1932

Abstract

Herpesviral mRNAs are produced and translated by cellular machinery, rendering them susceptible to the network of regulatory events that impact translation. In response, these viruses have evolved to infiltrate and hijack translational control pathways as well as to integrate specialized host translation strategies into their own repertoire. They are robust systems to dissect mechanisms of mammalian translational regulation and continue to offer insight into -acting mRNA features that impact assembly and activity of the translation apparatus. Here, I discuss recent advances revealing the extent to which the three herpesvirus subfamilies regulate both host and viral translation, thereby dramatically impacting the landscape of protein synthesis in infected cells.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-virology-100114-054839
2015-11-09
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/virology/2/1/annurev-virology-100114-054839.html?itemId=/content/journals/10.1146/annurev-virology-100114-054839&mimeType=html&fmt=ahah

Literature Cited

  1. Jackson RJ, Hellen CU, Pestova TV. 1.  2010. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11:113–27 [Google Scholar]
  2. Ma XM, Blenis J. 2.  2009. Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol. 10:307–18 [Google Scholar]
  3. Walsh D, Mohr I. 3.  2006. Assembly of an active translation initiation factor complex by a viral protein. Genes Dev. 20:461–72 [Google Scholar]
  4. Chuluunbaatar U, Roller R, Feldman ME, Brown S, Shokat KM, Mohr I. 4.  2010. Constitutive mTORC1 activation by a herpesvirus Akt surrogate stimulates mRNA translation and viral replication. Genes Dev. 24:2627–39A viral kinase mimics the substrate specificity of Akt to activate mTORC1. [Google Scholar]
  5. Benetti L, Roizman B. 5.  2006. Protein kinase B/Akt is present in activated form throughout the entire replicative cycle of ΔUS3 mutant virus but only at early times after infection with wild-type herpes simplex virus 1. J. Virol. 80:3341–48 [Google Scholar]
  6. Moorman NJ, Cristea IM, Terhune SS, Rout MP, Chait BT, Shenk T. 6.  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]
  7. Arias C, Walsh D, Harbell J, Wilson AC, Mohr I. 7.  2009. Activation of host translational control pathways by a viral developmental switch. PLOS Pathog. 5:e1000334 [Google Scholar]
  8. Sodhi A, Chaisuparat R, Hu J, Ramsdell AK, Manning BD. 8.  et al. 2006. The TSC2/mTOR pathway drives endothelial cell transformation induced by the Kaposi's sarcoma-associated herpesvirus G protein-coupled receptor. Cancer Cell 10:133–43 [Google Scholar]
  9. Moody CA, Scott RS, Amirghahari N, Nathan CO, Young LS. 9.  et al. 2005. Modulation of the cell growth regulator mTOR by Epstein-Barr virus-encoded LMP2A. J. Virol. 79:5499–506 [Google Scholar]
  10. Thoreen CC, Kang SA, Chang JW, Liu Q, Zhang J. 10.  et al. 2009. An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1. J. Biol. Chem. 284:8023–32 [Google Scholar]
  11. Clippinger AJ, Maguire TG, Alwine JC. 11.  2011. The changing role of mTOR kinase in the maintenance of protein synthesis during human cytomegalovirus infection. J. Virol. 85:3930–39 [Google Scholar]
  12. Moorman NJ, Shenk T. 12.  2010. Rapamycin-resistant mTORC1 kinase activity is required for herpesvirus replication. J. Virol. 84:5260–69 [Google Scholar]
  13. Lenarcic EM, Ziehr B, De Leon G, Mitchell D, Moorman NJ. 13.  2014. Differential role for host translation factors in host and viral protein synthesis during human cytomegalovirus infection. J. Virol. 88:1473–83Late in infection, HCMV RNAs but not cellular mRNAs become resistant to eIF4F disruption. [Google Scholar]
  14. Hsieh AC, Liu Y, Edlind MP, Ingolia NT, Janes MR. 14.  et al. 2012. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485:55–61 [Google Scholar]
  15. Thoreen CC, Chantranupong L, Keys HR, Wang T, Gray NS, Sabatini DM. 15.  2012. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485:109–13 [Google Scholar]
  16. Lee AS, Kranzusch PJ, Cate JH. 16.  2015. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522:111–14
  17. Perez C, McKinney C, Chulunbaatar U, Mohr I. 17.  2011. Translational control of the abundance of cytoplasmic poly(A) binding protein in human cytomegalovirus-infected cells. J. Virol. 85:156–64 [Google Scholar]
  18. Walsh D, Perez C, Notary J, Mohr I. 18.  2005. Regulation of the translation initiation factor eIF4F by multiple mechanisms in human cytomegalovirus-infected cells. J. Virol. 79:8057–64 [Google Scholar]
  19. Aoyagi M, Gaspar M, Shenk TE. 19.  2010. Human cytomegalovirus UL69 protein facilitates translation by associating with the mRNA cap-binding complex and excluding 4EBP1. PNAS 107:2640–45 [Google Scholar]
  20. Kuang E, Tang Q, Maul GG, Zhu F. 20.  2008. Activation of p90 ribosomal S6 kinase by ORF45 of Kaposi's sarcoma-associated herpesvirus and its role in viral lytic replication. J. Virol. 82:1838–50 [Google Scholar]
  21. Kuang E, Wu F, Zhu F. 21.  2009. Mechanism of sustained activation of ribosomal S6 kinase (RSK) and ERK by Kaposi sarcoma-associated herpesvirus ORF45: Multiprotein complexes retain active phosphorylated ERK and RSK and protect them from dephosphorylation. J. Biol. Chem. 284:13958–68 [Google Scholar]
  22. Kuang E, Fu B, Liang Q, Myoung J, Zhu F. 22.  2011. Phosphorylation of eukaryotic translation initiation factor 4B (EIF4B) by open reading frame 45/p90 ribosomal S6 kinase (ORF45/RSK) signaling axis facilitates protein translation during Kaposi sarcoma-associated herpesvirus (KSHV) lytic replication. J. Biol. Chem. 286:41171–82 [Google Scholar]
  23. Parsyan A, Svitkin Y, Shahbazian D, Gkogkas C, Lasko P. 23.  et al. 2011. mRNA helicases: the tacticians of translational control. Nat. Rev. Mol. Cell Biol. 12:235–45 [Google Scholar]
  24. Fu B, Kuang E, Li W, Avey D, Li X. 24.  et al. 2014. Activation of p90 ribosomal S6 kinases (RSKs) by ORF45 of Kaposi sarcoma-associated herpesvirus is critical for optimal production of infectious viruses. J. Virol. 89:195–207 [Google Scholar]
  25. Jacquemont B, Roizman B. 25.  1975. RNA synthesis in cells infected with herpes simplex virus. X. Properties of viral symmetric transcripts and of double-stranded RNA prepared from them. J. Virol. 15:707–13 [Google Scholar]
  26. Langland JO, Jacobs BL. 26.  2002. The role of the PKR-inhibitory genes, E3L and K3L, in determining vaccinia virus host range. Virology 299:133–41 [Google Scholar]
  27. Taylor DR, Lee SB, Romano PR, Marshak DR, Hinnebusch AG. 27.  et al. 1996. Autophosphorylation sites participate in the activation of the double-stranded-RNA-activated protein kinase PKR. Mol. Cell. Biol. 16:6295–302 [Google Scholar]
  28. Silverman RH. 28.  2007. Viral encounters with 2′,5′-oligoadenylate synthetase and RNase L during the interferon antiviral response. J. Virol. 81:12720–29 [Google Scholar]
  29. Sciortino MT, Parisi T, Siracusano G, Mastino A, Taddeo B, Roizman B. 29.  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]
  30. He B, Gross M, Roizman B. 30.  1997. The γ134.5 protein of herpes simplex virus 1 complexes with protein phosphatase 1α to dephosphorylate the α subunit of the eukaryotic translation initiation factor 2 and preclude the shutoff of protein synthesis by double-stranded RNA-activated protein kinase. PNAS 94:843–48 [Google Scholar]
  31. Li Y, Zhang C, Chen X, Yu J, Wang Y. 31.  et al. 2011. ICP34.5 protein of herpes simplex virus facilitates the initiation of protein translation by bridging eukaryotic initiation factor 2α (eIF2α) and protein phosphatase 1. J. Biol. Chem. 286:24785–92 [Google Scholar]
  32. Van Opdenbosch N, Van den Broeke C, De Regge N, Tabares E, Favoreel HW. 32.  2012. The IE180 protein of pseudorabies virus suppresses phosphorylation of translation initiation factor eIF2α. J. Virol. 86:7235–40 [Google Scholar]
  33. Sanchez R, Mohr I. 33.  2007. Inhibition of cellular 2′-5′ oligoadenylate synthetase by the herpes simplex virus type 1 Us11 protein. J. Virol. 81:3455–64 [Google Scholar]
  34. Budt M, Niederstadt L, Valchanova RS, Jonjic S, Brune W. 34.  2009. Specific inhibition of the PKR-mediated antiviral response by the murine cytomegalovirus proteins m142 and m143. J. Virol. 83:1260–70 [Google Scholar]
  35. Child SJ, Geballe AP. 35.  2009. Binding and relocalization of protein kinase R by murine cytomegalovirus. J. Virol. 83:1790–99 [Google Scholar]
  36. Marshall EE, Bierle CJ, Brune W, Geballe AP. 36.  2009. Essential role for either TRS1 or IRS1 in human cytomegalovirus replication. J. Virol. 83:4112–20 [Google Scholar]
  37. Hakki M, Geballe AP. 37.  2005. Double-stranded RNA binding by human cytomegalovirus pTRS1. J. Virol. 79:7311–18 [Google Scholar]
  38. Hakki M, Marshall EE, De Niro KL, Geballe AP. 38.  2006. Binding and nuclear relocalization of protein kinase R by human cytomegalovirus TRS1. J. Virol. 80:11817–26 [Google Scholar]
  39. Child SJ, Hakki M, De Niro KL, Geballe AP. 39.  2004. Evasion of cellular antiviral responses by human cytomegalovirus TRS1 and IRS1. J. Virol. 78:197–205 [Google Scholar]
  40. Tan JC, Avdic S, Cao JZ, Mocarski ES, White KL. 40.  et al. 2011. Inhibition of 2′,5′-oligoadenylate synthetase expression and function by the human cytomegalovirus ORF94 gene product. J. Virol. 85:5696–700 [Google Scholar]
  41. Clarke PA, Schwemmle M, Schickinger J, Hilse K, Clemens MJ. 41.  1991. Binding of Epstein-Barr virus small RNA EBER-1 to the double-stranded RNA-activated protein kinase DAI. Nucleic Acids Res. 19:243–48 [Google Scholar]
  42. McKenna SA, Kim I, Liu CW, Puglisi JD. 42.  2006. Uncoupling of RNA binding and PKR kinase activation by viral inhibitor RNAs. J. Mol. Biol. 358:1270–85 [Google Scholar]
  43. Sharp TV, Schwemmle M, Jeffrey I, Laing K, Mellor H. 43.  et al. 1993. Comparative analysis of the regulation of the interferon-inducible protein kinase PKR by Epstein-Barr virus RNAs EBER-1 and EBER-2 and adenovirus VAI RNA. Nucleic Acids Res. 21:4483–90 [Google Scholar]
  44. Ruf IK, Lackey KA, Warudkar S, Sample JT. 44.  2005. Protection from interferon-induced apoptosis by Epstein-Barr virus small RNAs is not mediated by inhibition of PKR. J. Virol. 79:14562–69 [Google Scholar]
  45. Burysek L, Pitha PM. 45.  2001. Latently expressed human herpesvirus 8-encoded interferon regulatory factor 2 inhibits double-stranded RNA-activated protein kinase. J. Virol. 75:2345–52 [Google Scholar]
  46. Poppers J, Mulvey M, Perez C, Khoo D, Mohr I. 46.  2003. Identification of a lytic-cycle Epstein-Barr virus gene product that can regulate PKR activation. J. Virol. 77:228–36 [Google Scholar]
  47. Brewer JW. 47.  2014. Regulatory crosstalk within the mammalian unfolded protein response. Cell. Mol. Life Sci. 71:1067–79 [Google Scholar]
  48. Isler JA, Skalet AH, Alwine JC. 48.  2005. Human cytomegalovirus infection activates and regulates the unfolded protein response. J. Virol. 79:6890–99 [Google Scholar]
  49. Qian Z, Xuan B, Chapa TJ, Gualberto N, Yu D. 49.  2012. Murine cytomegalovirus targets transcription factor ATF4 to exploit the unfolded-protein response. J. Virol. 86:6712–23 [Google Scholar]
  50. Xuan B, Qian Z, Torigoi E, Yu D. 50.  2009. Human cytomegalovirus protein pUL38 induces ATF4 expression, inhibits persistent JNK phosphorylation, and suppresses endoplasmic reticulum stress-induced cell death. J. Virol. 83:3463–74 [Google Scholar]
  51. Stahl S, Burkhart JM, Hinte F, Tirosh B, Mohr H. 51.  et al. 2013. Cytomegalovirus downregulates IRE1 to repress the unfolded protein response. PLOS Pathog. 9:e1003544 [Google Scholar]
  52. Buchkovich NJ, Maguire TG, Yu Y, Paton AW, Paton JC, Alwine JC. 52.  2008. Human cytomegalovirus specifically controls the levels of the endoplasmic reticulum chaperone BiP/GRP78, which is required for virion assembly. J. Virol. 82:31–39 [Google Scholar]
  53. Buchkovich NJ, Yu Y, Pierciey FJ Jr, Alwine JC. 53.  2010. Human cytomegalovirus induces the endoplasmic reticulum chaperone BiP through increased transcription and activation of translation by using the BiP internal ribosome entry site. J. Virol. 84:11479–86 [Google Scholar]
  54. Lee DY, Sugden B. 54.  2008. The LMP1 oncogene of EBV activates PERK and the unfolded protein response to drive its own synthesis. Blood 111:2280–89 [Google Scholar]
  55. Bhende PM, Dickerson SJ, Sun X, Feng WH, Kenney SC. 55.  2007. X-box-binding protein 1 activates lytic Epstein-Barr virus gene expression in combination with protein kinase D. J. Virol. 81:7363–70 [Google Scholar]
  56. Mulvey M, Arias C, Mohr I. 56.  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]
  57. Carpenter JE, Jackson W, Benetti L, Grose C. 57.  2011. Autophagosome formation during varicella-zoster virus infection following endoplasmic reticulum stress and the unfolded protein response. J. Virol. 85:9414–24 [Google Scholar]
  58. Gaglia MM, Glaunsinger BA. 58.  2010. Viruses and the cellular RNA decay machinery. WIRES RNA 1:47–59 [Google Scholar]
  59. Dauber B, Pelletier J, Smiley JR. 59.  2011. The herpes simplex virus 1 vhs protein enhances translation of viral true late mRNAs and virus production in a cell type-dependent manner. J. Virol. 85:5363–73mRNA degradation by the vhs endonuclease liberates translation machinery for late gene expression. [Google Scholar]
  60. Esclatine A, Taddeo B, Roizman B. 60.  2004. Herpes simplex virus 1 induces cytoplasmic accumulation of TIA-1/TIAR and both synthesis and cytoplasmic accumulation of tristetraprolin, two cellular proteins that bind and destabilize AU-rich RNAs. J. Virol. 78:8582–92 [Google Scholar]
  61. Dauber B, Saffran HA, Smiley JR. 61.  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]
  62. Saffran HA, Read GS, Smiley JR. 62.  2010. Evidence for translational regulation by the herpes simplex virus virion host shutoff protein. J. Virol. 84:6041–49 [Google Scholar]
  63. Abernathy E, Clyde K, Yeasmin R, Krug LT, Burlingame A. 63.  et al. 2014. Gammaherpesviral gene expression and virion composition are broadly controlled by accelerated mRNA degradation. PLOS Pathog. 10:e1003882 [Google Scholar]
  64. Richner JM, Clyde K, Pezda AC, Cheng BY, Wang T. 64.  et al. 2011. Global mRNA degradation during lytic gammaherpesvirus infection contributes to establishment of viral latency. PLOS Pathog. 7:e1002150 [Google Scholar]
  65. McKinney C, Zavadil J, Bianco C, Shiflett L, Brown S, Mohr I. 65.  2014. Global reprogramming of the cellular translational landscape facilitates cytomegalovirus replication. Cell Rep. 6:9–17HCMV impacts the subset of host mRNAs recruited to or excluded from polysomes. [Google Scholar]
  66. Park EH, Zhang F, Warringer J, Sunnerhagen P, Hinnebusch AG. 66.  2011. Depletion of eIF4G from yeast cells narrows the range of translational efficiencies genome-wide. BMC Genomics 12:68 [Google Scholar]
  67. Ceci M, Gaviraghi C, Gorrini C, Sala LA, Offenhauser N. 67.  et al. 2003. Release of eIF6 (p27BBP) from the 60S subunit allows 80S ribosome assembly. Nature 426:579–84 [Google Scholar]
  68. Nott A, Le Hir H, Moore MJ. 68.  2004. Splicing enhances translation in mammalian cells: an additional function of the exon junction complex. Genes Dev. 18:210–22 [Google Scholar]
  69. Boyne JR, Jackson BR, Taylor A, Macnab SA, Whitehouse A. 69.  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]
  70. Fontaine-Rodriguez EC, Knipe DM. 70.  2008. Herpes simplex virus ICP27 increases translation of a subset of viral late mRNAs. J. Virol. 82:3538–45 [Google Scholar]
  71. Larralde O, Smith RW, Wilkie GS, Malik P, Gray NK, Clements JB. 71.  2006. Direct stimulation of translation by the multifunctional herpesvirus ICP27 protein. J. Virol. 80:1588–91 [Google Scholar]
  72. Ricci EP, Mure F, Gruffat H, Decimo D, Medina-Palazon C. 72.  et al. 2009. Translation of intronless RNAs is strongly stimulated by the Epstein-Barr virus mRNA export factor EB2. Nucleic Acids Res. 37:4932–43 [Google Scholar]
  73. Fontaine-Rodriguez EC, Taylor TJ, Olesky M, Knipe DM. 73.  2004. Proteomics of herpes simplex virus infected cell protein 27: association with translation initiation factors. Virology 330:487–92 [Google Scholar]
  74. Bono F, Ebert J, Unterholzner L, Guttler T, Izaurralde E, Conti E. 74.  2004. Molecular insights into the interaction of PYM with the Mago-Y14 core of the exon junction complex. EMBO Rep. 5:304–10 [Google Scholar]
  75. Diem MD, Chan CC, Younis I, Dreyfuss G. 75.  2007. PYM binds the cytoplasmic exon-junction complex and ribosomes to enhance translation of spliced mRNAs. Nat. Struct. Mol. Biol. 14:1173–79 [Google Scholar]
  76. Ellison KS, Maranchuk RA, Mottet KL, Smiley JR. 76.  2005. Control of VP16 translation by the herpes simplex virus type 1 immediate-early protein ICP27. J. Virol. 79:4120–31 [Google Scholar]
  77. Sokolowski M, Scott JE, Heaney RP, Patel AH, Clements JB. 77.  2003. Identification of herpes simplex virus RNAs that interact specifically with regulatory protein ICP27 in vivo. J. Biol. Chem. 278:33540–49 [Google Scholar]
  78. Verma D, Li DJ, Krueger B, Renne R, Swaminathan S. 78.  2015. Identification of the physiological gene targets of the essential lytic replicative KSHV ORF57 protein. J. Virol. 89:1688–702 [Google Scholar]
  79. Kang JG, Pripuzova N, Majerciak V, Kruhlak M, Le SY, Zheng ZM. 79.  2011. Kaposi's sarcoma-associated herpesvirus ORF57 promotes escape of viral and human interleukin-6 from microRNA-mediated suppression. J. Virol. 85:2620–30 [Google Scholar]
  80. Sei E, Wang T, Hunter OV, Xie Y, Conrad NK. 80.  2015. HITS-CLIP analysis uncovers a link between the Kaposi's sarcoma associated herpesvirus ORF57 protein and host pre-mRNA metabolism. PLOS Pathog. 11:e1004652 [Google Scholar]
  81. Bourdetsky D, Schmelzer CE, Admon A. 81.  2014. The nature and extent of contributions by defective ribosome products to the HLA peptidome. PNAS 111:E1591–99 [Google Scholar]
  82. Qian SB, Reits E, Neefjes J, Deslich JM, Bennink JR, Yewdell JW. 82.  2006. Tight linkage between translation and MHC class I peptide ligand generation implies specialized antigen processing for defective ribosomal products. J. Immunol. 177:227–33 [Google Scholar]
  83. Yewdell JW. 83.  2005. Serendipity strikes twice: the discovery and rediscovery of defective ribosomal products (DRiPS). Cell. Mol. Biol. 51:635–41 [Google Scholar]
  84. Tellam J, Fogg MH, Rist M, Connolly G, Tscharke D. 84.  et al. 2007. Influence of translation efficiency of homologous viral proteins on the endogenous presentation of CD8+ T cell epitopes. J. Exp. Med. 204:525–32 [Google Scholar]
  85. Apcher S, Daskalogianni C, Manoury B, Fahraeus R. 85.  2010. Epstein Barr virus-encoded EBNA1 interference with MHC class I antigen presentation reveals a close correlation between mRNA translation initiation and antigen presentation. PLOS Pathog. 6:e1001151 [Google Scholar]
  86. Blake N. 86.  2010. Immune evasion by gammaherpesvirus genome maintenance proteins. J. Gen. Virol. 91:829–46 [Google Scholar]
  87. Tellam JT, Lekieffre L, Zhong J, Lynn DJ, Khanna R. 87.  2012. Messenger RNA sequence rather than protein sequence determines the level of self-synthesis and antigen presentation of the EBV-encoded antigen, EBNA1. PLOS Pathog. 8:e1003112 [Google Scholar]
  88. Cardinaud S, Starck SR, Chandra P, Shastri N. 88.  2010. The synthesis of truncated polypeptides for immune surveillance and viral evasion. PLOS ONE 5:e8692 [Google Scholar]
  89. Murat P, Zhong J, Lekieffre L, Cowieson NP, Clancy JL. 89.  et al. 2014. G-quadruplexes regulate Epstein-Barr virus-encoded nuclear antigen 1 mRNA translation. Nat. Chem. Biol. 10:358–64G-quadruplexes in the EBNA1 mRNA slow its translation, thereby potentially reducing antigen presentation. [Google Scholar]
  90. Ressing ME, Horst D, Griffin BD, Tellam J, Zuo J. 90.  et al. 2008. Epstein-Barr virus evasion of CD8+ and CD4+ T cell immunity via concerted actions of multiple gene products. Semin. Cancer Biol. 18:397–408 [Google Scholar]
  91. Tellam JT, Zhong J, Lekieffre L, Bhat P, Martinez M. 91.  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]
  92. Kwun HJ, da Silva SR, Qin H, Ferris RL, Tan R. 92.  et al. 2011. The central repeat domain 1 of Kaposi's sarcoma-associated herpesvirus (KSHV) latency associated-nuclear antigen 1 (LANA1) prevents cis MHC class I peptide presentation. Virology 412:357–65 [Google Scholar]
  93. Kwun HJ, da Silva SR, Shah IM, Blake N, Moore PS, Chang Y. 93.  2007. Kaposi's sarcoma-associated herpesvirus latency-associated nuclear antigen 1 mimics Epstein-Barr virus EBNA1 immune evasion through central repeat domain effects on protein processing. J. Virol. 81:8225–35 [Google Scholar]
  94. Rubio CA, Weisburd B, Holderfield M, Arias C, Fang E. 94.  et al. 2014. Transcriptome-wide characterization of the eIF4A signature highlights plasticity in translation regulation. Genome Biol. 15:476 [Google Scholar]
  95. Wolfe AL, Singh K, Zhong Y, Drewe P, Rajasekhar VK. 95.  et al. 2014. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513:65–70 [Google Scholar]
  96. Kwun HJ, Toptan T, Ramos da Silva S, Atkins JF, Moore PS, Chang Y. 96.  2014. Human DNA tumor viruses generate alternative reading frame proteins through repeat sequence recoding. PNAS 111:E4342–49 [Google Scholar]
  97. Toptan T, Fonseca L, Kwun HJ, Chang Y, Moore PS. 97.  2013. Complex alternative cytoplasmic protein isoforms of the Kaposi's sarcoma-associated herpesvirus latency-associated nuclear antigen 1 generated through noncanonical translation initiation. J. Virol. 87:2744–55 [Google Scholar]
  98. Zaldumbide A, Ossevoort M, Wiertz EJ, Hoeben RC. 98.  2007. In cis inhibition of antigen processing by the latency-associated nuclear antigen I of Kaposi sarcoma herpes virus. Mol. Immunol. 44:1352–60 [Google Scholar]
  99. Dresang LR, Teuton JR, Feng H, Jacobs JM, Camp DG II. 99.  et al. 2011. Coupled transcriptome and proteome analysis of human lymphotropic tumor viruses: insights on the detection and discovery of viral genes. BMC Genomics 12:625 [Google Scholar]
  100. Chaudhary PM, Jasmin A, Eby MT, Hood L. 100.  1999. Modulation of the NF-κB pathway by virally encoded death effector domains-containing proteins. Oncogene 18:5738–46 [Google Scholar]
  101. Thome M, Schneider P, Hofmann K, Fickenscher H, Meinl E. 101.  et al. 1997. Viral FLICE-inhibitory proteins (FLIPs) prevent apoptosis induced by death receptors. Nature 386:517–21 [Google Scholar]
  102. Bieleski L, Talbot SJ. 102.  2001. Kaposi's sarcoma-associated herpesvirus vCyclin open reading frame contains an internal ribosome entry site. J. Virol. 75:1864–69 [Google Scholar]
  103. Grundhoff A, Ganem D. 103.  2001. Mechanisms governing expression of the v-FLIP gene of Kaposi's sarcoma-associated herpesvirus. J. Virol. 75:1857–63 [Google Scholar]
  104. Low W, Harries M, Ye H, Du MQ, Boshoff C, Collins M. 104.  2001. Internal ribosome entry site regulates translation of Kaposi's sarcoma-associated herpesvirus FLICE inhibitory protein. J. Virol. 75:2938–45 [Google Scholar]
  105. Othman Z, Sulaiman MK, Willcocks MM, Ulryck N, Blackbourn DJ. 105.  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]
  106. Feoktistova K, Tuvshintogs E, Do A, Fraser CS. 106.  2013. Human eIF4E promotes mRNA restructuring by stimulating eIF4A helicase activity. PNAS 110:13339–44 [Google Scholar]
  107. Arias C, Weisburd B, Stern-Ginossar N, Mercier A, Madrid AS. 107.  et al. 2014. KSHV 2.0: a comprehensive annotation of the Kaposi's sarcoma-associated herpesvirus genome using next-generation sequencing reveals novel genomic and functional features. PLOS Pathog. 10:e1003847 [Google Scholar]
  108. Stern-Ginossar N, Weisburd B, Michalski A, Le VT, Hein MY. 108.  et al. 2012. Decoding human cytomegalovirus. Science 338:1088–93Discovery of hundreds of new ORFs and protein expression control through alternative transcript start sites. [Google Scholar]
  109. Ingolia NT, Ghaemmaghami S, Newman JR, Weissman JS. 109.  2009. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324:218–23 [Google Scholar]
  110. Jaber T, Yuan Y. 110.  2013. A virally encoded small peptide regulates RTA stability and facilitates Kaposi's sarcoma-associated herpesvirus lytic replication. J. Virol. 87:3461–70 [Google Scholar]
  111. Cao J, Geballe AP. 111.  1995. Translational inhibition by a human cytomegalovirus upstream open reading frame despite inefficient utilization of its AUG codon. J. Virol. 69:1030–36 [Google Scholar]
  112. Kozak M. 112.  1987. Effects of intercistronic length on the efficiency of reinitiation by eucaryotic ribosomes. Mol. Cell. Biol. 7:3438–45 [Google Scholar]
  113. Kozak M. 113.  2001. Constraints on reinitiation of translation in mammals. Nucleic Acids Res. 29:5226–32 [Google Scholar]
  114. Gaba A, Wang Z, Krishnamoorthy T, Hinnebusch AG, Sachs MS. 114.  2001. Physical evidence for distinct mechanisms of translational control by upstream open reading frames. EMBO J. 20:6453–63 [Google Scholar]
  115. Hinnebusch AG. 115.  1993. Gene-specific translational control of the yeast GCN4 gene by phosphorylation of eukaryotic initiation factor 2. Mol. Microbiol. 10:215–23 [Google Scholar]
  116. Asano K, Krishnamoorthy T, Phan L, Pavitt GD, Hinnebusch AG. 116.  1999. Conserved bipartite motifs in yeast eIF5 and eIF2Bε, GTPase-activating and GDP-GTP exchange factors in translation initiation, mediate binding to their common substrate eIF2. EMBO J. 18:1673–88 [Google Scholar]
  117. Kronstad LM, Brulois KF, Jung JU, Glaunsinger BA. 117.  2013. Dual short upstream open reading frames control translation of a herpesviral polycistronic mRNA. PLOS Pathog. 9:e1003156uORFs enable polycistronic gene expression in KSHV. [Google Scholar]
  118. Kronstad LM, Brulois KF, Jung JU, Glaunsinger BA. 118.  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]
  119. Brubaker SW, Gauthier AE, Mills EW, Ingolia NT, Kagan JC. 119.  2014. A bicistronic MAVS transcript highlights a class of truncated variants in antiviral immunity. Cell 156:800–11 [Google Scholar]
  120. Calkhoven CF, Muller C, Leutz A. 120.  2000. Translational control of C/EBPα and C/EBPβ isoform expression. Genes Dev. 14:1920–32 [Google Scholar]
  121. Ingolia NT, Brar GA, Stern-Ginossar N, Harris MS, Talhouarne GJ. 121.  et al. 2014. Ribosome profiling reveals pervasive translation outside of annotated protein-coding genes. Cell Rep. 8:1365–79 [Google Scholar]
  122. Lee AS, Burdeinick-Kerr R, Whelan SP. 122.  2013. A ribosome-specialized translation initiation pathway is required for cap-dependent translation of vesicular stomatitis virus mRNAs. PNAS 110:324–29The ribosomal protein RPL40 directs transcript-specific translation of vesicular stomatitis virus and select host mRNAs. [Google Scholar]
/content/journals/10.1146/annurev-virology-100114-054839
Loading
/content/journals/10.1146/annurev-virology-100114-054839
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