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.
Ras leads an important signaling pathway that is deregulated in neurofibromatosis type 1 and malignant peripheral nerve sheath tumor (MPNST). In this study, we show that overactivation of Ras and many of its downstream effectors occurred in only a fraction of MPNST cell lines. RalA, however, was overactivated in all MPNST cells and tumor samples compared to nontransformed Schwann cells. Silencing Ral or inhibiting it with a dominant-negative Ral (Ral S28N) caused a significant reduction in proliferation, invasiveness, and in vivo tumorigenicity of MPNST cells. Silencing Ral also reduced the expression of epithelial mesenchymal transition markers. Expression of the NF1-GTPase-related domain (NF1-GRD) diminished the levels of Ral activation, implicating a role for neurofibromin in regulating RalA activation. NF1-GRD treatment caused a significant decrease in proliferation, invasiveness, and cell cycle progression, but cell death increased. We propose Ral overactivation as a novel cell signaling abnormality in MPNST that leads to important biological outcomes with translational ramifications.The Ras family of guanine-nucleotide bound proteins exerts a fundamental role in cell biology and constitutes an important area of cancer research due to its significant involvement in the development and progression of malignancies (8, 10, 17, 18, 32). Ras-like (Ral) proteins are crucial members of this family and have been shown to play a pivotal role in human tumors (7, 28, 41, 66, 70). Because Ral guanine nucleotide exchange factors (Ral-GEFs) are direct effectors of Ras, the Ral signaling pathway has been traditionally considered a Ras-effector pathway. Activation of Ras (in resemblance to Ral) is regulated by two classes of proteins: Ras-GEFs (e.g., SOS) and Ras- GTPase activating proteins (Ras-GAPs such as neurofibromin). The latter induces hydrolysis of Ras from the active (GTP) form to the inactive (GDP) form (13). Ral-GEFs include two main groups: the proteins that are stimulated by Ras because of their carboxy-terminal Ras binding domain (RalGDS, RGL1, and RGL2) and the proteins that are activated by substrates of PI3K through a pleckstrin homology domain on their C-terminal (RALGPS1 and RALGPS2) (19). Although highly similar to Ras, Ral proteins (RalA and RalB) involve a series of distinctly different effectors that influence gene expression and translation through interaction with ZO-1-associated nucleic acid binding protein (ZONAB) and RalA binding protein 1 (RalBP1) (11, 23, 33). RalB directly interacts with the SEC5 subunit of exocyst to facilitate the host defense response (48, 58).In addition to overactivation of GEFs, inactivation of GAPs is another mechanism for overactivation of GTP-bound proteins. The lack of neurofibromin (encoded by NF1 on human chromosome 17q11.2), a Ras-GAP protein, is the main molecular event in neurofibromatosis type 1 (NF-1), an autosomal-dominant human genetic disease occurring in approximately 1 in 2,500 to 3,500 births (22, 27, 42). One of the main tumor-causing effects of inactivating mutations in the tumor suppressor NF1 gene is postulated to be the subsequent activation of Ras (3, 29, 53, 57, 69). With two main functional domains, SEC14 and Ras-GAP, neurofibromin is best known for its Ras-GAP function. Although the yeast SEC14p is shown to be involved in regulating intracellular proteins and lipid trafficking, the function of its homologous domain in neurofibromin is unknown (49, 62). Although neurofibromas are the most common tumors in NF-1, 10% of patients with plexiform develop malignant peripheral nerve sheath tumors (MPNSTs), which are typically high grade and often fatal (21, 34, 65).The molecular events involved in the malignant transformation of benign neurofibromas to MPNST are poorly defined. Usually arising in the third through sixth decades of life, these tumors are composed of tightly packed hyperchromatic spindle-shaped cells with frequent mitotic figures. Inactivation of both copies of the NF1 gene has been demonstrated in benign human neurofibromas and shown to cause tumors in murine models (56). Loss of heterozygosity of NF1 and p53 has frequently been observed in human MPNST (35, 47, 54). Recombinant mouse strains (NP mice), which harbor inactivated Nf1 and p53 alleles (cis-Nf1+/−:p53+/−), demonstrate the cumulative effects of loss of both Nf1 and p53 genes in the etiology of MPNST (14, 68).In the present study, we show that while both Ras activation and activation of a series of its downstream effector pathways are observed in a fraction of MPNST cells, RalA is activated globally in all studied mouse and human MPNST cells and tumor samples. Our results also explain the involvement of this signaling molecule in a series of key biological functions of MPNST cells, as shown in a variety of in vitro assays and an in vivo model of MPNST. Such information may play a role in designing novel therapies for treatment of MPNST or other tumors with overactivation of the Ral pathway.  相似文献   

2.
3.
Atk can be activated by two independent phosphorylation events. Growth factor-dependent phosphorylation of threonine 308 (Akt-308) by phosphatidylinositol 3-kinase-dependent PDK1 leads to activation of mammalian target of rapamycin (mTor) complex 1 (TORC1) and stimulation of protein synthesis. Phosphorylation on serine 473 (Akt-473) is catalyzed by mTor in a second complex (TORC2), and Akt-473 phosphorylates Foxo3a to inhibit apoptosis. Accumulation of both phosphorylated forms of Akt is frequent in cancer, and TORC2 activity is required for progression to prostate cancer with Pten mutation. Here, we link Akt-473 to the Rb1 pathway and show that mTor is overexpressed with loss of the Rb1 family pathway. This leads to constitutive Akt-473 and, in turn, phosphorylation of Foxo3a and resistance to cell adhesion-dependent apoptosis (anoikis). Additionally, Akt-473 accumulation blocks c-Raf activation, thereby preventing downstream Erk activation. This block cannot be overcome by constitutively active Ras, and it also prevents induction of the Arf tumor suppressor by Ras. These studies link inactivation of the Rb1 pathway, a hallmark of cancer, to accumulation of Akt-473, resistance to anoikis, and a block in c-Raf/Erk activation.Binding of growth factors to their cell surface receptors activates the Ras family of GTP-binding proteins, which in turn classically activates a cell proliferation signaling cascade, including c-Raf and Erk (27, 28). Mutations leading to constitutive Ras activation are among the most frequently found in cancer, highlighting the importance of this pathway in cancer initiation. Ras can also activate phosphatidylinositol 3-kinase (PI3K), which in turn leads to phosphorylation and activation of Akt (reviewed in reference 20). Akt activation can also stimulate cell proliferation by controlling the expression and subcellular localization of cell cycle regulators, such as cyclin D1 and p27Kip1, and it can regulate protein synthesis by controlling the activity of mammalian target of rapamycin (mTor) complex 1 (TORC1); inhibit apoptosis by controlling the subcellular localization of Foxo3a, which transactivates proapoptotic genes; and change cell shape and motility through regulation of the Rac-Rho-Ccd42 pathway (2). Constitutively active Akt is oncogenic, and activated Akt frequently accumulates in cancer cells (9, 22, 29, 38). As opposed to acting in concert, these two Ras-activated pathways appear to function separately or even antagonistically in at least some cells, because Akt can block activation of c-Raf, which in turn prevents Erk activation (20). Indeed, recent results suggest that the two pathways may function sequentially during tumor progression. While Ras-mediated activation of the MAP kinase/Erk pathway is important in tumor initiation, as tumorigenesis progresses, the major role of Ras activation may become the PI3K/Akt pathway (22).Akt is phosphorylated by two different kinases. PDK1 is activated via PI3K in response to growth factors, and it phosphorylates Akt on threonine 308 in the A loop (Akt-308) (1). The Pten phosphatase is negatively related to PI3K activity, and mutations in Pten are common in cancer (4). Akt-308 phosphorylates TSC2, thereby blocking the GAP activity of TSC1/TSC2 toward Rheb, whose GTPase activity is required for TORC1 function. mTor, in the context of TORC1, phosphorylates and activates the AGC kinase family member S6 to regulate protein synthesis, thereby linking growth factor stimulation to protein synthesis (5, 14, 36). Mutations in Ras and Pten are mutually exclusive in some tumors, and lack of Erk activation is an adverse prognostic factor in melanoma (4, 19, 21, 26, 38). Taken together, these results suggest that activation of PI3K and, in turn, Akt is a critical function of Ras mutation in some tumors.mTor forms a second complex (TORC2) with Sin1, Rictor, and mLST8 (24, 30), and mTor in the context of TORC2 phosphorylates Akt (which, like S6, is also an AGC family member) on serine 473 in the hydrophobic motif (11, 18, 32, 37). Knocking out components of TORC2 prevents Akt-473 without affecting Akt-308 (11, 18, 32, 37). Thus, Akt-473 and Akt-308 can be independently catalyzed. TORC2 also phosphorylates Akt and another AGC family member, protein kinase C, on the turn motif, and this phosphorylation is important for protein folding/stability (8, 17).Akt-308 and Akt-473 appear to be directed at distinct cellular pathways. Akt-308 is essential for regulation of TORC1 and thus protein synthesis, whereas Akt-473 is important for phosphorylation of Foxo3a and prevention of its translocation to the nucleus, where it activates apoptotic genes (16, 18). Accordingly, Akt-473 is associated with resistance of cells to stress and to anoikis (survival in the absence of normal matrix contact) (10). Although Akt-473 has been associated with cytoskeletal changes, no obvious effect of loss of TORC2 components and thus Akt-473 on the actin cytoskeleton or on cell proliferation was seen in culture (18), suggesting that these activities may be linked to Akt-308. Both Akt-308 and Akt-473 frequently accumulate in cancer cells, consistent with increased proliferation and protein synthesis and increased motility and survival as cells lose their normal matrix contacts in the forming tumor and during metastasis (9, 29). Such accumulation of Akt-308 can be facilitated by mutations, such as Ras, leading to constitutive activation of growth factor signaling or to mutation or epigenetic silencing of the Pten phosphatase, which inhibits PI3K activity (4, 9). However, less is known about how TORC2 is regulated or why Akt-473 accumulates in cancer.One target of the increase in cyclin-dependent kinase (cdk) activity resulting from Erk-dependent induction of genes such as the one encoding cyclin D1 is the Rb1 pathway, whose family members are hyperphosphorylated and inactivated by the resulting accumulation of cdk activity (13, 25). While Rb1 is mutated in some cancers (e.g., retinoblastoma, small-cell lung cancer, and osteosarcoma), more frequently the pathway, including all three Rb1 family members, is inactivated through hyperphosphorylation resulting from mutation or epigenetic silencing of a cdk inhibitor (13). Interestingly, mouse embryo fibroblasts (MEFs) mutated for only one of the three family members, Rb1, no longer require Erk signaling for proliferation in culture, but they do require Akt activity (6). Additionally, expression of the adenoviral oncoprotein E1a in cells blocks Erk activation, and this effect was eliminated with a point mutation in E1a that prevented its binding and by inhibition of the Rb1 family (3). These results suggest that even partial inactivation of the Rb1 pathway (e.g., mutation of Rb1) eliminates a requirement for Erk signaling, but further, the Rb1 pathway seems to be required for cells to activate Erk. As a model for Rb1 pathway inactivation in cancer, we examined the effects of mutation of the three Rb1 family members on Akt and Erk signaling pathways in MEFs.  相似文献   

4.
The hypoxia-regulated alternative TrkAIII splice variant expressed by human neuroblastomas exhibits oncogenic potential, driven by in-frame exon 6 and 7 alternative splicing, leading to omission of the receptor extracellular immunoglobulin C1 domain and several N-glycosylation sites. Here, we show that the TrkAIII oncogene promotes genetic instability by interacting with and exhibiting catalytic activity at the centrosome. This function depends upon intracellular TrkAIII accumulation and spontaneous interphase-restricted activation, in cytoplasmic tyrosine kinase (tk) domain orientation, predominantly within structures that closely associate with the fully assembled endoplasmic reticulum intermediate compartment and Golgi network. This facilitates TrkAIII tk-mediated binding of γ-tubulin, which is regulated by endogenous protein tyrosine phosphatases and geldanamycin-sensitive interaction with Hsp90, paving the way for TrkAIII recruitment to the centrosome. At the centrosome, TrkAIII differentially phosphorylates several centrosome-associated components, increases centrosome interaction with polo kinase 4, and decreases centrosome interaction with separase, the net results of which are centrosome amplification and increased genetic instability. The data characterize TrkAIII as a novel internal membrane-associated centrosome kinase, unveiling an important alternative mechanism to “classical” cell surface oncogenic receptor tk signaling through which stress-regulated alternative TrkAIII splicing influences the oncogenic process.Alternative splicing is fundamental for differential protein expression from the same gene and not only increases the proteomic complexity of higher organisms (29) but is also involved in cancer pathogenesis, activating several oncogenes and inactivating several oncosuppressors (17).The neurotrophin receptor tropomyosin-related kinase A (TrkA) is among the proto-oncogenes activated by alternative splicing, with a novel hypoxia-regulated oncogenic alternative TrkAIII splice variant recently identified in advanced-stage human neuroblastomas (NB) and primary glioblastomas (44, 45). In contrast to wild-type TrkAI/TrkAII, the expression of which is associated with better prognosis for NB, induces NB cell differentiation, exhibits a tumor suppressor function in NB models in vivo (9, 19, 22, 30, 44, 45), and may regulate both spontaneous and therapy-induced NB regression (30), TrkAIII is expressed by more-advanced-stage NB and exhibits oncogenic activity in NB models (44, 45). This has challenged the hypothesis of an exclusively tumor-suppressing function for TrkA in NB by providing a way through which tumor-suppressing signals from TrkA can be converted to oncogenic signals from TrkAIII during tumor progression.The oncogenic potential of TrkAIII, characterized by NIH 3T3 cell transforming and NB xenograft primary and metastatic tumor-promoting activity (44, 45), is driven by in-frame alternative splicing of exons 6 and 7. This results in the omission of the receptor extracellular immunoglobulin C1 (Ig C1) Ig-like domain and several N-glycosylation sites important in regulating TrkA cell surface expression and preventing ligand-independent activation (2, 44, 45, 48). As a consequence, and unlike TrkAI and TrkAII, TrkAIII is not expressed at the cell surface but accumulates in the intracellular membrane compartment, within which it exhibits spontaneous tyrosine kinase (tk) and phosphoinositol-3 kinase (PI3K) activity and induces chronic signaling through PI3K/Akt/NF-κB but not Ras/mitogen-activated protein kinase (MAPK), inducing a more stress-resistant, angiogenic, and tumorigenic NB cell phenotype (44, 45). This differs from ligand-activated cell surface TrkA, which signals transiently through Ras/MAPK in NB cells to induce differentiation and a less angiogenic and tumorigenic NB cell phenotype (9, 19, 22, 30, 44, 45). This difference in signaling provides a potential basis for the opposing tumor-suppressing and oncogenic effects of alternative TrkA splice variants, which may not only depend upon the dislocation of TrkAIII from cell surface caveolae, which are the sites of TrkAI expression and Ras/MAPK signal initiation (45, 48), but also TrkAIII-associated PI3K activity below the Ras/MAPK activation threshold and/or TrkAIII-associated PI3K antagonism of Raf/MEK/extracellular signal-regulated kinase signaling (44). Signal transduction from intracellular TrkAIII bears close resemblance to the transient signaling through PI3K/Akt/NF-κB but not Ras/MAPK induced by A2a adenosine receptor/c-Src-mediated transactivation of immature TrkAI within the Golgi network (GN) (37), suggesting that the intracellular localization of TrkAIII is a critical determinant of both differential signaling and oncogenic potential.Intracellular nonnuclear membranes are separated into the endoplasmic reticulum (ER), ER-GN intermediate compartment (ERGIC), GN, and transport vesicles, which assemble around, integrate, and interact with the centrosome (38). The centrosome, comprised of two centrioles embedded within a pericentriolar matrix of over 100 proteins, including γ-tubulin, acts as the major microtubule-organizing center and orchestrates the assembly, organization, and integration of the ER, ERGIC, GN, and associated vesicles (38). The centrosome also maintains genomic integrity by duplicating once per cell cycle S phase, ensuring bipolar mitotic spindle formation, accurate chromosome segregation, and the inheritance of a single centrosome by each daughter cell (23).Centrosome duplication is tightly regulated by protein kinases Aurora-A and -B, polo kinases 1 and 4 (Plk-1 and Plk-4), Cdk2, PI3K, Zyg-1, Syk, Nek2, regulators Pin-1 and separase, and related phosphatases (11, 12, 18, 31, 42, 47, 53). The deregulation of centrosome duplication leads to centrosome amplification and subsequently to aberrant mitotic spindle formation, which promotes aneuploidy and polyploidy. These manifestations of genetic instability represent hallmarks of malignancy and drive tumor progression by promoting a more malignant phenotype (11, 12, 26, 35, 49). Centrosome amplification and subsequent genetic instability are induced by kinases that target the centrosome, the loss of centrosome-associated kinase inhibitors, altered levels of centrosome-associated regulators, and oncosuppressor inactivation (6, 7, 11, 12, 35, 40, 42, 49, 53).The localization of TrkAIII to internal membranes, as a prerequisite for oncogenic activity (44, 45), makes identification of the membrane context within which TrkAIII exhibits activity and novel substrate interactions within this environment important in elucidating how TrkAIII exerts oncogenic potential. In the present study, we unveil a novel oncogenic mechanism for TrkAIII by demonstrating that TrkAIII activation within membranes that associate closely with the assembled ER/ERGIC/GN facilitates recruitment to the centrosome, results in centrosome amplification, and promotes genetic instability.  相似文献   

5.
6.
7.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

8.
9.
KSR1 is a mitogen-activated protein (MAP) kinase scaffold that enhances the activation of the MAP kinase extracellular signal-regulated kinase (ERK). The function of KSR1 in NK cell function is not known. Here we show that KSR1 is required for efficient NK-mediated cytolysis and polarization of cytolytic granules. Single-cell analysis showed that ERK is activated in an all-or-none fashion in both wild-type and KSR1-deficient cells. In the absence of KSR1, however, the efficiency of ERK activation is attenuated. Imaging studies showed that KSR1 is recruited to the immunological synapse during T-cell activation and that membrane recruitment of KSR1 is required for recruitment of active ERK to the synapse.Kinase suppressor of Ras was originally identified in Drosophila melanogaster (53) and Caenorhabditis elegans (19, 32, 52) as a positive regulator of the extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase signaling pathway. It is thought to function as a MAP kinase scaffold because it can bind to Raf, MEK, and ERK (18, 19, 27, 28, 44, 59). While the exact function of KSR is unknown, preassembling the three components of the ERK MAP kinase cascade could function to enhance the efficiency of ERK activation, potentially regulate the subcellular location of ERK activation, and promote access to specific subcellular substrates (16, 45, 46).While only one isoform of KSR is expressed in Drosophila (53), two KSR isoforms have been identified in C. elegans (19, 32, 52) and most higher organisms. They are referred to as KSR1 and KSR2 (32, 43). While KSR1 mRNA and protein are detectable in a wide variety of cells and tissues, including brain, thymus, and muscle (10, 11, 29), little is known about the expression pattern of KSR2.We previously reported the phenotype of KSR1-deficient mice (30). These mice are born at Mendelian ratios and develop without any obvious defects. Using gel filtration, we showed that KSR1 promotes the formation of large signaling complexes containing KSR1, Raf, MEK, and ERK (30). Using both primary T cells stimulated with antibodies to the T-cell receptor as well as fibroblasts stimulated with growth factors, we showed that KSR1-deficient cells exhibit an attenuation of ERK activation with defects in cell proliferation.Here we explored the role of KSR1 in NK cell-mediated cytolysis. The killing of a target cell by a cytolytic T cell or NK cell is a complicated process that involves cell polarization with microtubule-dependent movement of cytolytic granules to an area that is proximal to the contact surface or immunological synapse (7, 33, 34, 48-50, 54). A variety of different signaling molecules are also involved, including calcium (23), phosphatidylinositol-3,4,5-triphosphate (13, 17), and activation of the ERK MAP kinase (6, 42, 56). Recently, the recruitment of activated ERK to the immunological synapse (IS) has been shown to be a feature of successful killing of a target by cytotoxic T lymphocytes (58).How active ERK is recruited to the synapse is not known. Since KSR1 is known to be recruited to the plasma membrane by Ras activation (24), and since the immunological synapse is one of the major sites of Ras activation (26, 41), it seemed plausible to test the hypothesis that KSR1 recruitment to the plasma membrane functions to recruit ERK to the immunological synapse and facilitate its activation. We found that KSR1 was recruited to the immunological synapse and that KSR1 appeared to be required for the localization of active ERK at the contact site. As KSR1-deficient cells exhibit a defect in killing, this suggests that KSR1 recruitment to the synapse may be important in the cytolytic killing of target cells.  相似文献   

10.
GAP1IP4BP is a member of the GAP1 family of Ras GTPase-activating proteins (GAPs) that includes GAP1m, CAPRI, and RASAL. Composed of a central Ras GAP-related domain (RasGRD), surrounded by amino-terminal C2 domains and a carboxy-terminal PH/Btk domain, these proteins, with the notable exception of GAP1m, possess an unexpected arginine finger-dependent GAP activity on the Ras-related protein Rap1 (S. Kupzig, D. Deaconescu, D. Bouyoucef, S. A. Walker, Q. Liu, C. L. Polte, O. Daumke, T. Ishizaki, P. J. Lockyer, A. Wittinghofer, and P. J. Cullen, J. Biol. Chem. 281:9891-9900, 2006). Here, we have examined the mechanism through which GAP1IP4BP can function as a Rap1 GAP. We show that deletion of domains on either side of the RasGRD, while not affecting Ras GAP activity, do dramatically perturb Rap1 GAP activity. By utilizing GAP1IP4BP/GAP1m chimeras, we establish that although the C2 and PH/Btk domains are required to stabilize the RasGRD, it is this domain which contains the catalytic machinery required for Rap1 GAP activity. Finally, a key residue in Rap1-specific GAPs is a catalytic asparagine, the so-called asparagine thumb. By generating a molecular model describing the predicted Rap1-binding site in the RasGRD of GAP1IP4BP, we show that mutagenesis of individual asparagine or glutamine residues that lie in close proximity to the predicted binding site has no detectable effect on the in vivo Rap1 GAP activity of GAP1IP4BP. In contrast, we present evidence consistent with a model in which the RasGRD of GAP1IP4BP functions to stabilize the switch II region of Rap1, allowing stabilization of the transition state during GTP hydrolysis initiated by the arginine finger.The Ras-like family of small GTPases are ubiquitously expressed, evolutionarily conserved proteins that, by undergoing conformational changes in response to the alternate binding of GDP and GTP, function as binary switches (28, 31, 35). The GDP-bound “off” state and the GTP-bound “on” state recognize distinct effector proteins, thereby allowing the regulation of a variety of downstream signaling events (28, 31, 35). While Ras is the best-known and best-studied Ras-like GTPase, Rap1 has recently attracted considerable attention (reviewed in reference 20).Rap1 was originally identified through its ability, when overexpressed, to reverse the phenotype of K-Ras-transformed NIH 3T3 cells (19). As Ras and Rap1 have very similar effector regions, the ability of Rap1 to reverse the transformed phenotype appeared to arise through an ability to compete with K-Ras effectors. For example, Rap1 binds the Ras effector Raf1 but this does not lead to its activation (11). This is consistent with a simple model in which Rap1 functions as a Ras antagonist (6, 37). However, recent work has challenged this view. Increasing evidence points to Rap1 interacting with its own panel of effectors through which it controls cell-cell adhesion and cell-matrix interactions (reviewed in reference 20).Like that of other GTPases, the activation of Ras and Rap1 is regulated through guanine nucleotide exchange factors, which control activation by stimulating the exchange of GDP for GTP. Inactivation is driven by GTPase-activating proteins (GAPs). These enhance the intrinsic GTPase activity of Ras and Rap1, thereby leading to GTP hydrolysis. A wide variety of guanine nucleotide exchange factors and GAPs specific for these GTPases have been identified (14). Through the arrangement of different modular domains, these proteins are regulated following the activation of cell surface receptors. This occurs either through direct association with the activated receptor or indirectly through second messengers (4, 5, 14, 41).Mammalian proteins capable of functioning as Ras GAPs include NF1 (3, 27, 40), p120GAP (38), the semaphorin 4D receptor plexin-B1 (29), and members of the GAP1 (reviewed in reference 41) and SynGAP (DAB2IP, nGAP, and SynGAP) families (10, 18, 39). These function as Ras GAPs by supplying a catalytic arginine residue—the arginine finger—into the active site of Ras. This stabilizes the transition state of the GTPase reaction, increasing the reaction rate by more than 1,000-fold (1, 33, 34).Rap1 GAPs include Rap GAPs I and II, the SPA-1 family (SPA-1, SPAR, SPAL, and E6TP1), and tuberin (16, 17, 26, 32). Unlike Ras, Rap1 does not possess the catalytic glutamine residue that is critical for GTP hydrolysis in Ras. This fundamental difference means that the mechanisms by which Ras and Rap1 GAPs function are distinct. Rap1 GAPs do not employ a catalytic arginine residue (8, 9); instead, they provide a catalytic asparagine—the asparagine thumb—to stimulate GTP hydrolysis (15). Here the asparagine carboxamide side chain has a function similar to that of the glutamine residue in Ras, stabilizing the position of the nucleophilic water and γ-phosphate in the transition complex (15, 36).Given such distinct catalytic mechanisms, surprisingly, some Ras GAPs, while having no detectable sequence homology with any Rap1 GAPs, are capable of stimulating the GTPase activity of Rap1. The first protein found to display such dual activities was GAP1IP4BP (13) (also known as RASA3, GAPIII, and R-Ras GAP). This is a member of the GAP1 family, which also comprises GAP1m, CAPRI, and RASAL (2, 23-25). These proteins are characterized by a domain architecture comprising amino-terminal tandem C2 domains, a highly conserved central Ras GAP-related domain (RasGRD), and a carboxy-terminal pleckstrin homology (PH) domain that is associated with a Bruton''s tyrosine kinase (Btk) motif (41). Consistent with the presence of the RasGRD, all proteins display Ras GAP activity, although each is differentially regulated following receptor stimulation (41). With the notable exception of GAP1m, all GAP1 proteins also possess efficient Rap1 GAP activity (22). Such dual specificity is not restricted solely to GAP1 proteins. Recently, C2 domain-containing SynGAP—a neuronal Ras GAP—has also been shown to display Rap1 GAP activity (21), an activity that appears to require, alongside the RasGRD, the presence of a single C2 domain (30).Here we have examined the mechanism behind the dual Ras and Rap1 GAP activities of GAP1IP4BP. Through the generation of a series of GAP1IP4BP/GAP1m chimeras, we have established that while the C2 domains of GAP1IP4BP are required to stabilize the RasGRD, these domains do not supply catalytic residues required for Rap1 GAP activity. Rather, the Rap1 GAP catalytic machinery appears to reside solely within the RasGRD. By the site-directed mutagenesis of selected asparagine and glutamine residues within this domain—selected following the generation of a predicted molecular model of the GAP1IP4BP RasGRD-Ras(Rap1) complex—we establish that the ability of GAP1IP4BP to function as a Rap1 GAP does not occur via a mechanism that utilizes a classic asparagine thumb. Rather, we suggest that the GAP1IP4BP RasGRD functions to stabilize the switch II region of Rap1 in a manner that allows a catalytic arginine finger from GAP1IP4BP to drive the hydrolysis of GTP.  相似文献   

11.
12.
The mammalian target of rapamycin (mTOR) is a conserved Ser/Thr kinase that forms two functionally distinct complexes important for nutrient and growth factor signaling. While mTOR complex 1 (mTORC1) regulates mRNA translation and ribosome biogenesis, mTORC2 plays an important role in the phosphorylation and subsequent activation of Akt. Interestingly, mTORC1 negatively regulates Akt activation, but whether mTORC1 signaling directly targets mTORC2 remains unknown. Here we show that growth factors promote the phosphorylation of Rictor (rapamycin-insensitive companion of mTOR), an essential subunit of mTORC2. We found that Rictor phosphorylation requires mTORC1 activity and, more specifically, the p70 ribosomal S6 kinase 1 (S6K1). We identified several phosphorylation sites in Rictor and found that Thr1135 is directly phosphorylated by S6K1 in vitro and in vivo, in a rapamycin-sensitive manner. Phosphorylation of Rictor on Thr1135 did not affect mTORC2 assembly, kinase activity, or cellular localization. However, cells expressing a Rictor T1135A mutant were found to have increased mTORC2-dependent phosphorylation of Akt. In addition, phosphorylation of the Akt substrates FoxO1/3a and glycogen synthase kinase 3α/β (GSK3α/β) was found to be increased in these cells, indicating that S6K1-mediated phosphorylation of Rictor inhibits mTORC2 and Akt signaling. Together, our results uncover a new regulatory link between the two mTOR complexes, whereby Rictor integrates mTORC1-dependent signaling.The mammalian target of rapamycin (mTOR) is an evolutionarily conserved phosphatidylinositol 3-kinase (PI3K)-related Ser/Thr kinase that integrates signals from nutrients, energy sufficiency, and growth factors to regulate cell growth as well as organ and body size in a variety of organisms (reviewed in references 4, 38, 49, and 77). mTOR was discovered as the molecular target of rapamycin, an antifungal agent used clinically as an immunosuppressant and more recently as an anticancer drug (5, 20). Recent evidence indicates that deregulation of the mTOR pathway occurs in a majority of human cancers (12, 18, 25, 46), suggesting that rapamycin analogs may be potent antineoplastic therapeutic agents.mTOR forms two distinct multiprotein complexes, the rapamycin-sensitive and -insensitive mTOR complexes 1 and 2 (mTORC1 and mTORC2), respectively (6, 47). In cells, rapamycin interacts with FKBP12 and targets the FKBP12-rapamycin binding (FRB) domain of mTORC1, thereby inhibiting some of its function (13, 40, 66). mTORC1 is comprised of the mTOR catalytic subunit and four associated proteins, Raptor (regulatory associated protein of mTOR), mLST8 (mammalian lethal with sec13 protein 8), PRAS40 (proline-rich Akt substrate of 40 kDa), and Deptor (28, 43, 44, 47, 59, 73, 74). The small GTPase Rheb (Ras homolog enriched in brain) is a key upstream activator of mTORC1 that is negatively regulated by the tuberous sclerosis complex 1 (TSC1)/TSC2 GTPase-activating protein complex (reviewed in reference 35). mTORC1 is activated by PI3K and Ras signaling through direct phosphorylation and inactivation of TSC2 by Akt, extracellular signal-regulated kinase (ERK), and p90 ribosomal protein S6 kinase (RSK) (11, 37, 48, 53, 63). mTORC1 activity is also regulated at the level of Raptor. Whereas low cellular energy levels negatively regulate mTORC1 activity through phosphorylation of Raptor by AMP-activated protein kinase (AMPK) (27), growth signaling pathways activating the Ras/ERK pathway positively regulate mTORC1 activity through direct phosphorylation of Raptor by RSK (10). More recent evidence has also shown that mTOR itself positively regulates mTORC1 activity by directly phosphorylating Raptor at proline-directed sites (20a, 75). Countertransport of amino acids (55) and amino acid signaling through the Rag GTPases were also shown to regulate mTORC1 activity (45, 65). When activated, mTORC1 phosphorylates two main regulators of mRNA translation and ribosome biogenesis, the AGC (protein kinase A, G, and C) family kinase p70 ribosomal S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), and thus stimulates protein synthesis and cellular growth (50, 60).The second mTOR complex, mTORC2, is comprised of mTOR, Rictor (rapamycin-insensitive companion of mTOR), mSin1 (mammalian stress-activated mitogen-activated protein kinase-interacting protein 1), mLST8, PRR5 (proline-rich region 5), and Deptor (21, 39, 58, 59, 66, 76, 79). Rapamycin does not directly target and inhibit mTORC2, but long-term treatment with this drug was shown to correlate with mTORC2 disassembly and cytoplasmic accumulation of Rictor (21, 39, 62, 79). Whereas mTORC1 regulates hydrophobic motif phosphorylation of S6K1, mTORC2 has been shown to phosphorylate other members of the AGC family of kinases. Biochemical and genetic evidence has demonstrated that mTORC2 phosphorylates Akt at Ser473 (26, 39, 68, 70), thereby contributing to growth factor-mediated Akt activation (6, 7, 52). Deletion or knockdown of the mTORC2 components mTOR, Rictor, mSin1, and mLST8 has a dramatic effect on mTORC2 assembly and Akt phosphorylation at Ser473 (26, 39, 79). mTORC2 was also shown to regulate protein kinase Cα (PKCα) (26, 66) and, more recently, serum- and glucocorticoid-induced protein kinase 1 (SGK1) (4, 22). Recent evidence implicates mTORC2 in the regulation of Akt and PKCα phosphorylation at their turn motifs (19, 36), but whether mTOR directly phosphorylates these sites remains a subject of debate (4).Activation of mTORC1 has been shown to negatively regulate Akt phosphorylation in response to insulin or insulin-like growth factor 1 (IGF1) (reviewed in references 30 and 51). This negative regulation is particularly evident in cell culture models with defects in the TSC1/TSC2 complex, where mTORC1 and S6K1 are constitutively activated. Phosphorylation of insulin receptor substrate-1 (IRS-1) by mTORC1 (72) and its downstream target S6K1 has been shown to decrease its stability and lead to an inability of insulin or IGF1 to activate PI3K and Akt (29, 69). Although the mechanism is unknown, platelet-derived growth factor receptor β (PDGF-Rβ) has been found to be downregulated in TSC1- and TSC2-deficient murine embryonic fibroblasts (MEFs), contributing to a reduction of PI3K signaling (80). Interestingly, inhibition of Akt phosphorylation by mTORC1 has also been observed in the presence of growth factors other than IGF-1, insulin, or PDGF, suggesting that there are other mechanisms by which mTORC1 activation restricts Akt activity in cells (reviewed in references 6 and 31). Recent evidence demonstrates that rapamycin treatment causes a significant increase in Rictor electrophoretic mobility (2, 62), suggesting that phosphorylation of the mTORC2 subunit Rictor may be regulated by mTORC1 or downstream protein kinases.Herein, we demonstrate that Rictor is phosphorylated by S6K1 in response to mTORC1 activation. We demonstrate that Thr1135 is directly phosphorylated by S6K1 and found that a Rictor mutant lacking this phosphorylation site increases Akt phosphorylation induced by growth factor stimulation. Cells expressing the Rictor T1135A mutant were found to have increased Akt signaling to its substrates compared to Rictor wild-type- and T1135D mutant-expressing cells. Together, our results suggest that Rictor integrates mTORC1 signaling via its phosphorylation by S6K1, resulting in the inhibition of mTORC2 and Akt signaling.  相似文献   

13.
RASSF2 is a novel proapoptotic effector of K-Ras. Inhibition of RASSF2 expression enhances the transforming effects of K-Ras, and epigenetic inactivation of RASSF2 is frequently detected in mutant Ras-containing primary tumors. Thus, RASSF2 is implicated as a tumor suppressor whose inactivation facilitates transformation by disconnecting apoptotic responses from Ras. The mechanism of action of RASSF2 is not known. Here we show that RASSF2 forms a direct and endogenous complex with the prostate apoptosis response protein 4 (PAR-4) tumor suppressor. This interaction is regulated by K-Ras and is essential for the full apoptotic effects of PAR-4. RASSF2 is primarily a nuclear protein, and shuttling of PAR-4 from the cytoplasm to the nucleus is essential for its function. We show that RASSF2 modulates the nuclear translocation of PAR-4 in prostate tumor cells, providing a mechanism for its biological effects. Thus, we identify the first tumor suppressor signaling pathway emanating from RASSF2, we identify a novel mode of action of a RASSF protein, and we provide an explanation for the extraordinarily high frequency of RASSF2 inactivation we have observed in primary prostate tumors.Ras oncoproteins regulate a broad range of signaling pathways involved in the control of cell growth and transformation (reviewed in reference 31). Activating mutations in ras genes, primarily K-Ras, are found in approximately 30% of primary human tumors (31). Moreover, hyperactivation of Ras signaling pathways even in the absence of ras mutations has been reported in many tumor types (13). Thus, abnormal activation of Ras signaling appears to be a frequent component of tumor development.Although activated forms of Ras promote growth and transformation, they can also induce apoptotic cell death (10). This is particularly apparent with K-Ras. The main Ras effector proteins identified to date that are implicated in mediating apoptosis are the RASSF proteins (12, 44).The best-characterized member of the RASSF family is RASSF1A. RASSF1A is thought to act as a scaffold protein that may link Ras to multiple tumor suppressor pathways. In particular, RASSF1A has been shown to bind and activate the proapoptotic effectors MstI, a proapoptotic Ste20-related kinase (28), and MOAP-1, resulting in Bax activation (45). Overexpression of RASSF1A promotes apoptosis, and knockdown of RASSF1A impairs the apoptotic activity of activated K-Ras (45). Moreover, deletion of RASSF1A in transgenic mice promotes a modest increase in tumorigenesis (43). Thus, RASSF1A has the potential to be a Ras effector/tumor suppressor.Expression of the RASSF1A protein is frequently lost in primary tumors due to promoter methylation, an epigenetic mechanism of gene silencing that plays a major role in the development of many cancers. Inactivation of RASSF1A expression has been shown to correlate with activation of Ras in tumors, suggesting that loss of RASSF1A-mediated growth-inhibitory signals is essential to subvert Ras apoptotic pathways, facilitating Ras driven tumorigenesis in vivo.RASSF2 is structurally related to RASSF1A and may also serve as a proapoptotic, K-Ras-specific effector. RASSF2 binds to K-Ras in a GTP-dependent manner via the effector domain (47) and can be detected in an endogenous complex with K-Ras (4). Like RASSF1A it is inactivated in a variety of tumors by promoter hypermethylation (7, 16, 24, 25, 27, 32, 36, 47, 48). Overexpression of RASSF2 promotes apoptosis and cell cycle arrest (47). It also inhibits the growth of tumor cells and impairs tumor xenograft formation in nude mice (7, 47). Knockdown of RASSF2 expression by small interfering RNA (siRNA) leads to enhanced growth in soft agar and an enhanced transformation due to activated Ras (1). Thus, like RASSF1A, RASSF2 exhibits the properties of a Ras effector/tumor suppressor (47). However, the mechanism by which RASSF2 promotes cell death and tumor inhibition is completely unknown. It seems likely that it may differ from the mechanisms employed by RASSF1A, a primarily cytoplasmic protein, as RASSF2 localizes mostly to the nucleus (7, 29). As with RASSF1A, RASSF2 has no apparent intrinsic enzyme activity or DNA binding properties, and thus it may interact with other proapoptotic effectors/tumor suppressors to mediate cell death. We have found that prostate apoptosis response protein 4 (PAR-4), a key tumor suppressor in prostate cancer (5), may be one such protein.PAR-4 appears to act at multiple levels that include activating both the FAS- and tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL)-controlled proapoptotic pathways, as well as inhibiting the NF-κB antiapoptotic pathway (5). PAR-4 is of note as it appears to be selective for inducing apoptosis in cancer cells but not in normal or immortalized cells (15). However, not all cancer cells are sensitive to PAR-4-mediated apoptosis. Those cancer cells resistant to PAR-4-induced cell death are resistant to nuclear translocation of PAR-4, a process necessary for the inhibition of NF-κB activity by PAR-4 (15, 23). The domain of PAR-4 responsible for its apoptotic activity has been mapped to the central core region of the protein (15) and confers resistance to tumor formation in vivo (49).In order to determine the mechanism of action of RASSF2, we performed a two-hybrid screen. This screen identified PAR-4 as a direct binding partner of RASSF2. Further experiments confirmed that the interaction could be detected between the endogenous proteins and that it could be enhanced in the presence of activated K-Ras. Downregulation of RASSF2 impaired the ability of PAR-4 to kill cells. Thus, we established a K-Ras-RASSF2-PAR-4 signaling pathway.PAR-4 must be translocated to the nucleus to induce apoptosis (15, 23); however, the mechanism by which this is accomplished is not known. RASSF2 is primarily a nuclear protein that directly binds PAR-4. Here we demonstrate that RASSF2 plays an essential role in the nuclear localization of PAR-4 and that activated K-Ras promotes the nuclear localization of PAR-4 in a RASSF2-dependent manner. Moreover, loss of RASSF2 confers resistance to TRAIL-induced PAR-4 nuclear localization and cell death in prostate cancer cells. As PAR-4 is so important to the development of prostate cancer, we performed an extensive analysis of the frequency of epigenetic inactivation of RASSF2 in primary prostate cancer. We determined that RASSF2 is inactivated in prostate cancer at a higher frequency than in any other cancer type yet investigated. Thus, we have identified the first tumor suppressor signaling pathway emanating from RASSF2 and shown that RASSF2 can link Ras to the key prostate tumor suppressor PAR-4. This may explain the high levels of inactivation observed for RASSF2 in prostate tumors and identifies RASSF2 as a key target for epigenetic therapy in prostate cancer.  相似文献   

14.
15.
16.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

17.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

18.
Protein tyrosine kinase 6 (PTK6) is a nonmyristoylated Src-related intracellular tyrosine kinase. Although not expressed in the normal mammary gland, PTK6 is expressed in a majority of human breast tumors examined, and it has been linked to ErbB receptor signaling and AKT activation. Here we demonstrate that AKT is a direct substrate of PTK6 and that AKT tyrosine residues 315 and 326 are phosphorylated by PTK6. Association of PTK6 with AKT occurs through the SH3 domain of PTK6 and is enhanced through SH2 domain-mediated interactions following tyrosine phosphorylation of AKT. Using Src, Yes, and Fyn null mouse embryonic fibroblasts (SYF cells), we show that PTK6 phosphorylates AKT in a Src family kinase-independent manner. Introduction of PTK6 into SYF cells sensitized these cells to physiological levels of epidermal growth factor (EGF) and increased AKT activation. Stable introduction of active PTK6 into SYF cells also resulted in increased proliferation. Knockdown of PTK6 in the BPH-1 human prostate epithelial cell line led to decreased AKT activation in response to EGF. Our data indicate that in addition to promoting growth factor receptor-mediated activation of AKT, PTK6 can directly activate AKT to promote oncogenic signaling.Protein tyrosine kinase 6 (PTK6; also known as the breast tumor kinase BRK) is an intracellular Src-related tyrosine kinase (9, 48). Human PTK6 was identified in cultured human melanocytes (32) and breast tumor cells (39), while its mouse orthologue was cloned from normal small intestinal epithelial cell RNA (50). Although PTK6 shares overall structural similarity with Src family tyrosine kinases, it lacks an N-terminal myristoylation consensus sequence for membrane targeting (39, 51). As a consequence, PTK6 is localized to different cellular compartments, including the nucleus (14, 15). PTK6 is expressed in normal differentiated epithelial cells of the gastrointestinal tract (34, 42, 51), prostate (14), and skin (51-53). Expression of PTK6 is upregulated in different types of cancers, including breast carcinomas (6, 39, 54), colon cancer (34), ovarian cancer (47), head and neck cancers (33), and metastatic melanoma cells (16). The significance of apparent opposing signaling roles for PTK6 in normal differentiation and cancer is still poorly understood.In human breast tumor cells, PTK6 enhances signaling from members of the ErbB receptor family (10, 29, 30, 36, 40, 49, 54). In the HB4a immortalized human mammary gland luminal epithelial cell line, PTK6 promoted epidermal growth factor (EGF)-induced ErbB3 tyrosine phosphorylation and AKT activation (29). In response to EGF stimulation, PTK6 promoted phosphorylation of the focal adhesion protein paxillin and Rac1-mediated cell migration (10). PTK6 can be activated by the ErbB3 ligand heregulin and promotes activation of extracellular signal-regulated kinase 5 (ERK5) and p38 mitogen-activated protein kinase (MAPK) in breast cancer cells (40). PTK6 can also phosphorylate p190RhoGAP-A and stimulate its activity, leading to RhoA inactivation and Ras activation and thereby promoting EGF-dependent breast cancer cell migration and proliferation (49). Expression of PTK6 has been correlated with ErbB2 expression in human breast cancers (4, 5, 54).AKT (also called protein kinase B) is a serine-threonine kinase that is activated downstream of growth factor receptors (38). It is a key player in signaling pathways that regulate energy metabolism, proliferation, and cell survival (7, 45). Aberrant activation of AKT through diverse mechanisms has been discovered in different cancers (2). AKT activation requires phosphorylation of AKT on threonine residue 308 and serine residue 473. The significance of phosphorylation of AKT on tyrosine residues is less well understood. Src has been shown to phosphorylate AKT on conserved tyrosine residues 315 and 326 near the activation loop (11). Substitution of these two tyrosine residues with phenylalanine abolished AKT kinase activity stimulated by EGF (11). Use of the Src family inhibitor PP2 impaired AKT activation following IGF-1 stimulation of oligodendrocytes (13). The RET/PTC receptor tyrosine kinase that responds to glial cell-line-derived neurotrophic factor also phosphorylated AKT tyrosine residue 315 promoting activation of AKT (28). AKT tyrosine residue 474 was phosphorylated when cells were treated with the tyrosine phosphatase inhibitor pervanadate, and phosphorylation of tyrosine 474 contributed to full activation of AKT (12). Recently, the nonreceptor tyrosine kinase Ack1 was shown to regulate AKT tyrosine phosphorylation and activation (37).Here we show that AKT is a cytoplasmic substrate of the intracellular tyrosine kinase PTK6. We identify the tyrosine residues on AKT that are targeted by PTK6, and we demonstrate that tyrosine phosphorylation plays a role in regulating association between PTK6 and AKT. In addition, we show that PTK6 promotes AKT activation and cell proliferation in a Src-independent manner.  相似文献   

19.
The filovirus VP40 protein is capable of budding from mammalian cells in the form of virus-like particles (VLPs) that are morphologically indistinguishable from infectious virions. Ebola virus VP40 (eVP40) contains well-characterized overlapping L domains, which play a key role in mediating efficient virus egress. L domains represent only one component required for efficient budding and, therefore, there is a need to identify and characterize additional domains important for VP40 function. We demonstrate here that the 96LPLGVA101 sequence of eVP40 and the corresponding 84LPLGIM89 sequence of Marburg virus VP40 (mVP40) are critical for efficient release of VP40 VLPs. Indeed, deletion of these motifs essentially abolished the ability of eVP40 and mVP40 to bud as VLPs. To address the mechanism by which the 96LPLGVA101 motif of eVP40 contributes to egress, a series of point mutations were introduced into this motif. These mutants were then compared to the eVP40 wild type in a VLP budding assay to assess budding competency. Confocal microscopy and gel filtration analyses were performed to assess their pattern of intracellular localization and ability to oligomerize, respectively. Our results show that mutations disrupting the 96LPLGVA101 motif resulted in both altered patterns of intracellular localization and self-assembly compared to wild-type controls. Interestingly, coexpression of either Ebola virus GP-WT or mVP40-WT with eVP40-ΔLPLGVA failed to rescue the budding defective eVP40-ΔLPLGVA mutant into VLPs; however, coexpression of eVP40-WT with mVP40-ΔLPLGIM successfully rescued budding of mVP40-ΔLPLGIM into VLPs at mVP40-WT levels. In sum, our findings implicate the LPLGVA and LPLGIM motifs of eVP40 and mVP40, respectively, as being important for VP40 structure/stability and budding.Ebola and Marburg viruses are members of the family Filoviridae. Filoviruses are filamentous, negative-sense, single-stranded RNA viruses that cause lethal hemorrhagic fevers in both humans and nonhuman primates (5). Filoviruses encode seven viral proteins including: NP (major nucleoprotein), VP35 (phosphoprotein), VP40 (matrix protein), GP (glycoprotein), VP30 (minor nucleoprotein), VP24 (secondary matrix protein), and L (RNA-dependent RNA polymerase) (2, 5, 10, 12, 45). Numerous studies have shown that expression of Ebola virus VP40 (eVP40) alone in mammalian cells leads to the production of virus-like particles (VLPs) with filamentous morphology which is indistinguishable from infectious Ebola virus particles (12, 17, 18, 25, 26, 27, 30, 31, 34, 49). Like many enveloped viruses such as rhabdovirus (11) and arenaviruses (44), Ebola virus encodes late-assembly or L domains, which are sequences required for the membrane fission event that separates viral and cellular membranes to release nascent virion particles (1, 5, 7, 10, 12, 18, 25, 27, 34). Thus far, four classes of L domains have been identified which were defined by their conserved amino acid core sequences: the Pro-Thr/Ser-Ala-Pro (PT/SAP) motif (25, 27), the Pro-Pro-x-Tyr (PPxY) motif (11, 12, 18, 19, 41, 53), the Tyr-x-x-Leu (YxxL) motif (3, 15, 27, 37), and the Phe-Pro-Ile-Val (FPIV) motif (39). Both PTAP and the PPxY motifs are essential for efficient particle release for eVP40 (25, 27, 48, 49), whereas mVP40 contains only a PPxY motif. L domains are believed to act as docking sites for the recruitment of cellular proteins involved in endocytic trafficking and multivesicular body biogenesis to facilitate virus-cell separation (8, 13, 14, 16, 28, 29, 33, 36, 43, 50, 51).In addition to L domains, oligomerization, and plasma-membrane localization of VP40 are two functions of the protein that are critical for efficient budding of VLPs and virions. Specific sequences involved in self-assembly and membrane localization have yet to be defined precisely. However, recent reports have attempted to identify regions of VP40 that are important for its overall function in assembly and budding. For example, the amino acid region 212KLR214 located at the C-terminal region was found to be important for efficient release of eVP40 VLPs, with Leu213 being the most critical (30). Mutation of the 212KLR214 region resulted in altered patterns of cellular localization and oligomerization of eVP40 compared to those of the wild-type genotype (30). In addition, the proline at position 53 was also implicated as being essential for eVP40 VLP release and plasma-membrane localization (54).In a more recent study, a YPLGVG motif within the M protein of Nipah virus (NiV) was shown to be important for stability, membrane binding, and budding of NiV VLPs (35). Whether this NiV M motif represents a new class of L domain remains to be determined. However, it is clear that this YPLGVG motif of NiV M is important for budding, perhaps involving a novel mechanism (35). Our rationale for investigating the corresponding, conserved motifs present within the Ebola and Marburg virus VP40 proteins was based primarily on these findings with NiV. In addition, Ebola virus VP40 motif maps close to the hinge region separating the N- and C-terminal domains of VP40 (4). Thus, the 96LPLGVA101 motif of eVP40 is predicted to be important for the overall stability and function of VP40 during egress. Findings presented here indicate that disruption of these filovirus VP40 motifs results in a severe defect in VLP budding, due in part to impairment in overall VP40 structure, stability and/or intracellular localization.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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