• Keine Ergebnisse gefunden

Fig. S1: mESCs undergo morphological changes during induced neuronal differentiation. wt mESC and monoclonal lines A6, B4 and C6 during early neuronal differentiation. Pictures taken with an optical reflected light microscope.

day 1 day 2 day 3

wtclone B4clone C6clone A6

undifferentiated

193 5.2 Supplementary table

Tab. S1: List of selected candidates for cap-independent translation initiation. Candidates have been identified by proteomic analysis in mESCs after 24 h of inhibition of cap-dependent translation by dominant negative 4E-BP1 overexpression. Fold change compared to mESCs.

A6 B4 C6 combined

24 h candidates

5'UTR length

splice sites

fold

change p-value fold

change p-value fold

change p-value fold

change p-value GO term function

Atp5e 109 2,6 0,384 2,6 0,386 2,9 0,340 2,7 0,080 ATP synthase hydrolase

Brwd3 219 2,0 0,250 2,0 0,255 1,8 0,276 1,9 0,043 chromatin modifying?

Ccdc47 213 1 3,3 0,000 3,1 0,000 1,1 0,316 2,5 0,011 osteoblast differentiation, ER overload response

Cdkn1c 208 1 2,2 0,153 1,7 0,032 3,1 0,005 2,3 0,011 cell cycle

Chek1 513 -0,1 0,856 1,2 0,351 2,5 0,024 1,2 0,223 protein phosphorylation

Dclk2 682 1,9 0,051 2,4 0,003 2,3 0,058 2,2 0,005 phosphorylation, intracellular signal transduction

Ero1lb 194 0,5 0,013 2,4 0,002 -0,2 0,631 0,9 0,250 oxidation-reduction process

Fam96b 275 1,9 0,147 1,9 0,149 2,2 0,112 2,0 0,006 chromosome segregation,

iron-sulfur cluster assembly

Gtf2a1 445 2,0 0,147 2,4 0,133 2,0 0,170 2,1 0,042 transcription

Hsbp1 38 2,6 0,001 1,1 0,220 2,4 0,005 2,0 0,007

cellular response to heat, muscle contraction, endodermal cell differentiation

Lama1 70 3,3 0,047 3,1 0,057 3,1 0,059 3,2 0,000 cell adhesion

194

Lsm6 203 1 1,5 0,385 1,7 0,343 1,7 0,343 1,6 0,072 mRNA splicing factor

Lsm7 23 2,1 0,230 3,4 0,073 1,6 0,381 2,4 0,128 mRNA splicing factor

Luzp1 397 3 2,0 0,127 1,7 0,175 2,2 0,106 2,0 0,006 ---

Palld 203 1 2,8 0,000 0,9 0,112 0,3 0,571 1,3 0,114 cell migration, actin cytoskeleton organization

Pcdh1 230 1 1,7 0,241 3,1 0,003 3,0 0,003 2,6 0,006 cell adhesion

Pex5 75 1 2,5 0,014 0,5 0,468 0,6 0,517 1,2 0,136 cell development

Pqbp1 582 1,7 0,328 1,6 0,354 1,9 0,294 1,7 0,056 transcription cofactor

Rad51 230 2,1 0,353 3,2 0,109 2,0 0,361 2,4 0,081 DNA strand-pairing hydrolase

Riok1 472 2,1 0,125 2,2 0,121 2,0 0,145 2,1 0,005

rRNA processing, phosphorylation

Sh3glb2 111 0,9 0,314 0,7 0,374 2,1 0,039 1,2 0,070 ---

Stmn2 252 2,9 0,352 2,0 0,521 3,1 0,333 2,7 0,112 tubulin binding

Tfe3 718 2 2,5 0,130 2,8 0,101 3,1 0,083 2,8 0,003

transcription, osteoclast

differentiation, humoral immune response

Zfp808 110 2,0 0,020 1,3 0,129 0,3 0,656 1,2 0,071 transcription

Znhit6 136 1,6 0,153 2,0 0,109 1,5 0,182 1,7 0,010 transcription factor binding

195 6REFERENCES

[1] N. R. Pamudurti et al., “Translation of CircRNAs.,” Mol. Cell, vol. 66, no. 1, pp. 9-21.e7, Apr. 2017.

[2] F. Poulin and N. Sonenberg, “Mechanism of Translation Initiation in Eukaryotes - Madame Curie Bios ... Mechanism of Translation Initiation in Eukaryotes Mechanism of Translation Initiation in Eukaryotes - Madame Curie Bios ...,” in Madame Curie Bioscience Database [Internet], Austin (TX): Landes Bioscience, 2013.

[3] A. G. Hinnebusch and J. R. Lorsch, “The mechanism of eukaryotic translation initiation - new insights and challenges.pdf,” Cold Spring Harb. Perspect. Biol., vol. 4, no. 10, pp. 1–25, 2012.

[4] T. von der Haar, J. D. Gross, G. Wagner, and J. E. G. McCarthy, “The mRNA cap-binding protein eIF4E in post-transcriptional gene expression.,” Nat. Struct. Mol. Biol., vol.

11, no. 6, pp. 503–511, 2004.

[5] J. D. Gross et al., “Ribosome loading onto the mRNA cap is driven by conformational coupling between eIF4G and eIF4E.,” Cell, vol. 115, no. 6, pp. 739–50, Dec. 2003.

[6] L. Volpon, M. J. Osborne, I. Topisirovic, N. Siddiqui, and K. L. B. Borden, “Cap-free structure of eIF4E suggests a basis for conformational regulation by its ligands,” EMBO J., 2006.

[7] M. Hilbert, F. Kebbel, A. Gubaev, and D. Klostermeier, “eIF4G stimulates the activity of the DEAD box protein eIF4A by a conformational guidance mechanism,” Nucleic Acids Res., 2011.

[8] K. H. Nielsen et al., “Synergistic activation of eIF4A by eIF4B and eIF4G,” Nucleic Acids Res., 2011.

[9] A. R. Özeş, K. Feoktistova, B. C. Avanzino, and C. S. Fraser, “Duplex unwinding and ATPase activities of the DEAD-box helicase eIF4A are coupled by eIF4G and eIF4B,” J.

Mol. Biol., 2011.

[10] A. Jacobson, “Poly(A) Metabolism and Translation: The Closed-loop Model*,” Transl.

Control, 1996.

[11] F. Rozen, I. Edery, K. Meerovitch, T. E. Dever, W. C. Merrick, and N. Sonenberg,

“Bidirectional RNA helicase activity of eucaryotic translation initiation factors 4A and 4F.,” Mol. Cell. Biol., vol. 10, no. 3, pp. 1134–1144, 1990.

[12] F. Gebauer and M. W. Hentze, “Molecular mechanisms of translational control,” Nat. Rev.

Mol. Cell Biol., vol. 5, no. 10, pp. 827–835, 2004.

[13] R. J. Jackson, C. U. T. T. Hellen, and T. V. Pestova, “The mechanism of eukaryotic

translation initiation and principles of its regulation,” Nat. Rev. Mol. Cell Biol., vol. 11, no. 2, pp. 113–127, Feb. 2010.

[14] D. Matsuda and T. W. Dreher, “Close spacing of AUG initiation codons confers dicistronic character on a eukaryotic mRNA.,” RNA, vol. 12, no. 7, pp. 1338–49, Jul.

2006.

[15] Y. Yu et al., “Position of eukaryotic translation initiation factor eIF1A on the 40S

ribosomal subunit mapped by directed hydroxyl radical probing,” Nucleic Acids Res., 2009.

[16] A. G. Hinnebusch, “Structural Insights into the Mechanism of Scanning and Start Codon Recognition in Eukaryotic Translation Initiation,” Trends in Biochemical Sciences. 2017.

[17] M. Kozak, “Point mutations define a sequence flanking the AUG initiator codon that modulates translation by eukaryotic ribosomes,” Cell, vol. 44, no. 2, pp. 283–292, 1986.

[18] M. Kozak, “An analysis of 5’-noncoding sequences from 699 vertebrate messenger rNAS,” Nucleic Acids Res., vol. 15, no. 20, pp. 8125–8148, 1987.

[19] M. Kozak, “At least six nucleotides preceding the AUG initiator codon enhance translation in mammalian cells,” J. Mol. Biol., 1987.

196

[20] N. Sonenberg and A. G. Hinnebusch, “Regulation of Translation Initiation in Eukaryotes:

Mechanisms and Biological Targets,” Cell, vol. 136, no. 4, pp. 731–745, 2009.

[21] M. G. Acker, B. S. Shin, J. S. Nanda, A. K. Saini, T. E. Dever, and J. R. Lorsch, “Kinetic Analysis of Late Steps of Eukaryotic Translation Initiation,” J. Mol. Biol., 2009.

[22] A. Ramanathan, G. B. Robb, and S.-H. Chan, “mRNA capping: biological functions and applications.,” Nucleic Acids Res., vol. 44, no. 16, pp. 7511–26, Sep. 2016.

[23] J. Mailliot and F. Martin, “Viral internal ribosomal entry sites: four classes for one goal,”

Wiley Interdiscip. Rev. RNA, vol. 9, no. 2, 2018.

[24] J. Pelletier and N. Sonenberg, “Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA,” Nature, vol. 334, no. 6180, pp. 320–

325, Jul. 1988.

[25] D. Trono, J. Pelletier, N. Sonenberg, and D. Baltimore, “Translation in mammalian cells of a gene linked to the poliovirus 5’ noncoding region.,” Science, vol. 241, no. 4864, pp.

445–8, Jul. 1988.

[26] S. K. Jang, H. G. Kräusslich, M. J. Nicklin, G. M. Duke, A. C. Palmenberg, and E.

Wimmer, “A segment of the 5’ nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation.,” J. Virol., vol. 62, no. 8, pp.

2636–43, Aug. 1988.

[27] R. J. Jackson, “The current status of vertebrate cellular mRNA IRESs,” Cold Spring Harb.

Perspect. Biol., vol. 5, no. 2, 2013.

[28] E. Martinez-Salas, R. Francisco-Velilla, J. Fernandez-Chamorro, and A. M. Embarek,

“Insights into structural and mechanistic features of viral IRES elements,” Front. Microbiol., vol. 8, no. JAN, pp. 1–15, 2018.

[29] G. Lozano and E. Martínez-Salas, “Structural insights into viral IRES-dependent translation mechanisms,” Curr. Opin. Virol., vol. 12, pp. 113–120, Jun. 2015.

[30] K.-M. Lee, C.-J. Chen, and S.-R. Shih, “Regulation Mechanisms of Viral IRES-Driven Translation,” Trends Microbiol., vol. 25, no. 7, pp. 546–561, Jul. 2017.

[31] Y. Hashem et al., “Hepatitis-C-virus-like internal ribosome entry sites displace eIF3 to gain access to the 40S subunit.,” Nature, vol. 503, no. 7477, pp. 539–43, Nov. 2013.

[32] K. Meerovitch, J. Pelletier, and N. Sonenberg, “A cellular protein that binds to the 5’-noncoding region of poliovirus RNA: implications for internal translation initiation.,”

Genes Dev., vol. 3, no. 7, pp. 1026–34, Jul. 1989.

[33] N. Luz and E. Beck, “Interaction of a cellular 57-kilodalton protein with the internal translation initiation site of foot-and-mouth disease virus.,” J. Virol., vol. 65, no. 12, pp.

6486–94, Dec. 1991.

[34] M. Kozak, “Alternative ways to think about mRNA sequences and proteins that appear to promote internal initiation of translation,” Gene, vol. 318, no. 1–2, pp. 1–23, 2003.

[35] R. J. J. Jackson, “Alternative mechanisms of initiating translation of mammalian mRNAs,”

Biochem. Soc. Trans., vol. 33, no. 6, p. 1231, 2005.

[36] P. Sarnow, “Translation of glucose-regulated protein 78/immunoglobulin heavy-chain binding protein mRNA is increased in poliovirus-infected cells at a time when cap-dependent translation of cellular mRNAs is inhibited.,” Proc. Natl. Acad. Sci. U. S. A., vol.

86, no. 15, pp. 5795–9, Aug. 1989.

[37] D. G. Macejak and P. Sarnow, “Internal initiation of translation mediated by the 5’ leader of a cellular mRNA.,” Nature, vol. 353, pp. 90–94, 1991.

[38] G. Johannes, M. S. Carter, M. B. Eisen, P. O. Brown, and P. Sarnow, “Identification of eukaryotic mRNAs that are translated at reduced cap binding complex eIF4F

concentrations using a cDNA microarray.,” Proc. Natl. Acad. Sci. U. S. A., vol. 96, no. 23, pp. 13118–23, 1999.

[39] K. A. Spriggs, M. Bushell, S. A. Mitchell, and A. E. Willis, “Internal ribosome entry

segment-mediated translation during apoptosis: the role of IRES-trans-acting factors,” Cell Death Differ., vol. 12, no. 6, pp. 585–591, 2005.

197

[40] K. A. Spriggs, M. Stoneley, M. Bushell, and A. E. Willis, “Re-programming of translation following cell stress allows IRES-mediated translation to predominate,” Biol. Cell, vol. 100, no. 1, pp. 27–38, 2008.

[41] M. Bushell et al., “Polypyrimidine Tract Binding Protein Regulates IRES-Mediated Gene Expression during Apoptosis,” Mol. Cell, vol. 23, no. 3, pp. 401–412, 2006.

[42] X. Qin and P. Sarnow, “Preferential Translation of Internal Ribosome Entry

Site-containing mRNAs during the Mitotic Cycle in Mammalian Cells,” J. Biol. Chem., vol. 279, no. 14, pp. 13721–13728, 2004.

[43] T. Kawai et al., “Global mRNA Stabilization Preferentially Linked to Translational Repression during the Endoplasmic Reticulum Stress Response Global mRNA

Stabilization Preferentially Linked to Translational Repression during the Endoplasmic Reticulum Stress Response,” Mol. Cell. Biol., vol. 24, no. 15, pp. 6639–6646, 2004.

[44] J. D. Thomas and G. J. Johannes, “Identification of mRNAs that continue to associate with polysomes during hypoxia,” RNA, vol. 13, no. 7, pp. 1116–1131, Jul. 2007.

[45] S. D. Baird, M. Turcotte, R. G. Korneluk, and M. Holcik, “Searching for IRES.,” RNA, vol. 12, no. 10, pp. 1755–85, Oct. 2006.

[46] M. Mokrejs et al., “IRESite: the database of experimentally verified IRES structures (www.iresite.org).,” Nucleic Acids Res., vol. 34, no. Database issue, pp. D125-30, Jan. 2006.

[47] M. Mokrejs, T. Masek, V. Vopálensky, P. Hlubucek, P. Delbos, and M. Pospísek,

“IRESite--a tool for the examination of viral and cellular internal ribosome entry sites.,”

Nucleic Acids Res., vol. 38, no. Database issue, pp. D131-6, Jan. 2010.

[48] C. U. T. Hellen and P. Sarnow, “Internal ribosome entry sites in eukaryotic mRNA molecules.,” Genes Dev., vol. 15, no. 13, pp. 1593–612, Jul. 2001.

[49] P. Kolekar, A. Pataskar, U. Kulkarni-Kale, J. Pal, and A. Kulkarni, “IRESPred: Web Server for Prediction of Cellular and Viral Internal Ribosome Entry Site (IRES),” Sci. Rep., vol. 6, no. June, p. 27436, 2016.

[50] S. Y. Le and J. V Maizel, “A common RNA structural motif involved in the internal initiation of translation of cellular mRNAs.,” Nucleic Acids Res., vol. 25, no. 2, pp. 362–69, Jan. 1997.

[51] G. Grillo et al., “UTRdb and UTRsite (RELEASE 2010): a collection of sequences and regulatory motifs of the untranslated regions of eukaryotic mRNAs.,” Nucleic Acids Res., vol. 38, no. Database issue, pp. D75-80, Jan. 2010.

[52] M. Kozak, “MINIREVIEW New Ways of Initiating Translation in Eukaryotes ?,” vol. 21, no. 6, pp. 1899–1907, 2001.

[53] B. Han and J. Zhang, “Regulation of Gene Expression by Internal Ribosome Entry Sites or Cryptic Promoters : the eIF4G Story Regulation of Gene Expression by Internal Ribosome Entry Sites or Cryptic Promoters : the eIF4G Story,” Mol. Cell. Biol., vol. 22, no.

21, pp. 7372–7384, 2002.

[54] W. Gan and R. E. Rhoads, “Internal Initiation of Translation Directed by the 5 -Untranslated Region of the mRNA for eIF4G , a Factor Involved in the from Cap-dependent to Internal Initiation *,” Biochemistry, pp. 623–626, 1996.

[55] W. Gan, M. LaCelle, and R. E. Rhoads, “Functional characterization of the internal

ribosome entry site of eIF4G mRNA.,” J. Biol. Chem., vol. 273, no. 9, pp. 5006–5012, 1998.

[56] R. Schneider, M. Kozak, C. Biology, and M. Kozak, “New ways of initiating translation in eukaryotes?,” Mol. Cell. Biol., vol. 21, no. 6, pp. 1899–907, Mar. 2001.

[57] B. T. Baranick, N. A. Lemp, J. Nagashima, K. Hiraoka, N. Kasahara, and C. R. Logg,

“Splicing mediates the activity of four putative cellular internal ribosome entry sites.,” Proc.

Natl. Acad. Sci. U. S. A., vol. 105, no. 12, pp. 4733–4738, 2008.

[58] R. M. Young, S. J. Wang, J. D. Gordan, X. Ji, S. A. Liebhaber, and M. C. Simon,

“Hypoxia-mediated selective mRNA translation by an internal ribosome entry site-independent mechanism,” J. Biol. Chem., vol. 283, no. 24, pp. 16309–16319, 2008.

[59] M. Kozak, “A second look at cellular mRNA sequences said to function as internal

198

ribosome entry sites,” Nucleic Acids Res., vol. 33, no. 20, pp. 6593–6602, 2005.

[60] M. Kozak, “Lessons (not) learned from mistakes about translation,” Gene, vol. 403, no. 1–

2, pp. 194–203, 2007.

[61] I. N. Shatsky, S. E. Dmitriev, I. M. Terenin, and D. E. Andreev, “Cap- and

IRES-independent scanning mechanism of translation initiation as an alternative to the concept of cellular IRESs,” Mol. Cells, vol. 30, no. 4, pp. 285–293, 2010.

[62] H. Allam and N. Ali, “Initiation factor eIF2-independent mode of c-Src mRNA translation occurs via an internal ribosome entry site,” J. Biol. Chem., vol. 285, no. 8, pp.

5713–5725, 2010.

[63] O. S. Kwon et al., “An mRNA-specific tRNAi carrier eIF2A plays a pivotal role in cell proliferation under stress conditions: Stress-resistant translation of c-Src mRNA is mediated by eIF2A,” Nucleic Acids Res., vol. 45, no. 1, pp. 296–310, 2017.

[64] I. M. Terenin, V. V. Smirnova, D. E. Andreev, S. E. Dmitriev, and I. N. Shatsky, “A researcher’s guide to the galaxy of IRESs,” Cell. Mol. Life Sci., vol. 74, no. 8, pp. 1431–

1455, Apr. 2017.

[65] M. E. Van Eden, M. P. Byrd, K. W. Sherrill, and R. E. Lloyd, “Demonstrating internal ribosome entry sites in eukaryotic mRNAs using stringent RNA test procedures.,” RNA, vol. 10, no. 4, pp. 720–30, Apr. 2004.

[66] S. R. Thompson, “So you want to know if your message has an IRES?,” Wiley Interdiscip.

Rev. RNA, vol. 3, no. 5, pp. 697–705, 2012.

[67] D. E. Andreev, S. E. Dmitriev, I. M. Terenin, V. S. Prassolov, W. C. Merrick, and I. N.

Shatsky, “Differential contribution of the m7G-cap to the 5′ end-dependent translation initiation of mammalian mRNAs,” Nucleic Acids Res., vol. 37, no. 18, pp. 6135–6147, 2009.

[68] A. Unbehaun, S. I. Borukhov, C. U. T. Hellen, and T. V Pestova, “Release of initiation factors from 48S complexes during ribosomal subunit joining and the link between establishment of codon-anticodon base-pairing and hydrolysis of eIF2-bound GTP.,”

Genes Dev., vol. 18, no. 24, pp. 3078–93, Dec. 2004.

[69] D. E. Andreev et al., “Translation of 5’ leaders is pervasive in genes resistant to eIF2 repression,” Elife, vol. 2015, no. 4, pp. 1–21, 2015.

[70] A. Basu et al., “Requirement of rRNA Methylation for 80S Ribosome Assembly on a Cohort of Cellular Internal Ribosome Entry Sites,” Mol. Cell. Biol., vol. 31, no. 22, pp.

4482–4499, 2011.

[71] A. A. Komar, B. Mazumder, and W. C. Merrick, “A new framework for understanding IRES-mediated translation,” Gene, vol. 502, no. 2, pp. 75–86, 2012.

[72] W. C. Merrick, “Cap-dependent and cap-independent translation in eukaryotic systems,”

Gene, vol. 332, no. 1–2, pp. 1–11, 2004.

[73] A. G. Bert, R. Grépin, M. a Vadas, and G. J. Goodall, “Assessing IRES activity in the HIF-1alpha and other cellular 5’ UTRs.,” RNA, vol. 12, no. 6, pp. 1074–83, Jun. 2006.

[74] M. E. Robertson, R. A. Seamons, and G. J. Belsham, “A selection system for functional internal ribosome entry site (IRES) elements: analysis of the requirement for a conserved GNRA tetraloop in the encephalomyocarditis virus IRES.,” RNA, vol. 5, no. 9, pp. 1167–

79, 1999.

[75] J. S. Kieft, A. Grech, P. Adams, and J. A. Doudna, “Mechanisms of internal ribosome entry in translation initiation.,” Cold Spring Harb. Symp. Quant. Biol., vol. 66, pp. 277–83, 2001.

[76] R. J. Jackson and A. Kaminski, “Internal initiation of translation in eukaryotes: the picornavirus paradigm and beyond.,” RNA, vol. 1, no. 10, pp. 985–1000, Dec. 1995.

[77] A. A. Komar and M. Hatzoglou, “Internal Ribosome Entry Sites in Cellular mRNAs:

Mystery of Their Existence,” J. Biol. Chem., vol. 280, no. 25, pp. 23425–23428, Jun. 2005.

[78] M. Kozak, “Faulty old ideas about translational regulation paved the way for current confusion about how microRNAs function,” Gene, vol. 423, no. 2, pp. 108–115, 2008.

[79] C. C. Thoreen, L. Chantranupong, H. R. Keys, T. Wang, N. S. Gray, and D. M. Sabatini,

199

“A unifying model for mTORC1-mediated regulation of mRNA translation,” Nature, vol.

485, no. 7396, pp. 109–113, May 2012.

[80] M. E. Tanenbaum, N. Stern-Ginossar, J. S. Weissman, and R. D. Vale, “Regulation of mRNA translation during mitosis,” Elife, vol. 4, no. AUGUST2015, pp. 1–19, 2015.

[81] M. Kozak, “New Ways of Initiating Translation in Eukaryotes?,” Mol. Cell. Biol., vol. 21, no. 6, pp. 1899–1907, Mar. 2001.

[82] M. Matsui, N. Yachie, Y. Okada, R. Saito, and M. Tomita, “Bioinformatic analysis of post-transcriptional regulation by uORF in human and mouse,” FEBS Lett., vol. 581, no. 22, pp. 4184–4188, 2007.

[83] S. E. Calvo, D. J. Pagliarini, and V. K. Mootha, “Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans,” Proc.

Natl. Acad. Sci., vol. 106, no. 18, pp. 7507–7512, 2009.

[84] Y. Ye et al., “Analysis of human upstream open reading frames and impact on gene expression,” Hum. Genet., vol. 134, no. 6, pp. 605–612, 2015.

[85] T. G. Johnstone, A. A. Bazzini, and A. J. Giraldez, “Upstream ORFs are prevalent translational repressors in vertebrates,” EMBO J., vol. 35, no. 7, pp. 706–723, 2016.

[86] K. Wethmar, “The regulatory potential of upstream open reading frames in eukaryotic gene expression,” Wiley Interdiscip. Rev. RNA, vol. 5, no. 6, pp. 765–778, 2014.

[87] H. Mouilleron, V. Delcourt, and X. Roucou, “Death of a dogma: Eukaryotic mRNAs can code for more than one protein,” Nucleic Acids Res., vol. 44, no. 1, pp. 14–23, 2016.

[88] A. M. McGeachy and N. T. Ingolia, “Starting too soon: upstream reading frames repress downstream translation,” EMBO J., vol. 35, no. 7, pp. 699–700, 2016.

[89] A. P. Fields et al., “A Regression-Based Analysis of Ribosome-Profiling Data Reveals a Conserved Complexity to Mammalian Translation,” Mol. Cell, vol. 60, no. 5, pp. 816–827, 2015.

[90] B. Vanderperre et al., “Direct Detection of Alternative Open Reading Frames Translation Products in Human Significantly Expands the Proteome,” PLoS One, vol. 8, no. 8, 2013.

[91] N. T. Ingolia, L. F. Lareau, and J. S. Weissman, “Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes,” Cell, vol. 147, no. 4, pp. 789–802, 2011.

[92] H. Xu et al., “Length of the ORF, position of the first AUG and the Kozak motif are important factors in potential dual-coding transcripts,” Cell Res., vol. 20, no. 4, pp. 445–

457, 2010.

[93] S. Ribrioux, A. Brüngger, B. Baumgarten, K. Seuwen, and M. R. John, “Bioinformatics prediction of overlapping frameshifted translation products in mammalian transcripts,”

BMC Genomics, vol. 9, no. 1, p. 122, 2008.

[94] W.-Y. Chung, S. Wadhawan, R. Szklarczyk, S. K. Pond, and A. Nekrutenko, “A First Look at ARFome: Dual-Coding Genes in Mammalian Genomes,” PLoS Comput. Biol., vol. 3, no.

5, p. e91, 2007.

[95] X. Gao, J. Wan, B. Liu, M. Ma, B. Shen, and S.-B. Qian, “Quantitative profiling of initiating ribosomes in vivo.,” Nat. Methods, vol. 12, no. 2, pp. 147–53, Feb. 2015.

[96] S. Lee, B. Liu, S. Lee, S.-X. Huang, B. Shen, and S.-B. Qian, “Global mapping of translation initiation sites in mammalian cells at single-nucleotide resolution,” Proc. Natl.

Acad. Sci., vol. 109, no. 37, pp. E2424–E2432, 2012.

[97] A. Koch et al., “A proteogenomics approach integrating proteomics and ribosome profiling increases the efficiency of protein identification and enables the discovery of alternative translation start sites,” Proteomics, vol. 14, no. 23–24, pp. 2688–2698, Dec. 2014.

[98] K. J. Autio et al., “An ancient genetic link between vertebrate mitochondrial fatty acid synthesis and RNA processing,” FASEB J., vol. 22, no. 2, pp. 569–578, Feb. 2008.

[99] D. E. Andreev et al., “Oxygen and glucose deprivation induces widespread alterations in mRNA translation within 20minutes,” Genome Biol., vol. 16, no. 1, pp. 1–14, 2015.

[100] M. Kozak, “Pushing the limits of the scanning mechanism for initiation of translation,”