Skip to main content
Log in

Hydroxylation and translational adaptation to stress: some answers lie beyond the STOP codon

  • Review
  • Published:
Cellular and Molecular Life Sciences Aims and scope Submit manuscript

Abstract

Regulation of protein synthesis contributes to maintenance of homeostasis and adaptation to environmental changes. mRNA translation is controlled at various levels including initiation, elongation and termination, through post-transcriptional/translational modifications of components of the protein synthesis machinery. Recently, protein and RNA hydroxylation have emerged as important enzymatic modifications of tRNAs, elongation and termination factors, as well as ribosomal proteins. These modifications enable a correct STOP codon recognition, ensuring translational fidelity. Recent studies are starting to show that STOP codon read-through is related to the ability of the cell to cope with different types of stress, such as oxidative and chemical insults, while correlations between defects in hydroxylation of protein synthesis components and STOP codon read-through are beginning to emerge. In this review we will discuss our current knowledge of protein synthesis regulation through hydroxylation of components of the translation machinery, with special focus on STOP codon recognition. We speculate on the possibility that programmed STOP codon read-through, modulated by hydroxylation of components of the protein synthesis machinery, is part of a concerted cellular response to stress.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  1. Aoyagi Y et al (1988) Energy cost of whole-body protein synthesis measured in vivo in chicks. Comp Biochem Physiol A Comp Physiol 91(4):765–768

    Article  CAS  PubMed  Google Scholar 

  2. Bock FJ, Todorova TT, Chang P (2015) RNA regulation by poly(ADP-Ribose) polymerases. Mol Cell 58(6):959–969

    Article  CAS  PubMed  Google Scholar 

  3. Kang SA et al (2013) mTORC1 phosphorylation sites encode their sensitivity to starvation and rapamycin. Science 341(6144):1236566

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Carroll AJ et al (2008) Analysis of the Arabidopsis cytosolic ribosome proteome provides detailed insights into its components and their post-translational modification. Mol Cell Proteomics 7(2):347–369

    Article  CAS  PubMed  Google Scholar 

  5. Kearse MG, Chen AS, Ware VC (2011) Expression of ribosomal protein L22e family members in Drosophila melanogaster: rpL22-like is differentially expressed and alternatively spliced. Nucleic Acids Res 39(7):2701–2716

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lopes AM et al (2010) The human RPS4 paralogue on Yq11.223 encodes a structurally conserved ribosomal protein and is preferentially expressed during spermatogenesis. BMC Mol Biol 11:33

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Hinnebusch AG (2014) The scanning mechanism of eukaryotic translation initiation. Annu Rev Biochem 83:779–812

    Article  CAS  PubMed  Google Scholar 

  8. Jackson RJ, Hellen CU, Pestova TV (2010) The mechanism of eukaryotic translation initiation and principles of its regulation. Nat Rev Mol Cell Biol 11(2):113–127

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sonenberg N, Hinnebusch AG (2009) Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136(4):731–745

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Rodnina MV, Beringer M, Wintermeyer W (2007) How ribosomes make peptide bonds. Trends Biochem Sci 32(1):20–26

    Article  CAS  PubMed  Google Scholar 

  11. Nurenberg E, Tampe R (2013) Tying up loose ends: ribosome recycling in eukaryotes and archaea. Trends Biochem Sci 38(2):64–74

    Article  PubMed  CAS  Google Scholar 

  12. Young DJ et al (2015) Rli1/ABCE1 recycles terminating ribosomes and controls translation reinitiation in 3′UTRs in vivo. Cell 162(4):872–884

    Article  CAS  PubMed  Google Scholar 

  13. Passmore LA et al (2007) The eukaryotic translation initiation factors eIF1 and eIF1A induce an open conformation of the 40S ribosome. Mol Cell 26(1):41–50

    Article  CAS  PubMed  Google Scholar 

  14. Kunze G et al (2004) The N terminus of bacterial elongation factor Tu elicits innate immunity in Arabidopsis plants. Plant Cell. 16(12):3496–3507

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Dzialo MC et al (2014) Translational roles of elongation factor 2 protein lysine methylation. J Biol Chem 289(44):30511–30524

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Odintsova TI et al (2003) Characterization and analysis of posttranslational modifications of the human large cytoplasmic ribosomal subunit proteins by mass spectrometry and Edman sequencing. J Protein Chem 22(3):249–258

    Article  CAS  PubMed  Google Scholar 

  17. Arnold RJ, Reilly JP (2002) Analysis of methylation and acetylation in E. coli ribosomal proteins. Methods Mol Biol 194:205–210

    CAS  PubMed  Google Scholar 

  18. Lee SW et al (2002) Direct mass spectrometric analysis of intact proteins of the yeast large ribosomal subunit using capillary LC/FTICR. Proc Natl Acad Sci USA 99(9):5942–5947

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Cammarano P et al (1972) Characterization of unfolded and compact ribosomal subunits from plants and their relationship to those of lower and higher animals: evidence for physicochemical heterogeneity among eucaryotic ribosomes. Biochim Biophys Acta 281(4):571–596

    Article  CAS  PubMed  Google Scholar 

  20. Louie DF et al (1996) Mass spectrometric analysis of 40S ribosomal proteins from Rat-1 fibroblasts. J Biol Chem 271(45):28189–28198

    Article  CAS  PubMed  Google Scholar 

  21. Vladimirov SN et al (1996) Characterization of the human small-ribosomal-subunit proteins by N-terminal and internal sequencing, and mass spectrometry. Eur J Biochem 239(1):144–149

    Article  CAS  PubMed  Google Scholar 

  22. Ban N et al (2014) A new system for naming ribosomal proteins. Curr Opin Struct Biol 24:165–169

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ruvinsky I et al (2005) Ribosomal protein S6 phosphorylation is a determinant of cell size and glucose homeostasis. Genes Dev 19(18):2199–2211

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Fu Y et al (2010) The AlkB domain of mammalian ABH8 catalyzes hydroxylation of 5-methoxycarbonylmethyluridine at the wobble position of tRNA. Angew Chem Int Ed Engl 49(47):8885–8888

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Kato M et al (2011) Crystal structure of a novel JmjC-domain-containing protein, TYW5, involved in tRNA modification. Nucleic Acids Res 39(4):1576–1585

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Scotti JS et al (2014) Human oxygen sensing may have origins in prokaryotic elongation factor Tu prolyl-hydroxylation. Proc Natl Acad Sci USA 111(37):13331–13336

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Feng T et al (2014) Optimal translational termination requires C4 lysyl hydroxylation of eRF1. Mol Cell 53(4):645–654

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ge W et al (2012) Oxygenase-catalyzed ribosome hydroxylation occurs in prokaryotes and humans. Nat Chem Biol 8(12):960–962

    Article  CAS  PubMed  Google Scholar 

  29. van Staalduinen LM, Novakowski SK, Jia Z (2014) Structure and functional analysis of YcfD, a novel 2-oxoglutarate/Fe(2)(+)-dependent oxygenase involved in translational regulation in Escherichia coli. J Mol Biol 426(9):1898–1910

    Article  PubMed  CAS  Google Scholar 

  30. Loenarz C, Schofield CJ (2011) Physiological and biochemical aspects of hydroxylations and demethylations catalyzed by human 2-oxoglutarate oxygenases. Trends Biochem Sci 36(1):7–18

    Article  CAS  PubMed  Google Scholar 

  31. Markolovic S, Wilkins SE, Schofield CJ (2015) Protein hydroxylation catalyzed by 2-oxoglutarate-dependent oxygenases. J Biol Chem 290(34):20712–20722

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bugg TD (2001) Oxygenases: mechanisms and structural motifs for O(2) activation. Curr Opin Chem Biol 5(5):550–555

    Article  CAS  PubMed  Google Scholar 

  33. Mittal N et al (2013) Unique posttranslational modifications in eukaryotic translation factors and their roles in protozoan parasite viability and pathogenesis. Mol Biochem Parasitol 187(1):21–31

    Article  CAS  PubMed  Google Scholar 

  34. Dever TE, Green R (2012) The elongation, termination, and recycling phases of translation in eukaryotes. Cold Spring Harb Perspect Biol 4(7):a013706

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Gupta N et al (2007) Whole proteome analysis of post-translational modifications: applications of mass-spectrometry for proteogenomic annotation. Genome Res 17(9):1362–1377

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Blaha G, Stanley RE, Steitz TA (2009) Formation of the first peptide bond: the structure of EF-P bound to the 70S ribosome. Science 325(5943):966–970

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Pavlov MY et al (2009) Slow peptide bond formation by proline and other N-alkylamino acids in translation. Proc Natl Acad Sci USA 106(1):50–54

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Ude S et al (2013) Translation elongation factor EF-P alleviates ribosome stalling at polyproline stretches. Science 339(6115):82–85

    Article  CAS  PubMed  Google Scholar 

  39. Doerfel LK et al (2013) EF-P is essential for rapid synthesis of proteins containing consecutive proline residues. Science 339(6115):85–88

    Article  CAS  PubMed  Google Scholar 

  40. Roy H et al (2011) The tRNA synthetase paralog PoxA modifies elongation factor-P with (R)-beta-lysine. Nat Chem Biol 7(10):667–669

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Peil L et al (2012) Lys34 of translation elongation factor EF-P is hydroxylated by YfcM. Nat Chem Biol 8(8):695–697

    Article  CAS  PubMed  Google Scholar 

  42. Navarre WW et al (2010) PoxA, yjeK, and elongation factor P coordinately modulate virulence and drug resistance in Salmonella enterica. Mol Cell 39(2):209–221

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Bullwinkle TJ et al (2013) (R)-Beta-lysine-modified elongation factor P functions in translation elongation. J Biol Chem 288(6):4416–4423

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kemper WM, Berry KW, Merrick WC (1976) Purification and properties of rabbit reticulocyte protein synthesis initiation factors M2Balpha and M2Bbeta. J Biol Chem 251(18):5551–5557

    CAS  PubMed  Google Scholar 

  45. Gregio AP et al (2009) eIF5A has a function in the elongation step of translation in yeast. Biochem Biophys Res Commun 380(4):785–790

    Article  CAS  PubMed  Google Scholar 

  46. Kang HA, Hershey JW (1994) Effect of initiation factor eIF-5A depletion on protein synthesis and proliferation of Saccharomyces cerevisiae. J Biol Chem 269(6):3934–3940

    CAS  PubMed  Google Scholar 

  47. Saini P et al (2009) Hypusine-containing protein eIF5A promotes translation elongation. Nature 459(7243):118–121

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Dever TE, Gutierrez E, Shin BS (2014) The hypusine-containing translation factor eIF5A. Crit Rev Biochem Mol Biol 49(5):413–425

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Gutierrez E et al (2013) eIF5A promotes translation of polyproline motifs. Mol Cell 51(1):35–45

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Park MH, Joe YA, Kang KR (1998) Deoxyhypusine synthase activity is essential for cell viability in the yeast Saccharomyces cerevisiae. J Biol Chem 273(3):1677–1683

    Article  CAS  PubMed  Google Scholar 

  51. Schrader R et al (2006) Temperature-sensitive eIF5A mutant accumulates transcripts targeted to the nonsense-mediated decay pathway. J Biol Chem 281(46):35336–35346

    Article  CAS  PubMed  Google Scholar 

  52. Park JH et al (2006) Molecular cloning, expression, and structural prediction of deoxyhypusine hydroxylase: a HEAT-repeat-containing metalloenzyme. Proc Natl Acad Sci USA 103(1):51–56

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Patel PH et al (2009) The Drosophila deoxyhypusine hydroxylase homologue nero and its target eIF5A are required for cell growth and the regulation of autophagy. J Cell Biol 185(7):1181–1194

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Sievert H et al (2014) A novel mouse model for inhibition of DOHH-mediated hypusine modification reveals a crucial function in embryonic development, proliferation and oncogenic transformation. Dis Model Mech 7(8):963–976

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Taylor D et al (2012) Cryo-EM structure of the mammalian eukaryotic release factor eRF1-eRF3-associated termination complex. Proc Natl Acad Sci USA 109(45):18413–18418

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Carlson BA et al (1999) Transfer RNA modification status influences retroviral ribosomal frameshifting. Virology 255(1):2–8

    Article  CAS  PubMed  Google Scholar 

  57. Varani G, McClain WH (2000) The G x U wobble base pair. A fundamental building block of RNA structure crucial to RNA function in diverse biological systems. EMBO Rep 1(1):18–23

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Gorgoni B et al (2014) Controlling translation elongation efficiency: tRNA regulation of ribosome flux on the mRNA. Biochem Soc Trans 42(1):160–165

    Article  CAS  PubMed  Google Scholar 

  59. Ogle JM et al (2001) Recognition of cognate transfer RNA by the 30S ribosomal subunit. Science 292(5518):897–902

    Article  CAS  PubMed  Google Scholar 

  60. Sharma D et al (2007) Mutational analysis of S12 protein and implications for the accuracy of decoding by the ribosome. J Mol Biol 374(4):1065–1076

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  62. Kawakami M et al (1988) Chemical structure of a new modified nucleoside located in the anticodon of Bombyx mori glycine tRNA2. J Biochem 104(1):108–111

    CAS  PubMed  Google Scholar 

  63. Kawakami M et al (1980) Abnormal codon recognition of glycyl-tRNA from the posterior silk glands of Bombyx mori. J Biochem 88(4):1151–1157

    CAS  PubMed  Google Scholar 

  64. Konevega AL et al (2004) Purine bases at position 37 of tRNA stabilize codon-anticodon interaction in the ribosomal A site by stacking and Mg2+-dependent interactions. RNA 10(1):90–101

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Noma A et al (2010) Expanding role of the jumonji C domain as an RNA hydroxylase. J Biol Chem 285(45):34503–34507

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Chowdhury R et al (2014) Ribosomal oxygenases are structurally conserved from prokaryotes to humans. Nature 510(7505):422–426

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Suzuki C et al (2007) Identification of Myc-associated protein with JmjC domain as a novel therapeutic target oncogene for lung cancer. Mol Cancer Ther 6(2):542–551

    Article  CAS  PubMed  Google Scholar 

  68. Teye K et al (2004) Increased expression of a Myc target gene Mina53 in human colon cancer. Am J Pathol 164(1):205–216

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Wang C et al (2015) Structure of the JmjC domain-containing protein NO66 complexed with ribosomal protein Rpl8. Acta Crystallogr D Biol Crystallogr 71(Pt 9):1955–1964

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Yanshina DD et al (2015) Hydroxylated histidine of human ribosomal protein uL2 is involved in maintaining the local structure of 28S rRNA in the ribosomal peptidyl transferase center. FEBS J 282(8):1554–1566

    Article  CAS  PubMed  Google Scholar 

  71. Singleton RS et al (2014) OGFOD1 catalyzes prolyl hydroxylation of RPS23 and is involved in translation control and stress granule formation. Proc Natl Acad Sci USA 111(11):4031–4036

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Loenarz C et al (2014) Hydroxylation of the eukaryotic ribosomal decoding center affects translational accuracy. Proc Natl Acad Sci USA 111(11):4019–4024

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Katz MJ et al (2014) Sudestada1, a Drosophila ribosomal prolyl-hydroxylase required for mRNA translation, cell homeostasis, and organ growth. Proc Natl Acad Sci USA 111(11):4025–4030

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Keeling KM et al (2006) Tpa1p is part of an mRNP complex that influences translation termination, mRNA deadenylation, and mRNA turnover in Saccharomyces cerevisiae. Mol Cell Biol 26(14):5237–5248

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Torabi N, Kruglyak L (2012) Genetic basis of hidden phenotypic variation revealed by increased translational readthrough in yeast. PLoS Genet 8(3):e1002546

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Namy O, Duchateau-Nguyen G, Rousset JP (2002) Translational readthrough of the PDE2 stop codon modulates cAMP levels in Saccharomyces cerevisiae. Mol Microbiol 43(3):641–652

    Article  CAS  PubMed  Google Scholar 

  77. Freitag J, Ast J, Bolker M (2012) Cryptic peroxisomal targeting via alternative splicing and stop codon read-through in fungi. Nature 485(7399):522–525

    Article  CAS  PubMed  Google Scholar 

  78. Touriol C et al (2003) Generation of protein isoform diversity by alternative initiation of translation at non-AUG codons. Biol Cell 95(3–4):169–178

    Article  CAS  PubMed  Google Scholar 

  79. Casey JL, Gerin JL (1995) Hepatitis D virus RNA editing: specific modification of adenosine in the antigenomic RNA. J Virol 69(12):7593–7600

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Jacks T, Varmus HE (1985) Expression of the Rous sarcoma virus pol gene by ribosomal frameshifting. Science 230(4731):1237–1242

    Article  CAS  PubMed  Google Scholar 

  81. Weiner AM, Weber K (1971) Natural read-through at the UGA termination signal of Q-beta coat protein cistron. Nat New Biol 234(50):206–209

    Article  CAS  PubMed  Google Scholar 

  82. Geller AI, Rich A (1980) A UGA termination suppression tRNATrp active in rabbit reticulocytes. Nature 283(5742):41–46

    Article  CAS  PubMed  Google Scholar 

  83. Yamaguchi Y et al (2012) L-MPZ, a novel isoform of myelin P0, is produced by stop codon readthrough. J Biol Chem 287(21):17765–17776

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Beier H, Grimm M (2001) Misreading of termination codons in eukaryotes by natural nonsense suppressor tRNAs. Nucleic Acids Res 29(23):4767–4782

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Stiebler AC et al (2014) Ribosomal readthrough at a short UGA stop codon context triggers dual localization of metabolic enzymes in Fungi and animals. PLoS Genet 10(10):e1004685

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  86. Jungreis I et al (2011) Evidence of abundant stop codon readthrough in Drosophila and other metazoa. Genome Res 21(12):2096–2113

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Brown CM et al (1993) The translational termination signal database. Nucleic Acids Res 21(13):3119–3123

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Tate WP et al (1995) Translational termination efficiency in both bacteria and mammals is regulated by the base following the stop codon. Biochem Cell Biol 73(11–12):1095–1103

    Article  CAS  PubMed  Google Scholar 

  89. Namy O, Hatin I, Rousset JP (2001) Impact of the six nucleotides downstream of the stop codon on translation termination. EMBO Rep 2(9):787–793

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Berry MJ et al (1991) Recognition of UGA as a selenocysteine codon in type I deiodinase requires sequences in the 3′ untranslated region. Nature 353(6341):273–276

    Article  CAS  PubMed  Google Scholar 

  91. Kossinova O et al (2013) A novel insight into the mechanism of mammalian selenoprotein synthesis. RNA 19(8):1147–1158

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Jalajakumari MB et al (1989) Genes for biosynthesis and assembly of CS3 pili of CFA/II enterotoxigenic Escherichia coli: novel regulation of pilus production by bypassing an amber codon. Mol Microbiol 3(12):1685–1695

    Article  CAS  PubMed  Google Scholar 

  93. Schueren F et al (2014) Peroxisomal lactate dehydrogenase is generated by translational readthrough in mammals. Elife 3:e03640

    Article  PubMed  CAS  Google Scholar 

  94. Hofstetter H, Monstein HJ, Weissmann C (1974) The readthrough protein A1 is essential for the formation of viable Q beta particles. Biochim Biophys Acta 374(2):238–251

    Article  CAS  PubMed  Google Scholar 

  95. Pelham HR (1978) Leaky UAG termination codon in tobacco mosaic virus RNA. Nature 272(5652):469–471

    Article  CAS  PubMed  Google Scholar 

  96. Philipson L et al (1978) Translation of MuLV and MSV RNAs in nuclease-treated reticulocyte extracts: enhancement of the gag-pol polypeptide with yeast suppressor tRNA. Cell 13(1):189–199

    Article  CAS  PubMed  Google Scholar 

  97. Yoshinaka Y et al (1985) Murine leukemia virus protease is encoded by the gag-pol gene and is synthesized through suppression of an amber termination codon. Proc Natl Acad Sci USA 82(6):1618–1622

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Boylan M, Coleman DC, Smyth CJ (1987) Molecular cloning and characterization of the genetic determinant encoding CS3 fimbriae of enterotoxigenic Escherichia coli. Microb Pathog 2(3):195–209

    Article  CAS  PubMed  Google Scholar 

  99. Lin MF et al (2007) Revisiting the protein-coding gene catalog of Drosophila melanogaster using 12 fly genomes. Genome Res 17(12):1823–1836

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Dunn JG et al (2013) Ribosome profiling reveals pervasive and regulated stop codon readthrough in Drosophila melanogaster. Elife 2:e01179

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  101. Xue F, Cooley L (1993) kelch encodes a component of intercellular bridges in Drosophila egg chambers. Cell 72(5):681–693

    Article  CAS  PubMed  Google Scholar 

  102. Steneberg P et al (1998) Translational readthrough in the hdc mRNA generates a novel branching inhibitor in the drosophila trachea. Genes Dev 12(7):956–967

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Bergstrom DE et al (1995) Regulatory autonomy and molecular characterization of the Drosophila out at first gene. Genetics 139(3):1331–1346

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Klagges BR et al (1996) Invertebrate synapsins: a single gene codes for several isoforms in Drosophila. J Neurosci 16(10):3154–3165

    CAS  PubMed  Google Scholar 

  105. Samuels ME, Schedl P, Cline TW (1991) The complex set of late transcripts from the Drosophila sex determination gene sex-lethal encodes multiple related polypeptides. Mol Cell Biol 11(7):3584–3602

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Samson ML, Lisbin MJ, White K (1995) Two distinct temperature-sensitive alleles at the elav locus of Drosophila are suppressed nonsense mutations of the same tryptophan codon. Genetics 141(3):1101–1111

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Chao AT et al (2003) Mutations in eukaryotic release factors 1 and 3 act as general nonsense suppressors in Drosophila. Genetics 165(2):601–612

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Chittum HS et al (1998) Rabbit beta-globin is extended beyond its UGA stop codon by multiple suppressions and translational reading gaps. Biochemistry 37(31):10866–10870

    Article  CAS  PubMed  Google Scholar 

  109. Eswarappa SM et al (2014) Programmed translational readthrough generates antiangiogenic VEGF-Ax. Cell 157(7):1605–1618

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Carmeliet P et al (1996) Abnormal blood vessel development and lethality in embryos lacking a single VEGF allele. Nature 380(6573):435–439

    Article  CAS  PubMed  Google Scholar 

  111. Demirci H et al (2013) The central role of protein S12 in organizing the structure of the decoding site of the ribosome. RNA 19(12):1791–1801

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Ogle JM et al (2002) Selection of tRNA by the ribosome requires a transition from an open to a closed form. Cell 111(5):721–732

    Article  CAS  PubMed  Google Scholar 

  113. Demirci H et al (2013) A structural basis for streptomycin-induced misreading of the genetic code. Nat Commun 4:1355

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  114. O’Connor M et al (1995) Genetic probes of ribosomal RNA function. Biochem Cell Biol 73(11–12):859–868

    Article  PubMed  Google Scholar 

  115. Funatsu G, Wittmann HG (1972) Ribosomal proteins. 33. Location of amino-acid replacements in protein S12 isolated from Escherichia coli mutants resistant to streptomycin. J Mol Biol 68(3):547–550

    Article  CAS  PubMed  Google Scholar 

  116. Alksne LE et al (1993) An accuracy center in the ribosome conserved over 2 billion years. Proc Natl Acad Sci USA 90(20):9538–9541

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Gorini L, Kataja E (1964) Streptomycin-induced oversuppression in E. coli. Proc Natl Acad Sci USA 51:995–1001

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Chandramouli P et al (2008) Structure of the mammalian 80S ribosome at 8.7 A resolution. Structure 16(4):535–548

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Bonetti B et al (1995) The efficiency of translation termination is determined by a synergistic interplay between upstream and downstream sequences in Saccharomyces cerevisiae. J Mol Biol 251(3):334–345

    Article  CAS  PubMed  Google Scholar 

  120. Selmer M et al (2006) Structure of the 70S ribosome complexed with mRNA and tRNA. Science 313(5795):1935–1942

    Article  CAS  PubMed  Google Scholar 

  121. Noeske J et al (2015) High-resolution structure of the Escherichia coli ribosome. Nat Struct Mol Biol 22(4):336–341. doi:10.1038/nsmb.2994

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Nakamura Y, Ito K (2011) tRNA mimicry in translation termination and beyond. Wiley Interdiscip Rev RNA 2(5):647–668

    Article  CAS  PubMed  Google Scholar 

  123. Chavatte L et al (2002) The invariant uridine of stop codons contacts the conserved NIKSR loop of human eRF1 in the ribosome. EMBO J 21(19):5302–5311

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Bulygin KN et al (2010) Three distinct peptides from the N domain of translation termination factor eRF1 surround stop codon in the ribosome. RNA 16(10):1902–1914

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Brown A et al (2015) Structural basis for stop codon recognition in eukaryotes. Nature 524(7566):493–496

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Laplante M, Sabatini DM (2012) mTOR signaling in growth control and disease. Cell 149(2):274–293

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Vannuvel K et al (2013) Functional and morphological impact of ER stress on mitochondria. J Cell Physiol 228(9):1802–1818

    Article  CAS  PubMed  Google Scholar 

  128. Pimentel J, Boccaccio GL (2014) Translation and silencing in RNA granules: a tale of sand grains. Front Mol Neurosci 7:68

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  129. Thomas MG et al (2011) RNA granules: the good, the bad and the ugly. Cell Signal 23(2):324–334

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Bar-Peled L, Sabatini DM (2014) Regulation of mTORC1 by amino acids. Trends Cell Biol 24(7):400–406

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Meyuhas O (2008) Physiological roles of ribosomal protein S6: one of its kind. Int Rev Cell Mol Biol 268:1–37

    Article  CAS  PubMed  Google Scholar 

  132. Browne GJ, Proud CG (2004) A novel mTOR-regulated phosphorylation site in elongation factor 2 kinase modulates the activity of the kinase and its binding to calmodulin. Mol Cell Biol 24(7):2986–2997

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Keegan LP, Gallo A, O’Connell MA (2001) The many roles of an RNA editor. Nat Rev Genet 2(11):869–878

    Article  CAS  PubMed  Google Scholar 

  134. Yamasaki S, Anderson P (2008) Reprogramming mRNA translation during stress. Curr Opin Cell Biol 20(2):222–226

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Gerashchenko MV, Lobanov AV, Gladyshev VN (2012) Genome-wide ribosome profiling reveals complex translational regulation in response to oxidative stress. Proc Natl Acad Sci USA 109(43):17394–17399

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Ingolia NT, Lareau LF, Weissman JS (2011) Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 147(4):789–802

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Oren YS et al (2014) The unfolded protein response affects readthrough of premature termination codons. EMBO Mol Med 6(5):685–701

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Sideri TC et al (2010) Ribosome-associated peroxiredoxins suppress oxidative stress-induced de novo formation of the [PSI+] prion in yeast. Proc Natl Acad Sci USA 107(14):6394–6399

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Eaglestone SS, Cox BS, Tuite MF (1999) Translation termination efficiency can be regulated in Saccharomyces cerevisiae by environmental stress through a prion-mediated mechanism. EMBO J 18(7):1974–1981

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. True HL, Berlin I, Lindquist SL (2004) Epigenetic regulation of translation reveals hidden genetic variation to produce complex traits. Nature 431(7005):184–187

    Article  CAS  PubMed  Google Scholar 

  141. Ivan M et al (2001) HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 292(5516):464–468

    Article  CAS  PubMed  Google Scholar 

  142. Jaakkola P et al (2001) Targeting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292(5516):468–472

    Article  CAS  PubMed  Google Scholar 

  143. Lando D et al (2002) FIH-1 is an asparaginyl hydroxylase enzyme that regulates the transcriptional activity of hypoxia-inducible factor. Genes Dev 16(12):1466–1471

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Bruick RK, McKnight SL (2001) A conserved family of prolyl-4-hydroxylases that modify HIF. Science 294(5545):1337–1340

    Article  CAS  PubMed  Google Scholar 

  145. Ratcliffe PJ (2013) Oxygen sensing and hypoxia signalling pathways in animals: the implications of physiology for cancer. J Physiol 591(Pt 8):2027–2042

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Epstein AC et al (2001) C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 107(1):43–54

    Article  CAS  PubMed  Google Scholar 

  147. Karijolich J, Yu YT (2014) Therapeutic suppression of premature termination codons: mechanisms and clinical considerations (review). Int J Mol Med 34(2):355–362

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Keeling KM, Bedwell DM (2011) Suppression of nonsense mutations as a therapeutic approach to treat genetic diseases. Wiley Interdiscip Rev RNA 2(6):837–852

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We wish to thank all members of the Wappner’s lab for discussions and to Peter Ratcliffe for his continuous support. This work was funded by ANPCyT Grants PICT 2014 No. 0649 and PICT 2012 No. 0214. PW is a career investigator of CONICET.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. Wappner.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Katz, M.J., Gándara, L., De Lella Ezcurra, A.L. et al. Hydroxylation and translational adaptation to stress: some answers lie beyond the STOP codon. Cell. Mol. Life Sci. 73, 1881–1893 (2016). https://doi.org/10.1007/s00018-016-2160-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00018-016-2160-y

Keywords

Navigation