Skip to main content

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

  • Review Article
  • Published:

Understanding the contribution of synonymous mutations to human disease

Key Points

  • Synonymous mutations, once thought to be 'silent', are now increasingly acknowledged to be able to cause changes in protein expression, conformation and function.

  • Studies of the association of genetic variants with disease have revealed a substantial contribution of synonymous mutations to human disease risk and other complex traits. This article includes a compendium of human diseases or clinical conditions associated with synonymous mutations.

  • Concomitantly, there has been substantial progress in understanding the mechanisms by which synonymous substitutions effect changes in the phenotype.

  • Synonymous mutations can affect protein conformation and function by affecting post-transcriptional processing and regulation of RNA, altering the local and global structure of the mRNA and influencing the kinetics of translation.

  • Recent estimates suggest that 5–10% of human genes contain at least one region where synonymous mutations could be harmful. Thus, synonymous SNPs identified in genome-wide association studies should be included in follow-up functional and mechanistic studies.

  • Understanding the effect (or effects) of synonymous mutations could have important implications in the practice of medicine and biotechnology.

Abstract

Synonymous mutations — sometimes called 'silent' mutations — are now widely acknowledged to be able to cause changes in protein expression, conformation and function. The recent increase in knowledge about the association of genetic variants with disease, particularly through genome-wide association studies, has revealed a substantial contribution of synonymous SNPs to human disease risk and other complex traits. Here we review current understanding of the extent to which synonymous mutations influence disease, the various molecular mechanisms that underlie these effects and the implications for future research and biomedical applications.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Molecular mechanisms by which synonymous mutations can change in protein levels or conformation.

Similar content being viewed by others

References

  1. Anfinsen, C. B. Principles that govern folding of protein chains. Science 181, 223–230 (1973).

    Article  CAS  PubMed  Google Scholar 

  2. Kimura, M. Preponderance of synonymous changes as evidence for neutral theory of molecular evolution. Nature 267, 275–276 (1977).

    Article  CAS  PubMed  Google Scholar 

  3. Chamary, J. V., Parmley, J. L. & Hurst, L. D. Hearing silence: non-neutral evolution at synonymous sites in mammals. Nature Rev. Genet. 7, 98–108 (2006). This is an excellent early Review from a molecular evolution perspective that brought the importance of synonymous mutations on human health and disease to a larger audience.

    Article  CAS  PubMed  Google Scholar 

  4. Plotkin, J. B. & Kudla, G. Synonymous but not the same: the causes and consequences of codon bias. Nature Rev. Genet. 12, 32–42 (2011). This is a recent Review that explains global patterns in codon bias and how they influence protein folding.

    Article  CAS  PubMed  Google Scholar 

  5. Cartegni, L., Chew, S. L. & Krainer, A. R. Listening to silence and understanding nonsense: exonic mutations that affect splicing. Nature Rev. Genet. 3, 285–298 (2002).

    Article  CAS  PubMed  Google Scholar 

  6. Nackley, A. G. et al. Human catechol-O-methyltransferase haplotypes modulate protein expression by altering mRNA secondary structure. Science 314, 1930–1933 (2006). The first study to elucidate a detailed molecular mechanism based on mRNA structure for how synonymous mutations can have physiological consequences.

    Article  CAS  PubMed  Google Scholar 

  7. Kimchi-Sarfaty, C. et al. A “silent” polymorphism in the MDR1 gene changes substrate specificity. Science 315, 525–528 (2007). The first study to provide evidence that synonymous changes that do not affect mRNA levels have clinical consequences.

    Article  CAS  PubMed  Google Scholar 

  8. Gunderson, K. L., Steemers, F. J., Lee, G., Mendoza, L. G. & Chee, M. S. A genome-wide scalable SNP genotyping assay using microarray technology. Nature Genet. 37, 549–554 (2005).

    Article  CAS  PubMed  Google Scholar 

  9. Kennedy, G. et al. Large-scale genotyping of complex DNA. Am. J. Hum. Genet. 71, 204 (2002).

    Article  Google Scholar 

  10. Matsuzaki, H. et al. Genotyping over 100,000 SNPs on a pair of oligonucleotide arrays. Nature Methods 1, 109–111 (2004).

    Article  CAS  PubMed  Google Scholar 

  11. Steemers, F. J. et al. Whole-genome genotyping with the single-base extension assay. Nature Methods 3, 31–33 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Manolio, T. A., Brooks, L. D. & Collins, F. S. A HapMap harvest of insights into the genetics of common disease. J. Clin. Invest. 118, 1590–1605 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Stark, K. et al. Genetic association study identifies HSPB7 as a risk gene for idiopathic dilated cardiomyopathy. PLoS Genet. 6, e1001167 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Mannisto, P. T. & Kaakkola, S. Catechol-O-methyltransferase (COMT): biochemistry, molecular biology, pharmacology, and clinical efficacy of the new selective COMT inhibitors. Pharmacol. Rev. 51, 593–628 (1999).

    CAS  PubMed  Google Scholar 

  15. Winterer, G. & Goldman, D. Genetics of human prefrontal function. Brain Res. Rev. 43, 134–163 (2003).

    Article  CAS  PubMed  Google Scholar 

  16. Ellis, P. E., Dawson, M. & Dixon, M. J. Mutation testing in Treacher Collins syndrome. J. Orthodont. 29, 293–297 (2011).

    Article  Google Scholar 

  17. Macaya, D. et al. A synonymous mutation in TCOF1 causes Treacher Collins syndrome due to mis-splicing of a constitutive exon. Am. J. Med. Genet. A 149, 1624–1627 (2009).

    Article  CAS  Google Scholar 

  18. Bartoszewski, R. A. et al. A synonymous single nucleotide polymorphism in ΔF508 CFTR alters the secondary structure of the mRNA and the expression of the mutant protein. J. Biol. Chem. 285, 28741–28748 (2010). This paper reveals a very interesting finding. A common amino acid deletion has long been thought to be responsible for cystic fibrosis; this study shows that a coincident synonymous mutation results in a change in mRNA structure and protein level and may be responsible for the polymorphism.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Ramser, J. et al. Rare missense and synonymous variants in UBE1 are associated with X-linked infantile spinal muscular atrophy. Am. J. Hum. Genet. 82, 188–193 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. O'Driscoll, M., Ruiz-Perez, V. L., Woods, C. G., Jeggo, P. A. & Goodship, J. A. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nature Genet. 33, 497–501 (2003).

    Article  CAS  PubMed  Google Scholar 

  21. Ho, P. A. et al. WT1 synonymous single nucleotide polymorphism rs16754 correlates with higher mRNA expression and predicts significantly improved outcome in favorable-risk pediatric acute myeloid leukemia: a report from the children's oncology group. J. Clin. Oncol. 29, 704–711 (2011).

    Article  CAS  PubMed  Google Scholar 

  22. Hassan, H. E., Myers, A. L., Coop, A. & Eddington, N. D. Differential involvement of P-glycoprotein (ABCB1) in permeability, tissue distribution, and antinociceptive activity of methadone, buprenorphine, and diprenorphine: in vitro and in vivo evaluation. J. Pharm. Sci. 98, 4928–4940 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ambudkar, S. V., Kim, I. W. & Sauna, Z. E. The power of the pump: mechanisms of action of P-glycoprotein (ABCB1). Eur. J. Pharm. Sci. 27, 392–400 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. van der Veldt, A. A. M. et al. Genetic polymorphisms associated with a prolonged progression-free survival in patients with metastatic renal cell cancer treated with sunitinib. Clin. Cancer Res. 17, 620–629 (2011).

    Article  CAS  PubMed  Google Scholar 

  25. Herrlinger, K. R. et al. ABCB1 single-nucleotide polymorphisms determine tacrolimus response in patients with ulcerative colitis. Clin. Pharmacol. Ther. 89, 422–428 (2011).

    Article  CAS  PubMed  Google Scholar 

  26. Sauna, Z. E., Kim, I. W. & Ambudkar, S. V. Genomics and the mechanism of P-glycoprotein (ABCB1). J. Bioenerg. Biomembr. 39, 481–487 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. Chen, R., Davydov, E. V., Sirota, M. & Butte, A. J. Non-synonymous and synonymous coding SNPs show similar likelihood and effect size of human disease association. PLoS ONE 5, e13574 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. 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 

  29. Chamary, J. V. & Hurst, L. D. The price of silent mutations. Sci. Am. 300, 46–53 (2009).

    Article  CAS  PubMed  Google Scholar 

  30. Schwanhausser, B. et al. Global quantification of mammalian gene expression control. Nature 473, 337–342 (2011).

    Article  CAS  PubMed  Google Scholar 

  31. Vogel, C. et al. Sequence signatures and mRNA concentration can explain two-thirds of protein abundance variation in a human cell line. Mol. Syst. Biol. 6, 400 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Czech, A., Fedyunin, I., Zhang, G. & Ignatova, Z. Silent mutations in sight: co-variations in tRNA abundance as a key to unravel consequences of silent mutations. Mol. Biosyst. 6, 1767–1772 (2010).

    Article  CAS  PubMed  Google Scholar 

  33. Pagani, F., Raponi, M. & Baralle, F. E. Synonymous mutations in CFTR exon 12 affect splicing and are not neutral in evolution. Proc. Natl Acad. Sci. USA 102, 6368–6372 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wang, G. S. & Cooper, T. A. Splicing in disease: disruption of the splicing code and the decoding machinery. Nature Rev. Genet. 8, 749–761 (2007).

    Article  CAS  PubMed  Google Scholar 

  35. Friedman, R. C., Farh, K. K. H., Burge, C. B. & Bartel, D. P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 19, 92–105 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. 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. Nature Genet. 43, 242–245 (2011). This study demonstrates the presence of an miRNA binding site in an exon, a synonymous change in which alters suceptibility to Crohn's disease.

    Article  CAS  PubMed  Google Scholar 

  37. Parkes, M. et al. Sequence variants in the autophagy gene IRGM and multiple other replicating loci contribute to Crohn's disease susceptibility. Nature Genet. 39, 830–832 (2007).

    Article  CAS  PubMed  Google Scholar 

  38. Duan, J. B. et al. Synonymous mutations in the human dopamine receptor D2 (DRD2) affect mRNA stability and synthesis of the receptor. Hum. Mol. Genet. 12, 205–216 (2003).

    Article  CAS  PubMed  Google Scholar 

  39. Kudla, G., Lipinski, L., Caffin, F., Helwak, A. & Zylicz, M. High guanine and cytosine content increases mRNA levels in mammalian cells. PLoS Biol. 4, 933–942 (2006).

    Article  CAS  Google Scholar 

  40. Gu, W. J., 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  CAS  Google Scholar 

  41. Kudla, G., Murray, A. W., Tollervey, D. & Plotkin, J. B. Coding-sequence determinants of gene expression in Escherichia coli. Science 324, 255–258 (2009). This important study, which was carried out using synthetic genes, shows the importance mRNA structure near the start codon and the consequenes of synonymous mutations in this region.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Studer, S. M. & Joseph, S. Unfolding of mRNA secondary structure by the bacterial translation initiation complex. Mol. Cell 22, 105–115 (2006).

    Article  CAS  PubMed  Google Scholar 

  43. Dong, H. J., Nilsson, L. & Kurland, C. G. Co-variation of tRNA abundance and codon usage in Escherichia coli at different growth rates. J. Mol. Biol. 260, 649–663 (1996).

    Article  CAS  PubMed  Google Scholar 

  44. Duret, L. tRNA gene number and codon usage in the C. elegans genome are co-adapted for optimal translation of highly expressed genes. Trends Genet. 16, 287–289 (2000).

    Article  CAS  PubMed  Google Scholar 

  45. Zhang, G., Hubalewska, M. & Ignatova, Z. Transient ribosomal attenuation coordinates protein synthesis and co-translational folding. Nature Struct. Mol. Biol. 16, 274–280 (2009).

    Article  CAS  Google Scholar 

  46. Bonekamp, F., Dalboge, H., Christensen, T. & Jensen, K. F. Translation rates of individual codons are not correlated with transfer-RNA abundances or with frequencies of utilization in Escherichia coli. J. Bacteriol. 171, 5812–5816 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Higgs, P. G. & Ran, W. Q. Coevolution of codon usage and tRNA genes leads to alternative stable states of biased codon usage. Mol. Biol. Evol. 25, 2279–2291 (2008).

    Article  CAS  PubMed  Google Scholar 

  48. Ikemura, T. Codon usage and transfer-RNA content in unicellular and multicellular organisms. Mol. Biol. Evol. 2, 13–34 (1985).

    CAS  PubMed  Google Scholar 

  49. Komar, A. A. A pause for thought along the co-translational folding pathway. Trends Biochem. Sci. 34, 16–24 (2009).

    Article  CAS  PubMed  Google Scholar 

  50. Purvis, I. J. et al. The efficiency of folding of some proteins is increased by controlled rates of translation in vivo — a hypothesis. J. Mol. Biol. 193, 413–417 (1987). This study presents the first challenge to the concept that only the sequence of amino acids determines the final conformation of a protein.

    Article  CAS  PubMed  Google Scholar 

  51. Komar, A. A., Lesnik, T. & Reiss, C. Synonymous codon substitutions affect ribosome traffic and protein folding during in vitro translation. FEBS Lett. 462, 387–391 (1999).

    Article  CAS  PubMed  Google Scholar 

  52. Fedorov, A. N. & Baldwin, T. O. Contribution of cotranslational folding to the rate of formation of native protein-structure. Proc. Natl Acad. Sci. USA 92, 1227–1231 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Clarke, D. T., Doig, A. J., Stapley, B. J. & Jones, G. R. The α-helix folds on the millisecond time scale. Proc. Natl Acad. Sci. USA 96, 7232–7237 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Tuller, T. et al. An evolutionarily conserved mechanism for controlling the efficiency of protein translation. Cell 141, 344–354 (2010).

    Article  CAS  PubMed  Google Scholar 

  55. Ehrenberg, M., Dennis, P. P. & Bremer, H. Maximum rrn promoter activity in Escherichia coli at saturating concentrations of free RNA polymerase. Biochimie 92, 12–20 (2010).

    Article  CAS  PubMed  Google Scholar 

  56. Powers, E. T. & Balch, W. E. Costly mistakes: translational infidelity and protein homeostasis. Cell 134, 204–206 (2008).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Cannarrozzi, G. et al. A role for codon order in translation dynamics. Cell 141, 355–367 (2010).

    Article  CAS  PubMed  Google Scholar 

  59. Deutscher, M. P. The eukaryotic aminoacyl-transfer RNA-synthetase complex — suggestions for its structure and function. J. Cell Biol. 99, 373–377 (1984).

    Article  CAS  PubMed  Google Scholar 

  60. Kaminska, M. et al. Dynamic organization of aminoacyl-tRNA synthetase complexes in the cytoplasm of human cells. J. Biol. Chem. 284, 13746–13754 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Mirande, M., Lecorre, D. & Waller, J. P. A complex from cultured Chinese-hamster ovary cells containing 9 aminoacyl-transfer RNA-synthetases — thermolabile leucyl-transfer RNA-synthetase from the Tsh1 mutant-cell line is an integral component of this complex. Eur. J. Biochem. 147, 281–289 (1985).

    Article  CAS  PubMed  Google Scholar 

  62. Sosnick, T. R. Kinetic barriers and the role of topology in protein and RNA folding. Protein Sci. 17, 1308–1318 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Jha, S. & Komar, A. A. Birth, life and death of nascent polypeptide chains. Biotechnol. J. 6, 623–640 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Fedyukina, D. V. & Cavagnero, S. Protein folding at the exit tunnel. Annu. Rev. Biophys. 40, 337–359 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Netzer, W. J. & Hartl, F. U. Recombination of protein domains facilitated by co-translational folding in eukaryotes. Nature 388, 343–349 (1997).

    Article  CAS  PubMed  Google Scholar 

  66. Nicola, A. V., Chen, W. & Helenius, A. Co-translational folding of an alphavirus capsid protein in the cytosol of living cells. Nature Cell Biol. 1, 341–345 (1999).

    Article  CAS  PubMed  Google Scholar 

  67. Frydman, J., Erdjument-Bromage, H., Tempst, P. & Hartl, F. U. Co-translational domain folding as the structural basis for the rapid de novo folding of firefly luciferase. Nature Struct. Biol. 6, 697–705 (1999).

    Article  CAS  PubMed  Google Scholar 

  68. Tsai, C. J. et al. Synonymous mutations and ribosome stalling can lead to altered folding pathways and distinct minima. J. Mol. Biol. 383, 281–291 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Sauna, Z. E., Kimchi-Sarfaty, C., Ambudkar, S. V. & Gottesman, M. M. Silent polymorphisms speak: how they affect pharmacogenomics and the treatment of cancer. Cancer Res. 67, 9609–9612 (2007).

    Article  CAS  PubMed  Google Scholar 

  70. Balch, W. E., Morimoto, R. I., Dillin, A. & Kelly, J. W. Adapting proteostasis for disease intervention. Science 319, 916–919 (2008).

    Article  CAS  PubMed  Google Scholar 

  71. Gidalevitz, T., Ben-Zvi, A., Ho, K. H., Brignull, H. R. & Morimoto, R. I. Progressive disruption of cellular protein folding in models of polyglutamine diseases. Science 311, 1471–1474 (2006).

    Article  CAS  PubMed  Google Scholar 

  72. Pal, C., Papp, B. & Lercher, M. J. An integrated view of protein evolution. Nature Rev. Genet. 7, 337–348 (2006).

    Article  CAS  PubMed  Google Scholar 

  73. Wilke, C. O. & Drummond, D. A. Signatures of protein biophysics in coding sequence evolution. Curr. Opin. Struct. Biol. 20, 385–389 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Drummond, D. A. & Wilke, C. O. Mistranslation-induced protein misfolding as a dominant constraint on coding-sequence evolution. Cell 134, 341–352 (2008). This is an interesting global computational study conducted in a range of taxa using evolutionary simulation. It shows that selective pressures are exerted as a consequence of protein misfolding resulting from synonymous changes.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Cohen, E., Bieschke, J., Perciavalle, R. M., Kelly, J. W. & Dillin, A. Opposing activities protect against age-onset proteotoxicity. Science 313, 1604–1610 (2006).

    Article  CAS  PubMed  Google Scholar 

  76. Krejsa, C., Rogge, M. & Sadee, W. Protein therapeutics: new applications for pharmacogenetics. Nature Rev. Drug Discov. 5, 507–521 (2006).

    Article  CAS  Google Scholar 

  77. Ward, N. J. et al. Codon optimization of human factor VIII cDNAs leads to high-level expression. Blood 117, 798–807 (2011).

    Article  CAS  PubMed  Google Scholar 

  78. Maertens, B. et al. Gene optimization mechanisms: a multi-gene study reveals a high success rate of full-length human proteins expressed in Escherichia coli. Protein Sci. 19, 1312–1326 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. De Groot, A. S. & Scott, D. W. Immunogenicity of protein therapeutics. Trends Immunol. 28, 482–490 (2007).

    Article  CAS  PubMed  Google Scholar 

  80. Wang, J. H. et al. Neutralizing antibodies to therapeutic enzymes: considerations for testing, prevention and treatment. Nature Biotech. 26, 901–908 (2008).

    Article  CAS  Google Scholar 

  81. Wen, J. D. et al. Following translation by single ribosomes one codon at a time. Nature 452, 598–603 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Eden, E. et al. Proteome half-life dynamics in living human cells. Science 331, 764–768 (2011).

    Article  CAS  PubMed  Google Scholar 

  83. Plinke, C. et al. embCAB sequence variation among ethambutol-resistant Mycobacterium tuberculosis isolates without embB306 mutation. J. Antimicrob. Chemother. 65, 1359–1367 (2010).

    Article  CAS  PubMed  Google Scholar 

  84. Dissanayeke, S. R. et al. Polymorphic variation in TIRAP is not associated with susceptibility to childhood TB but may determine susceptibility to TBM in some ethnic groups. PLoS ONE 4, e6698 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Geslain, R. & Pan, T. Functional analysis of human tRNA isodecoders. J. Mol. Biol. 396, 821–831 (2010).

    Article  CAS  PubMed  Google Scholar 

  86. Gustilo, E. M., Franck, A. P. F. & Agris, P. F. tRNA's modifications bring order to gene expression. Curr. Opin. Microbiol. 11, 134–140 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Glinskii, A. B. et al. Identification of intergenic trans-regulatory RNAs containing a disease-linked SNP sequence and targeting cell cycle progression/differentiation pathways in multiple common human disorders. Cell Cycle 8, 3925–3942 (2009).

    Article  CAS  PubMed  Google Scholar 

  88. Treutlein, J. et al. Genome-wide association study of alcohol dependence. Arch. Gen. Psychiat. 66, 773–784 (2009).

    Article  CAS  PubMed  Google Scholar 

  89. Haiman, C. A. et al. Multiple regions within 8q24 independently affect risk for prostate cancer. Nature Genet. 39, 638–644 (2007).

    Article  CAS  PubMed  Google Scholar 

  90. Yeager, M. et al. Genome-wide association study of prostate cancer identifies a second risk locus at 8q24. Nature Genet. 39, 645–649 (2007).

    Article  CAS  PubMed  Google Scholar 

  91. Sharp, P. M., Tuohy, T. M. & Mosurski, K. R. Codon usage in yeast: cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res. 14, 5125–5143 (1986).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Bonekamp, F. & Jensen, K. F. The AGG codon is translated slowly in E. coli even at very low expression levels. Nucleic Acids Res. 16, 3013–3024 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Folley, L. S. & Yarus, M. Codon contexts from weakly expressed genes reduce expression in vivo. J. Mol. Biol. 209, 359–378 (1989).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank B. Golding for helpful discussions and G. S. Pandey for comments and help in the manuscript preparation. The assistance of E. Ayalp and I. Kimchi in generating the table in box 1 and J. Sauna for the illustrations in figure 1 is gratefully acknowledged. This work was supported by funds from the Laboratory of Hemostasis (C.K.-S.) and the Center for Biologics Evaluation and Research, US Food and Drug Administration's Modernization of Science programme (Z.E.S.). The findings and conclusions in this article have not been formally disseminated by the US Food and Drug Administration and should not be construed to represent any Agency determination or policy.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Zuben E. Sauna or Chava Kimchi-Sarfaty.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary information S1 (table)

Human diseases associated with synonymous mutations. (PDF 213 kb)

Supplementary information S2 (table)

Changes in RSCU values due to synonymous mutations associated with human diseases. (PDF 143 kb)

Related links

Related links

FURTHER INFORMATION

Zuben E. Sauna's homepage

Chava Kimchi-Sarfaty's homepage

Codon usage table

Glossary

Splicing

The transcribed precursor RNA consists of exons (which encode amino acids) and introns, which must be edited out: splicing is the process by which this occurs.

Synonymous SNPs

(sSNPs). Single-nucleotide changes that do not result in a change in the amino acid in the translated protein.

Protein therapeutics

Proteins used in the treatment of human diseases that are purified from animal or human sources or, increasingly, manufactured by recombinant DNA technology.

Non-synonymous SNPs

(nsSNPs). Single-nucleotide changes that result in a change in the amino acid in the translated protein.

Spliceosome

A complex of small nuclear RNAs and protein subunits that removes introns from the transcribed precursor mRNA.

Splicing enhancers

Short nucleotide sequences that flag the boundaries of the exon for the splicing machinery of the cell.

MicroRNAs

(miRNAs). Short RNA molecules that are post-transcriptional regulators. They bind to complementary sequences on mRNAs and result in gene repression.

Exon skipping

RNA splicing that results in one or more exons being eliminated from the final mRNA.

Relative synonymous codon usage

(RSCU). This is a simple measure of non-uniform usage of synonymous codons in a coding sequence. RSCU values show the number of times a particular codon is observed relative to the number of times that the codon would be observed if all the codons for a given amino acid had the same probability.

Proteostasis

(Protein homeostasis). This process is necessary for normal cellular function and is maintained by a highly conserved network of biological pathways.

Codon optimization

The use of the host organism's codon bias when proteins are often expressed in a foreign host.

Optical tweezers

Highly focused laser beams that can be used to hold and move microscopic objects, sometimes at the level of a single molecule.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Sauna, Z., Kimchi-Sarfaty, C. Understanding the contribution of synonymous mutations to human disease. Nat Rev Genet 12, 683–691 (2011). https://doi.org/10.1038/nrg3051

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg3051

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research