1932

Abstract

A successful infection by a plant virus results from the complex molecular interplay between the host plant and the invading virus. Thus, dissecting the molecular network of virus-host interactions advances the understanding of the viral infection process and may assist in the development of novel antiviral strategies. In the past decade, molecular identification and functional characterization of host factors in the virus life cycle, particularly single-stranded, positive-sense RNA viruses, have been a research focus in plant virology. As a result, a number of host factors have been identified. These host factors are implicated in all the major steps of the infection process. Some host factors are diverted for the viral genome translation, some are recruited to improvise the viral replicase complexes for genome multiplication, and others are components of transport complexes for cell-to-cell spread via plasmodesmata and systemic movement through the phloem. This review summarizes current knowledge about host factors and discusses future research directions.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-phyto-080614-120001
2015-08-04
2024-04-20
Loading full text...

Full text loading...

/deliver/fulltext/phyto/53/1/annurev-phyto-080614-120001.html?itemId=/content/journals/10.1146/annurev-phyto-080614-120001&mimeType=html&fmt=ahah

Literature Cited

  1. Agbeci M, Grangeon R, Nelson RS, Zheng H, Laliberté J-F. 1.  2013. Contribution of host intracellular transport machineries to intercellular movement of Turnip mosaic virus. PLOS Pathog. 9:e1003683 [Google Scholar]
  2. Ahlquist P, Noueiry AO, Lee WM, Kushner DB, Dye BT. 2.  2003. Host factors in positive-strand RNA virus genome replication. J. Virol. 15:8181–86 [Google Scholar]
  3. Alcaide-Loridan C, Jupin I. 3.  2012. Ubiquitin and plant viruses, let's play together. Plant Physiol. 160:72–82 [Google Scholar]
  4. Amari K, Boutant E, Hofmann C, Schmitt-Keichinger C, Fernandez-Calvino L. 4.  et al. 2010. A family of plasmodesmal proteins with receptor-like properties for plant viral movement proteins. PLOS Pathog. 6:e1001119 [Google Scholar]
  5. Amari K, Di Donato M, Dolja VV, Heinlein M. 5.  2014. Myosins VIII and XI play distinct roles in reproduction and transport of Tobacco mosaic virus. PLOS Pathog. 10:e1004448 [Google Scholar]
  6. Amari K, Lerich A, Schitt-Keichinger C, Dolja VV, Ritzenthaler C. 6.  2011. Tubule-guided cell-to-cell movement of a plant virus requires class XI myosin motors. PLOS Pathog. 7:e1002327 [Google Scholar]
  7. Ashby J, Boutant E, Seemanpilai M, Sambade A, Ritzenthaler C. 7.  et al. 2006. Tobacco mosaic virus movement protein functions as a structural microtubule-associated protein. J. Virol. 80:8329–44 [Google Scholar]
  8. Barajas D, de Castro Martín IF, Pogany J, Risco C, Nagy PD. 8.  2014. Noncanonical role for the host Vps AAA+ ATPase ESCRT protein in the formation of Tomato bushy stunt virus replicase. PLOS Pathog. 10:e1004087 [Google Scholar]
  9. Barajas D, Jiang Y, Nagy PD. 9.  2009. A unique role for the host ESCRT proteins in replication of Tomato bushy stunt virus. PLOS Pathog. 5:e1000705 [Google Scholar]
  10. Barajas D, Nagy PD. 10.  2010. Ubiquitination of tombusvirus p33 replication protein plays a role in virus replication and binding to the host Vps23p ESCRT protein. Virology 397:358–68 [Google Scholar]
  11. Barajas D, Xu K, de Castro Martín IF, Sasvari Z, Brandizzi F. 11.  et al. 2014. Co-opted oxysterol-binding ORD and VAP proteins channel sterols to RNA virus replication sites via membrane contact sites. PLOS Pathog. 10:e1004388 [Google Scholar]
  12. Beauchemin C, Boutet N, Laliberté JF. 12.  2007. Visualization of the interaction between the precursors of VPg, the viral protein linked to the genome of Turnip mosaic virus, and the translation eukaryotic initiation factor (iso) 4E in planta. J. Virol. 81:775–82 [Google Scholar]
  13. Boggio R, Chiocca S. 13.  2006. Viruses and SUMOylation: recent highlights. Curr. Opin. Microbiol. 9:430–36 [Google Scholar]
  14. Brandner K, Sambade A, Boutant E, Didier P, Mély Y. 14.  et al. 2008. Tobacco mosaic virus movement protein interacts with green fluorescent protein–tagged microtubule end-binding protein. Plant Physiol. 147:611–23 [Google Scholar]
  15. Brisco MJ, Hull R, Wilson TMA. 15.  1985. Southern bean mosaic virus–specific proteins are synthesized in an in vitro system supplemented with intact, treated virions. Virology 143:392–98 [Google Scholar]
  16. Bucher GL, Tarina C, Heinlein M, Di Serio F, Meins F Jr. 16.  et al. 2001. Local expression of enzymatically active class I β-1,3-glucanase enhances symptoms of TMV infection in tobacco. Plant J. 28:361–39 [Google Scholar]
  17. Camborde L, Planchais S, Tournier V, Jakubiec A, Drugeon G. 17.  et al. 2010. The ubiquitin-proteasome system regulates the accumulation of Turnip yellow mosaic virus RNA-dependent RNA polymerase during viral infection. Plant Cell 22:3142–52 [Google Scholar]
  18. Carroll D. 18.  2014. Genome engineering with targetable nucleases. Annu. Rev. Biochem. 83:409–39 [Google Scholar]
  19. Chen CE, Yeh KC, Wu SH, Wang HI, Yeh HH. 19.  2013. A vicilin-like seed storage protein, PAP85, is involved in Tobacco mosaic virus replication. J. Virol. 87:6888–900 [Google Scholar]
  20. Chen M-H, Citovsky V. 20.  2003. Systemic movement of a tobamovirus requires host cell pectin methylesterase. Plant J. 35:386–92 [Google Scholar]
  21. Chen M-H, Sheng J, Hind G, Handa AK, Citovsky V. 21.  2000. Interaction between the Tobacco mosaic virus movement protein and host cell pectin methylesterases is required for viral cell-to-cell movement. EMBO J. 19:913–20 [Google Scholar]
  22. Chenon M, Camborde L, Cheminant S, Jupin I. 22.  2012. A viral deubiquitylating enzyme targets viral RNA-dependent RNA polymerase and affects viral infectivity. EMBO J. 31:741–53 [Google Scholar]
  23. Chuang C, Barajas D, Qin J, Nagy PD. 23.  2014. Inactivation of the host Lipin gene accelerates RNA virus replication through viral exploitation of the expanded endoplasmic reticulum membrane. PLOS Pathog. 10:e1003944 [Google Scholar]
  24. Cohen P. 24.  2000. The regulation of protein function by multisite phosphorylation: a 25 year update. Trends Biochem. Sci. 25:596–601 [Google Scholar]
  25. Culver JN. 25.  2002. Tobacco mosaic virus assembly: determinants in pathogenicity and resistance. Annu. Rev. Phytopathol. 40:287–308 [Google Scholar]
  26. Decroocq V, Sicard O, Alamillo JM, Lansac M, Eyquard JP. 26.  et al. 2006. Multiple resistance traits control Plum pox virus infection in Arabidopsis thaliana. Mol. Plant-Microbe Interact. 19:541–49 [Google Scholar]
  27. den Boon JA, Diaz A, Ahlquist P. 27.  2010. Cytoplasmic viral replication complexes. Cell Host Microbe 8:77–85 [Google Scholar]
  28. Diaz A, Wang X. 28.  2014. Bromovirus-induced remodeling of host membranes during viral RNA replication. Curr. Opin. Virol. 9:104–10 [Google Scholar]
  29. Diaz A, Wang X, Ahlquist P. 29.  2010. Membrane-shaping host reticulon proteins play crucial roles in viral RNA replication compartment formation and function. Proc. Natl. Acad. Sci. USA 107:16291–96 [Google Scholar]
  30. Díez J, Ishikawa M, Kaido M, Ahlquist P. 30.  2000. Identification and characterization of a host protein required for efficient template selection in viral RNA replication. Proc. Natl. Acad. Sci. USA 97:3913–18 [Google Scholar]
  31. Dreher TW, Uhlenbeck OC, Browning KS. 31.  1999. Quantitative assessment of EF-1α GTP binding to aminoacyl-tRNAs, aminoacyl-viral RNA, and tRNA shows close correspondence to the RNA binding properties of EF-Tu. J. Biol. Chem. 274:666–72 [Google Scholar]
  32. D'Souza-Schorey C, Chavrier P. 32.  2006. ARF proteins: roles in membrane traffic and beyond. Nat. Rev. Mol. Cell Biol. 7:347–58 [Google Scholar]
  33. Dufresne PJ, Thivierge K, Cotton S, Beauchemin C, Ide C. 33.  et al. 2008. Heat shock 70 protein interaction with Turnip mosaic virus RNA-dependent RNA polymerase within virus-induced membrane vesicles. Virology 374:217–27 [Google Scholar]
  34. Dunoyer P, Thomas C, Harrison S, Revers F, Maule A. 34.  2004. A cysteine-rich plant protein potentiates potyvirus movement through an interaction with the virus genome-linked protein VPg. J. Virol. 78:2301–309 [Google Scholar]
  35. Faulkner C, Akman OE, Bell K, Jeffree C, Oparka K. 35.  2008. Peeking into pit fields: a multiple twinning model of secondary plasmodesmata formation in tobacco. Plant Cell 20:1504–18 [Google Scholar]
  36. Fridborg I, Grainger J, Page A, Coleman M, Findlay K, Angell S. 36.  2003. TIP, a novel host factor linking callose degradation with the cell-to-cell movement of Potato virus X. Mol. Plant-Microbe Interact. 16:132–40 [Google Scholar]
  37. Gebauer F, Hentze MW. 37.  2004. Molecular mechanisms of translational control. Nat. Rev. Mol. Cell Biol. 5:827–35 [Google Scholar]
  38. Geiss-Friedlander R, Melchior F. 38.  2007. Concepts in SUMOylation: a decade on. Nat. Rev. Mol. Cell Biol. 8:947–56 [Google Scholar]
  39. Hafrén A, Eskelin K, Mäkinen K. 39.  2013. Ribosomal protein P0 promotes Potato virus A infection and functions in viral translation together with VPg and eIF(iso)4E. J. Virol. 87:4302–12 [Google Scholar]
  40. Hafrén A, Hofius D, Rönnholm G, Sonnewald U, Mäkinen K. 40.  2010. HSP70 and its cochaperone CPIP promote potyvirus infection in Nicotiana benthamiana by regulating viral coat protein functions. Plant Cell 22:523–35 [Google Scholar]
  41. Hafrén A, Mäkinen K. 41.  2008. Purification of viral genome-linked protein VPg from potato virus A–infected plants reveals several posttranslationally modified forms of the protein. J. Gen. Virol. 89:1509–18 [Google Scholar]
  42. Harries P, Ding B. 42.  2011. Cellular factors in plant virus movement: at the leading edge of macromolecular trafficking in plants. Virology 411:237–43 [Google Scholar]
  43. Heinlein M. 43.  2015. Plasmodesmata: channels for viruses on the move. Methods Mol. Biol. 1217:25–52 [Google Scholar]
  44. Hilliker A, Gao Z, Jankowsky E, Parker R. 44.  2011. The DEAD-box protein Ded1 modulates translation by the formation and resolution of an eIF4F-mRNA complex. Mol. Cell 43:962–72 [Google Scholar]
  45. Hipper C, Brault V, Ziegler-Graff V, Revers F. 45.  2013. Viral and cellular factors involved in phloem transport of plant viruses. Front. Plant Sci. 4:154 [Google Scholar]
  46. Huang TS, Nagy PD. 46.  2011. Direct inhibition of tombusvirus plus-strand RNA synthesis by a dominant negative mutant of a host metabolic enzyme, glyceraldehyde-3-phosphate dehydrogenase, in yeast and plants. J. Virol. 85:9090–102 [Google Scholar]
  47. Huang TS, Wei T, Laliberté J-F, Wang A. 47.  2010. A host RNA helicase-like protein, AtRH8, interacts with the potyviral genome-linked protein, VPg, associates with the virus accumulation complex, and is essential for infection. Plant Physiol. 152:255–66 [Google Scholar]
  48. Huang YW, Hu CC, Liou MR, Chang BY, Tsai CH. 48.  et al. 2012. Hsp90 interacts specifically with viral RNA and differentially regulates replication initiation of Bamboo mosaic virus and associated satellite RNA. PLOS Pathog. 8e1002726
  49. Hwang J, Oh CS, Kang BC. 49.  2013. Translation elongation factor 1B (eEF1B) is an essential host factor for Tobacco mosaic virus infection in plants. Virology 439:105–14 [Google Scholar]
  50. Hyodo K, Mine A, Taniguchi T, Kaido M, Mise K. 50.  et al. 2013. ADP ribosylation factor 1 plays an essential role in the replication of a plant RNA virus. J. Virol. 87:163–76 [Google Scholar]
  51. Hyodo K, Okuno T. 51.  2014. Host factors used by positive-strand RNA plant viruses for genome replication. J. Gen. Plant Pathol. 80:123–35 [Google Scholar]
  52. Iglesias VA, Meins F Jr. 52.  2000. Movement of plant viruses is delayed in a β-1,3-glucanase-deficient mutant showing a reduced plasmodesmatal size exclusion limit and enhanced callose deposition. Plant J. 21:157–66 [Google Scholar]
  53. Ishikawa M, Okada Y. 53.  2004. Replication of tobamovirus RNA. Proc. Jpn. Acad. Ser. B 80:215–24 [Google Scholar]
  54. Jakubiec A, Jupin I. 54.  2007. Regulation of positive-strand RNA virus replication: the emerging role of phosphorylation. Virus Res. 129:73–79 [Google Scholar]
  55. Jakubiec A, Tournier V, Drugeon G, Pflieger S, Camborde L. 55.  et al. 2006. Phosphorylation of viral RNA-dependent RNA polymerase and its role in replication of a plus-strand RNA virus. J. Biol. Chem. 281:21236–49 [Google Scholar]
  56. Jiang Y, Serviene E, Gal J, Panavas T, Nagy PD. 56.  2006. Identification of essential host factors affecting tombusvirus RNA replication based on the yeast Tet promoters Hughes Collection. J. Virol. 80:7394–404 [Google Scholar]
  57. Jonczyk M, Pathak KB, Sharma M, Nagy PD. 57.  2007. Exploiting alternative subcellular location for replication: tombusvirus replication switches to the endoplasmic reticulum in the absence of peroxisomes. Virology 362:320–30 [Google Scholar]
  58. Kang HK, Yang SH, Lee YP, Park YI, Kim SH. 58.  2012. A tobacco CBL-interacting protein kinase homolog is involved in phosphorylation of the N-terminal domain of the cucumber mosaic virus polymerase 2a protein. Biosci. Biotechnol. Biochem. 76:2101–6 [Google Scholar]
  59. Kawakami S, Watanabe Y, Beachy RN. 59.  2004. Tobacco mosaic virus infection spreads cell to cell as intact replication complexes. Proc. Natl. Acad. Sci. USA 101:6291–96 [Google Scholar]
  60. Kawamura-Nagaya K, Ishibashi K, Huang Y-P, Miyashita S, Ishikawa M. 60.  2014. Replication protein of Tobacco mosaic virus cotranslationally binds 5′ untranslated region of genomic RNA to enable viral replication. Proc. Natl. Acad. Sci. USA 111:E1620–28 [Google Scholar]
  61. Kim MJ, Huh SU, Ham B-K, Paek K-H. 61.  2008. A novel methyltransferase methylates Cucumber mosaic virus 1a protein and promotes systemic spread. J. Virol. 82:4823–33 [Google Scholar]
  62. Kim SH, Macfarlane S, Kalinina NO, Rakitina DV, Ryabov EV. 62.  et al. 2007. Interaction of a plant virus–encoded protein with the major nucleolar protein fibrillarin is required for systemic virus infection. Proc. Natl. Acad. Sci. USA 104:11115–20 [Google Scholar]
  63. Kim SH, Palukaitis P, Park YI. 63.  2002. Phosphorylation of Cucumber mosaic virus RNA polymerase 2a protein inhibits formation of replicase complex. EMBO J. 21:2292–300 [Google Scholar]
  64. Kim SH, Ryabov EV, Kalinina NO, Rakitina DV, Gillespie T. 64.  et al. 2007. Cajal bodies and the nucleolus are required for a plant virus systemic infection. EMBO J. 26:2169–79 [Google Scholar]
  65. Koenig R, Lesemann D-E, Pfeilstetter E. 65.  2009. New isolates of Carnation Italian ringspot virus differ from the original one by having replication-associated proteins with a typical tombusvirus-like N-terminus and by inducing peroxisome- rather than mitochondrion-derived multivesicular bodies. Arch. Virol. 154:1695–98 [Google Scholar]
  66. Kovalev N, Barajas D, Nagy PD. 66.  2012. Similar roles for yeast Dbp2 and Arabidopsis RH20 DEAD-box RNA helicases to Ded1 helicase in tombusvirus plus-strand synthesis. Virology 432:470–84 [Google Scholar]
  67. Kovalev N, Nagy PD. 67.  2014. The expanding functions of cellular helicases: the tombusvirus RNA replication enhancer co-opts the plant eIFAIII-like AtRH2 and the DDX5-like AtRH5 DEAD-box RNA helicases to promote viral asymmetric RNA replication. PLOS Pathog. 10:e1004051 [Google Scholar]
  68. Kovalev N, Pogany J, Nagy PD. 68.  2012. A co-opted DEAD-box RNA helicase enhances tombusvirus plus-strand synthesis. PLOS Pathog. 8:e1002537 [Google Scholar]
  69. Kushner DB, Lindenbach BD, Grdzelishvili VZ, Noueiry AO, Paul SM. 69.  et al. 2003. Systematic, genome-wide identification of host genes affecting replication of a positive-strand RNA virus. Proc. Natl. Acad. Sci. USA 100:15764–69 [Google Scholar]
  70. Kusumanegara K, Mine A, Hyodo K, Kaido M, Mise K. 70.  et al. 2012. Identification of domains in p27 auxiliary replicase protein essential for its association with the endoplasmic reticulum membranes in Red clover necrotic mosaic virus. Virology 433:131–41 [Google Scholar]
  71. Laliberté JF, Sanfaçon H. 71.  2010. Cellular remodeling during plant virus infection. Annu. Rev. Phytopathol. 48:69–91 [Google Scholar]
  72. Lartey RT, Ghoshoroy S, Citovsky V. 72.  1998. Identification of an Arabidopsis thaliana mutation (vsm1) that restricts systemic movement of tobamoviruses. Mol. Plant-Microbe Interact. 11:706–9 [Google Scholar]
  73. Lee JY, Taoka K, Yoo BC, Ben-Nissan G, Kim DJ. 73.  et al. 2005. Plasmodesmal-associated protein kinase in tobacco and Arabidopsis recognizes a subset of non-cell-autonomous proteins. Plant Cell 17:2817–31 [Google Scholar]
  74. Lee WM, Ahlquist P. 74.  2003. Membrane synthesis, specific lipid requirements, and localized lipid composition changes associated with a positive-strand RNA virus RNA replication protein. J. Virol. 77:12819–28 [Google Scholar]
  75. Lee WM, Ishikawa M, Ahlquist P. 75.  2001. Mutation of host Δ9 fatty acid desaturase inhibits Brome mosaic virus RNA replication between template recognition and RNA synthesis. J. Virol. 75:2097–106 [Google Scholar]
  76. Lellis AD, Kasschau KD, Whitham SA, Carrington JC. 76.  2002. Loss-of-susceptibility mutants of Arabidopsis thaliana reveal an essential role for eIF(iso)4E during potyvirus infection. Curr. Biol. 12:1046–51 [Google Scholar]
  77. Levy A, Erlanger M, Rosenthal M, Epel BL. 77.  2007. A plasmodesmata-associated beta-1,3-glucanase in Arabidopsis. Plant J. 49:669–82 [Google Scholar]
  78. Lewis JD, Lazarowitz SG. 78.  2010. Arabidopsis synaptotagmin SYTA regulates endocytosis and virus movement protein cell-to-cell transport. Proc. Natl. Acad. Sci. USA 107:2491–96 [Google Scholar]
  79. Li Y, Wu MY, Song HH, Hu X, Qiu BS. 79.  2005. Identification of a tobacco protein interacting with Tomato mosaic virus coat protein and facilitating long-distance movement of virus. Arch. Virol. 150:1993–2008 [Google Scholar]
  80. Li Z, Barajas D, Panavas T, Herbst DA, Nagy PD. 80.  2008. Cdc34p ubiquitin-conjugating enzyme is a component of the tombusvirus replicase complex and ubiquitinates p33 replication protein. J. Virol. 82:6911–26 [Google Scholar]
  81. Li Z, Gonzalez PA, Sasvari Z, Kinzy TG, Nagy PD. 81.  2014. Methylation of translation elongation factor 1A by the METTL10-like See1 methyltransferase facilitates tombusvirus replication in yeast and plants. Virology 448:43–54 [Google Scholar]
  82. Li Z, Pogany J, Panavas T, Xu K, Esposito AM. 82.  et al. 2009. Translation elongation factor 1A is a component of the tombusvirus replicase complex and affects the stability of the p33 replication co-factor. Virology 385:245–60 [Google Scholar]
  83. Li Z, Pogany J, Tupman S, Esposito AM, Kinzy TG. 83.  et al. 2010. Translation elongation factor 1A facilitates the assembly of the tombusvirus replicase and stimulates minus-strand synthesis. PLOS Pathog. 6:e1001175 [Google Scholar]
  84. Lionetti V, Raiola A, Cervone F, Bellincampi D. 84.  2014. Transgenic expression of pectin methylesterase inhibitors limits tobamovirus spread in tobacco and Arabidopsis. Mol. Plant Pathol. 15:265–74 [Google Scholar]
  85. Matsuda D, Dreher TW. 85.  2004. The tRNA-like structure of Turnip yellow mosaic virus RNA is a 3′-translational enhancer. Virology 321:36–46 [Google Scholar]
  86. Matsuda D, Yoshinari S, Dreher TW. 86.  2004. eEF1A binding to aminoacylated viral RNA represses minus strand synthesis by TYMV RNA-dependent RNA polymerase. Virology 321:47–56 [Google Scholar]
  87. Maule AJ. 87.  2008. Plasmodesmata: structure, function and biogenesis. Curr. Opin. Plant Biol. 11:680–86 [Google Scholar]
  88. McCartney AW, Greenwood JS, Fabian MR, White KA, Mullen RT. 88.  2005. Localization of the tomato bushy stunt virus replication protein p33 reveals a peroxisome-to-endoplasmic reticulum sorting pathway. Plant Cell 17:3513–31 [Google Scholar]
  89. Michon T, Estevez Y, Walter J, German-Retana S, Le Gall O. 89.  2006. The potyviral virus genome-linked protein VPg forms a ternary complex with the eukaryotic initiation factors eIF4E and eIF4G and reduces eIF4E affinity for a mRNA cap analogue. FEBS J. 273:1312–22 [Google Scholar]
  90. Mine A, Hyodo K, Tajima Y, Kusumanegara K, Taniguchi T. 90.  et al. 2012. Differential roles of Hsp70 and Hsp90 in the assembly of the replicase complex of a positive-strand RNA plant virus. J. Virol. 86:12091–104 [Google Scholar]
  91. Mine A, Takeda A, Taniguchi T, Taniguchi H, Kaido M. 91.  et al. 2010. Identification and characterization of the 480-kilodalton template-specific RNA-dependent RNA polymerase complex of Red clover necrotic mosaic virus. J. Virol. 84:6070–81 [Google Scholar]
  92. Nagy PD, Pogany J. 92.  2012. The dependence of viral RNA replication on co-opted host factors. Nat. Rev. Microbiol. 10:137–49 [Google Scholar]
  93. Nicaise V, Gallois JL, Chafiai F, Allen LM, Schurdi-Levraud V. 93.  et al. 2007. Coordinated and selective recruitment of eIF4E and eIF4G factors for potyvirus infection in Arabidopsis thaliana. FEBS Lett. 581:1041–46 [Google Scholar]
  94. Nishikiori M, Dohi K, Mori M, Meshi T, Naito S. 94.  et al. 2006. Membrane-bound Tomato mosaic virus replication proteins participate in RNA synthesis and are associated with host proteins in a pattern distinct from those that are not membrane bound. J. Virol. 80:8459–68 [Google Scholar]
  95. Nishikiori M, Mori M, Dohi K, Okamura H, Katoh E. 95.  et al. 2011. A host small GTP-binding protein ARL8 plays crucial roles in tobamovirus RNA replication. PLOS Pathog. 7:e1002409 [Google Scholar]
  96. Noueiry AO, Chen J, Ahlquist P. 96.  2000. A mutant allele of essential, general translation initiation factor DED1 selectively inhibits translation of a viral mRNA. Proc. Natl. Acad. Sci. USA 97:12985–90 [Google Scholar]
  97. Noueiry AO, Díez J, Falk SP, Chen J, Ahlquist P. 97.  2003. Yeast Lsm1p-7p/Pat1p deadenylation-dependent mRNA-decapping factors are required for Brome mosaic virus genomic RNA translation. Mol. Cell. Biol. 23:4094–106 [Google Scholar]
  98. Nziengui H, Schoefs B. 98.  2009. Functions of reticulons in plants: what we can learn from animals and yeasts. Cell. Mol. Life Sci. 66:584–95 [Google Scholar]
  99. Osman TA, Buck KW. 99.  1997. The Tobacco mosaic virus RNA polymerase complex contains a plant protein related to the RNA-binding subunit of yeast eIF-3. J. Virol. 71:6075–82 [Google Scholar]
  100. Ouko MO, Sambade A, Brandner K, Niehl A, Peña E. 100.  et al. 2010. Tobacco mutants with reduced microtubule dynamics are less susceptible to TMV. Plant J. 62:829–39 [Google Scholar]
  101. Pagny G, Paulstephenraj PS, Poque S, Sicard O, Cosson P. 101.  et al. 2012. Family-based linkage and association mapping reveals novel genes affecting Plum pox virus infection in Arabidopsis thaliana. New Phytol. 196:873–86 [Google Scholar]
  102. Panavas T, Serviene E, Brasher J, Nagy PD. 102.  2005. Yeast genome-wide screen reveals dissimilar sets of host genes affecting replication of RNA viruses. Proc. Natl. Acad. Sci. USA 102:7326–31 [Google Scholar]
  103. Pathak KB, Sasvari Z, Nagy PD. 103.  2008. The host Pex19p plays a role in peroxisomal localization of tombusvirus replication proteins. Virology 379:294–305 [Google Scholar]
  104. Pogany J, Stork J, Li Z, Nagy PD. 104.  2008. In vitro assembly of the Tomato bushy stunt virus replicase requires the host heat shock protein 70. Proc. Natl. Acad. Sci. USA 105:19956–61 [Google Scholar]
  105. Prasanth KR, Huang YW, Liou MR, Wang RY, Hu CC. 105.  et al. 2011. Glyceraldehyde 3-phosphate dehydrogenase negatively regulates the replication of Bamboo mosaic virus and its associated satellite RNA. J. Virol. 85:8829–40 [Google Scholar]
  106. Prod'homme D, Le Panse S, Drugeon D, Jupin I. 106.  2001. Detection and subcellular localization of the Turnip yellow mosaic virus 66K replication protein in infected cells. Virology 281:88–101 [Google Scholar]
  107. Puustinen P, Mäkinen K. 107.  2002. Detection of the potyviral genome-linked protein VPg in virions and its phosphorylation by host kinases. J. Virol. 76:12703–11 [Google Scholar]
  108. Quadt R, Kao CC, Browning KS, Hershberger RP, Ahlquist P. 108.  1993. Characterization of a host protein associated with brome mosaic virus RNA-dependent RNA polymerase. Proc. Natl. Acad. Sci. USA 90:1498–502 [Google Scholar]
  109. Reichel C, Beachy RN. 109.  1998. Tobacco mosaic virus infection induces severe morphological changes of the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 95:11169–74 [Google Scholar]
  110. Roenhorst JW, Verduin BJM, Goldbach RW. 110.  1989. Virus ribosome complexes from cell-free translation systems supplemented with cowpea chlorotic mottle virus particles. Virology 168:138–46 [Google Scholar]
  111. Sambade A, Brandner K, Hofman C, Seemanpillai M, Mutterer J. 111.  et al. 2008. Transport of TMV movement protein particles associated with the targeting of RNA to plasmodesmata. Traffic 9:2073–88 [Google Scholar]
  112. Sasvari Z, Gonzalez PA, Rachubinski RA, Nagy PD. 112.  2013. Tombusvirus replication depends on Sec39p endoplasmic reticulum–associated transport protein. Virology 447:21–31 [Google Scholar]
  113. Sasvari Z, Izotova L, Kinzy TG, Nagy PD. 113.  2011. Synergistic roles of eukaryotic translation elongation factors 1Bγ and 1A in stimulation of tombusvirus minus-strand synthesis. PLOS Pathog. 7:e1002438 [Google Scholar]
  114. Saxena P, Lomonossoff GP. 114.  2014. Virus infection cycle events coupled to RNA replication. Annu. Rev. Phytopathol. 52:197–212 [Google Scholar]
  115. Schwartz M, Chen J, Janda M, Sullivan M, den Boon J. 115.  et al. 2002. A positive-strand RNA virus replication complex parallels form and function of retrovirus capsids. Mol. Cell 9:505–14 [Google Scholar]
  116. Schoelz JE, Harries PA, Nelson RS. 116.  2011. Intracellular transport of plant viruses: finding the door out of the cell. Mol. Plant 4:813–31 [Google Scholar]
  117. Serrano C, González-Cruz J, Jauregui F, Medina C, Mancilla P. 117.  et al. 2008. Genetic and histological studies on the delayed systemic movement of Tobacco mosaic virus in Arabidopsis thaliana. BMC Genet. 9:59 [Google Scholar]
  118. Serva S, Nagy PD. 118.  2006. Proteomics analysis of the tombusvirus replicase: Hsp70 molecular chaperone is associated with the replicase and enhances viral RNA replication. J. Virol. 80:2162–69 [Google Scholar]
  119. Serviene E, Shapka N, Cheng CP, Panavas T, Phuangrat B. 119.  et al. 2005. Genome-wide screen identifies host genes affecting viral RNA recombination. Proc. Natl. Acad. Sci. USA 102:10545–50 [Google Scholar]
  120. Shah Nawaz-ul-Rehman M, Martinez-Ochoa N, Pascal H, Sasvari Z, Herbst C. 120.  et al. 2012. Proteome-wide overexpression of host proteins for identification of factors affecting tombusvirus RNA replication: an inhibitory role of protein kinase C. J. Virol. 86:9384–95 [Google Scholar]
  121. Shapka N, Stork J, Nagy PD. 121.  2005. Phosphorylation of the p33 replication protein of cucumber necrosis tombusvirus adjacent to the RNA binding site affects viral RNA replication. Virology 343:65–78 [Google Scholar]
  122. Sharma M, Sasvari Z, Nagy PD. 122.  2010. Inhibition of sterol biosynthesis reduces tombusvirus replication in yeast and plants. J. Virol. 84:2270–81 [Google Scholar]
  123. Sharma M, Sasvari Z, Nagy PD. 123.  2011. Inhibition of phospholipid biosynthesis decreases the activity of the tombusvirus replicase and alters the subcellular localization of replication proteins. Virology 415:141–52 [Google Scholar]
  124. Stork J, Panaviene Z, Nagy PD. 124.  2005. Inhibition of in vitro RNA binding and replicase activity by phosphorylation of the p33 replication protein of Cucumber necrosis tombusvirus. Virology 343:79–92 [Google Scholar]
  125. Tagami Y, Watanabe Y. 125.  2007. Effects of brefeldin A on the localization of tobamovirus movement protein and cell-to-cell movement of the virus. Virology 361:133–40 [Google Scholar]
  126. Thivierge K, Cotton S, Dufresne PJ, Mathieu I, Beauchemin C. 126.  et al. 2008. Eukaryotic elongation factor 1A interacts with Turnip mosaic virus RNA-dependent RNA polymerase and VPg-Pro in virus-induced vesicles. Virology 377:216–25 [Google Scholar]
  127. Tilsner J, Oparka KJ. 127.  2012. Missing links? The connection between replication and movement of plant RNA viruses. Curr. Opin. Virol. 2:705–11 [Google Scholar]
  128. Tomita Y, Mizuno T, Díez J, Naito S, Ahlquist P. 128.  et al. 2003. Mutation of host DnaJ homolog inhibits brome mosaic virus negative-strand RNA synthesis. J. Virol. 77:2990–97 [Google Scholar]
  129. Torrance L, Andreev IA, Gabrenaite-Verhovskaya R, Cowan G, Mäkinen K. 129.  et al. 2005. An unusual structure at one end of potato potyvirus particles. J. Mol. Biol. 357:1–8 [Google Scholar]
  130. Tsujimoto Y, Numaga T, Ohshima K, Yano MA, Ohsawa R. 130.  et al. 2003. Arabidopsis TOBAMOVIRUS MULTIPLICATION (TOM) 2 locus encodes a transmembrane protein that interacts with TOM1. EMBO J. 22:335–43 [Google Scholar]
  131. Turner KA, Sit TL, Callaway AS, Allen NS, Lommel SA. 131.  2004. Red clover necrotic mosaic virus replication proteins accumulate at the endoplasmic reticulum. Virology 320:276–90 [Google Scholar]
  132. Ueki S, Sperkto R, Natale DM, Citovsky V. 132.  2010. ANK, a host cytoplasmic receptor for the Tobacco mosaic virus cell-to-cell movement protein, facilitates intercellular transport through plasmodesmata. PLOS Pathog. 6:e1001201 [Google Scholar]
  133. Verchot J. 133.  2012. Cellular chaperones and folding enzymes are vital contributors to membrane bound replication and movement complexes during plant RNA virus infection. Front. Plant Sci. 3:275 [Google Scholar]
  134. Vijayapalani P, Maeshima M, Nagasaki-Takekuchi N, Miller WA. 134.  2012. Interaction of the trans-frame potyvirus protein P3N-PIPO with host protein PCaP1 facilitates potyvirus movement. PLOS Pathog. 8e1002639
  135. Wang A, Krishnaswamy S. 135.  2012. Eukaryotic translation initiation factor 4E-mediated recessive resistance to plant viruses and its utility in crop improvement. Mol. Plant Pathol. 13:795–803 [Google Scholar]
  136. Wang RY, Nagy PD. 136.  2008. Tomato bushy stunt virus co-opts the RNA-binding function of a host metabolic enzyme for viral genomic RNA synthesis. Cell Host Microbe 3:178–87 [Google Scholar]
  137. Wang RY, Stork J, Nagy PD. 137.  2009. A key role for heat shock protein 70 in the localization and insertion of tombusvirus replication proteins to intracellular membranes. J. Virol. 83:3276–87 [Google Scholar]
  138. Wang X, Diaz A, Hao L, Gancarz B, den Boon JA. 138.  et al. 2011. Intersection of the multivesicular body pathway and lipid homeostasis in RNA replication by a positive-strand RNA virus. J. Virol. 85:5494–503 [Google Scholar]
  139. Wei T, Huang T-S, McNeil JR, Laliberté J-F, Hong J. 139.  et al. 2010. Sequential recruitment of the endoplasmic reticulum and chloroplasts for plant potyvirus replication. J. Virol. 84:799–809 [Google Scholar]
  140. Wei T, Wang A. 140.  2008. Biogenesis of cytoplasmic membranous vesicles for plant potyvirus replication occurs at endoplasmic reticulum exit sites in a COPI- and COPII-dependent manner. J. Virol. 82:12252–64 [Google Scholar]
  141. Wei T, Zhang C, Hong J, Xiong R, Kasschau KD. 141.  et al. 2010. Formation of complexes at plasmodesmata for potyvirus intercellular movement is mediated by the viral protein P3N-PIPO. PLOS Pathog. 6:e1000962 [Google Scholar]
  142. Wei T, Zhang C, Hou X, Sanfaçon H, Wang A. 142.  2013. The SNARE protein Syp71 is essential for Turnip mosaic virus infection by mediating fusion of virus-induced vesicles with chloroplasts. PLOS Pathog. 9:e1003378 [Google Scholar]
  143. Wilson TMA. 143.  1984. Cotranslational disassembly of tobacco mosaic virus in vitro. Virology 137:255–65 [Google Scholar]
  144. Wittmann S, Chatel H, Fortin MG, Laliberté J-F. 144.  1997. Interaction of the viral protein genome linked of Turnip mosaic potyvirus with the translational eukaryotic initiation factor (iso)4E of Arabidopsis thaliana using the yeast two-hybrid system. Virology 234:84–92 [Google Scholar]
  145. Wu X, Shaw JG. 145.  1997. Evidence that a viral replicase protein is involved in the disassembly of tobacco mosaic virus particles in vivo. Virology 239:426–34 [Google Scholar]
  146. Xiong R, Wang A. 146.  2013. SCE1, the SUMO-conjugating enzyme in plants that interacts with NIb, the RNA-dependent RNA polymerase of Turnip mosaic virus, is required for viral infection. J. Virol. 87:4704–15 [Google Scholar]
  147. Yamaji Y, Kobayashi T, Hamada K, Sakurai K, Yoshii A. 147.  et al. 2006. In vivo interaction between Tobacco mosaic virus RNA-dependent RNA polymerase and host translation elongation factor 1A. Virology 347:100–8 [Google Scholar]
  148. Yamanaka T, Imai T, Satoh R, Kawashima A, Takahashi M. 148.  et al. 2002. Complete inhibition of tobamovirus multiplication by simultaneous mutations in two homologous host genes. J. Virol. 76:2491–97 [Google Scholar]
  149. Zavaliev R, Levy A, Gera A, Epel BL. 149.  2013. Subcellular dynamics and role of Arabidopsis β-1,3-glucanases in cell-to-cell movement of tobamoviruses. Mol. Plant-Microbe Interact. 26:1016–30 [Google Scholar]
  150. Zhang J, Diaz A, Mao L, Ahlquist P, Wang X. 150.  2012. Host acyl coenzyme A binding protein regulates replication complex assembly and activity of a positive-strand RNA virus. J. Virol. 86:5110–21 [Google Scholar]
/content/journals/10.1146/annurev-phyto-080614-120001
Loading
/content/journals/10.1146/annurev-phyto-080614-120001
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error