1932

Abstract

A large body of evidence indicates that genome annotation pipelines have biased our view of coding sequences because they generally undersample small proteins and peptides. The recent development of genome-wide translation profiling reveals the prevalence of small/short open reading frames (smORFs or sORFs), which are scattered over all classes of transcripts, including both mRNAs and presumptive long noncoding RNAs. Proteomic approaches further confirm an unexpected variety of smORF-encoded peptides (SEPs), representing an overlooked reservoir of bioactive molecules. Indeed, functional studies in a broad range of species from yeast to humans demonstrate that SEPs can harbor key activities for the control of development, differentiation, and physiology. Here we summarize recent advances in the discovery and functional characterization of smORF/SEPs and discuss why these small players can no longer be ignored with regard to genome function.

Keyword(s): lncRNApeptideSEPssmORFsORFtranslationuORF
Loading

Article metrics loading...

/content/journals/10.1146/annurev-cellbio-100616-060516
2017-10-06
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/cellbio/33/1/annurev-cellbio-100616-060516.html?itemId=/content/journals/10.1146/annurev-cellbio-100616-060516&mimeType=html&fmt=ahah

Literature Cited

  1. Albrecht G, Mosch HU, Hoffmann B, Reusser U, Braus GH. 1998. Monitoring the Gcn4 protein-mediated response in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 273:12696–702 [Google Scholar]
  2. Alsanousi N, Sugiki T, Furuita K, So M, Lee YH. et al. 2016. Solution NMR structure and inhibitory effect against amyloid-β fibrillation of Humanin containing a d-isomerized serine residue. Biochem. Biophys. Res. Commun. 477:647–53 [Google Scholar]
  3. Anderson DM, Anderson KM, Chang CL, Makarewich CA, Nelson BR. et al. 2015. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell 160:595–606 [Google Scholar]
  4. Andreev DE, O'Connor PB, Fahey C, Kenny EM, Terenin IM. et al. 2015. Translation of 5′ leaders is pervasive in genes resistant to eIF2 repression. eLife 4:e03971 [Google Scholar]
  5. Andrews SJ, Rothnagel JA. 2014. Emerging evidence for functional peptides encoded by short open reading frames. Nat. Rev. Genet. 15:193–204 [Google Scholar]
  6. Apcher S, Millot G, Daskalogianni C, Scherl A, Manoury B, Fahraeus R. 2013. Translation of pre-spliced RNAs in the nuclear compartment generates peptides for the MHC class I pathway. PNAS 110:17951–56 [Google Scholar]
  7. Aspden JL, Eyre-Walker YC, Phillips RJ, Amin U, Mumtaz MA. et al. 2014. Extensive translation of small Open Reading Frames revealed by Poly-Ribo-Seq. eLife 3:e03528 [Google Scholar]
  8. Barbosa C, Peixeiro I, Romao L. 2013. Gene expression regulation by upstream open reading frames and human disease. PLOS Genet 9:e1003529 [Google Scholar]
  9. Bazzini AA, Johnstone TG, Christiano R, Mackowiak SD, Obermayer B. et al. 2014. Identification of small ORFs in vertebrates using ribosome footprinting and evolutionary conservation. EMBO J 33:981–93 [Google Scholar]
  10. Bi P, Ramirez-Martinez A, Li H, Cannavino J, McAnally JR. et al. 2017. Control of muscle formation by the fusogenic micropeptide myomixer. Science 356:323–27 [Google Scholar]
  11. Billingsley ML, Yun J, Reese BE, Davidson CE, Buck-Koehntop BA, Veglia G. 2006. Functional and structural properties of stannin: roles in cellular growth, selective toxicity, and mitochondrial responses to injury. J. Cell. Biochem. 98:243–50 [Google Scholar]
  12. Cabrera-Quio LE, Herberg S, Pauli A. 2016. Decoding sORF translation—from small proteins to gene regulation. RNA Biol 13:1051–59 [Google Scholar]
  13. Calkhoven CF, Bouwman PR, Snippe L, Ab G. 1994. Translation start site multiplicity of the CCAAT/enhancer binding protein α mRNA is dictated by a small 5′ open reading frame. Nucleic Acids Res 22:5540–47 [Google Scholar]
  14. Calkhoven CF, Muller C, Leutz A. 2000. Translational control of C/EBPα and C/EBPβ isoform expression. Genes Dev 14:1920–32 [Google Scholar]
  15. Calviello L, Mukherjee N, Wyler E, Zauber H, Hirsekorn A. et al. 2016. Detecting actively translated open reading frames in ribosome profiling data. Nat. Methods 13:165–70 [Google Scholar]
  16. Calvo SE, Pagliarini DJ, Mootha VK. 2009. Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. PNAS 106:7507–12 [Google Scholar]
  17. Carlevaro-Fita J, Rahim A, Guigo R, Vardy LA, Johnson R. 2016. Cytoplasmic long noncoding RNAs are frequently bound to and degraded at ribosomes in human cells. RNA 22:867–82 [Google Scholar]
  18. Carninci P, Kasukawa T, Katayama S, Gough J, Frith MC. et al. 2005. The transcriptional landscape of the mammalian genome. Science 309:1559–63 [Google Scholar]
  19. Cassidy L, Prasse D, Linke D, Schmitz RA, Tholey A. 2016. Combination of bottom-up 2D-LC-MS and semi-top-down GelFree-LC-MS enhances coverage of proteome and low molecular weight short open reading frame encoded peptides of the archaeon Methanosarcina mazei. J. Proteome Res. 15:3773–83 [Google Scholar]
  20. Casson SA, Chilley PM, Topping JF, Evans IM, Souter MA, Lindsey K. 2002. The POLARIS gene of Arabidopsis encodes a predicted peptide required for correct root growth and leaf vascular patterning. Plant Cell 14:1705–21 [Google Scholar]
  21. Cech TR, Steitz JA. 2014. The noncoding RNA revolution—trashing old rules to forge new ones. Cell 157:77–94 [Google Scholar]
  22. Chanut-Delalande H, Fernandes I, Roch F, Payre F, Plaza S. 2006. Shavenbaby couples patterning to epidermal cell shape control. PLOS Biol 4:e290 [Google Scholar]
  23. Chanut-Delalande H, Ferrer P, Payre F, Plaza S. 2012. Effectors of tridimensional cell morphogenesis and their evolution. Semin. Cell Dev. Biol. 23:341–49 [Google Scholar]
  24. Chanut-Delalande H, Hashimoto Y, Pelissier-Monier A, Spokony R, Dib A. et al. 2014. Pri peptides are mediators of ecdysone for the temporal control of development. Nat. Cell Biol. 16:1035–44 [Google Scholar]
  25. Chapman NA, Dupre DJ, Rainey JK. 2014. The apelin receptor: physiology, pathology, cell signalling, and ligand modulation of a peptide-activated class A GPCR. Biochem. Cell. Biol. 92:431–40 [Google Scholar]
  26. Chatterjee S, Pal JK. 2009. Role of 5′- and 3′-untranslated regions of mRNAs in human diseases. Biol. Cell 101:251–62 [Google Scholar]
  27. Chew GL, Pauli A, Rinn JL, Regev A, Schier AF, Valen E. 2013. Ribosome profiling reveals resemblance between long non-coding RNAs and 5′ leaders of coding RNAs. Development 140:2828–34 [Google Scholar]
  28. Chew GL, Pauli A, Schier AF. 2016. Conservation of uORF repressiveness and sequence features in mouse, human and zebrafish. Nat. Commun. 7:11663 [Google Scholar]
  29. Chilley PM, Casson SA, Tarkowski P, Hawkins N, Wang KL. et al. 2006. The POLARIS peptide of Arabidopsis regulates auxin transport and root growth via effects on ethylene signaling. Plant Cell 18:3058–72 [Google Scholar]
  30. Chng SC, Ho L, Tian J, Reversade B. 2013. ELABELA: a hormone essential for heart development signals via the apelin receptor. Dev. Cell 27:672–80 [Google Scholar]
  31. Chun SY, Rodriguez CM, Todd PK, Mills RE. 2016. SPECtre: a spectral coherence–based classifier of actively translated transcripts from ribosome profiling sequence data. BMC Bioinform 17:482 [Google Scholar]
  32. Cobb LJ, Lee C, Xiao J, Yen K, Wong RG. et al. 2016. Naturally occurring mitochondrial-derived peptides are age-dependent regulators of apoptosis, insulin sensitivity, and inflammatory markers. Aging 8:796–809 [Google Scholar]
  33. Crappe J, Ndah E, Koch A, Steyaert S, Gawron D. et al. 2015. PROTEOFORMER: deep proteome coverage through ribosome profiling and MS integration. Nucleic Acids Res 43:e29 [Google Scholar]
  34. Crappe J, Van Criekinge W, Trooskens G, Hayakawa E, Luyten W. et al. 2013. Combining in silico prediction and ribosome profiling in a genome-wide search for novel putatively coding sORFs. BMC Genom 14:648 [Google Scholar]
  35. D'Lima NG, Ma J, Winkler L, Chu Q, Loh KH. et al. 2017. A human microprotein that interacts with the mRNA decapping complex. Nat. Chem. Biol. 2:174–80 [Google Scholar]
  36. Davey NE, Seo MH, Yadav VK, Jeon J, Nim S. et al. 2017. Discovery of short linear motif–mediated interactions through phage display of intrinsically disordered regions of the human proteome. FEBS J 284:485–98 [Google Scholar]
  37. Delon I, Chanut-Delalande H, Payre F. 2003. The Ovo/Shavenbaby transcription factor specifies actin remodelling during epidermal differentiation in Drosophila. Mech. Dev. 120:747–58 [Google Scholar]
  38. Derrien T, Johnson R, Bussotti G, Tanzer A, Djebali S. et al. 2012. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res 22:1775–89 [Google Scholar]
  39. Dinger ME, Pang KC, Mercer TR, Mattick JS. 2008. Differentiating protein-coding and noncoding RNA: challenges and ambiguities. PLOS Comput. Biol. 4:e1000176 [Google Scholar]
  40. Dinkel H, Van Roey K, Michael S, Kumar M, Uyar B. et al. 2016. ELM 2016—data update and new functionality of the eukaryotic linear motif resource. Nucleic Acids Res 44:D294–300 [Google Scholar]
  41. Djakovic S, Dyachok J, Burke M, Frank MJ, Smith LG. 2006. BRICK1/HSPC300 functions with SCAR and the ARP2/3 complex to regulate epidermal cell shape in Arabidopsis. Development 133:1091–100 [Google Scholar]
  42. Dong X, Wang D, Liu P, Li C, Zhao Q. et al. 2013. Zm908p11, encoded by a short open reading frame (sORF) gene, functions in pollen tube growth as a profilin ligand in maize. J. Exp. Bot. 64:2359–72 [Google Scholar]
  43. Duncan CD, Mata J. 2014. The translational landscape of fission-yeast meiosis and sporulation. Nat. Struct. Mol. Biol. 21:641–47 [Google Scholar]
  44. Duvallet E, Boulpicante M, Yamazaki T, Daskalogianni C, Prado Martins R. et al. 2016. Exosome-driven transfer of tumor-associated Pioneer Translation Products (TA-PTPs) for the MHC class I cross-presentation pathway. Oncoimmunology 5:e1198865 [Google Scholar]
  45. Ebina I, Takemoto-Tsutsumi M, Watanabe S, Koyama H, Endo Y. et al. 2015. Identification of novel Arabidopsis thaliana upstream open reading frames that control expression of the main coding sequences in a peptide sequence–dependent manner. Nucleic Acids Res 43:1562–76 [Google Scholar]
  46. Fell VL, Schild-Poulter C. 2015. The Ku heterodimer: function in DNA repair and beyond. Mutat. Res. Rev. Mutat. Res. 763:15–29 [Google Scholar]
  47. Fernandes I, Chanut-Delalande H, Ferrer P, Latapie Y, Waltzer L. et al. 2010. Zona pellucida domain proteins remodel the apical compartment for localized cell shape changes. Dev. Cell 18:64–76 [Google Scholar]
  48. Ferreira JP, Overton KW, Wang CL. 2013. Tuning gene expression with synthetic upstream open reading frames. PNAS 110:11284–89 [Google Scholar]
  49. Fields AP, Rodriguez EH, Jovanovic M, Stern-Ginossar N, Haas BJ. et al. 2015. A regression-based analysis of ribosome-profiling data reveals a conserved complexity to mammalian translation. Mol. Cell 60:816–27 [Google Scholar]
  50. Firth AE, Brown CM. 2006. Detecting overlapping coding sequences in virus genomes. BMC Bioinform 7:75 [Google Scholar]
  51. Frank MJ, Cartwright HN, Smith LG. 2003. Three Brick genes have distinct functions in a common pathway promoting polarized cell division and cell morphogenesis in the maize leaf epidermis. Development 130:753–62 [Google Scholar]
  52. Frankel N, Erezyilmaz DF, McGregor AP, Wang S, Payre F, Stern DL. 2011. Morphological evolution caused by many subtle-effect substitutions in regulatory DNA. Nature 474:598–603 [Google Scholar]
  53. Galindo MI, Pueyo JI, Fouix S, Bishop SA, Couso JP. 2007. Peptides encoded by short ORFs control development and define a new eukaryotic gene family. PLOS Biol 5:e106 [Google Scholar]
  54. Gunisova S, Beznoskova P, Mohammad MP, Vlckova V, Valasek LS. 2016. In-depth analysis of cis-determinants that either promote or inhibit reinitiation on GCN4 mRNA after translation of its four short uORFs. RNA 22:542–58 [Google Scholar]
  55. Gunisova S, Valasek LS. 2014. Fail-safe mechanism of GCN4 translational control—uORF2 promotes reinitiation by analogous mechanism to uORF1 and thus secures its key role in GCN4 expression. Nucleic Acids Res 42:5880–93 [Google Scholar]
  56. Guo B, Zhai D, Cabezas E, Welsh K, Nouraini S. et al. 2003. Humanin peptide suppresses apoptosis by interfering with Bax activation. Nature 423:456–61 [Google Scholar]
  57. Guttman M, Russell P, Ingolia NT, Weissman JS, Lander ES. 2013. Ribosome profiling provides evidence that large noncoding RNAs do not encode proteins. Cell 154:240–51 [Google Scholar]
  58. Hanada K, Akiyama K, Sakurai T, Toyoda T, Shinozaki K, Shiu SH. 2010. sORF finder: a program package to identify small open reading frames with high coding potential. Bioinformatics 26:399–400 [Google Scholar]
  59. Hanada K, Zhang X, Borevitz JO, Li WH, Shiu SH. 2007. A large number of novel coding small open reading frames in the intergenic regions of the Arabidopsis thaliana genome are transcribed and/or under purifying selection. Genome Res 17:632–40 [Google Scholar]
  60. Hanyu-Nakamura K, Sonobe-Nojima H, Tanigawa A, Lasko P, Nakamura A. 2008. Drosophila Pgc protein inhibits P-TEFb recruitment to chromatin in primordial germ cells. Nature 451:730–33 [Google Scholar]
  61. Hao Y, Zhang L, Niu Y, Cai T, Luo J. et al. 2017. SmProt: a database of small proteins encoded by annotated coding and non-coding RNA loci. Brief. Bioinform. In press. https://doi.org/10.1093/bib/bbx005
  62. Hashimoto Y, Kurita M, Aiso S, Nishimoto I, Matsuoka M. 2009. Humanin inhibits neuronal cell death by interacting with a cytokine receptor complex or complexes involving CNTF receptor α/WSX-1/gp130. Mol. Biol. Cell 20:2864–73 [Google Scholar]
  63. Hazarika RR, De Coninck B, Yamamoto LR, Martin LR, Cammue BP, van Noort V. 2017. ARA-PEPs: a repository of putative sORF-encoded peptides in Arabidopsis thaliana. BMC Bioinform 18:37 [Google Scholar]
  64. Helker CS, Schuermann A, Pollmann C, Chng SC, Kiefer F. et al. 2015. The hormonal peptide Elabela guides angioblasts to the midline during vasculogenesis. eLife 4:e06726 [Google Scholar]
  65. Hinnebusch AG. 1984. Evidence for translational regulation of the activator of general amino acid control in yeast. PNAS 81:6442–46 [Google Scholar]
  66. Hinnebusch AG. 2005. Translational regulation of GCN4 and the general amino acid control of yeast. Annu. Rev. Microbiol. 59:407–50 [Google Scholar]
  67. Hinnebusch AG, Ivanov IP, Sonenberg N. 2016. Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science 352:1413–16 [Google Scholar]
  68. Ho L, Tan SY, Wee S, Wu Y, Tan SJ. et al. 2015. ELABELA is an endogenous growth factor that sustains hESC self-renewal via the PI3K/AKT pathway. Cell Stem Cell 17:435–47 [Google Scholar]
  69. Hsieh AC, Liu Y, Edlind MP, Ingolia NT, Janes MR. et al. 2012. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature 485:55–61 [Google Scholar]
  70. Hsu PY, Calviello L, Wu HL, Li FW, Rothfels CJ. et al. 2016. Super-resolution ribosome profiling reveals unannotated translation events in Arabidopsis. PNAS 113:E7126–35 [Google Scholar]
  71. Huang SK, Shin K, Sarker M, Rainey JK. 2017. Apela exhibits isoform- and headgroup-dependent modulation of micelle binding, peptide conformation and dynamics. Biochim. Biophys. Acta 1859:767–78 [Google Scholar]
  72. Hug N, Longman D, Caceres JF. 2016. Mechanism and regulation of the nonsense-mediated decay pathway. Nucleic Acids Res 44:1483–95 [Google Scholar]
  73. Ikeuchi M, Yamaguchi T, Kazama T, Ito T, Horiguchi G, Tsukaya H. 2011. ROTUNDIFOLIA4 regulates cell proliferation along the body axis in Arabidopsis shoot. Plant Cell Physiol 52:59–69 [Google Scholar]
  74. Ikonen M, Liu B, Hashimoto Y, Ma L, Lee KW. et al. 2003. Interaction between the Alzheimer's survival peptide humanin and insulin-like growth factor–binding protein 3 regulates cell survival and apoptosis. PNAS 100:13042–47 [Google Scholar]
  75. Ingolia NT. 2016. Ribosome footprint profiling of translation throughout the genome. Cell 165:22–33 [Google Scholar]
  76. Ingolia NT, Brar GA, Stern-Ginossar N, Harris MS, Talhouarne GJ. et al. 2014. Ribosome profiling reveals pervasive translation outside of annotated protein-coding genes. Cell Rep 8:1365–79 [Google Scholar]
  77. Ingolia NT, Ghaemmaghami S, Newman JR, Weissman JS. 2009. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324:218–23 [Google Scholar]
  78. Ingolia NT, Lareau LF, Weissman JS. 2011. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 147:789–802 [Google Scholar]
  79. Ito K, Chiba S. 2013. Arrest peptides: cis-acting modulators of translation. Annu. Rev. Biochem. 82:171–202 [Google Scholar]
  80. Jackson RJ. 2005. Alternative mechanisms of initiating translation of mammalian mRNAs. Biochem. Soc. Trans. 33:1231–41 [Google Scholar]
  81. Ji Z, Song R, Regev A, Struhl K. 2015. Many lncRNAs, 5′UTRs, and pseudogenes are translated and some are likely to express functional proteins. eLife 4:e08890 [Google Scholar]
  82. Johnstone TG, Bazzini AA, Giraldez AJ. 2016. Upstream ORFs are prevalent translational repressors in vertebrates. EMBO J 35:706–23 [Google Scholar]
  83. Kastenmayer JP, Ni L, Chu A, Kitchen LE, Au WC. et al. 2006. Functional genomics of genes with small open reading frames (sORFs) in S. cerevisiae. Genome Res. 16:365–73 [Google Scholar]
  84. Kessler MM, Zeng Q, Hogan S, Cook R, Morales AJ, Cottarel G. 2003. Systematic discovery of new genes in the Saccharomyces cerevisiae genome. Genome Res 13:264–71 [Google Scholar]
  85. Khila A, El Haidani A, Vincent A, Payre F, Souda SI. 2003. The dual function of ovo/shavenbaby in germline and epidermis differentiation is conserved between Drosophila melanogaster and the olive fruit fly Bactrocera oleae. Insect Biochem. Mol. Biol 33:691–99 [Google Scholar]
  86. Kondo T, Hashimoto Y, Kato K, Inagaki S, Hayashi S, Kageyama Y. 2007. Small peptide regulators of actin-based cell morphogenesis encoded by a polycistronic mRNA. Nat. Cell Biol. 9:660–65 [Google Scholar]
  87. Kondo T, Plaza S, Zanet J, Benrabah E, Valenti P. et al. 2010. Small peptides switch the transcriptional activity of Shavenbaby during Drosophila embryogenesis. Science 329:336–39 [Google Scholar]
  88. Kong L, Zhang Y, Ye ZQ, Liu XQ, Zhao SQ. et al. 2007. CPC: Assess the protein-coding potential of transcripts using sequence features and support vector machine. Nucleic Acids Res 35:W345–49 [Google Scholar]
  89. Kozak M. 1986. Point mutations define a sequence flanking the AUG initiator codon that modulates translation by eukaryotic ribosomes. Cell 44:283–92 [Google Scholar]
  90. Kumar M, Srinivas V, Patankar S. 2015. Upstream AUGs and upstream ORFs can regulate the downstream ORF in Plasmodium falciparum. Malar. J. 14:512 [Google Scholar]
  91. Ladoukakis E, Pereira V, Magny EG, Eyre-Walker A, Couso JP. 2011. Hundreds of putatively functional small open reading frames in Drosophila. Genome Biol. 12:R118 [Google Scholar]
  92. Laing WA, Martinez-Sanchez M, Wright MA, Bulley SM, Brewster D. et al. 2015. An upstream open reading frame is essential for feedback regulation of ascorbate biosynthesis in Arabidopsis. Plant Cell 27:772–86 [Google Scholar]
  93. Lauressergues D, Couzigou JM, Clemente HS, Martinez Y, Dunand C. et al. 2015. Primary transcripts of microRNAs encode regulatory peptides. Nature 520:90–93 [Google Scholar]
  94. Law GL, Raney A, Heusner C, Morris DR. 2001. Polyamine regulation of ribosome pausing at the upstream open reading frame of S-adenosylmethionine decarboxylase. J. Biol. Chem. 276:38036–43 [Google Scholar]
  95. Lee C, Kim KH, Cohen P. 2016. MOTS-c: a novel mitochondrial-derived peptide regulating muscle and fat metabolism. Free Radic. Biol. Med. 100:182–87 [Google Scholar]
  96. Lee C, Zeng J, Drew BG, Sallam T, Martin-Montalvo A. et al. 2015. The mitochondrial-derived peptide MOTS-c promotes metabolic homeostasis and reduces obesity and insulin resistance. Cell Metab 21:443–54 [Google Scholar]
  97. Li H, Hu C, Bai L, Li H, Li M. et al. 2016. Ultra-deep sequencing of ribosome-associated poly-adenylated RNA in early Drosophila embryos reveals hundreds of conserved translated sORFs. DNA Res 23:571–80 [Google Scholar]
  98. Lin MF, Jungreis I, Kellis M. 2011. PhyloCSF: a comparative genomics method to distinguish protein coding and non-coding regions. Bioinformatics 27:i275–82 [Google Scholar]
  99. Liu J, Mehdi S, Topping J, Friml J, Lindsey K. 2013. Interaction of PLS and PIN and hormonal crosstalk in Arabidopsis root development. Front. Plant Sci. 4:75 [Google Scholar]
  100. Liu L, Dilworth D, Gao L, Monzon J, Summers A. et al. 1999. Mutation of the CDKN2A 5′ UTR creates an aberrant initiation codon and predisposes to melanoma. Nat. Genet. 21:128–32 [Google Scholar]
  101. Ma J, Diedrich JK, Jungreis I, Donaldson C, Vaughan J. et al. 2016. Improved identification and analysis of small open reading frame encoded polypeptides. Anal. Chem. 88:3967–75 [Google Scholar]
  102. Ma J, Ward CC, Jungreis I, Slavoff SA, Schwaid AG. et al. 2014. Discovery of human sORF-encoded polypeptides (SEPs) in cell lines and tissue. J. Proteome Res. 13:1757–65 [Google Scholar]
  103. Ma J, Yan B, Qu Y, Qin F, Yang Y. et al. 2008. Zm401, a short-open reading-frame mRNA or noncoding RNA, is essential for tapetum and microspore development and can regulate the floret formation in maize. J. Cell. Biochem. 105:136–46 [Google Scholar]
  104. Mackowiak SD, Zauber H, Bielow C, Thiel D, Kutz K. et al. 2015. Extensive identification and analysis of conserved small ORFs in animals. Genome Biol 16:179 [Google Scholar]
  105. Magny EG, Pueyo JI, Pearl FM, Cespedes MA, Niven JE. et al. 2013. Conserved regulation of cardiac calcium uptake by peptides encoded in small open reading frames. Science 341:1116–20 [Google Scholar]
  106. Martinho RG, Kunwar PS, Casanova J, Lehmann R. 2004. A noncoding RNA is required for the repression of RNApolII-dependent transcription in primordial germ cells. Curr. Biol. 14:159–65 [Google Scholar]
  107. Matsumoto A, Pasut A, Matsumoto M, Yamashita R, Fung J. et al. 2017. mTORC1 and muscle regeneration are regulated by the LINC00961-encoded SPAR polypeptide. Nature 541:228–32 [Google Scholar]
  108. Matsuoka M. 2015. Protective effects of Humanin and calmodulin-like skin protein in Alzheimer's disease and broad range of abnormalities. Mol. Neurobiol. 51:1232–39 [Google Scholar]
  109. McGregor AP, Orgogozo V, Delon I, Zanet J, Srinivasan DG. et al. 2007. Morphological evolution through multiple cis-regulatory mutations at a single gene. Nature 448:587–90 [Google Scholar]
  110. Menoret D, Santolini M, Fernandes I, Spokony R, Zanet J. et al. 2013. Genome-wide analyses of Shavenbaby target genes reveals distinct features of enhancer organization. Genome Biol 14:R86 [Google Scholar]
  111. Menschaert G, Van Criekinge W, Notelaers T, Koch A, Crappe J. et al. 2013. Deep proteome coverage based on ribosome profiling aids mass spectrometry–based protein and peptide discovery and provides evidence of alternative translation products and near-cognate translation initiation events. Mol. Cell. Proteom. 12:1780–90 [Google Scholar]
  112. Michel AM, Andreev DE, Baranov PV. 2014a. Computational approach for calculating the probability of eukaryotic translation initiation from ribo-seq data that takes into account leaky scanning. BMC Bioinform 15:380 [Google Scholar]
  113. Michel AM, Fox G, Kiran AM, De Bo C, O'Connor PB. et al. 2014b. GWIPS-viz: development of a ribo-seq genome browser. Nucleic Acids Res 42:D859–64 [Google Scholar]
  114. Modell AE, Blosser SL, Arora PS. 2016. Systematic targeting of protein-protein interactions. Trends Pharmacol. Sci. 37:702–13 [Google Scholar]
  115. Mueller PP, Hinnebusch AG. 1986. Multiple upstream AUG codons mediate translational control of GCN4. Cell 45:201–7 [Google Scholar]
  116. Munzarova V, Panek J, Gunisova S, Danyi I, Szamecz B, Valasek LS. 2011. Translation reinitiation relies on the interaction between eIF3a/TIF32 and progressively folded cis-acting mRNA elements preceding short uORFs. PLOS Genet 7:e1002137 [Google Scholar]
  117. Nakamura A, Amikura R, Mukai M, Kobayashi S, Lasko PF. 1996. Requirement for a noncoding RNA in Drosophila polar granules for germ cell establishment. Science 274:2075–79 [Google Scholar]
  118. Narita NN, Moore S, Horiguchi G, Kubo M, Demura T. et al. 2004. Overexpression of a novel small peptide ROTUNDIFOLIA4 decreases cell proliferation and alters leaf shape in Arabidopsis thaliana. Plant J. 38:699–713 [Google Scholar]
  119. Natarajan K, Meyer MR, Jackson BM, Slade D, Roberts C. et al. 2001. Transcriptional profiling shows that Gcn4p is a master regulator of gene expression during amino acid starvation in yeast. Mol. Cell. Biol. 21:4347–68 [Google Scholar]
  120. Nelson BR, Makarewich CA, Anderson DM, Winders BR, Troupes CD. et al. 2016. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science 351:271–75 [Google Scholar]
  121. Olexiouk V, Crappe J, Verbruggen S, Verhegen K, Martens L, Menschaert G. 2016. sORFs.org: a repository of small ORFs identified by ribosome profiling. Nucleic Acids Res 44:D324–29 [Google Scholar]
  122. Ota T, Suzuki Y, Nishikawa T, Otsuki T, Sugiyama T. et al. 2004. Complete sequencing and characterization of 21,243 full-length human cDNAs. Nat. Genet. 36:40–45 [Google Scholar]
  123. Oyama M, Itagaki C, Hata H, Suzuki Y, Izumi T. et al. 2004. Analysis of small human proteins reveals the translation of upstream open reading frames of mRNAs. Genome Res 14:2048–52 [Google Scholar]
  124. Pauli A, Norris ML, Valen E, Chew GL, Gagnon JA. et al. 2014. Toddler: an embryonic signal that promotes cell movement via Apelin receptors. Science 343:1248636 [Google Scholar]
  125. Pauli A, Valen E, Lin MF, Garber M, Vastenhouw NL. et al. 2012. Systematic identification of long noncoding RNAs expressed during zebrafish embryogenesis. Genome Res 22:577–91 [Google Scholar]
  126. Pauli A, Valen E, Schier AF. 2015. Identifying (non-)coding RNAs and small peptides: challenges and opportunities. BioEssays 37:103–12 [Google Scholar]
  127. Payre F. 2004. Genetic control of epidermis differentiation in Drosophila. Int. J. Dev. Biol. 48:207–15 [Google Scholar]
  128. Payre F, Desplan C. 2016. Small peptides control heart activity. Science 351:226–27 [Google Scholar]
  129. Payre F, Vincent A, Carreno S. 1999. ovo/svb integrates Wingless and DER pathways to control epidermis differentiation. Nature 400:271–75 [Google Scholar]
  130. Peri S, Pandey A. 2001. A reassessment of the translation initiation codon in vertebrates. Trends Genet 17:685–87 [Google Scholar]
  131. Perjes A, Kilpio T, Ulvila J, Magga J, Alakoski T. et al. 2016. Characterization of apela, a novel endogenous ligand of apelin receptor, in the adult heart. Basic Res. Cardiol. 111:2 [Google Scholar]
  132. Prabakaran S, Hemberg M, Chauhan R, Winter D, Tweedie-Cullen RY. et al. 2014. Quantitative profiling of peptides from RNAs classified as noncoding. Nat. Commun. 5:5429 [Google Scholar]
  133. Pueyo JI, Magny EG, Couso JP. 2016a. New peptides under the s(ORF)ace of the genome. Trends Biochem. Sci. 41:665–78 [Google Scholar]
  134. Pueyo JI, Magny EG, Sampson CJ, Amin U, Evans IR. et al. 2016b. Hemotin, a regulator of phagocytosis encoded by a small ORF and conserved across metazoans. PLOS Biol 14:e1002395 [Google Scholar]
  135. Quinn ME, Goh Q, Kurosaka M, Gamage DG, Petrany MJ. et al. 2017. Myomerger induces fusion of non-fusogenic cells and is required for skeletal muscle development. Nat. Commun. 8:15665 [Google Scholar]
  136. Raj A, Wang SH, Shim H, Harpak A, Li YI. et al. 2016. Thousands of novel translated open reading frames in humans inferred by ribosome footprint profiling. eLife 5:e13328 [Google Scholar]
  137. Rohrig H, Schmidt J, Miklashevichs E, Schell J, John M. 2002. Soybean ENOD40 encodes two peptides that bind to sucrose synthase. PNAS 99:1915–20 [Google Scholar]
  138. Ruiz-Orera J, Messeguer X, Subirana JA, Alba MM. 2014. Long non-coding RNAs as a source of new peptides. eLife 3:e03523 [Google Scholar]
  139. Saghatelian A, Couso JP. 2015. Discovery and characterization of smORF-encoded bioactive polypeptides. Nat. Chem. Biol. 11:909–16 [Google Scholar]
  140. Savard J, Marques-Souza H, Aranda M, Tautz D. 2006. A segmentation gene in tribolium produces a polycistronic mRNA that codes for multiple conserved peptides. Cell 126:559–69 [Google Scholar]
  141. Schleich S, Strassburger K, Janiesch PC, Koledachkina T, Miller KK. et al. 2014. DENR-MCT-1 promotes translation re-initiation downstream of uORFs to control tissue growth. Nature 512:208–12 [Google Scholar]
  142. Schwaid AG, Shannon DA, Ma J, Slavoff SA, Levin JZ. et al. 2013. Chemoproteomic discovery of cysteine-containing human short open reading frames. J. Am. Chem. Soc. 135:16750–53 [Google Scholar]
  143. Sendoel A, Dunn JG, Rodriguez EH, Naik S, Gomez NC. et al. 2017. Translation from unconventional 5′ start sites drives tumour initiation. Nature 541:494–99 [Google Scholar]
  144. Sheth U, Parker R. 2003. Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science 300:805–8 [Google Scholar]
  145. Slavoff SA, Heo J, Budnik BA, Hanakahi LA, Saghatelian A. 2014. A human short open reading frame (sORF)-encoded polypeptide that stimulates DNA end joining. J. Biol. Chem. 289:10950–57 [Google Scholar]
  146. Slavoff SA, Mitchell AJ, Schwaid AG, Cabili MN, Ma J. et al. 2013. Peptidomic discovery of short open reading frame–encoded peptides in human cells. Nat. Chem. Biol. 9:59–64 [Google Scholar]
  147. Smith JE, Alvarez-Dominguez JR, Kline N, Huynh NJ, Geisler S. et al. 2014. Translation of small open reading frames within unannotated RNA transcripts in Saccharomyces cerevisiae. Cell Rep. 7:1858–66 [Google Scholar]
  148. Starck SR, Tsai JC, Chen K, Shodiya M, Wang L. et al. 2016. Translation from the 5′ untranslated region shapes the integrated stress response. Science 351:aad3867 [Google Scholar]
  149. Storz G, Wolf YI, Ramamurthi KS. 2014. Small proteins can no longer be ignored. Annu. Rev. Biochem. 83:753–77 [Google Scholar]
  150. Sucena E, Delon I, Jones I, Payre F, Stern DL. 2003. Regulatory evolution of shavenbaby/ovo underlies multiple cases of morphological parallelism. Nature 424:935–38 [Google Scholar]
  151. Tada M, Kirchberger MA, Katz AM. 1975. Phosphorylation of a 22,000-dalton component of the cardiac sarcoplasmic reticulum by adenosine 3′:5′-monophosphate-dependent protein kinase. J. Biol. Chem. 250:2640–47 [Google Scholar]
  152. Timinszky G, Bortfeld M, Ladurner AG. 2008. Repression of RNA polymerase II transcription by a Drosophila oligopeptide. PLOS ONE 3:e2506 [Google Scholar]
  153. Tupy JL, Bailey AM, Dailey G, Evans-Holm M, Siebel CW. et al. 2005. Identification of putative noncoding polyadenylated transcripts in Drosophila melanogaster. PNAS 102:5495–500 [Google Scholar]
  154. Uchiyama-Kadokura N, Murakami K, Takemoto M, Koyanagi N, Murota K. et al. 2014. Polyamine-responsive ribosomal arrest at the stop codon of an upstream open reading frame of the AdoMetDC1 gene triggers nonsense-mediated mRNA decay in Arabidopsis thaliana. Plant Cell Physiol. 55:1556–67 [Google Scholar]
  155. Van Damme P, Gawron D, Van Criekinge W, Menschaert G. 2014. N-terminal proteomics and ribosome profiling provide a comprehensive view of the alternative translation initiation landscape in mice and men. Mol. Cell. Proteom. 13:1245–61 [Google Scholar]
  156. Van de Sande K, Pawlowski K, Czaja I, Wieneke U, Schell J. et al. 1996. Modification of phytohormone response by a peptide encoded by ENOD40 of legumes and a nonlegume. Science 273:370–73 [Google Scholar]
  157. Vanderperre B, Lucier JF, Roucou X. 2012. HAltORF: a database of predicted out-of-frame alternative open reading frames in human. Database 2012:bas025 [Google Scholar]
  158. Vizcaino JA, Csordas A, Del-Toro N, Dianes JA, Griss J. et al. 2016. 2016 update of the PRIDE database and its related tools. Nucleic Acids Res 44:11033 [Google Scholar]
  159. Von Arnim AG, Jia Q, Vaughn JN. 2014. Regulation of plant translation by upstream open reading frames. Plant Sci 214:1–12 [Google Scholar]
  160. Von Bohlen AE, Böhm J, Pop R, Johnson DS, Tolmie J. et al. 2017. A mutation creating an upstream initiation codon in the SOX9 5′UTR causes acampomelic campomelic dysplasia. Mol. Genet. Genom. Med. 3:261–68 [Google Scholar]
  161. Wang Z, Gaba A, Sachs MS. 1999. A highly conserved mechanism of regulated ribosome stalling mediated by fungal arginine attenuator peptides that appears independent of the charging status of arginyl-tRNAs. J. Biol. Chem. 274:37565–74 [Google Scholar]
  162. Wang Z, Sachs MS. 1997. Arginine-specific regulation mediated by the Neurospora crassa arg-2 upstream open reading frame in a homologous, cell-free in vitro translation system. J. Biol. Chem. 272:255–61 [Google Scholar]
  163. Wawrzynow A, Theibert JL, Murphy C, Jona I, Martonosi A, Collins JH. 1992. Sarcolipin, the “proteolipid” of skeletal muscle sarcoplasmic reticulum, is a unique, amphipathic, 31-residue peptide. Arch. Biochem. Biophys. 298:620–23 [Google Scholar]
  164. Wen Y, Liu Y, Xu Y, Zhao Y, Hua R. et al. 2009. Loss-of-function mutations of an inhibitory upstream ORF in the human hairless transcript cause Marie Unna hereditary hypotrichosis. Nat. Genet. 41:228–33 [Google Scholar]
  165. Wethmar K, Begay V, Smink JJ, Zaragoza K, Wiesenthal V. et al. 2010a. C/EBPβΔuORF mice—a genetic model for uORF-mediated translational control in mammals. Genes Dev 24:15–20 [Google Scholar]
  166. Wethmar K, Smink JJ, Leutz A. 2010b. Upstream open reading frames: molecular switches in (patho)physiology. BioEssays 32:885–93 [Google Scholar]
  167. Xie SQ, Nie P, Wang Y, Wang H, Li H. et al. 2016. RPFdb: a database for genome wide information of translated mRNA generated from ribosome profiling. Nucleic Acids Res 44:D254–58 [Google Scholar]
  168. Xu G, Yuan M, Ai C, Liu L, Zhuang E. et al. 2017. uORF-mediated translation allows engineered plant disease resistance without fitness costs. Nature 545:491–94 [Google Scholar]
  169. Yang P, Read C, Kuc RE, Buonincontri G, Southwood M. et al. 2017. Elabela/toddler is an endogenous agonist of the apelin APJ receptor in the adult cardiovascular system, and exogenous administration of the peptide compensates for the downregulation of its expression in pulmonary arterial hypertension. Circulation 135:1160–73 [Google Scholar]
  170. Yang X, Tschaplinski TJ, Hurst GB, Jawdy S, Abraham PE. et al. 2011. Discovery and annotation of small proteins using genomics, proteomics, and computational approaches. Genome Res 21:634–41 [Google Scholar]
  171. Ye Y, Liang Y, Yu Q, Hu L, Li H. et al. 2015. Analysis of human upstream open reading frames and impact on gene expression. Hum. Genet. 134:605–12 [Google Scholar]
  172. Yen K, Lee C, Mehta H, Cohen P. 2013. The emerging role of the mitochondrial-derived peptide humanin in stress resistance. J. Mol. Endocrinol. 50:R11–19 [Google Scholar]
  173. Young SK, Wek RC. 2016. Upstream open reading frames differentially regulate gene-specific translation in the integrated stress response. J. Biol. Chem. 291:16927–35 [Google Scholar]
  174. Zanet J, Benrabah E, Li T, Pelissier-Monier A, Chanut-Delalande H. et al. 2015. Pri sORF peptides induce selective proteasome-mediated protein processing. Science 349:1356–58 [Google Scholar]
  175. Zanet J, Chanut-Delalande H, Plaza S, Payre F. 2016. Small peptides as newcomers in the control of Drosophila development. Curr. Top. Dev. Biol. 117:199–219 [Google Scholar]
  176. Zhang F, Hinnebusch AG. 2011. An upstream ORF with non-AUG start codon is translated in vivo but dispensable for translational control of GCN4 mRNA. Nucleic Acids Res 39:3128–40 [Google Scholar]
  177. Zhang Q, Vashisht AA, O'Rourke J, Corbel SY, Moran R. et al. 2017. The microprotein Minion controls cell fusion and muscle formation. Nat. Commun. 8:15664 [Google Scholar]
/content/journals/10.1146/annurev-cellbio-100616-060516
Loading
/content/journals/10.1146/annurev-cellbio-100616-060516
Loading

Data & Media loading...

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