9 research outputs found

    The mitochondrial and chloroplast dual targeting of a multifunctional plant viral protein modulates chloroplast-to-nucleus communication, RNA silencing suppressor activity, encapsidation, pathogenesis and tissue tropism

    Full text link
    [EN] Plant defense against melon necrotic spot virus (MNSV) is triggered by the viral auxiliary replicase p29 that is targeted to mitochondrial membranes causing morphological alterations, oxidative burst and necrosis. Here we show that MNSV coat protein (CP) was also targeted to mitochondria and mitochondrial-derived replication complexes [viral replication factories or complex (VRC)], in close association with p29, in addition to chloroplasts. CP import resulted in the cleavage of the R/arm domain previously implicated in genome binding during encapsidation and RNA silencing suppression (RSS). We also show that CP organelle import inhibition enhanced RSS activity, CP accumulation and VRC biogenesis but resulted in inhibition of systemic spreading, indicating that MNSV whole-plant infection requires CP organelle import. We hypothesize that to alleviate the p29 impact on host physiology, MNSV could moderate its replication and p29 accumulation by regulating CP RSS activity through organelle targeting and, consequently, eluding early-triggered antiviral response. Cellular and molecular events also suggested that S/P domains, which correspond to processed CP in chloroplast stroma or mitochondrion matrix, could mitigate host response inhibiting p29-induced necrosis. S/P deletion mainly resulted in a precarious balance between defense and counter-defense responses, generating either cytopathic alterations and MNSV cell-to-cell movement restriction or some degree of local movement. In addition, local necrosis and defense responses were dampened when RSS activity but not S/P organelle targeting was affected. Based on a robust biochemical and cellular analysis, we established that the mitochondrial and chloroplast dual targeting of MNSV CP profoundly impacts the viral infection cycle.The authors thank L. Corachan-Valencia for technical assistance. This work was funded by grant BIO2017¿88321-R from the Spanish Agencia Estatal de Investigacion (AEI) and Fondo Europeo de Desarrollo Regional (FEDER). J.A.N. and M.S.-B. are the recipients of a postdoctoral contract and a PhD fellowship from the Ministerio de Ciencia, Innovacion y Universidades of Spain, respectively. Ministerio de Ciencia e Innovacion (PID2020-115571RB-I00), European Regional Development FundNavarro Bohigues, JA.; Sáiz-Bonilla, M.; Sanchez Navarro, JA.; Pallás Benet, V. (2021). The mitochondrial and chloroplast dual targeting of a multifunctional plant viral protein modulates chloroplast-to-nucleus communication, RNA silencing suppressor activity, encapsidation, pathogenesis and tissue tropism. The Plant Journal. 108(1):197-218. https://doi.org/10.1111/tpj.15435S197218108

    Highly efficient construction of infectious viroid-derived clones

    Full text link
    [EN] Background Viroid research generally relies on infectious cDNA clones that consist of dimers of the entire viroid sequence. At present, those dimers are generated by self-ligation of monomeric cDNA, a strategy that presents several disadvantages: (i) low efficiency, (ii) it is a non-oriented reaction requiring tedious screenings and (iii) additional steps are required for cloning into a binary vector for agroinfiltration or for in vitro RNA production. Results We have developed a novel strategy for simultaneous construction of a viroid dimeric cDNA and cloning into a multipurpose binary vector ready for agroinfiltration or in vitro transcription. The assembly is based on IIs restriction enzymes and positive selection and supposes a universal procedure for obtaining infectious clones of a viroid independently of its sequence, with a high efficiency. Thus, infectious clones of one viroid of each family were obtained and its infectivity was analyzed by molecular hybridization. Conclusion This is a zero-background strategy for direct cloning into a binary vector, optimized for the generation of infectious viroids. As a result, this methodology constitutes a powerful tool for viroid research and exemplifies the applicability of type IIs restriction enzymes and the lethal gene ccdB to design efficient and affordable direct cloning approaches of PCR products into binary vectors.This work was supported by the Spanish Ministry of Economy and Competitiveness (co-supported by FEDER) Grants BIO2017-88321-R (VP) and AGL2016-79825-R (GG). The funders had no role in the experiment design, data analysis, decision to publish, or preparation of the manuscript.Márquez-Molins, J.; Navarro Bohigues, JA.; Pallás Benet, V.; Gomez, GG. (2019). Highly efficient construction of infectious viroid-derived clones. Plant Methods. 15:1-8. https://doi.org/10.1186/s13007-019-0470-4S1815Flores R, Minoia S, Carbonell A, Gisel A, Delgado S, López-Carrasco A, et al. Viroids, the simplest RNA replicons: how they manipulate their hosts for being propagated and how their hosts react for containing the infection. Virus Res. 2015;209:136–45.Flores R, Hernández C, de Alba AEM, Daròs J-A, Di Serio F. Viroids and viroid-host interactions. Annu Rev Phytopathol. 2005;43:117–39.Gómez G, Martínez G, Pallás V. Interplay between viroid-induced pathogenesis and RNA silencing pathways. Trends Plant Sci. 2009;14:264–9.Di Serio F, Flores R, Verhoeven JTJ, Li SF, Pallás V, Randles JW, et al. Current status of viroid taxonomy. Arch Virol. 2014;159:3467–78.Branch A, Robertson H. A replication cycle for viroids and other small infectious RNA’s. Science (80−). 1984;223:450–5.Daròs JA, Marcos JF, Hernández C, Flores R. Replication of avocado sunblotch viroid: evidence for a symmetric pathway with two rolling circles and hammerhead ribozyme processing. Proc Natl Acad Sci USA. 1994;91:12813–7.Gago S, Elena SF, Flores R, Sanjuán R. Extremely High Mutation Rate of a Hammerhead Viroid. Science (80-). 2009;323:1308–1308.Steger G, Riesner D. Viroid research and its significance for RNA technology and basic biochemistry. Nucleic Acids Res. 2018;46:10563–76.Gómez G, Torres H, Pallás V. Identification of translocatable RNA-binding phloem proteins from melon, potential components of the long-distance RNA transport system. Plant J. 2004;41:107–16.Takeda R, Petrov AI, Leontis NB, Ding B. A three-dimensional RNA motif in Potato spindle tuber viroid mediates trafficking from palisade mesophyll to spongy mesophyll in Nicotiana benthamiana. Plant Cell. 2011;23:258–72.Gómez G, Pallás V. Studies on subcellular compartmentalization of plant pathogenic noncoding RNAs give new insights into the intracellular RNA-traffic mechanisms. Plant Physiol. 2012;159:558–64.Wassenegger M, Heimes S, Riedel L, Sänger HL. RNA-directed de novo methylation of genomic sequences in plants. Cell. 1994;76:567–76.Martinez G, Castellano M, Tortosa M, Pallas V, Gomez G. A pathogenic non-coding RNA induces changes in dynamic DNA methylation of ribosomal RNA genes in host plants. Nucleic Acids Res. 2014;42:1553–62.Castellano M, Martinez G, Marques MC, Moreno-Romero J, Köhler C, Pallas V, et al. Changes in the DNA methylation pattern of the host male gametophyte of viroid-infected cucumber plants. J Exp Bot. 2016;67:5857–68.Cress DE, Kiefer MC, Owens RA. Construction of infectious potato spindle tuber viroid cDNA clones. Nucleic Acids Res. 1983;11:6821–35.Tabler M, Sänger HL. Infectivity studies on different potato spindle tuber viroid (PSTV) RNAs synthesized in vitro with the SP6 transcription system. EMBO J. 1985;4:2191–9.Visvader JE, Forster AC, Symons RH. Infectivity and in vitro mutagenesis of monomeric cDNA clones of citrus exocortis viroid indicates the site of processing of viroid precursors. Nucleic Acids Res. 1985;13:5843–56.Gardner RC, Chonoles KR, Owens RA. Potato spindle tuber viroid infections mediated by the Ti plasmid of Agrobacterium tumefaciens. Plant Mol Biol. 1986;6:221–8.Minoia S, Navarro B, Delgado S, Di Serio F, Flores R. Viroid RNA turnover: characterization of the subgenomic RNAs of potato spindle tuber viroid accumulating in infected tissues provides insights into decay pathways operating in vivo. Nucleic Acids Res. 2015;43:2313–25.López-Carrasco A, Ballesteros C, Sentandreu V, Delgado S, Gago-Zachert S, Flores R, et al. Different rates of spontaneous mutation of chloroplastic and nuclear viroids as determined by high-fidelity ultra-deep sequencing. PLoS Pathog. 2017;13:e1006547.Giguère T, Adkar-Purushothama CR, Perreault J-P. Comprehensive secondary structure elucidation of four genera of the family Pospiviroidae. PLoS ONE. 2014;9:e98655.Adkar-Purushothama CR, Brosseau C, Giguère T, Sano T, Moffett P, Perreault J-P. Small RNA derived from the virulence modulating region of the potato spindle tuber viroid silences callose synthase genes of tomato plants. Plant Cell. 2015;27:2178–94.Gibson DG, Young L, Chuang R-Y, Venter JC, Hutchison CA, Smith HO. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat Methods. 2009;6:343–5.Engler C, Gruetzner R, Kandzia R, Marillonnet S. Golden gate shuffling: a one-pot DNA shuffling method based on type IIs restriction enzymes. PLoS ONE. 2009;4:e5553.Carbonell A, Takeda A, Fahlgren N, Johnson SC, Cuperus JT, Carrington JC. New generation of artificial MicroRNA and synthetic trans-acting small interfering RNA vectors for efficient gene silencing in Arabidopsis. Plant Physiol. 2014;165:15–29.Genovés A, Navarro JA, Pallás V. Functional analysis of the five melon necrotic spot virus genome-encoded proteins. J Gen Virol. 2006;87:2371–80.Gómez G, Pallás V. A long-distance translocatable phloem protein from cucumber forms a ribonucleoprotein complex in vivo with Hop stunt viroid RNA. J Virol. 2004;78:10104–10.Herranz MC, Sanchez-Navarro JA, Aparicio F, Pallás V. Simultaneous detection of six stone fruit viruses by non-isotopic molecular hybridization using a unique riboprobe or ‘polyprobe’. J Virol Methods. 2005;124:49–55.Daròs JA. Eggplant latent viroid: a friendly experimental system in the family Avsunviroidae. Mol Plant Pathol. 2016;17:1170–7.Lin D, O’Callaghan CA. MetClo: methylase-assisted hierarchical DNA assembly using a single type IIS restriction enzyme. Nucleic Acids Res. 2018;46:e113.Sanjuán R, Daròs J-A. One-step site-directed mutagenesis of viroid dimeric cDNA. J Virol Methods. 2007;145:71–5

    A conserved motif in three viral movement proteins from different genera is required for host factor recruitment and cell-to-cell movement

    Full text link
    [EN] Due to their minimal genomes, plant viruses are forced to hijack specific cellular pathways to ensure host colonization, a condition that most frequently involves physical interaction between viral and host proteins. Among putative viral interactors are the movement proteins, responsible for plasmodesma gating and genome binding during viral transport. Two of them, DGBp1 and DGBp2, are required for alpha-, beta- and gammacarmovirus cell-to-cell movement, but the number of DGBp-host interactors identified at present is limited. By using two different approaches, yeast two-hybrid and bimolecular fluorescence complementation assays, we found three Arabidopsis factors, eIF3g1, RPP3A and WRKY36, interacting with DGBp1s from each genus mentioned above. eIF3g1 and RPP3A are mainly involved in protein translation initiation and elongation phases, respectively, while WRKY36 belongs to WRKY transcription factor family, important regulators of many defence responses. These host proteins are not expected to be associated with viral movement, but knocking out WRKY36 or silencing either RPP3A or eIF3g1 negatively affected Arabidopsis infection by Turnip crinkle virus. A highly conserved FNF motif at DGBp1 C-terminus was required for protein-protein interaction and cell-to-cell movement, suggesting an important biological role.We thank Dr. Anne Simon and Dr. Steve A. Lommel for providing an infectious cDNA clone of the Turnip crinkle virus strain M (TCV-M) and PZP-TCV-sGFP plasmid, respectively. This work was funded by grant BIO2017-88321-R from the Spanish Agencia Estatal de Investigacion (AEI) and Fondo Europeo de Desarrollo Regional (FEDER). J.A.N. and M.S.-S. are the recipients of a postdoctoral contract and a PhD fellowship from the Ministerio de Ciencia, Innovacion y Universidades of Spain.Navarro Bohigues, JA.; Serra-Soriano, M.; Corachán Valencia, L.; Pallás Benet, V. (2020). A conserved motif in three viral movement proteins from different genera is required for host factor recruitment and cell-to-cell movement. Scientific Reports. 10(1):1-15. https://doi.org/10.1038/s41598-020-61741-5S115101Sicard, A., Michalakis, Y., Gutiérrez, S. & Blanc, S. The strange lifestyle of multipartite viruses. PLoS Pathog. 12, e1005819–e1005819 (2016).Pita, J. S. & Roossinck, M. J. Virus populations, mutation rates and frequencies. In Plant Virus Evol. (ed. Roossinck, M. J.) 109–121 https://doi.org/10.1007/978-3-540-75763-4_6 (Springer Berlin Heidelberg, 2008).Ivanov, K. I. & Makinen, K. Coat proteins, host factors and plant viral replication. Curr. Opin. Virol. 2, 712–718 (2012).Wang, A. Dissecting the molecular network of virus-plant interactions: the complex roles of host factors. Annu. Rev. Phytopathol. 53, 45–66 (2015).García, J. A. & Pallás, V. Viral factors involved in plant pathogenesis. Curr. Opin. Virol. 10, 21–30 (2015).Garcia-Ruiz, H. Host factors against plant viruses. Mol. Plant Pathol. 20, 1588–1601 (2019).Hull, R. Induction of disease 1: virus movement through the plant and effects on plant metabolism. In Matthews’ Plant Virology (ed. Hull, R.) 373–436 (Academic Press, 2002).Navarro, J. A. & Pallas, V. An update on the intracellular and intercellular trafficking of carmoviruses. Front. Plant. Sci. 8, 1801 (2017).Azevedo, J. et al. Argonaute quenching and global changes in Dicer homeostasis caused by a pathogen-encoded GW repeat protein. Genes Dev. 24, 904–915 (2010).Zhang, X., Zhang, X., Singh, J., Li, D. & Qu, F. Temperature-dependent survival of Turnip crinkle virus-infected arabidopsis plants relies on an RNA silencing-based defense that requires dcl2, AGO2, and HEN1. J. Virol. 86, 6847–6854 (2012).Donze, T., Qu, F., Twigg, P. & Morris, T. J. 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 (2014).Lin, B. & Heaton, L. A. An Arabidopsis thaliana protein interacts with a movement protein of Turnip crinkle virus in yeast cells and in vitro. J. Gen. Virol. 82, 1245–1251 (2001).Navarro, J. A., Sanchez-Navarro, J. A. & Pallas, V. Key checkpoints in the movement of plant viruses through the host. Adv. Virus Res. 104, 1–64 (2019).Navarro, J. A. et al. RNA-binding properties and membrane insertion of Melon necrotic spot virus (MNSV) double gene block movement proteins. Virology 356, 57–67 (2006).Serra-Soriano, M., Pallás, V. & Navarro, J. A. A model for transport of a viral membrane protein through the early secretory pathway: minimal sequence and endoplasmic reticulum lateral mobility requirements. Plant J. 77, 863–879 (2014).Genoves, A., Navarro, J. A. & Pallas, V. Functional analysis of the five melon necrotic spot virus genome-encoded proteins. J. Gen. Virol. 87, 2371–2380 (2006).Li, W. Z., Qu, F. & Morris, T. J. Cell-to-cell movement of turnip crinkle virus is controlled by two small open reading frames that function in trans. Virology 244, 405–416 (1998).Martinez-Turino, S. & Hernandez, C. 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, 310–319 (2011).Molnár, A., Havelda, Z., Dalmay, T., Szutorisz, H. & Burgyán, J. Complete nucleotide sequence of tobacco necrosis virus strain D(H) and genes required for RNA replication and virus movement. J. Gen. Virol. 78(6), 1235–1239 (1997).Marcos, J. F., Vilar, M., Pérez-Payá, E. & Pallás, V. In vivo detection, RNA-binding properties and characterization of the RNA-binding domain of the p7 putative movement protein from Carnation mottle carmovirus (CarMV). Virology 255, 354–365 (1999).Vilar, M., Esteve, V., Pallás, V., Marcos, J. F. & Pérez-Payá, E. Structural properties of carnation mottle virus p7 movement protein and its RNA-binding domain. J. Biol. Chem. 276, 18122–18129 (2001).Vilar, M., Saurí, A., Marcos, J. F., Mingarro, I. & Pérez‐Payá, E. Transient structural ordering of the RNA‐binding domain of carnation mottle virus p7 movement protein modulates nucleic acid binding. Chembiochem 6, 1391–1396 (2005).Genoves, A., Navarro, J. A. & Pallas, V. A self-interacting carmovirus movement protein plays a role in binding of viral RNA during the cell-to-cell movement and shows an actin cytoskeleton dependent location in cell periphery. Virology 395, 133–142 (2009).Garcia-Castillo, S., Sanchez-Pina, M. A. & Pallas, V. Spatio-temporal analysis of the RNAs, coat and movement (p7) proteins of Carnation mottle virus in Chenopodium quinoa plants. J. Gen. Virol. 84, 745–749 (2003).Cohen, Y., Qu, F., Gisel, A., Morris, T. J. & Zambryski, P. C. Nuclear localization of turnip crinkle virus movement protein p8. Virology 273, 276–285 (2000).Martinez-Gil, L., Sauri, A., Vilar, M., Pallas, V. & Mingarro, I. Membrane insertion and topology of the p7B movement protein of Melon Necrotic Spot Virus (MNSV). Virology 367, 348–357 (2007).Sauri, A., Saksena, S., Salgado, J., Johnson, A. E. & Mingarro, I. Double-spanning plant viral movement protein integration into the endoplasmic reticulum membrane is signal recognition particle-dependent, translocon-mediated, and concerted. J. Biol. Chem. 280, 25907–25912 (2005).Genoves, A., Navarro, J. A. & Pallas, V. The intra- and intercellular movement of Melon necrotic spot virus (MNSV) depends on an active secretory pathway. Mol. Plant-Microbe Interact. 23, 263–272 (2010).Aparicio, F. & Pallás, V. 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. Mol. Plant Pathol. 18(2), 173–186 (2017).Yang, Y. et al. UVR8 interacts with WRKY36 to regulate HY5 transcription and hypocotyl elongation in Arabidopsis. Nat. Plants 4, 98–107 (2018).Cutler, S. R., Ehrhardt, D. W., Griffitts, J. S. & Somerville, C. R. Random GFP::cDNA fusions enable visualization of subcellular structures in cells of Arabidopsis at a high frequency. Proc. Natl. Acad. Sci. USA 97, 3718–3723 (2000).La Cour, T. et al. Analysis and prediction of leucine-rich nuclear export signals. Protein Eng. Des. Sel. 17, 527–536 (2004).Prieto, G., Fullaondo, A. & Rodriguez, J. A. Prediction of nuclear export signals using weighted regular expressions (Wregex). Bioinformatics 30, 1220–1227 (2014).Wobbe, K. K., Akgoz, M., Dempsey, D. A. & Klessig, D. F. A single amino acid change in turnip crinkle virus movement protein p8 affects RNA binding and virulence on Arabidopsis thaliana. J. Virol. 72, 6247–6250 (1998).Akgoz, M., Nguyen, Q. N., Talmadge, A. E., Drainville, K. E. & Wobbe, K. K. Mutational analysis of Turnip crinkle virus movement protein p8. Mol. Plant Pathol. 2(1), 37–48, https://doi.org/10.1046/j.1364-3703.2001.00048.x (2001).Ahlfors, R. et al. Arabidopsis RADICAL-INDUCED CELL DEATH1 belongs to the WWE protein-protein interaction domain protein family and modulates abscisic acid, ethylene, and methyl jasmonate responses. Plant Cell 16, 1925–1937 (2004).Aravind, L. The WWE domain: a common interaction module in protein ubiquitination and ADP ribosylation. Trends Biochem. Sci. 26, 273–275 (2001).Meyer, E. A., Castellano, R. K. & Diederich, F. Interactions with aromatic rings in chemical and biological recognition. Angew. Chem. Int. Ed. Engl. 42, 1210–1250 (2003).Profit, A. A., Felsen, V., Chinwong, J., Mojica, E. R. & Desamero, R. Z. Evidence of pi-stacking interactions in the self-assembly of hIAPP(22-29). Proteins 81, 690–703 (2013).Amari, K., Vazquez, F. & Heinlein, M. Manipulation of plant host susceptibility: an emerging role for viral movement proteins? Front. Plant. Sci. 3, 10 (2012).Morozov, S. Y. & Solovyev, A. G. Did silencing suppression counter-defensive strategy contribute to origin and evolution of the triple gene block coding for plant virus movement proteins? Front. Plant. Sci. 3, 136 (2012).Levy, A., Zheng, J. Y. & Lazarowitz, S. G. The tobamovirus turnip vein clearing virus 30-kilodalton movement protein localizes to novel nuclear filaments to enhance virus infection. J. Virol. 87, 6428–6440 (2013).Gonzalo, P. & Reboud, J. P. The puzzling lateral flexible stalk of the ribosome. Biol. Cell. 95, 179–193 (2003).Szick, K., Springer, M. & Bailey-Serres, J. Evolutionary analyses of the 12-kDa acidic ribosomal P-proteins reveal a distinct protein of higher plant ribosomes. Proc. Natl. Acad. Sci. USA 95, 2378–2383 (1998).Hafren, A., Eskelin, K. & Makinen, K. Ribosomal protein P0 promotes Potato virus A infection and functions in viral translation together with VPg and eIF(iso)4E. J. Virol. 87, 4302–4312 (2013).Sato, H. et al. Measles virus N protein inhibits host translation by binding to eIF3-p40. J. Virol. 81, 11569–11576 (2007).Bhardwaj, U., Powell, P. & Goss, D. J. Eukaryotic initiation factor (eIF) 3 mediates barley yellow dwarf viral mRNA 3′–5′ UTR interactions and 40S ribosomal subunit binding to facilitate cap-independent translation. Nucleic Acids Res. 47, 6225–6235Park, H. S., Himmelbach, A., Browning, K. S., Hohn, T. & Ryabova, L. A. A plant viral ‘reinitiation’ factor interacts with the host translational machinery. Cell 106, 723–733 (2001).Thiébeauld, O., Pooggin, M. & Ryabova, L. Alternative translation strategies in plant viruses. Plant Viruses 1, 1–20 (2007).Bureau, M. et al. P6 protein of Cauliflower mosaic virus, a translation reinitiator, interacts with ribosomal protein L13 from Arabidopsis thaliana. J. Gen. Virol. 85, 3765–3775 (2004).Ryabova, L. A., Pooggin, M. M. & Hohn, T. Translation reinitiation and leaky scanning in plant viruses. Virus Res. 119, 52–62 (2006).Chen, L., Zhang, L., Li, D., Wang, F. & Yu, D. WRKY8 transcription factor functions in the TMV-cg defense response by mediating both abscisic acid and ethylene signaling in Arabidopsis. Proc. Natl. Acad. Sci. USA 110, E1963–E1971 (2013).Huh, S. U., Choi, L. M., Lee, G. J., Kim, Y. J. & Paek, K. H. Capsicum annuum WRKY transcription factor d (CaWRKYd) regulates hypersensitive response and defense response upon Tobacco mosaic virus infection. Plant Sci. 197, 50–58 (2012).Menke, F. L. et al. Tobacco transcription factor WRKY1 is phosphorylated by the MAP kinase SIPK and mediates HR-like cell death in tobacco. Mol. Plant Microbe Interact. 18, 1027–1034 (2005).Park, H. S. & Kim, K. H. Virus-induced silencing of the WRKY1 transcription factor that interacts with the SL1 structure of Potato virus X leads to higher viral RNA accumulation and severe necrotic symptoms. Plant Pathol. J. 28, 40–48 (2012).Gao, R., Liu, P., Yong, Y. & Wong, S. M. Genome-wide transcriptomic analysis reveals correlation between higher WRKY61 expression and reduced symptom severity in Turnip crinkle virus infected Arabidopsis thaliana. Sci. Rep. 6, 24604 (2016).Zou, L. et al. Transcription factor WRKY30 mediates resistance to Cucumber mosaic virus in Arabidopsis. Biochem. Biophys. Res. Commun. 517, 118–124 (2019).Huang, Y. et al. Members of WRKY Group III transcription factors are important in TYLCV defense signaling pathway in tomato (Solanum lycopersicum). BMC Genomics 17, 788 (2016).Besseau, S., Li, J. & Palva, E. T. WRKY54 and WRKY70 co-operate as negative regulators of leaf senescence in Arabidopsis thaliana. J. Exp. Bot. 63, 2667–2679 (2012).Yan, L. et al. Auto- and cross-repression of three arabidopsis WRKY transcription factors WRKY18, WRKY40, and WRKY60 negatively involved in ABA signaling. J. Plant Growth Regul. 32, 399–416 (2013).Xu, E., Vaahtera, L. & Brosché, M. Roles of defense hormones in the regulation of ozone-induced changes in gene expression and cell death. Mol. Plant 8, 1776–1794 (2015).Li, S.-W., Leng, Y. & Shi, R.-F. Transcriptomic profiling provides molecular insights into hydrogen peroxide-induced adventitious rooting in mung bean seedlings. BMC Genomics 18, 188 (2017).Imran, Q. M. et al. Transcriptome wide identification and characterization of NO-responsive WRKY transcription factors in Arabidopsis thaliana L. Environ. Exp. Bot. 148, 128–143 (2018).Hernandez, J. et al. Oxidative stress and antioxidative responses in plant–virus interactions. Physiol. Mol. Plant Pathol. 94, 134–148 (2015).Ahlfors, R., Brosche, M., Kollist, H. & Kangasjarvi, J. Nitric oxide modulates ozone-induced cell death, hormone biosynthesis and gene expression in Arabidopsis thaliana. Plant J. 58, 1–12 (2009).Oh, J. W., Kong, Q., Song, C., Carpenter, C. D. & Simon, A. E. Open reading frames of Turnip crinkle virus involved in satellite symptom expression and incompatibility with Arabidopsis thaliana ecotype Dijon. Mol. Plant Microbe Interact. 8, 979–987 (1995).Gietz, R. D. & Woods, R. A. Transformation of yeast by lithium acetate/single-stranded carrier DNA/polyethylene glycol method. Methods Enzym. 350, 87–96 (2002).Nemeth, K. et al. Pleiotropic control of glucose and hormone responses by PRL1, a nuclear WD protein, in Arabidopsis. Genes Dev. 12, 3059–3073 (1998).Knoester, M. et al. Ethylene-insensitive tobacco lacks nonhost resistance against soil-borne fungi. Proc. Natl. Acad. Sci. USA 95, 1933–1937 (1998).Liu, Y., Schiff, M., Marathe, R. & Dinesh-Kumar, S. P. Tobacco Rar1, EDS1 and NPR1/NIM1 like genes are required for N-mediated resistance to tobacco mosaic virus. Plant J. 30, 415–429 (2002).Fernandez-Pozo, N., Rosli, H. G., Martin, G. B. & Mueller, L. A. The SGN VIGS tool: user-friendly software to design virus-induced gene silencing (VIGS) constructs for functional genomics. Mol. Plant 8, 486–488 (2015).Martínez-Pérez, M. et al. Arabidopsis m6A demethylase activity modulates viral infection of a plant virus and the m6A abundance in its genomic RNAs. Proc. Natl. Acad. Sci. USA 114, 10755–10760 (2017).Powers, J. G. et al. A versatile assay for the identification of RNA silencing suppressors based on complementation of viral movement. Mol. Plant-Microbe Interact. 21(7), 879–890, https://doi.org/10.1094/MPMI-21-7-0879 (2008).Pallas, V., Mas, P. & Sanchez-Navarro, J. A. Detection of plant RNA viruses by nonisotopic dot-blot hybridization. Methods Mol. Biol. 81, 461–468 (1998).Navarro, J. A., Serra-Soriano, M. & Pallás, V. A Protocol to Measure the Extent of Cell-to-cell Movement of RNA Viruses in Planta. Bio-protocol 4, e1269 (2014).Koressaar, T. et al. Primer3_masker: integrating masking of template sequence with primer design software. Bioinformatics 34, 1937–1938 (2018).Lilly, S. T., Drummond, R. S., Pearson, M. N. & MacDiarmid, R. M. Identification and validation of reference genes for normalization of transcripts from virus-infected Arabidopsis thaliana. Mol. Plant Microbe Interact. 24, 294–304 (2011)

    Might exogenous circular RNAs act as protein-coding transcripts in plants?

    Full text link
    [EN] Circular RNAs (circRNAs) are regulatory molecules involved in the modulation of gene expression. Although originally assumed as non-coding RNAs, recent studies have evidenced that animal circRNAs can act as translatable transcripts. The study of plant-circRNAs is incipient, and no autonomous coding plant-circRNA has been described yet. Viroids are the smallest plant-pathogenic circRNAs known to date. Since their discovery 50 years ago, viroids have been considered valuable systems for the study of the structure-function relationships in RNA, essentially because they have not been shown to have coding capacity. We used two pathogenic circRNAs (Hop stunt viroid and Eggplant latent viroid) as experimental tools to explore the coding potential of plant-circRNAs. Our work supports that the analysed viroids contain putative ORFs able to encode peptides carrying subcellular localization signals coincident with the corresponding replication-specific organelle. Bioassays in well-established hosts revealed that mutations in these ORFs diminish their biological efficiency. Interestingly, circular forms of HSVd and ELVd were found to co-sediment with polysomes, revealing their physical interaction with the translational machinery of the plant cell. Based on this evidence we hypothesize about the possibility that plant circRNAs in general, and viroids in particular, can act, under certain cellular conditions, as non-canonical translatable transcripts.This work was supported by the Spanish Ministry of Economy and Competitiveness (co-supported by FEDER) PID2020-115571RB-I00VP) and AGL2016-79825-R (GG), and by the Spanish Ministry of Science and Innovation (co-supported by FEDER) Grant PID2019104126RB-I00 (GG). JMM was the recipient of a predoctoral contract supported by the Conselleria d ' Educacio, Investigacio, Cultura i Esport Generalitat Valenciana -ACIF Programme (ACIF-2017-114). The funders had no role in the experiment design, data analysis, decision to publish, or preparation of the manuscript;Conselleria d'Educacio, Investigacio, Cultura i Esport [ACIF-2017-114];Ministerio de Economia, Industria y Competitividad, Gobierno de Espana [AGL201679825-R];Ministerio de Econom?a, Industria y Competitividad, Gobierno de Espana [PID2020-115571RB-I00];Ministerio de Ciencia, Innovacion y Universidades [PID2019-104126RB-I00].Márquez-Molins, J.; Navarro Bohigues, JA.; Cervera Seco, L.; Pallás Benet, V.; Gomez, GG. (2021). Might exogenous circular RNAs act as protein-coding transcripts in plants?. RNA Biology. 18:98-107. https://doi.org/10.1080/15476286.2021.1962670S981071

    Cucurbit chlorotic yellows virus p22 suppressor of RNA silencing binds single-, double-stranded long and short interfering RNA molecules in vitro

    Full text link
    [EN] Cucurbit chlorotic yellows virus (CCYV) is a new member of the genus Crinivirus (family Closteroviridae) with a bi-partite genome. CCYV RNA 1-encoded p22 has recently been reported to be a weak local suppressor of RNA silencing for which an interaction with cucumber SKP1LB1 through an F-box-like motif was demonstrated to be essential. Using a bacterially expressed maltose-binding protein (MBP) fusion of CCYV p22 in electrophoretic mobility shift assays (EMSA), we have examined in vitro its ability to bind different RNA templates. Our experiments showed that CCYV p22 is able to bind to ss and ds long RNAs, in addition to ss and ds small interfering (si) RNA molecules. CCYV p22 deletion mutants (MBP_CCYV DEL1-4) were produced that covered the entire protein, with MBP_CCYV DEL2 corresponding to the F-box motif and its flanking sequences. None of these deletions abolished the capacity of CCYV p22 to bind ss- and dsRNA molecules. However, deletions affecting the C-terminal half of the protein resulted in decreased binding efficiency for either ss- or dsRNA molecules indicating that essential elements for these interactions are located in this region. Taken together, our data add to current knowledge of the mode of action of suppressors of RNA silencing encoded by genes sited at the 3'-terminus of crinivirus genomic RNA 1, and shed light on the involvement of CCYV p22 in the suppression of RNA silencing and/or in another role in the virus life cycle via RNA binding.Salavert, F.; Navarro Bohigues, JA.; Owen, CA.; Khechmar, S.; Pallás Benet, V.; Livieratos, IC. (2020). Cucurbit chlorotic yellows virus p22 suppressor of RNA silencing binds single-, double-stranded long and short interfering RNA molecules in vitro. Virus Research. 279:1-8. https://doi.org/10.1016/j.virusres.2020.197887S18279Abrahamian, P. E., Seblani, R., Sobh, H., & Abou-Jawdah, Y. (2013). Detection and quantitation of two cucurbit criniviruses in mixed infection by real-time RT-PCR. Journal of Virological Methods, 193(2), 320-326. doi:10.1016/j.jviromet.2013.06.004Bananej, K., Menzel, W., Kianfar, N., Vahdat, A., & Winter, S. (2013). First Report of Cucurbit chlorotic yellows virus Infecting Cucumber, Melon, and Squash in Iran. Plant Disease, 97(7), 1005-1005. doi:10.1094/pdis-01-13-0125-pdnBrantley, J. D., & Hunt, A. G. (1993). The N-terminal protein of the polyprotein encoded by the potyvirus tobacco vein mottling virus is an RNA-binding protein. Journal of General Virology, 74(6), 1157-1162. doi:10.1099/0022-1317-74-6-1157BUCK, K. W., & KEMPSON-JONES, G. F. (1970). Three Types of Virus Particle in Penicillium stoloniferum. Nature, 225(5236), 945-946. doi:10.1038/225945a0Cañizares, M. C., Navas-Castillo, J., & Moriones, E. (2008). Multiple suppressors of RNA silencing encoded by both genomic RNAs of the crinivirus, Tomato chlorosis virus. Virology, 379(1), 168-174. doi:10.1016/j.virol.2008.06.020Chen, S., Sun, X., Shi, Y., Wei, Y., Han, X., Li, H., … Shi, Y. (2019). Cucurbit Chlorotic Yellows Virus p22 Protein Interacts with Cucumber SKP1LB1 and Its F-Box-Like Motif Is Crucial for Silencing Suppressor Activity. Viruses, 11(9), 818. doi:10.3390/v11090818Cuellar, W. J., Tairo, F., Kreuze, J. F., & Valkonen, J. P. T. (2008). Analysis of gene content in sweet potato chlorotic stunt virus RNA1 reveals the presence of the p22 RNA silencing suppressor in only a few isolates: implications for viral evolution and synergism. Journal of General Virology, 89(2), 573-582. doi:10.1099/vir.0.83471-0Cuellar, W. J., Kreuze, J. F., Rajamaki, M.-L., Cruzado, K. R., Untiveros, M., & Valkonen, J. P. T. (2009). Elimination of antiviral defense by viral RNase III. Proceedings of the National Academy of Sciences, 106(25), 10354-10358. doi:10.1073/pnas.0806042106GYOUTOKU, Y., OKAZAKI, S., FURUTA, A., ETOH, T., MIZOBE, M., KUNO, K., … OKUDA, M. (2009). Chlorotic yellows disease of melon caused by Cucurbit chlorotic yellows virus, a new crinivirus. Japanese Journal of Phytopathology, 75(2), 109-111. doi:10.3186/jjphytopath.75.109Herranz, M. C., & Pallás, V. (2004). RNA-binding properties and mapping of the RNA-binding domain from the movement protein of Prunus necrotic ringspot virus. Journal of General Virology, 85(3), 761-768. doi:10.1099/vir.0.19534-0Huang, L.-H., Tseng, H.-H., Li, J.-T., & Chen, T.-C. (2010). First Report of Cucurbit chlorotic yellows virus Infecting Cucurbits in Taiwan. Plant Disease, 94(9), 1168-1168. doi:10.1094/pdis-94-9-1168bKataya, A. R. A., Suliman, M. N. S., Kalantidis, K., & Livieratos, I. C. (2009). Cucurbit yellow stunting disorder virus p25 is a suppressor of post-transcriptional gene silencing. Virus Research, 145(1), 48-53. doi:10.1016/j.virusres.2009.06.010Klaassen, V. A., Mayhew, D., Fisher, D., & Falk, B. W. (1996). In VitroTranscripts from Cloned cDNAs of the Lettuce Infectious Yellows Closterovirus Bipartite Genomic RNAs Are Competent for Replication inNicotiana benthamianaProtoplasts. Virology, 222(1), 169-175. doi:10.1006/viro.1996.0407Kreuze, J. F., Savenkov, E. I., Cuellar, W., Li, X., & Valkonen, J. P. T. (2005). Viral Class 1 RNase III Involved in Suppression of RNA Silencing. Journal of Virology, 79(11), 7227-7238. doi:10.1128/jvi.79.11.7227-7238.2005Kubota, K., & Ng, J. C. K. (2016). Lettuce chlorosis virus P23 Suppresses RNA Silencing and Induces Local Necrosis with Increased Severity at Raised Temperatures. Phytopathology®, 106(6), 653-662. doi:10.1094/phyto-09-15-0219-rLandeo-Ríos, Y., Navas-Castillo, J., Moriones, E., & Cañizares, M. C. (2016). The p22 RNA silencing suppressor of the crinivirus Tomato chlorosis virus preferentially binds long dsRNAs preventing them from cleavage. Virology, 488, 129-136. doi:10.1016/j.virol.2015.11.008Marcos, J. F., Vilar, M., Pérez-Payá, E., & Pallás, V. (1999). In VivoDetection, RNA-Binding Properties and Characterization of the RNA-Binding Domain of the p7 Putative Movement Protein from Carnation Mottle Carmovirus (CarMV). Virology, 255(2), 354-365. doi:10.1006/viro.1998.9596Mashiko, T., Wang, W.-Q., Hartono, S., Suastica, G., Neriya, Y., Nishigawa, H., & Natsuaki, T. (2019). The p27 open reading frame of tomato infectious chlorosis virus encodes a suppressor of RNA silencing. Journal of General Plant Pathology, 85(4), 301-305. doi:10.1007/s10327-019-00850-0Navarro, J. A., Genovés, A., Climent, J., Saurí, A., Martínez-Gil, L., Mingarro, I., & Pallás, V. (2006). RNA-binding properties and membrane insertion of Melon necrotic spot virus (MNSV) double gene block movement proteins. Virology, 356(1-2), 57-67. doi:10.1016/j.virol.2006.07.040Okuda, M., Okazaki, S., Yamasaki, S., Okuda, S., & Sugiyama, M. (2010). Host Range and Complete Genome Sequence of Cucurbit chlorotic yellows virus, a New Member of the Genus Crinivirus. Phytopathology®, 100(6), 560-566. doi:10.1094/phyto-100-6-0560Orfanidou, C., Maliogka, V. I., & Katis, N. I. (2014). First Report of Cucurbit chlorotic yellows virus in Cucumber, Melon, and Watermelon in Greece. Plant Disease, 98(10), 1446-1446. doi:10.1094/pdis-03-14-0311-pdnOrfanidou, C. G., Mathioudakis, M. M., Katsarou, K., Livieratos, I., Katis, N., & Maliogka, V. I. (2019). Cucurbit chlorotic yellows virus p22 is a suppressor of local RNA silencing. Archives of Virology, 164(11), 2747-2759. doi:10.1007/s00705-019-04391-xOrílio, A. F., Fortes, I. M., & Navas-Castillo, J. (2014). Infectious cDNA clones of the crinivirus Tomato chlorosis virus are competent for systemic plant infection and whitefly-transmission. Virology, 464-465, 365-374. doi:10.1016/j.virol.2014.07.032Pumplin, N., & Voinnet, O. (2013). RNA silencing suppression by plant pathogens: defence, counter-defence and counter-counter-defence. Nature Reviews Microbiology, 11(11), 745-760. doi:10.1038/nrmicro3120Richmond, K. E., Chenault, K., Sherwood, J. L., & German, T. L. (1998). Characterization of the Nucleic Acid Binding Properties of Tomato Spotted Wilt Virus Nucleocapsid Protein. Virology, 248(1), 6-11. doi:10.1006/viro.1998.9223Salem, N. M., Chen, A. Y. S., Tzanetakis, I. E., Mongkolsiriwattana, C., & Ng, J. C. K. (2009). Further complexity of the genus Crinivirus revealed by the complete genome sequence of Lettuce chlorosis virus (LCV) and the similar temporal accumulation of LCV genomic RNAs 1 and 2. Virology, 390(1), 45-55. doi:10.1016/j.virol.2009.04.025Serra-Soriano, M., Antonio Navarro, J., & Pallás, V. (2016). Dissecting the multifunctional role of the N-terminal domain of theMelon necrotic spot viruscoat protein in RNA packaging, viral movement and interference with antiviral plant defence. Molecular Plant Pathology, 18(6), 837-849. doi:10.1111/mpp.12448Simon, A. E., & Miller, W. A. (2013). 3′ Cap-Independent Translation Enhancers of Plant Viruses. Annual Review of Microbiology, 67(1), 21-42. doi:10.1146/annurev-micro-092412-155609Voinnet, O. (2005). Induction and suppression of RNA silencing: insights from viral infections. Nature Reviews Genetics, 6(3), 206-220. doi:10.1038/nrg1555Wang, J., Stewart, L. R., Kiss, Z., & Falk, B. W. (2010). Lettuce infectious yellows virus (LIYV) RNA 1-encoded P34 is an RNA-binding protein and exhibits perinuclear localization. Virology, 403(1), 67-77. doi:10.1016/j.virol.2010.04.006Weinheimer, I., Boonrod, K., Moser, M., Wassenegger, M., Krczal, G., Butcher, S. J., & Valkonen, J. P. T. (2014). Binding and processing of small dsRNA molecules by the class 1 RNase III protein encoded by sweet potato chlorotic stunt virus. Journal of General Virology, 95(2), 486-495. doi:10.1099/vir.0.058693-0Yeh, H.-H., Tian, T., Rubio, L., Crawford, B., & Falk, B. W. (2000). Asynchronous Accumulation of Lettuce Infectious Yellows Virus RNAs 1 and 2 and Identification of an RNA 1 trans Enhancer of RNA 2 Accumulation. Journal of Virology, 74(13), 5762-5768. doi:10.1128/jvi.74.13.5762-5768.200

    In memoriam of Ricardo Flores: The career, achievements, and legacy of an inspirational plant virologist

    Get PDF
    [EN] Ricardo Flores (1947-2020) focused his research on the identification, replication, pathogenesis, and evolution of viroids, the minimal non-protein-coding circular RNAs (250-400 nt) able to replicate and incite diseases in plants that are remarkable for being at the lowest step of the biological scale. He and his collaborators initially identified and characterized additional group members, adding six new ones to the family Pospiviroidae, and expanding the Avsunviroidae from one to four members. They showed that members of the second family "encode" ribozymes, a property that, together with others, makes them candidates for being the most primitive replicons that emerged on our planet 3500 million years ago. He also made important contributions regarding how viroids replicate, providing relevant data on the templates, enzymes, and ribozymes that mediate this process and on the mutation rate, which turned out to be the highest reported for any biological entity. More recently, he concentrated on the role that RNA silencing could play on viroid-host interactions, describing details of this process. Ricardo also worked on citrus tristeza virus, a widely different type of subcellular pathogen, and made important contributions on the structure, localization and functions of its unique p23 protein. His research has produced 170 original articles and reviews, according to Web of Science. He encouraged the scientific careers of a large number of researchers, and collaborated with many others, some of whom have recapitulated his scientific legacy in this review and contributed with other chapters in this special issue.This work was supported by the Spanish Agencia Estatal de Investigaci ' on (AEI) and Fondo Europeo de Desarrollo Regional (FEDER), grant number PID2020-115571RB-100. We apologize to colleagues whose work was not cited in this review due to the page limit.Pallás Benet, V.; Hernandez Fort, C.; Marcos, JF.; Daròs, J.; Ambrós, S.; Navarro, B.; Navarro Bohigues, JA.... (2022). In memoriam of Ricardo Flores: The career, achievements, and legacy of an inspirational plant virologist. Virus Research. 312(198718):1-9. https://doi.org/10.1016/j.virusres.2022.1987181931219871

    A potyvirus vector efficiently targets recombinant proteins to chloroplasts, mitochondria and nuclei in plant cells when expressed at the amino terminus of the polyprotein

    Full text link
    Plant virus-based expression systems allow quick and efficient production of recombinant proteins in plant biofactories. Among them, a system derived from tobacco etch virus (TEV; genus potyvirus) permits coexpression of equimolar amounts of several recombinant proteins. This work analyzed how to target recombinant proteins to different subcellular localizations in the plant cell using this system. We constructed TEV clones in which green fluorescent protein (GFP), with a chloroplast transit peptide (cTP), a nuclear localization signal (NLS) or a mitochondrial targeting peptide (mTP) was expressed either as the most amino-terminal product or embedded in the viral polyprotein. Results showed that cTP and mTP mediated efficient translocation of GFP to the corresponding organelle only when present at the amino terminus of the viral polyprotein. In contrast, the NLS worked efficiently at both positions. Viruses expressing GFP in the amino terminus of the viral polyprotein produced milder symptoms. Untagged GFPs and cTP and NLS tagged amino-terminal GFPs accumulated to higher amounts in infected tissues. Finally, viral progeny from clones with internal GFPs maintained the extra gene better. These observations will help in the design of potyvirus-based vectors able to coexpress several proteins while targeting different subcellular localizations, as required in plant metabolic engineering.We thank Veronica Aragones for excellent technical assistance. We also thank Ash N. Watson (Smurfit Institute of Genetics, Trinity College Dublin, Ireland) for English edition. This research was supported by grants BIO2011-26741 and BIO2014-54269-R from Ministerio de Economia y Competitividad (MINECO, Spain) to J.-A.D. E.M. was the recipient of a predoctoral fellowship (AP2012-3751) from Ministerio de Educacion, Cultura y Deporte (Spain). J.-A.N was supported by grant BIO2011-25018 (MINECO).Majer, E.; Navarro Bohigues, JA.; Daros Arnau, JA. (2015). A potyvirus vector efficiently targets recombinant proteins to chloroplasts, mitochondria and nuclei in plant cells when expressed at the amino terminus of the polyprotein. Biotechnology Journal. 10(11):1792-1802. https://doi.org/10.1002/biot.201500042S17921802101

    A sensitive and rapid RNA silencing suppressor activity assay based on alfalfa mosaic virus expression vector

    Full text link
    [EN] Plant viruses express RNA silencing suppressor (RSS) proteins to counteract plant defence mechanisms. Here, we describe a method to assess the RSS activity based on an alfalfa mosaic virus (AMV) RNA 3 expression vector and transgenic Nicotiana tabacum plants that express the P1 and P2 subunits of the AMV replicase (P12 plants). Inoculation of P12 plants with different AMV RNA 3 constructs expressing different HC-Pro mutants that differ in their RSS capabilities, revealed a perfect correlation between necrotic lesions on inoculated leaves and RSS activity. Protoplast analysis showed that the RSS activity correlated with the accumulation of the AMV RNAs. A direct comparison between three RSS activity assays and the AMV-P12 system revealed that the coat protein of carnation mottle virus displays RSS activity with the four assays and reduced the accumulation of the siRNAs.We thank L. Corachan for her excellent technical assistance. We thank Dr. S. Elena, I2SysBio (Valencia, Spain), for supplying us the different HC-Pro mutants. This work was supported by grant BIO2017-88321-R from the Spanish Direccion General de Investigacion Cientifica y Tecnica (DGICYT) and the Prometeo Program GV2015/010 from the Generalitat Valenciana.Martínez-Pérez, M.; Navarro Bohigues, JA.; Pallás Benet, V.; Sanchez Navarro, JA. (2019). A sensitive and rapid RNA silencing suppressor activity assay based on alfalfa mosaic virus expression vector. Virus Research. 272:1-5. https://doi.org/10.1016/j.virusres.2019.197733S1527

    Molecular characterization, targeting and expression analysis of chloroplast and mitochondrion protein import components in Nicotiana benthamiana

    Full text link
    [EN] Improved bioinformatics tools for annotating gene function are becoming increasingly available, but such information must be considered theoretical until further experimental evidence proves it. In the work reported here, the genes for the main components of the translocons of the outer membrane of chloroplasts (Toc) and mitochondria (Tom), including preprotein receptors and protein-conducting channels of N. benthamiana, were identified. Sequence identity searches and phylogenetic relationships with functionally annotated sequences such as those of A. thaliana revealed that N. benthamiana orthologs mainly exist as recently duplicated loci. Only a Toc34 ortholog was found (NbToc34), while Toc159 receptor family was composed of four orthologs but somewhat different from those of A. thaliana. Except for NbToc90, the rest (NbToc120, NbToc159A and NbToc159B) had a molecular weight of about 150 kDa and an acidic domain similar in length. Only two orthologs of the Tom20 receptors, NbTom20-1 and NbTom20-2, were found. The number of the Toc and Tom receptor isoforms in N. benthamiana was comparable to that previously reported in tomato and what we found in BLAST searches in other species in the genera Nicotiana and Solanum. After cloning, the subcellular localization of N. benthamiana orthologs was studied, resulting to be identical to that of A. thaliana receptors. Phenotype analysis after silencing together with relative expression analysis in roots, stems and leaves revealed that, except for the Toc and Tom channel- forming components (NbToc75 and NbTom40) and NbToc34, functional redundancy could be observed either among Toc159 or mitochondrial receptors. Finally, heterodimer formation between NbToc34 and the NbToc159 family receptors was confirmed by two alternative techniques indicating that different Toc complexes could be assembled. Additional work needs to be addressed to know if this results in a functional specialization of each Toc complex.Funding This work was funded by grant PID2020-115571RB-I00 from the Spanish Agencia Estatal de Investigacion (AEI) and Fondo Europeo de Desarrollo Regional (FEDER). JN and MS-B are the recipients of a postdoctoral contract and a PhD fellowship from the Ministerio de Ciencia, Innovacion y Universidades of Spain, repectively.Sáiz-Bonilla, M.; Martín Merchán, A.; Pallás Benet, V.; Navarro Bohigues, JA. (2022). Molecular characterization, targeting and expression analysis of chloroplast and mitochondrion protein import components in Nicotiana benthamiana. Frontiers in Plant Science. 13:1-21. https://doi.org/10.3389/fpls.2022.10406881211
    corecore