Home Business Unconventional viral gene expression mechanisms as therapeutic targets

Unconventional viral gene expression mechanisms as therapeutic targets

0

[ad_1]

  • 1.

    Kozak, M. Point mutations define a sequence flanking the AUG initiator codon that modulates translation by eukaryotic ribosomes. Cell 44, 283–292 (1986).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 2.

    Chung, W. Y., Wadhawan, S., Szklarczyk, R., Pond, S. K. & Nekrutenko, A. A first look at ARFome: dual-coding genes in mammalian genomes. PLoS Comput. Biol. 3, e91 (2007).

    ADS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 3.

    Michel, A. M. et al. Observation of dually decoded regions of the human genome using ribosome profiling data. Genome Res. 22, 2219–2229 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 4.

    Schlub, T. E. & Holmes, E. C. Properties and abundance of overlapping genes in viruses. Virus Evol. 6, veaa009 (2020).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 5.

    Biswas, P., Jiang, X., Pacchia, A. L., Dougherty, J. P. & Peltz, S. W. The human immunodeficiency virus type 1 ribosomal frameshifting site is an invariant sequence determinant and an important target for antiviral therapy. J. Virol. 78, 2082–2087 (2004).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 6.

    Shehu-Xhilaga, M., Crowe, S. M. & Mak, J. Maintenance of the Gag/Gag-Pol ratio is important for human immunodeficiency virus type 1 RNA dimerization and viral infectivity. J. Virol. 75, 1834–1841 (2001).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 7.

    Jacks, T., Madhani, H. D., Masiarz, F. R. & Varmus, H. E. Signals for ribosomal frameshifting in the Rous sarcoma virus Gag-Pol region. Cell 55, 447–458 (1988). This article reports the discovery of the frameshift site and stem-loop structure as requirements for PRFs in the synthesis of Gag–Pol protein.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 8.

    van Knippenberg, I., Carlton-Smith, C. & Elliott, R. M. The N-terminus of Bunyamwera orthobunyavirus NSs protein is essential for interferon antagonism. J. Gen. Virol. 91, 2002–2006 (2010).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 9.

    McFadden, N. et al. Norovirus regulation of the innate immune response and apoptosis occurs via the product of the alternative open reading frame 4. PLoS Pathog. 7, e1002413 (2011).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 10.

    Scholthof, H. B. The Tombusvirus-encoded P19: from irrelevance to elegance. Nat. Rev. Microbiol. 4, 405–411 (2006).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 11.

    Chen, W. et al. A novel influenza A virus mitochondrial protein that induces cell death. Nat. Med. 7, 1306–1312 (2001).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 12.

    Rancurel, C., Khosravi, M., Dunker, A. K., Romero, P. R. & Karlin, D. Overlapping genes produce proteins with unusual sequence properties and offer insight into de novo protein creation. J. Virol. 83, 10719–10736 (2009). This paper was one of the first to suggest that overprinting is a mechanism for de novo gene and protein creation, and uses viruses as a model system to characterize the properties and the phylogenetic distributions of such proteins.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 13.

    Neme, R. & Tautz, D. Phylogenetic patterns of emergence of new genes support a model of frequent de novo evolution. BMC Genomics 14, 117 (2013).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 14.

    Penno, C., Kumari, R., Baranov, P. V., van Sinderen, D. & Atkins, J. F. Specific reverse transcriptase slippage at the HIV ribosomal frameshift sequence: potential implications for modulation of GagPol synthesis. Nucleic Acids Res. 45, 10156–10167 (2017).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 15.

    Hausmann, S., Garcin, D., Delenda, C. & Kolakofsky, D. The versatility of paramyxovirus RNA polymerase stuttering. J. Virol. 73, 5568–5576 (1999).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 16.

    Barr, J. N. & Wertz, G. W. Polymerase slippage at vesicular stomatitis virus gene junctions to generate poly(A) is regulated by the upstream 3′-AUAC-5′ tetranucleotide: implications for the mechanism of transcription termination. J. Virol. 75, 6901–6913 (2001).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 17.

    Thomas, S. M., Lamb, R. A. & Paterson, R. G. Two mRNAs that differ by two nontemplated nucleotides encode the amino coterminal proteins P and V of the paramyxovirus SV5. Cell 54, 891–902 (1988). A seminal paper that first identified transcriptional slippage as a mechanism of generating multiple transcripts from one gene in paramyxovirus SV5.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 18.

    Olspert, A., Chung, B. Y., Atkins, J. F., Carr, J. P. & Firth, A. E. Transcriptional slippage in the positive-sense RNA virus family Potyviridae. EMBO Rep. 16, 995–1004 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 19.

    Ratinier, M. et al. Transcriptional slippage prompts recoding in alternate reading frames in the hepatitis C virus (HCV) core sequence from strain HCV-1. J. Gen. Virol. 89, 1569–1578 (2008).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 20.

    Kolakofsky, D., Roux, L., Garcin, D. & Ruigrok, R. W. H. Paramyxovirus mRNA editing, the “rule of six” and error catastrophe: a hypothesis. J. Gen. Virol. 86, 1869–1877 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 21.

    Dillon, P. J. & Parks, G. D. Role for the phosphoprotein P subunit of the paramyxovirus polymerase in limiting induction of host cell antiviral responses. J. Virol. 81, 11116–11127 (2007).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 22.

    Gainey, M. D., Dillon, P. J., Clark, K. M., Manuse, M. J. & Parks, G. D. Paramyxovirus-induced shutoff of host and viral protein synthesis: role of the P and V proteins in limiting PKR activation. J. Virol. 82, 828–839 (2008).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 23.

    Didcock, L., Young, D. F., Goodbourn, S. & Randall, R. E. The V protein of simian virus 5 inhibits interferon signalling by targeting STAT1 for proteasome-mediated degradation. J. Virol. 73, 9928–9933 (1999).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 24.

    Kato, A., Kiyotani, K., Sakai, Y., Yoshida, T. & Nagai, Y. The paramyxovirus, Sendai virus, V protein encodes a luxury function required for viral pathogenesis. EMBO J. 16, 578–587 (1997).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 25.

    Vulliémoz, D. & Roux, L. “Rule of six”: how does the Sendai virus RNA polymerase keep count? J. Virol. 75, 4506–4518 (2001).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 26.

    Iseni, F. et al. Chemical modification of nucleotide bases and mRNA editing depend on hexamer or nucleoprotein phase in Sendai virus nucleocapsids. RNA 8, 1056–1067 (2002).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 27.

    Shabman, R. S. et al. Deep sequencing identifies noncanonical editing of Ebola and Marburg virus RNAs in infected cells. MBio 5, e02011-14 (2014).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 28.

    Mehedi, M. et al. A new Ebola virus nonstructural glycoprotein expressed through RNA editing. J. Virol. 85, 5406–5414 (2011).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 29.

    Sanchez, A., Trappier, S. G., Mahy, B. W., Peters, C. J. & Nichol, S. T. The virion glycoproteins of Ebola viruses are encoded in two reading frames and are expressed through transcriptional editing. Proc. Natl Acad. Sci. USA 93, 3602–3607 (1996).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 30.

    Sanchez, A. et al. Biochemical analysis of the secreted and virion glycoproteins of Ebola virus. J. Virol. 72, 6442–6447 (1998).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 31.

    Volchkov, V. E. et al. GP mRNA of Ebola virus is edited by the Ebola virus polymerase and by T7 and vaccinia virus polymerases. Virology 214, 421–430 (1995).

    CAS 

    Google Scholar
     

  • 32.

    Selman, M., Dankar, S. K., Forbes, N. E., Jia, J. J. & Brown, E. G. Adaptive mutation in influenza A virus non-structural gene is linked to host switching and induces a novel protein by alternative splicing. Emerg. Microbes Infect. 1, e42 (2012).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 33.

    Neumann, G., Hughes, M. T. & Kawaoka, Y. Influenza A virus NS2 protein mediates vRNP nuclear export through NES-independent interaction with hCRM1. EMBO J. 19, 6751–6758 (2000).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 34.

    Tsai, P. L. et al. Cellular RNA binding proteins NS1-BP and hnRNP K regulate influenza A virus RNA splicing. PLoS Pathog. 9, e1003460 (2013).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 35.

    Shih, S. R. & Krug, R. M. Novel exploitation of a nuclear function by influenza virus: the cellular SF2/ASF splicing factor controls the amount of the essential viral M2 ion channel protein in infected cells. EMBO J. 15, 5415–5427 (1996).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 36.

    Artarini, A. et al. Regulation of influenza A virus mRNA splicing by CLK1. Antiviral Res. 168, 187–196 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 37.

    Bogdanow, B. et al. The dynamic proteome of influenza A virus infection identifies M segment splicing as a host range determinant. Nat. Commun. 10, 5518 (2019).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 38.

    Jiang, T., Nogales, A., Baker, S. F., Martinez-Sobrido, L. & Turner, D. H. Mutations designed by ensemble defect to misfold conserved RNA structures of influenza A segments 7 and 8 affect splicing and attenuate viral replication in cell culture. PLoS ONE 11, e0156906 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 39.

    Cai, Z. et al. VirusCircBase: a database of virus circular RNAs. Brief Bioinform. 22, 2182–2190 (2021).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 40.

    Zhao, J. et al. Transforming activity of an oncoprotein-encoding circular RNA from human papillomavirus. Nat. Commun. 10, 2300 (2019). This paper reports how oncogenic strains of human papillomavirus use back-splicing to generate circular RNAs of oncogene E7, which have an essential role in the malignant transformation of infected cells.

    ADS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 41.

    Atkins, J. F., Loughran, G., Bhatt, P. R., Firth, A. E. & Baranov, P. V. Ribosomal frameshifting and transcriptional slippage: from genetic steganography and cryptography to adventitious use. Nucleic Acids Res. 44, 7007–7078 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 42.

    Loughran, G., Firth, A. E. & Atkins, J. F. Ribosomal frameshifting into an overlapping gene in the 2B-encoding region of the cardiovirus genome. Proc. Natl Acad. Sci. USA 108, E1111–E1119 (2011).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 43.

    Harger, J. W., Meskauskas, A. & Dinman, J. D. An “integrated model” of programmed ribosomal frameshifting. Trends Biochem. Sci. 27, 448–454 (2002).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 44.

    Dinman, J. D. Mechanisms and implications of programmed translational frameshifting. Wiley Interdiscip. Rev. RNA 3, 661–673 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 45.

    Jacks, T. et al. Characterization of ribosomal frameshifting in HIV-1 Gag-Pol expression. Nature 331, 280–283 (1988).

    ADS 
    CAS 

    Google Scholar
     

  • 46.

    Dulude, D., Baril, M. & Brakier-Gingras, L. Characterization of the frameshift stimulatory signal controlling a programmed −1 ribosomal frameshift in the human immunodeficiency virus type 1. Nucleic Acids Res. 30, 5094–5102 (2002).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 47.

    Cassan, M., Delaunay, N., Vaquero, C. & Rousset, J. P. Translational frameshifting at the Gag-Pol junction of human immunodeficiency virus type 1 is not increased in infected T-lymphoid cells. J. Virol. 68, 1501–1508 (1994).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 48.

    Brierley, I., Meredith, M. R., Bloys, A. J. & Hagervall, T. G. Expression of a coronavirus ribosomal frameshift signal in Escherichia coli: influence of tRNA anticodon modification on frameshifting. J. Mol. Biol. 270, 360–373 (1997).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 49.

    Firth, A. E., Blitvich, B. J., Wills, N. M., Miller, C. L. & Atkins, J. F. Evidence for ribosomal frameshifting and a novel overlapping gene in the genomes of insect-specific flaviviruses. Virology 399, 153–166 (2010).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 50.

    Baranov, P. V. et al. Programmed ribosomal frameshifting in decoding the SARS-CoV genome. Virology 332, 498–510 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 51.

    Plant, E. P., Rakauskaite, R., Taylor, D. R. & Dinman, J. D. Achieving a golden mean: mechanisms by which coronaviruses ensure synthesis of the correct stoichiometric ratios of viral proteins. J. Virol. 84, 4330–4340 (2010).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 52.

    Su, M. C., Chang, C. T., Chu, C. H., Tsai, C. H. & Chang, K. Y. An atypical RNA pseudoknot stimulator and an upstream attenuation signal for −1 ribosomal frameshifting of SARS coronavirus. Nucleic Acids Res. 33, 4265–4275 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 53.

    Dulude, D., Berchiche, Y. A., Gendron, K., Brakier-Gingras, L. & Heveker, N. Decreasing the frameshift efficiency translates into an equivalent reduction of the replication of the human immunodeficiency virus type 1. Virology 345, 127–136 (2006).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 54.

    Garcia-Miranda, P. et al. Stability of HIV frameshift site RNA correlates with frameshift efficiency and decreased virus infectivity. J. Virol. 90, 6906–6917 (2016).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 55.

    Karacostas, V., Wolffe, E. J., Nagashima, K., Gonda, M. A. & Moss, B. Overexpression of the HIV-1 Gag-Pol polyprotein results in intracellular activation of HIV-1 protease and inhibition of assembly and budding of virus-like particles. Virology 193, 661–671 (1993).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 56.

    Suzuki, Y. et al. Characterization of RyDEN (C19orf66) as an interferon-stimulated cellular inhibitor against dengue virus replication. PLoS Pathog. 12, e1005357 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 57.

    Wang, X. et al. Regulation of HIV-1 Gag-Pol expression by shiftless, an inhibitor of programmed −1 ribosomal frameshifting. Cell 176, 625–635.e14 (2019). This article describes how the host factor shiftless functions as a universal inhibitor of −1 PRFs by causing premature translational termination in HIV-1 and several other viruses.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 58.

    Schwartz, S., Felber, B. K., Fenyö, E. M. & Pavlakis, G. N. Env and Vpu proteins of human immunodeficiency virus type 1 are produced from multiple bicistronic mRNAs. J. Virol. 64, 5448–5456 (1990). The authors report the discovery of leaky scanning of an upstream viral protein as the mechanism of generating the envelope protein in HIV-1.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 59.

    Kolakofsky, D., Le Mercier, P., Iseni, F. & Garcin, D. Viral DNA polymerase scanning and the gymnastics of Sendai virus RNA synthesis. Virology 318, 463–473 (2004).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 60.

    Stacey, S. N. et al. Leaky scanning is the predominant mechanism for translation of human papillomavirus type 16 E7 oncoprotein from E6/E7 bicistronic mRNA. J. Virol. 74, 7284–7297 (2000).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 61.

    Fuller, F., Bhown, A. S. & Bishop, D. H. Bunyavirus nucleoprotein, N, and a non-structural protein, NSS, are coded by overlapping reading frames in the S RNA. J. Gen. Virol. 64, 1705–1714 (1983).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 62.

    Vera-Otarola, J. et al. The Andes hantavirus NSs protein is expressed from the viral small mRNA by a leaky scanning mechanism. J. Virol. 86, 2176–2187 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 63.

    Strebel, K., Klimkait, T. & Martin, M. A. A novel gene of HIV-1, vpu, and its 16-kilodalton product. Science 241, 1221–1223 (1988).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 64.

    Willey, R. L., Maldarelli, F., Martin, M. A. & Strebel, K. Human immunodeficiency virus type 1 Vpu protein induces rapid degradation of CD4. J. Virol. 66, 7193–7200 (1992).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 65.

    Terwilliger, E. F., Cohen, E. A., Lu, Y. C., Sodroski, J. G. & Haseltine, W. A. Functional role of human immunodeficiency virus type 1 Vpu. Proc. Natl Acad. Sci. USA 86, 5163–5167 (1989).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 66.

    Sauter, D. et al. Tetherin-driven adaptation of Vpu and Nef function and the evolution of pandemic and nonpandemic HIV-1 strains. Cell Host Microbe 6, 409–421 (2009).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 67.

    Wise, H. M. et al. Overlapping signals for translational regulation and packaging of influenza A virus segment 2. Nucleic Acids Res. 39, 7775–7790 (2011).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 68.

    Zell, R. et al. Prevalence of PB1-F2 of influenza A viruses. J. Gen. Virol. 88, 536–546 (2007).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 69.

    McAuley, J. L. et al. PB1-F2 proteins from H5N1 and 20th century pandemic influenza viruses cause immunopathology. PLoS Pathog. 6, e1001014 (2010).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 70.

    Zamarin, D., García-Sastre, A., Xiao, X., Wang, R. & Palese, P. Influenza virus PB1-F2 protein induces cell death through mitochondrial ANT3 and VDAC1. PLoS Pathog. 1, e4 (2005).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 71.

    Lulla, V. et al. An upstream protein-coding region in enteroviruses modulates virus infection in gut epithelial cells. Nat. Microbiol. 4, 280–292 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 72.

    Irigoyen, N. et al. High-resolution analysis of coronavirus gene expression by RNA sequencing and ribosome profiling. PLoS Pathog. 12, e1005473 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 73.

    Wu, H. Y., Guan, B. J., Su, Y. P., Fan, Y. H. & Brian, D. A. Reselection of a genomic upstream open reading frame in mouse hepatitis coronavirus 5′-untranslated-region mutants. J. Virol. 88, 846–858 (2014).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 74.

    Hofmann, M. A., Senanayake, S. D. & Brian, D. A. A translation-attenuating intraleader open reading frame is selected on coronavirus mRNAs during persistent infection. Proc. Natl Acad. Sci. USA 90, 11733–11737 (1993).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 75.

    Shabman, R. S. et al. An upstream open reading frame modulates ebola virus polymerase translation and virus replication. PLoS Pathog. 9, e1003147 (2013). This paper relates how an upstream ORF suppresses the translation of the downstream canonical L protein as a way to maintain protein expression level in ebolavirus.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 76.

    Degnin, C. R., Schleiss, M. R., Cao, J. & Geballe, A. P. Translational inhibition mediated by a short upstream open reading frame in the human cytomegalovirus gpUL4 (gp48) transcript. J. Virol. 67, 5514–5521 (1993).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 77.

    Chen, A., Kao, Y. F. & Brown, C. M. Translation of the first upstream ORF in the hepatitis B virus pregenomic RNA modulates translation at the core and polymerase initiation codons. Nucleic Acids Res. 33, 1169–1181 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 78.

    Calvo, S. E., Pagliarini, D. J. & Mootha, V. K. Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. Proc. Natl Acad. Sci. USA 106, 7507–7512 (2009).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 79.

    Ji, Z., Song, R., Regev, A. & Struhl, K. Many lncRNAs, 5’UTRs, and pseudogenes are translated and some are likely to express functional proteins. eLife 4, e08890 (2015).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 80.

    Jin, X., Turcott, E., Englehardt, S., Mize, G. J. & Morris, D. R. The two upstream open reading frames of oncogene mdm2 have different translational regulatory properties. J. Biol. Chem. 278, 25716–25721 (2003).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 81.

    Firth, A. E. & Brierley, I. Non-canonical translation in RNA viruses. J. Gen. Virol. 93, 1385–1409 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 82.

    Young, S. K. & Wek, R. C. Upstream open reading frames differentially regulate gene-specific translation in the integrated stress response. J. Biol. Chem. 291, 16927–16935 (2016).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 83.

    Gupta, K. C. & Patwardhan, S. ACG, the initiator codon for a Sendai virus protein. J. Biol. Chem. 263, 8553–8556 (1988). This paper represents one of the first observations of near-cognate start codon use in viral protein expression.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 84.

    Boeck, R., Curran, J., Matsuoka, Y., Compans, R. & Kolakofsky, D. The parainfluenza virus type 1 P/C gene uses a very efficient GUG codon to start its C′ protein. J. Virol. 66, 1765–1768 (1992).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 85.

    Prats, A. C., De Billy, G., Wang, P. & Darlix, J. L. CUG initiation codon used for the synthesis of a cell surface antigen coded by the murine leukemia virus. J. Mol. Biol. 205, 363–372 (1989).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 86.

    Corcelette, S., Massé, T. & Madjar, J. J. Initiation of translation by non-AUG codons in human T-cell lymphotropic virus type I mRNA encoding both Rex and Tax regulatory proteins. Nucleic Acids Res. 28, 1625–1634 (2000).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 87.

    Machkovech, H. M., Bloom, J. D. & Subramaniam, A. R. Comprehensive profiling of translation initiation in influenza virus infected cells. PLoS Pathog. 15, e1007518 (2019).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 88.

    Shirako, Y. Non-AUG translation initiation in a plant RNA virus: a forty-amino-acid extension is added to the N terminus of the soil-borne wheat mosaic virus capsid protein. J. Virol. 72, 1677–1682 (1998).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 89.

    Carroll, R. & Derse, D. Translation of equine infectious anemia virus bicistronic tat-rev mRNA requires leaky ribosome scanning of the tat CTG initiation codon. J. Virol. 67, 1433–1440 (1993).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 90.

    Sasaki, J. & Nakashima, N. Methionine-independent initiation of translation in the capsid protein of an insect RNA virus. Proc. Natl Acad. Sci. USA 97, 1512–1515 (2000). This paper reports the discovery that translation initiation of an insect virus mRNA is independent of both AUG and the methionine initiator tRNA.

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 91.

    Wilson, J. E., Powell, M. J., Hoover, S. E. & Sarnow, P. Naturally occurring dicistronic cricket paralysis virus RNA is regulated by two internal ribosome entry sites. Mol. Cell. Biol. 20, 4990–4999 (2000).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 92.

    Rialdi, A. et al. The RNA exosome syncs IAV-RNAPII transcription to promote viral ribogenesis and infectivity. Cell 169, 679–692.e14 (2017).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 93.

    Dias, A. et al. The cap-snatching endonuclease of influenza virus polymerase resides in the PA subunit. Nature 458, 914–918 (2009).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 94.

    Reich, S. et al. Structural insight into cap-snatching and RNA synthesis by influenza polymerase. Nature 516, 361–366 (2014).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 95.

    Koppstein, D., Ashour, J. & Bartel, D. P. Sequencing the cap-snatching repertoire of H1N1 influenza provides insight into the mechanism of viral transcription initiation. Nucleic Acids Res. 43, 5052–5064 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 96.

    Gu, W. et al. Influenza A virus preferentially snatches noncoding RNA caps. RNA 21, 2067–2075 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 97.

    Sikora, D., Rocheleau, L., Brown, E. G. & Pelchat, M. Influenza A virus cap-snatches host RNAs based on their abundance early after infection. Virology 509, 167–177 (2017).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 98.

    Ho, J. S. Y. et al. Hybrid gene origination creates human-virus chimeric proteins during infection. Cell 181, 1502–1517 (2020). This article describes the discovery of upstream AUGs in cap-snatched host sequences being transcribed and translated to generate host–virus hybrid proteins.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 99.

    Wise, H. M. et al. An alternative AUG codon that produces an N-terminally extended form of the influenza A virus NP is a virulence factor for a swine-derived virus. Preprint at https://doi.org/10.1101/738427 (2019).

  • 100.

    Kim, H., Webster, R. G. & Webby, R. J. Influenza virus: dealing with a drifting and shifting pathogen. Viral Immunol. 31, 174–183 (2018).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 101.

    Firth, A. E. & Brown, C. M. Detecting overlapping coding sequences in virus genomes. BMC Bioinformatics 7, 75 (2006).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 102.

    Firth, A. E. & Brown, C. M. Detecting overlapping coding sequences with pairwise alignments. Bioinformatics 21, 282–292 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 103.

    Jagger, B. W. et al. An overlapping protein-coding region in influenza A virus segment 3 modulates the host response. Science 337, 199–204 (2012).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 104.

    Mohamadi, M. et al. Hepatitis C virus alternative reading frame protein (ARFP): production, features, and pathogenesis. J. Med. Virol. 92, 2930–2937 (2020).

  • 105.

    Zanker, D. J. et al. Influenza A virus infection induces viral and cellular defective ribosomal products encoded by alternative reading frames. J. Immunol. 202, 3370–3380 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 106.

    Depledge, D. P. et al. Direct RNA sequencing on nanopore arrays redefines the transcriptional complexity of a viral pathogen. Nat. Commun. 10, 754 (2019).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 107.

    Di, H. et al. Expanded subgenomic mRNA transcriptome and coding capacity of a nidovirus. Proc. Natl Acad. Sci. USA 114, E8895–E8904 (2017).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 108.

    Bansal, A. et al. CD8 T cell response and evolutionary pressure to HIV-1 cryptic epitopes derived from antisense transcription. J. Exp. Med. 207, 51–59 (2010). This paper describes how an ARF in HIV encodes cryptic epitopes that contribute to the majority of the CD8 T cell response during infection.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 109.

    Vetsika, E. K. et al. Sequential administration of the native TERT572 cryptic peptide enhances the immune response initiated by its optimized variant TERT572Y in cancer patients. J. Immunother. 34, 641–650 (2011).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 110.

    Dolan, B. P., Li, L., Takeda, K., Bennink, J. R. & Yewdell, J. W. Defective ribosomal products are the major source of antigenic peptides endogenously generated from influenza A virus neuraminidase. J. Immunol. 184, 1419–1424 (2010).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 111.

    Bansal, A. et al. Enhanced recognition of HIV-1 cryptic epitopes restricted by HLA class I alleles associated with a favorable clinical outcome. J. Acquir. Immune Defic. Syndr. 70, 1–8 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 112.

    Bet, A., Sterrett, S., Sato, A., Bansal, A. & Goepfert, P. A. Characterization of T-cell responses to cryptic epitopes in recipients of a noncodon-optimized HIV-1 vaccine. J. Acquir. Immune Defic. Syndr. 65, 142–150 (2014).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 113.

    Peng, B. J. et al. Antisense-derived HIV-1 cryptic epitopes are not major drivers of viral evolution during the acute phase of infection. J. Virol. 92, e00711-18 (2018).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 114.

    Maness, N. J. et al. Robust, vaccine-induced CD8+ T lymphocyte response against an out-of-frame epitope. J. Immunol. 184, 67–72 (2010). Four articles111–114 demonstrate how the T cell response to cryptic epitopes produced from ARFs in HIV is robust, associated with improved clinical outcomes and associated with minimal viral escape.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 115.

    Seif, M., Einsele, H. & Löffler, J. CAR T cells beyond cancer: hope for immunomodulatory therapy of infectious diseases. Front. Immunol. 10, 2711 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 116.

    Park, S. J., Kim, Y. G. & Park, H. J. Identification of RNA pseudoknot-binding ligand that inhibits the -1 ribosomal frameshifting of SARS-coronavirus by structure-based virtual screening. J. Am. Chem. Soc. 133, 10094–10100 (2011). This paper describes how an RNA structure-based inhibitor screening revealed a candidate that targets and reduces the efficiency of −1 PRFs in SARS-CoV.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 117.

    Ritchie, D. B., Soong, J., Sikkema, W. K. & Woodside, M. T. Anti-frameshifting ligand reduces the conformational plasticity of the SARS virus pseudoknot. J. Am. Chem. Soc. 136, 2196–2199 (2014).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 118.

    Ritchie, D. B., Foster, D. A. & Woodside, M. T. Programmed −1 frameshifting efficiency correlates with RNA pseudoknot conformational plasticity, not resistance to mechanical unfolding. Proc. Natl Acad. Sci. USA 109, 16167–16172 (2012).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 119.

    Seth, P. P. et al. SAR by MS: discovery of a new class of RNA-binding small molecules for the hepatitis C virus: internal ribosome entry site IIA subdomain. J. Med. Chem. 48, 7099–7102 (2005).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 120.

    Parsons, J. et al. Conformational inhibition of the hepatitis C virus internal ribosome entry site RNA. Nat. Chem. Biol. 5, 823–825 (2009).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 121.

    Dibrov, S. M. et al. Structure of a hepatitis C virus RNA domain in complex with a translation inhibitor reveals a binding mode reminiscent of riboswitches. Proc. Natl Acad. Sci. USA 109, 5223–5228 (2012).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 122.

    Yamamoto, H. et al. Molecular architecture of the ribosome-bound hepatitis C virus internal ribosomal entry site RNA. EMBO J. 34, 3042–3058 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 123.

    Juliano, R., Alam, M. R., Dixit, V. & Kang, H. Mechanisms and strategies for effective delivery of antisense and siRNA oligonucleotides. Nucleic Acids Res. 36, 4158–4171 (2008).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 124.

    Soler, M., McHutchison, J. G., Kwoh, T. J., Dorr, F. A. & Pawlotsky, J. M. Virological effects of ISIS 14803, an antisense oligonucleotide inhibitor of hepatitis C virus (HCV) internal ribosome entry site (IRES), on HCV IRES in chronic hepatitis C patients and examination of the potential role of primary and secondary HCV resistance in the outcome of treatment. Antivir. Ther. 9, 953–968 (2004).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 125.

    Zhang, H. et al. Antisense oligonucleotide inhibition of hepatitis C virus (HCV) gene expression in livers of mice infected with an HCV–vaccinia virus recombinant. Antimicrob. Agents Chemother. 43, 347–353 (1999).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 126.

    Hayden, F. G. & Shindo, N. Influenza virus polymerase inhibitors in clinical development. Curr. Opin. Infect. Dis. 32, 176–186 (2019).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 127.

    Noshi, T. et al. In vitro characterization of baloxavir acid, a first-in-class cap-dependent endonuclease inhibitor of the influenza virus polymerase PA subunit. Antiviral Res. 160, 109–117 (2018).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 128.

    Fukao, K. et al. Combination treatment with the cap-dependent endonuclease inhibitor baloxavir marboxil and a neuraminidase inhibitor in a mouse model of influenza A virus infection. J. Antimicrob. Chemother. 74, 654–662 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 129.

    Mishin, V. P. et al. Susceptibility of influenza A, B, C, and D viruses to baloxavir. Emerg. Infect. Dis. 25, 1969–1972 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 130.

    de Farias, S. T., Dos Santos Junior, A. P., Rêgo, T. G. & José, M. V. Origin and evolution of RNA-dependent RNA polymerase. Front. Genet. 8, 125 (2017).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 131.

    Guedj, J. et al. Antiviral efficacy of favipiravir against ebola virus: a translational study in cynomolgus macaques. PLoS Med. 15, e1002535 (2018).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 132.

    Omoto, S. et al. Characterization of influenza virus variants induced by treatment with the endonuclease inhibitor baloxavir marboxil. Sci. Rep. 8, 9633 (2018).

    ADS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 133.

    Hofmann, W. P., Soriano, V. & Zeuzem, S. in Antiviral Stategies: Handbook of Experimental Pharmacology (eds Kräusslich H. G. & Bartenschlager, R.) 321–346 (2009).

  • 134.

    Arts, E. J. & Hazuda, D. J. HIV-1 antiretroviral drug therapy. Cold Spring Harb. Perspect. Med. 2, a007161 (2012).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 135.

    Bayat Mokhtari, R. et al. Combination therapy in combating cancer. Oncotarget 8, 38022–38043 (2017).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 136.

    Worthington, R. J. & Melander, C. Combination approaches to combat multidrug-resistant bacteria. Trends Biotechnol. 31, 177–184 (2013).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 137.

    Simmonds, P., Aiewsakun, P. & Katzourakis, A. Prisoners of war – host adaptation and its constraints on virus evolution. Nat. Rev. Microbiol. 17, 321–328 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 138.

    Krause-Kyora, B. et al. Neolithic and medieval virus genomes reveal complex evolution of hepatitis B. eLife 7, e36666 (2018).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 139.

    Mühlemann, B. et al. Ancient hepatitis B viruses from the Bronze Age to the Medieval period. Nature 557, 418–423 (2018). Two articles138,139 show that hepatitis B virus isolated from samples from the Bronze and Neolithic Age is minimally divergent from modern strains, which suggests that viral adaptation to the host may confer an upper limit to the mutations that can be accumulated over time.

    ADS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 140.

    Majzoub, K. et al. RACK1 controls IRES-mediated translation of viruses. Cell 159, 1086–1095 (2014).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 141.

    Ullah, H., Hou, W., Dakshanamurthy, S. & Tang, Q. Host targeted antiviral (HTA): functional inhibitor compounds of scaffold protein RACK1 inhibit herpes simplex virus proliferation. Oncotarget 10, 3209–3226 (2019).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 142.

    Heaton, N. S. et al. Targeting viral proteostasis limits influenza virus, HIV, and dengue virus infection. Immunity 44, 46–58 (2016). This paper reports that viruses exhibit host dependency of SEC61-mediated protein translocation and folding, and that inhibition of SEC61 provides a broad-spectrum inhibition on growth and infectivity in several viruses.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 143.

    Shah, P. S. et al. Comparative flavivirus–host protein interaction mapping reveals mechanisms of dengue and Zika virus pathogenesis. Cell 175, 1931–1945 (2018).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 144.

    Gordon, D. E. et al. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature 583, 459–468 (2020).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 145.

    Dalziel, M., Crispin, M., Scanlan, C. N., Zitzmann, N. & Dwek, R. A. Emerging principles for the therapeutic exploitation of glycosylation. Science 343, 1235681 (2014).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 146.

    Watanabe, Y., Bowden, T. A., Wilson, I. A. & Crispin, M. Exploitation of glycosylation in enveloped virus pathobiology. Biochim. Biophys. Acta, Gen. Subj. 1863, 1480–1497 (2019).

    CAS 

    Google Scholar
     

  • 147.

    Bojkova, D. et al. Proteomics of SARS-CoV-2-infected host cells reveals therapy targets. Nature 583, 469–472 (2020).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 148.

    Dhillon, P. et al. Cytoplasmic relocalization and colocalization with viroplasms of host cell proteins, and their role in rotavirus infection. J. Virol. 92, e00612-18 (2018).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 149.

    Zhao, N. et al. Influenza virus infection causes global RNAPII termination defects. Nat. Struct. Mol. Biol. 25, 885–893 (2018).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 150.

    Boudreault, S., Roy, P., Lemay, G. & Bisaillon, M. Viral modulation of cellular RNA alternative splicing: a new key player in virus-host interactions? Wiley Interdiscip. Rev. RNA 10, e1543 (2019).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 151.

    Bonenfant, G. et al. Zika virus subverts stress granules to promote and restrict viral gene expression. J. Virol. 93, e00520-19 (2019).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 152.

    Batra, R. et al. RNA-binding protein CPEB1 remodels host and viral RNA landscapes. Nat. Struct. Mol. Biol. 23, 1101–1110 (2016).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 153.

    Tremblay, M. P. et al. Global profiling of alternative RNA splicing events provides insights into molecular differences between various types of hepatocellular carcinoma. BMC Genomics 17, 683 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 154.

    Lin, J. C. Therapeutic applications of targeted alternative splicing to cancer treatment. Int. J. Mol. Sci. 19, E75 (2017).


    Google Scholar
     

  • 155.

    Urbanski, L. M., Leclair, N. & Anczuków, O. Alternative-splicing defects in cancer: splicing regulators and their downstream targets, guiding the way to novel cancer therapeutics. Wiley Interdiscip. Rev. RNA 9, e1476 (2018).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 156.

    Pelletier, J. & Sonenberg, N. Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 334, 320–325 (1988).

    ADS 
    CAS 

    Google Scholar
     

  • 157.

    Jang, S. K. et al. A segment of the 5′ nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J. Virol. 62, 2636–2643 (1988). This paper represents one of the first observations of IRESs as an alternative cap-independent translation mechanism that is used by encephalomyocarditis virus.

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 158.

    Griffiths, A. & Coen, D. M. An unusual internal ribosome entry site in the herpes simplex virus thymidine kinase gene. Proc. Natl Acad. Sci. USA 102, 9667–9672 (2005).

    ADS 
    CAS 

    Google Scholar
     

  • 159.

    Kang, S. T., Wang, H. C., Yang, Y. T., Kou, G. H. & Lo, C. F. The DNA virus white spot syndrome virus uses an internal ribosome entry site for translation of the highly expressed nonstructural protein ICP35. J. Virol. 87, 13263–13278 (2013).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 160.

    Zhao, J. et al. IRESbase: a comprehensive database of experimentally validated internal ribosome entry sites. Genomics Proteomics Bioinformatics 18, 129–139 (2020).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 161.

    Sweeney, T. R., Abaeva, I. S., Pestova, T. V. & Hellen, C. U. The mechanism of translation initiation on type 1 picornavirus IRESs. EMBO J. 33, 76–92 (2014).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 162.

    Kolupaeva, V. G., Lomakin, I. B., Pestova, T. V. & Hellen, C. U. Eukaryotic initiation factors 4G and 4A mediate conformational changes downstream of the initiation codon of the encephalomyocarditis virus internal ribosomal entry site. Mol. Cell. Biol. 23, 687–698 (2003).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 163.

    Fernández, I. S., Bai, X. C., Murshudov, G., Scheres, S. H. & Ramakrishnan, V. Initiation of translation by cricket paralysis virus IRES requires its translocation in the ribosome. Cell 157, 823–831 (2014).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 164.

    Deniz, N., Lenarcic, E. M., Landry, D. M. & Thompson, S. R. Translation initiation factors are not required for Dicistroviridae IRES function in vivo. RNA 15, 932–946 (2009).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 165.

    Fütterer, J., Kiss-László, Z. & Hohn, T. Nonlinear ribosome migration on cauliflower mosaic virus 35S RNA. Cell 73, 789–802 (1993).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 166.

    Pooggin, M. M., Fütterer, J., Skryabin, K. G. & Hohn, T. A short open reading frame terminating in front of a stable hairpin is the conserved feature in pregenomic RNA leaders of plant pararetroviruses. J. Gen. Virol. 80, 2217–2228 (1999).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 167.

    Yueh, A. & Schneider, R. J. Selective translation initiation by ribosome jumping in adenovirus-infected and heat-shocked cells. Genes Dev. 10, 1557–1567 (1996).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 168.

    Latorre, P., Kolakofsky, D. & Curran, J. Sendai virus Y proteins are initiated by a ribosomal shunt. Mol. Cell. Biol. 18, 5021–5031 (1998).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 169.

    Meyers, G. Translation of the minor capsid protein of a calicivirus is initiated by a novel termination-dependent reinitiation mechanism. J. Biol. Chem. 278, 34051–34060 (2003).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 170.

    Horvath, C. M., Williams, M. A. & Lamb, R. A. Eukaryotic coupled translation of tandem cistrons: identification of the influenza B virus BM2 polypeptide. EMBO J. 9, 2639–2647 (1990).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 171.

    Ahmadian, G., Randhawa, J. S. & Easton, A. J. Expression of the ORF-2 protein of the human respiratory syncytial virus M2 gene is initiated by a ribosomal termination-dependent reinitiation mechanism. EMBO J. 19, 2681–2689 (2000).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 172.

    Gould, P. S. & Easton, A. J. Coupled translation of the respiratory syncytial virus M2 open reading frames requires upstream sequences. J. Biol. Chem. 280, 21972–21980 (2005).

    CAS 

    Google Scholar
     

  • 173.

    Jeudy, S., Abergel, C., Claverie, J. M. & Legendre, M. Translation in giant viruses: a unique mixture of bacterial and eukaryotic termination schemes. PLoS Genet. 8, e1003122 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 174.

    Schueren, F. & Thoms, S. Functional translational readthrough: a systems biology perspective. PLoS Genet. 12, e1006196 (2016).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 175.

    Leinfelder, W., Zehelein, E., Mandrand-Berthelot, M. A. & Böck, A. Gene for a novel tRNA species that accepts l-serine and cotranslationally inserts selenocysteine. Nature 331, 723–725 (1988).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 176.

    Lee, B. J., Worland, P. J., Davis, J. N., Stadtman, T. C. & Hatfield, D. L. Identification of a selenocysteyl-tRNA(Ser) in mammalian cells that recognizes the nonsense codon, UGA. J. Biol. Chem. 264, 9724–9727 (1989).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 177.

    Ryan, M. D. & Drew, J. Foot-and-mouth disease virus 2A oligopeptide mediated cleavage of an artificial polyprotein. EMBO J. 13, 928–933 (1994).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 178.

    Donnelly, M. L. L. et al. Analysis of the aphthovirus 2A/2B polyprotein ‘cleavage’ mechanism indicates not a proteolytic reaction, but a novel translational effect: a putative ribosomal ‘skip’. J. Gen. Virol. 82, 1013–1025 (2001).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 179.

    Luke, G. A. et al. Occurrence, function and evolutionary origins of ‘2A-like’ sequences in virus genomes. J. Gen. Virol. 89, 1036–1042 (2008).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 180.

    Sharma, P. et al. 2A peptides provide distinct solutions to driving stop-carry on translational recoding. Nucleic Acids Res. 40, 3143–3151 (2012).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 181.

    Ogino, T. & Banerjee, A. K. Unconventional mechanism of mRNA capping by the RNA-dependent RNA polymerase of vesicular stomatitis virus. Mol. Cell 25, 85–97 (2007).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 182.

    Decroly, E., Ferron, F., Lescar, J. & Canard, B. Conventional and unconventional mechanisms for capping viral mRNA. Nat. Rev. Microbiol. 10, 51–65 (2011).

    PubMed 
    PubMed Central 

    Google Scholar
     

  • 183.

    Ahola, T. & Kääriäinen, L. Reaction in alphavirus mRNA capping: formation of a covalent complex of nonstructural protein nsP1 with 7-methyl-GMP. Proc. Natl Acad. Sci. USA 92, 507–511 (1995).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 184.

    Goodfellow, I. The genome-linked protein VPg of vertebrate viruses – a multifaceted protein. Curr. Opin. Virol. 1, 355–362 (2011).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 185.

    Sola, I., Almazán, F., Zúñiga, S. & Enjuanes, L. Continuous and discontinuous RNA synthesis in coronaviruses. Annu. Rev. Virol. 2, 265–288 (2015).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 186.

    Kim, D. et al. The architecture of SARS-CoV-2 transcriptome. Cell 181, 914–921.e10 (2020).

    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • 187.

    Finkel, Y. et al. The coding capacity of SARS-CoV-2. Nature 589, 125–130 (2021).

    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar
     

  • [ad_2]

    Source link