Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Functional 5′ UTR mRNA structures in eukaryotic translation regulation and how to find them

An Author Correction to this article was published on 15 August 2018

Key Points

  • mRNA 5′ untranslated regions (UTRs), which serve as the entry point for the ribosome during translation, can adopt elaborate RNA secondary and tertiary structures that may regulate translation initiation in a cap-dependent or cap-independent manner.

  • Complex RNA structures in 5′ UTRs, such as RNA G-quadruplexes, may serve as steric blocks to RNA structure unwinding by the scanning ribosome, eukaryotic initiation factor 4A (eIF4A) and other helicases.

  • Internal ribosome entry sites (IRESs) recruit ribosomes to 5′ UTRs in a cap-independent manner. Many cellular IRESs are activated upon stress, whereas some are required in regular physiological conditions. Cellular IRESs tend to integrate or interact with RNA chaperones, upstream open reading frames and/or G-quadruplex structures for function.

  • The initiation factor eIF3 can have a specialized function in translation by directly binding to structured or chemically modified 5′ UTRs in target mRNAs to mediate selective internal initiation.

  • Global in vivo RNA structure probing technologies use chemical modifiers to assess the transcriptome RNA structure inside cells.

  • Single-nucleotide-resolution chemical RNA structure probing is emerging to explore the folding state of the transcriptome in living cells; the resulting models for how 5′ UTR structures impact translation should be validated by compensatory mutagenesis.

Abstract

RNA molecules can fold into intricate shapes that can provide an additional layer of control of gene expression beyond that of their sequence. In this Review, we discuss the current mechanistic understanding of structures in 5′ untranslated regions (UTRs) of eukaryotic mRNAs and the emerging methodologies used to explore them. These structures may regulate cap-dependent translation initiation through helicase-mediated remodelling of RNA structures and higher-order RNA interactions, as well as cap-independent translation initiation through internal ribosome entry sites (IRESs), mRNA modifications and other specialized translation pathways. We discuss known 5′ UTR RNA structures and how new structure probing technologies coupled with prospective validation, particularly compensatory mutagenesis, are likely to identify classes of structured RNA elements that shape post-transcriptional control of gene expression and the development of multicellular organisms.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Interspecies variation in 5′ UTR lengths.
Figure 2: Cis-acting regulatory RNA elements and structures in eukaryotic 5′ UTRs influence mRNA translation.
Figure 3: Cellular IRES structures employ different mechanisms for ribosome recruitment.
Figure 4: The effects of N6-methyladenosine on mRNA translation and decay.
Figure 5: Global RNA structure probing to assess translation regulation.

Similar content being viewed by others

References

  1. Breaker, R. R. Riboswitches and the RNA world. Cold Spring Harb. Perspect. Biol. 4, 1–15 (2012).

    Article  CAS  Google Scholar 

  2. Bowman, J. C., Hud, N. V. & Williams, L. D. The ribosome challenge to the RNA world. J. Mol. Evol. 80, 143–161 (2015).

    Article  CAS  PubMed  Google Scholar 

  3. Nahvi, A. et al. Genetic control by a metabolite binding mRNA. Chem. Biol. 9, 1043–1049 (2002).

    Article  CAS  PubMed  Google Scholar 

  4. Wachter, A. Gene regulation by structured mRNA elements. Trends Genet. 30, 172–181 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Martin, W. & Koonin, E. V. Introns and the origin of nucleus-cytosol compartmentalization. Nature 440, 41–45 (2006).

    Article  CAS  PubMed  Google Scholar 

  6. Morris, K. V. & Mattick, J. S. The rise of regulatory RNA. Nat. Rev. Genet. 15, 423–437 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Jackson, R. J., Hellen, C. U. T. & Pestova, T. V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11, 113–127 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Simsek, D. et al. The mammalian ribo-interactome reveals ribosome functional diversity and heterogeneity. Cell 169, 1051–1057 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Shi, Z. et al. Heterogeneous ribosomes preferentially translate distinct subpools of mRNAs genome-wide. Mol. Cell 67, 71–83 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Topisirovic, I., Svitkin, Y. V., Sonenberg, N. & Shatkin, A. J. Cap and cap-binding proteins in the control of gene expression. Wiley Interdiscip. Rev. RNA 2, 277–298 (2011).

    Article  CAS  PubMed  Google Scholar 

  12. Pichon, X. et al. RNA binding protein/RNA element interactions and the control of translation. Curr. Protein Pept. Sci. 13, 294–304 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Manning, K. S. & Cooper, T. A. The roles of RNA processing in translating genotype to phenotype. Nat. Rev. Mol. Cell Biol. 18, 102–114 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 31, 3406–3415 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lorenz, R. et al. ViennaRNA package 2.0. Algorithms Mol. Biol. 6, 26 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Mayr, C. Regulation by 3′-untranslated regions. Annu. Rev. Genet. 51, 171–194 (2017).

    Article  CAS  PubMed  Google Scholar 

  17. Pesole, G. et al. Structural and functional features of eukaryotic mRNA untranslated regions. Gene 276, 73–81 (2001).

    Article  CAS  PubMed  Google Scholar 

  18. Mazumder, B., Seshadri, V. & Fox, P. L. Translational control by the 3′-UTR: the ends specify the means. Trends Biochem. Sci. 28, 91–98 (2003).

    Article  CAS  PubMed  Google Scholar 

  19. Lynch, M., Scofield, D. G. & Hong, X. The evolution of transcription-initiation sites. Mol. Biol. Evol. 22, 1137–1146 (2005).

    Article  CAS  PubMed  Google Scholar 

  20. Hinnebusch, A. G. The scanning mechanism of eukaryotic translation initiation. Annu. Rev. Biochem. 83, 779–812 (2014).

    Article  CAS  PubMed  Google Scholar 

  21. Hinnebusch, A. G., Ivanov, I. P. & Sonenberg, N. Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science 352, 1413–1416 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ferreira, J. P., Overton, K. W. & Wang, C. L. Tuning gene expression with synthetic upstream open reading frames. Proc. Natl Acad. Sci. USA 110, 11284–11289 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Somers, J., Pöyry, T. & Willis, A. E. A perspective on mammalian upstream open reading frame function. Int. J. Biochem. Cell Biol. 45, 1690–1700 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Araujo, P. R. et al. Before it gets started: regulating translation at the 5′ UTR. Comp. Funct. Genomics 2012, 1–8 (2012).

    Article  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  27. Noderer, W. L. et al. Quantitative analysis of mammalian translation initiation sites by FACS-seq. Mol. Syst. Biol. 10, 1–14 (2014).

    Article  CAS  Google Scholar 

  28. Mathews, M. B., Sonenberg, N. & Hershey, J. W. B. Origins and principles of translational control. Cold Spring Harbor Monogr. Ser. 48, 1–40 (2007).

    CAS  Google Scholar 

  29. Schott, J. et al. Translational regulation of specific mRNAs controls feedback inhibition and survival during macrophage activation. PLOS Genet. 10, e1004368 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hernández, G., Altmann, M. & Lasko, P. Origins and evolution of the mechanisms regulating translation initiation in eukaryotes. Trends Biochem. Sci. 35, 63–73 (2010).

    Article  CAS  PubMed  Google Scholar 

  31. Hernandez, G., Osnaya, V. G., Garcia, A. & Velasco, M. X. Evolution of the Protein Synthesis Machinery and its Regulation (eds Hernández, G. & Jagus, R.) 81–107 (Springer Int. Pub. 2016).

    Book  Google Scholar 

  32. Mignone, F., Gissi, C., Liuni, S. & Pesole, G. Untranslated regions of mRNAs. Genome Biol. 3, 4.1–4.10 (2002).

    Article  Google Scholar 

  33. Jan, C. H., Friedman, R. C., Ruby, J. G. & Bartel, D. P. Formation, regulation and evolution of Caenorhabditis elegans 3′UTRs. Nature 469, 97–101 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Nagalakshmi, U. et al. The transcriptional landscape of the yeast genome defined by RNA sequencing. Science 320, 1344–1349 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Pelletier, J. & Sonenberg, N. Insertion mutagenesis to increase secondary structure within the 5′ noncoding region of a eukaryotic mRNA reduces translational efficiency. Cell 40, 515–526 (1985). This is one of the first studies to show a role for a 5′ UTR RNA secondary structure in influencing mRNA translation.

    Article  CAS  PubMed  Google Scholar 

  36. Manzella, J. M. & Blackshear, P. J. Regulation of rat ornithine decarboxylase mRNA translation by its 5′-untranslated region. J. Biol. Chem. 265, 11817–11822 (1990).

    CAS  PubMed  Google Scholar 

  37. Taliaferro, J. M. et al. RNA sequence context effects measured in vitro predict in vivo protein binding and regulation. Mol. Cell 64, 294–306 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Montoya, J., Ojala, D. & Attardi, G. Distinctive features of the 5′-terminal sequences of the human mitochondrial mRNAs. Nature 290, 465–470 (1981).

    Article  CAS  PubMed  Google Scholar 

  39. Haimov, O., Sinvani, H. & Dikstein, R. Cap-dependent, scanning-free translation initiation mechanisms. Biochim. Biophys. Acta 1849, 1313–1318 (2015).

    Article  CAS  PubMed  Google Scholar 

  40. Hentze, M. W. et al. Identification of the iron responsive element for the translational regulation of human ferritin mRNA. Science 238, 1570–1573 (1987).

    Article  CAS  PubMed  Google Scholar 

  41. Muckenthaler, M. U., Rivella, S., Hentze, M. W. & Galy, B. A red carpet for iron metabolism. Cell 3, 1–18 (2017).

    Google Scholar 

  42. Gray, N. K. & Hentze, M. W. Iron regulatory protein prevents binding of the 43S translation pre-initiation complex to ferritin and eALAS mRNAs. EMBO J. 13, 3882–3891 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Muckenthaler, M., Gray, N. K. & Hentze, M. W. IRP-1 binding to ferritin mRNA prevents the recruitment of the small ribosomal subunit by the cap-binding complex eIF4F. Mol. Cell 2, 383–388 (1998).

    Article  CAS  PubMed  Google Scholar 

  44. Babendure, J. R., Babendure, J. L., Ding, J.-H. & Tsien, R. Y. Control of mammalian translation by mRNA structure near caps. RNA 12, 851–861 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Kozak, M. Influences of mRNA secondary structure on initiation by eukaryotic ribosomes. Proc. Natl Acad. Sci. USA 83, 2850–2854 (1986).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Parsyan, A. et al. mRNA helicases: the tacticians of translational control. Nat. Rev. Mol. Cell Biol. 12, 235–245 (2011).

    Article  CAS  PubMed  Google Scholar 

  47. Wolfe, A. L. et al. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513, 65–70 (2014). Ribosome profiling in cancer cells following eIF4A perturbation identifies stable RG4 structures in the 5′ UTR in eIF4A-sensitive target mRNAs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Rubio, C. A. et al. Transcriptome-wide characterization of the eIF4A signature highlights plasticity in translation regulation. Genome Biol. 15, 1–19 (2014).

    Article  CAS  Google Scholar 

  49. Iwasaki, S., Floor, S. N. & Ingolia, N. T. Rocaglates convert DEAD-box protein eIF4A into a sequence-selective translational repressor. Nature 534, 558–561 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Özes, A. R., Feoktistova, K., Avanzino, B. C. & Fraser, C. S. Duplex unwinding and ATPase activities of the DEAD-box helicase eIF4A are coupled by eIF4G and eIF4B. J. Mol. Biol. 412, 674–687 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Dmitriev, S. E. et al. Efficient translation initiation directed by the 900-nucleotide-long and GC-rich 5′ untranslated region of the human retrotransposon LINE-1 mRNA is strictly cap dependent rather than internal ribosome entry site mediated. Mol. Cell. Biol. 27, 4685–4697 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bordeleau, M. E. et al. Therapeutic suppression of translation initiation modulates chemosensitivity in a mouse lymphoma model. J. Clin. Invest. 118, 2651–2660 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Cencic, R. et al. Antitumor activity and mechanism of action of the cyclopenta[b]benzofuran, silvestrol. PLoS ONE 4, e5223 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Sadlish, H. et al. Evidence for a functionally relevant rocaglamide binding site on the eIF4A-RNA complex. ACS Chem. Biol. 8, 1519–1527 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Chu, J. & Pelletier, J. Targeting the eIF4A RNA helicase as an anti-neoplastic approach. Biochim. Biophys. Acta 1849, 781–791 (2014).

    Article  CAS  PubMed  Google Scholar 

  56. Gandin, V. et al. NanoCAGE reveals 5′ UTR features that define specific modes of translation of functionally related MTOR-sensitive mRNAs. Genome Res. 26, 636–648 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Modelska, A. et al. The malignant phenotype in breast cancer is driven by eIF4A1-mediated changes in the translational landscape. Cell Death Dis. 6, e1603 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Sen, N. D., Zhou, F., Ingolia, N. T. & Hinnebusch, A. G. Genome-wide analysis of translational efficiency reveals distinct but overlapping functions of yeast DEAD-box RNA helicases Ded1 and eIF4A. Genome Res. 25, 1196–1205 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Pisareva, V. P., Pisarev, A. V., Komar, A. A., Hellen, C. U. T. & Pestova, T. V. Translation initiation on mammalian mRNAs with structured 5′UTRs requires DexH-box protein DHX29. Cell 135, 1237–1250 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Sen, N. D., Zhou, F., Harris, M. S., Ingolia, N. T. & Hinnebusch, A. G. eIF4B stimulates translation of long mRNAs with structured 5′ UTRs and low closed-loop potential but weak dependence on eIF4G. Proc. Natl Acad. Sci. USA 113, 10464–10472 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Jungfleisch, J. et al. A novel translational control mechanism involving RNA structures within coding sequences. Genome Res. 27, 95–106 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Hänsel-Hertsch, R., Di Antonio, M. & Balasubramanian, S. DNA G-quadruplexes in the human genome: detection, functions and therapeutic potential. Nat. Rev. Mol. Cell Biol. 5, 279–284 (2017).

    Article  CAS  Google Scholar 

  63. Cammas, A. & Millevoi, S. RNA G-quadruplexes: emerging mechanisms in disease. Nucleic Acids Res. 45, 1584–1595 (2016).

    PubMed Central  Google Scholar 

  64. Fay, M. M., Lyons, S. M. & Ivanov, P. RNA G-quadruplexes in biology: principles and molecular mechanisms. J. Mol. Biol. 429, 2127–2147 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Bugaut, A. & Balasubramanian, S. 5′-UTR RNA G-quadruplexes: translation regulation and targeting. Nucleic Acids Res. 40, 4727–4741 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Song, J., Perreault, J.-P., Topisirovic, I. & Richard, S. RNA G-quadruplexes and their potential regulatory roles in translation. Translation (Austin) 4, e1244031 (2016).

    Google Scholar 

  67. Beaudoin, J. D. & Perreault, J. P. 5′-UTR G-quadruplex structures acting as translational repressors. Nucleic Acids Res. 38, 7022–7036 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Bolduc, F., Garant, J. M., Allard, F. & Perreault, J. P. Irregular G-quadruplexes found in the untranslated regions of human mRNAs influence translation. J. Biol. Chem. 291, 21751–21760 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Kumari, S., Bugaut, A. & Balasubramanian, S. Position and stability are determining factors for translation repression by an RNA G-quadruplex-forming sequence within the 5′ UTR of the NRAS proto-oncogene. Biochemistry 47, 12664–12669 (2008).

    Article  CAS  PubMed  Google Scholar 

  70. Halder, K., Wieland, M. & Hartig, J. S. Predictable suppression of gene expression by 5′-UTR-based RNA quadruplexes. Nucleic Acids Res. 37, 6811–6817 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kumari, S., Bugaut, A., Huppert, J. L. & Balasubramanian, S. An RNA G-quadruplex in the 5′ UTR of the NRAS proto-oncogene modulates translation. Nat. Chem. Biol. 3, 218–221 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Arora, A. et al. Inhibition of translation in living eukaryotic cells by an RNA G-quadruplex motif. RNA 14, 1290–1296 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Melko, M. & Bardoni, B. The role of G-quadruplex in RNA metabolism: involvement of FMRP and FMR2P. Biochimie 92, 919–926 (2010).

    Article  CAS  PubMed  Google Scholar 

  74. Darnell, J. C. et al. Fragile X mental retardation protein targets G quartet mRNAs important for neuronal function. Cell 107, 489–499 (2001).

    Article  CAS  PubMed  Google Scholar 

  75. Anderson, B. R. et al. Identification of consensus binding sites clarifies FMRP binding determinants. Nucleic Acids Res. 44, 6649–6659 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Schaeffer, C. et al. The fragile X mental retardation protein binds specifically to its mRNA via a purine quartet motif. EMBO J. 20, 4803–4813 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Didiot, M. C. et al. The G-quartet containing FMRP binding site in FMR1 mRNA is a potent exonic splicing enhancer. Nucleic Acids Res. 36, 4902–4912 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Castets, M. et al. FMRP interferes with the Rac1 pathway and controls actin cytoskeleton dynamics in murine fibroblasts. Hum. Mol. Genet. 14, 835–844 (2005).

    Article  CAS  PubMed  Google Scholar 

  79. Darnell, J. C. et al. FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 146, 247–261 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Chen, E., Sharma, M. R., Shi, X., Agrawal, R. K. & Joseph, S. Fragile X mental retardation protein regulates translation by binding directly to the ribosome. Mol. Cell 54, 407–417 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Benhalevy, D. et al. The human CCHC-type zinc finger nucleic acid-binding protein binds G-rich elements in target mRNA coding sequences and promotes translation. Cell Rep. 18, 2979–2990 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Ben-Asouli, Y., Banai, Y., Pel-Or, Y., Shir, A. & Kaempfer, R. Human interferon-γ mRNA autoregulates its translation through a pseudoknot that activates the interferon-inducible protein kinase PKR. Cell 108, 221–232 (2002).

    Article  CAS  PubMed  Google Scholar 

  83. Cohen-Chalamish, S. et al. Dynamic refolding of IFN-γ mRNA enables it to function as PKR activator and translation template. Nat. Chem. Biol. 5, 896–903 (2009). The studies in Refs 82 and 83 show how short helical segments assemble into a pseudoknot structure in a 5′ UTR, which is bound by PKR and activates it, lead to inhibition of translation initiation.

    Article  CAS  PubMed  Google Scholar 

  84. Nallagatla, S. R., Toroney, R. & Bevilacqua, P. C. Regulation of innate immunity through RNA structure and the protein kinase PKR. Curr. Opin. Struct. Biol. 21, 119–127 (2011).

    Article  CAS  PubMed  Google Scholar 

  85. Yoon, J.-H., Abdelmohsen, K. & Gorospe, M. Posttranscriptional gene regulation by long noncoding RNA. J. Mol. Biol. 425, 3723–3730 (2013).

    Article  CAS  PubMed  Google Scholar 

  86. Carrieri, C. et al. Long non-coding antisense RNA controls Uchl1 translation through an embedded SINEB2 repeat. Nature 491, 454–457 (2012).

    Article  CAS  PubMed  Google Scholar 

  87. Brierley, I., Gilbert, R. J. & Pennell, S. RNA pseudoknots and the regulation of protein synthesis. Biochem. Soc. Trans. 36, 684–689 (2008).

    Article  CAS  PubMed  Google Scholar 

  88. Osman, F., Jarrous, N., Ben-Asouli, Y. & Kaempfer, R. A cis-acting element in the 3′-untranslated region of human TNF-α mRNA renders splicing dependent on the activation of protein kinase PKR. Genes Dev. 13, 3280–3293 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Namer, L. S. et al. An ancient pseudoknot in TNF-α pre-mRNA activates PKR, inducing eIF2α phosphorylation that potently enhances splicing. Cell Rep. 20, 188–200 (2017).

    Article  CAS  PubMed  Google Scholar 

  90. Reenan, R. A. Molecular determinants and guided evolution of species-specific RNA editing. Nature 434, 409–413 (2005).

    Article  CAS  PubMed  Google Scholar 

  91. Farabaugh, P. J. Programmed translational frameshifting. Microbiol. Rev. 60, 103–134 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Advani, V. M. & Dinman, J. D. Reprogramming the genetic code: the emerging role of ribosomal frameshifting in regulating cellular gene expression. Bioessays 38, 21–26 (2016).

    Article  CAS  PubMed  Google Scholar 

  94. Brierley, I., Digard, P. & Inglis, S. C. Characterization of an efficient coronavirus ribosomal frameshifting signal: requirement for an RNA pseudoknot. Cell 57, 537–547 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Baril, M., Dulude, D., Steinberg, S. V. & Brakier-Gingras, L. The frameshift stimulatory signal of human immunodeficiency virus type 1 group O is a pseudoknot. J. Mol. Biol. 331, 571–583 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Namy, O., Moran, S. J., Stuart, D. I., Gilbert, R. J. C. & Brierley, I. A mechanical explanation of RNA pseudoknot function in programmed ribosomal frameshifting. Nature 441, 244–247 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Belew, A. T., Advani, V. M. & Dinman, J. D. Endogenous ribosomal frameshift signals operate as mRNA destabilizing elements through at least two molecular pathways in yeast. Nucleic Acids Res. 39, 2799–2808 (2011).

    Article  CAS  PubMed  Google Scholar 

  98. Belew, A. T. et al. Ribosomal frameshifting in the CCR5 mRNA is regulated by miRNAs and the NMD pathway. Nature 512, 265–269 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Holcik, M. & Sonenberg, N. Translational control in stress and apoptosis. Nat. Rev. Mol. Cell Biol. 6, 318–327 (2005).

    Article  CAS  PubMed  Google Scholar 

  100. Qin, X. & Sarnow, P. Preferential translation of internal ribosome entry site-containing mRNAs during the mitotic cycle in mammalian cells. J. Biol. Chem. 279, 13721–13728 (2004).

    Article  CAS  PubMed  Google Scholar 

  101. Jackson, R. J. The current status of vertebrate cellular mRNA IRESs. Cold Spring Harb. Perspect. Biol. 5, a011569 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Pestova, T. V. et al. Molecular mechanisms of translation initiation in eukaryotes. Proc. Natl Acad. Sci. USA 98, 7029–7036 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Stoneley, M. & Willis, A. E. Cellular internal ribosome entry segments: structures, trans-acting factors and regulation of gene expression. Oncogene 23, 3200–3207 (2004).

    Article  CAS  PubMed  Google Scholar 

  104. Spriggs, K. A., Stoneley, M., Bushell, M. & Willis, A. E. Re-programming of translation following cell stress allows IRES-mediated translation to predominate. Biol. Cell 100, 27–38 (2008).

    Article  CAS  PubMed  Google Scholar 

  105. Lacerda, R., Menezes, J. & Romão, L. More than just scanning: the importance of cap-independent mRNA translation initiation for cellular stress response and cancer. Cell. Mol. Life Sci. 74, 1659–1680 (2016).

    Article  CAS  PubMed  Google Scholar 

  106. Macejak, D. G. & Sarnow, P. Internal initiation of translation mediated by the 5′ leader of a cellular mRNA. Nature 353, 90–94 (1991). This study presents the first description of IRES activity in the 5′ UTR of a cellular mRNA.

    Article  CAS  PubMed  Google Scholar 

  107. Thoma, C., Bergamini, G., Galy, B., Hundsdoerfer, P. & Hentze, M. W. Enhancement of IRES-mediated translation of the c-myc and BiP mRNAs by the poly(A) tail is independent of intact eIF4G and PABP. Mol. Cell 15, 925–935 (2004).

    Article  CAS  PubMed  Google Scholar 

  108. Buchkovich, N. J., Yu, Y., Pierciey, F. J. & Alwine, J. C. 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, 11479–11486 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Mokrejs, M. et al. IRESite — a tool for the examination of viral and cellular internal ribosome entry sites. Nucleic Acids Res. 38, D131–D136 (2010).

    Article  CAS  PubMed  Google Scholar 

  110. Baranick, B. T. et al. Splicing mediates the activity of four putative cellular internal ribosome entry sites. Proc. Natl Acad. Sci. USA 105, 4733–4738 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Komar, A. A. & Hatzoglou, M. Internal ribosome entry sites in cellular mRNAs: mystery of their existence. J. Biol. Chem. 280, 23425–23428 (2005).

    Article  CAS  PubMed  Google Scholar 

  112. Komar, A. A. & Hatzoglou, M. Cellular IRES-mediated translation: the war of ITAFs in pathophysiological states. Cell Cycle 10, 229–240 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Baird, S. D., Lewis, S. M., Turcotte, M. & Holcik, M. A search for structurally similar cellular internal ribosome entry sites. Nucleic Acids Res. 35, 4664–4677 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Mitchell, S. A. et al. Identification of a motif that mediates polypyrimidine tract-binding protein-dependent internal ribosome entry. Genes Dev. 19, 1556–1571 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. King, H. A., Cobbold, L. C. & Willis, A. E. The role of IRES trans-acting factors in regulating translation initiation. Biochem. Soc. Trans. 38, 1581–1586 (2010).

    Article  CAS  PubMed  Google Scholar 

  116. Faye, M. D. & Holcik, M. The role of IRES trans-acting factors in carcinogenesis. Biochim. Biophys. Acta 1849, 887–897 (2014).

    Article  CAS  PubMed  Google Scholar 

  117. Lewis, S. M. & Holcik, M. For IRES trans-acting factors, it is all about location. Oncogene 27, 1033–1035 (2008).

    Article  CAS  PubMed  Google Scholar 

  118. Cobbold, L. C. et al. Upregulated c-myc expression in multiple myeloma by internal ribosome entry results from increased interactions with and expression of PTB-1 and YB-1. Oncogene 29, 2884–2891 (2010).

    Article  CAS  PubMed  Google Scholar 

  119. Mitchell, S. A., Spriggs, K. A., Coldwell, M. J., Jackson, R. J. & Willis, A. E. The Apaf-1 internal ribosome entry segment attains the correct structural conformation for function via interactions with PTB and unr. Mol. Cell 11, 757–771 (2003).

    Article  CAS  PubMed  Google Scholar 

  120. Pickering, B. M., Mitchell, S. A., Evans, J. R. & Willis, A. E. Polypyrimidine tract binding protein and poly r(C) binding protein 1 interact with the BAG-1 IRES and stimulates its activity in vitro and in vivo. Nucleic Acids Res. 31, 639–646 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Pickering, B. M., Mitchell, S. A., Spriggs, K. A., Willis, A. E. & Stoneley, M. Bag-1 internal ribosome entry segment activity is promoted by structural changes mediated by poly(rC) binding protein 1 and recruitment of polypyrimidine tract binding protein 1. Mol. Cell. Biol. 24, 5595–5605 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Le Quesne, J. P., Stoneley, M., Fraser, G. A. & Willis, A. E. Derivation of a structural model for the c-myc IRES. J. Mol. Biol. 310, 111–126 (2001).

    Article  CAS  PubMed  Google Scholar 

  123. Chappell, S. A. et al. A mutation in the c-myc-IRES leads to enhanced internal ribosome entry in multiple myeloma: a novel mechanism of oncogene de-regulation. Oncogene 19, 4437–4440 (2000).

    Article  CAS  PubMed  Google Scholar 

  124. Yaman, I. et al. The zipper model of translational control: a small upstream ORF is the switch that controls structural remodeling of an mRNA leader. Cell 113, 519–531 (2003).

    Article  CAS  PubMed  Google Scholar 

  125. Fernandez, J. et al. Ribosome stalling regulates ires-mediated translation in eukaryotes, a parallel to prokaryotic attenuation. Mol. Cell 17, 405–416 (2005).

    Article  CAS  PubMed  Google Scholar 

  126. Chen, T. M. et al. Overexpression of FGF9 in colon cancer cells is mediated by hypoxia-induced translational activation. Nucleic Acids Res. 42, 2932–2944 (2014).

    Article  CAS  PubMed  Google Scholar 

  127. Bastide, A. et al. An upstream open reading frame within an IRES controls expression of a specific VEGF-A isoform. Nucleic Acids Res. 36, 2434–2445 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Fernandez, J. et al. Regulation of internal ribosome entry site-mediated translation by eukaryotic initiation factor-2a phosphorylation and translation of a small upstream open reading frame. J. Biol. Chem. 277, 2050–2058 (2002).

    Article  CAS  PubMed  Google Scholar 

  129. Morris, M. J., Negishi, Y., Pazsint, C., Schonhoft, J. D. & Basu, S. An RNA G-quadruplex is essential for cap-independent translation initiation in human VEGF IRES. J. Am. Chem. Soc. 132, 17831–17839 (2010).

    Article  CAS  PubMed  Google Scholar 

  130. Cammas, A. et al. Stabilization of the G-quadruplex at the VEGF IRES represses cap-independent translation. RNA Biol. 12, 320–329 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  131. Bonnal, S. et al. A single internal ribosome entry site containing a G quartet RNA structure drives fibroblast growth factor 2 gene expression at four alternative translation initiation codons. J. Biol. Chem. 278, 39330–39336 (2003).

    Article  CAS  PubMed  Google Scholar 

  132. Bhattacharyya, D., Diamond, P. & Basu, S. An independently folding RNA G-quadruplex domain directly recruits the 40S ribosomal subunit. Biochemistry 54, 1879–1885 (2015).

    Article  CAS  PubMed  Google Scholar 

  133. Xue, S. et al. RNA regulons in Hox 5′ UTRs confer ribosome specificity to gene regulation. Nature 517, 33–38 (2015). The first targeted knockout of a cellular IRES in mice reveals a key role for IRES-driven translation of Hox mRNAs in vivo and identifies a new cis -regulatory element, the TIE, which blocks cap-dependent translation.

    Article  CAS  PubMed  Google Scholar 

  134. Landry, D. M., Hertz, M. I. & Thompson, S. R. RPS25 is essential for translation initiation by the Dicistroviridae and hepatitis C viral IRESs. Genes Dev. 23, 2753–2764 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Jack, K. et al. rRNA pseudouridylation defects affect ribosomal ligand binding and translational fidelity from yeast to human cells. Mol. Cell 44, 660–666 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Kondrashov, N. et al. Ribosome-mediated specificity in Hox mRNA translation and vertebrate tissue patterning. Cell 145, 383–397 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Martineau, Y. et al. Internal ribosome entry site structural motifs conserved among mammalian fibroblast growth factor 1 alternatively spliced mRNAs. Mol. Cell. Biol. 24, 7622–7635 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Cheng, C. Y. et al. Consistent global structures of complex RNA states through multidimensional chemical mapping. eLife 4, 1–38 (2015).

    CAS  Google Scholar 

  139. Créancier, L., Morello, D., Mercier, P. & Prats, A. C. Fibroblast growth factor 2 internal ribosome entry site (IRES) activity ex vivo and in transgenic mice reveals a stringent tissue-specific regulation. J. Cell Biol. 150, 275–281 (2000).

    Article  PubMed  PubMed Central  Google Scholar 

  140. Créancier, L., Mercier, P., Prats, A., Morello, D. & Cre, L. c-Myc internal ribosome entry site activity is developmentally controlled and subjected to a strong translational repression in adult transgenic mice. Mol. Cell. Biol. 21, 1833–1840 (2001).

    Article  PubMed  PubMed Central  Google Scholar 

  141. Morfoisse, F. et al. Hypoxia induces VEGF-C expression in metastatic tumor cells via a HIF-1a-independent translation-mediated mechanism. Cell Rep. 6, 155–167 (2014).

    Article  CAS  PubMed  Google Scholar 

  142. Audigier, S. et al. Potent activation of FGF-2 IRES-dependent mechanism of translation during brain development. RNA 14, 1852–1864 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Conte, C. et al. FGF2 translationally induced by hypoxia is involved in negative and positive feedback loops with HIF-1α. PLOS ONE 3, e3078 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Weingarten-Gabbay, S. et al. Systematic discovery of cap-independent translation sequences in human and viral genomes. Science 351, 1–24 (2016). This study reports the first genome-wide approach to identify novel IRES-like elements in vivo and finds short poly(U) sequences and small structured elements that harbour IRES activity.

    Article  CAS  Google Scholar 

  145. Lee, A. S. Y., Kranzusch, P. J. & Cate, J. H. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522, 111–114 (2015). This study identifies a novel role for eIF3 in interacting with RNA stem–loop structures in the 5′ UTR of mRNAs to directly recruit the ribosome.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Blau, L. et al. Aberrant expression of c-Jun in glioblastoma by internal ribosome entry site (IRES)-mediated translational activation. Proc. Natl Acad. Sci. USA 109, E2875–E2884 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Lee, A. S., Kranzusch, P. J., Doudna, J. A. & Cate, J. H. eIF3d is an mRNA cap-binding protein that is required for specialized translation initiation. Nature 536, 96–99 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Zhao, B. S., Roundtree, I. A. & He, C. Post-transcriptional gene regulation by mRNA modifications. Nat. Rev. Mol. Cell Biol. 18, 31–42 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Harcourt, E. M., Kietrys, A. M. & Kool, E. T. Chemical and structural effects of base modifications in messenger RNA. Nature 541, 339–346 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Dominissini, D. et al. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 485, 201–206 (2012).

    Article  CAS  PubMed  Google Scholar 

  151. Meyer, K. D. et al. 5′ UTR m6A promotes cap-independent translation. Cell 163, 999–1010 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Zhou, J. et al. Dynamic m(6)A mRNA methylation directs translational control of heat shock response. Nature 526, 591–594 (2015). Refs 151 and 152 demonstrate that m6A promotes cap-independent mRNA translation initiation, which is mediated by direct eIF3 recruitment. They also show that m6A levels in 5′ UTRs are increased during stress.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Liu, N. et al. N(6)-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions. Nature 518, 560–564 (2015). This study reports that the incorporation of m6A into a hairpin modulates the local secondary structure of RNA ('m6A switch'), thereby facilitating the binding of the indirect m6A reader hnRNPC.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Spitale, R. C. et al. Structural imprints in vivo decode RNA regulatory mechanisms. Nature 519, 486–490 (2015). This paper presents the first global in vivo RNA structure probing method, icSHAPE, and connects m6A modifications with destabilization of local RNA structure across the transcriptome.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Schwartz, S. et al. High-Resolution mapping reveals a conserved, widespread, dynamic mRNA methylation program in yeast meiosis. Cell 155, 1409–1421 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Roost, C. et al. Structure and thermodynamics of N6-methyladenosine in RNA: a spring-loaded base modification. J. Am. Chem. Soc. 137, 2107–2115 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Wang, X. et al. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 505, 117–120 (2014).

    Article  CAS  PubMed  Google Scholar 

  158. Wang, X. et al. N6-methyladenosine modulates messenger RNA translation efficiency. Cell 161, 1388–1399 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Lin, S., Choe, J., Du, P., Triboulet, R. & Gregory, R. I. The m6A methyltransferase METTL3 promotes translation in human cancer cells. Mol. Cell 62, 335–345 (2015).

    Article  CAS  Google Scholar 

  160. Dominissini, D. et al. The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA. Nature 530, 441–446 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Delatte, B. et al. Transcriptome-wide distribution and function of RNA hydroxymethylcytosine. Science 351, 282–285 (2016).

    Article  CAS  PubMed  Google Scholar 

  162. Slobodin, B. et al. Transcription impacts the efficiency of mRNA translation via co-transcriptional N6-adenosine methylation. Cell 169, 326–337 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Choi, J. et al. N(6)-methyladenosine in mRNA disrupts tRNA selection and translation-elongation dynamics. Nat. Struct. Mol. Biol. 23, 110–115 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Legnini, I. et al. Circ-ZNF609 is a circular RNA that can be translated and functions in myogenesis. Mol. Cell 66, 22–37 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Pamudurti, N. R. et al. Translation of CircRNAs. Mol. Cell 66, 9–21 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Zhou, C. et al. Genome-wide maps of m6A circRNAs identify widespread and cell-type-specific methylation patterns that are distinct from mRNAs. Cell Rep. 20, 2262–2276 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Mortimer, S. A., Kidwell, M. A. & Doudna, J. A. Insights into RNA structure and function from genome-wide studies. Nat. Rev. Genet. 15, 469–479 (2014).

    Article  CAS  PubMed  Google Scholar 

  168. Kwok, C. K., Tang, Y., Assmann, S. M. & Bevilacqua, P. C. The RNA structurome: transcriptome-wide structure probing with next-generation sequencing. Trends Biochem. Sci. 40, 221–232 (2015).

    Article  CAS  PubMed  Google Scholar 

  169. Kubota, M., Tran, C. & Spitale, R. C. Progress and challenges for chemical probing of RNA structure inside living cells. Nat. Chem. Biol. 11, 933–941 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Bevilacqua, P. C., Ritchey, L. E., Su, Z. & Assmann, S. M. Genome-wide analysis of RNA secondary structure. Annu. Rev. Genet. 50, 235–266 (2016).

    Article  CAS  PubMed  Google Scholar 

  171. Incarnato, D. & Oliviero, S. The RNA epistructurome: uncovering RNA function by studying structure and post-transcriptional modifications. Trends Biotechnol. 35, 318–333 (2017).

    Article  CAS  PubMed  Google Scholar 

  172. Rouskin, S., Zubradt, M., Washietl, S., Kellis, M. & Weissman, J. S. Genome-wide probing of RNA structure reveals active unfolding of mRNA structures in vivo. Nature 505, 701–705 (2014).

    Article  CAS  PubMed  Google Scholar 

  173. Ding, Y. et al. In vivo genome-wide profiling of RNA secondary structure reveals novel regulatory features. Nature 505, 696–700 (2014). Refs 172 and 173 introduce the first methods for in vivo DMS-based chemical modification of accessible RNA nucleotides to globally probe RNA secondary structure.

    Article  CAS  PubMed  Google Scholar 

  174. Kertesz, M. et al. Genome-wide measurement of RNA secondary structure in yeast. Nature 467, 103–107 (2010).

    Article  CAS  PubMed  Google Scholar 

  175. Siegfried, N. A., Busan, S., Rice, G. M., Nelson, J. A. & Weeks, K. M. RNA motif discovery by SHAPE and mutational profiling (SHAPE-MaP). Nat. Methods 11, 959–965 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Fang, R., Moss, W. N., Rutenberg-Schoenberg, M. & Simon, M. D. Probing xist RNA structure in cells using targeted structure-seq. PLoS Genet. 11, e1005668 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Zubradt, M. et al. DMS-MaPseq for genome-wide or targeted RNA structure probing in vivo. Nat. Methods 14, 75–82 (2016). At the current vanguard of in vivo chemical mapping techniques, DMS–MaPseq leverages several methodological advances to pilot applications to probe low-abundance mRNAs and pre-mRNAs and to derive structural models that inform compensatory mutagenesis experiments.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Homan, P. J. et al. Single-molecule correlated chemical probing of RNA. Proc. Natl Acad. Sci. USA 111, 13858–13863 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Sergiev, P. V., Dontsova, O. A. & Bogdanov, A. A. Chemical methods for the structural study of the ribosome: judgment day. Mol. Biol. 35, 472–495 (2001).

    Article  CAS  Google Scholar 

  180. Miao, Z. et al. RNA-puzzles round III: 3D RNA structure prediction of five riboswitches and one ribozyme. RNA 23, 655–672 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Kladwang, W., VanLang, C. C., Cordero, P. & Das, R. A two-dimensional mutate-and-map strategy for non-coding RNA structure. Nat. Chem. 3, 954–962 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Tian, S., Cordero, P., Kladwang, W. & Das, R. High-throughput mutate-map-rescue evaluates SHAPE-directed RNA structure and uncovers excited states. RNA 20, 1815–1826 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Tian, S. & Das, R. RNA structure through multidimensional chemical mapping. Q. Rev. Biophys. 49, 1–30 (2016).

    Article  Google Scholar 

  184. Smola, M. J., Calabrese, J. M. & Weeks, K. M. Detection of RNA-protein interactions in living cells with SHAPE. Biochemistry 54, 6867–6875 (2015).

    Article  CAS  PubMed  Google Scholar 

  185. Cheng, C., Kladwang, W., Yesselman, J. & Das, R. RNA structure inference through chemical mapping after accidental or intentional mutations. Proc. Natl Acad. Sci. USA 114, 9876–9881 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Risca, V. I., Denny, S. K., Straight, A. F. & Greenleaf, W. J. Variable chromatin structure revealed by in situ spatially correlated DNA cleavage mapping. Nature 541, 237–241 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Burkhardt, D. H. et al. Operon mRNAs are organized into ORF-centric structures that predict translation efficiency. eLife 6, e22037 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  188. Dreyfuss, G., Kim, V. N., Kataoka, N. & Medical, H. H. Messenger-RNA-binding proteins and the messages they carry. Nat. Rev. Mol. Cell Biol. 3, 195–205 (2002).

    Article  CAS  PubMed  Google Scholar 

  189. Ray, D. et al. A compendium of RNA-binding motifs for decoding gene regulation. Nature 499, 172–177 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. May, J., Johnson, P., Saleem, H. & Simon, A. E. A sequence-independent, unstructured IRES is responsible for internal expression of the coat protein of Turnip Crinkle Virus. J. Virol. 91, e02421–e02416 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Clements, J., Eddy, S. R. & Rivas, E. A statistical test for conserved RNA structure shows lack of evidence for structure in lncRNAs. Nat. Methods 14, 45–48 (2017).

    Article  CAS  PubMed  Google Scholar 

  192. Pedersen, J. S. et al. Identification and classification of conserved RNA secondary structures in the human genome. PLoS Comput. Biol. 2, 251–262 (2006).

    Article  CAS  Google Scholar 

  193. Parker, B. J. et al. New families of human regulatory RNA structures identified by comparative analysis of vertebrate genomes. Genome Res. 21, 1929–1943 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Goodarzi, H. et al. Systematic discovery of structural elements governing stability of mammalian messenger RNAs. Nature 485, 264–268 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Rousseau, D., Kaspar, R., Rosenwald, I., Gehrke, L. & Sonenberg, N. Translation initiation of ornithine decarboxylase and nucleocytoplasmic transport of cyclin D1 mRNA are increased in cells overexpressing eukaryotic initiation factor 4E. Proc. Natl Acad. Sci. USA 93, 1065–1070 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Hinnebusch, A. G. Structural insights into the mechanism of scanning and start codon recognition in eukaryotic translation initiation. Trends Biochem. Sci. 1348, 1–23 (2017).

    Google Scholar 

  197. Martin, F. et al. Ribosomal 18S rRNA base pairs with mRNA during eukaryotic translation initiation. Nat. Commun. 7, 1–7 (2016).

    Google Scholar 

  198. Mahamid, J. et al. Visualizing the molecular sociology at the HeLa cell nuclear periphery. Science 351, 969–972 (2016).

    Article  CAS  PubMed  Google Scholar 

  199. Eddy, S. R. Computational analysis of conserved RNA secondary structure in transcriptomes and genomes. Annu. Rev. Biophys. 43, 433–456 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Majumder, M. et al. The hnRNA-binding proteins hnRNP L and PTB are required for efficient translation of the Cat-1 arginine/lysine transporter mRNA during amino acid starvation. Mol. Cell. Biol. 29, 2899–2912 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Ke, S. et al. A majority of m6A residues are in the last exons, allowing the potential for 3′ UTR regulation. Genes Dev. 29, 2037–2053 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors apologize to those researchers whose work was not cited owing to space limitations and thank G. W. Byeon in the Barna laboratory for data analysis and figure preparation of eukaryotic 5′ UTR lengths, the Barna laboratory members for helpful comments on the manuscript, and H.-G. Wendel, J. Kieft, P. Sarnow and M. Hentze for thoughtful suggestions. RNA research in the authors' laboratories is supported by the New York Stem Cell Foundation (M.B.), Mallinckrodt Foundation Award (M.B.), Pew Scholars Award (M.B.) and NIH grants R01HD086634 (M.B.), R21HD086730 (M.B.), R01 GM102519 (R.D.) and R35 GM122579 (R.D.). M.B. is a New York Stem Cell Foundation Robertson Investigator. K.L. is supported by an EMBO Long-Term Fellowship (ALTF 539–2015) and is the Layton Family Fellow of the Damon Runyon Cancer Research Foundation (DRG-2237-15).

Author information

Authors and Affiliations

Authors

Contributions

All authors substantially contributed to researching data for the article, discussion of the content, writing and editing and/or reviewing the manuscript before submission.

Corresponding authors

Correspondence to Rhiju Das or Maria Barna.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary information S1 (box)

Genome-wide RNA-structure probing technologies (PDF 284 kb)

Supplementary information S2 (box)

Viral IRESs and systematic IRES discovery (PDF 4714 kb)

Supplementary information S3 (table)

Experimental probing of cellular IRES RNA structures (PDF 203 kb)

PowerPoint slides

Glossary

Secondary structures

Patterns of Watson–Crick base pairs (G–C, A–U and G–U) that define the double helices of an RNA.

Tertiary structures

Interactions that orient RNA double helices into specific three-dimensional arrangements, often involving non-Watson–Crick base pairs.

Ribozymes

RNA molecules that catalyse specific biochemical reactions.

Riboswitches

Non-coding mRNA structures that sense the environmental status of a cell by directly binding to small molecule ligands, such as a metabolite or an ion. This interaction changes the RNA conformation and controls gene expression through alternative splicing, transcription or translation.

Peptidyl transferase centre

The site in the large ribosomal subunit that catalyses peptide bond formation and peptide release; it is composed entirely of RNA.

Pseudoknots

RNA tertiary structures formed by base pairing of a single-stranded loop of a hairpin with a complementary sequence outside that hairpin.

Upstream open reading frames

(uORFs). Small ORFs located in the 5′ UTR of some mRNAs. Translation of uORFs can regulate the translation of the downstream ORF.

Kozak sequence

A favourable sequence (GCCRCCAUGG in mammals, where R is a purine) surrounding the translation start codon (AUG); also called the Kozak consensus or Kozak context.

GC content

Percentage of guanine and cytosine nucleotides in an RNA molecule. G–C base pairs are more stable than A–U base pairs and can form more stable RNA structures.

Free energy

G). The stability of an RNA secondary structure, estimated as the sum of the free energies assigned to all loops and base pair stacks of a folded RNA, based on a computational folding algorithm.

5′ cap

A modified guanine nucleotide, m7GpppN (where m7G is 7-methylguanosine, p is a phosphate group and N is any base), located at the 5′ end of eukaryotic mRNAs.

Ribonucleoprotein

(RNP). A complex of proteins in association with an RNA (RNP) or mRNA (mRNP).

Ribosome profiling

Sequencing of the RNA fragments protected by ribosomes, providing a quantitative signature of the translated mRNAs at a given time.

RNA G-quadruplex

(RG4). Extremely stable RNA structure formed in G-rich regions by the stacking of at least two G-tetrads, each of them forming a square-shaped structure by non-Watson–Crick interactions between two or more layers of paired G-quartets.

Toeprinting

Nucleotide-resolution assay that uses primer extension inhibition induced by the complex of a protein, or the ribosome, bound to an mRNA, to report its position on the mRNA.

In-line probing

Nucleotide-resolution structure probing method that uses the natural instability of RNA: the 2′ hydroxyl of each nucleotide can attack its phosphodiester backbone at a rate dependent on local structure.

No-go decay

An mRNA and translation quality-control mechanism that recognizes and degrades mRNAs following prolonged ribosome stalling during translation.

Nonsense-mediated decay

An RNA surveillance mechanism that recognizes and degrades mRNAs containing premature termination codons to prevent their translation.

Selective 2′-hydroxyl acylation analysed by primer extension

(SHAPE). SHAPE and in vivo click SHAPE (icSHAPE) use cell-permeable reagents to acylate the 2′-OH of accessible, single-stranded RNA at all four nucleotide bases.

Mutate-and-map

Two-dimensional expansion of chemical probing that infers RNA base pairing by mutating each nucleotide and detecting increased chemical accessibility at other nucleotides.

m6A switches

mRNA sequences that adopt a different secondary structure depending on N6-adenosine methylation.

Dimethyl sulfate

(DMS). A cell-permeable reagent that methylates adenine and cytosine nucleotides that are not protected by base pairing. Modification sites stall or induce a mutation during primer extension by reverse transcriptase, and sequencing the resulting complementary DNA indicates the folding state of a nucleotide in an RNA.

Psoralen

A natural plant product that intercalates into DNA and RNA and reversibly crosslinks interacting RNA duplex molecules at UpA motifs upon irradiation with ultraviolet light.

Sequence covariation

Nucleotide substitutions that differ between two or more homologous genes but retain the potential for RNA base pairing in each sequence.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Leppek, K., Das, R. & Barna, M. Functional 5′ UTR mRNA structures in eukaryotic translation regulation and how to find them. Nat Rev Mol Cell Biol 19, 158–174 (2018). https://doi.org/10.1038/nrm.2017.103

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm.2017.103

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing