8 research outputs found

    Polycipiviridae: a proposed new family of polycistronic picorna-like RNA viruses

    Get PDF
    Solenopsis invicta virus 2 is a single-stranded positive-sense picorna-like RNA virus with an unusual genome structure. The monopartite genome of approximately 11 kb contains four open reading frames in its 5′ one third, three of which encode proteins with homology to picornavirus-like jelly-roll fold capsid proteins. These are followed by an intergenic region, and then a single long open reading frame that covers the 3′ two thirds of the genome. The polypeptide translation of this 3′ open reading frame contains motifs characteristic of picornavirus-like helicase, protease and RNA-dependent RNA polymerase domains. Inspection of public transcriptome shotgun assembly sequences revealed five related apparently nearly complete virus genomes isolated from ant species and one from a dipteran insect. By high-throughput sequencing and in silico assembly of RNA isolated from Solenopsis invicta and four other ant species, followed by targeted Sanger sequencing, we obtained nearly complete genomes for four further viruses in the group. Four further sequences were obtained from a recent large-scale invertebrate virus study. The 15 sequences are highly divergent (pairwise amino acid identities as low as 17% in the non-structural polyprotein), but possess the same overall polycistronic genome structure distinct from all other characterized picorna-like viruses. Consequently we propose the formation of a new virus family, Polycipiviridae, to classify this clade of arthropod-infecting polycistronic picorna-like viruses. We further propose that this family be divided into three genera: Chipolycivirus (2 species), Hupolycivirus (2 species), and Sopolycivirus (11 species), with members of the latter infecting ants in at least three different subfamilies.This work was supported by a Wellcome Trust grant [106207] and a European Research Council (ERC) European Union's Horizon 2020 research and innovation programme grant [646891] to A.E.F

    The coat protein of Alfalfa mosaic virus interacts and interferes with the transcriptional activity of the bHLH transcription factor ILR3 promoting salicylic acid-dependent defence signalling response

    Full text link
    [EN] During virus infection, specific viral component-host factor interaction elicits the transcriptional reprogramming of diverse cellular pathways. Alfalfa mosaic virus (AMV) can establish a compatible interaction in tobacco and Arabidopsis hosts. We show that the coat protein (CP) of AMV interacts directly with transcription factor (TF) ILR3 of both species. ILR3 is a basic helix-loop-helix (bHLH) family member of TFs, previously proposed to participate in diverse metabolic pathways. ILR3 has been shown to regulate NEET in Arabidopsis, a critical protein in plant development, senescence, iron metabolism and reactive oxygen species (ROS) homeostasis. We show that the AMV CP-ILR3 interaction causes a fraction of this TF to relocate from the nucleus to the nucleolus. ROS, pathogenesis-related protein 1 (PR1) mRNAs, salicylic acid (SA) and jasmonic acid (JA) contents are increased in healthy Arabidopsis loss-of-function ILR3 mutant (ilr3.2) plants, which implicates ILR3 in the regulation of plant defence responses. In AMV-infected wild-type (wt) plants, NEET expression is reduced slightly, but is induced significantly in ilr3.2 mutant plants. Furthermore, the accumulation of SA and JA is induced in Arabidopsis wt-infected plants. AMV infection in ilr3.2 plants increases JA by over 10-fold, and SA is reduced significantly, indicating an antagonist crosstalk effect. The accumulation levels of viral RNAs are decreased significantly in ilr3.2 mutants, but the virus can still systemically invade the plant. The AMV CP-ILR3 interaction may down-regulate a host factor, NEET, leading to the activation of plant hormone responses to obtain a hormonal equilibrium state, where infection remains at a level that does not affect plant viability.F.A. was the recipient of a contract Ramon y Cajal (RYC-2010-06169) program of the Ministerio de Educacion, Cultura y Deporte of Spain. We thank L. Corachan for excellent technical assistance. This work was supported by Grants BIO2014-54862-R from the Spanish grant agency Direccion General de Investigacion Cientifica y Tecnica (DGICT) the Prometeo Program GV2015/010 from the Generalitat Valenciana and PAID-06-10-1496 from the Universitat Politecnica de Valencia (Spain).Aparicio Herrero, F.; Pallás Benet, V. (2017). The coat protein of Alfalfa mosaic virus interacts and interferes with the transcriptional activity of the bHLH transcription factor ILR3 promoting salicylic acid-dependent defence signalling response. Molecular Plant Pathology. 18(2):173-186. https://doi.org/10.1111/mpp.12388S173186182Abbink, T. E. M., Peart, J. R., Mos, T. N. M., Baulcombe, D. C., Bol, J. F., & Linthorst, H. J. M. (2002). Silencing of a Gene Encoding a Protein Component of the Oxygen-Evolving Complex of Photosystem II Enhances Virus Replication in Plants. Virology, 295(2), 307-319. doi:10.1006/viro.2002.1332Alazem, M., & Lin, N. (2014). Roles of plant hormones in the regulation of host–virus interactions. Molecular Plant Pathology, 16(5), 529-540. doi:10.1111/mpp.12204Aparicio, F., Vilar, M., Perez-Payá, E., & Pallás, V. (2003). The coat protein of prunus necrotic ringspot virus specifically binds to and regulates the conformation of its genomic RNA. Virology, 313(1), 213-223. doi:10.1016/s0042-6822(03)00284-8Aparicio, F., Thomas, C. L., Lederer, C., Niu, Y., Wang, D., & Maule, A. J. (2005). Virus Induction of Heat Shock Protein 70 Reflects a General Response to Protein Accumulation in the Plant Cytosol. Plant Physiology, 138(1), 529-536. doi:10.1104/pp.104.058958Aparicio, F., Sánchez-Navarro, J. A., & Pallás, V. (2006). In vitro and in vivo mapping of the Prunus necrotic ringspot virus coat protein C-terminal dimerization domain by bimolecular fluorescence complementation. Journal of General Virology, 87(6), 1745-1750. doi:10.1099/vir.0.81696-0Balasubramaniam, M., Kim, B.-S., Hutchens-Williams, H. M., & Loesch-Fries, L. S. (2014). The Photosystem II Oxygen-Evolving Complex Protein PsbP Interacts With the Coat Protein of Alfalfa mosaic virus and Inhibits Virus Replication. Molecular Plant-Microbe Interactions®, 27(10), 1107-1118. doi:10.1094/mpmi-02-14-0035-rBhat, S., Folimonova, S. Y., Cole, A. B., Ballard, K. D., Lei, Z., Watson, B. S., … Nelson, R. S. (2012). Influence of Host Chloroplast Proteins on Tobacco mosaic virus Accumulation and Intercellular Movement. Plant Physiology, 161(1), 134-147. doi:10.1104/pp.112.207860Bol, J. F. (2005). Replication of Alfamo- and Ilarviruses: Role of the Coat Protein. Annual Review of Phytopathology, 43(1), 39-62. doi:10.1146/annurev.phyto.43.101804.120505Callaway, A., Giesman-Cookmeyer, D., Gillock, E. T., Sit, T. L., & Lommel, S. A. (2001). THEMULTIFUNCTIONALCAPSIDPROTEINS OFPLANTRNA VIRUSES. Annual Review of Phytopathology, 39(1), 419-460. doi:10.1146/annurev.phyto.39.1.419Collum, T. D., & Culver, J. N. (2016). The impact of phytohormones on virus infection and disease. Current Opinion in Virology, 17, 25-31. doi:10.1016/j.coviro.2015.11.003Culver, J. N., & Padmanabhan, M. S. (2007). Virus-Induced Disease: Altering Host Physiology One Interaction at a Time. Annual Review of Phytopathology, 45(1), 221-243. doi:10.1146/annurev.phyto.45.062806.094422Donze, T., Qu, F., Twigg, P., & Morris, T. J. (2014). Turnip crinkle virus coat protein inhibits the basal immune response to virus invasion in Arabidopsis by binding to the NAC transcription factor TIP. Virology, 449, 207-214. doi:10.1016/j.virol.2013.11.018Fryer, M. J., Ball, L., Oxborough, K., Karpinski, S., Mullineaux, P. M., & Baker, N. R. (2003). Control of Ascorbate Peroxidase 2 expression by hydrogen peroxide and leaf water status during excess light stress reveals a functional organisation of Arabidopsis leaves. The Plant Journal, 33(4), 691-705. doi:10.1046/j.1365-313x.2003.01656.xGarcía, J. A., & Pallás, V. (2015). Viral factors involved in plant pathogenesis. Current Opinion in Virology, 11, 21-30. doi:10.1016/j.coviro.2015.01.001Heim, M. A. (2003). The Basic Helix-Loop-Helix Transcription Factor Family in Plants: A Genome-Wide Study of Protein Structure and Functional Diversity. Molecular Biology and Evolution, 20(5), 735-747. doi:10.1093/molbev/msg088Herranz, M. C., Pallas, V., & Aparicio, F. (2012). Multifunctional Roles for the N-Terminal Basic Motif of Alfalfa mosaic virus Coat Protein: Nucleolar/Cytoplasmic Shuttling, Modulation of RNA-Binding Activity, and Virion Formation. Molecular Plant-Microbe Interactions®, 25(8), 1093-1103. doi:10.1094/mpmi-04-12-0079-rHuang, Z., Yeakley, J. M., Garcia, E. W., Holdridge, J. D., Fan, J.-B., & Whitham, S. A. (2005). Salicylic Acid-Dependent Expression of Host Genes in Compatible Arabidopsis-Virus Interactions. Plant Physiology, 137(3), 1147-1159. doi:10.1104/pp.104.056028Inaba, J., Kim, B. M., Shimura, H., & Masuta, C. (2011). Virus-Induced Necrosis Is a Consequence of Direct Protein-Protein Interaction between a Viral RNA-Silencing Suppressor and a Host Catalase. Plant Physiology, 156(4), 2026-2036. doi:10.1104/pp.111.180042Jiménez, I., López, L., Alamillo, J. M., Valli, A., & García, J. A. (2006). Identification of a Plum pox virus CI-Interacting Protein from Chloroplast That Has a Negative Effect in Virus Infection. Molecular Plant-Microbe Interactions®, 19(3), 350-358. doi:10.1094/mpmi-19-0350Kim, K.-C., Lai, Z., Fan, B., & Chen, Z. (2008). Arabidopsis WRKY38 and WRKY62 Transcription Factors Interact with Histone Deacetylase 19 in Basal Defense. The Plant Cell, 20(9), 2357-2371. doi:10.1105/tpc.107.055566Kim, S. A., Punshon, T., Lanzirotti, A., Li, L., Alonso, J. M., Ecker, J. R., … Guerinot, M. L. (2006). Localization of Iron in Arabidopsis Seed Requires the Vacuolar Membrane Transporter VIT1. Science, 314(5803), 1295-1298. doi:10.1126/science.1132563Liu, Z., Zhang, Z., Faris, J. D., Oliver, R. P., Syme, R., McDonald, M. C., … Friesen, T. L. (2012). The Cysteine Rich Necrotrophic Effector SnTox1 Produced by Stagonospora nodorum Triggers Susceptibility of Wheat Lines Harboring Snn1. PLoS Pathogens, 8(1), e1002467. doi:10.1371/journal.ppat.1002467Long, T. A., Tsukagoshi, H., Busch, W., Lahner, B., Salt, D. E., & Benfey, P. N. (2010). The bHLH Transcription Factor POPEYE Regulates Response to Iron Deficiency in Arabidopsis Roots. The Plant Cell, 22(7), 2219-2236. doi:10.1105/tpc.110.074096Lukhovitskaya, N. I., Solovieva, A. D., Boddeti, S. K., Thaduri, S., Solovyev, A. G., & Savenkov, E. I. (2013). An RNA Virus-Encoded Zinc-Finger Protein Acts as a Plant Transcription Factor and Induces a Regulator of Cell Size and Proliferation in Two Tobacco Species. The Plant Cell, 25(3), 960-973. doi:10.1105/tpc.112.106476Mandadi, K. K., & Scholthof, K.-B. G. (2013). Plant Immune Responses Against Viruses: How Does a Virus Cause Disease? The Plant Cell, 25(5), 1489-1505. doi:10.1105/tpc.113.111658Maule, A. J., Escaler, M., & Aranda, M. A. (2000). Programmed responses to virus replication in plants. Molecular Plant Pathology, 1(1), 9-15. doi:10.1046/j.1364-3703.2000.00002.xNechushtai, R., Conlan, A. R., Harir, Y., Song, L., Yogev, O., Eisenberg-Domovich, Y., … Mittler, R. (2012). Characterization of Arabidopsis NEET Reveals an Ancient Role for NEET Proteins in Iron Metabolism. The Plant Cell, 24(5), 2139-2154. doi:10.1105/tpc.112.097634Nelson, B. K., Cai, X., & Nebenführ, A. (2007). A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants. The Plant Journal, 51(6), 1126-1136. doi:10.1111/j.1365-313x.2007.03212.xNemeth, K., Salchert, K., Putnoky, P., Bhalerao, R., Koncz-Kalman, Z., Stankovic-Stangeland, B., … Koncz, C. (1998). Pleiotropic control of glucose and hormone responses by PRL1, a nuclear WD protein, in Arabidopsis. Genes & Development, 12(19), 3059-3073. doi:10.1101/gad.12.19.3059Ni, P., & Cheng Kao, C. (2013). Non-encapsidation activities of the capsid proteins of positive-strand RNA viruses. Virology, 446(1-2), 123-132. doi:10.1016/j.virol.2013.07.023Olsen, A. N., Ernst, H. A., Leggio, L. L., & Skriver, K. (2005). NAC transcription factors: structurally distinct, functionally diverse. Trends in Plant Science, 10(2), 79-87. doi:10.1016/j.tplants.2004.12.010Paddock, M. L., Wiley, S. E., Axelrod, H. L., Cohen, A. E., Roy, M., Abresch, E. C., … Jennings, P. A. (2007). MitoNEET is a uniquely folded 2Fe 2S outer mitochondrial membrane protein stabilized by pioglitazone. Proceedings of the National Academy of Sciences, 104(36), 14342-14347. doi:10.1073/pnas.0707189104Pallas, V., & García, J. A. (2011). How do plant viruses induce disease? Interactions and interference with host components. Journal of General Virology, 92(12), 2691-2705. doi:10.1099/vir.0.034603-0Pallas, V., Aparicio, F., Herranz, M. C., Sanchez-Navarro, J. A., & Scott, S. W. (2013). The Molecular Biology of Ilarviruses. Advances in Virus Research, 139-181. doi:10.1016/b978-0-12-407698-3.00005-3Palukaitis, P., Groen, S. C., & Carr, J. P. (2013). The Rumsfeld paradox: some of the things we know that we don’t know about plant virus infection. Current Opinion in Plant Biology, 16(4), 513-519. doi:10.1016/j.pbi.2013.06.004Peng, X., Hu, Y., Tang, X., Zhou, P., Deng, X., Wang, H., & Guo, Z. (2012). Constitutive expression of rice WRKY30 gene increases the endogenous jasmonic acid accumulation, PR gene expression and resistance to fungal pathogens in rice. Planta, 236(5), 1485-1498. doi:10.1007/s00425-012-1698-7Pieterse, C. M. J., Van der Does, D., Zamioudis, C., Leon-Reyes, A., & Van Wees, S. C. M. (2012). Hormonal Modulation of Plant Immunity. Annual Review of Cell and Developmental Biology, 28(1), 489-521. doi:10.1146/annurev-cellbio-092910-154055Puranik, S., Sahu, P. P., Srivastava, P. S., & Prasad, M. (2012). NAC proteins: regulation and role in stress tolerance. Trends in Plant Science, 17(6), 369-381. doi:10.1016/j.tplants.2012.02.004Rampey, R. A., Woodward, A. W., Hobbs, B. N., Tierney, M. P., Lahner, B., Salt, D. E., & Bartel, B. (2006). An Arabidopsis Basic Helix-Loop-Helix Leucine Zipper Protein Modulates Metal Homeostasis and Auxin Conjugate Responsiveness. Genetics, 174(4), 1841-1857. doi:10.1534/genetics.106.061044Ren, T., Qu, F., & Morris, T. J. (2005). The nuclear localization of the Arabidopsis transcription factor TIP is blocked by its interaction with the coat protein of Turnip crinkle virus. Virology, 331(2), 316-324. doi:10.1016/j.virol.2004.10.039Rodrigo, G., Carrera, J., Ruiz-Ferrer, V., del Toro, F. J., Llave, C., Voinnet, O., & Elena, S. F. (2012). A Meta-Analysis Reveals the Commonalities and Differences in Arabidopsis thaliana Response to Different Viral Pathogens. PLoS ONE, 7(7), e40526. doi:10.1371/journal.pone.0040526Sanchez-Navarro, J., Miglino, R., Ragozzino, A., & Bol, J. F. (2001). Engineering of Alfalfa mosaic virus RNA 3 into an expression vector. Archives of Virology, 146(5), 923-939. doi:10.1007/s007050170125Sánchez-Navarro, J. A., Carmen Herranz, M., & Pallás, V. (2006). Cell-to-cell movement of Alfalfa mosaic virus can be mediated by the movement proteins of Ilar-, bromo-, cucumo-, tobamo- and comoviruses and does not require virion formation. Virology, 346(1), 66-73. doi:10.1016/j.virol.2005.10.024Selth, L. A., Dogra, S. C., Rasheed, M. S., Healy, H., Randles, J. W., & Rezaian, M. A. (2004). A NAC Domain Protein Interacts with Tomato leaf curl virus Replication Accessory Protein and Enhances Viral Replication. The Plant Cell, 17(1), 311-325. doi:10.1105/tpc.104.027235Seo, M., Jikumaru, Y., & Kamiya, Y. (2011). Profiling of Hormones and Related Metabolites in Seed Dormancy and Germination Studies. Methods in Molecular Biology, 99-111. doi:10.1007/978-1-61779-231-1_7Su, L.-W., Chang, S. H., Li, M.-Y., Huang, H.-Y., Jane, W.-N., & Yang, J.-Y. (2013). Purification and biochemical characterization of Arabidopsis At-NEET, an ancient iron-sulfur protein, reveals a conserved cleavage motif for subcellular localization. Plant Science, 213, 46-54. doi:10.1016/j.plantsci.2013.09.001Taschner, P. E. M., Van Der Kuyl, A. C., Neeleman, L., & Bol, J. F. (1991). Replication of an incomplete alfalfa mosaic virus genome in plants transformed with viral replicase genes. Virology, 181(2), 445-450. doi:10.1016/0042-6822(91)90876-dThompson, J. D., Higgins, D. G., & Gibson, T. J. (1994). CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Research, 22(22), 4673-4680. doi:10.1093/nar/22.22.4673Toledo-Ortiz, G., Huq, E., & Quail, P. H. (2003). The Arabidopsis Basic/Helix-Loop-Helix Transcription Factor Family. The Plant Cell, 15(8), 1749-1770. doi:10.1105/tpc.013839Torres, M. A. (2010). ROS in biotic interactions. Physiologia Plantarum, 138(4), 414-429. doi:10.1111/j.1399-3054.2009.01326.xUhrig, J. F., Canto, T., Marshall, D., & MacFarlane, S. A. (2004). Relocalization of Nuclear ALY Proteins to the Cytoplasm by the Tomato Bushy Stunt Virus P19 Pathogenicity Protein. Plant Physiology, 135(4), 2411-2423. doi:10.1104/pp.104.046086Weber, P. H., & Bujarski, J. J. (2015). Multiple functions of capsid proteins in (+) stranded RNA viruses during plant–virus interactions. Virus Research, 196, 140-149. doi:10.1016/j.virusres.2014.11.014Whitham, S. A., Quan, S., Chang, H.-S., Cooper, B., Estes, B., Zhu, T., … Hou, Y.-M. (2003). Diverse RNA viruses elicit the expression of common sets of genes in susceptibleArabidopsis thalianaplants. The Plant Journal, 33(2), 271-283. doi:10.1046/j.1365-313x.2003.01625.xWhitham, S. A., Yang, C., & Goodin, M. M. (2006). Global Impact: Elucidating Plant Responses to Viral Infection. Molecular Plant-Microbe Interactions®, 19(11), 1207-1215. doi:10.1094/mpmi-19-120

    Mixed Infections of Four Viruses, the Incidence and Phylogenetic Relationships of Sweet Potato Chlorotic Fleck Virus (Betaflexiviridae) Isolates in Wild Species and Sweetpotatoes in Uganda and Evidence of Distinct Isolates in East Africa

    No full text
    corecore