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:

New frontiers in translational control of the cancer genome

A Corrigendum to this article was published on 24 April 2017

This article has been updated

Key Points

  • Oncogenic signalling rewires the translational machinery to support the expression of specific functional classes of pro-tumorigenic genes.

  • An expanding list of structural and sequence-specific cis-regulatory elements control expression of the oncogenic translation programme.

  • Specific trans-acting factors, such as RNA-binding proteins, interact with cis-regulatory elements to dictate mRNA translation in cancer cells.

  • Tumour cells alter tRNA expression and modification patterns to match the codon-usage bias of mRNA networks that support the transformed phenotype.

  • Translational control is crucial for the adaptive response to tumour-associated stress and represents a new therapeutic target for selectively killing cancer cells.

  • Recent technological advances in profiling genome-wide translation, RNA–protein interactions, RNA structure, RNA modifications and tRNA expression are driving a new understanding of the molecular mechanisms underlying translational control in cancer.

Abstract

The past several years have seen dramatic leaps in our understanding of how gene expression is rewired at the translation level during tumorigenesis to support the transformed phenotype. This work has been driven by an explosion in technological advances and is revealing previously unimagined regulatory mechanisms that dictate functional expression of the cancer genome. In this Review we discuss emerging trends and exciting new discoveries that reveal how this translational circuitry contributes to specific aspects of tumorigenesis and cancer cell function, with a particular focus on recent insights into the role of translational control in the adaptive response to oncogenic stress conditions.

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: Oncogenic activation of mRNA translation.
Figure 2: mRNA regulatory elements direct specialized translation of the oncogenic programme.
Figure 3: Codon usage: a new layer of control in translation of the cancer genome.
Figure 4: Translational responses and cancer cell adaptation to tumour-associated stress.

Similar content being viewed by others

Change history

  • 06 April 2017

    In this Review, in Figure 1, for both the figure and figure legend 'eIF4B' was incorrectly labelled as 'eIF2B'. The paper has been corrected online.

References

  1. Miluzio, A. et al. Impairment of cytoplasmic eIF6 activity restricts lymphomagenesis and tumor progression without affecting normal growth. Cancer Cell 19, 765–775 (2011). This study demonstrated that mice happloinsufficient for eIF6, which regulates the formation of functional 80S ribosomes, show delayed in vivo tumorigenesis and reduced tumour growth, thus uncovering a rate-limiting role for translation initiation independent of the eIF4F complex.

    Article  CAS  PubMed  Google Scholar 

  2. Barna, M. et al. Suppression of Myc oncogenic activity by ribosomal protein haploinsufficiency. Nature 456, 971–975 (2008). This is the first study to genetically demonstrate that the ability of MYC to drive increased protein synthesis is a key determinant of oncogenicity.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Furic, L. et al. eIF4E phosphorylation promotes tumorigenesis and is associated with prostate cancer progression. Proc. Natl Acad. Sci. USA 107, 14134–14139 (2010). This paper describes the generation of a knock-in mouse that expresses a non-phosphorylatable form of eIF4E and demonstrates a crucial role for eIF4E phosphorylation during in vivo tumorigenesis.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Hsieh, A. C. et al. Genetic dissection of the oncogenic mTOR pathway reveals druggable addiction to translational control via 4EBP–eIF4E. Cancer Cell 17, 249–261 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Faller, W. J. et al. mTORC1-mediated translational elongation limits intestinal tumour initiation and growth. Nature 517, 497–500 (2015). This exciting study demonstrates that oncogenic activation of translation elongation through eEF2 can be rate limiting for tumorigenesis.

    Article  CAS  PubMed  Google Scholar 

  6. Wolfe, A. L. et al. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513, 65–70 (2014). In this study, ribosome profiling in cancer cells revealed genome-wide translational effects of eIF4A inhibiton and uncovered the G-quadruplex as a novel 5′UTR cis -regulatory element.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Rubio, C. A. et al. Transcriptome-wide characterization of the eIF4A signature highlights plasticity in translation regulation. Genome Biol. 15, 476 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Hsieh, A. C. et al. Cell type-specific abundance of 4EBP1 primes prostate cancer sensitivity or resistance to PI3K pathway inhibitors. Sci. Signal. 8, ra116 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Pourdehnad, M. et al. Myc and mTOR converge on a common node in protein synthesis control that confers synthetic lethality in Myc-driven cancers. Proc. Natl Acad. Sci. USA 110, 11988–11993 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Truitt, M. L. et al. Differential requirements for eIF4E dose in normal development and cancer. Cell 162, 59–71 (2015). This paper describes the generation of the first genetic loss-of-function mouse for eIF4E and unexpectedly reveals that eIF4E dose is limiting for tumorigenesis but not for normal development. Translation profiling further uncovers an eIF4E-dependent oncogenic translation programme enriched for oxidative stress response genes and marked by a novel functional cis -regulatory element termed the CERT.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Leprivier, G. et al. The eEF2 kinase confers resistance to nutrient deprivation by blocking translation elongation. Cell 153, 1064–1079 (2013). This paper shows that cancer cells can reprogramme translation at the elongation step through AMPK–eEF2K signalling in order to adapt to metabolic stress conditions.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Badura, M., Braunstein, S., Zavadil, J. & Schneider, R. J. DNA damage and eIF4G1 in breast cancer cells reprogram translation for survival and DNA repair mRNAs. Proc. Natl Acad. Sci. USA 109, 18767–18772 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Boussemart, L. et al. eIF4F is a nexus of resistance to anti-BRAF and anti-MEK cancer therapies. Nature 513, 105–109 (2014).

    Article  CAS  PubMed  Google Scholar 

  14. Waskiewicz, A. J., Flynn, A., Proud, C. G. & Cooper, J. A. Mitogen-activated protein kinases activate the serine/threonine kinases Mnk1 and Mnk2. EMBO J. 16, 1909–1920 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Robichaud, N. et al. Phosphorylation of eIF4E promotes EMT and metastasis via translational control of SNAIL and MMP-3. Oncogene 34, 2032–2042 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Jones, R. M. et al. An essential E box in the promoter of the gene encoding the mRNA cap-binding protein (eukaryotic initiation factor 4E) is a target for activation by c-myc. Mol. Cell. Biol. 16, 4754–4764 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Rosenwald, I. B., Rhoads, D. B., Callanan, L. D., Isselbacher, K. J. & Schmidt, E. V. Increased expression of eukaryotic translation initiation factors eIF-4E and eIF-2α in response to growth induction by c-myc. Proc. Natl Acad. Sci. USA 90, 6175–6178 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Haghighat, A., Mader, S., Pause, A. & Sonenberg, N. Repression of cap-dependent translation by 4E-binding protein 1: competition with p220 for binding to eukaryotic initiation factor-4E. EMBO J. 14, 5701–5709 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Bah, A. et al. Folding of an intrinsically disordered protein by phosphorylation as a regulatory switch. Nature 519, 106–109 (2015).

    Article  CAS  PubMed  Google Scholar 

  20. She, Q. B. et al. 4E-BP1 is a key effector of the oncogenic activation of the AKT and ERK signaling pathways that integrates their function in tumors. Cancer Cell 18, 39–51 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Dorrello, N. V. et al. S6K1- and βTRCP-mediated degradation of PDCD4 promotes protein translation and cell growth. Science 314, 467–471 (2006).

    Article  CAS  PubMed  Google Scholar 

  22. Raught, B. et al. Phosphorylation of eucaryotic translation initiation factor 4B Ser422 is modulated by S6 kinases. EMBO J. 23, 1761–1769 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Petroulakis, E. et al. p53-dependent translational control of senescence and transformation via 4E-BPs. Cancer Cell 16, 439–446 (2009).

    Article  CAS  PubMed  Google Scholar 

  24. Wang, X. et al. Eukaryotic elongation factor 2 kinase activity is controlled by multiple inputs from oncogenic signaling. Mol. Cell. Biol. 34, 4088–4103 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Vogel, C. & Marcotte, E. M. Insights into the regulation of protein abundance from proteomic and transcriptomic analyses. Nat. Rev. Genet. 13, 227–232 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Hsieh, A. C. et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485, 55–61 (2012). This study, along with reference 117, uses novel ATP active-site inhibitors of mTOR and unbiased genome-wide profiling to define the mTOR-dependent translatome and identifies the 5′TOP and the PRTE as cis -regulatory elements that control translation of key mRNA subsets. It also describes a new role for mTOR–4EBP signalling in specializing translation of the cancer genome to direct tumour invasion and metastasis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Rajasekhar, V. K. et al. Oncogenic Ras and Akt signaling contribute to glioblastoma formation by differential recruitment of existing mRNAs to polysomes. Mol. Cell 12, 889–901 (2003).

    Article  CAS  PubMed  Google Scholar 

  28. Cunningham, J. T., Moreno, M. V., Lodi, A., Ronen, S. M. & Ruggero, D. Protein and nucleotide biosynthesis are coupled by a single rate-limiting enzyme, PRPS2, to drive cancer. Cell 157, 1088–1103 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Silvera, D., Formenti, S. C. & Schneider, R. J. Translational control in cancer. Nat. Rev. Cancer 10, 254–266 (2010).

    Article  CAS  PubMed  Google Scholar 

  30. Bhat, M. et al. Targeting the translation machinery in cancer. Nat. Rev. Drug Discov. 14, 261–278 (2015).

    Article  CAS  PubMed  Google Scholar 

  31. Blagden, S. P. & Willis, A. E. The biological and therapeutic relevance of mRNA translation in cancer. Nat. Rev. Clin. Oncol. 8, 280–291 (2011).

    Article  CAS  PubMed  Google Scholar 

  32. 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 

  33. 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 5′UTR secondary structure in influencing mRNA translation.

    Article  CAS  PubMed  Google Scholar 

  34. Koromilas, A. E., Lazaris-Karatzas, A. & Sonenberg, N. mRNAs containing extensive secondary structure in their 5′ non-coding region translate efficiently in cells overexpressing initiation factor eIF-4E. EMBO J. 11, 4153–4158 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Ray, B. K. et al. ATP-dependent unwinding of messenger RNA structure by eukaryotic initiation factors. J. Biol. Chem. 260, 7651–7658 (1985).

    CAS  PubMed  Google Scholar 

  36. Svitkin, Y. V. et al. The requirement for eukaryotic initiation factor 4A (elF4A) in translation is in direct proportion to the degree of mRNA 5′ secondary structure. RNA 7, 382–394 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Rozen, F. et al. Bidirectional RNA helicase activity of eucaryotic translation initiation factors 4A and 4F. Mol. Cell. Biol. 10, 1134–1144 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Feoktistova, K., Tuvshintogs, E., Do, A. & Fraser, C. S. Human eIF4E promotes mRNA restructuring by stimulating eIF4A helicase activity. Proc. Natl Acad. Sci. USA 110, 13339–13344 (2013). This study describes a new function of eIF4E in stimulating eIF4A helicase activity that occurs independently of eIF4E binding to the 5′ cap.

    Article  PubMed  PubMed Central  Google Scholar 

  39. 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 

  40. 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 

  41. Kevil, C., Carter, P., Hu, B. & DeBenedetti, A. Translational enhancement of FGF-2 by eIF-4 factors, and alternate utilization of CUG and AUG codons for translation initiation. Oncogene 11, 2339–2348 (1995).

    CAS  PubMed  Google Scholar 

  42. Kevil, C. G. et al. Translational regulation of vascular permeability factor by eukaryotic initiation factor 4E: implications for tumor angiogenesis. Int. J. Cancer 65, 785–790 (1996).

    Article  CAS  PubMed  Google Scholar 

  43. 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 

  44. Wiegering, A. et al. Targeting translation initiation bypasses signaling crosstalk mechanisms that maintain high MYC levels in colorectal cancer. Cancer Discov. 5, 768–781 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. 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 

  46. 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 

  47. Gerlitz, G., Jagus, R. & Elroy-Stein, O. Phosphorylation of initiation factor-2α is required for activation of internal translation initiation during cell differentiation. Eur. J. Biochem. 269, 2810–2819 (2002).

    Article  CAS  PubMed  Google Scholar 

  48. Holcik, M., Lefebvre, C., Yeh, C., Chow, T. & Korneluk, R. G. A new internal-ribosome-entry-site motif potentiates XIAP-mediated cytoprotection. Nat. Cell Biol. 1, 190–192 (1999).

    Article  CAS  PubMed  Google Scholar 

  49. Coldwell, M. J., Mitchell, S. A., Stoneley, M., MacFarlane, M. & Willis, A. E. Initiation of Apaf-1 translation by internal ribosome entry. Oncogene 19, 899–905 (2000).

    Article  CAS  PubMed  Google Scholar 

  50. Sherrill, K. W., Byrd, M. P., Van Eden, M. E. & Lloyd, R. E. BCL-2 translation is mediated via internal ribosome entry during cell stress. J. Biol. Chem. 279, 29066–29074 (2004).

    Article  CAS  PubMed  Google Scholar 

  51. Nanbru, C. et al. Alternative translation of the proto-oncogene c-myc by an internal ribosome entry site. J. Biol. Chem. 272, 32061–32066 (1997).

    Article  CAS  PubMed  Google Scholar 

  52. Kullmann, M., Gopfert, U., Siewe, B. & Hengst, L. ELAV/Hu proteins inhibit p27 translation via an IRES element in the p27 5′UTR. Genes Dev. 16, 3087–3099 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Ray, P. S., Grover, R. & Das, S. Two internal ribosome entry sites mediate the translation of p53 isoforms. EMBO Rep. 7, 404–410 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Yang, D. Q., Halaby, M. J. & Zhang, Y. The identification of an internal ribosomal entry site in the 5′-untranslated region of p53 mRNA provides a novel mechanism for the regulation of its translation following DNA damage. Oncogene 25, 4613–4619 (2006).

    Article  CAS  PubMed  Google Scholar 

  55. Graber, T. E. & Holcik, M. Cap-independent regulation of gene expression in apoptosis. Mol. Biosyst. 3, 825–834 (2007).

    Article  CAS  PubMed  Google Scholar 

  56. Wurth, L. & Gebauer, F. RNA-binding proteins, multifaceted translational regulators in cancer. Biochim. Biophys. Acta 1849, 881–886 (2015).

    Article  CAS  PubMed  Google Scholar 

  57. 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 

  58. Mukhopadhyay, R., Jia, J., Arif, A., Ray, P. S. & Fox, P. L. The GAIT system: a gatekeeper of inflammatory gene expression. Trends Biochem. Sci. 34, 324–331 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Chaudhury, A. et al. TGF-β-mediated phosphorylation of hnRNP E1 induces EMT via transcript-selective translational induction of Dab2 and ILEI. Nat. Cell Biol. 12, 286–293 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Hussey, G. S. et al. Identification of an mRNP complex regulating tumorigenesis at the translational elongation step. Mol. Cell 41, 419–431 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Lee, A. S., Kranzusch, P. J. & Cate, J. H. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522, 111–114 (2015). Using transcriptome-wide PAR-CLIP analysis, this study identified a novel role for eIF3 in interacting non-canonically with stem–loops in the 5′UTR of select mRNAs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. 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 

  63. Somers, J., Poyry, 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 

  64. Hopkins, B. D. et al. A secreted PTEN phosphatase that enters cells to alter signaling and survival. Science 341, 399–402 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. 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  PubMed  PubMed Central  Google Scholar 

  66. Ingolia, N. T. et al. Ribosome profiling reveals pervasive translation outside of annotated protein-coding genes. Cell Rep. 8, 1365–1379 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Ingolia, N. T., Lareau, L. F. & Weissman, J. S. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 147, 789–802 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Lee, S. et al. Global mapping of translation initiation sites in mammalian cells at single-nucleotide resolution. Proc. Natl Acad. Sci. USA 109, E2424–E2432 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Kozak, M. An analysis of 5′-noncoding sequences from 699 vertebrate messenger RNAs. Nucleic Acids Res. 15, 8125–8148 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ye, Y. et al. Analysis of human upstream open reading frames and impact on gene expression. Hum. Genet. 134, 605–612 (2015).

    Article  CAS  PubMed  Google Scholar 

  71. 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).

    Article  CAS  PubMed  Google Scholar 

  72. Brown, C. Y., Mize, G. J., Pineda, M., George, D. L. & Morris, D. R. Role of two upstream open reading frames in the translational control of oncogene mdm2. Oncogene 18, 5631–5637 (1999).

    Article  CAS  PubMed  Google Scholar 

  73. Landers, J. E., Cassel, S. L. & George, D. L. Translational enhancement of mdm2 oncogene expression in human tumor cells containing a stabilized wild-type p53 protein. Cancer Res. 57, 3562–3568 (1997).

    CAS  PubMed  Google Scholar 

  74. Mehta, A., Trotta, C. R. & Peltz, S. W. Derepression of the Her-2 uORF is mediated by a novel post-transcriptional control mechanism in cancer cells. Genes Dev. 20, 939–953 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Liu, L. et al. Mutation of the CDKN2A 5′ UTR creates an aberrant initiation codon and predisposes to melanoma. Nat. Genet. 21, 128–132 (1999).

    Article  CAS  PubMed  Google Scholar 

  76. Wethmar, K. et al. C/EBPβΔORF mice — a genetic model for uORF-mediated translational control in mammals. Genes Dev. 24, 15–20 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Gomis, R. R., Alarcon, C., Nadal, C., Van Poznak, C. & Massague, J. C/EBPβ at the core of the TGFβ cytostatic response and its evasion in metastatic breast cancer cells. Cancer Cell 10, 203–214 (2006).

    Article  CAS  PubMed  Google Scholar 

  78. Begay, V. et al. Deregulation of the endogenous C/EBPβ LIP isoform predisposes to tumorigenesis. J. Mol. Med. (Berl.) 93, 39–49 (2015).

    Article  CAS  Google Scholar 

  79. Jia, J., Yao, P., Arif, A. & Fox, P. L. Regulation and dysregulation of 3′UTR-mediated translational control. Curr. Opin. Genet. Dev. 23, 29–34 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Elkon, R., Ugalde, A. P. & Agami, R. Alternative cleavage and polyadenylation: extent, regulation and function. Nat. Rev. Genet. 14, 496–506 (2013).

    Article  CAS  PubMed  Google Scholar 

  81. Tian, B., Hu, J., Zhang, H. & Lutz, C. S. A large-scale analysis of mRNA polyadenylation of human and mouse genes. Nucleic Acids Res. 33, 201–212 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Yan, J. & Marr, T. G. Computational analysis of 3′-ends of ESTs shows four classes of alternative polyadenylation in human, mouse, and rat. Genome Res. 15, 369–375 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Sandberg, R., Neilson, J. R., Sarma, A., Sharp, P. A. & Burge, C. B. Proliferating cells express mRNAs with shortened 3′ untranslated regions and fewer microRNA target sites. Science 320, 1643–1647 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Mayr, C. & Bartel, D. P. Widespread shortening of 3′UTRs by alternative cleavage and polyadenylation activates oncogenes in cancer cells. Cell 138, 673–684 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Iwakawa, H. O. & Tomari, Y. The functions of microRNAs: mRNA decay and translational repression. Trends Cell Biol. 25, 651–665 (2015).

    Article  CAS  PubMed  Google Scholar 

  86. Jonas, S. & Izaurralde, E. Towards a molecular understanding of microRNA-mediated gene silencing. Nat. Rev. Genet. 16, 421–433 (2015).

    Article  CAS  PubMed  Google Scholar 

  87. Bazzini, A. A., Lee, M. T. & Giraldez, A. J. Ribosome profiling shows that miR-430 reduces translation before causing mRNA decay in zebrafish. Science 336, 233–237 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Djuranovic, S., Nahvi, A. & Green, R. miRNA-mediated gene silencing by translational repression followed by mRNA deadenylation and decay. Science 336, 237–240 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Meijer, H. A. et al. Translational repression and eIF4A2 activity are critical for microRNA-mediated gene regulation. Science 340, 82–85 (2013).

    Article  CAS  PubMed  Google Scholar 

  90. Mathonnet, G. et al. MicroRNA inhibition of translation initiation in vitro by targeting the cap-binding complex eIF4F. Science 317, 1764–1767 (2007).

    Article  CAS  PubMed  Google Scholar 

  91. Eichhorn, S. W. et al. mRNA destabilization is the dominant effect of mammalian microRNAs by the time substantial repression ensues. Mol. Cell 56, 104–115 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Bethune, J., Artus-Revel, C. G. & Filipowicz, W. Kinetic analysis reveals successive steps leading to miRNA-mediated silencing in mammalian cells. EMBO Rep. 13, 716–723 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Subtelny, A. O., Eichhorn, S. W., Chen, G. R., Sive, H. & Bartel, D. P. Poly(A)-tail profiling reveals an embryonic switch in translational control. Nature 508, 66–71 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Fukao, A. et al. MicroRNAs trigger dissociation of eIF4AI and eIF4AII from target mRNAs in humans. Mol. Cell 56, 79–89 (2014).

    Article  CAS  PubMed  Google Scholar 

  95. Fukaya, T., Iwakawa, H. O. & Tomari, Y. MicroRNAs block assembly of eIF4F translation initiation complex in Drosophila. Mol. Cell 56, 67–78 (2014).

    Article  CAS  PubMed  Google Scholar 

  96. Thermann, R. & Hentze, M. W. Drosophila miR2 induces pseudo-polysomes and inhibits translation initiation. Nature 447, 875–878 (2007).

    Article  CAS  PubMed  Google Scholar 

  97. Johnson, S. M. et al. RAS is regulated by the let-7 microRNA family. Cell 120, 635–647 (2005).

    Article  CAS  PubMed  Google Scholar 

  98. Lu, J. et al. MicroRNA expression profiles classify human cancers. Nature 435, 834–838 (2005).

    Article  CAS  PubMed  Google Scholar 

  99. Thomson, J. M. et al. Extensive post-transcriptional regulation of microRNAs and its implications for cancer. Genes Dev. 20, 2202–2207 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Novoa, I., Gallego, J., Ferreira, P. G. & Mendez, R. Mitotic cell-cycle progression is regulated by CPEB1 and CPEB4-dependent translational control. Nat. Cell Biol. 12, 447–456 (2010).

    Article  CAS  PubMed  Google Scholar 

  101. Bava, F. A. et al. CPEB1 coordinates alternative 3′-UTR formation with translational regulation. Nature 495, 121–125 (2013). This study identifies a novel function of CPEB1 in regulating 3′UTR shortening through APA in hundreds of mRNAs involved in cellular proliferation, cancer progression and pre-mRNA splicing.

    Article  CAS  PubMed  Google Scholar 

  102. Masamha, C. P. et al. CFIm25 links alternative polyadenylation to glioblastoma tumour suppression. Nature 510, 412–416 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Glisovic, T., Bachorik, J. L., Yong, J. & Dreyfuss, G. RNA-binding proteins and post-transcriptional gene regulation. FEBS Lett. 582, 1977–1986 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Izquierdo, J. M. Hu antigen R (HuR) functions as an alternative pre-mRNA splicing regulator of Fas apoptosis-promoting receptor on exon definition. J. Biol. Chem. 283, 19077–19084 (2008).

    Article  CAS  PubMed  Google Scholar 

  105. Dixon, D. A. et al. Altered expression of the mRNA stability factor HuR promotes cyclooxygenase-2 expression in colon cancer cells. J. Clin. Invest. 108, 1657–1665 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Mazan-Mamczarz, K. et al. RNA-binding protein HuR enhances p53 translation in response to ultraviolet light irradiation. Proc. Natl Acad. Sci. USA 100, 8354–8359 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Eswarappa, S. M. et al. Programmed translational readthrough generates antiangiogenic VEGF-Ax. Cell 157, 1605–1618 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Gerstberger, S., Hafner, M. & Tuschl, T. A census of human RNA-binding proteins. Nat. Rev. Genet. 15, 829–845 (2014).

    Article  CAS  PubMed  Google Scholar 

  109. Kechavarzi, B. & Janga, S. C. Dissecting the expression landscape of RNA-binding proteins in human cancers. Genome Biol. 15, R14 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Park, S. M. et al. Musashi2 sustains the mixed-lineage leukemia-driven stem cell regulatory program. J. Clin. Invest. 125, 1286–1298 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  111. Ito, T. et al. Regulation of myeloid leukaemia by the cell-fate determinant Musashi. Nature 466, 765–768 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Kharas, M. G. et al. Musashi-2 regulates normal hematopoiesis and promotes aggressive myeloid leukemia. Nat. Med. 16, 903–908 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Castello, A. et al. Insights into RNA biology from an atlas of mammalian mRNA-binding proteins. Cell 149, 1393–1406 (2012). This important study was the first to use quantitative MS in combination with UV crosslinking to identify the complete RBP–mRNA interactome, uncovering an unexpected number and diversity of RBPs within the cell.

    Article  CAS  PubMed  Google Scholar 

  114. Kwon, S. C. et al. The RNA-binding protein repertoire of embryonic stem cells. Nat. Struct. Mol. Biol. 20, 1122–1130 (2013).

    Article  CAS  PubMed  Google Scholar 

  115. Baltz, A. G. et al. The mRNA-bound proteome and its global occupancy profile on protein-coding transcripts. Mol. Cell 46, 674–690 (2012).

    Article  CAS  PubMed  Google Scholar 

  116. Preitner, N. et al. APC is an RNA-binding protein, and its interactome provides a link to neural development and microtubule assembly. Cell 158, 368–382 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Thoreen, C. C. et al. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485, 109–113 (2012). This paper, along with reference 26, identifies the 5′TOP and the PRTE as cis -regulatory elements that control translation of key mRNA subsets downstream of mTOR.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Xue, S. et al. RNA regulons in Hox 5′ UTRs confer ribosome specificity to gene regulation. Nature 517, 33–38 (2015). This paper describes the generation of the first targeted knockout of a cellular IRES in mice and demonstrates a key functional role for IRES-driven translation in vivo . It also reveals that IRES-dependent translation is enabled by a new cis -regulatory element termed the TIE, which blocks cap-dependent translation.

    Article  CAS  PubMed  Google Scholar 

  119. Levy, S., Avni, D., Hariharan, N., Perry, R. P. & Meyuhas, O. Oligopyrimidine tract at the 5′ end of mammalian ribosomal protein mRNAs is required for their translational control. Proc. Natl Acad. Sci. USA 88, 3319–3323 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Avni, D., Shama, S., Loreni, F. & Meyuhas, O. Vertebrate mRNAs with a 5′-terminal pyrimidine tract are candidates for translational repression in quiescent cells: characterization of the translational cis-regulatory element. Mol. Cell. Biol. 14, 3822–3833 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Jefferies, H. B., Reinhard, C., Kozma, S. C. & Thomas, G. Rapamycin selectively represses translation of the “polypyrimidine tract” mRNA family. Proc. Natl Acad. Sci. USA 91, 4441–4445 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Pende, M. et al. S6K1−/−/S6K2−/− mice exhibit perinatal lethality and rapamycin-sensitive 5′-terminal oligopyrimidine mRNA translation and reveal a mitogen-activated protein kinase-dependent S6 kinase pathway. Mol. Cell. Biol. 24, 3112–3124 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Ruvinsky, I. et al. Ribosomal protein S6 phosphorylation is a determinant of cell size and glucose homeostasis. Genes Dev. 19, 2199–2211 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Miloslavski, R. et al. Oxygen sufficiency controls TOP mRNA translation via the TSC-Rheb-mTOR pathway in a 4E-BP-independent manner. J. Mol. Cell Biol. 6, 255–266 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Patursky-Polischuk, I. et al. The TSC-mTOR pathway mediates translational activation of TOP mRNAs by insulin largely in a raptor- or rictor-independent manner. Mol. Cell. Biol. 29, 640–649 (2009).

    Article  CAS  PubMed  Google Scholar 

  126. Patursky-Polischuk, I. et al. Reassessment of the role of TSC, mTORC1 and microRNAs in amino acids-meditated translational control of TOP mRNAs. PLoS ONE 9, e109410 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Feldman, M. E. et al. Active-site inhibitors of mTOR target rapamycin-resistant outputs of mTORC1 and mTORC2. PLoS Biol. 7, e38 (2009).

    Article  CAS  PubMed  Google Scholar 

  128. Thoreen, C. C. et al. An ATP-competitive mammalian target of rapamycin inhibitor reveals rapamycin-resistant functions of mTORC1. J. Biol. Chem. 284, 8023–8032 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  129. Sheridan, C. M. et al. YB-1 and MTA1 protein levels and not DNA or mRNA alterations predict for prostate cancer recurrence. Oncotarget 6, 7470–7480 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  130. Tcherkezian, J. et al. Proteomic analysis of cap-dependent translation identifies LARP1 as a key regulator of 5′TOP mRNA translation. Genes Dev. 28, 357–371 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Mura, M. et al. LARP1 post-transcriptionally regulates mTOR and contributes to cancer progression. Oncogene 34, 5025–5036 (2015).

    Article  CAS  PubMed  Google Scholar 

  132. Fonseca, B. D. et al. La-related protein 1 (LARP1) represses terminal oligopyrimidine (TOP) mRNA translation downstream of mTOR complex 1 (mTORC1). J. Biol. Chem. 290, 15996–16020 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Meyer, K. D. et al. Comprehensive analysis of mRNA methylation reveals enrichment in 3′ UTRs and near stop codons. Cell 149, 1635–1646 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Zhou, J. et al. Dynamic m6. Nature 526, 591–594 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Dominissini, D. et al. Topology of the human and mouse m6. Nature 485, 201–206 (2012).

    Article  CAS  PubMed  Google Scholar 

  137. Carlile, T. M. et al. Pseudouridine profiling reveals regulated mRNA pseudouridylation in yeast and human cells. Nature 515, 143–146 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Li, X. et al. Chemical pulldown reveals dynamic pseudouridylation of the mammalian transcriptome. Nat. Chem. Biol. 11, 592–597 (2015).

    Article  CAS  PubMed  Google Scholar 

  139. Lovejoy, A. F., Riordan, D. P. & Brown, P. O. Transcriptome-wide mapping of pseudouridines: pseudouridine synthases modify specific mRNAs in S. cerevisiae. PLoS ONE 9, e110799 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Schwartz, S. et al. Transcriptome-wide mapping reveals widespread dynamic-regulated pseudouridylation of ncRNA and mRNA. Cell 159, 148–162 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Pavon-Eternod, M. et al. tRNA over-expression in breast cancer and functional consequences. Nucleic Acids Res. 37, 7268–7280 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Winter, A. G. et al. RNA polymerase III transcription factor TFIIIC2 is overexpressed in ovarian tumors. Proc. Natl Acad. Sci. USA 97, 12619–12624 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Zhou, Y., Goodenbour, J. M., Godley, L. A., Wickrema, A. & Pan, T. High levels of tRNA abundance and alteration of tRNA charging by bortezomib in multiple myeloma. Biochem. Biophys. Res. Commun. 385, 160–164 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Pavon-Eternod, M., Gomes, S., Rosner, M. R. & Pan, T. Overexpression of initiator methionine tRNA leads to global reprogramming of tRNA expression and increased proliferation in human epithelial cells. RNA 19, 461–466 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Gomez-Roman, N., Grandori, C., Eisenman, R. N. & White, R. J. Direct activation of RNA polymerase III transcription by c-Myc. Nature 421, 290–294 (2003).

    Article  CAS  PubMed  Google Scholar 

  146. Felton-Edkins, Z. A. et al. The mitogen-activated protein (MAP) kinase ERK induces tRNA synthesis by phosphorylating TFIIIB. EMBO J. 22, 2422–2432 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Wei, Y., Tsang, C. K. & Zheng, X. F. Mechanisms of regulation of RNA polymerase III-dependent transcription by TORC1. EMBO J. 28, 2220–2230 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Kantidakis, T., Ramsbottom, B. A., Birch, J. L., Dowding, S. N. & White, R. J. mTOR associates with TFIIIC, is found at tRNA and 5S rRNA genes, and targets their repressor Maf1. Proc. Natl Acad. Sci. USA 107, 11823–11828 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  149. Lampson, B. L. et al. Rare codons regulate KRas oncogenesis. Curr. Biol. 23, 70–75 (2013).

    Article  CAS  PubMed  Google Scholar 

  150. Dittmar, K. A., Goodenbour, J. M. & Pan, T. Tissue-specific differences in human transfer RNA expression. PLoS Genet. 2, e221 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Plotkin, J. B., Robins, H. & Levine, A. J. Tissue-specific codon usage and the expression of human genes. Proc. Natl Acad. Sci. USA 101, 12588–12591 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Gingold, H. et al. A dual program for translation regulation in cellular proliferation and differentiation. Cell 158, 1281–1292 (2014). This is a key study showing a link between tRNA expression patterns and codon usage bias in gene expression programmes that support cancer cell behaviours.

    Article  CAS  PubMed  Google Scholar 

  153. Phizicky, E. M. & Hopper, A. K. tRNA biology charges to the front. Genes Dev. 24, 1832–1860 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  154. Spinola, M. et al. Identification and functional characterization of the candidate tumor suppressor gene TRIT1 in human lung cancer. Oncogene 24, 5502–5509 (2005).

    Article  CAS  PubMed  Google Scholar 

  155. Begley, U. et al. Trm9-catalyzed tRNA modifications link translation to the DNA damage response. Mol. Cell 28, 860–870 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Chan, C. T. et al. Reprogramming of tRNA modifications controls the oxidative stress response by codon-biased translation of proteins. Nat. Commun. 3, 937 (2012).

    Article  CAS  PubMed  Google Scholar 

  157. Van den Born, E. et al. ALKBH8-mediated formation of a novel diastereomeric pair of wobble nucleosides in mammalian tRNA. Nat. Commun. 2, 172 (2011).

    Article  CAS  PubMed  Google Scholar 

  158. Songe-Moller, L. et al. Mammalian ALKBH8 possesses tRNA methyltransferase activity required for the biogenesis of multiple wobble uridine modifications implicated in translational decoding. Mol. Cell. Biol. 30, 1814–1827 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Fu, D. et al. Human AlkB homolog ABH8 Is a tRNA methyltransferase required for wobble uridine modification and DNA damage survival. Mol. Cell. Biol. 30, 2449–2459 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Kryukov, G. V. et al. Characterization of mammalian selenoproteomes. Science 300, 1439–1443 (2003).

    Article  CAS  PubMed  Google Scholar 

  161. Endres, L. et al. Alkbh8 regulates selenocysteine-protein expression to protect against reactive oxygen species damage. PLoS ONE 10, e0131335 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Cozen, A. E. et al. ARM-seq: AlkB-facilitated RNA methylation sequencing reveals a complex landscape of modified tRNA fragments. Nat. Methods 12, 879–884 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Luo, J., Solimini, N. L. & Elledge, S. J. Principles of cancer therapy: oncogene and non-oncogene addiction. Cell 136, 823–837 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Gaillard, H., Garcia-Muse, T. & Aguilera, A. Replication stress and cancer. Nat. Rev. Cancer 15, 276–289 (2015).

    Article  CAS  PubMed  Google Scholar 

  165. Gorrini, C., Harris, I. S. & Mak, T. W. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 12, 931–947 (2013).

    Article  CAS  PubMed  Google Scholar 

  166. Sabharwal, S. S. & Schumacker, P. T. Mitochondrial ROS in cancer: initiators, amplifiers or an Achilles' heel? Nat. Rev. Cancer 14, 709–721 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Leprivier, G., Rotblat, B., Khan, D., Jan, E. & Sorensen, P. H. Stress-mediated translational control in cancer cells. Biochim. Biophys. Acta 1849, 845–860 (2015).

    Article  CAS  PubMed  Google Scholar 

  168. Zoncu, R., Efeyan, A. & Sabatini, D. M. mTOR: from growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 12, 21–35 (2011).

    Article  CAS  PubMed  Google Scholar 

  169. Brugarolas, J. et al. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 18, 2893–2904 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Sancak, Y. et al. The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320, 1496–1501 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Walter, P. & Ron, D. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334, 1081–1086 (2011).

    Article  CAS  PubMed  Google Scholar 

  172. Dong, J., Qiu, H., Garcia-Barrio, M., Anderson, J. & Hinnebusch, A. G. Uncharged tRNA activates GCN2 by displacing the protein kinase moiety from a bipartite tRNA-binding domain. Mol. Cell 6, 269–279 (2000).

    Article  CAS  PubMed  Google Scholar 

  173. Williams, B. R. PKR; a sentinel kinase for cellular stress. Oncogene 18, 6112–6120 (1999).

    Article  CAS  PubMed  Google Scholar 

  174. Chen, J. J. Regulation of protein synthesis by the heme-regulated eIF2α kinase: relevance to anemias. Blood 109, 2693–2699 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  175. Braunstein, S. et al. A hypoxia-controlled cap-dependent to cap-independent translation switch in breast cancer. Mol. Cell 28, 501–512 (2007). This study identifies a hypoxia-induced switch from cap-dependent to IRES-dependent translation that supports tumorigenesis by driving the expression of key pro-angiogenic and pro-survival mRNAs.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  177. Gu, L. et al. Regulation of XIAP translation and induction by MDM2 following irradiation. Cancer Cell 15, 363–375 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. 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 

  179. Shi, Y. et al. Therapeutic potential of targeting IRES-dependent c-myc translation in multiple myeloma cells during ER stress. Oncogene 35, 1015–1024 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Andreev, D. E. et al. Translation of 5′ leaders is pervasive in genes resistant to eIF2 repression. eLife 4, e03971 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  181. Somers, J. et al. A common polymorphism in the 5′ UTR of ERCC5 creates an upstream ORF that confers resistance to platinum-based chemotherapy. Genes Dev. 29, 1891–1896 (2015). This article describes a common polymorphic variant that creates a new uORF in the 5′UTR of the DNA repair enzyme ERCC5 and promotes tumour resistance to platinum-based chemotherapeutics.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Vattem, K. M. & Wek, R. C. Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells. Proc. Natl Acad. Sci. USA 101, 11269–11274 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Harding, H. P. et al. An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Mol. Cell 11, 619–633 (2003).

    Article  CAS  PubMed  Google Scholar 

  184. Pola, C., Formenti, S. C. & Schneider, R. J. Vitronectin-αvβ3 integrin engagement directs hypoxia-resistant mTOR activity and sustained protein synthesis linked to invasion by breast cancer cells. Cancer Res. 73, 4571–4578 (2013).

    Article  CAS  PubMed  Google Scholar 

  185. Connolly, E., Braunstein, S., Formenti, S. & Schneider, R. J. Hypoxia inhibits protein synthesis through a 4E-BP1 and elongation factor 2 kinase pathway controlled by mTOR and uncoupled in breast cancer cells. Mol. Cell. Biol. 26, 3955–3965 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Hart, L. S. et al. ER stress-mediated autophagy promotes Myc-dependent transformation and tumor growth. J. Clin. Invest. 122, 4621–4634 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Cho, P. F. et al. A new paradigm for translational control: inhibition via 5′-3′ mRNA tethering by Bicoid and the eIF4E cognate 4EHP. Cell 121, 411–423 (2005).

    Article  CAS  PubMed  Google Scholar 

  188. Morita, M. et al. A novel 4EHP-GIGYF2 translational repressor complex is essential for mammalian development. Mol. Cell. Biol. 32, 3585–3593 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. von Stechow, L. et al. The E3 ubiquitin ligase ARIH1 protects against genotoxic stress by initiating a 4EHP-mediated mRNA translation arrest. Mol. Cell. Biol. 35, 1254–1268 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Yi, T., Papadopoulos, E., Hagner, P. R. & Wagner, G. Hypoxia-inducible factor-1α (HIF-1α) promotes cap-dependent translation of selective mRNAs through up-regulating initiation factor eIF4E1 in breast cancer cells under hypoxia conditions. J. Biol. Chem. 288, 18732–18742 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Uniacke, J. et al. An oxygen-regulated switch in the protein synthesis machinery. Nature 486, 126–129 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Avdulov, S. et al. eIF4E threshold levels differ in governing normal and neoplastic expansion of mammary stem and luminal progenitor cells. Cancer Res. 75, 687–697 (2015).

    Article  CAS  PubMed  Google Scholar 

  193. Martinez, A. et al. Phosphorylation of eIF4E confers resistance to cellular stress and DNA-damaging agents through an interaction with 4E-T: a rationale for novel therapeutic approaches. PLoS ONE 10, e0123352 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Kenney, J. W., Moore, C. E., Wang, X. & Proud, C. G. Eukaryotic elongation factor 2 kinase, an unusual enzyme with multiple roles. Adv. Biol. Regul. 55, 15–27 (2014).

    Article  CAS  PubMed  Google Scholar 

  195. Ilic, N., Utermark, T., Widlund, H. R. & Roberts, T. M. PI3K-targeted therapy can be evaded by gene amplification along the MYC-eukaryotic translation initiation factor 4E (eIF4E) axis. Proc. Natl Acad. Sci. USA 108, E699–E708 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  196. Graff, J. R. et al. Therapeutic suppression of translation initiation factor eIF4E expression reduces tumor growth without toxicity. J. Clin. Invest. 117, 2638–2648 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Drygin, D. et al. Targeting RNA polymerase I with an oral small molecule CX-5461 inhibits ribosomal RNA synthesis and solid tumor growth. Cancer Res. 71, 1418–1430 (2011).

    Article  CAS  PubMed  Google Scholar 

  198. Bywater, M. J. et al. Inhibition of RNA polymerase I as a therapeutic strategy to promote cancer-specific activation of p53. Cancer Cell 22, 51–65 (2012). This surprising study showed that therapeutic targeting of ribosome biogenesis through the inhibition of RNA polymerase I could selectively kill cancer cells in vivo.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. [No authors listed.] CX-5461 inhibits RNA Pol I in blood cancers. Cancer Discov. 4, OF5 (2014).

  200. Duncan, R., Milburn, S. C. & Hershey, J. W. Regulated phosphorylation and low abundance of HeLa cell initiation factor eIF-4F suggest a role in translational control. Heat shock effects on eIF-4F. J. Biol. Chem. 262, 380–388 (1987).

    CAS  PubMed  Google Scholar 

  201. Hiremath, L. S., Webb, N. R. & Rhoads, R. E. Immunological detection of the messenger RNA cap-binding protein. J. Biol. Chem. 260, 7843–7849 (1985).

    CAS  PubMed  Google Scholar 

  202. Lazaris-Karatzas, A., Montine, K. S. & Sonenberg, N. Malignant transformation by a eukaryotic initiation factor subunit that binds to mRNA 5′ cap. Nature 345, 544–547 (1990). This manuscript provides the first functional evidence that eIF4E overexpression can be oncogenic and demonstrates that deregulated translation can act as a primary driver of tumorigenesis.

    Article  CAS  PubMed  Google Scholar 

  203. Lazaris-Karatzas, A. & Sonenberg, N. The mRNA 5′ cap-binding protein, eIF-4E, cooperates with v-myc or E1A in the transformation of primary rodent fibroblasts. Mol. Cell. Biol. 12, 1234–1238 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Ruggero, D. et al. The translation factor eIF-4E promotes tumor formation and cooperates with c-Myc in lymphomagenesis. Nat. Med. 10, 484–486 (2004). This study is the first demonstration that eIF4E overexpression is sufficient to drive spontaneous tumorigenesis in vivo.

    Article  CAS  PubMed  Google Scholar 

  205. Ryazanov, A. G., Shestakova, E. A. & Natapov, P. G. Phosphorylation of elongation factor 2 by EF-2 kinase affects rate of translation. Nature 334, 170–173 (1988).

    Article  CAS  PubMed  Google Scholar 

  206. Cheng, E. H., Gorelick, F. S., Czernik, A. J., Bagaglio, D. M. & Hait, W. N. Calmodulin-dependent protein kinases in rat glioblastoma. Cell Growth Differ. 6, 615–621 (1995).

    CAS  PubMed  Google Scholar 

  207. Parmer, T. G. et al. Activity and regulation by growth factors of calmodulin-dependent protein kinase III (elongation factor 2-kinase) in human breast cancer. Br. J. Cancer 79, 59–64 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Stumpf, C. R. & Ruggero, D. The cancerous translation apparatus. Curr. Opin. Genet. Dev. 21, 474–483 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Boon, K. et al. N-myc enhances the expression of a large set of genes functioning in ribosome biogenesis and protein synthesis. EMBO J. 20, 1383–1393 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Kim, S., Li, Q., Dang, C. V. & Lee, L. A. Induction of ribosomal genes and hepatocyte hypertrophy by adenovirus-mediated expression of c-Myc in vivo. Proc. Natl Acad. Sci. USA 97, 11198–11202 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Grandori, C. et al. c-Myc binds to human ribosomal DNA and stimulates transcription of rRNA genes by RNA polymerase I. Nat. Cell Biol. 7, 311–318 (2005).

    Article  CAS  PubMed  Google Scholar 

  212. Arabi, A. et al. c-Myc associates with ribosomal DNA and activates RNA polymerase I transcription. Nat. Cell Biol. 7, 303–310 (2005).

    Article  CAS  PubMed  Google Scholar 

  213. Cairns, C. A. & White, R. J. p53 is a general repressor of RNA polymerase III transcription. EMBO J. 17, 3112–3123 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. White, R. J., Trouche, D., Martin, K., Jackson, S. P. & Kouzarides, T. Repression of RNA polymerase III transcription by the retinoblastoma protein. Nature 382, 88–90 (1996).

    Article  CAS  PubMed  Google Scholar 

  215. Hannan, K. M. et al. mTOR-dependent regulation of ribosomal gene transcription requires S6K1 and is mediated by phosphorylation of the carboxy-terminal activation domain of the nucleolar transcription factor UBF. Mol. Cell. Biol. 23, 8862–8877 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Mayer, C., Zhao, J., Yuan, X. & Grummt, I. mTOR-dependent activation of the transcription factor TIF-IA links rRNA synthesis to nutrient availability. Genes Dev. 18, 423–434 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Stefanovsky, V. Y. et al. An immediate response of ribosomal transcription to growth factor stimulation in mammals is mediated by ERK phosphorylation of UBF. Mol. Cell 8, 1063–1073 (2001).

    Article  CAS  PubMed  Google Scholar 

  218. Zhao, J., Yuan, X., Frodin, M. & Grummt, I. ERK-dependent phosphorylation of the transcription initiation factor TIF-IA is required for RNA polymerase I transcription and cell growth. Mol. Cell 11, 405–413 (2003).

    Article  CAS  PubMed  Google Scholar 

  219. Xue, S. & Barna, M. Specialized ribosomes: a new frontier in gene regulation and organismal biology. Nat. Rev. Mol. Cell Biol. 13, 355–369 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. 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 

  221. Marcel, V. et al. p53 acts as a safeguard of translational control by regulating fibrillarin and rRNA methylation in cancer. Cancer Cell 24, 318–330 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Ruggero, D. et al. Dyskeratosis congenita and cancer in mice deficient in ribosomal RNA modification. Science 299, 259–262 (2003).

    Article  CAS  PubMed  Google Scholar 

  223. Yoon, A. et al. Impaired control of IRES-mediated translation in X-linked dyskeratosis congenita. Science 312, 902–906 (2006).

    Article  CAS  PubMed  Google Scholar 

  224. Bellodi, C., Kopmar, N. & Ruggero, D. Deregulation of oncogene-induced senescence and p53 translational control in X-linked dyskeratosis congenita. EMBO J. 29, 1865–1876 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Bellodi, C. et al. Loss of function of the tumor suppressor DKC1 perturbs p27 translation control and contributes to pituitary tumorigenesis. Cancer Res. 70, 6026–6035 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Bellodi, C. et al. H/ACA small RNA dysfunctions in disease reveal key roles for noncoding RNA modifications in hematopoietic stem cell differentiation. Cell Rep. 3, 1493–1502 (2013).

    Article  CAS  PubMed  Google Scholar 

  227. Gao, L. et al. Genome-wide small nucleolar RNA expression analysis of lung cancer by next-generation deep sequencing. Int. J. Cancer 136, E623–E629 (2015).

    Article  CAS  PubMed  Google Scholar 

  228. Jha, P. et al. Genome-wide small noncoding RNA profiling of pediatric high-grade gliomas reveals deregulation of several miRNAs, identifies downregulation of snoRNA cluster HBII-52 and delineates H3F3A and TP53 mutant-specific miRNAs and snoRNAs. Int. J. Cancer 137, 2343–2353 (2015).

    Article  CAS  PubMed  Google Scholar 

  229. Martens-Uzunova, E. S. et al. Diagnostic and prognostic signatures from the small non-coding RNA transcriptome in prostate cancer. Oncogene 31, 978–991 (2012).

    Article  CAS  PubMed  Google Scholar 

  230. Ravo, M. et al. Small non-coding RNA deregulation in endometrial carcinogenesis. Oncotarget 6, 4677–4691 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  231. Ronchetti, D. et al. Small nucleolar RNAs as new biomarkers in chronic lymphocytic leukemia. BMC Med. Genom. 6, 27 (2013).

    Article  Google Scholar 

  232. Wang, X. et al. Regulation of elongation factor 2 kinase by p90(RSK1) and p70 S6 kinase. EMBO J. 20, 4370–4379 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Browne, G. J. & Proud, C. G. A novel mTOR-regulated phosphorylation site in elongation factor 2 kinase modulates the activity of the kinase and its binding to calmodulin. Mol. Cell. Biol. 24, 2986–2997 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Johannes, G., Carter, M. S., Eisen, M. B., Brown, P. O. & Sarnow, P. Identification of eukaryotic mRNAs that are translated at reduced cap binding complex eIF4F concentrations using a cDNA microarray. Proc. Natl Acad. Sci. USA 96, 13118–13123 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223 (2009). This article provides the first description of ribosome profiling, a technology that has provided unparalleled insights into translational control of the cancer genome.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Jan, C. H., Williams, C. C. & Weissman, J. S. Principles of ER cotranslational translocation revealed by proximity-specific ribosome profiling. Science 346, 1257521 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Licatalosi, D. D. et al. HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature 456, 464–469 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Hafner, M. et al. Transcriptome-wide identification of RNA-binding protein and microRNA target sites by PAR-CLIP. Cell 141, 129–141 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Konig, J. et al. iCLIP reveals the function of hnRNP particles in splicing at individual nucleotide resolution. Nat. Struct. Mol. Biol. 17, 909–915 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Chu, C. et al. Systematic discovery of Xist RNA binding proteins. Cell 161, 404–416 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. McHugh, C. A. et al. The Xist lncRNA interacts directly with SHARP to silence transcription through HDAC3. Nature 521, 232–236 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Kudla, G., Granneman, S., Hahn, D., Beggs, J. D. & Tollervey, D. Cross-linking, ligation, and sequencing of hybrids reveals RNA–RNA interactions in yeast. Proc. Natl Acad. Sci. USA 108, 10010–10015 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  244. Underwood, J. G. et al. FragSeq: transcriptome-wide RNA structure probing using high-throughput sequencing. Nat. Methods 7, 995–1001 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Spitale, R. C. et al. RNA SHAPE analysis in living cells. Nat. Chem. Biol. 9, 18–20 (2013).

    Article  CAS  PubMed  Google Scholar 

  246. Lucks, J. B. et al. Multiplexed RNA structure characterization with selective 2′-hydroxyl acylation analyzed by primer extension sequencing (SHAPE-Seq). Proc. Natl Acad. Sci. USA 108, 11063–11068 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  247. Spitale, R. C. et al. Structural imprints in vivo decode RNA regulatory mechanisms. Nature 519, 486–490 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Sugimoto, Y. et al. hiCLIP reveals the in vivo atlas of mRNA secondary structures recognized by Staufen 1. Nature 519, 491–494 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Khoddami, V. & Cairns, B. R. Identification of direct targets and modified bases of RNA cytosine methyltransferases. Nat. Biotechnol. 31, 458–464 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Hussain, S. et al. NSun2-mediated cytosine-5 methylation of vault noncoding RNA determines its processing into regulatory small RNAs. Cell Rep. 4, 255–261 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Squires, J. E. et al. Widespread occurrence of 5-methylcytosine in human coding and non-coding RNA. Nucleic Acids Res. 40, 5023–5033 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  252. Chan, C. T. et al. A quantitative systems approach reveals dynamic control of tRNA modifications during cellular stress. PLoS Genet. 6, e1001247 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Basanta-Sanchez, M., Temple, S., Ansari, S. A., D'Amico, A. & Agris, P. F. Attomole quantification and global profile of RNA modifications: epitranscriptome of human neural stem cells. Nucleic Acids Res. 44, e26 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. Zheng, G. et al. Efficient and quantitative high-throughput tRNA sequencing. Nat. Methods 12, 835–837 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank M. Barna for critical discussion and reading of the manuscript. This work was supported by funding from the US National Institutes of Health (R01 CA140456, R01 CA184624, R01 DK098057, R01 CA154916, RO1 HL119439 and PO1 CA165997). The authors apologize to those whose work could not be mentioned owing to space limitations. D.R. is a Leukaemia and Lymphoma Society Scholar.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Davide Ruggero.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

mTOR

A serine/threonine kinase that forms two distinct molecular complexes (mTORC1 and mTORC2) and acts as a master regulator of protein synthesis, largely through mTORC1-dependent phosphorylation of eukaryotic translation initiation factor 4E-binding proteins (4EBPs) and ribosomal protein S6 kinase.

tRNA

Non-coding RNA with a unique L-shaped tertiary structure that contains a 'charged' amino acid covalently linked to the tRNA by aminoacyl transferases at one end and the tRNA anticodon loop that recognizes distinct mRNA codons through base-pairing interactions at the other end.

Internal ribosome entry site

(IRES). A complex structural element first discovered and characterized in viruses that facilitates translation initiation by recruiting the 40S ribosome to the mRNA in a cap-independent manner frequently aided by specific RNA-binding proteins known as IRES trans-acting factors (ITAFs).

Messenger ribonucleoprotein (mRNP) complex

A complex of mRNA and RNA-binding proteins that can vary throughout the life of the mRNA and act to either promote or inhibit mRNA splicing, stability and translation depending on the nature of the bound proteins.

Upstream open reading frame

(uORF). An ORF comprising a start codon upstream of the primary ORF with an in-frame stop codon that can lie either 5′ or 3′ of the primary start codon. uORFs typically block translation of downstream ORFs, as ribosomes generally initiate translation at the first start codon they encounter and then disassociate from the mRNA upon translation termination.

Alternative cleavage and polyadenylation

(APA). A mechanism for generating mRNAs with different lengths of 3′ untranslated region that relies on the recognition of alternative poly(A) signals in association with U- and GU-rich downstream sequence elements by the cleavage and polyadenylation specificity factor (CPSF) complex and the cleavage stimulating factor (CSTF) complex.

Programmed translational readthrough

(PTR). A phenomenon that enables translation to continue past the normal stop codon in favour of termination at a downstream stop codon, potentially generating an elongated protein with a novel 3′ extension.

5′-terminal oligopyrimidine tract

(5′TOP). A sequence-specific element located at the +1 position of the 5′ untranslated region that consists of a 5′ cytosine residue followed by a stretch of 7–13 pyrimidine nucleotides.

Pyrimidine-rich translational element

(PRTE). A 5′ untranslated region (5′UTR) sequence-specific motif related to the 5′-terminal oligopyrimidine tract that is not restricted to the +1 position of the 5′UTR and contains an invariant uridine at position 6 flanked by pyrimidines.

Ribosomopathies

Inherited human cancer susceptibility disorders — such as cartilage–hair hypoplasia syndrome, Shwachman–Diamond syndrome, 5q deletion syndrome, dyskeratosis congenita and Diamond–Blackfan anaemia — that are characterized by mutations in distinct components of the translational machinery, including enzymes involved in the modification and processing of rRNA, ribosome assembly factors and ribosomal proteins.

Methionine initiator tRNA

(tRNAiMet). A distinct species of tRNA charged with methionine that interacts with the ribosome P site as part of the eIF2–GTP–Met-tRNAiMet ternary complex that initiates translation at start codons.

Wobble position

The first position in a tRNA anticodon loop that can enable non-Watson–Crick base interactions with the third nucleotide in the codon triplet and be a site for tRNA modifications.

Unfolded protein response

(UPR). A conserved pathway activated by the accumulation of unfolded proteins in the endoplasmic reticulum (ER) that signals through three downstream effector arms (PRKR-like ER kinase (PERK), inositol-requiring enzyme 1α (IRE1α) and activating transcription factor 6 (ATF6)) to either relieve protein misfolding stress in the ER through increased production of chaperone proteins, enhanced protein degradation and dampened protein synthesis or commit the cell to apoptosis if stress is unresolvable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Truitt, M., Ruggero, D. New frontiers in translational control of the cancer genome. Nat Rev Cancer 16, 288–304 (2016). https://doi.org/10.1038/nrc.2016.27

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc.2016.27

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer