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:

Translational control in cancer

Key Points

  • Alterations in the expression and activity of specific translation factors, and the uncoupling of translational regulation from inhibition by tumour cell stresses, such as hypoxia and nutrient deprivation, are a common feature of human cancers. These alterations are associated with specific types of human cancers and different stages of disease and cellular transformation.

  • Most, if not all, of the pathways that are crucial for cancer development and progression affect translational control. These include the selective translation of mRNAs encoding proteins that are engaged in tumour cell proliferation, growth, angiogenesis, response to mitogenic stimuli and the tumour microenvironment, and tumour stress responses such as hypoxia and nutrient deprivation.

  • Targeting specific translation factors that are altered in expression or activity in human cancers, and the specialized mRNA translation elements used preferentially in tumours, offers great promise for the development of a new generation of cancer therapeutics.

Abstract

Remarkable progress has been made in defining a new understanding of the role of mRNA translation and protein synthesis in human cancer. Translational control is a crucial component of cancer development and progression, directing both global control of protein synthesis and selective translation of specific mRNAs that promote tumour cell survival, angiogenesis, transformation, invasion and metastasis. Translational control of cancer is multifaceted, involving alterations in translation factor levels and activities unique to different types of cancers, disease stages and the tumour microenvironment. Several clinical efforts are underway to target specific components of the translation apparatus or unique mRNA translation elements for cancer therapeutics.

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: Eukaryotic cap-dependent mRNA translation.
Figure 2: A cap-dependent to cap-independent translation switch mediated by the Akt–mTOR–4E-BP pathway.
Figure 3: Key signal transduction pathways that regulate protein synthesis.

Similar content being viewed by others

References

  1. Pianese, G. Beitrag zur histologie und Aetiologie der carcinoma histologische und experimentelle. Beitr. Pathol. Anat. Allgem. Pathol. 142, 1–193 (1896).

    Google Scholar 

  2. Gani, R. The nucleoli of cultured human lymphocytes. I. Nucleolar morphology in relation to transformation and the DNA cycle. Exp. Cell Res. 97, 249–258 (1976).

    Article  CAS  PubMed  Google Scholar 

  3. Johnson, L. F., Levis, R., Abelson, H. T., Green, H. & Penman, S. Changes in RNA in relation to growth of the fibroblast. IV. Alterations in theproduction and processing of mRNA and rRNA in resting and growing cells. J. Cell Biol. 71, 933–938 (1976).

    Article  CAS  PubMed  Google Scholar 

  4. Zetterberg, A., Larsson, O. & Wiman, K. G. What is the restriction point? Curr. Opin. Cell. Biol. 7, 835–842 (1995).

    Article  CAS  PubMed  Google Scholar 

  5. Barna, M. et al. Suppression of Myc oncogenic activity by ribosomal protein haploinsufficiency. Nature 456, 971–975 (2008). This paper directly links Myc overexpression to cancer development through alteration of translational control. MYC reduces IRES-dependent Cdk11 translation during mitosis and increases chromosome instability.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Getz, M. J., Elder, P. K., Benz, E. W. Jr, Stephens, R. E. & Moses, H. L. Effect of cell proliferation on levels and diversity of poly(A)-containing mRNA. Cell 7, 255–265 (1976).

    Article  CAS  PubMed  Google Scholar 

  7. 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 

  8. Wek, R. C., Jiang, H. Y. & Anthony, T. G. Coping with stress: eIF2 kinases and translational control. Biochem. Soc. Trans. 34, 7–11 (2006).

    Article  CAS  PubMed  Google Scholar 

  9. Donze, O., Jagus, R., Koromilas, A. E., Hershey, J. W. & Sonenberg, N. Abrogation of translation initiation factor eIF-2 phosphorylation causes malignant transformation of NIH 3T3 cells. EMBO J. 14, 3828–3834 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Koromilas, A. E., Roy, S., Barber, G. N., Katze, M. G. & Sonenberg, N. Malignant transformation by a mutant of the IFN-inducible dsRNA-dependent protein kinase. Science 257, 1685–1689 (1992).

    Article  CAS  PubMed  Google Scholar 

  11. Tejada, S. et al. Eukaryotic initiation factors (eIF) 2α and 4E expression, localization, and phosphorylation in brain tumors. J. Histochem. Cytochem. 57, 503–512 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Abraham, N. et al. Characterization of transgenic mice with targeted disruption of the catalytic domain of the double-stranded RNA-dependent protein kinase, PKR. J. Biol. Chem. 274, 5953–5962 (1999).

    Article  CAS  PubMed  Google Scholar 

  13. Yang, Y. L. et al. Deficient signaling in mice devoid of double-stranded RNA-dependent protein kinase. EMBO J. 14, 6095–6106 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bi, M. et al. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. EMBO J. 24, 3470–3481 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Koritzinsky, M. et al. Gene expression during acute and prolonged hypoxia is regulated by distinct mechanisms of translational control. EMBO J. 25, 1114–1125 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Wouters, B. G. & Koritzinsky, M. Hypoxia signalling through mTOR and the unfolded protein response in cancer. Nature Rev. Cancer 8, 851–864 (2008).

    Article  CAS  Google Scholar 

  17. Schewe, D. M. & Aguirre-Ghiso, J. A. Inhibition of eIF2α dephosphorylation maximizes bortezomib efficiency and eliminates quiescent multiple myeloma cells surviving proteasome inhibitor therapy. Cancer Res. 69, 1545–1552 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Dong, Z. & Zhang, J. T. Initiation factor eIF3 and regulation of mRNA translation, cell growth, and cancer. Crit. Rev. Oncol. Hematol. 59, 169–180 (2006).

    Article  PubMed  Google Scholar 

  19. Zhang, L., Pan, X. & Hershey, J. W. Individual overexpression of five subunits of human translation initiation factor eIF3 promotes malignant transformation of immortal fibroblast cells. J. Biol. Chem. 282, 5790–5800 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Dong, Z., Liu, L. H., Han, B., Pincheira, R. & Zhang, J. T. Role of eIF3 p170 in controlling synthesis of ribonucleotide reductase M2 and cell growth. Oncogene 23, 3790–3801 (2004).

    Article  CAS  PubMed  Google Scholar 

  21. Dong, Z. & Zhang, J. T. EIF3 p170, a mediator of mimosine effect on protein synthesis and cell cycle progression. Mol. Biol. Cell 14, 3942–3951 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Scoles, D. R., Yong, W. H., Qin, Y., Wawrowsky, K. & Pulst, S. M. Schwannomin inhibits tumorigenesis through direct interaction with the eukaryotic initiation factor subunit c (eIF3c). Hum. Mol. Genet. 15, 1059–1070 (2006).

    Article  CAS  PubMed  Google Scholar 

  23. Nupponen, N. N. et al. Amplification and overexpression of p40 subunit of eukaryotic translation initiation factor 3 in breast and prostate cancer. Am. J. Pathol. 154, 1777–1783 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Saramaki, O. et al. Amplification of EIF3S3 gene is associated with advanced stage in prostate cancer. Am. J. Pathol. 159, 2089–2094 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Cappuzzo, F. et al. MYC and EIF3H coamplification significantly improve response and survival of non-small cell lung cancer patients (NSCLC) treated with gefitinib. J. Thorac Oncol. 4, 472–478 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Shi, J. et al. Decreased expression of eukaryotic initiation factor 3f deregulates translation and apoptosis in tumor cells. Oncogene 25, 4923–4936 (2006).

    Article  CAS  PubMed  Google Scholar 

  27. Marchetti, A. et al. Int-6, a highly conserved, widely expressed gene, is mutated by mouse mammary tumor virus in mammary preneoplasia. J. Virol. 69, 1932–1938 (1995). This paper remains the clearest example of transformation by a retrovirus that alters translational control. It shows that MMTV inserts within EIf3e (also known as Int6 ) to truncate its expression and induce mammary tumorigenesis in the mouse.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Rasmussen, S. B., Kordon, E., Callahan, R. & Smith, G. H. Evidence for the transforming activity of a truncated Int6 gene, in vitro. Oncogene 20, 5291–5301 (2001).

    Article  CAS  PubMed  Google Scholar 

  29. Mayeur, G. L. & Hershey, J. W. Malignant transformation by the eukaryotic translation initiation factor 3 subunit p48 (eIF3e). FEBS Lett. 514, 49–54 (2002).

    Article  CAS  PubMed  Google Scholar 

  30. Mack, D. L., Boulanger, C. A., Callahan, R. & Smith, G. H. Expression of truncated Int6/eIF3e in mammary alveolar epithelium leads to persistent hyperplasia and tumorigenesis. Breast Cancer Res. 9, R42 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Marchetti, A., Buttitta, F., Pellegrini, S., Bertacca, G. & Callahan, R. Reduced expression of INT-6/eIF3-p48 in human tumors. Int. J. Oncol. 18, 175–179 (2001).

    CAS  PubMed  Google Scholar 

  32. Umar, A. et al. Identification of a putative protein profile associating with tamoxifen therapy-resistance in breast cancer. Mol. Cell. Proteomics 8, 1278–1294 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Zanelli, C. F. & Valentini, S. R. Is there a role for eIF5A in translation? Amino Acids 33, 351–358 (2007).

    Article  CAS  PubMed  Google Scholar 

  34. Jenkins, Z. A., Haag, P. G. & Johansson, H. E. Human eIF5A2 on chromosome 3q25-q27 is a phylogenetically conserved vertebrate variant of eukaryotic translation initiation factor 5A with tissue-specific expression. Genomics 71, 101–109 (2001).

    Article  CAS  PubMed  Google Scholar 

  35. Benne, R., Brown-Luedi, M. L. & Hershey, J. W. Purification and characterization of protein synthesis initiation factors eIF-1, eIF-4C, eIF-4D, and eIF-5 from rabbit reticulocytes. J. Biol. Chem. 253, 3070–3077 (1978).

    Article  CAS  PubMed  Google Scholar 

  36. Kemper, W. M., Berry, K. W. & Merrick, W. C. Purification and properties of rabbit reticulocyte protein synthesis initiation factors M2Bα and M2Bβ. J. Biol. Chem. 251, 5551–5557 (1976).

    Article  CAS  PubMed  Google Scholar 

  37. Saini, P., Eyler, D. E., Green, R. & Dever, T. E. Hypusine-containing protein eIF5A promotes translation elongation. Nature 459, 118–121 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Guan, X. Y. et al. Oncogenic role of eIF-5A2 in the development of ovarian cancer. Cancer Res. 64, 4197–4200 (2004).

    Article  CAS  PubMed  Google Scholar 

  39. Zender, L. et al. An oncogenomics-based in vivo RNAi screen identifies tumor suppressors in liver cancer. Cell 135, 852–864 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Hanauske-Abel, H. M. et al. Detection of a sub-set of polysomal mRNAs associated with modulation of hypusine formation at the G1-S boundary. Proposal of a role for eIF-5A in onset of DNA replication. FEBS Lett. 366, 92–98 (1995).

    Article  CAS  PubMed  Google Scholar 

  41. Cracchiolo, B. M. et al. Eukaryotic initiation factor 5A-1 (eIF5A-1) as a diagnostic marker for aberrant proliferation in intraepithelial neoplasia of the vulva. Gynecol. Oncol. 94, 217–222 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Balabanov, S. et al. Hypusination of eukaryotic initiation factor 5A (eIF5A): a novel therapeutic target in BCR-ABL-positive leukemias identified by a proteomics approach. Blood 109, 1701–1711 (2007).

    Article  CAS  PubMed  Google Scholar 

  43. Miluzio, A., Beugnet, A., Volta, V. & Biffo, S. Eukaryotic initiation factor 6 mediates a continuum between 60S ribosome biogenesis and translation. EMBO Rep. 10, 459–465 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Rosso, P. et al. Overexpression of p27BBP in head and neck carcinomas and their lymph node metastases. Head Neck 26, 408–417 (2004).

    Article  PubMed  Google Scholar 

  45. Sanvito, F. et al. Expression of a highly conserved protein, p27BBP, during the progression of human colorectal cancer. Cancer Res. 60, 510–516 (2000).

    CAS  PubMed  Google Scholar 

  46. Flavin, R. J. et al. Altered eIF6 and Dicer expression is associated with clinicopathological features in ovarian serous carcinoma patients. Mod. Pathol. 21, 676–684 (2008).

    Article  CAS  PubMed  Google Scholar 

  47. Gandin, V. et al. Eukaryotic initiation factor 6 is rate-limiting in translation, growth and transformation. Nature 455, 684–688 (2008). This paper expanded the understanding of the role of protein synthesis in cancer to include regulation of the 60S ribosome subunit and its ability to form translation-capable 80S ribosomes through external signals acting on eIF6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. 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 

  49. Graff, J. R., Konicek, B. W., Carter, J. H. & Marcusson, E. G. Targeting the eukaryotic translation initiation factor 4E for cancer therapy. Cancer Res. 68, 631–634 (2008).

    Article  CAS  PubMed  Google Scholar 

  50. 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).

    Article  CAS  PubMed  Google Scholar 

  51. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Lazaris-Karatzas, A. et al. Ras mediates translation initiation factor 4E-induced malignant transformation. Genes Dev. 6, 1631–1642 (1992). The first study to clearly identify a translation factor (eIF4E) as an effector of an oncogenic pathway. It showed that eIF4E overexpression, as occurs in many cancers, is stimulated by the Ras pathway, and is an essential partner in tumorigenesis through increased translation.

    Article  CAS  PubMed  Google Scholar 

  53. Ruggero, D. et al. The translation factor eIF-4E promotes tumor formation and cooperates with c-Myc in lymphomagenesis. Nature Med. 10, 484–486 (2004).

    Article  CAS  PubMed  Google Scholar 

  54. Wendel, H. G. et al. Survival signalling by Akt and eIF4E in oncogenesis and cancer therapy. Nature 428, 332–337 (2004).

    Article  CAS  PubMed  Google Scholar 

  55. 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 

  56. 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 

  57. Mamane, Y. et al. Epigenetic activation of a subset of mRNAs by eIF4E explains its effects on cell proliferation. PLoS One 2, e242 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Larsson, O. et al. Eukaryotic translation initiation factor 4E induced progression of primary human mammary epithelial cells along the cancer pathway is associated with targeted translational deregulation of oncogenic drivers and inhibitors. Cancer Res. 67, 6814–6824 (2007).

    Article  CAS  PubMed  Google Scholar 

  59. Schneider, R. J. & Sonenberg, N. in Translational Control in Biology and Medicine (eds. Mathews, M. B., Sonenberg, N. & Hershey, J. W. B.) (Cold Spring Harbor Laboratory Press, Cold Spring Harbor, 2007).

    Google Scholar 

  60. Waskiewicz, A. J. et al. Phosphorylation of the cap-binding protein eukaryotic translation initiation factor 4E by protein kinase Mnk1 in vivo. Mol. Cell. Biol. 19, 1871–1880 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kim, S. H., Miller, F. R., Tait, L., Zheng, J. & Novak, R. F. Proteomic and phosphoproteomic alterations in benign, premalignant and tumor human breast epithelial cells and xenograft lesions: biomarkers of progression. Int. J. Cancer 124, 2813–2828 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Graff, J. R. et al. eIF4E activation is commonly elevated in advanced human prostate cancers and significantly related to reduced patient survival. Cancer Res. 69, 3866–3873 (2009).

    Article  CAS  PubMed  Google Scholar 

  63. Wendel, H. G. et al. Dissecting eIF4E action in tumorigenesis. Genes Dev. 21, 3232–3237 (2007). Using a mouse lymphoma model, this study demonstrated an essential role for eIF4E phosphorylation in promoting tumorigenesis and anti-apoptosis, in part by increased translation of Mcl1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Rahmani, M., Davis, E. M., Bauer, C., Dent, P. & Grant, S. Apoptosis induced by the kinase inhibitor BAY 43–9006 in human leukemia cells involves down-regulation of Mcl-1 through inhibition of translation. J. Biol. Chem. 280, 35217–35227 (2005).

    Article  CAS  PubMed  Google Scholar 

  65. Liu, L. et al. Sorafenib blocks the RAF/MEK/ERK pathway, inhibits tumor angiogenesis, and induces tumor cell apoptosis in hepatocellular carcinoma model PLC/PRF/5. Cancer Res. 66, 11851–11858 (2006).

    Article  CAS  PubMed  Google Scholar 

  66. Noske, A. et al. Activation of mTOR in a subgroup of ovarian carcinomas: correlation with p-eIF-4E and prognosis. Oncol. Rep. 20, 1409–1417 (2008).

    CAS  PubMed  Google Scholar 

  67. Fan, S. et al. Phosphorylated eukaryotic translation initiation factor 4 (eIF4E) is elevated in human cancer tissues. Cancer Biol. Ther. 8, 1463–1469 (2009).

    Article  PubMed  Google Scholar 

  68. Coleman, L. J. et al. Combined analysis of eIF4E and 4E-binding protein expression predicts breast cancer survival and estimates eIF4E activity. Br. J. Cancer 100, 1393–1399 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Salehi, Z. & Mashayekhi, F. Expression of the eukaryotic translation initiation factor 4E (eIF4E) and 4E-BP1 in esophageal cancer. Clin. Biochem. 39, 404–409 (2006).

    Article  CAS  PubMed  Google Scholar 

  70. 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 showed that most LABCs overexpress eIF4G1 and 4E-BP1, which was shown in animal models to comprise a hypoxia-activated switch that stimulates IRES-dependent translation of survival, angiogenesis and hypoxia-orchestrating mRNAs, promoting development of locally advanced tumours.

    Article  CAS  PubMed  Google Scholar 

  71. Fukuchi-Shimogori, T. et al. Malignant transformation by overproduction of translation initiation factor eIF4G. Cancer Res. 57, 5041–5044 (1997).

    CAS  PubMed  Google Scholar 

  72. Comtesse, N. et al. Frequent overexpression of the genes FXR1, CLAPM1 and EIF4G located on amplicon 3q26–27 in squamous cell carcinoma of the lung. Int. J. Cancer 120, 2538–2544 (2007).

    Article  CAS  PubMed  Google Scholar 

  73. Bauer, C. et al. Translation initiation factor eIF-4G is immunogenic, overexpressed, and amplified in patients with squamous cell lung carcinoma. Cancer 92, 822–829 (2001).

    Article  CAS  PubMed  Google Scholar 

  74. Rosenwald, I. B., Hutzler, M. J., Wang, S., Savas, L. & Fraire, A. E. Expression of eukaryotic translation initiation factors 4E and 2α is increased frequently in bronchioloalveolar but not in squamous cell carcinomas of the lung. Cancer 92, 2164–2171 (2001).

    Article  CAS  PubMed  Google Scholar 

  75. Silvera, D. et al. Essential role for eIF4GI overexpression in the pathogenesis of inflammatory breast cancer. Naure Cell Biol. 11, 903–908 (2009). This paper showed that overexpression of eIF4G1 occurs in IBC and is crucial for disease development. High levels of eIF4G1 promoted IRES-dependent translation of P120CTN and formation of metastatic emboli.

    CAS  Google Scholar 

  76. Silvera, D. & Schneider, R. J. Inflammatory breast cancer cells are constitutively adapted to hypoxia. Cell Cycle 8, 3091–3096 (2009).

    Article  CAS  PubMed  Google Scholar 

  77. Shuda, M. et al. Enhanced expression of translation factor mRNAs in hepatocellular carcinoma. Anticancer Res. 20, 2489–2494 (2000).

    CAS  PubMed  Google Scholar 

  78. Eberle, J., Krasagakis, K. & Orfanos, C. E. Translation initiation factor eIF-4A1 mRNA is consistently overexpressed in human melanoma cells in vitro. Int. J. Cancer 71, 396–401 (1997).

    Article  CAS  PubMed  Google Scholar 

  79. Bordeleau, M. E. et al. Therapeutic suppression of translation initiation modulates chemosensitivity in a mouse lymphoma model. J. Clin. Invest. 118, 2651–2660 (2008). This study used small-molecule library screening to identify a specific inhibitor of eIF4A (silvestrol), demonstrating that initiation factors could be targeted for anticancer therapy. Silvestrol enhanced chemosensitivity in a mouse lymphoma model.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  81. Lankat-Buttgereit, B. & Goke, R. The tumour suppressor Pdcd4: recent advances in the elucidation of function and regulation. Biol. Cell 101, 309–317 (2009).

    Article  CAS  PubMed  Google Scholar 

  82. Yang, H. S. et al. The transformation suppressor Pdcd4 is a novel eukaryotic translation initiation factor 4A binding protein that inhibits translation. Mol. Cell. Biol. 23, 26–37 (2003).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. LaRonde-LeBlanc, N., Santhanam, A. N., Baker, A. R., Wlodawer, A. & Colburn, N. H. Structural basis for inhibition of translation by the tumor suppressor Pdcd4. Mol. Cell. Biol. 27, 147–156 (2007).

    Article  CAS  PubMed  Google Scholar 

  84. Holcik, M. Targeting translation for treatment of cancer—a novel role for IRES? Curr. Cancer Drug Targets. 4, 299–311 (2004).

    Article  CAS  PubMed  Google Scholar 

  85. Jopling, C. L. & Willis, A. E. N. -myc translation is initiated via an internal ribosome entry segment that displays enhanced activity in neuronal cells. Oncogene 20, 2664–26670 (2001).

    Article  CAS  PubMed  Google Scholar 

  86. Spriggs, K. A. et al. Canonical initiation factor requirements of the Myc family of internal ribosome entry segments. Mol. Cell. Biol. 29, 1565–1574 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  88. Evans, J. R. et al. Members of the poly (rC) binding protein family stimulate the activity of the c-myc internal ribosome entry segment in vitro and in vivo. Oncogene 22, 8012–8020 (2003).

    Article  PubMed  CAS  Google Scholar 

  89. Shi, Y. et al. IL-6-induced stimulation of c-myc translation in multiple myeloma cells is mediated by myc internal ribosome entry site function and the RNA-binding protein, hnRNP A1. Cancer Res. 68, 10215–10222 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Holcik, M., Yeh, C., Korneluk, R. G. & Chow, T. Translational upregulation of X-linked inhibitor of apoptosis (XIAP) increases resistance to radiation induced cell death. Oncogene 19, 4174–4177 (2000).

    Article  CAS  PubMed  Google Scholar 

  91. Desplanques, G. et al. Impact of XIAP protein levels on the survival of myeloma cells. Haematologica 94, 87–93 (2009).

    Article  CAS  PubMed  Google Scholar 

  92. Engelman, J. A. Targeting PI3K signalling in cancer: opportunities, challenges and limitations. Nature Rev. Cancer 9, 550–562 (2009).

    Article  CAS  Google Scholar 

  93. Ma, X. M. & Blenis, J. Molecular mechanisms of mTOR-mediated translational control. Nature Rev. Mol. Cell. Biol. 10, 307–318 (2009).

    Article  CAS  Google Scholar 

  94. 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 

  95. Pyronnet, S. et al. Human eukaryotic translation initiation factor 4G (eIF4G) recruits mnk1 to phosphorylate eIF4E. EMBO J. 18, 270–279 (1999). This study demonstrated that MNK1 is part of the eIF4F complex, resides on initiation factor eIF4G and phosphorylates eIF4E after assembly of eIF4F to promote translation initiation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Guertin, D. A. & Sabatini, D. M. Defining the role of mTOR in cancer. Cancer Cell 12, 9–22 (2007).

    Article  CAS  PubMed  Google Scholar 

  97. Lin, C. J., Malina, A. & Pelletier, J. c-Myc and eIF4F constitute a feedforward loop that regulates cell growth: implications for anticancer therapy. Cancer Res. 69, 7491–7494 (2009).

    Article  CAS  PubMed  Google Scholar 

  98. Mavrakis, K. J. et al. Tumorigenic activity and therapeutic inhibition of Rheb GTPase. Genes Dev. 22, 2178–2188 (2008). This study demonstrated that RHEB, an activator of mTOR in the mTORC1 complex, promotes lymphoma in PTEN-deficient mice and requires eIF4E as an effector.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Wu, X., Senechal, K., Neshat, M. S., Whang, Y. E. & Sawyers, C. L. The PTEN/MMAC1 tumor suppressor phosphatase functions as a negative regulator of the phosphoinositide 3-kinase/Akt pathway. Proc. Natl. Acad. Sci. USA 95, 15587–15591 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Zhou, X. et al. Activation of the Akt/mammalian target of rapamycin/4E-BP1 pathway by ErbB2 overexpression predicts tumor progression in breast cancers. Clin. Cancer Res. 10, 6779–6788 (2004).

    Article  CAS  PubMed  Google Scholar 

  101. Ruggero, D. & Pandolfi, P. P. Does the ribosome translate cancer? Nature Rev. Cancer 3, 179–192 (2003).

    Article  CAS  Google Scholar 

  102. Nakamura, J. L., Garcia, E. & Pieper, R. O. S6K1 plays a key role in glial transformation. Cancer Res. 68, 6516–6523 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Holz, M. K., Ballif, B. A., Gygi, S. P. & Blenis, J. mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events. Cell 123, 569–580 (2005).

    Article  CAS  PubMed  Google Scholar 

  104. Dorrello, N. V. et al. S6K1- and βTRCP-mediated degradation of PDCD4 promotes protein translation and cell growth. Science 314, 467–471 (2006). This paper showed that eIF4A is a target of the tumour suppressor PDCD4. In response to mitogens, S6K1 phosphorylates PDCD4, which is then rapidly degraded in the proteasome, releasing eIF4A activity and promoting translation.

    Article  CAS  PubMed  Google Scholar 

  105. Carayol, N. et al. Suppression of programmed cell death 4 (PDCD4) protein expression by BCR-ABL-regulated engagement of the mTOR/p70 S6 kinase pathway. J. Biol. Chem. 283, 8601–8610 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Shahbazian, D. et al. The mTOR/PI3K and MAPK pathways converge on eIF4B to control its phosphorylation and activity. EMBO J. 25, 2781–2791 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Maira, S. M., Stauffer, F., Schnell, C. & Garcia-Echeverria, C. PI3K inhibitors for cancer treatment: where do we stand? Biochem. Soc. Trans. 37, 265–272 (2009).

    Article  CAS  PubMed  Google Scholar 

  108. Choo, A. Y. & Blenis, J. Not all substrates are treated equally: implications for mTOR, rapamycin-resistance and cancer therapy. Cell Cycle 8, 567–572 (2009).

    Article  CAS  PubMed  Google Scholar 

  109. Yu, K. et al. Beyond rapalog therapy: preclinical pharmacology and antitumor activity of WYE-125132, an ATP-competitive and specific inhibitor of mTORC1 and mTORC2. Cancer Res. 70, 621–631 (2010).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  111. 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  CAS  PubMed  PubMed Central  Google Scholar 

  112. Inoki, K., Li, Y., Zhu, T., Wu, J. & Guan, K. L. TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nature Cell Biol. 4, 648–657 (2002). This study demonstrated that a primary mechanism for downregulation of mTOR is through the upstream TSC1 and TSC2 and that TSC2 is inhibited by activated Akt.

    Article  CAS  PubMed  Google Scholar 

  113. Avruch, J. et al. Activation of mTORC1 in two steps: Rheb-GTP activation of catalytic function and increased binding of substrates to raptor. Biochem. Soc. Trans. 37, 223–226 (2009).

    Article  CAS  PubMed  Google Scholar 

  114. Vander Haar, E., Lee, S. I., Bandhakavi, S., Griffin, T. J. & Kim, D. H. Insulin signalling to mTOR mediated by the Akt/PKB substrate PRAS40. Nature Cell Biol. 9, 316–323 (2007).

    Article  CAS  PubMed  Google Scholar 

  115. Madhunapantula, S. V., Sharma, A. & Robertson, G. P. PRAS40 deregulates apoptosis in malignant melanoma. Cancer Res. 67, 3626–3636 (2007).

    Article  CAS  PubMed  Google Scholar 

  116. Johnson, M. D., O'Connell, M., Vito, F. & Bakos, R. S. Increased STAT-3 and synchronous activation of Raf-1-MEK-1-MAPK, and phosphatidylinositol 3-Kinase-Akt-mTOR pathways in atypical and anaplastic meningiomas. J. Neurooncol. 92, 129–136 (2009).

    Article  CAS  PubMed  Google Scholar 

  117. Kwiatkowski, D. J. Tuberous sclerosis: from tubers to mTOR. Ann. Hum. Genet. 67, 87–96 (2003).

    Article  CAS  PubMed  Google Scholar 

  118. Budanov, A. V. & Karin, M. p53 target genes sestrin1 and sestrin2 connect genotoxic stress and mTOR signaling. Cell 134, 451–460 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Braunstein, S., Badura, M. L., Xi, Q., Formenti, S. C. & Schneider, R. J. Regulation of protein synthesis by ionizing radiation. Mol. Cell. Biol. 29, 5645–5656 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Haines, G. K. 3rd. et al. Expression of the double-stranded RNA-dependent protein kinase (p68) in squamous cell carcinoma of the head and neck region. Arch. Otolaryngol. Head Neck Surg. 119, 1142–1147 (1993).

    Article  PubMed  Google Scholar 

  121. Haines, G. K. et al. Correlation of the expression of double-stranded RNA-dependent protein kinase (p68) with differentiation in head and neck squamous cell carcinoma. Virchows Arch. B Cell Pathol. Incl. Mol. Pathol. 63, 289–295 (1993).

    Article  CAS  PubMed  Google Scholar 

  122. Haines, G. K. 3rd. et al. Interferon-responsive protein kinase (p68) and proliferating cell nuclear antigen are inversely distributed in head and neck squamous cell carcinoma. Tumour Biol. 19, 52–59 (1998).

    Article  PubMed  Google Scholar 

  123. Terada, T., Maeta, H., Endo, K. & Ohta, T. Protein expression of double-stranded RNA-activated protein kinase in thyroid carcinomas: correlations with histologic types, pathologic parameters, and Ki-67 labeling. Hum. Pathol. 31, 817–821 (2000).

    Article  CAS  PubMed  Google Scholar 

  124. Kim, S. H., Forman, A. P., Mathews, M. B. & Gunnery, S. Human breast cancer cells contain elevated levels and activity of the protein kinase, PKR. Oncogene 19, 3086–3094 (2000).

    Article  CAS  PubMed  Google Scholar 

  125. Haines, G. K. et al. Expression of the double-stranded RNA-dependent protein kinase (p68) in human breast tissues. Tumour Biol. 17, 5–12 (1996).

    Article  CAS  PubMed  Google Scholar 

  126. Shimada, A. et al. Aberrant expression of double-stranded RNA-dependent protein kinase in hepatocytes of chronic hepatitis and differentiated hepatocellular carcinoma. Cancer Res. 58, 4434–4438 (1998).

    CAS  PubMed  Google Scholar 

  127. Rosenwald, I. B., Wang, S., Savas, L., Woda, B. & Pullman, J. Expression of translation initiation factor eIF-2α is increased in benign and malignant melanocytic and colonic epithelial neoplasms. Cancer 98, 1080–1088 (2003).

    Article  CAS  PubMed  Google Scholar 

  128. Wang, S. et al. Expression of eukaryotic translation initiation factors 4E and 2α correlates with the progression of thyroid carcinoma. Thyroid 11, 1101–1107 (2001).

    Article  CAS  PubMed  Google Scholar 

  129. Wang, S. et al. Expression of the eukaryotic translation initiation factors 4E and 2α in non-Hodgkin's lymphomas. Am. J. Pathol. 155, 247–255 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Rosenwald, I. B. et al. Expression of the translation initiation factors eIF-4E and eIF-2* is frequently increased in neoplastic cells of Hodgkin lymphoma. Hum. Pathol. 39, 910–916 (2008).

    Article  CAS  PubMed  Google Scholar 

  131. Bachmann, F., Banziger, R. & Burger, M. M. Cloning of a novel protein overexpressed in human mammary carcinoma. Cancer Res. 57, 988–994 (1997).

    CAS  PubMed  Google Scholar 

  132. Dellas, A. et al. Expression of p150 in cervical neoplasia and its potential value in predicting survival. Cancer 83, 1376–1383 (1998).

    Article  CAS  PubMed  Google Scholar 

  133. Pincheira, R., Chen, Q. & Zhang, J. T. Identification of a 170-kDa protein over-expressed in lung cancers. Br. J. Cancer 84, 1520–1527 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Chen, G. & Burger, M. M. p150 expression and its prognostic value in squamous-cell carcinoma of the esophagus. Int. J. Cancer 84, 95–100 (1999).

    Google Scholar 

  135. Chen, G. & Burger, M. M. p150 overexpression in gastric carcinoma: the association with p53, apoptosis and cell proliferation. Int. J. Cancer 112, 393–398 (2004).

    Article  CAS  PubMed  Google Scholar 

  136. Rothe, M., Ko, Y., Albers, P. & Wernert, N. Eukaryotic initiation factor 3 p110 mRNA is overexpressed in testicular seminomas. Am. J. Pathol. 157, 1597–1604 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Guan, X. Y. et al. Isolation of a novel candidate oncogene within a frequently amplified region at 3q26 in ovarian cancer. Cancer Res. 61, 3806–3809 (2001).

    CAS  PubMed  Google Scholar 

  138. Xie, D. et al. Overexpression of EIF-5A2 is associated with metastasis of human colorectal carcinoma. Hum. Pathol. 39, 80–86 (2008).

    Article  CAS  PubMed  Google Scholar 

  139. Luo, J. H. et al. Overexpression of EIF-5A2 predicts tumor recurrence and progression in pTa/pT1 urothelial carcinoma of the bladder. Cancer Sci. 100, 896–902 (2009).

    Article  CAS  PubMed  Google Scholar 

  140. Chen, W. et al. Overexpression of EIF-5A2 is an independent predictor of outcome in patients of urothelial carcinoma of the bladder treated with radical cystectomy. Cancer Epidemiol. Biomarkers Prev. 18, 400–408 (2009).

    Article  CAS  PubMed  Google Scholar 

  141. De Benedetti, A. & Graff, J. R. eIF-4E expression and its role in malignancies and metastases. Oncogene 23, 3189–3199 (2004).

    Article  CAS  PubMed  Google Scholar 

  142. Thumma, S. C. & Kratzke, R. A. Translational control: a target for cancer therapy. Cancer Lett. 258, 1–8 (2007).

    Article  CAS  PubMed  Google Scholar 

  143. Bauer, C. et al. Overexpression of the eukaryotic translation initiation factor 4G (eIF4G-1) in squamous cell lung carcinoma. Int. J. Cancer 98, 181–185 (2002).

    Article  CAS  PubMed  Google Scholar 

  144. Brass, N. et al. Translation initiation factor eIF-4γ is encoded by an amplified gene and induces an immune response in squamous cell lung carcinoma. Hum. Mol. Genet. 6, 33–39 (1997).

    Article  CAS  PubMed  Google Scholar 

  145. Konicek, B. W., Dumstorf, C. A. & Graff, J. R. Targeting the eIF4F translation initiation complex for cancer therapy. Cell Cycle 7, 2466–2471 (2008).

    Article  CAS  PubMed  Google Scholar 

  146. Moerke, N. J. et al. Small-molecule inhibition of the interaction between the translation initiation factors eIF4E and eIF4G. Cell 128, 257–267 (2007).

    Article  CAS  PubMed  Google Scholar 

  147. Tamburini, J. et al. Protein synthesis is resistant to rapamycin and constitutes a promising therapeutic target in acute myeloid leukemia. Blood 114, 1618–1627 (2009).

    Article  CAS  PubMed  Google Scholar 

  148. Faivre, S., Kroemer, G. & Raymond, E. Current development of mTOR inhibitors as anticancer agents. Nature Rev. Drug Discov. 5, 671–688 (2006).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Department of Defense Breast Cancer Research grant W81XWH-04-1 (R.J.S. and S.C.F.), Breast Cancer Research Foundation grant (R.J.S. and S.C.F.) and New York State Health grant (D.S.).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert J. Schneider.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

National Cancer Institute Drug Dictionary 

gefitinib

FURTHER INFORMATION

Robert J. Schneider's homepage

Glossary

Eμ-Myc lymphomagenesis mouse model

One of the first mouse cancer models, it uses overexpression of Myc in B cells of these mice driven by the immunoglobulin heavy chain enhancer ().

Polyribosome

An mRNA–ribosome complex that generally contains actively translating mRNAs bound to multiple 80S ribosomes. The greater the number of ribosomes associated with an mRNA the greater the translation activity of the mRNA.

m7GpppN cap

A post-transcriptional, covalently linked inverted m7GTP (m7G) attached to the first 5′ nucleotide (N) of the mRNA. It protects the mRNA from degradation and binds eIF4E, recruiting eIF4F and 40S ribosome subunits.

Unfolded protein response

A cellular response to stress that is unique to the endoplasmic reticulum (ER), and which senses the misfolding of proteins in the ER. It activates pathways that help cells survive proteotoxicity caused by unfolded proteins or activate mechanisms of cell death.

Tumour quiescence

A mechanism whereby tumour cells can adapt by genetic and epigenetic alterations to unfavourable conditions in the microenvironment before proliferating again.

Gleason score

Pathological classification of the grade of prostate cancer. A lower score indicates a well-differentiated tumour with slower growth. A higher score indicates a poorly differentiated tumour with aggressive growth.

Mouse mammary tumour virus (MMTV)

MMTV causes mammary adenocarcinoma in infected mice and selectively replicates in the alveolar epithelial cells of the mammary gland. The MMTV promoter and enhancer are often used to direct the expression of transgenes specifically in the mammary gland.

Mouse tumour xenograft models

Hetero-transplantation of human tumour cells into immunodeficient mice, in either the orthotopic (same organ) site or ectopic (foreign) site. Mice are typically athymic nu/nu T cell deficient or severe combined immunodeficient (SCID), lacking B cell and T cell functions.

BCR–ABL

A leukaemogenic fusion protein that is a result of reciprocal translocation between chromosome 9 and chromosome 22. The 5′ section of BCR from chromosome 9 is fused to most of the proto-oncogene ABL from chromosome 22.

5′ untranslated region (5′UTR)

Also known as the 5′ noncoding region in eukaryotes, this includes the region of the mRNA from the transcription start site (m7GpppN cap) to the start of translation initiation, typically an AUG codon.

Internal ribosome entry site (IRES)

A nucleotide sequence in the mRNA 5′ noncoding region that confers internal ribosome initiation of translation without eIF4E but typically with other initiation factors, bypassing the m7GpppN cap and cap-dependent regulation.

Flavagline drug

A strong cell cycle inhibiting chemotherapeutic flavonol-cinnamate drug, a natural product of the plant genus Aglaia. It impairs ribosome recruitment by acting on eIF4A.

Rapamycin

An immunosuppressive drug that inhibits mTOR. Rapamycin and other mTOR inhibitors are being assessed as potential therapeutics in several cancers.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Silvera, D., Formenti, S. & Schneider, R. Translational control in cancer. Nat Rev Cancer 10, 254–266 (2010). https://doi.org/10.1038/nrc2824

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc2824

This article is cited by

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