37 research outputs found

    Extracellular vesicles : where the amyloid precursor protein carboxyl-terminal fragments accumulate and amyloid-β oligomerizes

    Get PDF
    Pleiotropic roles are proposed for brain extracellular vesicles (EVs) in the development of Alzheimer's disease (AD). Our previous studies have suggested a beneficial role for EVs in AD, where the endosomal system in vulnerable neurons is compromised, contributing to the removal of accumulated material from neurons. However, the involvement of EVs in propagating AD amyloidosis throughout the brain has been considered because the amyloid-β precursor protein (APP), APP metabolites, and key APP cleaving enzymes were identified in association with EVs. Here, we undertook to determine whether the secretase machinery is actively processing APP in EVs isolated from the brains of wild-type and APP overexpressing Tg2576 mice. We found that full-length APP is cleaved in EVs incubated in the absence of cells. The resulting metabolites, both α- and β-APP carboxyl-terminal fragments and APP intracellular domain accumulate in EVs over time and amyloid-β dimerizes. Thus, EVs contribute to the removal from neurons and transport of APP-derived neurotoxic peptides. While this is potentially a venue for propagation of the pathology throughout the brain, it may contribute to efficient removal of neurotoxic peptides from the brain

    An Internal Ribosome Entry Site Directs Translation of the 39-Gene from Pelargonium Flower Break Virus Genomic RNA: Implications for Infectivity

    Get PDF
    [EN] Pelargonium flower break virus (PFBV, genus Carmovirus) has a single-stranded positive-sense genomic RNA (gRNA) which contains five ORFs. The two 59-proximal ORFs encode the replicases, two internal ORFs encode movement proteins, and the 39-proximal ORF encodes a polypeptide (p37) which plays a dual role as capsid protein and as suppressor of RNA silencing. Like other members of family Tombusviridae, carmoviruses express ORFs that are not 59-proximal from subgenomic RNAs. However, in one case, corresponding to Hisbiscus chlorotic ringspot virus, it has been reported that the 39-proximal gene can be translated from the gRNA through an internal ribosome entry site (IRES). Here we show that PFBV also holds an IRES that mediates production of p37 from the gRNA, raising the question of whether this translation strategy may be conserved in the genus. The PFBV IRES was functional both in vitro and in vivo and either in the viral context or when inserted into synthetic bicistronic constructs. Through deletion and mutagenesis studies we have found that the IRES is contained within a 80 nt segment and have identified some structural traits that influence IRES function. Interestingly, mutations that diminish IRES activity strongly reduced the infectivity of the virus while the progress of the infection was favoured by mutations potentiating such activity. These results support the biological significance of the IRES-driven p37 translation and suggest that production of the silencing suppressor from the gRNA might allow the virus to early counteract the defence response of the host, thus facilitating pathogen multiplication and spread.This research was supported by grants BFU2006-11230 and BFU2009-11699 from the Spanish Ministerio de Ciencia e Innovacion (MICINN) and by grants ACOM/2006/210 and ACOMP/2009/040 (to CH) and GVPRE/2008/121 (to OF-M) from the Generalitat Valenciana. The latter was the recipient of an I3P postdoctoral contract from the Spanish Consejo Superior de Investigaciones Cientificas and an additional contract from MICINN. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.Fernandez Miragall, O.; Hernandez Fort, C. (2011). An Internal Ribosome Entry Site Directs Translation of the 39-Gene from Pelargonium Flower Break Virus Genomic RNA: Implications for Infectivity. PLoS ONE. 6(7):22617-22617. https://doi.org/10.1371/journal.pone.0022617S226172261767Gallie, D. R. (1996). Translational control of cellular and viral mRNAs. Plant Molecular Biology, 32(1-2), 145-158. doi:10.1007/bf00039381Kozak, M. (2002). Pushing the limits of the scanning mechanism for initiation of translation. Gene, 299(1-2), 1-34. doi:10.1016/s0378-1119(02)01056-9Sachs, A. B., Sarnow, P., & Hentze, M. W. (1997). Starting at the Beginning, Middle, and End: Translation Initiation in Eukaryotes. Cell, 89(6), 831-838. doi:10.1016/s0092-8674(00)80268-8Kozak, M. (1992). Regulation of Translation in Eukaryotic Systems. Annual Review of Cell Biology, 8(1), 197-225. doi:10.1146/annurev.cb.08.110192.001213Sonenberg, N., & Hinnebusch, A. G. (2009). Regulation of Translation Initiation in Eukaryotes: Mechanisms and Biological Targets. Cell, 136(4), 731-745. doi:10.1016/j.cell.2009.01.042F�tterer, J., & Hohn, T. (1996). Translation in plants-rules and exceptions. Plant Molecular Biology, 32(1-2), 159-189. doi:10.1007/bf00039382Gale, M., Tan, S.-L., & Katze, M. G. (2000). Translational Control of Viral Gene Expression in Eukaryotes. Microbiology and Molecular Biology Reviews, 64(2), 239-280. doi:10.1128/mmbr.64.2.239-280.2000Kozak, M. (2001). Constraints on reinitiation of translation in mammals. Nucleic Acids Research, 29(24), 5226-5232. doi:10.1093/nar/29.24.5226Pelletier, J., & Sonenberg, N. (1988). Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature, 334(6180), 320-325. doi:10.1038/334320a0Mokrejš, M., Mašek, T., Vopálenský, V., Hlubuček, P., Delbos, P., & Pospíšek, M. (2009). IRESite—a tool for the examination of viral and cellular internal ribosome entry sites. Nucleic Acids Research, 38(suppl_1), D131-D136. doi:10.1093/nar/gkp981Basso, J., Dallaire, P., Charest, P. J., Devantier, Y., & Laliberte, J.-F. (1994). Evidence for an Internal Ribosome Entry Site Within the 5’ Non-translated Region of Turnip Mosaic Potyvirus RNA. Journal of General Virology, 75(11), 3157-3165. doi:10.1099/0022-1317-75-11-3157Levis, C., & Astier-Manifacier, S. (1993). The 5′ untranslated region of PVY RNA, even located in an internal position, enables initiation of translation. Virus Genes, 7(4), 367-379. doi:10.1007/bf01703392Karetnikov, A., & Lehto, K. (2007). The RNA2 5’ leader of Blackcurrant reversion virus mediates efficient in vivo translation through an internal ribosomal entry site mechanism. Journal of General Virology, 88(1), 286-297. doi:10.1099/vir.0.82307-0Ivanov, P. A., Karpova, O. V., Skulachev, M. V., Tomashevskaya, O. L., Rodionova, N. P., Dorokhov, Y. L., & Atabekov, J. G. (1997). A Tobamovirus Genome That Contains an Internal Ribosome Entry Site Functionalin Vitro. Virology, 232(1), 32-43. doi:10.1006/viro.1997.8525Skulachev, M. V., Ivanov, P. A., Karpova, O. V., Korpela, T., Rodionova, N. P., Dorokhov, Y. L., & Atabekov, J. G. (1999). Internal Initiation of Translation Directed by the 5′-Untranslated Region of the Tobamovirus Subgenomic RNA I2. Virology, 263(1), 139-154. doi:10.1006/viro.1999.9928Jaag, H. M., Kawchuk, L., Rohde, W., Fischer, R., Emans, N., & Prufer, D. (2003). An unusual internal ribosomal entry site of inverted symmetry directs expression of a potato leafroll polerovirus replication-associated protein. Proceedings of the National Academy of Sciences, 100(15), 8939-8944. doi:10.1073/pnas.1332697100Balvay, L., Rifo, R. S., Ricci, E. P., Decimo, D., & Ohlmann, T. (2009). Structural and functional diversity of viral IRESes. Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms, 1789(9-10), 542-557. doi:10.1016/j.bbagrm.2009.07.005Kneller, E. L. P., Rakotondrafara, A. M., & Miller, W. A. (2006). Cap-independent translation of plant viral RNAs. Virus Research, 119(1), 63-75. doi:10.1016/j.virusres.2005.10.010Rico, P., & Hern�ndez, C. (2004). Complete nucleotide sequence and genome organization of Pelargonium flower break virus. Archives of Virology, 149(3), 641-651. doi:10.1007/s00705-003-0231-5Martinez-Turino, S., & Hernandez, C. (2010). Identification and characterization of RNA-binding activity in the ORF1-encoded replicase protein of Pelargonium flower break virus. Journal of General Virology, 91(12), 3075-3084. doi:10.1099/vir.0.023093-0Martínez-Turiño, S., & Hernández, C. (2011). A membrane-associated movement protein of Pelargonium flower break virus shows RNA-binding activity and contains a biologically relevant leucine zipper-like motif. Virology, 413(2), 310-319. doi:10.1016/j.virol.2011.03.001Martinez-Turino, S., & Hernandez, C. (2009). Inhibition of RNA silencing by the coat protein of Pelargonium flower break virus: distinctions from closely related suppressors. Journal of General Virology, 90(2), 519-525. doi:10.1099/vir.0.006098-0Rico, P., & Hernández, C. (2009). Characterization of the subgenomic RNAs produced by Pelargonium flower break virus: Identification of two novel RNAs species. Virus Research, 142(1-2), 100-107. doi:10.1016/j.virusres.2009.01.018Koh, D. C.-Y., Wong, S.-M., & Liu, D. X. (2003). Synergism of the 3′-Untranslated Region and an Internal Ribosome Entry Site Differentially Enhances the Translation of a Plant Virus Coat Protein. Journal of Biological Chemistry, 278(23), 20565-20573. doi:10.1074/jbc.m210212200Hellen, C. U. T. (2001). Internal ribosome entry sites in eukaryotic mRNA molecules. Genes & Development, 15(13), 1593-1612. doi:10.1101/gad.891101Martínez-Salas, E. (1999). Internal ribosome entry site biology and its use in expression vectors. Current Opinion in Biotechnology, 10(5), 458-464. doi:10.1016/s0958-1669(99)00010-5Dobrikova, E., Florez, P., Bradrick, S., & Gromeier, M. (2003). Activity of a type 1 picornavirus internal ribosomal entry site is determined by sequences within the 3’ nontranslated region. Proceedings of the National Academy of Sciences, 100(25), 15125-15130. doi:10.1073/pnas.2436464100Belsham, G. J. (2009). Divergent picornavirus IRES elements. Virus Research, 139(2), 183-192. doi:10.1016/j.virusres.2008.07.001Fernández-Miragall, O., Quinto, S. L. de, & Martínez-Salas, E. (2009). Relevance of RNA structure for the activity of picornavirus IRES elements. Virus Research, 139(2), 172-182. doi:10.1016/j.virusres.2008.07.009Pestova, T. V., Kolupaeva, V. G., Lomakin, I. B., Pilipenko, E. V., Shatsky, I. N., Agol, V. I., & Hellen, C. U. T. (2001). Molecular mechanisms of translation initiation in eukaryotes. Proceedings of the National Academy of Sciences, 98(13), 7029-7036. doi:10.1073/pnas.111145798FERNANDEZ-MIRAGALL, O. (2003). Structural organization of a viral IRES depends on the integrity of the GNRA motif. RNA, 9(11), 1333-1344. doi:10.1261/rna.5950603ROBERTSON, M. E. M., SEAMONS, R. A., & BELSHAM, G. J. (1999). A selection system for functional internal ribosome entry site (IRES) elements: Analysis of the requirement for a conserved GNRA tetraloop in the encephalomyocarditis virus IRES. RNA, 5(9), 1167-1179. doi:10.1017/s1355838299990301Dorokhov, Y. L., Skulachev, M. V., Ivanov, P. A., Zvereva, S. D., Tjulkina, L. G., Merits, A., … Atabekov, J. G. (2002). Polypurine (A)-rich sequences promote cross-kingdom conservation of internal ribosome entry. Proceedings of the National Academy of Sciences, 99(8), 5301-5306. doi:10.1073/pnas.082107599Xia, X., & Holcik, M. (2009). Strong Eukaryotic IRESs Have Weak Secondary Structure. PLoS ONE, 4(1), e4136. doi:10.1371/journal.pone.0004136Lu, J., Zhang, J., Wang, X., Jiang, H., Liu, C., & Hu, Y. (2006). In vitro and in vivo identification of structural and sequence elements in the 5’ untranslated region of Ectropis obliqua picorna-like virus required for internal initiation. Journal of General Virology, 87(12), 3667-3677. doi:10.1099/vir.0.82090-0Yang, L. J., Hidaka, M., Sonoda, J., Masaki, H., & Uozumi, T. (1997). Mutational Analysis of the Potato Virus Y 5′ Untranslated Region for Alteration in Translational Enhancement in Tobacco Protoplasts. Bioscience, Biotechnology, and Biochemistry, 61(12), 2131-2133. doi:10.1271/bbb.61.2131BERGAMINI, G., PREISS, T., & HENTZE, M. W. (2000). Picornavirus IRESes and the poly(A) tail jointly promote cap-independent translation in a mammalian cell-free system. RNA, 6(12), 1781-1790. doi:10.1017/s1355838200001679Bradrick, S. S. (2006). The hepatitis C virus 3’-untranslated region or a poly(A) tract promote efficient translation subsequent to the initiation phase. Nucleic Acids Research, 34(4), 1293-1303. doi:10.1093/nar/gkl019Lopez de Quinto, S. (2002). IRES-driven translation is stimulated separately by the FMDV 3’-NCR and poly(A) sequences. Nucleic Acids Research, 30(20), 4398-4405. doi:10.1093/nar/gkf569Song, Y., Friebe, P., Tzima, E., Junemann, C., Bartenschlager, R., & Niepmann, M. (2006). The Hepatitis C Virus RNA 3’-Untranslated Region Strongly Enhances Translation Directed by the Internal Ribosome Entry Site. Journal of Virology, 80(23), 11579-11588. doi:10.1128/jvi.00675-06Koh, D. C.-Y., Liu, D. X., & Wong, S.-M. (2002). A Six-Nucleotide Segment within the 3’ Untranslated Region of Hibiscus Chlorotic Ringspot Virus Plays an Essential Role in Translational Enhancement. Journal of Virology, 76(3), 1144-1153. doi:10.1128/jvi.76.3.1144-1153.2002Stupina, V. A., Meskauskas, A., McCormack, J. C., Yingling, Y. G., Shapiro, B. A., Dinman, J. D., & Simon, A. E. (2008). The 3’ proximal translational enhancer of Turnip crinkle virus binds to 60S ribosomal subunits. RNA, 14(11), 2379-2393. doi:10.1261/rna.1227808Truniger, V., Nieto, C., González-Ibeas, D., & Aranda, M. (2008). Mechanism of plant eIF4E-mediated resistance against a Carmovirus (Tombusviridae): cap-independent translation of a viral RNA controlledin cisby an (a)virulence determinant. The Plant Journal, 56(5), 716-727. doi:10.1111/j.1365-313x.2008.03630.xMiller, W. A., Wang, Z., & Treder, K. (2007). The amazing diversity of cap-independent translation elements in the 3′-untranslated regions of plant viral RNAs. Biochemical Society Transactions, 35(6), 1629-1633. doi:10.1042/bst0351629Miller, W. A., & White, K. A. (2006). Long-Distance RNA-RNA Interactions in Plant Virus Gene Expression and Replication. Annual Review of Phytopathology, 44(1), 447-467. doi:10.1146/annurev.phyto.44.070505.143353Koh, D. C.-Y., Wang, X., Wong, S.-M., & Liu, D. X. (2006). Translation initiation at an upstream CUG codon regulates the expression of Hibiscus chlorotic ringspot virus coat protein. Virus Research, 122(1-2), 35-44. doi:10.1016/j.virusres.2006.06.008Castaño, A., Ruiz, L., & Hernández, C. (2009). Insights into the translational regulation of biologically active open reading frames of Pelargonium line pattern virus. Virology, 386(2), 417-426. doi:10.1016/j.virol.2009.01.017Fraser, C. S., & Doudna, J. A. (2006). Structural and mechanistic insights into hepatitis C viral translation initiation. Nature Reviews Microbiology, 5(1), 29-38. doi:10.1038/nrmicro1558LÓPEZ-LASTRA, M., RIVAS, A., & BARRÍA, M. I. (2005). Protein synthesis in eukaryotes: The growing biological relevance of cap-independent translation initiation. Biological Research, 38(2-3). doi:10.4067/s0716-97602005000200003Pacheco, A., & Martinez-Salas, E. (2010). Insights into the Biology of IRES Elements through Riboproteomic Approaches. Journal of Biomedicine and Biotechnology, 2010, 1-12. doi:10.1155/2010/458927Bernstein, J., Sella, O., Le, S.-Y., & Elroy-Stein, O. (1997). PDGF2/c-sismRNA Leader Contains a Differentiation-linked Internal Ribosomal Entry Site (D-IRES). Journal of Biological Chemistry, 272(14), 9356-9362. doi:10.1074/jbc.272.14.9356Scheper, G. C., Voorma, H. O., & Thomas, A. A. M. (1994). Basepairing with 18S ribosomal RNA in internal initiation of translation. FEBS Letters, 352(3), 271-275. doi:10.1016/0014-5793(94)00975-9Dresios, J., Chappell, S. A., Zhou, W., & Mauro, V. P. (2005). An mRNA-rRNA base-pairing mechanism for translation initiation in eukaryotes. Nature Structural & Molecular Biology, 13(1), 30-34. doi:10.1038/nsmb1031Reigadas, S., Pacheco, A., Ramajo, J., de Quinto, S. L., & Martinez-Salas, E. (2005). Specific interference between two unrelated internal ribosome entry site elements impairs translation efficiency. FEBS Letters, 579(30), 6803-6808. doi:10.1016/j.febslet.2005.11.015Ishitani, M., Xiong, L., Stevenson, B., & Zhu, J. K. (1997). Genetic analysis of osmotic and cold stress signal transduction in Arabidopsis: interactions and convergence of abscisic acid-dependent and abscisic acid-independent pathways. The Plant Cell, 9(11), 1935-1949. doi:10.1105/tpc.9.11.1935Knoester, M., van Loon, L. C., van den Heuvel, J., Hennig, J., Bol, J. F., & Linthorst, H. J. M. (1998). Ethylene-insensitive tobacco lacks nonhost resistance against soil-borne fungi. Proceedings of the National Academy of Sciences, 95(4), 1933-1937. doi:10.1073/pnas.95.4.1933Mathews, D. H., Sabina, J., Zuker, M., & Turner, D. H. (1999). Expanded sequence dependence of thermodynamic parameters improves prediction of RNA secondary structure. Journal of Molecular Biology, 288(5), 911-940. doi:10.1006/jmbi.1999.2700Zuker, M. (2003). Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Research, 31(13), 3406-3415. doi:10.1093/nar/gkg59
    corecore