is an important step in the development of tumors and metastasis. Arf GTPase-activating proteins (GAPs) are nonredundant regulators of specialized membrane surfaces implicated in cell migration. Part of Arf GAP function is mediated by regulating the ADP ribosylation factor (Arf) family GTP-binding proteins. However, Arf GAPs can also function independently of their GAP enzymatic activity, in some cases working as Arf effectors. In this commentary, we discuss examples of Arf GAPs that function either as regulators of Arfs or independently of the GTPase activity to regulate membrane structures that mediate cell adhesion and movement.Key words: Arf GAP, Arf, effector, ADP-ribosylation factor, GTPase-activating protein, focal adhesions, podosomes, invadopodia, cell migrationCell migration involves adhesive structures in which the cell membrane is integrated with the actin cytoskeleton.1 Cells acquire a spatial asymmetry to enable them to turn intracellular generated forces into a net cell body translocation. With the asymmetry, there is a clear distinction between the cell front and rear. Active membrane processes, including lamellipodia and filopodia, take place primarily around the cell front. Extension of both filopodia and lamellipodia is coupled with local actin polymerization, which generates protrusive force. In some cells, focal complexes form at the leading edge of lamellipodia and filopodia. Focal complexes are specialized surfaces of the plasma membrane that mediate attachment to the substratum, providing traction and allowing the cell edge to protrude. Focal complexes mature with cell migration to form another specialized surface in the plasma membrane, focal adhesions (FAs). FAs localize to the termini of stress fiber bundles and serve in longer-term anchorage at the rear of the cell.2 A contractile force is generated at the rear of the cell by the myosin motors to move the cell forward and cell-substratum (extracellular matrix) attachments are released to retract the cell rear. In some cells, podosomes are adhesive structures that mediate cell migration and sometimes invasion.The structures involved in cell migration that are affected by Arf GAPs are FAs, podosomes and invadopodia. FAs contain multiple proteins, including integrins, which are transmembrane proteins.3 The extracellular part of integrins binds to the extracellular matrix. The cytoplasmic domains of integrins associate with multiple signaling proteins as well as proteins that are part of the actin cytoskeleton, thereby coordinating signaling events involved in cell migration and linking the extracellular matrix to the cytoskeleton. Cytoplasmic proteins critical to the function of FAs and that are often used as markers of FAs include vinculin, paxillin and focal adhesion kinase. At least five distinct Arf GAPs have been found to associate with FAs, including GIT1, GIT2, ASAP1, ASAP3 and ARAP2.4Podosomes and invadopodia are related structures induced by action of Src (reviewed in ref. 5). They contain some proteins in common with FAs, but do have some differences that likely reflect different function and/or regulation. For example, podosomes contain ASAP1 but not ASAP3.6 Podosomes and invadopodia have not been examined for the presence of other Arf GAPs. Like FAs, podosomes and invadopodia mediate adhesion to extracellular surfaces. In addition, they are points of degradation of the extracellular matrix and may transfer tension along the extracellular matrix to enable the cell to move. Consistent with the function in motility, podosomes and invadopodia are dynamic structures, turning over in minutes. Podosomes are found in normal physiology of cells including smooth muscle cells, osteoclasts and macrophages and in Src-transformed fibroblasts. Invadopodia are observed in transformed cells, such as cells derived from breast cancers.Two families of GTP-binding proteins within the Ras superfamily, Rho and Arf, are involved in both actin and membrane remodeling. RhoA regulates stress fibers (bundles of actin filaments that traverse the cell and are linked to the extracellular matrix through FAs) and the assembly of FAs.7 Rac1 regulates membrane ruffling and lamellipodia formation.8 Cdc42 regulates filopodia formation.9 Activation of Cdc42 has been shown to lead to the sequential activation of Rac1 and then RhoA in growth factor stimulated fibroblasts.Arf proteins regulate membrane traffic and the actin cytoskeleton.10 There are six mammalian Arf proteins, divided into three classes based on their amino-acid sequence. Arf12 and 3 are class I, Arf4 and Arf5 are class II and Arf6 is the single member of the class III group. Arf1 and Arf6 have been the most extensively studied. Most work has focused on Arf1 function in the Golgi apparatus and endocytic compartments although Arf1 has been found to affect paxillin recruitment to FAs and trafficking of epidermal growth factor receptor from the plasma membrane. Arf6 affects the endocytic pathway and the peripheral actin cytoskeleton.The function of Rho and Arf family proteins depends on a cycle of binding and hydrolyzing GTP. However, Rho and Arf family proteins have slow intrinsic nucleotide exchange. Rho family proteins have slow intrinsic GTPase activity and Arf family proteins have no detectable intrinsic GTPase activity. The cycle of GTP binding and hydrolysis is driven by accessory proteins called guanine nucleotide exchange factors (GEFs) and GTPase-activating proteins (GAPs). Rho family proteins are also regulated by guanine nucleotide dissociation inhibitors, which prevent spontaneous activation in the cytoplasm.Arf GAPs are enzymes that catalyze the hydrolysis of GTP bound to Arf proteins, thereby converting Arf•GTP to Arf•GDP. Thirty-one genes in human encode proteins with Arf GAP domains (Fig. 1). The Arf GAP family is divided into ten subgroups based on domain structure and phylogenetic analysis.11 Six subgroups contain the Arf GAP domain at the N-terminus of the protein. Four groups contain a tandem of a PH, Arf GAP and Ankyrin repeat domains. The Arf GAP nomenclature is mostly based on the protein domain structure. For instance, the ASAP first identified, ASAP1, contains Arf GAP, SH3, Ank repeat and PH domains; ARAPs contain Arf GAP, Rho GAP, Ank repeat and PH domains; ACAPs contain Arf GAP, coiled-coil (later identified as BAR domain), Ank repeat and PH domains; and AGAPs contain Arf GAP, GTP-binding protein-like, Ank repeat and PH domains.Open in a separate windowFigure 1Domain structure of the Arf GAP family. The schematic representation of the ten groups of proteins containing the Arf GAP domain is not drawn to scale. Abbreviations used are: ALPS, ArfGAP1 lipid-packing sensor domain; Ank, Ankyrin repeats; Arf GAP, Arf GTPase activating domain; BAR, Bin/Amphiphysin/Rvs domain; CALM, CALM binding domain; CB, clathrin box; CC, coiled-coiled domain; FG repeats, multiple copies of the XXFG motif; GLD, GTP-binding protein-like domain; PBS, paxillin binding site; PH, pleckstrin homology domain; Pro (PxxP)3, cluster of three proline-rich (PxxP) motifs; Pro (D/ELPPKP)8, eigth tandem Prolin-rich (D/ELPPKP) motifs; RA, Ras association motif; Rho GAP, Rho GTPase activating domain; SAM, sterile α-motif; SH3, Src homology 3 domain; SHD, Spa homology domain. *ASAP2 and ASAP3 lack the Pro (D/ELPPKP)8 motifs. ASAP3 has no SH3 domain. &AGAP2 has a splice variant with three N-terminal PxxP motifs, called PIKE-L. @ARAP2 has an inactive Rho GAP domain.The subcellular localization and function of a number of Arf GAPs have been identified. Arf GAP1, Arf GAP2 and Arf GAP3 are found in the Golgi apparatus where they control membrane traffic by regulating Arf1•GTP levels.12,13 Arf GAP1 has also been proposed to directly contribute to the formation of transport intermediates.14 SMAPs and AGAP1 and AGAP2 are associated with endosomes and regulate endocytic trafficking.14,15 ASAPs, ARAPs and Gits are associated with FAs. ASAPs, ARAPs and ACAPs are found in actin-rich membrane ruffles. ASAP1 is also found in invadopodia and podosomes.4 We propose that common to all Arf GAPs is that they laterally organize membranes, which maintain surfaces of specialized functions such as FAs and podosomes/invadopodia. Some Arf GAPs function primarily as Arf effectors with the turnover rate of the specialized membrane surface being determined by the catalytic rate of the GAP. Other Arf GAPs function as Arf regulators that integrate several signals.ASAP1 is an example of an Arf GAP that may function as an Arf effector to regulate podosomes and invadopodia. ASAP1 is encoded by a gene on the short arm of chromosome 8. The gene is amplified in aggressive forms of uveal melanoma and cell migration rates correlate with ASAP1 expression levels in uveal melanoma16 and other cell types. ASAP1 function depends on cycling among four cellular locations, cytosol, FAs, lamellipodia and podosomes/invadopodia. ASAP1 is necessary for the formation of podosomes/invadopodia.17,18The structural features of ASAP1 that are required to support podosome formation have been examined.17,18 ASAP1 contains, from the N-terminus, BAR, PH, Arf GAP, Ank repeat, proline rich and SH3 domains (Fig. 2A).19 There are two major isoforms, ASAP1a and ASAP1b that differ in the proline rich domain. ASAP1a contains three SH3 binding motifs within the proline rich region including an atypical SH3 binding motif with 6 consecutive prolines. The atypical SH3 binding motif is absent in ASAP1b (Fig. 2A). ASAP1 also has a highly conserved tyrosine between the Ank repeat and proline rich domains that is a site of phosphorylation by the oncogene Src.18Open in a separate windowFigure 2ASAP1 function in podosome and invadopodia formation. (A) Domain structure of ASAP1 splice variants. ASAP1a contains three proline-rich motifs, P1, P2 and P3. P1 and P3 contain a typical (PxxP) motif. P2 contains six prolines. ASAP1b contains only P1 and P3. (B) Model of ASAP1 functioning as an Arf effector to regulate podosome and invadopodia formation. ASAP1 integrates signals from Src, PIP2 and Arf•GTP. For abbreviations of the domain structure of ASAP1 see Figure 1. Other abbreviations: PIP2, phosphoinositides 4,5-biphosphate; Arf1, ADP-ribosylation factor 1.The BAR domain is a bundle of 3 α-helices that homodimerizes to form a boomerang-shaped structure.20,21 BAR domains sense or induce membrane curvature.20 ASAP1 has been found to induce curvature dependent on its BAR domain.22 BAR domains are also protein binding sites.21 The BAR domain of ASAP1 binds to FIP3, a Rab11 and Arf6 binding proteins.23 Arf6-dependent targeting of ASAP1 is likely mediated by FIP3.23 Deletion of or introduction of point mutations into the BAR domain render ASAP1 inactive in supporting podosome formation. The relative role of membrane tubulation and protein binding in mediating the effect of the BAR domain on podosome formation has not been explored.The SH3 domain of ASAP1 binds to focal adhesion kinase24 and pyk2.25 Either deletion of or introduction of point mutations into the SH3 domain abrogates the ability of ASAP1 to support podosome formation.18 The molecular basis for the function of the SH3 domain in podosome formation is not known. The proline rich domain binds to Src19 and CrkL.26 Whether it also binds to cortactin has not been resolved. Reports also conflict regarding the importance of the proline rich domain for podosomes/invadopodia formation.17,18Three signals impinge on ASAP1 to drive podosome formation (Fig. 2B). A conserved tyrosine between the Ank and proline rich motifs is phosphorylated by Src.18,25 Mutation of the tyrosine to phenylalanine results in a protein that functions as a dominant negative blocking podosome formation. ASAP1 with the tyrosine changed to glutamate can support podosome formation, but the mutant ASAP1 is not sufficient to drive podosome formation.18 Based on these results, phosphorylation of the conserved tyrosine is necessary but not sufficient to support podosome formation. Phosphatidylinositol 4,5-bisphosphate (PIP2) binds to the PH domain, which stimulates GAP activity in vitro.27 ASAP1 with mutations in the PH domain that abrogate binding, does not support podosome formation (Jian, Bharti and Randazzo PA, unpublished observations). Point mutations in the PH domain affect both the Km and the kcat for GAP activity. The effect of mutating the PH domain on the ability of ASAP1 to support podosome formation may be consequent to changes in binding Arf1•GTP; it is not likely the result of loss of GAP activity. ASAP1 with a point mutation in the GAP domain that prevents GAP activity but not Arf1•GTP binding is able to support podosome formation whereas a point mutant of ASAP1 that cannot bind Arf1•GTP does not (Jian, Bharti and Randazzo PA, unpublished observations).18 These data support the idea that ASAP1 integrates three signals, (1) PIP2, (2) Src and (3) Arf1•GTP. In response to the signals, ASAP1 functions as a scaffold and directly alters the lipid bilayer to create a domain within the plasma membrane that becomes a podosome. In this model, ASAP1 is functioning as an Arf effector and the GAP activity may regulate the turnover of podosomes.ASAP3, another ASAP-type protein, is found in FAs.6 Reducing ASAP3 expression also reduces cell migration and invasion of Arf GAPs in cell migration mammary carcinoma cells through matrigel. Although ASAP3 does not affect the ability to form FAs, it does affect stress fiber formation and may affect focal adhesion maturation (Ha, Chen and Randazzo PA, unpublished observations).6 The molecular mechanisms underlying the effects of ASAP3 on the cytoskeleton are being examined including the possibility that, like ASAP1, ASAP3 integrates several signals and functions as an Arf effector.ARAPs are examples of Arf GAP family proteins that function as Arf regulators. In common with ASAPs, they integrate a number of signaling pathways and affect the actin cytoskeleton. Three genes encode ARAPs in humans.11 Each of the ARAPs is comprised of a SAM, five PH, Arf GAP, Rho GAP, Ank repeat and Ras association domains. Two of the five PH domains have the consensus sequence for binding to the signaling lipid phosphoinositide 3,4,5-triphosphate (PIP3); however, when examined for ARAP1, PIP3 was not involved in membrane targeting (Campa F, Balla and Randazzo PA, unpublished observations).Examination of the role of ARAP2 in FA formation has provided information about the function of the GAP activity in the cellular function of an Arf GAP. ARAP2 selectively uses Arf6 as a substrate and, different from ARAP1 and ARAP3, has an inactive Rho GAP domain. The Rho GAP domain, however, retains the ability to selectively bind to RhoA•GTP. Also different from ARAP1 and ARAP3, ARAP2 associates with FAs. Cells with reduced expression of ARAP2, consequent to siRNA treatment, have fewer FAs and stress fibers and more focal complexes than control cells. The formation of FAs and stress fibers can be restored by expressing recombinant wild type ARAP2. A mutant of ARAP2 that lacks Arf GAP activity, while retaining the ability to bind to Arf6•GTP, cannot restore FA and stress fiber formation. Similarly, expression of a mutant of ARAP2 that is not able to bind RhoA•GTP cannot reverse the effect of reducing expression of endogenous ARAP2.28 These results support the idea that ARAP2 functions as an Arf GAP that is an effector of RhoA.The model of ARAP2 functioning as a RhoA effector can explain the effects of ARAP2 on FAs (Fig. 3). Arf6•GTP is involved in the formation of Rac1•GTP.29 Rac1•GTP drives lamellipodia and focal complex formation. The conversion of focal complexes to FAs is accompanied by an increase in RhoA•GTP and a decrease in Rac1•GTP. ARAP2 could function to mediate the reciprocal changes in RhoA and Rac1. RhoA•GTP formation leads to the activation of ARAP2. As a consequence of Arf6 GAP activity, Arf6•GTP is converted to Arf6•GDP. With reduced Arf6•GTP, Rac1•GTP concentration also decreases.Open in a separate windowFigure 3Model of ARAP2 as an Arf regulator that controls focal adhesion formation. In this model, ARAP2 functions as a RhoA effector. The inactive Rho GAP domain of ARAP2 binds to RhoA•GTP, which contributes to activation of Arf6 GAP activity. ARAP2 hydrolyzes its substrate Arf6•GTP into Arf6•GDP. Subsequent to Arf6•GTP hydrolysis, Rac1•GTP concentration decreases. For abbreviations of the domain structure of ARAP2 see Figure 1.The Arf GAP activity of other ARAPs may also be critical for cellular functions of the protein. Furthermore, the Rho GAP activity is slow for ARAP1 and ARAP3. It is possible that ARAP1 and ARAP3 can function as Rho effectors with an active Rho GAP domain analogously to ASAP1 functioning as an Arf effector. Further definition of the cellular function of ARAP1 and ARAP3 will provide opportunities to test this idea.We have provided two examples of Arf GAPs that affect cell adhesion and migration. In one case, the Arf GAP appears to function as an Arf effector. In the other case, the Arf GAP functions as a regulator of Arf. The difference in function was discerned using Arf GAP mutants. If functioning as an Arf effector, an Arf GAP mutant that can bind Arf•GTP but not induce hydrolysis can reverse the effect of reduced endogenous Arf GAP, whereas a mutant that cannot bind Arf•GTP cannot replace endogenous Arf GAP. When working as an Arf regulator, a mutant that can bind Arf•GTP but not induce GTP hydrolysis cannot replace endogenous Arf GAP. Whether functioning as an effector or regulator, the rate of GAP activity determines the turnover rate of a specialized membrane surface maintained by Arf.The Arf GAPs have specific sites of action within cells. Some contribute to malignancy, such as ASAP1, ASAP3, AGAP2 and SMAP1.30 The molecular basis of cellular function of each Arf GAPs is distinct. Here, we describe one Arf GAP that functions as an Arf effector and another that functions as an Arf regulator. Each class of Arf GAP has distinct sets of protein binding partners. Furthermore, catalytic mechanism differs among the GAPs. Because of these differences, Arf GAPs may be useful therapeutic targets for cancer therapy.  相似文献   

4.
ARAP1 is a phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GTPase-activating protein (GAP) with five PH domains that regulates endocytic trafficking of the epidermal growth factor receptor (EGFR). Two tandem PH domains are immediately N-terminal of the Arf GAP domain, and one of these fits the consensus sequence for PtdIns(3,4,5)P3 binding. Here, we tested the hypothesis that PtdIns(3,4,5)P3-dependent recruitment mediated by the first PH domain of ARAP1 regulates the in vivo and in vitro function of ARAP1. We found that PH1 of ARAP1 specifically bound to PtdIns(3,4,5)P3, but with relatively low affinity (≈1.6 μm), and the PH domains did not mediate PtdIns(3,4,5)P3-dependent recruitment to membranes in cells. However, PtdIns(3,4,5)P3 binding to the PH domain stimulated GAP activity and was required for in vivo function of ARAP1 as a regulator of endocytic trafficking of the EGFR. Based on these results, we propose a variation on the model for the function of phosphoinositide-binding PH domains. In our model, ARAP1 is recruited to membranes independently of PtdIns(3,4,5)P3, the subsequent production of which triggers enzymatic activity.Pleckstrin homology (PH)2 domains are a common structural motif encoded by the human genome (1, 2). Approximately 10% of PH domains bind to phosphoinositides. These PH domains are thought to mediate phosphoinositide-dependent recruitment to membranes (13). Most PH domains likely have functions other than or in addition to phosphoinositide binding. For example, PH domains have been found to bind to protein and DNA (412). In addition, some PH domains have been found to be structurally and functionally integrated with adjacent domains (13, 14). A small fraction of PH domain-containing proteins (about 9% of the human proteins) have multiple PH domains arranged in tandem, which have been proposed to function as adaptors but have only been examined in one protein (15, 16). Arf GTPase-activating proteins (GAPs) of the ARAP family are phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GAPs with tandem PH domains (17, 18). The function of specific PH domains in regulating Arf GAP activity and for biologic activity has not been described.Arf GAPs are proteins that induce the hydrolysis of GTP bound to Arfs (1923). The Arf proteins are members of the Ras superfamily of GTP-binding proteins (2427). The six Arf proteins in mammals (five in humans) are divided into three classes based on primary sequence: Arf1, -2, and -3 are class 1, Arf4 and -5 are class 2, and Arf6 is class 3 (23, 24, 2729). Class 1 and class 3 Arf proteins have been studied more extensively than class 2. They have been found to regulate membrane traffic and the actin cytoskeleton.The Arf GAPs are a family of proteins with diverse domain structures (20, 21, 23, 30). ARAPs, the most structurally complex of the Arf GAPs, contain, in addition to an Arf GAP domain, the sterile α motif (SAM), five PH, Rho GAP, and Ras association domains (17, 18, 31, 32). The first and second and the third and fourth PH domains are tandem (Fig. 1). The first and third PH domains of the ARAPs fit the consensus for PtdIns(3,4,5)P3 binding (3335). ARAPs have been found to affect actin and membrane traffic (21, 23). ARAP3 regulates growth factor-induced ruffling of porcine aortic endothelial cells (31, 36, 37). The function is dependent on the Arf GAP and Rho GAP domains. ARAP2 regulates focal adhesions, an actin cytoskeletal structure (17). ARAP2 function requires Arf GAP activity and a Rho GAP domain capable of binding RhoA·GTP. ARAP1 has been found to have a role in membrane traffic (18). The protein associates with pre-early endosomes involved in the attenuation of EGFR signals. The function of the tandem PH domains in the ARAPs has not been examined.Open in a separate windowFIGURE 1.ARAP1 binding to phospholipids. A, schematic of the recombinant proteins used in this study. Domain abbreviations: Ank, ankyrin repeat; PLCδ-PH, PH domain of phospholipase C δ; RA, Ras association motif; RhoGAP, Rho GTPase-activating domain. B, ARAP1 phosphoinositide binding specificity. 500 nm PH1-Ank recombinant protein was incubated with sucrose-loaded LUVs formed by extrusion through a 1-μm pore filter. LUVs contained PtdIns alone or PtdIns with 2.5 μm PtdIns(3,4,5)P3, 2.5 μm PtdIns(3)P, 2.5 μm PtdIns(4)P, 2.5 μm PtdIns(5)P, 2.5 μm PtdIns(3,4)P2, 2.5 μm PtdIns(3,5)P2, or 2.5 μm PtdIns(4,5)P2 with a total phosphoinositide concentration of 50 μm and a total phospholipid concentration of 500 μm. Vesicles were precipitated by ultracentrifugation, and associated proteins were separated by SDS-PAGE. The amount of precipitated protein was determined by densitometry of the Coomassie Blue-stained gels with standards on each gel. C, PtdIns(3,4,5)P3-dependent binding of ARAP1 to LUVs. 1 μm PH1-Ank or ArfGAP-Ank recombinant protein was incubated with 1 mm sucrose-loaded LUVs formed by extrusion through a 1-μm pore size filter containing varying concentration of PtdIns(3,4,5)P3. Precipitation of LUVs and analysis of associated proteins were performed as described in B. The average ± S.E. of three independent experiments is presented.Here we investigated the role of the first two PH domains of ARAP1 for catalysis and in vivo function. The first PH domain fits the consensus sequence for PtdIns(3,4,5)P3 binding (3335). The second does not fit a phosphoinositide binding consensus but is immediately N-terminal to the GAP domain. We have previously reported that the PH domain that occurs immediately N-terminal of the Arf GAP domain of ASAP1 is critical for the catalytic function of the protein (38, 39). We tested the hypothesis that the two PH domains of ARAP1 function independently; one recruits ARAP1 to PtdIns(3,4,5)P3-rich membranes, and the other functions with the catalytic domain. As predicted, PH1 interacted specifically with PtdIns(3,4,5)P3, and PH2 did not. However, both PH domains contributed to catalysis independently of recruitment to membranes. None of the PH domains in ARAP1 mediated PtdIns(3,4,5)P3-dependent targeting to plasma membranes (PM). PtdIns(3,4,5)P3 stimulated GAP activity, and the ability to bind PtdIns(3,4,5)P3 was required for ARAP1 to regulate membrane traffic. We propose that ARAP1 is recruited independently of PtdIns(3,4,5)P3 to the PM where PtdIns(3,4,5)P3 subsequently regulates its GAP activity to control endocytic events.  相似文献   

5.
The presence of extensive reciprocal conformational freedom between the catalytic and the hemopexin-like domains of full-length matrix metalloproteinase-1 (MMP-1) is demonstrated by NMR and small angle x-ray scattering experiments. This finding is discussed in relation to the essentiality of the hemopexin-like domain for the collagenolytic activity of MMP-1. The conformational freedom experienced by the present system, having the shortest linker between the two domains, when compared with similar findings on MMP-12 and MMP-9 having longer and the longest linker within the family, respectively, suggests this type of conformational freedom to be a general property of all MMPs.Matrix metalloproteinases (MMP)2 are extracellular hydrolytic enzymes involved in a variety of processes including connective tissue cleavage and remodeling (13). All 23 members of the family are able to cleave simple peptides derived from connective tissue components such as collagen, gelatin, elastin, etc. A subset of MMPs is able to hydrolyze more resistant polymeric substrates, such as cross-linked elastin, and partially degraded collagen forms, such as gelatin and type IV collagens (4). Intact triple helical type I–III collagen is only attacked by collagenases MMP-1, MMP-8, and MMP-13 and by MMP-2 and MMP-14 (512). Although the detailed mechanism of cleavage of single chain peptides by MMP has been largely elucidated (1319), little is known about the process of hydrolysis of triple helical collagen. In fact, triple helical collagen cannot be accommodated in the substrate-binding groove of the catalytic site of MMPs (9).All MMPs (but MMP-7) in their active form are constituted by a catalytic domain (CAT) and a hemopexin-like domain (HPX) (2022). The CAT domain contains two zinc ions and one to three calcium ions. One zinc ion is at the catalytic site and is responsible for the activity, whereas the other metal ions have structural roles. The isolated CAT domains retain full catalytic activity toward simple peptides and single chain polymeric substrates such as elastin, whereas hydrolysis of triple helical collagen also requires the presence of the HPX domain (9, 2325). It has been shown that the isolated CAT domain regains a small fraction of the activity of the full-length (FL) protein when high amounts of either inactivated full-length proteins or isolated HPX domains are added to the assay solution (9). Finally, it has been shown that the presence of the HPX domain alone alters the CD spectrum of triple helical collagen in a way that suggests its partial unwinding (26, 27). It is tempting to speculate that full-length collagenases attack collagen by first locally unwinding the triple helical structure with the help of the HPX domain and then cleaving the resulting, exposed, single filaments (9, 28).Until 2007, three-dimensional structures of full-length MMPs had been reported only for collagenase MMP-1 (2931) and gelatinase MMP-2 (32). The structures of the two proteins are very similar and show a compact arrangement of the two domains, which are connected by a short linker (14 and 20 amino acids, respectively). It is difficult to envisage that rigid and compact molecules of this type can interact with triple helical collagen in a way that can lead to first unwinding and then cleavage of individual filaments. It has been recently suggested that such concerted action could occur much more easily if the two domains could enjoy at least a partial conformational independence (9). Slight differences in the reciprocal orientation of the CAT and HPX domains of MMP-1 in the presence (29) and absence (30, 31) of the prodomain were indeed taken as a hint that the two domains could experience relative mobility (29).Two recent solution studies have shown that conformational independence is indeed occurring in gelatinase MMP-9 (33) and elastase MMP-12 (34), whereas the x-ray structure of the latter (34) is only slightly less compact than those of MMP-1 (2931) and MMP-2 (32). Among MMPs, MMP-9 features an exceptionally long linker (68 amino acid) (33, 35), which in fact constitutes a small domain by itself (the O-glycosylated domain) (33), and therefore, this inspiring observation can hardly be taken as evidence that conformational freedom is a general characteristic of the two-domain MMPs. MMP-12 features a much more normal 16-amino acid linker, thereby making more probable a general functional role for this conformational freedom (34). However, both MMP-9 and MMP-12 retain their full catalytic activity against their substrates even when deprived of the HPX domain (9). Therefore, the question remains of whether conformational freedom is also a required characteristic for those MMPs that are only active as full-length proteins, i.e. collagenases. Interestingly, the three collagenases (MMP-1, MMP-8, and MMP-13) have the shortest linker (14 amino acids) among all MMPs. Demonstrating or negating the presence of conformational freedom in one of these collagenases would therefore constitute a significant step forward to formulate mechanistic hypotheses on their collagenolytic activity.Our recent studies on MMP-12 in solution (34) have shown that a combination of NMR relaxation studies and small angle x-ray scattering (SAXS) is enough to show the presence and the extent of the relative conformational freedom of the two domains of MMPs. Here we apply the same strategy to full-length MMP-1 and show that sizable conformational freedom is indeed experienced even by this prototypical collagenase, although somewhat less pronounced than that observed for MMP-12.  相似文献   

6.
Previous studies showed that ADP-ribosylation factor 6 (Arf6) is important for platelet function; however, little is known about which signaling events regulate this small GTP-binding protein. Arf6-GTP was monitored in platelets stimulated with a number of agonists (TRAP, thrombin, convulxin, collagen, PMA, thapsigargin, or A23187) and all led to a time-dependent decrease in Arf6-GTP. ADP and U46619 were without effect. Using inhibitors, it was shown that the decrease of Arf6-GTP is a direct consequence of known signaling cascades. Upon stimulation via PAR receptors, Arf6-GTP loss could be blocked by treatment with U-73122, BAPTA/AM, Ro-31-8220, or Gö6976, indicating requirements for phospholipase C, calcium, and protein kinase C (PKC) α/β, respectively. The Arf6-GTP decrease in convulxin-stimulated platelets showed similar requirements and was also sensitive to piceatannol, wortmannin, and LY294002, indicating additional requirements for Syk and phosphatidylinositol 3-kinase. The convulxin-induced decrease was sensitive to both PKCα/β and δ inhibitors. Outside-in signaling, potentially via integrin engagement, caused a second wave of signaling that affected Arf6. Inclusion of RGDS peptides or EGTA, during activation, led to a biphasic response; Arf6-GTP levels partially recovered upon continued incubation. A similar response was seen in β3 integrin-null platelets. These data show that Arf6-GTP decreases in response to known signaling pathways associated with PAR and GPVI. They further reveal a second, aggregation-dependent, process that dampens Arf6-GTP recovery. This study demonstrates that the nucleotide state of Arf6 in platelets is regulated during the initial phases of activation and during the later stages of aggregation.Platelet activation is initiated through several classes of membrane receptors, which are stimulated by agonists produced at the vascular lesion (13). A second wave of signaling, caused by engagement of integrins, occurs as platelets bind to the lesion surface and aggregate (4). Together, these plasma membrane proteins initiate the platelet processes important for thrombosis (e.g. adhesion, spreading, secretion, and clot retraction). Small GTP-binding proteins, specifically members of the Ras superfamily, link signaling events from various platelet receptors to defined outcomes, such as shape change (57), aggregation (8, 9), and secretion (1012). Rab proteins play roles in granule secretion, with Rab4 and Rab6 being involved in alpha granule release (10, 11) and Rab27a/b in dense core granule release (12, 13). RalA is activated in response to various stimuli (1416) and may play a role in secretion by anchoring the exocyst complex to specific membrane sites (17). Rap1 plays a role in integrin αIIbβ3 activation (8, 9). Rho family GTPases (Rho, Rac, and Cdc42) play roles in platelet phosphoinositide signaling and in the regulation of the actin cytoskeleton (57). While these small GTP-binding proteins are clearly important to platelet function, it is equally clear that other small G proteins are present and functional in platelets (18).The ADP-ribosylation factor (Arf)2 family are Ras-related, small GTPases that affect both vesicular transport and cytoskeletal dynamics (19, 20). Based on their primary sequences, this family is divided into three classes, with Arf6 as the only member of class III (19). Arf6-GTP is considered the “active state” and can interact with downstream effectors, such as phospholipase D (PLD) (21), phosphatidylinositol 4-phosphate 5-kinase type α (22), and arfaptin 2 (23, 24), resulting in the recruitment of these effectors to the plasma membrane. The Arf6 GTP/GDP cycle is mediated by interactions with guanine nucleotide exchange factors (GEFs) and GTPase-activating proteins (GAPs). The large number of Arf-GEF and -GAP proteins have been discussed in recent reviews where it was noted that, unlike other small GTPases, Arf functions are generally not mediated solely by the GTP-bound state but through its cycling between states (19, 20, 25, 26).The effects that Arf6 has on the secretion and actin dynamics in nucleated cells make it an ideal candidate for function in platelets. Arf6 influences cortical actin and is important for spreading, ruffling, migration, and phagocytosis (reviewed in Ref. 19). Our previous work (27) showed that Arf6 is present on platelet membranes and is important for platelet function. Unlike other small G proteins, the Arf6 GTP-bound form is readily detectible in resting platelets and upon activation with collagen or convulxin there is a rapid conversion to the GDP-bound form. Acylated peptides, which mimic the myristoylated N terminus of Arfs have been used as isoform-specific inhibitors (28). In platelets, a myristoylated-Arf6 (myr-Arf6) peptide specifically blocks the activation-dependent loss of Arf6-GTP. This peptide also blocks aggregation, spreading on collagen, and activation of the Rho family of GTPases. Other GTPases, such as Ral and Rap, were unaffected. The simplest explanation for these data is that platelet activation stimulates the GTPase activity of Arf6, perhaps through activation of an Arf6-GAP. Alternatively, platelet activation could affect an Arf6-GEF thus reducing the production of Arf6-GTP. Regardless of mechanism, disruption of the activation-dependent loss of Arf6-GTP, with the myr-Arf6 peptide, profoundly affects the actin-based cytoskeletal rearrangements associated with platelet activation. While our initial report (27) established a role for Arf6 in platelet function, it was not clear what platelet signaling events were required to induce the loss of Arf6-GTP.In this article, we delineate the signaling cascades required for the activation-dependent loss of Arf6-GTP. We show that the Arf6-GTP to -GDP conversion was stimulated by primary agonists (thrombin, TRAP, collagen, or convulxin) but not by ADP or U46619. The decrease in Arf6-GTP, downstream of thrombin and convulxin, required PLC, and PKC activity. Loss of Arf6-GTP, via stimulation of GPVI with convulxin, additionally required Syk and PI3K activities. Pretreatment with passivators, nitric oxide (NO), and prostaglandin I2 (PGI2) blocked thrombin- and convulxin-induced loss of Arf6-GTP. Further experiments suggested a role for “outside-in” signaling, especially once platelet aggregates begin to form. Inclusion of RGDS peptide, EGTA, or the deletion of the β3 integrin had only minimal effects on the initial loss of Arf6-GTP but led to the partial recovery of Arf6-GTP levels. This biphasic change in Arf6-GTP levels was not seen when aggregation was allowed to occur normally. Taken together, these data show that the Arf6 nucleotide state is responsive to both initial agonist-mediated signaling and to a second wave of integrin-mediated signaling that occurs upon aggregation.  相似文献   

7.
Formin-homology (FH) 2 domains from formin proteins associate processively with the barbed ends of actin filaments through many rounds of actin subunit addition before dissociating completely. Interaction of the actin monomer-binding protein profilin with the FH1 domain speeds processive barbed end elongation by FH2 domains. In this study, we examined the energetic requirements for fast processive elongation. In contrast to previous proposals, direct microscopic observations of single molecules of the formin Bni1p from Saccharomyces cerevisiae labeled with quantum dots showed that profilin is not required for formin-mediated processive elongation of growing barbed ends. ATP-actin subunits polymerized by Bni1p and profilin release the γ-phosphate of ATP on average >2.5 min after becoming incorporated into filaments. Therefore, the release of γ-phosphate from actin does not drive processive elongation. We compared experimentally observed rates of processive elongation by a number of different FH2 domains to kinetic computer simulations and found that actin subunit addition alone likely provides the energy for fast processive elongation of filaments mediated by FH1FH2-formin and profilin. We also studied the role of FH2 structure in processive elongation. We found that the flexible linker joining the two halves of the FH2 dimer has a strong influence on dissociation of formins from barbed ends but only a weak effect on elongation rates. Because formins are most vulnerable to dissociation during translocation along the growing barbed end, we propose that the flexible linker influences the lifetime of this translocative state.Formins are multidomain proteins that assemble unbranched actin filament structures for diverse processes in eukaryotic cells (reviewed in Ref. 1). Formins stimulate nucleation of actin filaments and, in the presence of the actin monomer-binding protein profilin, speed elongation of the barbed ends of filaments (2-6). The ability of formins to influence elongation depends on the ability of single formin molecules to remain bound to a growing barbed end through multiple rounds of actin subunit addition (7, 8). To stay associated during subunit addition, a formin molecule must translocate processively on the barbed end as each actin subunit is added (1, 9-12). This processive elongation of a barbed end by a formin is terminated when the formin dissociates stochastically from the growing end during translocation (4, 10).The formin-homology (FH)2 1 and 2 domains are the best conserved domains of formin proteins (2, 13, 14). The FH2 domain is the signature domain of formins, and in many cases, is sufficient for both nucleation and processive elongation of barbed ends (2-4, 7, 15). Head-to-tail homodimers of FH2 domains (12, 16) encircle the barbed ends of actin filaments (9). In vitro, association of barbed ends with FH2 domains slows elongation by limiting addition of free actin monomers. This “gating” behavior is usually explained by a rapid equilibrium of the FH2-associated end between an open state competent for actin monomer association and a closed state that blocks monomer binding (4, 9, 17).Proline-rich FH1 domains located N-terminal to FH2 domains are required for profilin to stimulate formin-mediated elongation. Individual tracks of polyproline in FH1 domains bind 1:1 complexes of profilin-actin and transfer the actin directly to the FH2-associated barbed end to increase processive elongation rates (4-6, 8, 10, 17).Rates of elongation and dissociation from growing barbed ends differ widely for FH1FH2 fragments from different formin homologs (4). We understand few aspects of FH1FH2 domains that influence gating, elongation or dissociation. In this study, we examined the source of energy for formin-mediated processive elongation, and the influence of FH2 structure on elongation and dissociation from growing ends. In contrast to previous proposals (6, 18), we found that fast processive elongation mediated by FH1FH2-formins is not driven by energy from the release of the γ-phosphate from ATP-actin filaments. Instead, the data show that the binding of an actin subunit to the barbed end provides the energy for processive elongation. We found that in similar polymerizing conditions, different natural FH2 domains dissociate from growing barbed ends at substantially different rates. We further observed that the length of the flexible linker between the subunits of a FH2 dimer influences dissociation much more than elongation.  相似文献   

8.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

9.
NHE5 is a brain-enriched Na+/H+ exchanger that dynamically shuttles between the plasma membrane and recycling endosomes, serving as a mechanism that acutely controls the local pH environment. In the current study we show that secretory carrier membrane proteins (SCAMPs), a group of tetraspanning integral membrane proteins that reside in multiple secretory and endocytic organelles, bind to NHE5 and co-localize predominantly in the recycling endosomes. In vitro protein-protein interaction assays revealed that NHE5 directly binds to the N- and C-terminal cytosolic extensions of SCAMP2. Heterologous expression of SCAMP2 but not SCAMP5 increased cell-surface abundance as well as transporter activity of NHE5 across the plasma membrane. Expression of a deletion mutant lacking the SCAMP2-specific N-terminal cytosolic domain, and a mini-gene encoding the N-terminal extension, reduced the transporter activity. Although both Arf6 and Rab11 positively regulate NHE5 cell-surface targeting and NHE5 activity across the plasma membrane, SCAMP2-mediated surface targeting of NHE5 was reversed by dominant-negative Arf6 but not by dominant-negative Rab11. Together, these results suggest that SCAMP2 regulates NHE5 transit through recycling endosomes and promotes its surface targeting in an Arf6-dependent manner.Neurons and glial cells in the central and peripheral nervous systems are especially sensitive to perturbations of pH (1). Many voltage- and ligand-gated ion channels that control membrane excitability are sensitive to changes in cellular pH (1-3). Neurotransmitter release and uptake are also influenced by cellular and organellar pH (4, 5). Moreover, the intra- and extracellular pH of both neurons and glia are modulated in a highly transient and localized manner by neuronal activity (6, 7). Thus, neurons and glia require sophisticated mechanisms to finely tune ion and pH homeostasis to maintain their normal functions.Na+/H+ exchangers (NHEs)3 were originally identified as a class of plasma membrane-bound ion transporters that exchange extracellular Na+ for intracellular H+, and thereby regulate cellular pH and volume. Since the discovery of NHE1 as the first mammalian NHE (8), eight additional isoforms (NHE2-9) that share 25-70% amino acid identity have been isolated in mammals (9, 10). NHE1-5 commonly exhibit transporter activity across the plasma membrane, whereas NHE6-9 are mostly found in organelle membranes and are believed to regulate organellar pH in most cell types at steady state (11). More recently, NHE10 was identified in human and mouse osteoclasts (12, 13). However, the cDNA encoding NHE10 shares only a low degree of sequence similarity with other known members of the NHE gene family, raising the possibility that this sodium-proton exchanger may belong to a separate gene family distantly related to NHE1-9 (see Ref. 9).NHE gene family members contain 12 putative transmembrane domains at the N terminus followed by a C-terminal cytosolic extension that plays a role in regulation of the transporter activity by protein-protein interactions and phosphorylation. NHEs have been shown to regulate the pH environment of synaptic nerve terminals and to regulate the release of neurotransmitters from multiple neuronal populations (14-16). The importance of NHEs in brain function is further exemplified by the findings that spontaneous or directed mutations of the ubiquitously expressed NHE1 gene lead to the progression of epileptic seizures, ataxia, and increased mortality in mice (17, 18). The progression of the disease phenotype is associated with loss of specific neuron populations and increased neuronal excitability. However, NHE1-null mice appear to develop normally until 2 weeks after birth when symptoms begin to appear. Therefore, other mechanisms may compensate for the loss of NHE1 during early development and play a protective role in the surviving neurons after the onset of the disease phenotype.NHE5 was identified as a unique member of the NHE gene family whose mRNA is expressed almost exclusively in the brain (19, 20), although more recent studies have suggested that NHE5 might be functional in other cell types such as sperm (21, 22) and osteosarcoma cells (23). Curiously, mutations found in several forms of congenital neurological disorders such as spinocerebellar ataxia type 4 (24-26) and autosomal dominant cerebellar ataxia (27-29) have been mapped to chromosome 16q22.1, a region containing NHE5. However, much remains unknown as to the molecular regulation of NHE5 and its role in brain function.Very few if any proteins work in isolation. Therefore identification and characterization of binding proteins often reveal novel functions and regulation mechanisms of the protein of interest. To begin to elucidate the biological role of NHE5, we have started to explore NHE5-binding proteins. Previously, β-arrestins, multifunctional scaffold proteins that play a key role in desensitization of G-protein-coupled receptors, were shown to directly bind to NHE5 and promote its endocytosis (30). This study demonstrated that NHE5 trafficking between endosomes and the plasma membrane is regulated by protein-protein interactions with scaffold proteins. More recently, we demonstrated that receptor for activated C-kinase 1 (RACK1), a scaffold protein that links signaling molecules such as activated protein kinase C, integrins, and Src kinase (31), directly interacts with and activates NHE5 via integrin-dependent and independent pathways (32). These results further indicate that NHE5 is partly associated with focal adhesions and that its targeting to the specialized microdomain of the plasma membrane may be regulated by various signaling pathways.Secretory carrier membrane proteins (SCAMPs) are a family of evolutionarily conserved tetra-spanning integral membrane proteins. SCAMPs are found in multiple organelles such as the Golgi apparatus, trans-Golgi network, recycling endosomes, synaptic vesicles, and the plasma membrane (33, 34) and have been shown to play a role in exocytosis (35-38) and endocytosis (39). Currently, five isoforms of SCAMP have been identified in mammals. The extended N terminus of SCAMP1-3 contain multiple Asn-Pro-Phe (NPF) repeats, which may allow these isoforms to participate in clathrin coat assembly and vesicle budding by binding to Eps15 homology (EH)-domain proteins (40, 41). Further, SCAMP2 was shown recently to bind to the small GTPase Arf6 (38), which is believed to participate in traffic between the recycling endosomes and the cell surface (42, 43). More recent studies have suggested that SCAMPs bind to organellar membrane type NHE7 (44) and the serotonin transporter SERT (45) and facilitate targeting of these integral membrane proteins to specific intracellular compartments. We show in the current study that SCAMP2 binds to NHE5, facilitates the cell-surface targeting of NHE5, and elevates Na+/H+ exchange activity at the plasma membrane, whereas expression of a SCAMP2 deletion mutant lacking the N-terminal domain containing the NPF repeats suppresses the effect. Further we show that this activity of SCAMP2 requires an active form of a small GTPase Arf6, but not Rab11. We propose a model in which SCAMPs bind to NHE5 in the endosomal compartment and control its cell-surface abundance via an Arf6-dependent pathway.  相似文献   

10.
Cyclic di-GMP (c-di-GMP) is a ubiquitous bacterial second messenger that is involved in the regulation of cell surface-associated traits and the persistence of infections. Omnipresent GGDEF and EAL domains, which occur in various combinations with regulatory domains, catalyze c-di-GMP synthesis and degradation, respectively. The crystal structure of full-length YkuI from Bacillus subtilis, composed of an EAL domain and a C-terminal PAS-like domain, has been determined in its native form and in complex with c-di-GMP and Ca2+. The EAL domain exhibits a triose-phosphate isomerase-barrel fold with one antiparallel β-strand. The complex with c-di-GMP-Ca2+ defines the active site of the putative phosphodiesterase located at the C-terminal end of the β-barrel. The EAL motif is part of the active site with Glu-33 of the motif being involved in cation coordination. The structure of the complex allows the proposal of a phosphodiesterase mechanism, in which the divalent cation and the general base Glu-209 activate a catalytic water molecule for nucleophilic in-line attack on the phosphorus. The C-terminal domain closely resembles the PAS-fold. Its pocket-like structure could accommodate a yet unknown ligand. YkuI forms a tight dimer via EAL-EAL and trans EAL-PAS-like domain association. The possible regulatory significance of the EAL-EAL interface and a mechanism for signal transduction between sensory and catalytic domains of c-di-GMP-specific phosphodiesterases are discussed.The dinucleotide cyclic di-GMP (c-di-GMP) was discovered about 20 years ago when it was found to regulate the activity of cellulase synthase in Acetobacter xylinum (1). However, its prominent role as a global second messenger has been realized only upon the recent recognition of the omnipresence of genes coding for domains that catalyze c-di-GMP biosynthesis and degradation in eubacteria (2). GGDEF domains catalyze the condensation of two GTP molecules to the cyclic 2-fold symmetric dinucleotide (diguanylate cyclase activity (3-6)), whereas EAL domains are involved in its degradation to yield the linear dinucleotide pGpG (phosphodiesterase (PDE)4 A activity) (3, 7-9). Recently, also HD-GYP domains have been implicated in c-di-GMP-specific PDE activity (10). All the domains have been named according to their sequence signature motifs. They are typically found in combinations with various other, mostly sensory or regulatory, domains. It is thought that the balance between antagonistic diguanylate cyclase and PDE-A activities determines the cellular level of c-di-GMP and, thus, affects a variety of physiological processes in bacteria.It has been shown that, in general, c-di-GMP regulates cell surface-associated traits and community behavior such as biofilm formation (for reviews see Refs. 11-12), and its relevance to the virulence of pathogenic bacteria has been demonstrated (11, 13, 14). In particular, the dinucleotide has been proposed to orchestrate the switch between acute and persistent phase of infection.The best characterized diguanylate cyclase is PleD from Caulobacter crescentus with a Rec-Rec-GGDEF domain architecture (Rec indicates response regulator receiver domain). The structure of its GGDEF domain revealed a single GTP-binding site and suggested that dimerization is the prerequisite for enzymatic activity (4). This has been corroborated recently by crystallography showing directly that modification of the first Rec domain, mimicking phosphorylation by the cognate kinase, induces formation of a tightly packed dimer (15). Additionally, an upper limit of c-di-GMP levels in the cell seems to be ensured by potent allosteric product inhibition of the PleD cyclase (4, 15, 16). Recently, the crystal structure of another diguanylate cyclase, WspR from Pseudomonas aeruginosa with a Rec-GGDEF domain architecture, has been determined (17), which showed a tetrameric quaternary structure and active and feedback inhibition sites that are very similar to those in PleD.For EAL domains, it has been demonstrated that genetic knock-out results in phenotypes that are in line with the paradigm that an elevated cellular c-di-GMP concentration corresponds to a sessile and a low concentration to a motile bacterial life style (13, 18, 19). Only recently, EAL-mediated PDE-A activity has been measured in vitro (7-9, 20-22).The Bacillus subtilis YkuI protein was targeted for structure determination by the Midwest Center for Structural Genomics as a member of the large sequence family that contains EAL (Pfam number PF00563) domains. Here we report the crystal structure of YkuI showing the fold of the N-terminal EAL domain and the C-terminal PAS-like domain. Co-crystallization with c-di-GMP revealed the substrate binding mode and allows the proposal of a catalytic mechanism. The PAS-like domain most probably has regulatory function, which is discussed. Recently, another EAL structure has been deposited in the Protein Data Bank by the Midwest Center for Structural Genomics, the EAL domain of a GGDEF-EAL protein from Thiobacillus denitrificans (tdEAL; PDB code 2r6o). Comparison of the two structures suggests a possible regulatory mechanism.  相似文献   

11.
The carboxyl-terminal cholinesterase-like (ChEL) domain of thyroglobulin (Tg) has been identified as critically important in Tg export from the endoplasmic reticulum. In a number of human kindreds suffering from congenital hypothyroidism, and in the cog congenital goiter mouse and rdw rat dwarf models, thyroid hormone synthesis is inhibited because of mutations in the ChEL domain that block protein export from the endoplasmic reticulum. We hypothesize that Tg forms homodimers through noncovalent interactions involving two predicted α-helices in each ChEL domain that are homologous to the dimerization helices of acetylcholinesterase. This has been explored through selective epitope tagging of dimerization partners and by inserting an extra, unpaired Cys residue to create an opportunity for intermolecular disulfide pairing. We show that the ChEL domain is necessary and sufficient for Tg dimerization; specifically, the isolated ChEL domain can dimerize with full-length Tg or with itself. Insertion of an N-linked glycan into the putative upstream dimerization helix inhibits homodimerization of the isolated ChEL domain. However, interestingly, co-expression of upstream Tg domains, either in cis or in trans, overrides the dimerization defect of such a mutant. Thus, although the ChEL domain provides a nidus for Tg dimerization, interactions of upstream Tg regions with the ChEL domain actively stabilizes the Tg dimer complex for intracellular transport.The synthesis of thyroid hormone in the thyroid gland requires secretion of thyroglobulin (Tg)2 to the apical luminal cavity of thyroid follicles (1). Once secreted, Tg is iodinated via the activity of thyroid peroxidase (2). A coupling reaction involving a quinol-ether linkage especially engages di-iodinated tyrosyl residues 5 and 130 to form thyroxine within the amino-terminal portion of the Tg polypeptide (3, 4). Preferential iodination of Tg hormonogenic sites is dependent not on the specificity of the peroxidase (5) but upon the native structure of Tg (6, 7). To date, no other thyroidal proteins have been shown to effectively substitute in this role for Tg.The first 80% of the primary structure of Tg (full-length murine Tg: 2,746 amino acids) involves three regions called I-II-III comprised of disulfide-rich repeat domains held together by intradomain disulfide bonds (8, 9). The final 581 amino acids of Tg are strongly homologous to acetylcholinesterase (1012). Rate-limiting steps in the overall process of Tg secretion involve its structural maturation within the endoplasmic reticulum (ER) (13). Interactions between regions I-II-III and the cholinesterase-like (ChEL) domain have recently been suggested to be important in this process, with ChEL functioning as an intramolecular chaperone and escort for I-II-III (14). In addition, Tg conformational maturation culminates in Tg homodimerization (15, 16) with progression to a cylindrical, and ultimately, a compact ovoid structure (1719).In human congenital hypothyroidism with deficient Tg, the ChEL domain is a commonly affected site of mutation, including the recently described A2215D (20, 21), R2223H (22), G2300D, R2317Q (23), G2355V, G2356R, and the skipping of exon 45 (which normally encodes 36 amino acids), as well as the Q2638stop mutant (24) (in addition to polymorphisms including P2213L, W2482R, and R2511Q that may be associated with thyroid overgrowth (25)). As best as is currently known, all of the congenital hypothyroidism-inducing Tg mutants are defective for intracellular transport (26). A homozygous G2300R mutation (equivalent to residue 2,298 of mouse Tg) in the ChEL domain is responsible for congenital hypothyroidism in rdw rats (27, 28), whereas we identified the Tg-L2263P point mutation as the cause of hypothyroidism in the cog mouse (29). Such mutations perturb intradomain structure (30), and interestingly, block homodimerization (31). Acquisition of quaternary structure has long been thought to be required for efficient export from the ER (32) as exemplified by authentic acetylcholinesterase (33, 34) in which dimerization enhances protein stability and export (35).Tg comprised only of regions I-II-III (truncated to lack the ChEL domain) is blocked within the ER (30), whereas a secretory version of the isolated ChEL domain of Tg devoid of I-II-III undergoes rapid and efficient intracellular transport and secretion (14). A striking homology positions two predicted α-helices of the ChEL domain to the identical relative positions of the dimerization helices in acetylcholinesterase. This raises the possibility that ChEL may serve as a homodimerization domain for Tg, providing a critical function in maturation for Tg transport to the site of thyroid hormone synthesis (1).In this study, we provide unequivocal evidence for homodimerization of the ChEL domain and “hetero”-dimerization of that domain with full-length Tg, and we provide significant evidence that the predicted ChEL dimerization helices provide a nidus for Tg assembly. On the other hand, our data also suggest that upstream Tg regions known to interact with ChEL (14) actively stabilize the Tg dimer complex. Together, I-II-III and ChEL provide unique contributions to the process of intracellular transport of Tg through the secretory pathway.  相似文献   

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

14.
The causative agent of Legionnaires disease, Legionella pneumophila, forms a replicative vacuole in phagocytes by means of the intracellular multiplication/defective organelle trafficking (Icm/Dot) type IV secretion system and translocated effector proteins, some of which subvert host GTP and phosphoinositide (PI) metabolism. The Icm/Dot substrate SidC anchors to the membrane of Legionella-containing vacuoles (LCVs) by specifically binding to phosphatidylinositol 4-phosphate (PtdIns(4)P). Using a nonbiased screen for novel L. pneumophila PI-binding proteins, we identified the Rab1 guanine nucleotide exchange factor (GEF) SidM/DrrA as the predominant PtdIns(4)P-binding protein. Purified SidM specifically and directly bound to PtdIns(4)P, whereas the SidM-interacting Icm/Dot substrate LidA preferentially bound PtdIns(3)P but also PtdIns(4)P, and the L. pneumophila Arf1 GEF RalF did not bind to any PIs. The PtdIns(4)P-binding domain of SidM was mapped to the 12-kDa C-terminal sequence, termed “P4M” (PtdIns4P binding of SidM/DrrA). The isolated P4M domain is largely helical and displayed higher PtdIns(4)P binding activity in the context of the α-helical, monomeric full-length protein. SidM constructs containing P4M were translocated by Icm/Dot-proficient L. pneumophila and localized to the LCV membrane, indicating that SidM anchors to PtdIns(4)P on LCVs via its P4M domain. An L. pneumophila ΔsidM mutant strain displayed significantly higher amounts of SidC on LCVs, suggesting that SidM and SidC compete for limiting amounts of PtdIns(4)P on the vacuole. Finally, RNA interference revealed that PtdIns(4)P on LCVs is specifically formed by host PtdIns 4-kinase IIIβ. Thus, L. pneumophila exploits PtdIns(4)P produced by PtdIns 4-kinase IIIβ to anchor the effectors SidC and SidM to LCVs.The Gram-negative pathogen Legionella pneumophila is the causative agent of Legionnaires disease, but it evolved as a parasite of various species of environmental predatory protozoa, including the social amoeba Dictyostelium discoideum (1, 2). The human disease is linked to the inhalation of contaminated aerosols, followed by replication in alveolar macrophages. To accommodate the transfer between host cells, L. pneumophila alternates between replicative and transmissive phases, the regulation of which includes an apparent quorum-sensing system (35).In macrophages and amoebae, L. pneumophila forms a replicative compartment, the Legionella-containing vacuole (LCV).3 LCVs avoid fusion with lysosomes (6), intercept vesicular traffic at endoplasmic reticulum (ER) exit sites (7), and fuse with the ER (810). The uptake of L. pneumophila and formation of LCVs in macrophages and amoebae depends on the Icm/Dot type IV secretion system (T4SS) (1114). Although more than 100 Icm/Dot substrates (“effector” proteins) have been identified to date, only few are functionally characterized, including effectors that interfere with host cell signal transduction, vesicle trafficking, or apoptotic pathways (1518).Two Icm/Dot-translocated substrates, SidM/DrrA (19, 20) and RalF (21), have been characterized as guanine nucleotide exchange factors (GEFs) for the Rho subfamily of small GTPases. These bacterial GEFs are recruited to and activate their targets on LCVs. Small GTPases of the Rho subfamily are involved in many eukaryotic signal transduction pathways and in actin cytoskeleton regulation (22). Inactive Rho GTPases bind GDP and a guanine nucleotide dissociation inhibitor (GDI). The GTPases are activated by removal of the GDI and the exchange of GDP with GTP by GEFs, which promotes the interaction with downstream effector proteins, such as protein or lipid kinases and various adaptor proteins. The cycle is closed by hydrolysis of the bound GTP, which is mediated by GTPase-activating proteins.SidM is a GEF for Rab1, which is essential for ER to Golgi vesicle transport, and additionally, SidM acts as a GDI displacement factor (GDF) to activate Rab1 (23, 24). The function of SidM is assisted by the Icm/Dot substrate LidA, which also localizes to LCVs. LidA preferentially binds to activated Rab1, thus supporting the recruitment of early secretory vesicles by SidM (19, 20, 23, 25, 26). Another Icm/Dot substrate, LepB (27), contributes to Rab1-mediated membrane cycling by inactivating Rab1 through its GTPase-activating protein function, thus acting as an antagonist of SidM (24).The Icm/Dot substrate RalF recruits and activates the small GTPase ADP-ribosylation factor 1 (Arf1), which is involved in retrograde vesicle transport from Golgi to ER (21). Dominant negative Arf1 (7, 28) or knockdown of Arf1 by RNA interference (29) impairs the formation of LCVs, as well as the recruitment of the Icm/Dot substrate SidC to the LCV (30).SidC and its paralogue SdcA localize to the LCV membrane (31), where the proteins specifically bind to the host cell lipid phosphatidylinositol 4-phosphate (PtdIns(4)P) (32, 33). Phosphoinositides (PIs) regulate eukaryotic receptor-mediated signal transduction, actin remodeling, and membrane dynamics (34, 35). PtdIns(4)P is present on the cytoplasmic membrane, but localizes preferentially to the trans-Golgi network (TGN), where this PI is produced by an Arf-dependent recruitment of PtdIns(4)P kinase IIIβ (PI4K IIIβ) (36) to promote trafficking along the secretory pathway. Recently, PtdIns(4)P was found to also mediate the export of early secretory vesicles from ER exit sites (37). At present, the L. pneumophila effector proteins that mediate exploitation of host PI signaling remain ill defined.In a nonbiased screen for L. pneumophila PI-binding proteins using different PIs coupled to agarose beads, we identified SidM as a major PtdIns(4)P-binding effector. We mapped its PtdIns(4)P binding activity to a novel P4M domain within a 12-kDa C-terminal sequence. SidM constructs, including the P4M domain, were found to be translocated and bind the LCV membrane, where the levels of PtdIns(4)P are controlled by PI4K IIIβ.  相似文献   

15.
16.
Protein-disulfide isomerase (PDI), an endoplasmic reticulum (ER)-resident protein, is primarily known as a catalyst of oxidative protein folding but also has a protein unfolding activity. We showed previously that PDI unfolds the cholera toxin A1 (CTA1) polypeptide to facilitate the ER-to-cytosol retrotranslocation of the toxin during intoxication. We now provide insight into the mechanism of this unfoldase activity. PDI includes two redox-active (a and a′) and two redox-inactive (b and b′) thioredoxin-like domains, a linker (x), and a C-terminal domain (c) arranged as abbxac. Using recombinant PDI fragments, we show that binding of CTA1 by the continuous PDIbbxa′ fragment is necessary and sufficient to trigger unfolding. The specific linear arrangement of bbxa′ and the type a domain (aversus a) C-terminal to bbx are additional determinants of activity. These data suggest a general mechanism for the unfoldase activity of PDI: the concurrent and specific binding of bbxa′ to particular regions along the CTA1 molecule triggers its unfolding. Furthermore, we show the bb′ domains of PDI are indispensable to the unfolding reaction, whereas the function of its a′ domain can be substituted partially by the a′ domain from ERp57 (abbxac) or ERp72 (ca°abbxa′), PDI-like proteins that do not unfold CTA1 normally. However, the bb′ domains of PDI were insufficient to convert full-length ERp57 into an unfoldase because the a domain of ERp57 inhibited toxin binding. Thus, we propose that generating an unfoldase from thioredoxin-like domains requires the bb′(x) domains of PDI followed by an a′ domain but not preceded by an inhibitory a domain.Protein-disulfide isomerase (PDI)2 is a multifunctional protein that resides in the endoplasmic reticulum (ER) lumen of all eukaryotic cells (reviewed in Ref. 1). Mammalian PDI was first identified as a catalyst of oxidative protein folding (2), but it is now also known to mediate viral infection (3, 4), antigen processing (5), collagen assembly (6), and ER-associated degradation (79). To participate in this variety of cellular processes, PDI performs multiple activities. For example, during oxidative protein folding, PDI catalyzes the oxidation and isomerization of disulfide bonds and induces conformational changes in non-native polypeptides (10). Independently of redox chemistry, PDI is a molecular chaperone, binding polypeptides to prevent their aggregation (1113). PDI also acts as a structural subunit of the prolyl 4-hydroxylase (P4H) and microsomal triglyceride transfer protein complexes; however, this function is similar to its chaperone activity (1419). In contrast to its protein folding activities, PDI unfolds the catalytic A1 polypeptide of cholera toxin (CTA1) in preparation for the retrotranslocation of the toxin from the ER lumen into the cytosol (8, 20).Cholera toxin (CT) is a pathogenic factor that causes secretory diarrhea in animals (reviewed in Ref. 21). The holotoxin includes a single catalytic A subunit (CTA) and a homopentameric B subunit (CTB) joined noncovalently (22). Upon secretion from the bacterium Vibrio cholerae, CTA is cleaved into the A1 and A2 polypeptides, which are joined by a disulfide bond and noncovalent interactions (22, 23). To intoxicate a cell, CTB binds the ganglioside GM1 on the surface of the cell, and the holotoxin is transported in a retrograde manner to the ER lumen (24). In the ER, CTA is reduced to generate CTA1, and PDI unfolds and dissociates CTA1 from the holotoxin (20). The unfolded toxin is subsequently transported across the ER membrane (25, 26). Upon reaching the cytosol, CTA1 refolds and induces toxicity (27, 28).We showed previously that PDI acts as a redox-dependent chaperone to unfold CTA1 (20). In the reduced state of PDI, it binds and unfolds the toxin. Subsequent oxidation of PDI by ER oxidase 1 causes PDI to release unfolded CTA1 (25). Aside from this information, nothing is known about the mechanism of the unfolding activity of PDI.PDI is a modular protein comprising two a-type thioredoxin-like domains (a and a′), two b-type thioredoxin-like domains (b and b′), a flexible linker (x), and an extended C-terminal domain (c) arranged as abbxac (2931). The a-type domains are characterized by the presence of the catalytic sequence CXXC and are therefore redox-active, whereas the b-type domains lack this sequence and are redox-inactive (32). The thioredoxin-like domains of PDI differ from each other in primary structure despite having a common fold. The crystal structure of yeast PDI shows the bb′ domains form a rigid base from which the a-type domains extend like flexible arms (33, 34). This base is thought to be the core of a substrate-binding groove formed by all four thioredoxin-like domains (30, 33, 35).To understand the mechanism of the unfoldase activity of PDI, we analyzed the contribution of each domain to the ability of PDI to bind and unfold CTA1 using recombinant PDI fragments. Unfolded CTA1 was detected by an established in vitro trypsin sensitivity assay that relies on tryptic cleavage sites hidden in the folded toxin to be exposed in the unfolded toxin (20). Because CTA1 likely mimics a misfolded host cell protein for its recognition and unfolding by PDI (22, 36, 37), this study has implications for how PDI unfolds endogenous misfolded proteins in preparation for their retrotranslocation and subsequent ER-associated degradation.There are nearly 20 mammalian PDI-like proteins, characterized by the presence of one or more thioredoxin-like domains and ER localization (reviewed in Refs. 38, 39). We previously demonstrated that two PDI-like proteins, ERp72 and ERp57, do not facilitate CTA1 retrotranslocation (8). In contrast to PDI, ERp72 retains CTA1 in the ER and either stabilizes its native conformation or renders it more compact (8). To understand how these structurally homologous proteins are functionally unique, we tested whether the various thioredoxin-like domains of ERp57 and ERp72 could functionally replace the corresponding PDI domains to unfold CTA1. Thus, in addition to suggesting a general mechanism for the unfoldase activity of PDI, our data indicate functional similarities and differences among thioredoxin-like domains of PDI family proteins.  相似文献   

17.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

18.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  相似文献   

19.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

20.
Protein kinase D (PKD) is a serine/threonine protein kinase rapidly activated by G protein-coupled receptor (GPCR) agonists via a protein kinase C (PKC)-dependent pathway. Recently, PKD has been implicated in the regulation of long term cellular activities, but little is known about the mechanism(s) of sustained PKD activation. Here, we show that cell treatment with the preferential PKC inhibitors GF 109203X or Gö 6983 blocked rapid (1–5-min) PKD activation induced by bombesin stimulation, but this inhibition was greatly diminished at later times of bombesin stimulation (e.g. 45 min). These results imply that GPCR-induced PKD activation is mediated by early PKC-dependent and late PKC-independent mechanisms. Western blot analysis with site-specific antibodies that detect the phosphorylated state of the activation loop residues Ser744 and Ser748 revealed striking PKC-independent phosphorylation of Ser748 as well as Ser744 phosphorylation that remained predominantly but not completely PKC-dependent at later times of bombesin or vasopressin stimulation (20–90 min). To determine the mechanisms involved, we examined activation loop phosphorylation in a set of PKD mutants, including kinase-deficient, constitutively activated, and PKD forms in which the activation loop residues were substituted for alanine. Our results show that PKC-dependent phosphorylation of the activation loop Ser744 and Ser748 is the primary mechanism involved in early phase PKD activation, whereas PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies identify a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.A rapid increase in the synthesis of lipid-derived second messengers with subsequent activation of protein phosphorylation cascades has emerged as a fundamental signal transduction mechanism triggered by multiple extracellular stimuli, including hormones, neurotransmitters, chemokines, and growth factors (1). Many of these agonists bind to G protein-coupled receptors (GPCRs),4 activate heterotrimeric G proteins and stimulate isoforms of the phospholipase C family, including β, γ, δ, and ε (reviewed in Refs. 1 and 2). Activated phospholipase Cs catalyze the hydrolysis of phosphatidylinositol 4,5-bisphosphate to produce the second messengers inositol 1,4,5-trisphosphate and diacylglycerol (DAG). Inositol 1,4,5-trisphosphate mobilizes Ca2+ from intracellular stores (3, 4) whereas DAG directly activates the classic (α, β, and γ) and novel (δ, ε, η, and θ) isoforms of PKC (57). Although it is increasingly recognized that each PKC isozyme has specific functions in vivo (58), the mechanisms by which PKC-mediated signals are propagated to critical downstream targets remain incompletely defined.PKD, also known initially as PKCμ (9, 10), and two recently identified serine protein kinases termed PKD2 (11) and PKCν/PKD3 (12, 13), which are similar in overall structure and primary amino acid sequence to PKD (14), constitute a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase group (15) and separate from the previously identified PKCs (14). Salient features of PKD structure include an N-terminal regulatory region containing a tandem repeat of cysteine-rich zinc finger-like motifs (termed the cysteine-rich domain) that confers high affinity binding to phorbol esters and DAG (9, 16, 17), followed by a pleckstrin homology (PH) domain that negatively regulates catalytic activity (18, 19). The C-terminal region of the PKDs contains its catalytic domain, which is distantly related to Ca2+-regulated kinases.In unstimulated cells, PKD is in a state of low kinase catalytic activity maintained by the N-terminal domain, which represses the catalytic activity of the enzyme by autoinhibition. Consistent with this model, deletions or single amino acid substitutions in the PH domain result in constitutive kinase activity (1820). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (21). In response to cellular stimuli, PKD is converted from a low activity form into a persistently active form that is retained during isolation from cells, as shown by in vitro kinase assays performed in the absence of lipid co-activators (21, 22). PKD activation has been demonstrated in response to engagement of specific GPCRs either by regulatory peptides (2330) or lysophosphatidic acid (27, 31, 32); signaling through Gq, G12, Gi, and Rho (27, 3134); activation of receptor tyrosine kinases, such as the platelet-derived growth factor receptor (23, 35, 36); cross-linking of B-cell receptor and T-cell receptor in B and T lymphocytes, respectively (3740); and oxidative stress (4144).Throughout these studies, multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF 109203X or Gö 6983) that do not directly inhibit PKD catalytic activity (21, 22), implying that PKD activation in intact cells is mediated, directly or indirectly, through PKCs. In line with this conclusion, cotransfection of PKD with active mutant forms of “novel” PKCs (PKCs δ, ε, η, and θ) resulted in robust PKD activation in the absence of cell stimulation (21, 4446). Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade in response to multiple GPCR agonists in a broad range of cell types, including normal and cancer cells (reviewed in Ref. 14). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as the activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (reviewed in Ref. 14). Collectively, these findings demonstrated the existence of rapidly activated PKC-PKD protein kinase cascade(s) and raised the possibility that some PKC-dependent biological responses involve PKD acting as a downstream effector.PKD has been reported recently to mediate several important cellular activities and processes, including signal transduction (30, 4749), chromatin modification (50), Golgi organization and function (51, 52), c-Jun function (47, 53, 54), NFκB-mediated gene expression (43, 55, 56), and cell survival, migration, and differentiation and DNA synthesis and proliferation (reviewed in Ref. 14). Thus, mounting evidence indicates that PKD has a remarkable diversity of both its signal generation and distribution and its potential for complex regulatory interactions with multiple downstream pathways, leading to multiple responses, including long term cellular events. Despite increasing recognition of its importance, very little is known about the mechanism(s) of sustained PKD activation as opposed to the well documented rapid, PKC-dependent PKD activation.The results presented here demonstrate that prolonged GPCR-induced PKD activation is mediated by sequential PKC-dependent and PKC-independent phases of regulation. We report here, for the first time, that PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies expand previous models of PKD regulation by identifying a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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