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 of stem cell function

Abstract

Stem cells are characterized by their ability to self-renew and differentiate into many different cell types. Research has focused primarily on how these processes are regulated at a transcriptional level. However, recent studies have indicated that stem cell behaviour is strongly coupled to the regulation of protein synthesis by the ribosome. In this Review, we discuss how different translation mechanisms control the function of adult and embryonic stem cells. Stem cells are characterized by low global translation rates despite high levels of ribosome biogenesis. The maintenance of pluripotency, the commitment to a specific cell fate and the switch to cell differentiation depend on the tight regulation of protein synthesis and ribosome biogenesis. Translation regulatory mechanisms that impact on stem cell function include mTOR signalling, ribosome levels, and mRNA and tRNA features and amounts. Understanding these mechanisms important for stem cell self-renewal and differentiation may also guide our understanding of cancer grade and metastasis.

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

Fig. 1: Overview of stem cell types and functions.
Fig. 2: Translation and ribosome biogenesis in stem cells.
Fig. 3: mTOR signalling coordinates protein synthesis and ribosome biogensis to stem cell growth, survival and differentiation.
Fig. 4: The differentiation ‘switch’.
Fig. 5: mRNA and tRNA features regulate translation and stem cell function.
Fig. 6: Translational control of tumour grade and metastasis.

Similar content being viewed by others

References

  1. Theunissen, T. W. & Jaenisch, R. Mechanisms of gene regulation in human embryos and pluripotent stem cells. Development 144, 4496–4509 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Young, R. A. Control of the embryonic stem cell state. Cell 144, 940–954 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Bulut-Karslioglu, A. et al. The transcriptionally permissive chromatin state of embryonic stem cells is acutely tuned to translational output. Cell Stem Cell 22, 369–383.e8 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Iwafuchi-Doi, M. & Zaret, K. S. Cell fate control by pioneer transcription factors. Development 143, 1833–1837 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Freimer, J. W., Hu, T. J. & Blelloch, R. Decoupling the impact of microRNAs on translational repression versus RNA degradation in embryonic stem cells. eLife 7, e38014 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  6. Atlasi, Y. et al. The translational landscape of ground state pluripotency. Nat. Commun. 11, 1617 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Tahmasebi, S., Khoutorsky, A., Mathews, M. B. & Sonenberg, N. Translation deregulation in human disease. Nat. Rev. Mol. Cell Biol. 19, 791–807 (2018).

    Article  CAS  PubMed  Google Scholar 

  8. Buszczak, M., Signer, R. A. J. & Morrison, S. J. Cellular differences in protein synthesis regulate tissue homeostasis. Cell 159, 242–251 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Gabut, M., Bourdelais, F. & Durand, S. Ribosome and translational control in stem cells. Cells 9, 497 (2020).

    Article  CAS  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Dever, T. E., Dinman, J. D. & Green, R. Translation elongation and recoding in eukaryotes. Cold Spring Harb. Perspect. Biol. 10, a032649 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Dever, T. E. & Green, R. The elongation, termination, and recycling phases of translation in eukaryotes. Cold Spring Harb. Perspect. Biol. 4, a013706 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Pisarev, A. V. et al. The role of ABCE1 in eukaryotic posttermination ribosomal recycling. Mol. Cell 37, 196–210 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Sampath, P. et al. A hierarchical network controls protein translation during murine embryonic stem cell self-renewal and differentiation. Cell Stem Cell 2, 448–460 (2008). Sampath et al. (2008) show for the first time that protein synthesis rates are low in mouse ESCs relative to differentiated cell types.

    Article  CAS  PubMed  Google Scholar 

  15. Fortier, S., MacRae, T., Bilodeau, M., Sargeant, T. & Sauvageau, G. Haploinsufficiency screen highlights two distinct groups of ribosomal protein genes essential for embryonic stem cell fate. Proc. Natl Acad. Sci. USA 112, 2127–2132 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. 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). Ingolia et al. (2011) perform ribosome profiling for the first time in a stem cell lineage to reveal that the translational efficiency of many genes is upregulated during ESC differentiation through regulation by uORFs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Blair, J. D., Hockemeyer, D., Doudna, J. A., Bateup, H. S. & Floor, S. N. Widespread translational remodeling during human neuronal differentiation. Cell Rep. 21, 2005–2016 (2017). Blair et al. (2017) perform transcript-isoform analysis that demonstrates that 3′ UTR regulation of translation increases as ESCs differentiate to neurons.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Notingher, I. et al. In situ spectral monitoring of mRNA translation in embryonic stem cells during differentiation in vitro. Anal. Chem. 76, 3185–3193 (2004).

    Article  CAS  PubMed  Google Scholar 

  19. Easley, C. A. 4th et al. mTOR-mediated activation of p70 S6K induces differentiation of pluripotent human embryonic stem cells. Cell. Reprogram. 12, 263–273 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Tan, S. M. et al. Divergent LIN28-mRNA associations result in translational suppression upon the initiation of differentiation. Nucleic Acids Res. 42, 7997–8007 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Signer, R. A. J., Magee, J. A., Salic, A. & Morrison, S. J. Haematopoietic stem cells require a highly regulated protein synthesis rate. Nature 509, 49–54 (2014). This is one of the early studies to demonstrate that global protein synthesis rates are low in a stem cell lineage (HSCs) and that tight regulation of protein synthesis is required for HSC engraftment and prevention of cancer.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Jarzebowski, L. et al. Mouse adult hematopoietic stem cells actively synthesize ribosomal RNA. RNA 24, 1803–1812 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Signer, R. A. J. et al. The rate of protein synthesis in hematopoietic stem cells is limited partly by 4E-BPs. Genes Dev. 30, 1698–1703 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Blanco, S. et al. Stem cell function and stress response are controlled by protein synthesis. Nature 534, 335–340 (2016). Blanco et al. (2016) show that protein synthesis rates are low in epidermal hair follicle stem cells and that this regulation occurs through tRFs, which when dysregulated can contribute to tumorigenesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Liakath-Ali, K. et al. An evolutionarily conserved ribosome-rescue pathway maintains epidermal homeostasis. Nature 556, 376–380 (2018). Liakath-Ali et al. (2018) are the first to demonstrate the role of the ribosome rescue factor Pelota in an adult stem cell lineage.

    Article  CAS  PubMed  Google Scholar 

  26. Baser, A. et al. Onset of differentiation is post-transcriptionally controlled in adult neural stem cells. Nature 566, 100–104 (2019). Baser et al. (2019) demonstrate dynamic protein synthesis rates over the course of neural differentiation.

    Article  CAS  PubMed  Google Scholar 

  27. Zismanov, V. et al. Phosphorylation of eIF2α is a translational control mechanism regulating muscle stem cell quiescence and self-renewal. Cell Stem Cell 18, 79–90 (2016).

    Article  CAS  PubMed  Google Scholar 

  28. Sanchez, C. G. et al. Regulation of ribosome biogenesis and protein synthesis controls germline stem cell differentiation. Cell Stem Cell 18, 276–290 (2016).

    Article  CAS  PubMed  Google Scholar 

  29. Baßler, J. & Hurt, E. Eukaryotic ribosome assembly. Annu. Rev. Biochem. 88, 281–306 (2019).

    Article  PubMed  CAS  Google Scholar 

  30. Zaidi, S. K. et al. Expression of ribosomal RNA and protein genes in human embryonic stem cells is associated with the activating H3K4me3 histone mark. J. Cell. Physiol. 231, 2007–2013 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Cartwright, P. et al. LIF/STAT3 controls ES cell self-renewal and pluripotency by a Myc-dependent mechanism. Development 132, 885–896 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Watanabe-Susaki, K. et al. Biosynthesis of ribosomal RNA in nucleoli regulates pluripotency and differentiation ability of pluripotent stem cells. Stem Cell 32, 3099–3111 (2014).

    Article  CAS  Google Scholar 

  33. Corsini, N. S. et al. Coordinated control of mRNA and rRNA processing controls embryonic stem cell pluripotency and differentiation. Cell Stem Cell 22, 543–558.e12 (2018). Corsini et al. (2018) reveal dynamic protein synthesis rates and RiBi as ESCs differentiate to EBs regulated by the factor HTATSF1.

    Article  CAS  PubMed  Google Scholar 

  34. Saez, I. et al. The E3 ubiquitin ligase UBR5 interacts with the H/ACA ribonucleoprotein complex and regulates ribosomal RNA biogenesis in embryonic stem cells. FEBS Lett. 594, 175–188 (2020).

    Article  CAS  PubMed  Google Scholar 

  35. Salomon-Kent, R. et al. New face for chromatin-related mesenchymal modulator: n-CHD9 localizes to nucleoli and interacts with ribosomal genes. J. Cell. Physiol. 230, 2270–2280 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. van Riggelen, J., Yetil, A. & Felsher, D. W. MYC as a regulator of ribosome biogenesis and protein synthesis. Nat. Rev. Cancer 10, 301–309 (2010).

    Article  PubMed  CAS  Google Scholar 

  37. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Chappell, J. & Dalton, S. Roles for MYC in the establishment and maintenance of pluripotency. Cold Spring Harb. Perspect. Med. 3, a014381 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Herrlinger, S. et al. Lin28-mediated temporal promotion of protein synthesis is crucial for neural progenitor cell maintenance and brain development in mice. Development 146, dev173765 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Peng, S. et al. Genome-wide studies reveal that Lin28 enhances the translation of genes important for growth and survival of human embryonic stem cells. Stem Cells 29, 496–504 (2011).

    Article  CAS  PubMed  Google Scholar 

  41. Zhang, Q., Shalaby, N. A. & Buszczak, M. Changes in rRNA transcription influence proliferation and cell fate within a stem cell lineage. Science 343, 298–301 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Stedman, A. et al. Ribosome biogenesis dysfunction leads to p53-mediated apoptosis and goblet cell differentiation of mouse intestinal stem/progenitor cells. Cell Death Differ. 22, 1865–1876 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Le Bouteiller, M. et al. Notchless-dependent ribosome synthesis is required for the maintenance of adult hematopoietic stem cells. J. Exp. Med. 210, 2351–2369 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Cai, X. et al. Runx1 deficiency decreases ribosome biogenesis and confers stress resistance to hematopoietic stem and progenitor cells. Cell Stem Cell 17, 165–177 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Pereira, I. T. et al. Cardiomyogenic differentiation is fine-tuned by differential mRNA association with polysomes. BMC Genomics 20, 219 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  46. de Klerk, E. et al. Assessing the translational landscape of myogenic differentiation by ribosome profiling. Nucleic Acids Res. 43, 4408–4428 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Marcon, B. H. et al. Downregulation of the protein synthesis machinery is a major regulatory event during early adipogenic differentiation of human adipose-derived stromal cells. Stem Cell Res. 25, 191–201 (2017).

    Article  CAS  PubMed  Google Scholar 

  48. Thomson, J. A. Embryonic stem cell lines derived from human blastocysts. Science 282, 1145–1147 (1998).

    Article  CAS  PubMed  Google Scholar 

  49. Liu, G. Y. & Sabatini, D. M. mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol. 21, 183–203 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Thoreen, C. C. et al. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485, 109–113 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Hsieh, A. C. et al. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485, 55–61 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Morita, M. et al. mTORC1 controls mitochondrial activity and biogenesis through 4E-BP-dependent translational regulation. Cell Metab. 18, 698–711 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Larsson, O. et al. Distinct perturbation of the translatome by the antidiabetic drug metformin. Proc. Natl Acad. Sci. USA 109, 8977–8982 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Liu, X. et al. Regulation of mitochondrial biogenesis in erythropoiesis by mTORC1-mediated protein translation. Nat. Cell Biol. 19, 626–638 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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 

  57. 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 

  58. Michels, A. A. et al. mTORC1 directly phosphorylates and regulates human MAF1. Mol. Cell. Biol. 30, 3749–3757 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Moustafa-Kamal, M. et al. The mTORC1/S6K/PDCD4/eIF4A axis determines outcome of mitotic arrest. Cell Rep. 33, 108230 (2020).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Wang, X. et al. Regulation of elongation factor 2 kinase by p90RSK1 and p70 S6 kinase. EMBO J. 20, 4370–4379 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Chauvin, C. et al. Ribosomal protein S6 kinase activity controls the ribosome biogenesis transcriptional program. Oncogene 33, 474–483 (2014).

    Article  CAS  PubMed  Google Scholar 

  63. Meyuhas, O. Ribosomal protein S6 phosphorylation: four decades of research. Int. Rev. Cell Mol. Biol. 320, 41–73 (2015).

    Article  CAS  PubMed  Google Scholar 

  64. Raught, B. et al. Serum-stimulated, rapamycin-sensitive phosphorylation sites in the eukaryotic translation initiation factor 4GI. EMBO J. 19, 434–444 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Gangloff, Y.-G. et al. Disruption of the mouse mTOR gene leads to early postimplantation lethality and prohibits embryonic stem cell development. Mol. Cell. Biol. 24, 9508–9516 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Hentges, K. E. et al. FRAP/mTOR is required for proliferation and patterning during embryonic development in the mouse. Proc. Natl Acad. Sci. USA 98, 13796–13801 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Murakami, M. et al. mTOR is essential for growth and proliferation in early mouse embryos and embryonic stem cells. Mol. Cell. Biol. 24, 6710–6718 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Jansova, D. et al. Regulation of 4E-BP1 activity in the mammalian oocyte. Cell Cycle 16, 927–939 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Guo, J. et al. Oocyte stage-specific effects of MTOR determine granulosa cell fate and oocyte quality in mice. Proc. Natl Acad. Sci. USA 115, E5326–E5333 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Bulut-Karslioglu, A. et al. Inhibition of mTOR induces a paused pluripotent state. Nature 540, 119–123 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Hartman, N. W. et al. mTORC1 targets the translational repressor 4E-BP2, but not S6 kinase 1/2, to regulate neural stem cell self-renewal in vivo. Cell Rep. 5, 433–444 (2013).

    Article  CAS  PubMed  Google Scholar 

  72. Arvidsson, A., Collin, T., Kirik, D., Kokaia, Z. & Lindvall, O. Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat. Med. 8, 963–970 (2002).

    Article  CAS  PubMed  Google Scholar 

  73. Llorens-Bobadilla, E. et al. Single-cell transcriptomics reveals a population of dormant neural stem cells that become activated upon brain injury. Cell Stem Cell 17, 329–340 (2015).

    Article  CAS  PubMed  Google Scholar 

  74. Rodgers, J. T. et al. mTORC1 controls the adaptive transition of quiescent stem cells from G 0 to G Alert. Nature 510, 393–396 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Ding, X. et al. mTORC1 and mTORC2 regulate skin morphogenesis and epidermal barrier formation. Nat. Commun. 7, 13226 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Ding, X. et al. Epidermal mammalian target of rapamycin complex 2 controls lipid synthesis and filaggrin processing in epidermal barrier formation. J. Allergy Clin. Immunol. 145, 283–300.e3 (2020).

    Article  CAS  PubMed  Google Scholar 

  77. He, J. et al. An elaborate regulation of mammalian target of rapamycin activity is required for somatic cell reprogramming induced by defined transcription factors. Stem Cell Dev. 21, 2630–2641 (2012).

    Article  CAS  Google Scholar 

  78. Haller, S. et al. mTORC1 activation during repeated regeneration impairs somatic stem cell maintenance. Cell Stem Cell 21, 806–818.e5 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Fingar, D. C. et al. mTOR controls cell cycle progression through its cell growth effectors S6K1 and 4E-BP1/eukaryotic translation initiation factor 4E. Mol. Cell. Biol. 24, 200–216 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Fingar, D. C., Salama, S., Tsou, C., Harlow, E. & Blenis, J. Mammalian cell size is controlled by mTOR and its downstream targets S6K1 and 4EBP1/eIF4E. Genes Dev. 16, 1472–1487 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Cho, J. et al. LIN28A is a suppressor of ER-associated translation in embryonic stem cells. Cell 151, 765–777 (2012).

    Article  CAS  PubMed  Google Scholar 

  82. Magee, J. A. & Signer, R. A. J. Developmental stage-specific changes in protein synthesis differentially sensitize hematopoietic stem cells and erythroid progenitors to impaired ribosome biogenesis. Stem Cell Rep. 16, 20–28 (2021).

    Article  CAS  Google Scholar 

  83. Copley, M. R. et al. The Lin28b–let-7–Hmga2 axis determines the higher self-renewal potential of fetal haematopoietic stem cells. Nat. Cell Biol. 15, 916–925 (2013).

    Article  CAS  PubMed  Google Scholar 

  84. Basak, A. et al. Control of human hemoglobin switching by LIN28B-mediated regulation of BCL11A translation. Nat. Genet. 52, 138–145 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Yuan, J., Nguyen, C. K., Liu, X., Kanellopoulou, C. & Muljo, S. A. Lin28b reprograms adult bone marrow hematopoietic progenitors to mediate fetal-like lymphopoiesis. Science 335, 1195–1200 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Wang, S. et al. Transformation of the intestinal epithelium by the MSI2 RNA-binding protein. Nat. Commun. 6, 6517 (2015).

    Article  CAS  PubMed  Google Scholar 

  87. 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 

  88. Vu, L. P. et al. Functional screen of MSI2 interactors identifies an essential role for SYNCRIP in myeloid leukemia stem cells. Nat. Genet. 49, 866–875 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Ye, J. & Blelloch, R. Regulation of pluripotency by RNA binding proteins. Cell Stem Cell 15, 271–280 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Hung, S. S. C. et al. Repression of global protein synthesis by Eif1a-like genes that are expressed specifically in the two-cell embryos and the transient Zscan4-positive state of embryonic stem cells. DNA Res. 20, 391–402 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Buckley, S. M. et al. Regulation of pluripotency and cellular reprogramming by the ubiquitin-proteasome system. Cell Stem Cell 11, 783–798 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Chen, Z. Y., Wang, X., Zhou, Y., Offner, G. & Tseng, C.-C. Destabilization of Krüppel-like factor 4 protein in response to serum stimulation involves the ubiquitin-proteasome pathway. Cancer Res. 65, 10394–10400 (2005).

    Article  CAS  PubMed  Google Scholar 

  93. Vilchez, D. et al. Increased proteasome activity in human embryonic stem cells is regulated by PSMD11. Nature 489, 304–308 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Koyuncu, S. et al. The ubiquitin ligase UBR5 suppresses proteostasis collapse in pluripotent stem cells from Huntington’s disease patients. Nat. Commun. 9, 2886 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  95. Noormohammadi, A. et al. Mechanisms of protein homeostasis (proteostasis) maintain stem cell identity in mammalian pluripotent stem cells. Cell. Mol. Life Sci. 75, 275–290 (2018).

    Article  CAS  PubMed  Google Scholar 

  96. Cui, C.-P. et al. Dynamic ubiquitylation of Sox2 regulates proteostasis and governs neural progenitor cell differentiation. Nat. Commun. 9, 4648 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  97. Nguyen, A. T. et al. UBE2O remodels the proteome during terminal erythroid differentiation. Science 357, eaan0218 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  98. Dez, C. & Tollervey, D. Ribosome synthesis meets the cell cycle. Curr. Opin. Microbiol. 7, 631–637 (2004).

    Article  CAS  PubMed  Google Scholar 

  99. Cheng, Z. et al. Small and large ribosomal subunit deficiencies lead to distinct gene expression signatures that reflect cellular growth rate. Mol. Cell 73, 36–47.e10 (2019).

    Article  CAS  PubMed  Google Scholar 

  100. Grollman, A. P. Cytotoxic inhibitors of protein synthesis. in Antineoplastic and Immunosuppressive Agents: Part II (eds Sartorelli, A. C. & Johns, D. G.) 554–570 (Springer, 1975).

  101. Morgado-Palacin, L., Llanos, S. & Serrano, M. Ribosomal stress induces L11- and p53-dependent apoptosis in mouse pluripotent stem cells. Cell Cycle 11, 503–510 (2012).

    Article  CAS  PubMed  Google Scholar 

  102. Zambetti, N. A. et al. Deficiency of the ribosome biogenesis gene Sbds in hematopoietic stem and progenitor cells causes neutropenia in mice by attenuating lineage progression in myelocytes. Haematologica 100, 1285–1293 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Garçon, L. et al. Ribosomal and hematopoietic defects in induced pluripotent stem cells derived from Diamond Blackfan anemia patients. Blood 122, 912–921 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  104. Ludwig, L. S. et al. Altered translation of GATA1 in Diamond-Blackfan anemia. Nat. Med. 20, 748–753 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Khajuria, R. K. et al. Ribosome levels selectively regulate translation and lineage commitment in human hematopoiesis. Cell 173, 90–103.e19 (2018). Khajuria et al. (2018) demonstrate that ribosome levels are essential for proper stem cell differentiation and lineage commitment in the haematopoietic system.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Hesling, C., Oliveira, C. C., Castilho, B. A. & Zanchin, N. I. T. The Shwachman–Bodian–Diamond syndrome associated protein interacts with HsNip7 and its down-regulation affects gene expression at the transcriptional and translational levels. Exp. Cell Res. 313, 4180–4195 (2007).

    Article  CAS  PubMed  Google Scholar 

  107. Sinha, N. K. et al. EDF1 coordinates cellular responses to ribosome collisions. eLife 9, e58828 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Wu, C. C.-C., Peterson, A., Zinshteyn, B., Regot, S. & Green, R. Ribosome collisions trigger general stress responses to regulate cell fate. Cell 182, 404–416.e14 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Yilmaz, O. H. et al. Pten dependence distinguishes haematopoietic stem cells from leukaemia-initiating cells. Nature 441, 475–482 (2006).

    Article  CAS  PubMed  Google Scholar 

  110. Elkenani, M. et al. Pelota regulates epidermal differentiation by modulating BMP and PI3K/AKT signaling pathways. J. Invest. Dermatol. 136, 1664–1671 (2016).

    Article  CAS  PubMed  Google Scholar 

  111. de Klerk, E. & ’t Hoen, P. A. C. Alternative mRNA transcription, processing, and translation: insights from RNA sequencing. Trends Genet. 31, 128–139 (2015).

    Article  PubMed  CAS  Google Scholar 

  112. Wong, Q. W.-L. et al. Embryonic stem cells exhibit mRNA isoform specific translational regulation. PLoS ONE 11, e0143235 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Tahmasebi, S. et al. Control of embryonic stem cell self-renewal and differentiation via coordinated alternative splicing and translation of YY2. Proc. Natl Acad. Sci. USA 113, 12360–12367 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Tahmasebi, S. et al. Multifaceted regulation of somatic cell reprogramming by mRNA translational control. Cell Stem Cell 14, 606–616 (2014).

    Article  CAS  PubMed  Google Scholar 

  115. 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 

  116. Garcia-Garcia, C., Frieda, K. L., Feoktistova, K., Fraser, C. S. & Block, S. M. Factor-dependent processivity in human eIF4A DEAD-box helicase. Science 348, 1486–1488 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Wolfe, A. L. et al. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513, 65–70 (2014). Wolfe et al. (2014) identify distinct translational signatures for eIF4A-regulated transcripts compared with mTOR/4E-BP/eIF4E-regulated transcripts in a cancer model system.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Nguyen, T. M. et al. FGFR1-activated translation of WNT pathway components with structured 5′ UTRs is vulnerable to inhibition of eif4a-dependent translation initiation. Cancer Res. 78, 4229–4240 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Kerr, C. L., Bol, G. M., Vesuna, F. & Raman, V. Targeting RNA helicase DDX3 in stem cell maintenance and teratoma formation. Genes Cancer 10, 11–20 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Lai, M.-C., Chang, W.-C., Shieh, S.-Y. & Tarn, W.-Y. DDX3 regulates cell growth through translational control of cyclin E1. Mol. Cell. Biol. 30, 5444–5453 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Soto-Rifo, R. et al. DEAD-box protein DDX3 associates with eIF4F to promote translation of selected mRNAs. EMBO J. 31, 3745–3756 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Truitt, M. L. et al. Differential requirements for eIF4E dose in normal development and cancer. Cell 162, 59–71 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Sugiyama, H. et al. Nat1 promotes translation of specific proteins that induce differentiation of mouse embryonic stem cells. Proc. Natl Acad. Sci. USA 114, 340–345 (2017).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. de la Parra, C. et al. A widespread alternate form of cap-dependent mRNA translation initiation. Nat. Commun. 9, 3068 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  128. Yang, G., Smibert, C. A., Kaplan, D. R. & Miller, F. D. An eIF4E1/4E-T complex determines the genesis of neurons from precursors by translationally repressing a proneurogenic transcription program. Neuron 84, 723–739 (2014).

    Article  CAS  PubMed  Google Scholar 

  129. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Bazzini, A. A. et al. Codon identity regulates mRNA stability and translation efficiency during the maternal-to-zygotic transition. EMBO J. 35, 2087–2103 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Morris, D. R. & Geballe, A. P. Upstream open reading frames as regulators of mRNA translation. Mol. Cell. Biol. 20, 8635–8642 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Starck, S. R. et al. Translation from the 5′untranslated region shapes the integrated stress response. Science 351, aad3867 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  133. Pakos-Zebrucka, K. et al. The integrated stress response. EMBO Rep. 17, 1374–1395 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Hinnebusch, A. G. Translational regulation of yeast GCN4. A window on factors that control initiator-tRNA binding to the ribosome. J. Biol. Chem. 272, 21661–21664 (1997).

    Article  CAS  PubMed  Google Scholar 

  135. 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 

  136. Popa, A., Lebrigand, K., Barbry, P. & Waldmann, R. Pateamine A-sensitive ribosome profiling reveals the scope of translation in mouse embryonic stem cells. BMC Genomics 17, 52 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  137. Fujii, K., Shi, Z., Zhulyn, O., Denans, N. & Barna, M. Pervasive translational regulation of the cell signalling circuitry underlies mammalian development. Nat. Commun. 8, 14443 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Paolini, N. A. et al. Ribosome profiling uncovers selective mRNA translation associated with eIF2 phosphorylation in erythroid progenitors. PLoS ONE 13, e0193790 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Friend, K., Brooks, H. A., Propson, N. E., Thomson, J. A. & Kimble, J. Embryonic stem cell growth factors regulate eIF2α phosphorylation. PLoS ONE 10, e0139076 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Chen, J. et al. Pervasive functional translation of noncanonical human open reading frames. Science 367, 1140–1146 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  142. Floor, S. N. & Doudna, J. A. Tunable protein synthesis by transcript isoforms in human cells. eLife 5, e10921 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  143. 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 

  144. 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 

  145. Ji, Z., Lee, J. Y., Pan, Z., Jiang, B. & Tian, B. Progressive lengthening of 3′ untranslated regions of mRNAs by alternative polyadenylation during mouse embryonic development. Proc. Natl Acad. Sci. USA 106, 7028–7033 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Göpferich, M. et al. Single cell 3′UTR analysis identifies changes in alternative polyadenylation throughout neuronal differentiation and in autism. Preprint at bioRxiv https://doi.org/10.1101/2020.08.12.247627 (2020).

  147. Wang, Y., Arribas-Layton, M., Chen, Y., Lykke-Andersen, J. & Sen, G. L. DDX6 orchestrates mammalian progenitor function through the mRNA degradation and translation pathways. Mol. Cell 60, 118–130 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Boutet, S. C. et al. Alternative polyadenylation mediates microRNA regulation of muscle stem cell function. Cell Stem Cell 10, 327–336 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Spangenberg, L. et al. Polysome profiling shows extensive posttranscriptional regulation during human adipocyte stem cell differentiation into adipocytes. Stem Cell Res. 11, 902–912 (2013).

    Article  CAS  PubMed  Google Scholar 

  150. Lodish, H. F. Model for the regulation of mRNA translation applied to haemoglobin synthesis. Nature 251, 385–388 (1974).

    Article  CAS  PubMed  Google Scholar 

  151. Mills, E. W. & Green, R. Ribosomopathies: there’s strength in numbers. Science 358, eaan2755 (2017).

    Article  PubMed  CAS  Google Scholar 

  152. Chennupati, V. et al. Ribonuclease inhibitor 1 regulates erythropoiesis by controlling GATA1 translation. J. Clin. Invest. 128, 1597–1614 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  153. Popis, M. C., Blanco, S. & Frye, M. Posttranscriptional methylation of transfer and ribosomal RNA in stress response pathways, cell differentiation, and cancer. Curr. Opin. Oncol. 28, 65–71 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Bornelöv, S., Selmi, T., Flad, S., Dietmann, S. & Frye, M. Codon usage optimization in pluripotent embryonic stem cells. Genome Biol. 20, 119 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  155. Presnyak, V. et al. Codon optimality is a major determinant of mRNA stability. Cell 160, 1111–1124 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Radhakrishnan, A. et al. The DEAD-box protein Dhh1p couples mRNA decay and translation by monitoring codon optimality. Cell 167, 122–132.e9 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Goncalves, K. A. et al. Angiogenin promotes hematopoietic regeneration by dichotomously regulating quiescence of stem and progenitor cells. Cell 166, 894–906 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Guzzi, N. et al. Pseudouridylation of tRNA-derived fragments steers translational control in stem cells. Cell 173, 1204–1216.e26 (2018). Guzzi et al. (2018) show that tRF generation in ESCs contributes to the inhibition of global protein synthesis.

    Article  CAS  PubMed  Google Scholar 

  159. Gingold, H. et al. A dual program for translation regulation in cellular proliferation and differentiation. Cell 158, 1281–1292 (2014).

    Article  CAS  PubMed  Google Scholar 

  160. Lin, S. et al. Mettl1/Wdr4-mediated m7G tRNA methylome is required for normal mRNA translation and embryonic stem cell self-renewal and differentiation. Mol. Cell 71, 244–255.e5 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Tuorto, F. et al. The tRNA methyltransferase Dnmt2 is required for accurate polypeptide synthesis during haematopoiesis. EMBO J. 34, 2350–2362 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Su, Z., Kuscu, C., Malik, A., Shibata, E. & Dutta, A. Angiogenin generates specific stress-induced tRNA halves and is not involved in tRF-3-mediated gene silencing. J. Biol. Chem. 294, 16930–16941 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Yamasaki, S., Ivanov, P., Hu, G.-F. & Anderson, P. Angiogenin cleaves tRNA and promotes stress-induced translational repression. J. Cell Biol. 185, 35–42 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Ivanov, P., Emara, M. M., Villen, J., Gygi, S. P. & Anderson, P. Angiogenin-induced tRNA fragments inhibit translation initiation. Mol. Cell 43, 613–623 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Blanco, S. et al. The RNA-methyltransferase Misu (NSun2) poises epidermal stem cells to differentiate. PLoS Genet. 7, e1002403 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Hussain, S. et al. The mouse cytosine-5 RNA methyltransferase NSun2 is a component of the chromatoid body and required for testis differentiation. Mol. Cell. Biol. 33, 1561–1570 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Kahvejian, A., Roy, G. & Sonenberg, N. The mRNA closed-loop model: the function of PABP and PABP-interacting proteins in mRNA translation. Cold Spring Harb. Symp. Quant. Biol. 66, 293–300 (2001).

    Article  CAS  PubMed  Google Scholar 

  168. Pinkard, O., McFarland, S., Sweet, T. & Coller, J. Quantitative tRNA-sequencing uncovers metazoan tissue-specific tRNA regulation. Nat. Commun. 11, 4104 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Xu, Y. & Ruggero, D. The Role of translation control in tumorigenesis and its therapeutic implications. Annu. Rev. Cancer Biol. 4, 437–457 (2020).

    Article  Google Scholar 

  170. Robichaud, N., Sonenberg, N., Ruggero, D. & Schneider, R. J. Translational control in cancer. Cold Spring Harb. Perspect. Biol. 11, 254–266 (2019).

    Article  CAS  Google Scholar 

  171. Morral, C. et al. Zonation of ribosomal DNA transcription defines a stem cell hierarchy in colorectal cancer. Cell Stem Cell 26, 845–861.e12 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Chen, L. et al. Circ-MALAT1 functions as both an mRNA translation brake and a microRNA sponge to promote self-renewal of hepatocellular cancer stem cells. Adv. Sci. 7, 1900949 (2020).

    Article  CAS  Google Scholar 

  173. Meacham, C. E. & Morrison, S. J. Tumour heterogeneity and cancer cell plasticity. Nature 501, 328–337 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Lim, S. et al. Targeting of the MNK-eIF4E axis in blast crisis chronic myeloid leukemia inhibits leukemia stem cell function. Proc. Natl Acad. Sci. USA 110, E2298–E2307 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Jögi, A., Vaapil, M., Johansson, M. & Påhlman, S. Cancer cell differentiation heterogeneity and aggressive behavior in solid tumors. Ups. J. Med. Sci. 117, 217–224 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  176. de Thé, H. Differentiation therapy revisited. Nat. Rev. Cancer 18, 117–127 (2018).

    Article  PubMed  CAS  Google Scholar 

  177. Enane, F. O., Saunthararajah, Y. & Korc, M. Differentiation therapy and the mechanisms that terminate cancer cell proliferation without harming normal cells. Cell Death Dis. 9, 912 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  178. Sendoel, A. et al. Translation from unconventional 5′ start sites drives tumour initiation. Nature 541, 494–499 (2017). Sendoel et al. (2017) show that SOX2 induction in the embryonic epidermis drives tumorigenesis through inhibition of eIF2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Zhang, S., Xiong, X. & Sun, Y. Functional characterization of SOX2 as an anticancer target. Signal. Transduct. Target. Ther. 5, 135 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Nguyen, H. G. et al. Development of a stress response therapy targeting aggressive prostate cancer. Sci. Transl. Med. 10, eaar2036 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  181. Manier, S. et al. Inhibiting the oncogenic translation program is an effective therapeutic strategy in multiple myeloma. Sci. Transl. Med. 9, eaal2668 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  182. 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 

  183. Rosenwald, I. B., Lazaris-Karatzas, A., Sonenberg, N. & Schmidt, E. V. Elevated levels of cyclin D1 protein in response to increased expression of eukaryotic initiation factor 4E. Mol. Cell. Biol. 13, 7358–7363 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Suresh, S. et al. eIF5B drives integrated stress response-dependent translation of PD-L1 in lung cancer. Nat. Cancer 1, 533–545 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  185. Xu, Y. et al. Translation control of the immune checkpoint in cancer and its therapeutic targeting. Nat. Med. 25, 301–311 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Vazquez-Arango, P. et al. Variant U1 snRNAs are implicated in human pluripotent stem cell maintenance and neuromuscular disease. Nucleic Acids Res. 44, 10960–10973 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Oh, J.-M. et al. U1 snRNP regulates cancer cell migration and invasion in vitro. Nat. Commun. 11, 1 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Fares, J., Fares, M. Y., Khachfe, H. H., Salhab, H. A. & Fares, Y. Molecular principles of metastasis: a hallmark of cancer revisited. Signal. Transduct. Target. Ther. 5, 28 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  189. Yuan, S., Norgard, R. J. & Stanger, B. Z. Cellular plasticity in cancer. Cancer Discov. 9, 837–851 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Feng, Y.-X. et al. Epithelial-to-mesenchymal transition activates PERK–eIF2α and sensitizes cells to endoplasmic reticulum stress. Cancer Discov. 4, 702–715 (2014).

    Article  CAS  PubMed  Google Scholar 

  191. Goodarzi, H. et al. Modulated expression of specific tRNAs drives gene expression and cancer progression. Cell 165, 1416–1427 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Ebright, R. Y. et al. Deregulation of ribosomal protein expression and translation promotes breast cancer metastasis. Science 367, 1468–1473 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. van Heesch, S. et al. The translational landscape of the human heart. Cell 178, 242–260.e29 (2019).

    Article  PubMed  CAS  Google Scholar 

  194. Guttman, M. et al. lincRNAs act in the circuitry controlling pluripotency and differentiation. Nature 477, 295–300 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Gong, C. et al. A long non-coding RNA, LncMyoD, regulates skeletal muscle differentiation by blocking IMP2-mediated mRNA translation. Dev. Cell 34, 181–191 (2015).

    Article  CAS  PubMed  Google Scholar 

  196. Batista, P. J. et al. m6A RNA modification controls cell fate transition in mammalian embryonic stem cells. Cell Stem Cell 15, 707–719 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Vu, L. P. et al. The N6-methyladenosine (m6A)-forming enzyme METTL3 controls myeloid differentiation of normal hematopoietic and leukemia cells. Nat. Med. 23, 1369–1376 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Lin, S., Choe, J., Du, P., Triboulet, R. & Gregory, R. I. The m6A methyltransferase METTL3 promotes translation in human cancer cells. Mol. Cell 62, 335–345 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Shi, H. et al. YTHDF3 facilitates translation and decay of N6-methyladenosine-modified RNA. Cell Res. 27, 315–328 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Choe, J. et al. mRNA circularization by METTL3-eIF3h enhances translation and promotes oncogenesis. Nature 561, 556–560 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Ishimura, R., Nagy, G., Dotu, I., Chuang, J. H. & Ackerman, S. L. Activation of GCN2 kinase by ribosome stalling links translation elongation with translation initiation. eLife 5, e14295 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  202. Mahdessian, D. et al. Spatiotemporal dissection of the cell cycle with single-cell proteogenomics. Nature 590, 649–654 (2021).

    Article  CAS  PubMed  Google Scholar 

  203. Regot, S., Hughey, J. J., Bajar, B. T., Carrasco, S. & Covert, M. W. High-sensitivity measurements of multiple kinase activities in live single cells. Cell 157, 1724–1734 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Aikin, T. J., Peterson, A. F., Pokrass, M. J., Clark, H. R. & Regot, S. MAPK activity dynamics regulate non-cell autonomous effects of oncogene expression. eLife 9, e60541 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Lewandowski, B. et al. Sequence-specific peptide synthesis by an artificial small-molecule machine. Science 339, 189–193 (2013).

    Article  CAS  PubMed  Google Scholar 

  206. O’Connell, A. E. et al. Mammalian Hbs1L deficiency causes congenital anomalies and developmental delay associated with Pelota depletion and 80S monosome accumulation. PLoS Genet. 15, e1007917 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  207. Posfai, E. et al. Evaluating totipotency using criteria of increasing stringency. Nat. Cell Biol. 23, 49–60 (2021).

    Article  CAS  PubMed  Google Scholar 

  208. De Paepe, C., Krivega, M., Cauffman, G., Geens, M. & Van de Velde, H. Totipotency and lineage segregation in the human embryo. Mol. Hum. Reprod. 20, 599–618 (2014).

    Article  PubMed  Google Scholar 

  209. Davidson, K. C., Mason, E. A. & Pera, M. F. The pluripotent state in mouse and human. Development 142, 3090–3099 (2015).

    Article  CAS  PubMed  Google Scholar 

  210. Weinberger, L., Ayyash, M., Novershtern, N. & Hanna, J. H. Dynamic stem cell states: naive to primed pluripotency in rodents and humans. Nat. Rev. Mol. Cell Biol. 17, 155–169 (2016).

    Article  CAS  PubMed  Google Scholar 

  211. Neagu, A. et al. In vitro capture and characterization of embryonic rosette-stage pluripotency between naive and primed states. Nat. Cell Biol. 22, 534–545 (2020).

    Article  CAS  PubMed  Google Scholar 

  212. Ying, Q.-L. et al. The ground state of embryonic stem cell self-renewal. Nature 453, 519–523 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Huang, J. et al. Pivotal role for glycogen synthase kinase-3 in hematopoietic stem cell homeostasis in mice. J. Clin. Invest. 119, 3519–3529 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  214. Mendoza, M. C., Er, E. E. & Blenis, J. The Ras-ERK and PI3K-mTOR pathways: cross-talk and compensation. Trends Biochem. Sci. 36, 320–328 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Hu, Z. et al. Transient inhibition of mTOR in human pluripotent stem cells enables robust formation of mouse-human chimeric embryos. Sci. Adv. 6, eaaz0298 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Scognamiglio, R. et al. Myc depletion induces a pluripotent dormant state mimicking diapause. Cell 164, 668–680 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Marks, H. et al. The transcriptional and epigenomic foundations of ground state pluripotency. Cell 149, 590–604 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Pandolfini, L. et al. RISC-mediated control of selected chromatin regulators stabilizes ground state pluripotency of mouse embryonic stem cells. Genome Biol. 17, 94 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  219. Lane, S. W., Williams, D. A. & Watt, F. M. Modulating the stem cell niche for tissue regeneration. Nat. Biotechnol. 32, 795–803 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Vining, K. H. & Mooney, D. J. Mechanical forces direct stem cell behaviour in development and regeneration. Nat. Rev. Mol. Cell Biol. 18, 728–742 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Watt, F. M. & Huck, W. T. S. Role of the extracellular matrix in regulating stem cell fate. Nat. Rev. Mol. Cell Biol. 14, 467–473 (2013).

    Article  CAS  PubMed  Google Scholar 

  222. Fletcher, D. A. & Mullins, R. D. Cell mechanics and the cytoskeleton. Nature 463, 485–492 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Gross, S. R. & Kinzy, T. G. Improper organization of the actin cytoskeleton affects protein synthesis at initiation. Mol. Cell. Biol. 27, 1974–1989 (2007).

    Article  CAS  PubMed  Google Scholar 

  224. Stapulionis, R., Kolli, S. & Deutscher, M. P. Efficient mammalian protein synthesis requires an intact F-actin system. J. Biol. Chem. 272, 24980–24986 (1997).

    Article  CAS  PubMed  Google Scholar 

  225. Silva, R. C., Sattlegger, E. & Castilho, B. A. Perturbations in actin dynamics reconfigure protein complexes that modulate GCN2 activity and promote an eIF2 response. J. Cell Sci. 129, 4521–4533 (2016).

    CAS  PubMed  Google Scholar 

  226. Martin, K. C. & Ephrussi, A. mRNA localization: gene expression in the spatial dimension. Cell 136, 719–730 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Kim, S., Wong, P. & Coulombe, P. A. A keratin cytoskeletal protein regulates protein synthesis and epithelial cell growth. Nature 441, 362–365 (2006).

    Article  CAS  PubMed  Google Scholar 

  228. Coulombe, P. A. & Wong, P. Cytoplasmic intermediate filaments revealed as dynamic and multipurpose scaffolds. Nat. Cell Biol. 6, 699–706 (2004).

    Article  CAS  PubMed  Google Scholar 

  229. Kim, S. & Coulombe, P. A. Emerging role for the cytoskeleton as an organizer and regulator of translation. Nat. Rev. Mol. Cell Biol. 11, 75–81 (2010).

    Article  PubMed  CAS  Google Scholar 

  230. Chicurel, M. E., Singer, R. H., Meyer, C. J. & Ingber, D. E. Integrin binding and mechanical tension induce movement of mRNA and ribosomes to focal adhesions. Nature 392, 730–733 (1998).

    Article  CAS  PubMed  Google Scholar 

  231. Chung, J. & Kim, T. H. Integrin-dependent translational control: implication in cancer progression. Microsc. Res. Tech. 71, 380–386 (2008).

    Article  CAS  PubMed  Google Scholar 

  232. Willett, M., Pollard, H. J., Vlasak, M. & Morley, S. J. Localization of ribosomes and translation initiation factors to talin/β3-integrin-enriched adhesion complexes in spreading and migrating mammalian cells. Biol. Cell 102, 265–276 (2010).

    Article  CAS  PubMed  Google Scholar 

  233. Lian, I. et al. The role of YAP transcription coactivator in regulating stem cell self-renewal and differentiation. Genes Dev. 24, 1106–1118 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Walko, G. et al. A genome-wide screen identifies YAP/WBP2 interplay conferring growth advantage on human epidermal stem cells. Nat. Commun. 8, 14744 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Dupont, S. et al. Role of YAP/TAZ in mechanotransduction. Nature 474, 179–183 (2011).

    Article  CAS  PubMed  Google Scholar 

  236. Goodman, C. A. et al. Yes-associated protein is up-regulated by mechanical overload and is sufficient to induce skeletal muscle hypertrophy. FEBS Lett. 589, 1491–1497 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Judson, R. N. et al. The Hippo pathway member Yap plays a key role in influencing fate decisions in muscle satellite cells. J. Cell Sci. 125, 6009–6019 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  238. Watt, K. I. et al. The Hippo pathway effector YAP is a critical regulator of skeletal muscle fibre size. Nat. Commun. 6, 6048 (2015).

    Article  CAS  PubMed  Google Scholar 

  239. Park, Y.-Y. et al. Yes-associated protein 1 and transcriptional coactivator with PDZ-binding motif activate the mammalian target of rapamycin complex 1 pathway by regulating amino acid transporters in hepatocellular carcinoma. Hepatology 63, 159–172 (2016).

    Article  CAS  PubMed  Google Scholar 

  240. Hansen, C. G., Ng, Y. L. D., Lam, W.-L. M., Plouffe, S. W. & Guan, K.-L. The Hippo pathway effectors YAP and TAZ promote cell growth by modulating amino acid signaling to mTORC1. Cell Res. 25, 1299–1313 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. Tumaneng, K. et al. YAP mediates crosstalk between the Hippo and PI(3)K–TOR pathways by suppressing PTEN via miR-29. Nat. Cell Biol. 14, 1322–1329 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Ruwhof, C. & van der Laarse, A. Mechanical stress-induced cardiac hypertrophy: mechanisms and signal transduction pathways. Cardiovasc. Res. 47, 23–37 (2000).

    Article  CAS  PubMed  Google Scholar 

  243. Sherwood, D. J. et al. Differential regulation of MAP kinase, p70S6K, and Akt by contraction and insulin in rat skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 276, E870–E878 (1999).

    Article  CAS  Google Scholar 

  244. Sakamoto, K., Hirshman, M. F., Aschenbach, W. G. & Goodyear, L. J. Contraction regulation of Akt in rat skeletal muscle. J. Biol. Chem. 277, 11910–11917 (2002).

    Article  CAS  PubMed  Google Scholar 

  245. Sugimoto, A. et al. Piezo type mechanosensitive ion channel component 1 functions as a regulator of the cell fate determination of mesenchymal stem cells. Sci. Rep. 7, 17696 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  246. Del Mármol, J. I., Touhara, K. K., Croft, G. & MacKinnon, R. Piezo1 forms a slowly-inactivating mechanosensory channel in mouse embryonic stem cells. eLife 7, e33149 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  247. Pathak, M. M. et al. Stretch-activated ion channel Piezo1 directs lineage choice in human neural stem cells. Proc. Natl Acad. Sci. USA 111, 16148–16153 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Wang, L. et al. Mechanical sensing protein PIEZO1 regulates bone homeostasis via osteoblast-osteoclast crosstalk. Nat. Commun. 11, 282 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Han, Y. et al. Mechanosensitive ion channel Piezo1 promotes prostate cancer development through the activation of the Akt/mTOR pathway and acceleration of cell cycle. Int. J. Oncol. 55, 629–644 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Green and Watt laboratories for helpful advice and feedback.

Author information

Authors and Affiliations

Authors

Contributions

J.A.S., K.L.-A., R.G. and F.M.W. contributed to the conceptualization, writing and revisions of the manuscript. J.A.S. and K.L.-A. prepared the figures.

Corresponding authors

Correspondence to Rachel Green or Fiona M. Watt.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Molecular Cell Biology thanks Vijay Sankaran, Ana Martin-Villalba and Jianlong Wang for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Glossary

Self-renewal

A process by which stem cells divide and maintain stemness.

Quiescent

A reversible state in which cells exit the cell cycle but can re-enter it in response to stimuli such as injury to the tissue in which they reside.

Induced pluripotent stem cells

(iPSCs). Somatic cells that are reprogrammed by defined factors to acquire embryonic stem cell-like, pluripotent features.

Dedifferentiation

A process by which terminally or partially differentiated cells revert to less-differentiated cells within the same lineage.

Ribosome biogenesis

(RiBi). A concerted molecular process to build a ribosome involving more than 200 proteins and other factors.

tRNA

A small RNA molecule that links decoding of the mRNA to incorporation of the appropriate amino acid into the nascent peptide strand during translation elongation.

Polysome profiling

An experimental technique in which a cell lysate is separated by density gradient centrifugation and mRNAs associated with multiple ribosomes (polyribosomes or ‘polysomes’) can be fractionated; not to be confused with the sequencing method ribosome profiling.

H/ACA complex

A small nucleolar ribonucleoprotein complex involved in the pseudouridylation of ribosomal RNA.

Cap-dependent translation

Translation initiation by the binding of an mRNA cap protein, usually eukaryotic initiation factor 4E (eIF4E), to a modified guanine structure at the 5′ end of the mRNA. This interaction is required for canonical mRNA translation.

Cell lineages

The developmental paths of a tissue or organ from single or multiple cell types.

Ribosomopathies

Diseases caused by abnormalities in the structure and function of ribosomal proteins or genes encoding ribosomal RNA or genes involved in ribosome biogenesis.

Polysome

Multiple ribosomes loaded onto a single mRNA (‘polyribosome’).

Hypomorphic mutation

A mutation in a gene that confers less function than the wild-type copy of that gene but retains more function than a complete loss-of-function allele.

Ribosome rescue factor

A protein factor that is involved in dissociating the large and small ribosomal subunits of a ribosome stalled on an mRNA (for example, the Pelota–HBS1L complex).

Ribosome profiling

A method to deep sequence ribosome-protected mRNA fragments, also known as ribo-seq.

G-quadruplexes

RNA secondary structures canonically formed by the stacking of planar guanine tetrads and stabilized by Hoogsteen base pairing and a central cation.

Upstream open reading frames

(uORFs). Translatable open reading frame sequences within the 5′ untranslated region of an mRNA.

Integrated stress response

(IRS). An extensive intracellular signalling network activated in response to various stresses to maintain cellular homeostasis.

Codon usage

The use of one codon instead of another one to encode the same amino acid; two codons encoding the same amino acid may be recognized by different tRNAs.

Wobble position

The third nucleotide of a codon in which recognition by the cognate tRNA may occur by certain non-Watson–Crick base pairing.

Mitotic index

A measure of proliferating cells defined as the percentage of cells in mitosis. Used for tumour grading.

Epithelial-to-mesenchymal transition

A process by which epithelial cells lose cell polarity and cell–cell adhesion properties and become mesenchymal-like cells with increased migratory and invasive potential.

Transdifferentiation

The transformation of cells other than stem cells into a different cell type.

Synthetic ribosomes

Engineered artificial small molecules that can mimic ribosome function by synthesizing peptides in a sequence-specific manner.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Saba, J.A., Liakath-Ali, K., Green, R. et al. Translational control of stem cell function. Nat Rev Mol Cell Biol 22, 671–690 (2021). https://doi.org/10.1038/s41580-021-00386-2

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-021-00386-2

This article is cited by

Search

Quick links

Nature Briefing

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

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