Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Perspective
  • Published:

Selection on synonymous sites: the unwanted transcript hypothesis

Abstract

Although translational selection to favour codons that match the most abundant tRNAs is not readily observed in humans, there is nonetheless selection in humans on synonymous mutations. We hypothesize that much of this synonymous site selection can be explained in terms of protection against unwanted RNAs — spurious transcripts, mis-spliced forms or RNAs derived from transposable elements or viruses. We propose not only that selection on synonymous sites functions to reduce the rate of creation of unwanted transcripts (for example, through selection on exonic splice enhancers and cryptic splice sites) but also that high-GC content (but low-CpG content), together with intron presence and position, is both particular to functional native mRNAs and used to recognize transcripts as native. In support of this hypothesis, transcription, nuclear export, liquid phase condensation and RNA degradation have all recently been shown to promote GC-rich transcripts and suppress AU/CpG-rich ones. With such ‘traps’ being set against AU/CpG-rich transcripts, the codon usage of native genes has, in turn, evolved to avoid such suppression. That parallel filters against AU/CpG-rich transcripts also affect the endosomal import of RNAs further supports the unwanted transcript hypothesis of synonymous site selection and explains the similar design rules that have enabled the successful use of transgenes and RNA vaccines.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Processes acting on the transcription, processing and filtering of mRNA transcripts, with emphasis on roles of codon content or intronic signals.
Fig. 2: Dinucleotide frequencies of human coding sequences, single-exon coding sequences and genomic DNA, and as expected from mononucleotide frequencies.
Fig. 3: Native human exons predicted to be targeted by the HUSH complex.
Fig. 4: The mode of action of disease-associated synonymous mutations.

Similar content being viewed by others

References

  1. King, J. L. & Jukes, T. H. Non-Darwinian evolution. Science 164, 788–798 (1969).

    Article  CAS  PubMed  Google Scholar 

  2. Sharp, P. M., Averof, M., Lloyd, A. T., Matassi, G. & Peden, J. F. DNA sequence evolution: the sounds of silence. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 349, 241–247 (1995).

    Article  CAS  PubMed  Google Scholar 

  3. Ikemura, T. Correlation between the abundance of Escherichia coli transfer RNAs and the occurrence of the respective codons in its protein genes: a proposal for a synonymous codon choice that is optimal for the E. coli translational system. J. Mol. Biol. 151, 389–409 (1981).

    Article  CAS  PubMed  Google Scholar 

  4. Ikemura, T. Codon usage and tRNA content in unicellular and multicellular organisms. Mol. Biol. Evol. 2, 13–34 (1985).

    CAS  PubMed  Google Scholar 

  5. Sharp, P. M. & Li, W.-H. The codon adaptation index — a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 15, 1281–1295 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Qian, W., Yang, J. R., Pearson, N. M., Maclean, C. & Zhang, J. Balanced codon usage optimizes eukaryotic translational efficiency. PLoS Genet. 8, e1002603 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Akashi, H. Synonymous codon usage in Drosophila melanogaster: natural selection and translational accuracy. Genetics 136, 927–935 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Stoletzki, N. & Eyre-Walker, A. Synonymous codon usage in Escherichia coli: selection for translational accuracy. Mol. Biol. Evol. 24, 374–381 (2007).

    Article  CAS  PubMed  Google Scholar 

  9. Sharp, P. M., Emery, L. R. & Zeng, K. Forces that influence the evolution of codon bias. Philos. Trans. R. Soc. Lond. B Biol. Sci. 365, 1203–1212 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. dos Reis, M. & Wernisch, L. Estimating translational selection in eukaryotic genomes. Mol. Biol. Evol. 26, 451–461 (2009).

    Article  PubMed  Google Scholar 

  11. Lynch, M. & Conery, J. S. The origins of genome complexity. Science 302, 1401–1404 (2003).

    Article  CAS  PubMed  Google Scholar 

  12. Duret, L. Evolution of synonymous codon usage in metazoans. Curr. Opin. Genet. Dev. 12, 640–649 (2002).

    Article  CAS  PubMed  Google Scholar 

  13. Hunt, R. C., Simhadri, V. L., Iandoli, M., Sauna, Z. E. & Kimchi-Sarfaty, C. Exposing synonymous mutations. Trends Genet. 30, 308–321 (2014).

    Article  CAS  PubMed  Google Scholar 

  14. Bali, V. & Bebok, Z. Decoding mechanisms by which silent codon changes influence protein biogenesis and function. Int. J. Biochem. Cell Biol. 64, 58–74 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kudla, G., Lipinski, L., Caffin, F., Helwak, A. & Zylicz, M. High guanine and cytosine content increases mRNA levels in mammalian cells. PLoS Biol. 4, e180 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Mordstein, C. et al. Codon usage and splicing jointly influence mRNA localization. Cell Syst. 10, 351–362.e8 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Zuckerman, B., Ron, M., Mikl, M., Segal, E. & Ulitsky, I. Gene architecture and sequence composition underpin selective dependency of nuclear export of long RNAs on NXF1 and the TREX complex. Mol. Cell 79, 251–267.e6 (2020).

    Article  CAS  PubMed  Google Scholar 

  18. Lin, M. F. et al. Locating protein-coding sequences under selection for additional, overlapping functions in 29 mammalian genomes. Genome Res. 21, 1916–1928 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Caceres, E. F. & Hurst, L. D. The evolution, impact and properties of exonic splice enhancers. Genome Biol. 14, R143 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Savisaar, R. & Hurst, L. D. Exonic splice regulation imposes strong selection at synonymous sites. Genome Res. 28, 1442–1454 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Keightley, P. D. & Halligan, D. L. Inference of site frequency spectra from high-throughput sequence data: quantification of selection on nonsynonymous and synonymous sites in humans. Genetics 188, 931–940 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Eory, L., Halligan, D. L. & Keightley, P. D. Distributions of selectively constrained sites and deleterious mutation rates in the hominid and murid genomes. Mol. Biol. Evol. 27, 177–192 (2010).

    Article  CAS  PubMed  Google Scholar 

  23. Wen, P., Xiao, P. & Xia, J. dbDSM: a manually curated database for deleterious synonymous mutations. Bioinformatics 32, 1914–1916 (2016).

    Article  CAS  PubMed  Google Scholar 

  24. Statello, L., Guo, C.-J., Chen, L.-L. & Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 22, 96–118 (2021).

    Article  CAS  PubMed  Google Scholar 

  25. Hurst, L. D. Evolutionary genomics and the reach of selection. J. Biol. 8, 12 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Andrews, G. et al. Mammalian evolution of human cis-regulatory elements and transcription factor binding sites. Science 380, eabn7930 (2023).

    Article  CAS  PubMed  Google Scholar 

  27. Luthra, I. et al. Biochemical activity is the default DNA state in eukaryotes. Preprint at bioRxiv https://doi.org/10.1101/2022.12.16.520785 (2022).

  28. Camellato, B., Brosh, R., Maurano, M. T. & Boeke, J. D. Genomic analysis of a synthetic reversed sequence reveals default chromatin states in yeast and mammalian cells. Preprint at bioRxiv https://doi.org/10.1101/2023.06.20.545713v2 (2022).

  29. Xu, H., Li, C., Xu, C. & Zhang, J. Chance promoter activities illuminate the origins of eukaryotic intergenic transcriptions. Nat. Commun. 14, 1826 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Preker, P. et al. PROMoter uPstream transcripts share characteristics with mRNAs and are produced upstream of all three major types of mammalian promoters. Nucleic Acids Res. 39, 7179–7193 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Schuler, A., Ghanbarian, A. T. & Hurst, L. D. Purifying selection on splice-related motifs, not expression level nor RNA folding, explains nearly all constraint on human lincRNAs. Mol. Biol. Evol. 31, 3164–3183 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Managadze, D. et al. Negative correlation between expression level and evolutionary rate of long intergenic noncoding RNAs. Genome Biol. Evol. 3, 1390–1404 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Haerty, W. & Ponting, C. P. Mutations within lncRNAs are effectively selected against in fruitfly but not in human. Genome Biol. 14, R49 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Johnsson, P., Lipovich, L., Grandér, D. & Morris, K. V. Evolutionary conservation of long non-coding RNAs; sequence, structure, function. Biochim. Biophys. Acta 1840, 1063–1071 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Ponting, C. P. & Haerty, W. Genome-wide analysis of human long noncoding RNAs: a provocative review. Annu. Rev. Genomics Hum. Genet. 23, 153–172 (2022).

    Article  CAS  PubMed  Google Scholar 

  36. Wyers, F. et al. Cryptic Pol II transcripts are degraded by a nuclear quality control pathway involving a new poly(A) polymerase. Cell 121, 725–737 (2005).

    Article  CAS  PubMed  Google Scholar 

  37. Liu, S. J. et al. CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells. Science 355, aah7111 (2017).

    Article  PubMed  Google Scholar 

  38. Schlackow, M. et al. Distinctive patterns of transcription and RNA processing for human lincRNAs. Mol. Cell 65, 25–38 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Chuong, E. B., Elde, N. C. & Feschotte, C. Regulatory activities of transposable elements: from conflicts to benefits. Nat. Rev. Genet. 18, 71–86 (2017).

    Article  CAS  PubMed  Google Scholar 

  40. Wang, J. et al. Primate-specific endogenous retrovirus-driven transcription defines naive-like stem cells. Nature 516, 405–409 (2014).

    Article  CAS  PubMed  Google Scholar 

  41. Raskó, T. et al. A novel gene controls a new structure: piggybac transposable element-derived 1,unique to mammals, controls mammal-specific neuronal paraspeckles. Mol. Biol. Evol. 39, msac175 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  42. The ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

    Article  PubMed Central  Google Scholar 

  43. Carlevaro-Fita, J. et al. Ancient exapted transposable elements promote nuclear enrichment of human long noncoding RNAs. Genome Res. 29, 208–222 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Pickrell, J. K., Pai, A. A., Gilad, Y. & Pritchard, J. K. Noisy splicing drives mRNA isoform diversity in human cells. PLoS Genet. 6, e1001236 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Bénitìere, F., Necsulea, A. & Duret, L. Random genetic drift sets an upper limit on mRNA splicing accuracy in metazoans. Preprint at bioRxiv https://doi.org/10.1101/2022.12.09.519597v5 (2023).

  46. Irimia, M. et al. Complex selection on 5′ splice sites in intron-rich organisms. Genome Res. 19, 2021–2027 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Savisaar, R. & Hurst, L. D. Estimating the prevalence of functional exonic splice regulatory information. Hum. Genet. 136, 1059–1078 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Wagner, A. Energy constraints on the evolution of gene expression. Mol. Biol. Evol. 22, 1365–1374 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Kudla, G., Murray, A. W., Tollervey, D. & Plotkin, J. B. Coding-sequence determinants of gene expression in Escherichia coli. Science 324, 255–258 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Cambray, G., Guimaraes, J. C. & Arkin, A. P. Evaluation of 244,000 synthetic sequences reveals design principles to optimize translation in Escherichia coli. Nat. Biotechnol. 36, 1005–1015 (2018).

    Article  CAS  PubMed  Google Scholar 

  51. Mittal, P., Brindle, J., Stephen, J., Plotkin, J. B. & Kudla, G. Codon usage influences fitness through RNA toxicity. Proc. Natl Acad. Sci. USA 115, 8639–8644 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bourque, G. et al. Ten things you should know about transposable elements. Genome Biol. 19, 199 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Lionetti, M. et al. A compendium of DIS3 mutations and associated transcriptional signatures in plasma cell dyscrasias. Oncotarget 6, 26129–26141 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Fasken, M. B. et al. The RNA exosome and human disease. Methods Mol. Biol. 2062, 3–33 (2020).

    Article  CAS  PubMed  Google Scholar 

  55. Morton, D. J. et al. The RNA exosome and RNA exosome-linked disease. RNA 24, 127–142 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Giunta, M. et al. Altered RNA metabolism due to a homozygous RBM7 mutation in a patient with spinal motor neuropathy. Hum. Mol. Genet. 25, 2985–2996 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Insco, M. L. et al. Oncogenic CDK13 mutations impede nuclear RNA surveillance. Science 380, eabn7625 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Luo, S. et al. The evolutionary arms race between transposable elements and piRNAs in Drosophila melanogaster. BMC Evol. Biol. 20, 14 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Bertozzi, T. M., Elmer, J. L., Macfarlan, T. S. & Ferguson-Smith, A. C. KRAB zinc finger protein diversification drives mammalian interindividual methylation variability. Proc. Natl Acad. Sci. USA 117, 31290–31300 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Fox, A. H. & Lamond, A. I. Paraspeckles. Cold Spring Harb. Perspect. Biol. 2, a000687 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Kaneko, H. et al. DICER1 deficit induces Alu RNA toxicity in age-related macular degeneration. Nature 471, 325–330 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Muotri, A. R. et al. L1 retrotransposition in neurons is modulated by MeCP2. Nature 468, 443–446 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Tsivion-Visbord, H. et al. Increased RNA editing in maternal immune activation model of neurodevelopmental disease. Nat. Commun. 11, 5236 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Ansell, B. R. E. et al. A survey of RNA editing at single-cell resolution links interneurons to schizophrenia and autism. RNA 27, 1482–1496 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Li, P. et al. Aicardi–Goutieres syndrome protein TREX1 suppresses L1 and maintains genome integrity through exonuclease-independent ORF1p depletion. Nucleic Acids Res. 45, 4619–4631 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Stearrett, N. et al. Expression of human endogenous retroviruses in systemic lupus erythematosus: multiomic integration with gene expression. Front. Immunol. 12, 661437 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Dembny, P. et al. Human endogenous retrovirus HERV-K(HML-2) RNA causes neurodegeneration through Toll-like receptors. JCI Insight 5, e131093 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Ramirez, P. et al. Pathogenic tau accelerates aging-associated activation of transposable elements in the mouse central nervous system. Prog. Neurobiol. 208, 102181 (2022).

    Article  CAS  PubMed  Google Scholar 

  69. Grundy, E. E., Diab, N. & Chiappinelli, K. B. Transposable element regulation and expression in cancer. FEBS J. 289, 1160–1179 (2022).

    Article  CAS  PubMed  Google Scholar 

  70. Van Meter, M. et al. SIRT6 represses LINE1 retrotransposons by ribosylating KAP1 but this repression fails with stress and age. Nat. Commun. 5, 5011 (2014).

    Article  PubMed  Google Scholar 

  71. Hastings, M. L. & Krainer, A. R. Pre-mRNA splicing in the new millennium. Curr. Opin. Cell Biol. 13, 302–309 (2001).

    Article  CAS  PubMed  Google Scholar 

  72. Liu, H. X., Cartegni, L., Zhang, M. Q. & Krainer, A. R. A mechanism for exon skipping caused by nonsense or missense mutations in BRCA1 and other genes. Nat. Genet. 27, 55–58 (2001).

    Article  CAS  PubMed  Google Scholar 

  73. Neri, F. et al. Intragenic DNA methylation prevents spurious transcription initiation. Nature 543, 72–77 (2017).

    Article  CAS  PubMed  Google Scholar 

  74. Ilinskaya, O. N. & Mahmud, R. S. Ribonucleases as antiviral agents. Mol. Biol. 48, 615–623 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Meola, N. et al. Identification of a nuclear exosome decay pathway for processed transcripts. Mol. Cell 64, 520–533 (2016).

    Article  CAS  PubMed  Google Scholar 

  76. Ogami, K. et al. An Mtr4/ZFC3H1 complex facilitates turnover of unstable nuclear RNAs to prevent their cytoplasmic transport and global translational repression. Genes. Dev. 31, 1257–1271 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Lubas, M. et al. Interaction profiling identifies the human nuclear exosome targeting complex. Mol. Cell 43, 624–637 (2011).

    Article  CAS  PubMed  Google Scholar 

  78. Chen, L. L., DeCerbo, J. N. & Carmichael, G. G. Alu element-mediated gene silencing. EMBO J. 27, 1694–1705 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Monaghan, L., Longman, D. & Cáceres, J. F. Translation-coupled mRNA quality control mechanisms. EMBO J. 42, e114378 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Anderson, P. & Kedersha, N. Stress granules: the Tao of RNA triage. Trends Biochem. Sci. 33, 141–150 (2008).

    Article  CAS  PubMed  Google Scholar 

  81. Ding, S. W. & Voinnet, O. Antiviral immunity directed by small RNAs. Cell 130, 413–426 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Gao, G. X., Guo, X. M. & Goff, S. P. Inhibition of retroviral RNA production by ZAP, a CCCH-type zinc finger protein. Science 297, 1703–1706 (2002).

    Article  CAS  PubMed  Google Scholar 

  83. Kesner, J. S. et al. Noncoding translation mitigation. Nature 617, 395–402 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Liu, J. et al. The RNA m6A reader YTHDC1 silences retrotransposons and guards ES cell identity. Nature 591, 322–326 (2021).

    Article  CAS  PubMed  Google Scholar 

  85. Ries, R. J., Pickering, B. F., Poh, H. X., Namkoong, S. & Jaffrey, S. R. m6A governs length-dependent enrichment of mRNAs in stress granules. Nat. Struct. Mol. Biol. 30, 1525–1535 (2023).

    Article  CAS  PubMed  Google Scholar 

  86. Lee, E. S. et al. N6-Methyladenosine (m6A) promotes the nuclear retention of mRNAs with intact 5′ splice site motifs. Preprint at bioRxiv https://doi.org/10.1101/2023.06.20.545713v2 (2023).

  87. He, P. C. et al. Exon architecture controls mRNA m6A suppression and gene expression. Science 379, 677–682 (2023).

    Article  CAS  PubMed  Google Scholar 

  88. Delaunay, S., Helm, M. & Frye, M. RNA modifications in physiology and disease: towards clinical applications. Nat. Rev. Genet. https://doi.org/10.1038/s41576-41023-00645-41572 (2023).

  89. Sun, T. et al. Crosstalk between RNA m6A and DNA methylation regulates transposable element chromatin activation and cell fate in human pluripotent stem cells. Nat. Genet. 55, 1324–1335 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Janeway, C. A. Jr Approaching the asymptote? Evolution and revolution in immunology. Cold Spring Harb. Symp. Quant. Biol. 54 Pt 1, 1–13 (1989).

    Article  CAS  PubMed  Google Scholar 

  91. Logsdon, J. M. The recent origins of spliceosomal introns revisited. Curr. Opin. Genet. Dev. 8, 637–648 (1998).

    Article  CAS  PubMed  Google Scholar 

  92. Sakharkar, M. K., Chow, V. T. & Kangueane, P. Distributions of exons and introns in the human genome. Silico Biol. 4, 387–393 (2004).

    CAS  Google Scholar 

  93. Zhang, J., Sun, X. L., Qian, Y. M., LaDuca, J. P. & Maquat, L. E. At least one intron is required for the nonsense-mediated decay of triosephosphate isomerase mRNA: a possible link between nuclear splicing and cytoplasmic translation. Mol. Cell Biol. 18, 5272–5283 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Le Hir, H., Nott, A. & Moore, M. J. How introns influence and enhance eukaryotic gene expression. Trends Biochem. Sci. 28, 215–220 (2003).

    Article  PubMed  Google Scholar 

  95. Brocke, K. S., Neu-Yilik, G., Gehring, N. H., Hentze, M. W. & Kulozik, A. E. The human intronless melanocortin 4-receptor gene is NMD insensitive. Hum. Mol. Genet. 11, 331–335 (2002).

    Article  CAS  PubMed  Google Scholar 

  96. Savisaar, R. & Hurst, L. D. Purifying selection on exonic splice enhancers in intronless genes. Mol. Biol. Evol. 33, 1396–1418 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Long, H. et al. Evolutionary determinants of genome-wide nucleotide composition. Nat. Ecol. Evol. 2, 237–240 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  98. Ho, A. T. & Hurst, L. D. Unusual mammalian usage of TGA stop codons reveals that sequence conservation need not imply purifying selection. PLoS Biol. 20, e3001588 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Charneski, C. A., Honti, F., Bryant, J. M., Hurst, L. D. & Feil, E. J. Atypical at skew in Firmicute genomes results from selection and not from mutation. PLoS Genet. 7, e1002283 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Seczynska, M., Bloor, S., Cuesta, S. M. & Lehner, P. J. Genome surveillance by HUSH-mediated silencing of intronless mobile elements. Nature 601, 440–445 (2022).

    Article  CAS  PubMed  Google Scholar 

  101. Duret, L. & Galtier, N. Biased gene conversion and the evolution of mammalian genomic landscapes. Annu. Rev. Genomics Hum. Genet. 10, 285–311 (2009).

    Article  CAS  PubMed  Google Scholar 

  102. Liu, H. et al. Tetrad analysis in plants and fungi finds large differences in gene conversion rates but no GC bias. Nat. Ecol. Evol. 2, 164–173 (2018).

    Article  PubMed  Google Scholar 

  103. Galtier, N. Gene conversion drives GC content evolution in mammalian histones. Trends Genet. 19, 65–68 (2003).

    Article  CAS  PubMed  Google Scholar 

  104. D’Onofrio, G., Mouchiroud, D., Aissani, B., Gautier, C. & Bernardi, G. Correlations between the compositional properties of human genes, codon usage, and amino-acid-composition of proteins. J. Mol. Evol. 32, 504–510 (1991).

    Article  PubMed  Google Scholar 

  105. Duret, L. & Hurst, L. D. The elevated GC content at exonic third sites is not evidence against neutralist models of isochore evolution. Mol. Biol. Evol. 18, 757–762 (2001).

    Article  CAS  PubMed  Google Scholar 

  106. Duret, L. & Galtier, N. The covariation between TpA deficiency, CpG deficiency, and G + C content of human isochores is due to a mathematical artifact. Mol. Biol. Evol. 17, 1620–1625 (2000).

    Article  CAS  PubMed  Google Scholar 

  107. Morales, A. C. et al. Causes and consequences of purifying selection on SARS-CoV-2. Genome Biol. Evol. 13, 17 (2021).

    Article  Google Scholar 

  108. Zhou, Z. et al. Codon usage is an important determinant of gene expression levels largely through its effects on transcription. Proc. Natl Acad. Sci. USA 113, E6117–E6125 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Newman, Z. R., Young, J. M., Ingolia, N. T. & Barton, G. M. Differences in codon bias and GC content contribute to the balanced expression of TLR7 and TLR9. Proc. Natl Acad. Sci. USA 113, E1362–E1371 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Zhao, F. et al. Genome-wide role of codon usage on transcription and identification of potential regulators. Proc. Natl Acad. Sci. USA 118, e2022590118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Vlaming, H., Mimoso, C. A., Field, A. R., Martin, B. J. E. & Adelman, K. Screening thousands of transcribed coding and non-coding regions reveals sequence determinants of RNA polymerase II elongation potential. Nat. Struct. Mol. Biol. 29, 613–620 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Pantier, R. et al. SALL4 controls cell fate in response to DNA base composition. Mol. Cell 81, 845–858 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Hisano, M., Ohta, H., Nishimune, Y. & Nozaki, M. Methylation of CpG dinucleotides in the open reading frame of a testicular germ cell-specific intronless gene, Tact1/Actl7b, represses its expression in somatic cells. Nucleic Acids Res. 31, 4797–4804 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Hodges, B. L., Taylor, K. M., Joseph, M. F., Bourgeois, S. A. & Scheule, R. K. Long-term transgene expression from plasmid DNA gene therapy vectors is negatively affected by CpG dinucleotides. Mol. Ther. 10, 269–278 (2004).

    Article  CAS  PubMed  Google Scholar 

  115. Parmley, J. L., Chamary, J. V. & Hurst, L. D. Evidence for purifying selection against synonymous mutations in mammalian exonic splicing enhancers. Mol. Biol. Evol. 23, 301–309 (2006).

    Article  CAS  PubMed  Google Scholar 

  116. Carlini, D. B. & Genut, J. E. Synonymous SNPs provide evidence for selective constraint on human exonic splicing enhancers. J. Mol. Evol. 62, 89–98 (2006).

    Article  CAS  PubMed  Google Scholar 

  117. Sauna, Z. E. & Kimchi-Sarfaty, C. Understanding the contribution of synonymous mutations to human disease. Nat. Rev. Genet. 12, 683–691 (2011).

    Article  CAS  PubMed  Google Scholar 

  118. Savisaar, R. & Hurst, L. D. Both maintenance and avoidance of RNA-binding protein interactions constrain coding sequence evolution. Mol. Biol. Evol. 34, 1110–1126 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Parmley, J. L. & Hurst, L. D. Exonic splicing regulatory elements skew synonymous codon usage near intron–exon boundaries in mammals. Mol. Biol. Evol. 24, 1600–1603 (2007).

    Article  CAS  PubMed  Google Scholar 

  120. Willie, E. & Majewski, J. Evidence for codon bias selection at the pre-mRNA level in eukaryotes. Trends Genet. 20, 534–538 (2004).

    Article  CAS  PubMed  Google Scholar 

  121. Warnecke, T. & Hurst, L. D. Evidence for a trade-off between translational efficiency and splicing regulation in determining synonymous codon usage in Drosophila melanogaster. Mol. Biol. Evol. 24, 2755–2762 (2007).

    Article  CAS  PubMed  Google Scholar 

  122. Eskesen, S. T., Eskesen, F. N. & Ruvinsky, A. Natural selection affects frequencies of AG and GT dinucleotides at the 5′ and 3′ ends of exons. Genetics 167, 543–550 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Livingstone, M. et al. Investigating DNA-, RNA-, and protein-based features as a means to discriminate pathogenic synonymous variants. Hum. Mutat. 38, 1336–1347 (2017).

    Article  CAS  PubMed  Google Scholar 

  124. Eyre-Walker, A. & Hurst, L. D. The evolution of isochores. Nat. Rev. Genet. 2, 549–555 (2001).

    Article  CAS  PubMed  Google Scholar 

  125. Potrzebowski, L. et al. Chromosomal gene movements reflect the recent origin and biology of therian sex chromosomes. PLoS Biol. 6, e80 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  126. Mordstein, C. et al. Transcription, mRNA export, and immune evasion shape the codon usage of viruses. Genome Biol. Evol. 13, evab106 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  127. Goodman, D. B., Church, G. M. & Kosuri, S. Causes and effects of N-terminal codon bias in bacterial genes. Science 342, 475–479 (2013).

    Article  CAS  PubMed  Google Scholar 

  128. Gu, W., Zhou, T. & Wilke, C. O. A universal trend of reduced mRNA stability near the translation-initiation site in prokaryotes and eukaryotes. PLoS Comput. Biol. 6, e1000664 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  129. Palazzo, A. F. et al. The signal sequence coding region promotes nuclear export of mRNA. PLoS Biol. 5, e322 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  130. Huang, Y., Gattoni, R., Stevenin, J. & Steitz, J. A. SR splicing factors serve as adapter proteins for TAP-dependent mRNA export. Mol. Cell 11, 837–843 (2003).

    Article  CAS  PubMed  Google Scholar 

  131. Huang, Y. & Steitz, J. A. Splicing factors SRp20 and 9G8 promote the nucleocytoplasmic export of mRNA. Mol. Cell 7, 899–905 (2001).

    Article  CAS  PubMed  Google Scholar 

  132. Courel, M. et al. GC content shapes mRNA storage and decay in human cells. eLife 8, e49708 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  133. Chen, C.-Y. et al. AU binding proteins recruit the exosome to degrade ARE-containing mRNAs. Cell 107, 451–464 (2001).

    Article  CAS  PubMed  Google Scholar 

  134. Namkoong, S., Ho, A., Woo, Y. M., Kwak, H. & Lee, J. H. Systematic characterization of stress-induced RNA granulation. Mol. Cell 70, 175–187.e8 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Khong, A. et al. The stress granule transcriptome reveals principles of mRNA accumulation in stress granules. Mol. Cell 68, 808–820.e5 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Takata, M. A. et al. CG dinucleotide suppression enables antiviral defence targeting non-self RNA. Nature 550, 124–127 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Duan, J. & Antezana, M. A. Mammalian mutation pressure, synonymous codon choice, and mRNA degradation. J. Mol. Evol. 57, 694–701 (2003).

    Article  CAS  PubMed  Google Scholar 

  138. Hrossova, D. et al. RBM7 subunit of the NEXT complex binds U-rich sequences and targets 3′-end extended forms of snRNAs. Nucleic Acids Res. 43, 4236–4248 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Lubas, M. et al. The human nuclear exosome targeting complex is loaded onto newly synthesized RNA to direct early ribonucleolysis. Cell Rep. 10, 178–192 (2015).

    Article  CAS  PubMed  Google Scholar 

  140. Pechmann, S. & Frydman, J. Evolutionary conservation of codon optimality reveals hidden signatures of cotranslational folding. Nat. Struct. Mol. Biol. 20, 237–243 (2013).

    Article  CAS  PubMed  Google Scholar 

  141. Kimchi-Sarfaty, C. et al. A “silent” polymorphism in the MDR1 gene changes substrate specificity. Science 315, 525–528 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Radhakrishnan, A. & Green, R. Connections underlying translation and mRNA stability. J. Mol. Biol. 428, 3558–3564 (2016).

    Article  CAS  PubMed  Google Scholar 

  144. Buschauer, R. et al. The Ccr4–Not complex monitors the translating ribosome for codon optimality. Science 368, eaay6912 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Medina-Munoz, S. G. et al. Crosstalk between codon optimality and cis-regulatory elements dictates mRNA stability. Genome Biol. 22, 14 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Shu, H. et al. FMRP links optimal codons to mRNA stability in neurons. Proc. Natl Acad. Sci. USA 117, 30400–30411 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Kumar, A. et al. The slowing rate of CpG depletion in SARS-CoV-2 genomes is consistent with adaptations to the human host. Mol. Biol. Evol. 39, msac029 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Ficarelli, M. et al. CpG dinucleotides inhibit HIV-1 replication through zinc finger antiviral protein (ZAP)-dependent and -independent mechanisms. J. Virol. 94, e01337-19 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  149. Hurst, L. D. et al. A simple metric of promoter architecture robustly predicts expression breadth of human genes suggesting that most transcription factors are positive regulators. Genome Biol. 15, 413 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  150. Bestor, T. H. DNA methylation: evolution of a bacterial immune function into a regulator of gene expression and genome structure in higher eukaryotes. Philos. Trans. R. Soc. Lond. B Biol. Sci. 326, 179–187 (1990).

    Article  CAS  PubMed  Google Scholar 

  151. Voo, K. S., Carlone, D. L., Jacobsen, B. M., Flodin, A. & Skalnik, D. G. Cloning of a mammalian transcriptional activator that binds unmethylated CpG motifs and shares a CXXC domain with DNA methyltransferase, human trithorax, and methyl-CpG binding domain protein 1. Mol. Cell Biol. 20, 2108–2121 (2000).

    Article  CAS  PubMed  Google Scholar 

  152. Bauer, A. P. et al. The impact of intragenic CpG content on gene expression. Nucleic Acids Res. 38, 3891–3908 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Singer, T., McConnell, M. J., Marchetto, M. C., Coufal, N. G. & Gage, F. H. LINE-1 retrotransposons: mediators of somatic variation in neuronal genomes? Trends Neurosci. 33, 345–354 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Singh, M. et al. A new human embryonic cell type associated with activity of young transposable elements allows definition of the inner cell mass. PLoS Biol. 21, e3002162 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Mirihana Arachchilage, G., Hetti Arachchilage, M., Venkataraman, A., Piontkivska, H. & Basu, S. Stable G-quadruplex enabling sequences are selected against by the context-dependent codon bias. Gene 696, 149–161 (2019).

    Article  CAS  PubMed  Google Scholar 

  156. Varshney, D., Spiegel, J., Zyner, K., Tannahill, D. & Balasubramanian, S. The regulation and functions of DNA and RNA G-quadruplexes. Nat. Rev. Mol. Cell Biol. 21, 459–474 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Wang, Y., Qiu, C. & Cui, Q. A large-scale analysis of the relationship of synonymous SNPs changing microRNA regulation withfunctionality and disease. Int. J. Mol. Sci. 16, 23545–23555 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Brest, P. et al. A synonymous variant in IRGM alters a binding site for miR-196 and causes deregulation of IRGM-dependent xenophagy in Crohn’s disease. Nat. Genet. 43, 242–245 (2011).

    Article  CAS  PubMed  Google Scholar 

  159. Gartner, J. J. et al. Whole-genome sequencing identifies a recurrent functional synonymous mutation in melanoma. Proc. Natl Acad. Sci. USA 110, 13481–13486 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Hamdorf, M. et al. miR-128 represses L1 retrotransposition by binding directly to L1 RNA. Nat. Struct. Mol. Biol. 22, 824–831 (2015).

    Article  CAS  PubMed  Google Scholar 

  161. Eyre-Walker, A. & Keightley, P. D. The distribution of fitness effects of new mutations. Nat. Rev. Genet. 8, 610–618 (2007).

    Article  CAS  PubMed  Google Scholar 

  162. Fairbrother, W. G., Holste, D., Burge, C. B. & Sharp, P. A. Single nucleotide polymorphism-based validation of exonic splicing enhancers. PLoS Biol. 2, E268 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

  163. Chamary, J. V. & Hurst, L. D. Evidence for selection on synonymous mutations affecting stability of mRNA secondary structure in mammals. Genome Biol. 6, R75 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Nackley, A. G. et al. Human catechol-O-methyltransferase haplotypes modulate protein expression by altering mRNA secondary structure. Science 314, 1930–1933 (2006).

    Article  CAS  PubMed  Google Scholar 

  165. Schattner, P. & Diekhans, M. Regions of extreme synonymous codon selection in mammalian genes. Nucleic Acids Res. 34, 1700–1710 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Wu, X. M. & Hurst, L. D. Why selection might be stronger when populations are small: intron size and density predict within and between-species usage of exonic splice associated cis-motifs. Mol. Biol. Evol. 32, 1847–1861 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  167. Prats-Ejarque, G., Lu, L., Salazar, V. A., Moussaoui, M. & Boix, E. Evolutionary trends in RNA base selectivity within the RNase A superfamily. Front. Pharmacol. 10, 1170 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Mendell, J. T., Sharifi, N. A., Meyers, J. L., Martinez-Murillo, F. & Dietz, H. C. Nonsense surveillance regulates expression of diverse classes of mammalian transcripts and mutes genomic noise. Nat. Genet. 36, 1073–1078 (2004).

    Article  CAS  PubMed  Google Scholar 

  169. Torres, M. et al. Paraspeckles as rhythmic nuclear mRNA anchorages responsible for circadian gene expression. Nucleus 8, 249–254 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Prasanth, K. V. et al. Regulating gene expression through RNA nuclear retention. Cell 123, 249–263 (2005).

    Article  CAS  PubMed  Google Scholar 

  171. Lucks, J. B., Nelson, D. R., Kudla, G. R. & Plotkin, J. B. Genome landscapes and bacteriophage codon usage. PLoS Comput. Biol. 4, e1000001 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  172. De Vlugt, C., Sikora, D. & Pelchat, M. Insight into influenza: a virus cap-snatching. Viruses 10, 641 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  173. Jalkanen, A. L., Coleman, S. J. & Wilusz, J. Determinants and implications of mRNA poly(A) tail size—does this protein make my tail look big? Semin. Cell Dev. Biol. 34, 24–32 (2014).

    Article  CAS  PubMed  Google Scholar 

  174. Mauro, V. P. Codon optimization in the production of recombinant biotherapeutics: potential risks and considerations. Biodrugs 32, 69–81 (2018).

    Article  CAS  PubMed  Google Scholar 

  175. Ho, A. T. & Hurst, L. D. Effective population size predicts local rates but not local mitigation of read-through errors. Mol. Biol. Evol. 38, 244–262 (2021).

    Article  CAS  PubMed  Google Scholar 

  176. Allert, M., Cox, J. C. & Hellinga, H. W. Multifactorial determinants of protein expression in prokaryotic open reading frames. J. Mol. Biol. 402, 905–918 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Wu, X. & Hurst, L. D. Determinants of the usage of splice-associated cis-motifs predict the distribution of human pathogenic SNPs. Mol. Biol. Evol. 33, 518–529 (2016).

    Article  CAS  PubMed  Google Scholar 

  178. Abrahams, L. et al. Evidence in disease and non-disease contexts that nonsense mutations cause altered splicing via motif disruption. Nucleic Acids Res. 49, 9665–9685 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Soemedi, R. et al. Pathogenic variants that alter protein code often disrupt splicing. Nat. Genet. 49, 848–855 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Mühlhausen, S. & Hurst, L. D. Transgene-design: a web application for the design of mammalian transgenes. Bioinformatics 38, 2626–2627 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  181. Sharp, C. P. et al. CpG dinucleotide enrichment in the influenza A virus genome as a live attenuated vaccine development strategy. PLoS Pathog. 19, e1011357 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Yew, N. S. et al. CpG-depleted plasmid DNA vectors with enhanced safety and long-term gene expression in vivo. Mol. Ther. 5, 731–738 (2002).

    Article  CAS  PubMed  Google Scholar 

  183. Kariko, K., Buckstein, M., Ni, H. & Weissman, D. Suppression of RNA recognition by Toll-like receptors: the impact of nucleoside modification and the evolutionary origin of RNA. Immunity 23, 165–175 (2005).

    Article  CAS  PubMed  Google Scholar 

  184. Vaidyanathan, S. et al. Uridine depletion and chemical modification increase Cas9 mRNA activity and reduce immunogenicity without HPLC purification. Mol. Ther. Nucleic Acids 12, 530–542 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Ohta, T. The nearly neutral theory of molecular evolution. Ann. Rev. Ecol. Syst. 23, 263–286 (1992).

    Article  Google Scholar 

  186. Christmas, M. J. et al. Evolutionary constraint and innovation across hundreds of placental mammals. Science 380, eabn3943 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Lynch, M. et al. Genetic drift, selection and the evolution of the mutation rate. Nat. Rev. Genet. 17, 704–714 (2016).

    Article  CAS  PubMed  Google Scholar 

  188. Sung, W., Ackerman, M. S., Miller, S. F., Doak, T. G. & Lynch, M. Drift-barrier hypothesis and mutation-rate evolution. Proc. Natl Acad. Sci. USA 109, 18488–18492 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Seczynska, M. & Lehner, P. J. The sound of silence: mechanisms and implications of HUSH complex function. Trends Genet. 39, 251–267 (2023).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank the Humboldt Foundation for the award of the Humboldt Prize to L.D.H. to enable him to spend time in Germany. S.R. is funded by the Evolution Education Trust.

Author information

Authors and Affiliations

Authors

Contributions

L.D.H. and S.R. researched data for the article. All authors contributed substantially to discussion of the content. L.D.H. wrote the article. All authors reviewed and/or edited the manuscript before submission. Z.I. contributed to discussion of issues regarding transposable elements specifically.

Corresponding author

Correspondence to Laurence D. Hurst.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Reviews Genetics thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

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

Supplementary information

Glossary

Codon optimality-mediated RNA decay

A process by which mRNAs with abundant non-optimal codons (those matching rare tRNAs) are subject to decay, probably mediated by slow ribosomal progression.

Codon usage bias

A pattern in which synonymous codons occur at frequencies different from some null expectation (often the null expectation is of equal frequency).

CpG islands

Long (>200 bp) stretches of DNA rich in CpG dinucleotides that commonly occur in the vicinity of mammalian promoters. They are commonly unmethylated if the associated gene is expressed.

Cryptic splice sites

Splice sites within exons away from the canonical splice site that can result in alternative splicing.

Effective population size

(Ne). The number of individuals who an idealized neutrally evolving population would require for it to have properties (such as genetic diversity) equivalent to the real population. Ne is often smaller than the census population size (N). In simple cases, it approximates to the number of breeding individuals.

Exonic splice enhancers

(ESEs). Short (6–8 bp) RNA motifs that enable more accurate splicing, particularly in the vicinity of weak splice sites. ESEs often function as RNA-binding sites for serine–arginine-rich proteins.

Exon-junction complex

A protein complex that forms on a pre-mRNA molecule at the junction between two exons joined by RNA splicing.

Exosome complex

A multi-protein intracellular complex that degrades many types of RNA. In eukaryotes, it is present in the cytoplasm, the nucleus and the nucleolus. In humans, the cytoplasmic form is associated with DIS3L, whereas the nuclear complex contains DIS3. The nuclear form is termed the nuclear exosome complex, that which is targeted by the nuclear exosome targeting (NEXT) complex. The (RNA) exosome complex is not to be confused with exosomes, extracellular vesicles generated by cells.

GC-biased gene conversion

(gBGC). Biased gene conversion describes the recombination of short stretches of genetic material from a donor sequence to an acceptor sequence that is biased as to which is the donor strand and which is the acceptor strand. In gBGC, AT:GC mismatches in recombinant sections are resolved with a preference towards G or C.

G-Quadruplexes

Stable secondary structures formed in guanine-rich DNA and RNA sequences.

Inverted repeat Alu elements

(IRAlus). Single-stranded Alu elements that are followed downstream by their reverse complement. This characteristic often allows inverted repeats to fold on themselves and form double-stranded structures.

LINE1

(L1). LINE elements, of which L1 is an example, are the most abundant class of transposable elements in the human genome, formed by autonomous retrotransposition.

Mutational equilibrium

The frequency (for example, of nucleotides) in a population at which, if only mutation bias and neutral evolution affect the frequency, the frequency does not change.

Neutral evolution

The process by which allele frequency is determined by chance events alone operating on alleles of the same fitness.

No-go mRNA decay

A process by which RNAs with stacks of stalled ribosomes are degraded.

Nonsense-mediated decay

(NMD). A process by which mRNA molecules containing premature stop codons are degraded.

Non-stop RNA decay

A process by which mRNAs without a proper stop codon are identified and targeted for decay. In eukaryotes, this process discharges ribosomes stalled at the 3′ end of mRNAs, directing those mRNAs to the exosome complex.

Processing bodies

Ribonucleoprotein bodies found in the cytoplasm that retain mRNA molecules and contain proteins required for mRNA decay. A similar role is carried out by cytoplasmic stress granules.

Retrogenes

Processed copies of genes formed from reverse transcription of mature mRNA molecules of a parental gene (hence without introns).

Spurious transcripts

Transcripts generated by functionally irrelevant cellular events (such as transcription factor binding to random sequence).

Synonymous mutations

Single base-pair mutations in a protein-coding exon that change the codon to a different but synonymous one and hence do not a priori change the amino acid sequence of the encoded protein.

Translational selection

Selection that favours synonymous mutations owing to their effects on translation, typically assumed to be mediated by faster or more accurate translation. Commonly evidenced by a codon usage bias that favours synonymous codons matching more abundant iso-acceptor tRNAs.

Transposable elements

DNA sequences that can move within genomes and replicate without depending on gene replication of the host cell.

Zinc finger antiviral protein

(ZAP). A protein involved in preventing retroviral infection by binding of CpG-rich sequences and recruitment of the RNA degradation machinery.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Radrizzani, S., Kudla, G., Izsvák, Z. et al. Selection on synonymous sites: the unwanted transcript hypothesis. Nat Rev Genet (2024). https://doi.org/10.1038/s41576-023-00686-7

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41576-023-00686-7

Search

Quick links

Nature Briefing

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

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