1932

Abstract

Although viruses require cellular functions to replicate, their absolute dependence upon the host translation machinery to produce polypeptides indispensable for their reproduction is most conspicuous. Despite their incredible diversity, the mRNAs produced by all viruses must engage cellular ribosomes. This has proven to be anything but a passive process and has revealed a remarkable array of tactics for rapidly subverting control over and dominating cellular regulatory pathways that influence translation initiation, elongation, and termination. Besides enforcing viral mRNA translation, these processes profoundly impact host cell-intrinsic immune defenses at the ready to deny foreign mRNA access to ribosomes and block protein synthesis. Finally, genome size constraints have driven the evolution of resourceful strategies for maximizing viral coding capacity. Here, we review the amazing strategies that work to regulate translation in virus-infected cells, highlighting both virus-specific tactics and the tremendous insight they provide into fundamental translational control mechanisms in health and disease.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-virology-100114-055014
2016-09-29
2024-04-23
Loading full text...

Full text loading...

/deliver/fulltext/virology/3/1/annurev-virology-100114-055014.html?itemId=/content/journals/10.1146/annurev-virology-100114-055014&mimeType=html&fmt=ahah

Literature Cited

  1. Decroly E, Ferron F, Lescar J, Canard B. 1.  2012. Conventional and unconventional mechanisms for capping viral mRNA. Nat. Rev. Microbiol. 10:51–65 [Google Scholar]
  2. Dougherty JD, Reineke LC, Lloyd RE. 2.  2014. mRNA decapping enzyme 1a (Dcp1a)-induced translational arrest through protein kinase R (PKR) activation requires the N-terminal enabled vasodilator-stimulated protein homology 1 (EVH1) domain. J. Biol. Chem. 289:3936–49 [Google Scholar]
  3. Hopkins KC, McLane LM, Maqbool T, Panda D, Gordesky-Gold B, Cherry S. 3.  2013. A genome-wide RNAi screen reveals that mRNA decapping restricts bunyaviral replication by limiting the pools of Dcp2-accessible targets for cap-snatching. Genes Dev 27:1511–25 [Google Scholar]
  4. Mohr I, Sonenberg N. 4.  2012. Host translation at the nexus of infection and immunity. Cell Host Microbe 12:470–83 [Google Scholar]
  5. Walsh D, Mohr I. 5.  2006. Assembly of an active translation initiation factor complex by a viral protein. Genes Dev 20:461–72 [Google Scholar]
  6. Walsh D, Perez C, Notary J, Mohr I. 6.  2005. Regulation of the translation initiation factor eIF4F by multiple mechanisms in human cytomegalovirus-infected cells. J. Virol. 79:8057–64 [Google Scholar]
  7. Tellam JT, Zhong J, Lekieffre L, Bhat P, Martinez M. 7.  et al. 2014. mRNA Structural constraints on EBNA1 synthesis impact on in vivo antigen presentation and early priming of CD8+ T cells. PLOS Pathog 10:e1004423 [Google Scholar]
  8. Zhao Y, Pang TY, Wang Y, Wang S, Kang HX. 8.  et al. 2014. LMP1 stimulates the transcription of eIF4E to promote the proliferation, migration and invasion of human nasopharyngeal carcinoma. FEBS J 281:3004–18 [Google Scholar]
  9. Castelló A, Quintas A, Sánchez EG, Sabina P, Nogal M. 9.  et al. 2009. Regulation of host translational machinery by African swine fever virus. PLOS Pathog 5:e1000562 [Google Scholar]
  10. Zaborowska I, Kellner K, Henry M, Meleady P, Walsh D. 10.  2012. Recruitment of host translation initiation factor eIF4G by the vaccinia virus ssDNA binding protein I3. Virology 425:11–22 [Google Scholar]
  11. Walsh D, Arias C, Perez C, Halladin D, Escandon M. 11.  et al. 2008. Eukaryotic translation initiation factor 4F architectural alterations accompany translation initiation factor redistribution in poxvirus-infected cells. Mol. Cell. Biol. 28:2648–58 [Google Scholar]
  12. Katsafanas GC, Moss B. 12.  2007. Colocalization of transcription and translation within cytoplasmic poxvirus factories coordinates viral expression and subjugates host functions. Cell Host Microbe 2:221–28 [Google Scholar]
  13. Larralde O, Smith RW, Wilkie GS, Malik P, Gray NK, Clements JB. 13.  2006. Direct stimulation of translation by the multifunctional herpesvirus ICP27 protein. J. Virol. 80:1588–91 [Google Scholar]
  14. Boyne JR, Jackson BR, Taylor A, Macnab SA, Whitehouse A. 14.  2010. Kaposi's sarcoma-associated herpesvirus ORF57 protein interacts with PYM to enhance translation of viral intronless mRNAs. EMBO J 29:1851–64 [Google Scholar]
  15. Yanguez E, Rodriguez P, Goodfellow I, Nieto A. 15.  2012. Influenza virus polymerase confers independence of the cellular cap-binding factor eIF4E for viral mRNA translation. Virology 422:297–307 [Google Scholar]
  16. Mir MA, Panganiban AT. 16.  2008. A protein that replaces the entire cellular eIF4F complex. EMBO J 27:3129–39 [Google Scholar]
  17. Daughenbaugh KF, Fraser CS, Hershey JW, Hardy ME. 17.  2003. The genome-linked protein VPg of the Norwalk virus binds eIF3, suggesting its role in translation initiation complex recruitment. EMBO J 22:2852–59 [Google Scholar]
  18. Leen EN, Sorgeloos F, Correia S, Chaudhry Y, Cannac F. 18.  et al. 2016. A conserved interaction between a C-terminal motif in norovirus VPg and the HEAT-1 domain of eIF4G is essential for translation initiation. PLOS Pathog 12:e1005379 [Google Scholar]
  19. Silva LC, Almeida GM, Assis FL, Albarnaz JD, Boratto PV. 19.  et al. 2015. Modulation of the expression of mimivirus-encoded translation-related genes in response to nutrient availability during Acanthamoeba castellanii infection. Front. Microbiol. 6:539 [Google Scholar]
  20. Walsh D, Mohr I. 20.  2011. Viral subversion of the host protein synthesis machinery. Nat. Rev. Microbiol. 9:860–75 [Google Scholar]
  21. Buchkovich NJ, Yu Y, Zampieri CA, Alwine JC. 21.  2008. The TORrid affairs of viruses: effects of mammalian DNA viruses on the PI3K-Akt-mTOR signalling pathway. Nat. Rev. Microbiol. 6:266–75 [Google Scholar]
  22. Zaborowska I, Walsh D. 22.  2009. PI3K signaling regulates rapamycin-insensitive translation initiation complex formation in vaccinia virus-infected cells. J. Virol. 83:3988–92 [Google Scholar]
  23. Werden SJ, Barrett JW, Wang G, Stanford MM, McFadden G. 23.  2007. M-T5, the ankyrin repeat, host range protein of myxoma virus, activates Akt and can be functionally replaced by cellular PIKE-A. J. Virol. 81:2340–48 [Google Scholar]
  24. Chuluunbaatar U, Roller R, Feldman ME, Brown S, Shokat KM, Mohr I. 24.  2010. Constitutive mTORC1 activation by a herpesvirus Akt surrogate stimulates mRNA translation and viral replication. Genes Dev 24:2627–39 [Google Scholar]
  25. Moorman NJ, Cristea IM, Terhune SS, Rout MP, Chait BT, Shenk T. 25.  2008. Human cytomegalovirus protein UL38 inhibits host cell stress responses by antagonizing the tuberous sclerosis protein complex. Cell Host Microbe 3:253–62 [Google Scholar]
  26. Clippinger AJ, Alwine JC. 26.  2012. Dynein mediates the localization and activation of mTOR in normal and human cytomegalovirus-infected cells. Genes Dev 26:2015–26 [Google Scholar]
  27. Peng L, Liang D, Tong W, Li J, Yuan Z. 27.  2010. Hepatitis C virus NS5A activates the mammalian target of rapamycin (mTOR) pathway, contributing to cell survival by disrupting the interaction between FK506-binding protein 38 (FKBP38) and mTOR. J. Biol. Chem. 285:20870–81 [Google Scholar]
  28. Shives KD, Beatman EL, Chamanian M, O'Brien C, Hobson-Peters J, Beckham JD. 28.  2014. West Nile virus-induced activation of mammalian target of rapamycin complex 1 supports viral growth and viral protein expression. J. Virol. 88:9458–71 [Google Scholar]
  29. Shuda M, Kwun HJ, Feng H, Chang Y, Moore PS. 29.  2011. Human Merkel cell polyomavirus small T antigen is an oncoprotein targeting the 4E-BP1 translation regulator. J. Clin. Investig. 121:3623–34 [Google Scholar]
  30. Spangle JM, Munger K. 30.  2010. The human papillomavirus type 16 E6 oncoprotein activates mTORC1 signaling and increases protein synthesis. J. Virol. 84:9398–407 [Google Scholar]
  31. Abernathy E, Glaunsinger B. 31.  2015. Emerging roles for RNA degradation in viral replication and antiviral defense. Virology 479–80:600–8 [Google Scholar]
  32. Parrish S, Moss B. 32.  2007. Characterization of a second vaccinia virus mRNA-decapping enzyme conserved in poxviruses. J. Virol. 81:12973–78 [Google Scholar]
  33. Dauber B, Saffran HA, Smiley JR. 33.  2014. The herpes simplex virus 1 virion host shutoff protein enhances translation of viral late mRNAs by preventing mRNA overload. J. Virol. 88:9624–32 [Google Scholar]
  34. Muller M, Hutin S, Marigold O, Li KH, Burlingame A, Glaunsinger BA. 34.  2015. A ribonucleoprotein complex protects the interleukin-6 mRNA from degradation by distinct herpesviral endonucleases. PLOS Pathog 11:e1004899 [Google Scholar]
  35. Zhu Y, Wang X, Goff SP, Gao G. 35.  2012. Translational repression precedes and is required for ZAP-mediated mRNA decay. EMBO J 31:4236–46 [Google Scholar]
  36. Herdy B, Jaramillo M, Svitkin YV, Rosenfeld AB, Kobayashi M. 36.  et al. 2012. Translational control of the activation of transcription factor NF-κB and production of type I interferon by phosphorylation of the translation factor eIF4E. Nat. Immunol. 13:543–50 [Google Scholar]
  37. Su X, Yu Y, Zhong Y, Giannopoulou EG, Hu X. 37.  et al. 2015. Interferon-γ regulates cellular metabolism and mRNA translation to potentiate macrophage activation. Nat. Immunol. 16:838–49 [Google Scholar]
  38. Mizutani T, Fukushi S, Saijo M, Kurane I, Morikawa S. 38.  2004. Phosphorylation of p38 MAPK and its downstream targets in SARS coronavirus-infected cells. Biochem. Biophys. Res. Commun. 319:1228–34 [Google Scholar]
  39. Royall E, Doyle N, Abdul-Wahab A, Emmott E, Morley SJ. 39.  et al. 2015. Murine norovirus 1 (MNV1) replication induces translational control of the host by regulating eIF4E activity during infection. J. Biol. Chem. 290:4748–58 [Google Scholar]
  40. Arias C, Walsh D, Harbell J, Wilson AC, Mohr I. 40.  2009. Activation of host translational control pathways by a viral developmental switch. PLOS Pathog 5:e1000334 [Google Scholar]
  41. Tirosh O, Cohen Y, Shitrit A, Shani O, Le-Trilling VT. 41.  et al. 2015. The transcription and translation landscapes during human cytomegalovirus infection reveal novel host-pathogen interactions. PLOS Pathog 11:e1005288 [Google Scholar]
  42. McKinney C, Zavadil J, Bianco C, Shiflett L, Brown S, Mohr I. 42.  2014. Global reprogramming of the cellular translational landscape facilitates cytomegalovirus replication. Cell Rep 6:9–17 [Google Scholar]
  43. McKinney C, Perez C, Mohr I. 43.  2012. Poly(A) binding protein abundance regulates eukaryotic translation initiation factor 4F assembly in human cytomegalovirus-infected cells. PNAS 109:5627–32 [Google Scholar]
  44. Hopkins KC, Tartell MA, Herrmann C, Hackett BA, Taschuk F. 44.  et al. 2015. Virus-induced translational arrest through 4EBP1/2-dependent decay of 5′-TOP mRNAs restricts viral infection. PNAS 112:E2920–29 [Google Scholar]
  45. Mata MA, Satterly N, Versteeg GA, Frantz D, Wei S. 45.  et al. 2011. Chemical inhibition of RNA viruses reveals REDD1 as a host defense factor. Nat. Chem. Biol. 7:712–19 [Google Scholar]
  46. Patel RK, Hardy RW. 46.  2012. Role for the phosphatidylinositol 3-kinase-Akt-TOR pathway during Sindbis virus replication in arthropods. J. Virol. 86:3595–604 [Google Scholar]
  47. Shaheen ZR, Naatz A, Corbett JA. 47.  2015. CCR5-dependent activation of mTORC1 regulates translation of inducible NO synthase and COX-2 during encephalomyocarditis virus infection. J. Immunol. 195:4406–14 [Google Scholar]
  48. Joubert PE, Stapleford K, Guivel-Benhassine F, Vignuzzi M, Schwartz O, Albert ML. 48.  2015. Inhibition of mTORC1 enhances the translation of chikungunya proteins via the activation of the MnK/eIF4E pathway. PLOS Pathog 11:e1005091 [Google Scholar]
  49. Colina R, Costa-Mattioli M, Dowling RJ, Jaramillo M, Tai LH. 49.  et al. 2008. Translational control of the innate immune response through IRF-7. Nature 452:323–28 [Google Scholar]
  50. White JP, Reineke LC, Lloyd RE. 50.  2011. Poliovirus switches to an eIF2-independent mode of translation during infection. J. Virol. 85:8884–93 [Google Scholar]
  51. Sukarieh R, Sonenberg N, Pelletier J. 51.  2010. Nuclear assortment of eIF4E coincides with shut-off of host protein synthesis upon poliovirus infection. J. Gen. Virol. 91:1224–28 [Google Scholar]
  52. Groppo R, Brown BA, Palmenberg AC. 52.  2010. Mutational analysis of the EMCV 2A protein identifies a nuclear localization signal and an eIF4E binding site. Virology 410:257–67 [Google Scholar]
  53. de Breyne S, Chamond N, Decimo D, Trabaud MA, Andre P. 53.  et al. 2012. In vitro studies reveal that different modes of initiation on HIV-1 mRNA have different levels of requirement for eukaryotic initiation factor 4F. FEBS J 279:3098–111 [Google Scholar]
  54. Soto-Rifo R, Rubilar PS, Ohlmann T. 54.  2013. The DEAD-box helicase DDX3 substitutes for the cap-binding protein eIF4E to promote compartmentalized translation initiation of the HIV-1 genomic RNA. Nucleic Acids Res 41:6286–99 [Google Scholar]
  55. Sharma A, Yilmaz A, Marsh K, Cochrane A, Boris-Lawrie K. 55.  2012. Thriving under stress: selective translation of HIV-1 structural protein mRNA during Vpr-mediated impairment of eIF4E translation activity. PLOS Pathog 8:e1002612 [Google Scholar]
  56. Lee AS, Burdeinick-Kerr R, Whelan SP. 56.  2014. A genome-wide small interfering RNA screen identifies host factors required for vesicular stomatitis virus infection. J. Virol. 88:8355–60 [Google Scholar]
  57. Xi Q, Cuesta R, Schneider RJ. 57.  2004. Tethering of eIF4G to adenoviral mRNAs by viral 100k protein drives ribosome shunting. Genes Dev 18:1997–2009 [Google Scholar]
  58. Pooggin MM, Rajeswaran R, Schepetilnikov MV, Ryabova LA. 58.  2012. Short ORF-dependent ribosome shunting operates in an RNA picorna-like virus and a DNA pararetrovirus that cause rice tungro disease. PLOS Pathog 8:e1002568 [Google Scholar]
  59. Au HH, Jan E. 59.  2014. Novel viral translation strategies. Wiley Interdiscip. Rev. RNA 5:779–801 [Google Scholar]
  60. Sweeney TR, Abaeva IS, Pestova TV, Hellen CU. 60.  2014. The mechanism of translation initiation on Type 1 picornavirus IRESs. EMBO J 33:76–92 [Google Scholar]
  61. Kafasla P, Morgner N, Robinson CV, Jackson RJ. 61.  2010. Polypyrimidine tract-binding protein stimulates the poliovirus IRES by modulating eIF4G binding. EMBO J 29:3710–22 [Google Scholar]
  62. Yu Y, Abaeva IS, Marintchev A, Pestova TV, Hellen CU. 62.  2011. Common conformational changes induced in type 2 picornavirus IRESs by cognate trans-acting factors. Nucleic Acids Res 39:4851–65 [Google Scholar]
  63. Rojas-Araya B, Ohlmann T, Soto-Rifo R. 63.  2015. Translational control of the HIV unspliced genomic RNA. Viruses 7:4326–51 [Google Scholar]
  64. Locker N, Chamond N, Sargueil B. 64.  2011. A conserved structure within the HIV gag open reading frame that controls translation initiation directly recruits the 40S subunit and eIF3. Nucleic Acids Res 39:2367–77 [Google Scholar]
  65. Plank TD, Whitehurst JT, Kieft JS. 65.  2013. Cell type specificity and structural determinants of IRES activity from the 5′ leaders of different HIV-1 transcripts. Nucleic Acids Res 41:6698–714 [Google Scholar]
  66. Othman Z, Sulaiman MK, Willcocks MM, Ulryck N, Blackbourn DJ. 66.  et al. 2014. Functional analysis of Kaposi's sarcoma-associated herpesvirus vFLIP expression reveals a new mode of IRES-mediated translation. RNA 20:1803–14 [Google Scholar]
  67. Quade N, Boehringer D, Leibundgut M, van den Heuvel J, Ban N. 67.  2015. Cryo-EM structure of hepatitis C virus IRES bound to the human ribosome at 3.9-Å resolution. Nat. Commun. 6:7646 [Google Scholar]
  68. Spahn CM, Kieft JS, Grassucci RA, Penczek PA, Zhou K. 68.  et al. 2001. Hepatitis C virus IRES RNA-induced changes in the conformation of the 40S ribosomal subunit. Science 291:1959–62 [Google Scholar]
  69. Hashem Y, des Georges A, Dhote V, Langlois R, Liao HY. 69.  et al. 2013. Hepatitis-C-virus-like internal ribosome entry sites displace eIF3 to gain access to the 40S subunit. Nature 503:539–43 [Google Scholar]
  70. Matsuda D, Mauro VP. 70.  2014. Base pairing between hepatitis C virus RNA and 18S rRNA is required for IRES-dependent translation initiation in vivo. PNAS 111:15385–89 [Google Scholar]
  71. Muhs M, Hilal T, Mielke T, Skabkin MA, Sanbonmatsu KY. 71.  et al. 2015. Cryo-EM of ribosomal 80S complexes with termination factors reveals the translocated cricket paralysis virus IRES. Mol. Cell 57:422–32 [Google Scholar]
  72. Fernandez IS, Bai XC, Murshudov G, Scheres SH, Ramakrishnan V. 72.  2014. Initiation of translation by cricket paralysis virus IRES requires its translocation in the ribosome. Cell 157:823–31 [Google Scholar]
  73. Colussi TM, Costantino DA, Zhu J, Donohue JP, Korostelev AA. 73.  et al. 2015. Initiation of translation in bacteria by a structured eukaryotic IRES RNA. Nature 519:110–13 [Google Scholar]
  74. Spahn CM, Jan E, Mulder A, Grassucci RA, Sarnow P, Frank J. 74.  2004. Cryo-EM visualization of a viral internal ribosome entry site bound to human ribosomes: The IRES functions as an RNA-based translation factor. Cell 118:465–75 [Google Scholar]
  75. Au HH, Cornilescu G, Mouzakis KD, Ren Q, Burke JE. 75.  et al. 2015. Global shape mimicry of tRNA within a viral internal ribosome entry site mediates translational reading frame selection. PNAS 112:E6446–55 [Google Scholar]
  76. Li S, Peters GA, Ding K, Zhang X, Qin J, Sen GC. 76.  2006. Molecular basis for PKR activation by PACT or dsRNA. PNAS 103:10005–10 [Google Scholar]
  77. Tycowski KT, Guo YE, Lee N, Moss WN, Vallery TK. 77.  et al. 2015. Viral noncoding RNAs: more surprises. Genes Dev 29:567–84 [Google Scholar]
  78. Goodman AG, Fornek JL, Medigeshi GR, Perrone LA, Peng X. 78.  et al. 2009. P58IPK: a novel “CIHD” member of the host innate defense response against pathogenic virus infection. PLOS Pathog 5:e1000438 [Google Scholar]
  79. Mudhasani R, Tran JP, Retterer C, Kota KP, Whitehouse CA, Bavari S. 79.  2016. Protein kinase R degradation is essential for Rift Valley fever virus infection and is regulated by SKP1-CUL1-F-box (SCF)FBXW11-NSs E3 ligase. PLOS Pathog 12:e1005437 [Google Scholar]
  80. Hakki M, Marshall EE, De Niro KL, Geballe AP. 80.  2006. Binding and nuclear relocalization of protein kinase R by human cytomegalovirus TRS1. J. Virol. 80:11817–26 [Google Scholar]
  81. Li JJ, Cao C, Fixsen SM, Young JM, Ono C. 81.  et al. 2015. Baculovirus protein PK2 subverts eIF2α kinase function by mimicry of its kinase domain C-lobe. PNAS 112:E4364–73 [Google Scholar]
  82. Seo EJ, Liu F, Kawagishi-Kobayashi M, Ung TL, Cao C. 82.  et al. 2008. Protein kinase PKR mutants resistant to the poxvirus pseudosubstrate K3L protein. PNAS 105:16894–99 [Google Scholar]
  83. Peng C, Haller SL, Rahman MM, McFadden G, Rothenburg S. 83.  2016. Myxoma virus M156 is a specific inhibitor of rabbit PKR but contains a loss-of-function mutation in Australian virus isolates. PNAS 113:3855–60 [Google Scholar]
  84. Mulvey M, Arias C, Mohr I. 84.  2007. Maintenance of endoplasmic reticulum (ER) homeostasis in herpes simplex virus type 1-infected cells through the association of a viral glycoprotein with PERK, a cellular ER stress sensor. J. Virol. 81:3377–90 [Google Scholar]
  85. Pavio N, Romano PR, Graczyk TM, Feinstone SM, Taylor DR. 85.  2003. Protein synthesis and endoplasmic reticulum stress can be modulated by the hepatitis C virus envelope protein E2 through the eukaryotic initiation factor 2α kinase PERK. J. Virol. 77:3578–85 [Google Scholar]
  86. Berlanga JJ, Ventoso I, Harding HP, Deng J, Ron D. 86.  et al. 2006. Antiviral effect of the mammalian translation initiation factor 2α kinase GCN2 against RNA viruses. EMBO J 25:1730–40 [Google Scholar]
  87. Won S, Eidenschenk C, Arnold CN, Siggs OM, Sun L. 87.  et al. 2012. Increased susceptibility to DNA virus infection in mice with a GCN2 mutation. J. Virol. 86:1802–8 [Google Scholar]
  88. Rojas M, Vasconcelos G, Dever TE. 88.  2015. An eIF2α-binding motif in protein phosphatase 1 subunit GADD34 and its viral orthologs is required to promote dephosphorylation of eIF2α.. PNAS 112:E3466–75 [Google Scholar]
  89. Clavarino G, Claudio N, Couderc T, Dalet A, Judith D. 89.  et al. 2012. Induction of GADD34 is necessary for dsRNA-dependent interferon-β production and participates in the control of chikungunya virus infection. PLOS Pathog 8:e1002708 [Google Scholar]
  90. Zhang F, Moon A, Childs K, Goodbourn S, Dixon LK. 90.  2010. The African swine fever virus DP71L protein recruits the protein phosphatase 1 catalytic subunit to dephosphorylate eIF2α and inhibits CHOP induction but is dispensable for these activities during virus infection. J. Virol. 84:10681–89 [Google Scholar]
  91. Sanchez R, Mohr I. 91.  2007. Inhibition of cellular 2′-5′ oligoadenylate synthetase by the herpes simplex virus type 1 Us11 protein. J. Virol. 81:3455–64 [Google Scholar]
  92. Sciortino MT, Parisi T, Siracusano G, Mastino A, Taddeo B, Roizman B. 92.  2013. The virion host shutoff RNase plays a key role in blocking the activation of protein kinase R in cells infected with herpes simplex virus 1. J. Virol. 87:3271–76 [Google Scholar]
  93. White SD, Jacobs BL. 93.  2012. The amino terminus of the vaccinia virus E3 protein is necessary to inhibit the interferon response. J. Virol. 86:5895–904 [Google Scholar]
  94. Elde NC, Child SJ, Eickbush MT, Kitzman JO, Rogers KS. 94.  et al. 2012. Poxviruses deploy genomic accordions to adapt rapidly against host antiviral defenses. Cell 150:831–41 [Google Scholar]
  95. Rothenburg S, Seo EJ, Gibbs JS, Dever TE, Dittmar K. 95.  2009. Rapid evolution of protein kinase PKR alters sensitivity to viral inhibitors. Nat. Struct. Mol. Biol. 16:63–70 [Google Scholar]
  96. Burgess HM, Mohr I. 96.  2015. Cellular 5′-3′ mRNA exonuclease Xrn1 controls double-stranded RNA accumulation and anti-viral responses. Cell Host Microbe 17:332–44 [Google Scholar]
  97. Liu SW, Katsafanas GC, Liu R, Wyatt LS, Moss B. 97.  2015. Poxvirus decapping enzymes enhance virulence by preventing the accumulation of dsRNA and the induction of innate antiviral responses. Cell Host Microbe 17:320–31 [Google Scholar]
  98. Shabman RS, Hoenen T, Groseth A, Jabado O, Binning JM. 98.  et al. 2013. An upstream open reading frame modulates Ebola virus polymerase translation and virus replication. PLOS Pathog 9:e1003147 [Google Scholar]
  99. Robert F, Kapp LD, Khan SN, Acker MG, Kolitz S. 99.  et al. 2006. Initiation of protein synthesis by hepatitis C virus is refractory to reduced eIF2·GTP·Met-tRNAiMet ternary complex availability. Mol. Biol. Cell 17:4632–44 [Google Scholar]
  100. Dmitriev SE, Terenin IM, Andreev DE, Ivanov PA, Dunaevsky JE. 100.  et al. 2010. GTP-independent tRNA delivery to the ribosomal P-site by a novel eukaryotic translation factor. J. Biol. Chem. 285:26779–87 [Google Scholar]
  101. Kim JH, Park SM, Park JH, Keum SJ, Jang SK. 101.  2011. eIF2A mediates translation of hepatitis C viral mRNA under stress conditions. EMBO J 30:2454–64 [Google Scholar]
  102. Ventoso I, Sanz MA, Molina S, Berlanga JJ, Carrasco L, Esteban M. 102.  2006. Translational resistance of late alphavirus mRNA to eIF2α phosphorylation: a strategy to overcome the antiviral effect of protein kinase PKR. Genes Dev 20:87–100 [Google Scholar]
  103. Skabkin MA, Skabkina OV, Dhote V, Komar AA, Hellen CU, Pestova TV. 103.  2010. Activities of ligatin and MCT-1/DENR in eukaryotic translation initiation and ribosomal recycling. Genes Dev 24:1787–801 [Google Scholar]
  104. Pestova TV, de Breyne S, Pisarev AV, Abaeva IS, Hellen CU. 104.  2008. eIF2-dependent and eIF2-independent modes of initiation on the CSFV IRES: a common role of domain II. EMBO J 27:1060–72 [Google Scholar]
  105. Garcia-Moreno M, Sanz MA, Pelletier J, Carrasco L. 105.  2013. Requirements for eIF4A and eIF2 during translation of Sindbis virus subgenomic mRNA in vertebrate and invertebrate host cells. Cell Microbiol 15:823–40 [Google Scholar]
  106. Ventoso I. 106.  2012. Adaptive changes in alphavirus mRNA translation allowed colonization of vertebrate hosts. J. Virol. 86:9484–94 [Google Scholar]
  107. Garrey JL, Lee YY, Au HH, Bushell M, Jan E. 107.  2010. Host and viral translational mechanisms during cricket paralysis virus infection. J. Virol. 84:1124–38 [Google Scholar]
  108. Ruehle MD, Zhang H, Sheridan RM, Mitra S, Chen Y. 108.  et al. 2015. A dynamic RNA loop in an IRES affects multiple steps of elongation factor-mediated translation initiation. eLife 4:e08146 [Google Scholar]
  109. Colussi TM, Costantino DA, Hammond JA, Ruehle GM, Nix JC, Kieft JS. 109.  2014. The structural basis of transfer RNA mimicry and conformational plasticity by a viral RNA. Nature 511:366–69 [Google Scholar]
  110. Andreev DE, Hirnet J, Terenin IM, Dmitriev SE, Niepmann M, Shatsky IN. 110.  2012. Glycyl-tRNA synthetase specifically binds to the poliovirus IRES to activate translation initiation. Nucleic Acids Res 40:5602–14 [Google Scholar]
  111. Li M, Kao E, Gao X, Sandig H, Limmer K. 111.  et al. 2012. Codon-usage-based inhibition of HIV protein synthesis by human schlafen 11. Nature 491:125–28 [Google Scholar]
  112. Wang Q, Lee I, Ren J, Ajay SS, Lee YS, Bao X. 112.  2013. Identification and functional characterization of tRNA-derived RNA fragments (tRFs) in respiratory syncytial virus infection. Mol. Ther. 21:368–79 [Google Scholar]
  113. Costafreda MI, Perez-Rodriguez FJ, D'Andrea L, Guix S, Ribes E. 113.  et al. 2014. Hepatitis A virus adaptation to cellular shutoff is driven by dynamic adjustments of codon usage and results in the selection of populations with altered capsids. J. Virol. 88:5029–41 [Google Scholar]
  114. Ishimaru D, Plant EP, Sims AC, Yount BL Jr., Roth BM. 114.  et al. 2013. RNA dimerization plays a role in ribosomal frameshifting of the SARS coronavirus. Nucleic Acids Res 41:2594–608 [Google Scholar]
  115. Gao F, Simon AE. 115.  2016. Multiple cis-acting elements modulate programmed −1 ribosomal frameshifting in pea enation mosaic virus. Nucleic Acids Res 44:878–95 [Google Scholar]
  116. Jagger BW, Wise HM, Kash JC, Walters KA, Wills NM. 116.  et al. 2012. An overlapping protein-coding region in influenza A virus segment 3 modulates the host response. Science 337:199–204 [Google Scholar]
  117. Li Y, Treffers EE, Napthine S, Tas A, Zhu L. 117.  et al. 2014. Transactivation of programmed ribosomal frameshifting by a viral protein. PNAS 111:E2172–81 [Google Scholar]
  118. Fang Y, Treffers EE, Li Y, Tas A, Sun Z. 118.  et al. 2012. Efficient −2 frameshifting by mammalian ribosomes to synthesize an additional arterivirus protein. PNAS 109:E2920–28 [Google Scholar]
  119. Loughran G, Firth AE, Atkins JF. 119.  2011. Ribosomal frameshifting into an overlapping gene in the 2B-encoding region of the cardiovirus genome. PNAS 108:E1111–19 [Google Scholar]
  120. Ren Q, Wang QS, Firth AE, Chan MM, Gouw JW. 120.  et al. 2012. Alternative reading frame selection mediated by a tRNA-like domain of an internal ribosome entry site. PNAS 109:E630–39 [Google Scholar]
  121. Zinoviev A, Hellen CU, Pestova TV. 121.  2015. Multiple mechanisms of reinitiation on bicistronic calicivirus mRNAs. Mol. Cell 57:1059–73 [Google Scholar]
  122. Thiebeauld O, Schepetilnikov M, Park HS, Geldreich A, Kobayashi K. 122.  et al. 2009. A new plant protein interacts with eIF3 and 60S to enhance virus-activated translation re-initiation. EMBO J 28:3171–84 [Google Scholar]
  123. Watanabe Y, Ohtaki N, Hayashi Y, Ikuta K, Tomonaga K. 123.  2009. Autogenous translational regulation of the Borna disease virus negative control factor X from polycistronic mRNA using host RNA helicases. PLOS Pathog 5:e1000654 [Google Scholar]
  124. Kronstad LM, Brulois KF, Jung JU, Glaunsinger BA. 124.  2014. Reinitiation after translation of two upstream open reading frames (ORF) governs expression of the ORF35–37 Kaposi's sarcoma-associated herpesvirus polycistronic mRNA. J. Virol. 88:6512–18 [Google Scholar]
  125. Nicholson BL, White KA. 125.  2014. Functional long-range RNA-RNA interactions in positive-strand RNA viruses. Nat. Rev. Microbiol. 12:493–504 [Google Scholar]
  126. Jeudy S, Abergel C, Claverie JM, Legendre M. 126.  2012. Translation in giant viruses: a unique mixture of bacterial and eukaryotic termination schemes. PLOS Genet 8:e1003122 [Google Scholar]
  127. Houck-Loomis B, Durney MA, Salguero C, Shankar N, Nagle JM. 127.  et al. 2011. An equilibrium-dependent retroviral mRNA switch regulates translational recoding. Nature 480:561–64 [Google Scholar]
  128. Machida K, Mikami S, Masutani M, Mishima K, Kobayashi T, Imataka H. 128.  2014. A translation system reconstituted with human factors proves that processing of encephalomyocarditis virus proteins 2A and 2B occurs in the elongation phase of translation without eukaryotic release factors. J. Biol. Chem. 289:31960–71 [Google Scholar]
  129. Brubaker SW, Gauthier AE, Mills EW, Ingolia NT, Kagan JC. 129.  2014. A bicistronic MAVS transcript highlights a class of truncated variants in antiviral immunity. Cell 156:800–11 [Google Scholar]
  130. Belew AT, Meskauskas A, Musalgaonkar S, Advani VM, Sulima SO. 130.  et al. 2014. Ribosomal frameshifting in the CCR5 mRNA is regulated by miRNAs and the NMD pathway. Nature 512:265–69 [Google Scholar]
  131. Hafren A, Eskelin K, Makinen K. 131.  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]
  132. Cervantes-Salazar M, Angel-Ambrocio AH, Soto-Acosta R, Bautista-Carbajal P, Hurtado-Monzon AM. 132.  et al. 2015. Dengue virus NS1 protein interacts with the ribosomal protein RPL18: This interaction is required for viral translation and replication in Huh-7 cells. Virology 484:113–26 [Google Scholar]
  133. Green L, Houck-Loomis B, Yueh A, Goff SP. 133.  2012. Large ribosomal protein 4 increases efficiency of viral recoding sequences. J. Virol. 86:8949–58 [Google Scholar]
  134. Wu X, He WT, Tian S, Meng D, Li Y. 134.  et al. 2014. pelo is required for high efficiency viral replication. PLOS Pathog 10:e1004034 [Google Scholar]
  135. Majzoub K, Hafirassou ML, Meignin C, Goto A, Marzi S. 135.  et al. 2014. RACK1 controls IRES-mediated translation of viruses. Cell 159:1086–95 [Google Scholar]
  136. Hertz MI, Landry DM, Willis AE, Luo G, Thompson SR. 136.  2013. Ribosomal protein S25 dependency reveals a common mechanism for diverse internal ribosome entry sites and ribosome shunting. Mol. Cell. Biol. 33:1016–26 [Google Scholar]
  137. Lee AS, Burdeinick-Kerr R, Whelan SP. 137.  2013. A ribosome-specialized translation initiation pathway is required for cap-dependent translation of vesicular stomatitis virus mRNAs. PNAS 110:324–29 [Google Scholar]
  138. Kamitani W, Huang C, Narayanan K, Lokugamage KG, Makino S. 138.  2009. A two-pronged strategy to suppress host protein synthesis by SARS coronavirus Nsp1 protein. Nat. Struct. Mol. Biol. 16:1134–40 [Google Scholar]
  139. Jack K, Bellodi C, Landry DM, Niederer RO, Meskauskas A. 139.  et al. 2011. rRNA pseudouridylation defects affect ribosomal ligand binding and translational fidelity from yeast to human cells. Mol. Cell 44:660–66 [Google Scholar]
  140. Zhang R, Jha BK, Ogden KM, Dong B, Zhao L. 140.  et al. 2013. Homologous 2′,5′-phosphodiesterases from disparate RNA viruses antagonize antiviral innate immunity. PNAS 110:13114–19 [Google Scholar]
  141. Sorgeloos F, Jha BK, Silverman RH, Michiels T. 141.  2013. Evasion of antiviral innate immunity by Theiler's virus L* protein through direct inhibition of RNase L. PLOS Pathog 9:e1003474 [Google Scholar]
  142. McKinney C, Yu D, Mohr I. 142.  2013. A new role for the cellular PABP repressor Paip2 as an innate restriction factor capable of limiting productive cytomegalovirus replication. Genes Dev 27:1809–20 [Google Scholar]
  143. Tahiri-Alaoui A, Zhao Y, Sadigh Y, Popplestone J, Kgosana L. 143.  et al. 2014. Poly(A) binding protein 1 enhances cap-independent translation initiation of neurovirulence factor from avian herpesvirus. PLOS ONE 9:e114466 [Google Scholar]
  144. Suzuki Y, Chin WX, Han Q, Ichiyama K, Lee CH. 144.  et al. 2016. Characterization of RyDEN (C19orf66) as an interferon-stimulated cellular inhibitor against dengue virus replication. PLOS Pathog 12:e1005357 [Google Scholar]
  145. Blakqori G, van Knippenberg I, Elliott RM. 145.  2009. Bunyamwera orthobunyavirus S-segment untranslated regions mediate poly(A) tail-independent translation. J. Virol. 83:3637–46 [Google Scholar]
  146. Gratia M, Sarot E, Vende P, Charpilienne A, Baron CH. 146.  et al. 2015. Rotavirus nsp3 is a translational surrogate of the poly(A) binding protein-poly(A) complex. J. Virol. 89:8773–82 [Google Scholar]
  147. Arnold MM, Brownback CS, Taraporewala ZF, Patton JT. 147.  2012. Rotavirus variant replicates efficiently although encoding an aberrant NSP3 that fails to induce nuclear localization of poly(A)-binding protein. J. Gen. Virol. 93:1483–94 [Google Scholar]
  148. Montero H, Arias CF, Lopez S. 148.  2006. Rotavirus nonstructural protein NSP3 is not required for viral protein synthesis. J. Virol. 80:9031–38 [Google Scholar]
  149. Polacek C, Friebe P, Harris E. 149.  2009. Poly(A)-binding protein binds to the non-polyadenylated 3′ untranslated region of dengue virus and modulates translation efficiency. J. Gen. Virol. 90:687–92 [Google Scholar]
  150. Bai Y, Zhou K, Doudna JA. 150.  2013. Hepatitis C virus 3′UTR regulates viral translation through direct interactions with the host translation machinery. Nucleic Acids Res 41:7861–74 [Google Scholar]
  151. Garcia-Moreno M, Sanz MA, Carrasco L. 151.  2016. A viral mRNA motif at the 3′-untranslated region that confers translatability in a cell-specific manner. Implications for virus evolution. Sci. Rep. 6:19217 [Google Scholar]
  152. Simon AE, Miller WA. 152.  2013. 3′ cap–independent translation enhancers of plant viruses. Annu. Rev. Microbiol. 67:21–42 [Google Scholar]
  153. Gao F, Kasprzak WK, Szarko C, Shapiro BA, Simon AE. 153.  2014. The 3′ untranslated region of pea enation mosaic virus contains two T-shaped, ribosome-binding, cap-independent translation enhancers. J. Virol. 88:11696–712 [Google Scholar]
  154. Zuo X, Wang J, Yu P, Eyler D, Xu H. 154.  et al. 2010. Solution structure of the cap-independent translational enhancer and ribosome-binding element in the 3′ UTR of turnip crinkle virus. PNAS 107:1385–90 [Google Scholar]
  155. Mazumder B, Poddar D, Basu A, Kour R, Verbovetskaya V, Barik S. 155.  2014. Extraribosomal L13a is a specific innate immune factor for antiviral defense. J. Virol. 88:9100–10 [Google Scholar]
  156. Kobayashi M, Wilson AC, Chao MV, Mohr I. 156.  2012. Control of viral latency in neurons by axonal mTOR signaling and the 4E-BP translation repressor. Genes Dev 26:1527–32 [Google Scholar]
  157. Sanfacon H. 157.  2015. Plant translation factors and virus resistance. Viruses 7:3392–419 [Google Scholar]
  158. Parsons J, Castaldi MP, Dutta S, Dibrov SM, Wyles DL, Hermann T. 158.  2009. Conformational inhibition of the hepatitis C virus internal ribosome entry site RNA. Nat. Chem. Biol. 5:823–25 [Google Scholar]
  159. McMahon R, Zaborowska I, Walsh D. 159.  2011. Noncytotoxic inhibition of viral infection through eIF4F-independent suppression of translation by 4EGi-1. J. Virol. 85:853–64 [Google Scholar]
  160. Ritchie DB, Soong J, Sikkema WK, Woodside MT. 160.  2014. Anti-frameshifting ligand reduces the conformational plasticity of the SARS virus pseudoknot. J. Am. Chem. Soc. 136:2196–99 [Google Scholar]
  161. Kaufman HL, Kohlhapp FJ, Zloza A. 161.  2015. Oncolytic viruses: a new class of immunotherapy drugs. Nat. Rev. Drug Discov. 14:642–62 [Google Scholar]
  162. Mohr IJ, Pe'ery T, Mathews MB. 162.  2007. Protein synthesis and translational control during viral infection. Translational Control in Biology and Medicine MB Mathews, N Sonenberg, JWB Hershey 545–95 Cold Spring Harbor, NY: Cold Spring Harb. Lab. Press [Google Scholar]
/content/journals/10.1146/annurev-virology-100114-055014
Loading
/content/journals/10.1146/annurev-virology-100114-055014
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • 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