1932

Abstract

As a convergent mechanism downstream of most oncogenic signals, control of mRNA translation has emerged as a key driver in establishing and tuning gene expression at specific steps in cancer development. Translation control is the most energetically expensive molecular process in the cell that needs to be modulated upon adaption to limited cellular resources, such as cellular stress. It thereby serves as the Achilles’ heel for cancer cells, particularly in response to changes in the microenvironment as well as to nutrient and metabolic shifts characteristic of cancer cell growth and metastasis. In this review, we discuss emerging discoveries that reveal how cancer cells modulate the translation machinery to adapt to oncogenic stress, the mechanisms that guide mRNA translation specificity in cancer, and how this selective mode of gene regulation provides advantages for cancer progression. We also provide an overview of promising preclinical and clinical efforts aimed at targeting the unique vulnerabilities of cancer cells that rely on the remodeling of mRNA translation for their infinite growth and survival.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-cancerbio-030419-033420
2020-03-04
2024-04-24
Loading full text...

Full text loading...

/deliver/fulltext/cancerbio/4/1/annurev-cancerbio-030419-033420.html?itemId=/content/journals/10.1146/annurev-cancerbio-030419-033420&mimeType=html&fmt=ahah

Literature Cited

  1. Anand N, Murthy S, Amann G, Wernick M, Porter LA et al. 2002. Protein elongation factor EEF1A2 is a putative oncogene in ovarian cancer. Nat. Genet. 31:301–5
    [Google Scholar]
  2. Barbieri I, Tzelepis K, Pandolfini L, Shi J, Millan-Zambrano G et al. 2017. Promoter-bound METTL3 maintains myeloid leukaemia by m6A-dependent translation control. Nature 552:126–31
    [Google Scholar]
  3. Barna M, Pusic A, Zollo O, Costa M, Kondrashov N et al. 2008. Suppression of Myc oncogenic activity by ribosomal protein haploinsufficiency. Nature 456:971–75
    [Google Scholar]
  4. Bauer C, Diesinger I, Brass N, Steinhart H, Iro H, Meese EU 2001. Translation initiation factor eIF-4G is immunogenic, overexpressed, and amplified in patients with squamous cell lung carcinoma. Cancer 92:822–29
    [Google Scholar]
  5. Begley U, Dyavaiah M, Patil A, Rooney JP, DiRenzo D et al. 2007. Trm9-catalyzed tRNA modifications link translation to the DNA damage response. Mol. Cell 28:860–70
    [Google Scholar]
  6. Bellodi C, Kopmar N, Ruggero D 2010. Deregulation of oncogene-induced senescence and p53 translational control in X-linked dyskeratosis congenita. EMBO J 29:1865–76
    [Google Scholar]
  7. Benjamin D, Colombi M, Moroni C, Hall MN 2011. Rapamycin passes the torch: a new generation of mTOR inhibitors. Nat. Rev. Drug Discov. 10:868–80
    [Google Scholar]
  8. Boon K, Caron HN, van Asperen R, Valentijn L, Hermus MC et al. 2001. N-myc enhances the expression of a large set of genes functioning in ribosome biogenesis and protein synthesis. EMBO J 20:1383–93
    [Google Scholar]
  9. Bordeleau ME, Robert F, Gerard B, Lindqvist L, Chen SM et al. 2008. Therapeutic suppression of translation initiation modulates chemosensitivity in a mouse lymphoma model. J. Clin. Investig. 118:2651–60
    [Google Scholar]
  10. Braunstein S, Karpisheva K, Pola C, Goldberg J, Hochman T et al. 2007. A hypoxia-controlled cap-dependent to cap-independent translation switch in breast cancer. Mol. Cell 28:501–12
    [Google Scholar]
  11. Buchan JR, Parker R. 2009. Eukaryotic stress granules: the ins and outs of translation. Mol. Cell 36:932–41
    [Google Scholar]
  12. Buttgereit F, Brand MD. 1995. A hierarchy of ATP-consuming processes in mammalian cells. Biochem. J. 312:Pt. 1163–67
    [Google Scholar]
  13. Carelli JD, Sethofer SG, Smith GA, Miller HR, Simard JL et al. 2015. Ternatin and improved synthetic variants kill cancer cells by targeting the elongation factor-1A ternary complex. eLife 4:e10222
    [Google Scholar]
  14. Cerezo M, Guemiri R, Druillennec S, Girault I, Malka-Mahieu H et al. 2018. Translational control of tumor immune escape via the eIF4F-STAT1-PD-L1 axis in melanoma. Nat. Med. 24:1877–86
    [Google Scholar]
  15. Chan CT, Pang YL, Deng W, Babu IR, Dyavaiah M et al. 2012. Reprogramming of tRNA modifications controls the oxidative stress response by codon-biased translation of proteins. Nat. Commun. 3:937
    [Google Scholar]
  16. Chen L, Aktas BH, Wang Y, He X, Sahoo R et al. 2012. Tumor suppression by small molecule inhibitors of translation initiation. Oncotarget 3:869–81
    [Google Scholar]
  17. Chen T, Ozel D, Qiao Y, Harbinski F, Chen L et al. 2011. Chemical genetics identify eIF2α kinase heme-regulated inhibitor as an anticancer target. Nat. Chem. Biol. 7:610–16
    [Google Scholar]
  18. Choe J, Lin S, Zhang W, Liu Q, Wang L et al. 2018. mRNA circularization by METTL3-eIF3h enhances translation and promotes oncogenesis. Nature 561:556–60
    [Google Scholar]
  19. Cole MD, Cowling VH. 2009. Specific regulation of mRNA cap methylation by the c-Myc and E2F1 transcription factors. Oncogene 28:1169–75
    [Google Scholar]
  20. Cunningham JT, Moreno MV, Lodi A, Ronen SM, Ruggero D 2014. Protein and nucleotide biosynthesis are coupled by a single rate-limiting enzyme, PRPS2, to drive cancer. Cell 157:1088–103
    [Google Scholar]
  21. Dave B, Granados-Principal S, Zhu R, Benz S, Rabizadeh S et al. 2014. Targeting RPL39 and MLF2 reduces tumor initiation and metastasis in breast cancer by inhibiting nitric oxide synthase signaling. PNAS 111:8838–43
    [Google Scholar]
  22. De Benedetti A, Graff JR 2004. eIF-4E expression and its role in malignancies and metastases. Oncogene 23:3189–99
    [Google Scholar]
  23. De Gassart A, Demaria O, Panes R, Zaffalon L, Ryazanov AG et al. 2016. Pharmacological eEF2K activation promotes cell death and inhibits cancer progression. EMBO Rep 17:1471–84
    [Google Scholar]
  24. De Keersmaecker K, Atak ZK, Li N, Vicente C, Patchett S et al. 2013. Exome sequencing identifies mutation in CNOT3 and ribosomal genes RPL5 and RPL10 in T-cell acute lymphoblastic leukemia. Nat. Genet. 45:186–90
    [Google Scholar]
  25. Delaunay S, Rapino F, Tharun L, Zhou Z, Heukamp L et al. 2016. Elp3 links tRNA modification to IRES-dependent translation of LEF1 to sustain metastasis in breast cancer. J. Exp. Med. 213:2503–23
    [Google Scholar]
  26. Draptchinskaia N, Gustavsson P, Andersson B, Pettersson M, Willig TN et al. 1999. The gene encoding ribosomal protein S19 is mutated in Diamond-Blackfan anaemia. Nat. Genet. 21:169–75
    [Google Scholar]
  27. Drygin D, Siddiqui-Jain A, O'Brien S, Schwaebe M, Lin A et al. 2009. Anticancer activity of CX-3543: a direct inhibitor of rRNA biogenesis. Cancer Res 69:7653–61
    [Google Scholar]
  28. Duffy AG, Makarova-Rusher OV, Ulahannan SV, Rahma OE, Fioravanti S et al. 2016. Modulation of tumor eIF4E by antisense inhibition: a phase I/II translational clinical trial of ISIS 183750—an antisense oligonucleotide against eIF4E—in combination with irinotecan in solid tumors and irinotecan-refractory colorectal cancer. Int. J. Cancer 139:1648–57
    [Google Scholar]
  29. Eberle J, Krasagakis K, Orfanos CE 1997. Translation initiation factor eIF-4A1 mRNA is consistently overexpressed in human melanoma cells in vitro. Int. J. Cancer 71:396–401
    [Google Scholar]
  30. Ebert BL, Pretz J, Bosco J, Chang CY, Tamayo P et al. 2008. Identification of RPS14 as a 5q syndrome gene by RNA interference screen. Nature 451:335–39
    [Google Scholar]
  31. Feng YX, Sokol ES, Del Vecchio CA, Sanduja S, Claessen JH et al. 2014. Epithelial-to-mesenchymal transition activates PERK-eIF2α and sensitizes cells to endoplasmic reticulum stress. Cancer Discov 4:702–15
    [Google Scholar]
  32. Fonseca BD, Zakaria C, Jia JJ, Graber TE, Svitkin Y et al. 2015. La-related protein 1 (LARP1) represses terminal oligopyrimidine (TOP) mRNA translation downstream of mTOR complex 1 (mTORC1). J. Biol. Chem. 290:15996–6020
    [Google Scholar]
  33. Franchini DM, Lanvin O, Tosolini M, Patras de Campaigno E, Cammas A et al. 2019. Microtubule-driven stress granule dynamics regulate inhibitory immune checkpoint expression in T cells. Cell Rep 26:94–107.e7
    [Google Scholar]
  34. Fruman DA, Rommel C. 2014. PI3K and cancer: lessons, challenges and opportunities. Nat. Rev. Drug Discov. 13:140–56
    [Google Scholar]
  35. Furic L, Rong LW, Larsson O, Koumakpayi IH, Yoshida K et al. 2010. eIF4E phosphorylation promotes tumorigenesis and is associated with prostate cancer progression. PNAS 107:14134–39
    [Google Scholar]
  36. Gandin V, Miluzio A, Barbieri AM, Beugnet A, Kiyokawa Het al. 2008. Eukaryotic initiation factor 6 is rate-limiting in translation, growth and transformation. Nature 455:68488
    [Google Scholar]
  37. Gatza ML, Silva GO, Parker JS, Fan C, Perou CM 2014. An integrated genomics approach identifies drivers of proliferation in luminal-subtype human breast cancer. Nat. Genet 46:105159
    [Google Scholar]
  38. Genuth NR, Barna M. 2018. The discovery of ribosome heterogeneity and its implications for gene regulation and organismal life. Mol. Cell 71:364–74
    [Google Scholar]
  39. Ghosh B, Benyumov AO, Ghosh P, Jia Y, Avdulov S et al. 2009. Nontoxic chemical interdiction of the epithelial-to-mesenchymal transition by targeting cap-dependent translation. ACS Chem. Biol. 4:367–77
    [Google Scholar]
  40. Gingold H, Tehler D, Christoffersen NR, Nielsen MM, Asmar F et al. 2014. A dual program for translation regulation in cellular proliferation and differentiation. Cell 158:1281–92
    [Google Scholar]
  41. Gingras AC, Raught B, Sonenberg N 1999. eIF4 initiation factors: effectors of mRNA recruitment to ribosomes and regulators of translation. Annu. Rev. Biochem. 68:913–63
    [Google Scholar]
  42. Girardi T, Vereecke S, Sulima SO, Khan Y, Fancello L et al. 2018. The T-cell leukemia-associated ribosomal RPL10 R98S mutation enhances JAK-STAT signaling. Leukemia 32:809–19
    [Google Scholar]
  43. Goodarzi H, Nguyen HCB, Zhang S, Dill BD, Molina H, Tavazoie SF 2016. Modulated expression of specific tRNAs drives gene expression and cancer progression. Cell 165:1416–27
    [Google Scholar]
  44. Grandori C, Gomez-Roman N, Felton-Edkins ZA, Ngouenet C, Galloway DA et al. 2005. c-Myc binds to human ribosomal DNA and stimulates transcription of rRNA genes by RNA polymerase I. Nat. Cell Biol. 7:311–18
    [Google Scholar]
  45. Gu L, Zhu N, Zhang H, Durden DL, Feng Y, Zhou M 2009. Regulation of XIAP translation and induction by MDM2 following irradiation. Cancer Cell 15:363–75
    [Google Scholar]
  46. Guan XY, Fung JM, Ma NF, Lau SH, Tai LS et al. 2004. Oncogenic role of eIF-5A2 in the development of ovarian cancer. Cancer Res 64:4197–200
    [Google Scholar]
  47. Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A et al. 2008. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol. Cell 30:214–26
    [Google Scholar]
  48. Han DL, Liu J, Chen CY, Dong LH, Liu Y et al. 2019. Anti-tumour immunity controlled through mRNA m6A methylation and YTHDF1 in dendritic cells. Nature 566:270–74
    [Google Scholar]
  49. He LR, Zhao HY, Li BK, Liu YH, Liu MZ et al. 2011. Overexpression of eIF5A-2 is an adverse prognostic marker of survival in stage I non-small cell lung cancer patients. Int. J. Cancer 129:143–50
    [Google Scholar]
  50. Heiss NS, Knight SW, Vulliamy TJ, Klauck SM, Wiemann S et al. 1998. X-linked dyskeratosis congenita is caused by mutations in a highly conserved gene with putative nucleolar functions. Nat. Genet. 19:32–38
    [Google Scholar]
  51. Holcik M, Yeh C, Korneluk RG, Chow T 2000. Translational upregulation of X-linked inhibitor of apoptosis (XIAP) increases resistance to radiation induced cell death. Oncogene 19:4174–77
    [Google Scholar]
  52. Hong DS, Kurzrock R, Oh Y, Wheler J, Naing A et al. 2011. A phase 1 dose escalation, pharmacokinetic, and pharmacodynamic evaluation of eIF-4E antisense oligonucleotide LY2275796 in patients with advanced cancer. Clin. Cancer Res. 17:6582–91
    [Google Scholar]
  53. Hopkins BD, Fine B, Steinbach N, Dendy M, Rapp Z et al. 2013. A secreted PTEN phosphatase that enters cells to alter signaling and survival. Science 341:399–402
    [Google Scholar]
  54. Hsieh AC, Liu Y, Edlind MP, Ingolia NT, Janes MR et al. 2012. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485:55–61
    [Google Scholar]
  55. Iadevaia V, Caldarola S, Tino E, Amaldi F, Loreni F 2008. All translation elongation factors and the e, f, and h subunits of translation initiation factor 3 are encoded by 5′-terminal oligopyrimidine (TOP) mRNAs. RNA 14:1730–36
    [Google Scholar]
  56. Iritani BM, Eisenman RN. 1999. c-Myc enhances protein synthesis and cell size during B lymphocyte development. PNAS 96:13180–85
    [Google Scholar]
  57. Iwasaki S, Floor SN, Ingolia NT 2016. Rocaglates convert DEAD-box protein eIF4A into a sequence-selective translational repressor. Nature 534:558–61
    [Google Scholar]
  58. Jasiulionis MG, Luchessi AD, Moreira AG, Souza PP, Suenaga AP et al. 2007. Inhibition of eukaryotic translation initiation factor 5A (eIF5A) hypusination impairs melanoma growth. Cell Biochem. Funct. 25:109–14
    [Google Scholar]
  59. Jefferies HBJ, Reinhard C, Kozma SC, Thomas G 1994. Rapamycin selectively represses translation of the “polypyrimidine tract” mRNA family. PNAS 91:4441–45
    [Google Scholar]
  60. Khalaileh A, Dreazen A, Khatib A, Apel R, Swisa A et al. 2013. Phosphorylation of ribosomal protein S6 attenuates DNA damage and tumor suppression during development of pancreatic cancer. Cancer Res 73:1811–20
    [Google Scholar]
  61. Konicek BW, Stephens JR, McNulty AM, Robichaud N, Peery RB et al. 2011. Therapeutic inhibition of MAP kinase interacting kinase blocks eukaryotic initiation factor 4E phosphorylation and suppresses outgrowth of experimental lung metastases. Cancer Res 71:1849–57
    [Google Scholar]
  62. Koromilas AE, Lazaris-Karatzas A, Sonenberg N 1992. mRNAs containing extensive secondary structure in their 5′ non-coding region translate efficiently in cells overexpressing initiation factor eIF-4E. EMBO J 11:4153–58
    [Google Scholar]
  63. Kozak M. 1986. Influences of mRNA secondary structure on initiation by eukaryotic ribosomes. PNAS 83:2850–54
    [Google Scholar]
  64. Krishnamoorthy GP, Davidson NR, Leach SD, Zhao Z, Lowe SW et al. 2018. EIF1AX and RAS mutations cooperate to drive thyroid tumorigenesis through ATF4 and c-MYC. Cancer Discov 9:264–81
    [Google Scholar]
  65. Lampson BL, Pershing NL, Prinz JA, Lacsina JR, Marzluff WF et al. 2013. Rare codons regulate KRas oncogenesis. Curr. Biol. 23:70–75
    [Google Scholar]
  66. Lazaris-Karatzas A, Montine KS, Sonenberg N 1990. Malignant transformation by a eukaryotic initiation factor subunit that binds to mRNA 5′ cap. Nature 345:544–47
    [Google Scholar]
  67. Lee ASY, Kranzusch PJ, Cate JH 2015. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522:111–14
    [Google Scholar]
  68. Lee ASY, Kranzusch PJ, Doudna JA, Cate JHD 2016. eIF3d is an mRNA cap-binding protein that is required for specialized translation initiation. Nature 536:96–99
    [Google Scholar]
  69. Leprivier G, Remke M, Rotblat B, Dubuc A, Mateo AR et al. 2013. The eEF2 kinase confers resistance to nutrient deprivation by blocking translation elongation. Cell 153:1064–79
    [Google Scholar]
  70. Levy S, Avni D, Hariharan N, Perry RP, Meyuhas O 1991. Oligopyrimidine tract at the 5′ end of mammalian ribosomal-protein mRNAs is required for their translational control. PNAS 88:3319–23
    [Google Scholar]
  71. Li M, Kao E, Malone D, Gao X, Wang JYJ, David M 2018. DNA damage-induced cell death relies on SLFN11-dependent cleavage of distinct type II tRNAs. Nat. Struct. Mol. Biol. 25:1047–58
    [Google Scholar]
  72. Lin SB, Choe J, Du P, Triboulet R, Gregory RI 2016. The m6A methyltransferase METTL3 promotes translation in human cancer cells. Mol. Cell 62:335–45
    [Google Scholar]
  73. Lipton JM, Federman N, Khabbaze Y, Schwartz CL, Hilliard LM et al. 2001. Osteogenic sarcoma associated with Diamond–Blackfan anemia: a report from the Diamond–Blackfan Anemia Registry. J. Pediatr. Hematol. Oncol. 23:39–44
    [Google Scholar]
  74. Liu L, Dilworth D, Gao L, Monzon J, Summers A et al. 1999. Mutation of the CDKN2A 5′ UTR creates an aberrant initiation codon and predisposes to melanoma. Nat. Genet. 21:128–32
    [Google Scholar]
  75. Liu S, Hausmann S, Carlson SM, Fuentes ME, Francis JW et al. 2019. METTL13 methylation of eEF1A increases translational output to promote tumorigenesis. Cell 176:491–504.e21
    [Google Scholar]
  76. Liwak U, Thakor N, Jordan LE, Roy R, Lewis SM et al. 2012. Tumor suppressor PDCD4 represses internal ribosome entry site-mediated translation of antiapoptotic proteins and is regulated by S6 kinase 2. Mol. Cell. Biol. 32:1818–29
    [Google Scholar]
  77. Ljungstrom V, Cortese D, Young E, Pandzic T, Mansouri L et al. 2016. Whole-exome sequencing in relapsing chronic lymphocytic leukemia: clinical impact of recurrent RPS15 mutations. Blood 127:1007–16
    [Google Scholar]
  78. Lovett PS, Rogers EJ. 1996. Ribosome regulation by the nascent peptide. Microbiol. Rev. 60:366–85
    [Google Scholar]
  79. Ma L, Chen Z, Erdjument-Bromage H, Tempst P, Pandolfi PP 2005. Phosphorylation and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and cancer pathogenesis. Cell 121:179–93
    [Google Scholar]
  80. Manier S, Huynh D, Shen YJ, Zhou J, Yusufzai T et al. 2017. Inhibiting the oncogenic translation program is an effective therapeutic strategy in multiple myeloma. Sci. Transl. Med. 9:eaal2668
    [Google Scholar]
  81. Marcel V, Ghayad SE, Belin S, Therizols G, Morel AP et al. 2013. p53 acts as a safeguard of translational control by regulating fibrillarin and rRNA methylation in cancer. Cancer Cell 24:318–30
    [Google Scholar]
  82. Marchetti A, Buttitta F, Pellegrini S, Bertacca G, Callahan R 2001. Reduced expression of INT-6/eIF3-p48 in human tumors. Int. J. Oncol. 18:175–79
    [Google Scholar]
  83. Martineau Y, Azar R, Muller D, Lasfargues C, El Khawand S et al. 2014. Pancreatic tumours escape from translational control through 4E-BP1 loss. Oncogene 33:1367–74
    [Google Scholar]
  84. Mazumder B, Li X, Barik S 2010. Translation control: a multifaceted regulator of inflammatory response. J. Immunol. 184:3311–19
    [Google Scholar]
  85. McGillivray P, Ault R, Pawashe M, Kitchen R, Balasubramanian S, Gerstein M 2018. A comprehensive catalog of predicted functional upstream open reading frames in humans. Nucleic Acids Res 46:3326–38
    [Google Scholar]
  86. McMahon M, Contreras A, Holm M, Uechi T, Forester CM et al. 2019. A single H/ACA small nucleolar RNA mediates tumor suppression downstream of oncogenic RAS. eLife 8:e48847
    [Google Scholar]
  87. Mehta A, Trotta CR, Peltz SW 2006. Derepression of the Her-2 uORF is mediated by a novel post-transcriptional control mechanism in cancer cells. Genes Dev 20:939–53
    [Google Scholar]
  88. Miluzio A, Beugnet A, Grosso S, Brina D, Mancino Met al. 2011. Impairment of cytoplasmic eIF6 activity restricts lymphomagenesis and tumor progression without affecting normal growth. Cancer Cell 19:76575
    [Google Scholar]
  89. Mizrachy-Schwartz S, Kravchenko-Balasha N, Ben-Bassat H, Klein S, Levitzki A 2007. Optimization of energy-consuming pathways towards rapid growth in HPV-transformed cells. PLOS ONE 2:e628
    [Google Scholar]
  90. Montanaro L, Brigotti M, Clohessy J, Barbieri S, Ceccarelli C et al. 2006. Dyskerin expression influences the level of ribosomal RNA pseudo-uridylation and telomerase RNA component in human breast cancer. J. Pathol. 210:10–18
    [Google Scholar]
  91. Nakamura J, Aoyagi S, Nanchi I, Nakatsuka S, Hirata E et al. 2009. Overexpression of eukaryotic elongation factor eEF2 in gastrointestinal cancers and its involvement in G2/M progression in the cell cycle. Int. J. Oncol. 34:1181–89
    [Google Scholar]
  92. Nakanishi S, Cleveland JL. 2016. Targeting the polyamine-hypusine circuit for the prevention and treatment of cancer. Amino Acids 48:2353–62
    [Google Scholar]
  93. Nguyen HG, Conn CS, Kye Y, Xue L, Forester CM et al. 2018. Development of a stress response therapy targeting aggressive prostate cancer. Sci. Transl. Med. 10:eaar2036
    [Google Scholar]
  94. Nishizawa Y, Konno M, Asai A, Koseki J, Kawamoto K et al. 2018. Oncogene c-Myc promotes epitranscriptome m6A reader YTHDF1 expression in colorectal cancer. Oncotarget 9:7476–86
    [Google Scholar]
  95. Northcott PA, Buchhalter I, Morrissy AS, Hovestadt V, Weischenfeldt J et al. 2017. The whole-genome landscape of medulloblastoma subtypes. Nature 547:311–17
    [Google Scholar]
  96. Nupponen NN, Porkka K, Kakkola L, Tanner M, Persson K et al. 1999. Amplification and overexpression of p40 subunit of eukaryotic translation initiation factor 3 in breast and prostate cancer. Am. J. Pathol. 154:1777–83
    [Google Scholar]
  97. Occhi G, Regazzo D, Trivellin G, Boaretto F, Ciato D et al. 2013. A novel mutation in the upstream open reading frame of the CDKN1B gene causes a MEN4 phenotype. PLOS Genet 9:e1003350
    [Google Scholar]
  98. Oh S, Flynn RA, Floor SN, Purzner J, Martin L et al. 2016. Medulloblastoma-associated DDX3 variant selectively alters the translational response to stress. Oncotarget 7:28169–82
    [Google Scholar]
  99. Ortiz-Zapater E, Pineda D, Martinez-Bosch N, Fernandez-Miranda G, Iglesias M et al. 2011. Key contribution of CPEB4-mediated translational control to cancer progression. Nat. Med. 18:83–90
    [Google Scholar]
  100. Palangat M, Anastasakis DG, Fei DL, Lindblad KE, Bradley R et al. 2019. The splicing factor U2AF1 contributes to cancer progression through a noncanonical role in translation regulation. Genes Dev 33:482–97
    [Google Scholar]
  101. Parsa AT, Waldron JS, Panner A, Crane CA, Parney IF et al. 2007. Loss of tumor suppressor PTEN function increases B7-H1 expression and immunoresistance in glioma. Nat. Med. 13:84–88
    [Google Scholar]
  102. Pavon-Eternod M, Gomes S, Geslain R, Dai Q, Rosner MR, Pan T 2009. tRNA over-expression in breast cancer and functional consequences. Nucleic Acids Res 37:7268–80
    [Google Scholar]
  103. Pelletier J, Sonenberg N. 1985. Insertion mutagenesis to increase secondary structure within the 5′ noncoding region of a eukaryotic mRNA reduces translational efficiency. Cell 40:515–26
    [Google Scholar]
  104. Peltonen K, Colis L, Liu H, Trivedi R, Moubarek MS et al. 2014. A targeting modality for destruction of RNA polymerase I that possesses anticancer activity. Cancer Cell 25:77–90
    [Google Scholar]
  105. Pereira B, Billaud M, Almeida R 2017. RNA-binding proteins in cancer: old players and new actors. Trends Cancer 3:506–28
    [Google Scholar]
  106. Rao S, Lee SY, Gutierrez A, Perrigoue J, Thapa RJ et al. 2012. Inactivation of ribosomal protein L22 promotes transformation by induction of the stemness factor, Lin28B. Blood 120:3764–73
    [Google Scholar]
  107. Ray BK, Lawson TG, Kramer JC, Cladaras MH, Grifo JA et al. 1985. ATP-dependent unwinding of messenger RNA structure by eukaryotic initiation factors. J. Biol. Chem. 260:7651–58
    [Google Scholar]
  108. Robichaud N, del Rincon SV, Huor B, Alain T, Petruccelli LA et al. 2015. Phosphorylation of eIF4E promotes EMT and metastasis via translational control of SNAIL and MMP-3. Oncogene 34:2032–42
    [Google Scholar]
  109. Robichaud N, Hsu BE, Istomine R, Alvarez F, Blagih J et al. 2018. Translational control in the tumor microenvironment promotes lung metastasis: phosphorylation of eIF4E in neutrophils. PNAS 115:E2202–9
    [Google Scholar]
  110. Robinson G, Parker M, Kranenburg TA, Lu C, Chen X et al. 2012. Novel mutations target distinct subgroups of medulloblastoma. Nature 488:43–48
    [Google Scholar]
  111. Roux PP, Ballif BA, Anjum R, Gygi SP, Blenis J 2004. Tumor-promoting phorbol esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via p90 ribosomal S6 kinase. PNAS 101:13489–94
    [Google Scholar]
  112. Ruggero D, Montanaro L, Ma L, Xu W, Londei P et al. 2004. The translation factor eIF-4E promotes tumor formation and cooperates with c-Myc in lymphomagenesis. Nat. Med. 10:484–86
    [Google Scholar]
  113. Saramaki O, Willi N, Bratt O, Gasser TC, Koivisto P et al. 2001. Amplification of EIF3S3 gene is associated with advanced stage in prostate cancer. Am. J. Pathol. 159:2089–94
    [Google Scholar]
  114. Schmidt EV. 2004. The role of c-myc in regulation of translation initiation. Oncogene 23:3217–21
    [Google Scholar]
  115. Sendoel A, Dunn JG, Rodriguez EH, Naik S, Gomez NC et al. 2017. Translation from unconventional 5′ start sites drives tumour initiation. Nature 541:494–99
    [Google Scholar]
  116. Shi J, Kahle A, Hershey JW, Honchak BM, Warneke JA et al. 2006. Decreased expression of eukaryotic initiation factor 3f deregulates translation and apoptosis in tumor cells. Oncogene 25:4923–36
    [Google Scholar]
  117. Shi Z, Fujii K, Kovary KM, Genuth NR, Rost HL et al. 2017. Heterogeneous ribosomes preferentially translate distinct subpools of mRNAs genome-wide. Mol. Cell 67:71–83.e7
    [Google Scholar]
  118. Shimamura A, Alter BP. 2010. Pathophysiology and management of inherited bone marrow failure syndromes. Blood Rev 24:101–22
    [Google Scholar]
  119. Silvera D, Arju R, Darvishian F, Levine PH, Zolfaghari L et al. 2009. Essential role for eIF4GI overexpression in the pathogenesis of inflammatory breast cancer. Nat. Cell Biol. 11:903–8
    [Google Scholar]
  120. Simsek D, Tiu GC, Flynn RA, Byeon GW, Leppek K et al. 2017. The mammalian ribo-interactome reveals ribosome functional diversity and heterogeneity. Cell 169:1051–65.e18
    [Google Scholar]
  121. Starck SR, Tsai JC, Chen K, Shodiya M, Wang L et al. 2016. Translation from the 5′ untranslated region shapes the integrated stress response. Science 351:aad3867
    [Google Scholar]
  122. Sulima SO, Hofman IJF, De Keersmaecker K, Dinman JD 2017. How ribosomes translate cancer. Cancer Discov 7:1069–87
    [Google Scholar]
  123. Tang DJ, Dong SS, Ma NF, Xie D, Chen L et al. 2010. Overexpression of eukaryotic initiation factor 5A2 enhances cell motility and promotes tumor metastasis in hepatocellular carcinoma. Hepatology 51:1255–63
    [Google Scholar]
  124. Tang H, Hornstein E, Stolovich M, Levy G, Livingstone M et al. 2001. Amino acid-induced translation of TOP mRNAs is fully dependent on phosphatidylinositol 3-kinase-mediated signaling, is partially inhibited by rapamycin, and is independent of S6K1 and rpS6 phosphorylation. Mol. Cell. Biol. 21:8671–83
    [Google Scholar]
  125. Tcherkezian J, Cargnello M, Romeo Y, Huttlin EL, Lavoie G et al. 2014. Proteomic analysis of cap-dependent translation identifies LARP1 as a key regulator of 5′TOP mRNA translation. Genes Dev 28:357–71
    [Google Scholar]
  126. Thoreen CC, Chantranupong L, Keys HR, Wang T, Gray NS, Sabatini DM 2012. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485:109–13
    [Google Scholar]
  127. Truitt ML, Conn CS, Shi Z, Pang X, Tokuyasu T et al. 2015. Differential requirements for eIF4E dose in normal development and cancer. Cell 162:59–71
    [Google Scholar]
  128. Ueda T, Sasaki M, Elia AJ, Chio II, Hamada K et al. 2010. Combined deficiency for MAP kinase-interacting kinase 1 and 2 (Mnk1 and Mnk2) delays tumor development. PNAS 107:13984–90
    [Google Scholar]
  129. Vu LP, Pickering BF, Cheng Y, Zaccara S, Nguyen D et al. 2017. The N6-methyladenosine (m6A)-forming enzyme METTL3 controls myeloid differentiation of normal hematopoietic and leukemia cells. Nat. Med. 23:1369–76
    [Google Scholar]
  130. Wang X, Zhao BS, Roundtree IA, Lu ZK, Han DL et al. 2015. N6-methyladenosine modulates messenger RNA translation efficiency. Cell 161:1388–99
    [Google Scholar]
  131. Wang ZY, Feng XD, Molinolo AA, Martin D, Vitale-Cross L et al. 2019. 4E-BP1 is a tumor suppressor protein reactivated by mTOR inhibition in head and neck cancer. Cancer Res 79:1438–50
    [Google Scholar]
  132. Waskiewicz AJ, Johnson JC, Penn B, Mahalingam M, Kimball SR, Cooper JA 1999. Phosphorylation of the cap-binding protein eukaryotic translation initiation factor 4E by protein kinase Mnk1 in vivo. Mol. Cell. Biol. 19:1871–80
    [Google Scholar]
  133. Wei J, Kishton RJ, Angel M, Conn CS, Dalla-Venezia N et al. 2019. Ribosomal proteins regulate MHC class I peptide generation for immunosurveillance. Mol. Cell 73:1162–73.e5
    [Google Scholar]
  134. Wek RC. 2018. Role of eIF2α kinases in translational control and adaptation to cellular stress. Cold Spring Harb. Perspect. Biol. 10:a032870
    [Google Scholar]
  135. Wolfe AL, Singh K, Zhong Y, Drewe P, Rajasekhar VK et al. 2014. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513:65–70
    [Google Scholar]
  136. Xu Y, Poggio M, Jin HY, Shi Z, Forester CM et al. 2019. Translation control of the immune checkpoint in cancer and its therapeutic targeting. Nat. Med. 25:301–11
    [Google Scholar]
  137. Yaffe MB. 2019. Why geneticists stole cancer research even though cancer is primarily a signaling disease. Sci. Signal. 12:eaaw3483
    [Google Scholar]
  138. Yamasaki S, Anderson P. 2008. Reprogramming mRNA translation during stress. Curr. Opin. Cell Biol. 20:222–26
    [Google Scholar]
  139. Zhao X, Zhang C, Zhou H, Xiao B, Cheng Y et al. 2016. Synergistic antitumor activity of the combination of salubrinal and rapamycin against human cholangiocarcinoma cells. Oncotarget 7:85492–501
    [Google Scholar]
  140. Zoncu R, Efeyan A, Sabatini DM 2011. mTOR: from growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 12:21–35
    [Google Scholar]
/content/journals/10.1146/annurev-cancerbio-030419-033420
Loading
/content/journals/10.1146/annurev-cancerbio-030419-033420
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error