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.
The mammalian target of rapamycin (mTOR) complex 1 (mTORC1) functions as a rapamycin-sensitive environmental sensor that promotes cellular biosynthetic processes in response to growth factors and nutrients. While diverse physiological stimuli modulate mTORC1 signaling, the direct biochemical mechanisms underlying mTORC1 regulation remain poorly defined. Indeed, while three mTOR phosphorylation sites have been reported, a functional role for site-specific mTOR phosphorylation has not been demonstrated. Here we identify a new site of mTOR phosphorylation (S1261) by tandem mass spectrometry and demonstrate that insulin-phosphatidylinositol 3-kinase signaling promotes mTOR S1261 phosphorylation in both mTORC1 and mTORC2. Here we focus on mTORC1 and show that TSC/Rheb signaling promotes mTOR S1261 phosphorylation in an amino acid-dependent, rapamycin-insensitive, and autophosphorylation-independent manner. Our data reveal a functional role for mTOR S1261 phosphorylation in mTORC1 action, as S1261 phosphorylation promotes mTORC1-mediated substrate phosphorylation (e.g., p70 ribosomal protein S6 kinase 1 [S6K1] and eukaryotic initiation factor 4E binding protein 1) and cell growth to increased cell size. Moreover, Rheb-driven mTOR S2481 autophosphorylation and S6K1 phosphorylation require S1261 phosphorylation. These data provide the first evidence that site-specific mTOR phosphorylation regulates mTORC1 function and suggest a model whereby insulin-stimulated mTOR S1261 phosphorylation promotes mTORC1 autokinase activity, substrate phosphorylation, and cell growth.The mammalian target of rapamycin (mTOR), an evolutionarily conserved serine/threonine protein kinase, senses and integrates signals from diverse environmental cues (14, 31, 50, 74). mTOR associates with different partner proteins to form functionally distinct signaling complexes (4). The immunosuppressive drug rapamycin acutely inhibits signaling by mTOR complex 1 (mTORC1) (22), which contains mTOR, mLST8/GβL, raptor, and PRAS40 (24, 33, 34, 54, 67). Rapamycin fails to acutely inhibit signaling by mTORC2, which contains mTOR, mLST8/GβL, rictor, mSin1, and PRR5/Protor (18, 32, 47, 55, 73, 76). mTORC1 promotes various biosynthetic processes, including protein synthesis, cell growth (an increase in cell mass and size), and cell proliferation (an increase in cell number) (14, 40, 74). During growth factor (e.g., insulin) and nutrient (e.g., amino acids and glucose) sufficiency, mTORC1 phosphorylates the translational regulators p70 ribosomal protein S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E binding protein 1 (4EBP1) to coordinately upregulate protein biosynthesis (40). Both S6K1 and 4EBP1 contain a TOR signaling motif, which mediates their interaction with raptor and thus facilitates their recruitment to the mTOR kinase (10, 44, 57, 58). In addition to regulating protein synthesis, mTORC1-mediated phosphorylation of S6K1 and 4EBP also promotes cell growth and cell cycle progression (15, 16). While more recently identified and thus less well characterized than mTORC1, mTORC2 mediates the phosphorylation of AGC kinase family members (e.g., Akt [also known as protein kinase B, PKB], PKCα, and SGK1) on their hydrophobic motifs and modulates the organization of the actin cytoskeleton (20, 26, 32, 55, 56).The insulin pathway represents the best-characterized activator of mTORC1 signaling to date, and thus many signaling intermediates that link insulin receptor activation to mTORC1 have been identified (12, 31). Complementary work using Drosophila melanogaster genetics and mammalian cell culture identified TSC1 (hamartin) and TSC2 (tuberin) as upstream negative regulators of mTORC1 (27). Inactivation of either the TSC1 or TSC2 genes, whose protein products heterodimerize to form a tumor suppressor complex, causes the development of benign tumors in diverse organs in both humans and rodents, a disease known as tuberous sclerosis complex (TSC) (36). TSC2 contains a GTPase-activating protein domain that acts on Rheb, a Ras-like GTP binding protein that activates mTORC1 (27). Thus, in TSC-deficient cells, constitutive Rheb-GTP leads to chronically high mTORC1 signaling. While the mechanism by which Rheb-GTP activates mTORC1 remains incompletely understood, Rheb coimmunoprecipitates with mTOR and directly activates mTORC1 kinase activity in vivo and in vitro when GTP bound (2, 38, 54). Rheb has been reported to augment the activity of PLD1, an enzyme that catalyzes the production of the lipid second messenger phosphatidic acid, which contributes to the mitogenic activation of mTORC1 signaling (13, 62). Additionally, Rheb-GTP was reported to induce the dissociation of the endogenous mTOR inhibitor FKBP38 (3), although aspects of this model have been questioned (72). Insulin/phosphatidylinositol 3-kinase (PI3K) signaling reduces the inhibitory effect of TSC on mTORC1 via Akt-mediated phosphorylation of TSC2 (29, 42, 64). Additionally, Ras-regulated signaling via mitogen-activated protein kinase (MAPK) and RSK also inhibits TSC via PI3K/Akt-independent phosphorylation of TSC2 (39, 51, 63). In contrast, glucose deprivation enhances TSC''s inhibitory effect on mTORC1 signaling via AMP-activated protein kinase (AMPK)-mediated phosphorylation of TSC2 (on different sites) (30). Thus, TSC functions as a central nexus of diverse physiological signals to fine-tune mTORC1 signaling depending on environmental conditions (27). While the mechanism by which amino acids promote mTORC1 signaling has remained elusive, compelling new data reveal that the Rag GTPases link amino acid sensing to mTORC1 activation (35, 52, 53). During amino acid sufficiency, GTP-bound Rag heterodimers bind raptor and recruit mTORC1 to an endomembrane compartment that contains the mTORC1 activator Rheb; thus, amino acid sufficiency may function to prime mTORC1 for subsequent growth factor-mediated activation via a dynamic subcellular redistribution mechanism (52).Despite the well-characterized regulation of mTORC1 signaling by growth factors (e.g., insulin), nutrients (e.g., amino acids and glucose), and cellular stress (e.g., hypoxia) and the identification of numerous signaling mediators of these pathways, the direct molecular mechanisms by which cellular signals modulate mTORC1 action remain obscure (31). While three phosphorylation sites (P-sites) on mTOR have been reported to date (T2446, S2448, and S2481), no function has yet been ascribed to any site (7, 43, 49, 59). Here we identify S1261 as a novel mTOR phosphorylation site in vivo in cultured mammalian cells and provide the first evidence that site-specific mTOR phosphorylation regulates mTORC1 function. We show that insulin signals via the PI3K/TSC/Rheb pathway in an amino acid-dependent and rapamycin-insensitive manner to promote mTOR S1261 phosphorylation, which regulates mTORC1 autokinase activity, biochemical signaling to downstream substrates, and cell growth to increased cell size, a major cellular function of mTORC1. Elucidation of the molecular mechanisms underlying mTORC1 regulation will enable us to better understand how mTORC1 senses environmental stimuli to control cellular physiology. As aberrantly upregulated mTORC1 signaling likely contributes to cancer, insulin-resistant diabetes, and cardiovascular diseases, understanding mTORC1 regulation may aid in the development of novel therapeutics for these prevalent human diseases (11, 21, 28).  相似文献   

2.
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   

3.
4.
5.
The mammalian target of rapamycin (mTOR) interacts with raptor to form the protein complex mTORC1 (mTOR complex 1), which plays a central role in the regulation of cell growth in response to environmental cues. Given that glucose is a primary fuel source and a biosynthetic precursor, how mTORC1 signaling is coordinated with glucose metabolism has been an important question. Here, we found that the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) binds Rheb and inhibits mTORC1 signaling. Under low-glucose conditions, GAPDH prevents Rheb from binding to mTOR and thereby inhibits mTORC1 signaling. High glycolytic flux suppresses the interaction between GAPDH and Rheb and thus allows Rheb to activate mTORC1. Silencing of GAPDH or blocking of the Rheb-GAPDH interaction desensitizes mTORC1 signaling to changes in the level of glucose. The GAPDH-dependent regulation of mTORC1 in response to glucose availability occurred even in TSC1-deficient cells and AMPK-silenced cells, supporting the idea that the GAPDH-Rheb pathway functions independently of the AMPK axis. Furthermore, we show that glyceraldehyde-3-phosphate, a glycolytic intermediate that binds GAPDH, destabilizes the Rheb-GAPDH interaction even under low-glucose conditions, explaining how high-glucose flux suppresses the interaction and activates mTORC1 signaling. Taken together, our results suggest that the glycolytic flux regulates mTOR''s access to Rheb by regulating the Rheb-GAPDH interaction, thereby allowing mTORC1 to coordinate cell growth with glucose availability.The mTOR complex 1 (mTORC1) signal transduction pathway acts as a central controller of cell growth in mammals (20, 23, 29). mTORC1 integrates a wide range of intracellular and extracellular signals, including insulin, availability of nutrients (glucose and amino acids), cellular energy status, and hypoxia, to regulate protein synthesis and cell growth (11, 12, 17, 36, 46). Many of these environmental cues are integrated into tuberous sclerosis complex (TSC1-TSC2), the major upstream regulator of mTORC1. In response to the absence of insulin and to the low-energy status of cells, the TSC1-TSC2 complex stimulates the GTPase function of Rheb, a small GTPase that acts as a proximal key activator of mTORC1, which leads to the inhibition of Rheb-mediated mTORC1 activation. In contrast, inactivation of the TSC1-TSC2 complex results in the accumulation of GTP-bound Rheb and thus activation of mTORC1 (3, 13, 21, 27, 32, 39). For this reason, both the loss of TSC proteins and the overexpression of Rheb cause hyperactivation of mTORC1 signaling, which is frequently observed in many common human cancers (2, 5, 19, 25, 33). Therefore, a tight regulation of Rheb activity is critical for the proper operation of the mTORC1 pathway in response to environmental cues.Rheb is an atypical member of the Ras superfamily of GTPases (1, 10, 47). As with other small GTPases, the activity of Rheb is regulated by its guanine nucleotide binding status. However, the negative control of GTP-bound Rheb by the TSC1-TSC2 complex has only recently been investigated, and the regulation of the nucleotide binding status of Rheb is not fully understood. A recent study proposed that translationally controlled tumor protein may function as a guanine nucleotide exchange factor for Rheb that causes the accumulation of GTP-bound Rheb (18). GTP-bound Rheb is essential for activating mTOR kinase (21, 28, 38). However, the interaction between Rheb and mTOR does not depend on the GTP binding status of Rheb (30), raising questions regarding the mechanism by which Rheb activates mTORC1. Recently, FKBP38 (immunophilin FK506-binding protein, 38 kDa) was found to be a direct binding partner of Rheb and an inhibitor of mTORC1 (4). GTP-bound Rheb binds FKBP38 and releases FKBP38 from mTORC1, resulting in activation of the mTORC1 pathway. However, there have been conflicting results regarding the effects of nutrient availability on Rheb activity (31, 37, 42, 50) and the effect of these newly identified regulators of Rheb function (44, 45). Thus, the precise molecular mechanisms underlying Rheb regulation and Rheb-mediated mTORC1 activation have remained unclear.In this study, we identified glyceraldehyde-3-phosphate (Gly-3-P) dehydrogenase (GAPDH) as a novel Rheb binding protein and a negative regulator of Rheb. We found that the interaction between GAPDH and Rheb is induced when the glycolytic flux is suppressed under low-glucose conditions to inhibit mTORC1. Here, we provide a molecular mechanism underlying the cross talk between the glycolytic flux and the mTORC1 signaling.  相似文献   

6.
7.
8.
Proteomic identification of human papillomavirus type 16 (HPV16) E6-interacting proteins revealed several proteins involved in ubiquitin-mediated proteolysis. In addition to the well-characterized E6AP ubiquitin-protein ligase, a second HECT domain protein (HERC2) and a deubiquitylating enzyme (USP15) were identified by tandem affinity purification of HPV16 E6-associated proteins. This study focuses on the functional consequences of the interaction of E6 with USP15. Overexpression of USP15 resulted in increased levels of the E6 protein, and the small interfering RNA-mediated knockdown of USP15 decreased E6 protein levels. These results implicate USP15 directly in the regulation of E6 protein stability and suggest that ubiquitylated E6 could be a substrate for USP15 ubiquitin peptidase activity. It remains possible that E6 could affect the activity of USP15 on specific cellular substrates, a hypothesis that can be tested as more is learned about the substrates and pathways controlled by USP15.Human papillomaviruses (HPVs) are associated with several human cancers, most notably human cervical cancer, the second most common cancer among women worldwide (43). Papillomaviruses cause proliferative squamous epithelial lesions, and more than 100 HPV types have been described (14). The HPV types associated with mucosal squamous epithelial lesions have been further classified into high- or low-risk types based on the propensity for the lesions with which they are associated to progress to cancer. Among the high-risk HPV types, HPV type 16 (HPV16) and HPV18 account for approximately 70% of cervical cancers (43). The high-risk HPV types carry two genes, the E6 and E7 genes, which have oncogenic properties and are always expressed in HPV-positive cancers. E6 and E7 interfere with the p53 and retinoblastoma (pRB) tumor suppressor pathways, respectively, and contribute directly to cell cycle alterations, protection from apoptosis, and transformation (14). The dysregulated expression of the E6 and E7 oncoproteins is an important step in the progression from a preneoplastic stage to cancer in HPV-infected cells and is often a consequence of the integration of the viral genome into the host chromosome.The interaction between E6 and p53 is mediated by the E3 ubiquitin ligase E6AP (15). E6, p53, and E6AP form a complex in which E6 directs the ligase activity of E6AP to p53, thereby targeting p53 for ubiquitin-mediated degradation (36). E6, however, has a number of other cellular partners and other functions. For instance, the C terminus of the high-risk E6 protein contains a PDZ binding motif (20, 25) that mediates the interaction with several PDZ domain-containing proteins, including discs large (Dlg), Scribble (Scrib), the MAGI family of proteins, MUPP1, and PATJ (9, 10, 29). Some of these proteins are also targeted for degradation in an E6AP-dependent manner (22, 29). While the major mechanism of oncogenesis revolves around E6''s ability to inhibit the proapoptotic effects of p53, recent work involving the PDZ domain proteins indicates that these interactions are also important to the oncogenic potential of E6 (38, 41). Furthermore, E6 has been reported to bind a number of other cellular proteins, including but not limited to Bak, CBP/p300, c-Myc, E6TP1, hADA3, IRF3, MCM7, PTPH1, and TNF-R1 (7, 8, 17, 23, 24, 32, 35, 39, 40). The importance of the binding of several of these proteins with regard to the transformation or other functions of E6 remains to be established. E6 itself is thought to be targeted for degradation by an ubiquitin-proteasome pathway (18), although how E6 protein stability is regulated has not been well studied.Many of the E6 binding partners have been identified using purified bacterially expressed E6 fusion proteins and cell lysates from various cell types or using yeast two-hybrid screenings. While some of these interactions with E6 have been validated, the physiologic relevance of a number of proposed E6 targets remains undetermined. In an effort to identify E6-interacting proteins, perhaps under more physiologic conditions, we employed tandem affinity purification (TAP) using tagged HPV16 E6 stably expressed in the HPV16-positive cervical cancer cell line SiHa. We have discovered several new interacting proteins, including an interaction between E6 and the cellular deubiquitylating enzyme (DUB) USP15. USP15 is not targeted for degradation by E6, but we found that USP15 stabilizes E6 protein levels, suggesting that E6 may itself be a target for USP15 DUB activity.  相似文献   

9.
10.
11.
The human papillomavirus type 16 E5 oncoprotein (16E5) enhances acute, ligand-dependent activation of the epidermal growth factor receptor (EGFR) and concomitantly alkalinizes endosomes, presumably by binding to the 16-kDa “c” subunit of the V-ATPase proton pump (16K) and inhibiting V-ATPase function. However, the relationship between 16K binding, endosome alkalinization, and altered EGFR signaling remains unclear. Using an antibody that we generated against 16K, we found that 16E5 associated with only a small fraction of endogenous 16K in keratinocytes, suggesting that it was unlikely that E5 could significantly affect V-ATPase function by direct inhibition. Nevertheless, E5 inhibited the acidification of endosomes, as determined by a new assay using a biologically active, pH-sensitive fluorescent EGF conjugate. Since we also found that 16E5 did not alter cell surface EGF binding, the number of EGFRs on the cell surface, or the endocytosis of prebound EGF, we postulated that it might be blocking the fusion of early endosomes with acidified vesicles. Our studies with pH-sensitive and -insensitive fluorescent EGF conjugates and fluorescent dextran confirmed that E5 prevented endosome maturation (acidification and enlargement) by inhibiting endosome fusion. The E5-dependent defect in vesicle fusion was not due to detectable disruption of actin, tubulin, vimentin, or cytokeratin filaments, suggesting that membrane fusion was being directly affected rather than vesicle transport. Perhaps most importantly, while bafilomycin A1 (like E5) binds to 16K and inhibits endosome acidification, it did not mimic the ability of E5 to inhibit endosome enlargement or the trafficking of EGF. Thus, 16E5 alters EGF endocytic trafficking via a pH-independent inhibition of vesicle fusion.High-risk human papillomaviruses (HPVs) are the causative agent of cervical cancer (63) and HPV type 16 (HPV-16) is associated with a majority of cervical malignancies worldwide (13). HPV-16 encodes three oncoproteins: E5, E6, and E7. While the contributions of E6 and E7 to cellular immortalization and transformation have been characterized in detail (20), the role of HPV-16 E5 (16E5) is poorly understood (53). Nevertheless, a number of studies suggest that 16E5 does contribute to the development of cervical cancer. Most high-risk HPV types encode an E5 protein (48), and targeted expression of the three HPV-16 oncogenes in basal epithelial cells of transgenic mice (4) leads to a higher incidence of cervical cancer than does the expression of E6 and E7 alone (44). In addition, targeted epithelial expression of 16E5 (without E6 and E7) in transgenic mice induces skin tumors (21). It may be noteworthy that unlike high-risk HPV-18, which integrates into the host DNA and potentially disrupts E5 gene expression (20, 64), the HPV-16 genome often persists in episomal form in malignant lesions (12, 16, 24, 36, 42).Biological activities of 16E5 that may facilitate carcinogenesis include evading host immune detection by interfering with the transport of antigen-presenting major histocompatibility complex (MHC) class I molecules to the cell surface (6), promoting anchorage-independent growth (33, 41, 52) and disrupting gap junctions responsible for cell-cell communication (37, 58). The 16E5 phenotype most frequently linked to the development of cancer is enhanced ligand-dependent activation of the epidermal growth factor receptor (EGFR) (15, 41, 46, 52). 16E5 stimulates EGF-dependent cell proliferation in vitro (7, 33, 40, 41, 52, 60) and in vivo (21), which might expand the population of basal or stemlike keratinocytes and thereby increase the probability that some of these cells would undergo malignant transformation. A number of studies indicate that 16E5 may enhance ligand-dependent EGFR activation by interfering with the acidification of early endosomes containing EGF bound to activated EGFRs (17, 51, 57). It has been hypothesized that 16E5 inhibits the H+ V-ATPase responsible for maintaining an acidic luminal pH in late endosomes and lysosomes (28) by associating with the V-ATPase 16-kDa “c” subunit (16K) (1, 5, 14, 22, 46) and disrupting assembly of the V-ATPase integral (Vo) and peripheral (Vi) subcomplexes (10). In contrast, Thomsen et al. (57) reported that 16E5 inhibits early endosome trafficking in fibroblasts by completely depolymerizing actin microfilaments.Due to the unavailability of antibodies that recognize native 16E5 and 16K, direct association of 16E5 with 16K has only been observed by overexpressing epitope-tagged forms of both proteins in vitro (5, 46) or in vivo (1, 14, 22). It is uncertain, therefore, whether these associations occur when the proteins are expressed at “physiological” levels. In yeast, both wild-type 16E5 (10) and several 16E5 mutants that associate with 16K in COS cells (1) inhibit vacuolar acidification, although another study in yeast concludes the opposite (5). 16K is a component of the V-ATPase Vo subcomplex, which is assembled in the endoplasmic reticulum (ER) (28), and 16E5 localizes to the ER and nuclear envelope in epithelial cells (32, 54). Thus, the export of Vo from the ER could potentially be inhibited by a significant level of 16K binding to 16E5, although the differential alkalinization of endosomes rather than the Golgi apparatus (17) would require specificity for those proton pumps directed to those sites.In the present study, we generated an antibody against native 16K and used it to determine whether 16K/16E5 complexes formed in primary keratinocytes. We also synthesized a new pH-sensitive fluorescent EGF conjugate to evaluate whether there was a correlation between E5-induced EGFR activation, trafficking and endosome alkalinization. Finally, we simultaneously monitored EGFR endocytic trafficking (using pH-insensitive fluorescent EGF), endosome fusion (using fluorescent EGF and dextran), and the status of cellular filaments and microtubules to evaluate whether E5 might disrupt some of these structures that mediate vesicle transport.  相似文献   

12.
13.
14.
15.
Inactivating mutations of the neurofibromatosis 2 (NF2) gene, NF2, result predominantly in benign neurological tumors, schwannomas and meningiomas, in humans; however, mutations in murine Nf2 lead to a broad spectrum of cancerous tumors. The tumor-suppressive function of the NF2 protein, merlin, a membrane-cytoskeleton linker, remains unclear. Here, we identify the mammalian target of rapamycin complex 1 (mTORC1) as a novel mediator of merlin''s tumor suppressor activity. Merlin-deficient human meningioma cells and merlin knockdown arachnoidal cells, the nonneoplastic cell counterparts of meningiomas, exhibit rapamycin-sensitive constitutive mTORC1 activation and increased growth. NF2 patient tumors and Nf2-deficient mouse embryonic fibroblasts demonstrate elevated mTORC1 signaling. Conversely, the exogenous expression of wild-type merlin isoforms, but not a patient-derived L64P mutant, suppresses mTORC1 signaling. Merlin does not regulate mTORC1 via the established mechanism of phosphoinositide 3-kinase-Akt or mitogen-activated protein kinase/extracellular signal-regulated kinase-mediated TSC2 inactivation and may instead regulate TSC/mTOR signaling in a novel fashion. In conclusion, the deregulation of mTORC1 activation underlies the aberrant growth and proliferation of NF2-associated tumors and may restrain the growth of these lesions through negative feedback mechanisms, suggesting that rapamycin in combination with phosphoinositide 3-kinase inhibitors may be therapeutic for NF2.Meningiomas are mesenchymal tumors that arise from the arachnoid layer covering the brain and spinal cord and account for approximately 30% of all primary intracranial neoplasms (30). Most sporadic meningiomas (60%) display somatic inactivation of the NF2 gene. Germ line mutations of NF2 are associated with neurofibromatosis 2 (NF2), a dominantly inherited disorder characterized by multiple nervous system tumors, including schwannomas and meningiomas (33). Although most meningiomas are benign (WHO grade I), they often cause significant morbidity due to compression of the adjacent brain or spinal cord. Benign meningiomas also have recurrence rates of up to 20% over 10 years. Ten percent of meningiomas are classified as atypical (WHO grade II) or anaplastic (WHO grade III) and display more aggressive clinical behavior, with rapid growth and increased recurrence rates (6, 21). The current standard of care is maximal surgical resection, with adjuvant radiation reserved for progressive tumors or those with aggressive features (e.g., WHO grade II or III). The treatment strategy for meningiomas that progress despite surgery and radiation remains limited, and currently there is no effective chemotherapy.The development of effective therapies has been hampered, in part, by our incomplete understanding of the signals influencing meningioma cell growth. Enhanced expression of certain peptide and steroid growth factors and receptors in meningioma tissue suggests that specific autocrine growth-stimulatory loops may be functionally important in meningioma cell proliferation (20, 38). The scarcity of established meningioma models that would allow for the assessment of growth-regulatory mechanisms has also hampered progress. Recently, we have developed reliable meningioma models that overcome the challenges of the low growth rates and senescence of primary benign meningioma cells (19).Biallelic inactivation of the NF2 gene is detected in the majority of sporadic meningiomas and nearly all schwannomas (11). The tumor suppressor gene NF2 encodes merlin (also called schwannomin), a member of the ezrin-radixin-moesin (ERM) protein family that functions to link membrane proteins to the cortical actin cytoskeleton (31, 41). Like the ERM proteins, merlin has been implicated in the regulation of membrane organization and cytoskeleton-based cellular processes such as adhesion, migration, cell-cell contact, spreading, proliferation, and signal transduction (27). The loss of contact-dependent inhibition of proliferation is seen in several types of NF2-deficient cells (23, 29). Merlin controls cell proliferation in response to cell contact via CD44 (28) and functions together with the related tumor suppressor Expanded via the Hippo/Mst pathway in both Drosophila and some types of mammalian cells (14, 49). Although merlin is implicated in a wide range of cellular activities, the precise mechanism by which merlin mediates growth-inhibitory functions in human arachnoidal and Schwann cells and the way in which its loss results in tumor formation in NF2 remain poorly understood.We recently reported that primary human merlin-deficient meningioma cells exhibit a striking, enlarged-cell phenotype compared to nonneoplastic arachnoidal cell counterparts derived from the same patient (19). Interestingly, the tuberous sclerosis complex (TSC) tumor suppressor syndrome is characterized by widespread benign tumors that possess abnormally large cells (22). Mutations in the tumor suppressor genes TSC1 and TSC2 result in TSC syndrome, and the corresponding protein products, hamartin and tuberin (referred to as TSC1 and TSC2), function together as a complex that potently inhibits mammalian target of rapamycin complex 1 (mTORC1) (17). mTOR is an evolutionarily conserved Ser/Thr kinase that exists in one of two distinct functional complexes, TORC1 and TORC2. TORC1, which regulates autophagy, protein translation, and ribosome biogenesis, is potently and specifically inhibited by rapamycin (10, 46). TORC2, which is less sensitive to rapamycin, is important for cytoskeletal regulation and Akt/protein kinase B activation (16, 18, 36).The TSC1-TSC2 complex inhibits mTORC1 by acting as a GTPase-activating protein for the small GTPase Rheb (Ras homolog enriched in brain). Inactivation of the TSC1-TSC2 complex results in the accumulation of GTP-bound Rheb, which activates mTORC1 (10). In addition to naturally occurring mutations in the TSC1 and TSC2 genes, growth factor stimulation of the phosphoinositide 3-kinase (PI3K)-Akt pathway, as well as Ras/mitogen-activated protein kinase (MAPK) pathways, leads to the phosphorylation and inactivation of the TSC1-TSC2 complex and consequent activation of mTORC1 (17). The activation of mTORC1 results in the phosphorylation of two well-characterized effectors, eukaryotic translation initiation factor 4E-binding protein 1 (4EBP1) and S6 kinase 1 (S6K1), leading to an increase in ribosomal biogenesis and the selective translation of specific mRNA populations. As a critical regulator of cell growth and proliferation, the mTORC1 pathway is dysregulated in several hamartoma syndromes, as well as in many cancers (10).In this report, we identify the NF2 tumor suppressor protein, merlin, as a novel negative regulator of the mTORC1 pathway to control cell growth (cell size). We show that mTORC1 is constitutively activated in merlin-deficient human meningioma cells, leading to increased cell size. Furthermore, we suggest that the slow growth of merlin-deficient meningioma cells is due to a rapamycin-sensitive, mTORC1-S6K-dependent negative feedback loop that diminishes PI3K-Akt signaling in response to growth factor stimulation. The findings of these studies provide insight into the mechanism of merlin tumor suppressor activity and, moreover, indicate that rapamycin or rapamycin analogs in combination with PI3K inhibitors may provide promise as new therapeutics in the treatment of meningiomas and schwannomas.  相似文献   

16.
High-risk types of human papillomavirus (HPV) are considered the major causative agents of cervical carcinoma. The transforming ability of HPV resides in the E6 and E7 oncogenes, yet the pathway to transformation is not well understood. Cells expressing the oncogene E7 from high-risk HPVs have a high incidence of polyploidy, which has been shown to occur as an early event in cervical carcinogenesis and predisposes the cells to aneuploidy. The mechanism through which E7 contributes to polyploidy is not known. It has been hypothesized that E7 induces polyploidy in response to mitotic stress by abrogating the mitotic spindle assembly checkpoint. It was also proposed that E7 may stimulate rereplication to induce polyploidy. We have tested these hypotheses by using human epithelial cells in which E7 expression induces a significant amount of polyploidy. We find that E7-expressing cells undergo normal mitoses with an intact spindle assembly checkpoint and that they are able to complete cytokinesis. Our results also exclude DNA rereplication as a major mechanism of polyploidization in E7-expressing cells upon microtubule disruption. Instead, we have shown that while normal cells arrest at the postmitotic checkpoint after adaptation to the spindle assembly checkpoint, E7-expressing cells replicate their DNA and propagate as polyploid cells. Thus, abrogation of the postmitotic checkpoint leads to polyploidy formation in E7-expressing human epithelial cells. Our results suggest that downregulation of pRb is important for E7 to induce polyploidy and abrogation of the postmitotic checkpoint.An important hallmark of human cancers is aneuploidy, the state in which a cell has extra or missing chromosomes (12, 25). Polyploidy is the state in which cells have more than two equal sets of chromosomes and is thought to be an early event in multistep carcinogenesis that can lead to aneuploidy (1, 24), as exemplified in Barrett''s esophagus (11). Polyploidy has recently been shown to occur as an early event in cervical carcinogenesis and to predispose the cells to aneuploidy (26). Other recent studies have shown that tetraploid but not diploid mouse or human cells induce tumor formation in mice (3, 9). These studies highlight the potential importance of polyploidy in carcinogenesis.The cellular mechanisms responsible for this polyploidy formation are as of yet undetermined, but several models have been proposed. First, abrogation of the spindle assembly checkpoint followed by cleavage failure may lead to polyploidy formation (36, 40). A second proposed model is rereplication, a process of multiple rounds of DNA replication without an intervening mitosis. Third, cells that adapt to the mitotic spindle checkpoint halt in a G1-like state with 4C DNA content. Abrogation of this postmitotic checkpoint allows the cells to replicate their 4C DNA content, leading to polyploidy formation. This has been shown in cells that express the human papillomavirus type 16 (HPV-16) E6 oncogene that degrades p53 (21). Finally, cleavage failure, which yields binucleate cells with 4C DNA content, is also a potential mechanism for polyploidy formation (31).The postmitotic checkpoint becomes activated when cells with an intact spindle assembly checkpoint become arrested during mitosis for a prolonged period of time and eventually adapt to the checkpoint, exit mitosis without cleavage, and progress into a G1-like state with 4C DNA content (19, 22). The cells are prevented from continuing through the cell cycle and replicating their DNA by a proposed p53- and pRb-dependent postmitotic checkpoint (18, 19).High-risk types of HPV (of which HPV-16 is the most prevalent) are commonly associated with lesions that can progress to cervical carcinoma, which is one of the leading causes of cancer death in women worldwide (42). The transforming properties of high-risk HPVs primarily reside in the E6 and E7 oncogenes (reviewed in reference 7). The ability of high-risk HPV E6 and E7 proteins to promote the degradation of p53 and pRb, respectively, has been suggested as a mechanism by which HPV induces cellular transformation (6, 30). Expression of the high-risk HPV E6 and E7 oncogenes in human keratinocytes leads to polyploidy, which is enhanced by DNA damage and by activation of the spindle checkpoint through microtubule disruption (15, 27, 37, 38).Previously, it was thought but not directly shown that high-risk E6 and E7 induce polyploidy in response to microtubule disruption by abrogating the spindle checkpoint and that degradation of the tumor suppressor p53 by E6 is the mechanism by which E6 accomplishes this polyploidy formation (27, 37, 38). Others have proposed that E7 may play a role in stimulating DNA rereplication that occurs prior to mitosis initiation and polyploidy formation (20). Our recent studies demonstrate that E6 does not affect the mitotic spindle checkpoint (21). Instead, E6 abrogates the postmitotic checkpoint to induce polyploidy after microtubule disruption. Interestingly, E6 mutant proteins defective in inducing p53 degradation also induce polyploidy (21). The mechanism by which HPV E7 induces polyploidy remains to be determined. In this study, we investigate these possible mechanisms through which HPV-16 E7 induces polyploidy formation.  相似文献   

17.
The recently discovered Canis familiaris papillomavirus (PV) type 2 (CfPV2) provides a unique opportunity to study PV gene functions in vitro and in vivo. Unlike the previously characterized canine oral PV, CfPV2 contains an E5 open reading frame and is associated with progression to squamous cell carcinoma. In the current study, we have expressed and characterized the CfPV2-encoded E5 protein, a small, hydrophobic, 41-amino-acid polypeptide. We demonstrate that, similar to the E5 protein from high-risk human PV type 16, the CfPV2 E5 protein is localized in the endoplasmic reticulum (ER) and that its expression decreases keratinocyte proliferation and cell life span. E5 expression also increases the percentage of cells in the G1 phase of the cell cycle, with a concomitant decrease in the percentage of cells in S phase. To identify a potential mechanism for E5-mediated growth inhibition from the ER, we developed a real-time PCR method to quantify the splicing of XBP1 mRNA as a measure of ER stress. We found that the CfPV2 E5 protein induced ER stress and that this, as well as the observed growth inhibition, is tempered significantly by coexpression of the CfPV2 E6 and E7 genes. It is possible that the spatial/temporal regulation of E6/E7 gene expression during keratinocyte differentiation might therefore modulate E5 activity and ER stress.Papillomaviruses (PVs) are a large group of DNA tumor viruses that infect differentiated cutaneous and mucosal epithelia in a wide variety of mammalian species. There are nearly 200 types of human PVs (HPVs) (61), some of which are termed high risk (e.g., HPV type 16 [HPV-16]) and have the potential to immortalize primary cells and facilitate malignant progression to cervical cancer (52). An estimated 20 million cases of HPV infection occur each year in the United States alone, and cervical cancer is the second most common cause of cancer deaths among women worldwide. In general, PV infections are species specific, making it impossible to study the in vivo life cycle of HPV and the roles of its encoded proteins in viral replication and tumorigenesis. However, a few animal models do exist and the canine oral PV (COPV) has been helpful in mimicking certain biological properties of the high-risk mucosatropic HPVs, leading to the development of highly effective prophylactic vaccines (39, 49, 56). Although COPV mimics the mucosal tropism of the high-risk HPVs, it rarely progresses to cancer and lacks one of the early viral genes that may play an important role in tumorigenesis, E5. Recently, a new canine PV (Canis familiaris PV type 2 [CfPV2]) was isolated from the footpads of dogs (43). Unlike COPV, CfPV2 induces epidermal tumors and, when persistent, these benign infections progress to squamous cell carcinoma and metastasize widely. CfPV2 also encodes an E5 protein. In general, PV E5 proteins are small hydrophobic oncoproteins that localize to the endoplasmic reticulum (ER) or Golgi membranes (11, 16) but have limited amino acid sequence homology. Numerous cellular binding partners have been described for HPV-16 E5 proteins, including the V-ATPase 16-kDa subunit (1, 16), the nuclear import protein karyopherin beta 3 (25), the ER-resident protein Bap31 (40), proteins involved in zinc transport (ZnT1, EVER1, and EVER2) (27, 35), erbB4 (24), and HLA I (2). The HPV-16 E5 protein alters signaling pathways, predominantly the epidermal growth factor receptor (EGFR) pathway (17, 21, 46, 58); induces koilocytosis in cooperation with the E6 protein (26); and alters the plasma membrane expression of caveolin (47), HLA (3), and ganglioside GM1 (47). The last two changes might explain the ability of HPV-16-infected cells to circumvent detection by the host immune response and initiate tumor formation (3, 4, 21, 36, 46, 47).To provide a foundation for future in vivo studies, we initiated a series of in vitro experiments to define the intracellular localization and biological activity of CfPV2 E5. The current study demonstrates that CfPV2 E5 exhibits several properties of the HPV-16 E5 protein, including ER localization and inhibition of cell proliferation. A novel finding is that CfPV2 E5 activates the ER stress-signaling pathway, which may explain some of E5''s growth-related activities.  相似文献   

18.
Human papillomavirus type 16 (HPV16) has been identified as being the most common etiological agent leading to cervical cancer. Despite having a clear understanding of the role of HPV16 in oncogenesis, details of how HPV16 traffics during infection are poorly understood. HPV16 has been determined to enter via clathrin-mediated endocytosis, but the subsequent steps of HPV16 infection remain unclear. There is emerging evidence that several viruses take advantage of cross talk between routes of endocytosis. Specifically, JCV and bovine papillomavirus type 1 have been shown to enter cells by clathrin-dependent endocytosis and then require caveolin-1-mediated trafficking for infection. In this paper, we show that HPV16 is dependent on caveolin-1 after clathrin-mediated endocytosis. We provide evidence for the first time that HPV16 infection is dependent on trafficking to the endoplasmic reticulum (ER). This novel trafficking may explain the requirement for the caveolar pathway in HPV16 infection because clathrin-mediated endocytosis typically does not lead to the ER. Our data indicate that the infectious route for HPV16 following clathrin-mediated entry is caveolin-1 and COPI dependent. An understanding of the steps involved in HPV16 sorting and trafficking opens up the possibility of developing novel approaches to interfere with HPV16 infection and reduce the burden of papillomavirus diseases including cervical cancer.Human papillomavirus (PV) type 16 (HPV16) is a member of the family Papillomaviridae, a group of double-stranded DNA (dsDNA) viruses with a tropism for squamous epithelia (70). Most PV infections result in benign lesions, although a subset of high-risk HPVs are capable of malignant transformation, resulting in various cancers including cervical carcinoma (21, 38). Infection with HPV16 is responsible for causing approximately half of the cases of invasive cervical cancer (7). In spite of the link between HPV16 and cervical cancer, the intracellular movement of HPV16 through target keratinocyte cells during infection has not been defined in detail.Viruses can enter into target cells by taking advantage of the cell''s natural endocytosis machinery (60). One of the best-characterized modes of internalization is by receptor-mediated, clathrin-dependent endocytosis. In this mode of entry, clathrin-coated pits internalize cargo into clathrin-coated vesicles, which are pinched from the plasma membrane by dynamin-2 in order to internalize (68). The process of clathrin-mediated endocytosis occurs rapidly, resulting in the delivery of cargo to early/sorting endosomes within seconds to minutes (23, 31). From the sorting endosome, most clathrin-dependent ligands are trafficked back to the plasma membrane in recycling endosomes or to lysosomes for degradation (35, 56). Another well-studied model of ligand entry is caveolin-1-mediated endocytosis. The caveolar pathway typically involves entry via cholesterol-rich caveolae at the plasma membrane, which deliver their contents to pH-neutral organelles known as caveosomes (44, 65). The delivery of cargo from caveosomes to the Golgi apparatus and the endoplasmic reticulum (ER) was demonstrated previously (44, 46, 50). The traffickings of cargo internalized via clathrin- and caveolin-1-mediated endocytosis were once thought to be separate; however, it is becoming evident that viruses including bovine PV type 1 (BPV1), JCV, HPV31, and BKV rely on both pathways depending on the stage of infection (29, 32, 50, 63).PV internalization is preceded by virion attachment to the extracellular matrix, followed by binding to heparan sulfate (14, 15, 25). The involvement of a secondary receptor has been suggested, putatively an alpha-6 integrin (24, 37). Postbinding, a conformational change in the PV capsid results in a furin cleavage event at the N terminus of the minor capsid protein L2, which has been suggested to play a role in the endosomal escape of the viral genome (19, 30, 52). An increasing body of evidence supports the entry of HPV16 by clathrin-mediated endocytosis (9, 27, 62). Electron microscopy of HPV16 infection in COS-7 cells demonstrated HPV16 pseudovirions in clathrin-coated vesicles 20 min after entry and within structures resembling endosomes by 1 h postentry (9). HPV16 infection of HaCaT keratinocyte, COS-7, and 293TT cells has been blocked by chlorpromazine, an inhibitor of the formation of clathrin-coated pits (9, 27, 62, 67). Importantly, those studies showed that two inhibitors of caveolin-1-mediated internalization, filipin and nystatin, did not interfere with HPV16 infection (9, 27, 62). Our laboratory demonstrated the importance of dynamin in HPV16 infection, presumably in the scission of clathrin-coated vesicles from the plasma membrane (1). Recently, a clathrin-, caveolin-, and dynamin-independent endocytosis of HPV16 was suggested, although the use of the HPV18-positive, heteroploid HeLa cell line calls into question the relevance of this finding to natural infection (64).In a previous study, we described the postentry trafficking of BPV1 from endosomes to caveolin-1-positive vesicles, similarly to a related nonenveloped dsDNA virus, JCV (32, 50). Our data demonstrated that the infectious route of BPV1 involved entry by clathrin-mediated endocytosis followed by transport to the caveolar pathway in order to traffic to the ER (32). We found that BPV1 infection was neutralized by an antibody that prevented viral particle transport to the ER (33). The movement of BPV1 from the endosome to the caveosome provides a possible explanation for why BPV1 trafficking is so slow compared to those of other ligands of clathrin-mediated endocytosis (20, 26). The kinetics of BPV1 and HPV16 entry were previously reported to be identical, and the coincident internalization of HPV16 and BPV1 virus-like particles (VLPs) showed colocalization between the VLPs during infection (20, 62). These data suggest that HPV16 and BPV1 infection may be occurring by a similar mechanism.Our goal in the present study was to determine the intracellular trafficking events leading to HPV16 infection. The use of reporter virion technology has allowed the production of high-titer HPV16 virions by a method previously shown to yield virions that are infectious in vivo (16). In this study, we used HPV16 reporter virions to study HPV16 infection in the spontaneously immortalized human HaCaT keratinocyte cell line. Our data show that the infectious route of HPV16 is from early endosomes to caveolin-1-positive vesicles and then to the ER. Using immunofluorescence and short hairpin RNA (shRNA) against caveolin-1, we demonstrate the importance of the caveolar pathway after HPV16 has been internalized. We show that HPV16 infection was blocked by inhibiting the formation of COPI transport vesicles, which function in trafficking between the ER and the Golgi apparatus and from caveosomes to the ER (5, 39). We provide evidence that after reaching the caveosome, HPV16 requires passage to the ER for successful infection, a trafficking event made possible by COPI vesicle-mediated movement from the caveosome to the ER.  相似文献   

19.
The E5 protein of human papillomavirus type 16 is a small, hydrophobic protein that localizes predominantly to membranes of the endoplasmic reticulum (ER). To define the orientation of E5 in these membranes, we employed a differential, detergent permeabilization technique that makes use of the ability of low concentrations of digitonin to selectively permeabilize the plasma membrane and saponin to permeabilize all cellular membranes. We then generated a biologically active E5 protein that was epitope tagged at both its N and C termini and determined the accessibility of these termini to antibodies in the presence and absence of detergents. In both COS cells and human ectocervical cells, the C terminus of E5 was exposed to the cytoplasm, whereas the N terminus was restricted to the lumen of the ER. Finally, the deletion of the E5 third transmembrane domain (and terminal hydrophilic amino acids) resulted in a protein with its C terminus in the ER lumen. Taken together, these topology findings are compatible with a model of E5 being a 3-pass transmembrane protein and with studies demonstrating its C terminus interacting with cytoplasmic proteins.Human papillomaviruses (HPVs) are small, nonenveloped, double-stranded DNA viruses (25) that are the causative agents of benign and malignant tumors in humans (43). Most cancers of the cervix, vagina, and anus are caused by HPVs, as are a fraction of oropharyngeal cancers (29, 44). HPV type 16 (HPV-16) is the type most frequently found in anogenital cancers (15, 29), including cervical cancer, the most common cancer of women worldwide (44).Some of the biological activities of the HPV-16 E5 protein (16E5) include the augmentation of epidermal growth factor (EGF) signaling pathways (8), stimulation of anchorage-independent growth (38), alkalinization of endosomal pH (11), and alteration of membrane lipid composition (39). 16E5 also exhibits weak transforming activity in vitro (12), induces epithelial tumors in transgenic mice (13), and plays an important role in koilocytosis (20). There are multiple documented intracellular binding targets for 16E5 such as the 16-kDa subunit of the vacuolar H+-ATPase (7, 36), the heavy chain of HLA type I (1), EGF receptor family member ErbB4 (6), calnexin (16), the zinc transporter ZnT-1 (21), the EVER1 and EVER2 transmembrane channel-like proteins that modulate zinc homeostasis (21, 31), the nuclear import receptor family member karyopherin β3 (KNβ3) (19), and BAP31, which was previously reported to contribute to B-cell receptor activation (35).16E5 is a small, hydrophobic protein that localizes to intracellular membranes. When overexpressed in COS cells, it is present in the endoplasmic reticulum (ER) and, to a lesser extent, in the Golgi apparatus (7). At a lower level of expression in human foreskin keratinocytes and human ectocervical cells (HECs), 16E5 is present predominantly in the ER (10, 39). 16E5 contains three hydrophobic regions (14, 16, 22, 30, 41), and it was reported previously that the first hydrophobic region determines various biological properties of the protein (16, 22). It was also shown previously that the 16E5 C terminus plays a role in binding to karyopherin β3 (19) and in the formation of koilocytes (20). While theoretical predictions have been made for the topology of E5 in membranes (16), no experimental data exist. However, a recent study suggested that some highly expressed 16E5 localizes to the plasma membrane, with its C terminus exposed externally (18).The aim of the present study was to establish the orientation of 16E5 in the ER membrane. By using immunofluorescence microscopy coupled with differential membrane permeabilization (24, 34), we demonstrate the membrane orientation of an N- and C-terminally tagged, biologically active 16E5 protein. Our results indicate that the N terminus is intralumenal and that the C terminus is cytoplasmic, consistent with a model of E5 being a three-pass transmembrane protein and with current data on the interaction of its C terminus with cytoplasmic proteins.  相似文献   

20.
Papillomavirus capsids are composed of 72 pentamers reinforced through inter- and intrapentameric disulfide bonds. Recent research suggests that virus-like particles and pseudovirions (PsV) can undergo a redox-dependent conformational change involving disulfide interactions. We present here evidence that native virions exploit a tissue-spanning redox gradient that facilitates assembly events in the context of the complete papillomavirus life cycle. DNA encapsidation and infectivity titers are redox dependent in that they can be temporally modulated via treatment of organotypic cultures with oxidized glutathione. These data provide evidence that papillomavirus assembly and maturation is redox-dependent, utilizing multiple steps within both suprabasal and cornified layers.Human papillomaviruses (HPVs) exclusively infect cutaneous or mucosal epithelial tissues (14, 15, 30). HPV types that infect the mucosal epithelia can lead to the development of benign or malignant neoplasms, thus allowing for their categorization into low-risk or high-risk HPV types, respectively (14, 15, 30). A small subset of the more than 200 HPV types now identified are the causative agents of over 75% of all cervical cancers. HPV16 is the most prevalent type worldwide, found in ca. 50 to 62% of squamous cell carcinomas (14, 50).HPV16 virions contain a single, circular double-stranded DNA genome of ∼8 kb which associates with histones to form a chromatin-like structure. This minichromosome is packaged within a nonenveloped, icosahedral capsid composed of the major capsid protein L1 and the minor capsid protein L2. Similar to polyomaviruses, 72 capsomeres of L1 are geometrically arranged on a T=7 icosahedral lattice (2, 9, 17, 19, 36, 42). Recent cryoelectron microscopy images of HPV16 pseudovirions (PsV) suggest that L2 is arranged near the inner conical hollow of each L1 pentamer, although it is not known whether each L1 pentamer is occupied with a single L2 protein (5, 42).Due to technical constraints in the production of native HPV virions in organotypic culture, assembly studies of HPV particles have largely been restricted to the utilization of in vitro-derived particles such as virus-like particles (VLPs), PsV, and quasivirions (QV) (6, 12, 25, 40, 43). Recent research suggests that HPV and bovine papillomavirus PsV can undergo a redox-dependent conformational change that takes place over the course of many hours. This conformational change is characterized by resistance to proteolysis and chemical reduction and the appearance of a more orderly capsid structure via transmission electron microscopy (TEM) (7, 20).We present evidence that native virions, in the context of the complete papillomavirus life cycle, utilize a tissue-spanning redox gradient that facilitates multiple redox-dependent assembly and maturation events over the course of many days. We show that stability and specific infectivity of 20-day virions increases over 10-day virions, 20-day virions are more susceptible to neutralization than 10-day virions, and both viral DNA encapsidation and infectivity of HPV-infected tissues are redox dependent in that they can be manipulated via the treatment of organotypic tissues with oxidized glutathione (GSSG), which is concentration and temporally dependent.  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