1932

Abstract

Codon usage bias, the preference for certain synonymous codons, is found in all genomes. Although synonymous mutations were previously thought to be silent, a large body of evidence has demonstrated that codon usage can play major roles in determining gene expression levels and protein structures. Codon usage influences translation elongation speed and regulates translation efficiency and accuracy. Adaptation of codon usage to tRNA expression determines the proteome landscape. In addition, codon usage biases result in nonuniform ribosome decoding rates on mRNAs, which in turn influence the cotranslational protein folding process that is critical for protein function in diverse biological processes. Conserved genome-wide correlations have also been found between codon usage and protein structures. Furthermore, codon usage is a major determinant of mRNA levels through translation-dependent effects on mRNA decay and translation-independent effects on transcriptional and posttranscriptional processes. Here, we discuss the multifaceted roles and mechanisms of codon usage in different gene regulatory processes.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-071320-112701
2021-06-20
2024-04-26
Loading full text...

Full text loading...

/deliver/fulltext/biochem/90/1/annurev-biochem-071320-112701.html?itemId=/content/journals/10.1146/annurev-biochem-071320-112701&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Ikemura T. 1985. Codon usage and tRNA content in unicellular and multicellular organisms. Mol. Biol. Evol. 2:13–34
    [Google Scholar]
  2. 2. 
    Sharp PM, Tuohy TM, Mosurski KR 1986. Codon usage in yeast: Cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res 14:5125–43
    [Google Scholar]
  3. 3. 
    Plotkin JB, Kudla G. 2011. Synonymous but not the same: the causes and consequences of codon bias. Nat. Rev. Genet. 12:32–42
    [Google Scholar]
  4. 4. 
    Hershberg R, Petrov DA. 2008. Selection on codon bias. Annu. Rev. Genet. 42:287–99
    [Google Scholar]
  5. 5. 
    Zhou M, Guo J, Cha J, Chae M, Chen S et al. 2013. Non-optimal codon usage affects expression, structure and function of clock protein FRQ. Nature 495:111–15
    [Google Scholar]
  6. 6. 
    Chaney JL, Clark PL. 2015. Roles for synonymous codon usage in protein biogenesis. Annu. Rev. Biophys. 44:143–66
    [Google Scholar]
  7. 7. 
    Hanson G, Coller J. 2018. Codon optimality, bias and usage in translation and mRNA decay. Nat. Rev. Mol. Cell Biol. 19:20–30
    [Google Scholar]
  8. 8. 
    Quax TE, Claassens NJ, Soll D, van der Oost J 2015. Codon bias as a means to fine-tune gene expression. Mol. Cell 59:149–61
    [Google Scholar]
  9. 9. 
    Duret L. 2000. tRNA gene number and codon usage in the C. elegans genome are co-adapted for optimal translation of highly expressed genes. Trends Genet 16:287–89
    [Google Scholar]
  10. 10. 
    Sabi R, Tuller T. 2014. Modelling the efficiency of codon-tRNA interactions based on codon usage bias. DNA Res 21:511–26
    [Google Scholar]
  11. 11. 
    Spencer PS, Siller E, Anderson JF, Barral JM 2012. Silent substitutions predictably alter translation elongation rates and protein folding efficiencies. J. Mol. Biol. 422:328–35
    [Google Scholar]
  12. 12. 
    Zhou M, Wang T, Fu J, Xiao G, Liu Y 2015. Nonoptimal codon usage influences protein structure in intrinsically disordered regions. Mol. Microbiol. 97:974–87
    [Google Scholar]
  13. 13. 
    Zhou Z, Dang Y, Zhou M, Li L, Yu CH et al. 2016. Codon usage is an important determinant of gene expression levels largely through its effects on transcription. PNAS 113:E6117–25
    [Google Scholar]
  14. 14. 
    Duret L, Mouchiroud D. 1999. Expression pattern and, surprisingly, gene length shape codon usage in Caenorhabditis, Drosophila, and Arabidopsis. PNAS 96:4482–87
    [Google Scholar]
  15. 15. 
    Xu Y, Ma P, Shah P, Rokas A, Liu Y, Johnson CH 2013. Non-optimal codon usage is a mechanism to achieve circadian clock conditionality. Nature 495:116–20
    [Google Scholar]
  16. 16. 
    Hense W, Anderson N, Hutter S, Stephan W, Parsch J, Carlini DB 2010. Experimentally increased codon bias in the Drosophila Adh gene leads to an increase in larval, but not adult, alcohol dehydrogenase activity. Genetics 184:547–55
    [Google Scholar]
  17. 17. 
    Lampson BL, Pershing NL, Prinz JA, Lacsina JR, Marzluff WF et al. 2013. Rare codons regulate KRas oncogenesis. Curr. Biol. 23:70–75
    [Google Scholar]
  18. 18. 
    Fu J, Dang Y, Counter C, Liu Y 2018. Codon usage regulates human KRAS expression at both transcriptional and translational levels. J. Biol. Chem. 293:17929–40
    [Google Scholar]
  19. 19. 
    Sorensen MA, Kurland CG, Pedersen S 1989. Codon usage determines translation rate in Escherichia coli. J. Mol. Biol. 207:365–77
    [Google Scholar]
  20. 20. 
    Pedersen S. 1984. Escherichia coli ribosomes translate in vivo with variable rate. EMBO J 3:2895–98
    [Google Scholar]
  21. 21. 
    Siller E, DeZwaan DC, Anderson JF, Freeman BC, Barral JM 2010. Slowing bacterial translation speed enhances eukaryotic protein folding efficiency. J. Mol. Biol. 396:1310–18
    [Google Scholar]
  22. 22. 
    Bonekamp F, Dalboge H, Christensen T, Jensen KF 1989. Translation rates of individual codons are not correlated with tRNA abundances or with frequencies of utilization in Escherichia coli. J. Bacteriol. 171:5812–16
    [Google Scholar]
  23. 23. 
    Chevance FF, Le Guyon S, Hughes KT 2014. The effects of codon context on in vivo translation speed. PLOS Genet 10:e1004392
    [Google Scholar]
  24. 24. 
    Varenne S, Buc J, Lloubes R, Lazdunski C 1984. Translation is a non-uniform process: effect of tRNA availability on the rate of elongation of nascent polypeptide chains. J. Mol. Biol. 180:549–76
    [Google Scholar]
  25. 25. 
    Ingolia NT, Ghaemmaghami S, Newman JR, Weissman JS 2009. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324:218–23
    [Google Scholar]
  26. 26. 
    Ingolia NT, Lareau LF, Weissman JS 2011. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 147:789–802
    [Google Scholar]
  27. 27. 
    Li GW, Oh E, Weissman JS 2012. The anti-Shine-Dalgarno sequence drives translational pausing and codon choice in bacteria. Nature 484:538–41
    [Google Scholar]
  28. 28. 
    Qian W, Yang JR, Pearson NM, Maclean C, Zhang J 2012. Balanced codon usage optimizes eukaryotic translational efficiency. PLOS Genet 8:e1002603
    [Google Scholar]
  29. 29. 
    Dana A, Tuller T. 2014. The effect of tRNA levels on decoding times of mRNA codons. Nucleic Acids Res 42:9171–81
    [Google Scholar]
  30. 30. 
    Artieri CG, Fraser HB. 2014. Accounting for biases in riboprofiling data indicates a major role for proline in stalling translation. Genome Res 24:2011–21
    [Google Scholar]
  31. 31. 
    Yu CH, Dang Y, Zhou Z, Wu C, Zhao F et al. 2015. Codon usage influences the local rate of translation elongation to regulate co-translational protein folding. Mol. Cell 59:744–54
    [Google Scholar]
  32. 32. 
    Zhao F, Yu CH, Liu Y 2017. Codon usage regulates protein structure and function by affecting translation elongation speed in Drosophila cells. Nucleic Acids Res 45:8484–92
    [Google Scholar]
  33. 33. 
    Yang Q, Yu CH, Zhao F, Dang Y, Wu C et al. 2019. eRF1 mediates codon usage effects on mRNA translation efficiency through premature termination at rare codons. Nucleic Acids Res 47:9243–58
    [Google Scholar]
  34. 34. 
    Zhang G, Hubalewska M, Ignatova Z 2009. Transient ribosomal attenuation coordinates protein synthesis and co-translational folding. Nat. Struct. Mol. Biol. 16:274–80
    [Google Scholar]
  35. 35. 
    Weinberg DE, Shah P, Eichhorn SW, Hussmann JA, Plotkin JB, Bartel DP 2016. Improved ribosome-footprint and mRNA measurements provide insights into dynamics and regulation of yeast translation. Cell Rep 14:1787–99
    [Google Scholar]
  36. 36. 
    Hussmann JA, Patchett S, Johnson A, Sawyer S, Press WH 2015. Understanding biases in ribosome profiling experiments reveals signatures of translation dynamics in yeast. PLOS Genet 11:e1005732
    [Google Scholar]
  37. 37. 
    Wu CC, Zinshteyn B, Wehner KA, Green R 2019. High-resolution ribosome profiling defines discrete ribosome elongation states and translational regulation during cellular stress. Mol. Cell 73:95970.e5
    [Google Scholar]
  38. 38. 
    Yan X, Hoek TA, Vale RD, Tanenbaum ME 2016. Dynamics of translation of single mRNA molecules in vivo. Cell 165:976–89
    [Google Scholar]
  39. 39. 
    Jeacock L, Faria J, Horn D 2018. Codon usage bias controls mRNA and protein abundance in trypanosomatids. eLife7:e32496
    [Google Scholar]
  40. 40. 
    Sonenberg N, Hinnebusch AG. 2009. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136:731–45
    [Google Scholar]
  41. 41. 
    Kudla G, Murray AW, Tollervey D, Plotkin JB 2009. Coding-sequence determinants of gene expression in Escherichia coli. Science 324:255–58
    [Google Scholar]
  42. 42. 
    Lu P, Vogel C, Wang R, Yao X, Marcotte EM 2007. Absolute protein expression profiling estimates the relative contributions of transcriptional and translational regulation. Nat. Biotechnol. 25:117–24
    [Google Scholar]
  43. 43. 
    Tuller T, Carmi A, Vestsigian K, Navon S, Dorfan Y et al. 2010. An evolutionarily conserved mechanism for controlling the efficiency of protein translation. Cell 141:344–54
    [Google Scholar]
  44. 44. 
    Pop C, Rouskin S, Ingolia NT, Han L, Phizicky EM et al. 2014. Causal signals between codon bias, mRNA structure, and the efficiency of translation and elongation. Mol. Syst. Biol. 10:770
    [Google Scholar]
  45. 45. 
    Dittmar KA, Goodenbour JM, Pan T 2006. Tissue-specific differences in human transfer RNA expression. PLOS Genet 2:e221
    [Google Scholar]
  46. 46. 
    Goodarzi H, Nguyen HCB, Zhang S, Dill BD, Molina H, Tavazoie SF 2016. Modulated expression of specific tRNAs drives gene expression and cancer progression. Cell 165:1416–27
    [Google Scholar]
  47. 47. 
    Chan CT, Pang YL, Deng W, Babu IR, Dyavaiah M et al. 2012. Reprogramming of tRNA modifications controls the oxidative stress response by codon-biased translation of proteins. Nat. Commun. 3:937
    [Google Scholar]
  48. 48. 
    Chionh YH, McBee M, Babu IR, Hia F, Lin W et al. 2016. tRNA-mediated codon-biased translation in mycobacterial hypoxic persistence. Nat. Commun. 7:13302
    [Google Scholar]
  49. 49. 
    Goffena J, Lefcort F, Zhang Y, Lehrmann E, Chaverra M et al. 2018. Elongator and codon bias regulate protein levels in mammalian peripheral neurons. Nat. Commun. 9:889
    [Google Scholar]
  50. 50. 
    Lyu X, Yang Q, Li L, Dang Y, Zhou Z et al. 2020. Adaptation of codon usage to tRNA I34 modification controls translation kinetics and proteome landscape. PLOS Genet 16:e1008836
    [Google Scholar]
  51. 51. 
    Letzring DP, Wolf AS, Brule CE, Grayhack EJ 2013. Translation of CGA codon repeats in yeast involves quality control components and ribosomal protein L1. RNA 19:1208–17
    [Google Scholar]
  52. 52. 
    Wada M, Ito K. 2018. Misdecoding of rare CGA codon by translation termination factors, eRF1/eRF3, suggests novel class of ribosome rescue pathway in S. cerevisiae. FEBS J 286:788–802
    [Google Scholar]
  53. 53. 
    Curran JF. 1995. Decoding with the A:I wobble pair is inefficient. Nucleic Acids Res 23:683–88
    [Google Scholar]
  54. 54. 
    Subramaniam AR, Zid BM, O'Shea EK 2014. An integrated approach reveals regulatory controls on bacterial translation elongation. Cell 159:1200–11
    [Google Scholar]
  55. 55. 
    Jin H, Bjornsson A, Isaksson LA 2002. Cis control of gene expression in E. coli by ribosome queuing at an inefficient translational stop signal. EMBO J 21:4357–67
    [Google Scholar]
  56. 56. 
    Diament A, Feldman A, Schochet E, Kupiec M, Arava Y, Tuller T 2018. The extent of ribosome queuing in budding yeast. PLOS Comput. Biol. 14:e1005951
    [Google Scholar]
  57. 57. 
    Chu D, Kazana E, Bellanger N, Singh T, Tuite MF, von der Haar T 2014. Translation elongation can control translation initiation on eukaryotic mRNAs. EMBO J 33:21–34
    [Google Scholar]
  58. 58. 
    Ferrin MA, Subramaniam AR. 2017. Kinetic modeling predicts a stimulatory role for ribosome collisions at elongation stall sites in bacteria. eLife6:e23629
    [Google Scholar]
  59. 59. 
    Ishimura R, Nagy G, Dotu I, Zhou H, Yang XL et al. 2014. RNA function. Ribosome stalling induced by mutation of a CNS-specific tRNA causes neurodegeneration. Science 345:455–59
    [Google Scholar]
  60. 60. 
    Shah P, Ding Y, Niemczyk M, Kudla G, Plotkin JB 2013. Rate-limiting steps in yeast protein translation. Cell 153:1589–601
    [Google Scholar]
  61. 61. 
    Bhattacharyya S, Jacobs WM, Adkar BV, Yan J, Zhang W, Shakhnovich EI 2018. Accessibility of the Shine-Dalgarno sequence dictates N-terminal codon bias in E. coli. Mol. Cell 70:894905.e5
    [Google Scholar]
  62. 62. 
    Yarus M, Folley LS. 1985. Sense codons are found in specific contexts. J. Mol. Biol. 182:529–40
    [Google Scholar]
  63. 63. 
    Gutman GA, Hatfield GW. 1989. Nonrandom utilization of codon pairs in Escherichia coli. PNAS 86:3699–703
    [Google Scholar]
  64. 64. 
    Alexaki A, Kames J, Holcomb DD, Athey J, Santana-Quintero LV et al. 2019. Codon and codon-pair usage tables (CoCoPUTs): facilitating genetic variation analyses and recombinant gene design. J. Mol. Biol. 431:2434–41
    [Google Scholar]
  65. 65. 
    Pouyet F, Mouchiroud D, Duret L, Semon M 2017. Recombination, meiotic expression and human codon usage. eLife6:e27344
    [Google Scholar]
  66. 66. 
    Kunec D, Osterrieder N. 2016. Codon pair bias is a direct consequence of dinucleotide bias. Cell Rep 14:55–67
    [Google Scholar]
  67. 67. 
    Irwin B, Heck JD, Hatfield GW 1995. Codon pair utilization biases influence translational elongation step times. J. Biol. Chem. 270:22801–6
    [Google Scholar]
  68. 68. 
    Buchan JR, Aucott LS, Stansfield I 2006. tRNA properties help shape codon pair preferences in open reading frames. Nucleic Acids Res 34:1015–27
    [Google Scholar]
  69. 69. 
    Gamble CE, Brule CE, Dean KM, Fields S, Grayhack EJ 2016. Adjacent codons act in concert to modulate translation efficiency in yeast. Cell 166:679–90
    [Google Scholar]
  70. 70. 
    Curran JF, Poole ES, Tate WP, Gross BL 1995. Selection of aminoacyl-tRNAs at sense codons: The size of the tRNA variable loop determines whether the immediate 3′ nucleotide to the codon has a context effect. Nucleic Acids Res 23:4104–8
    [Google Scholar]
  71. 71. 
    Chin JX, Chung BK, Lee DY 2014. Codon Optimization OnLine (COOL): a web-based multi-objective optimization platform for synthetic gene design. Bioinformatics 30:2210–12
    [Google Scholar]
  72. 72. 
    Coleman JR, Papamichail D, Skiena S, Futcher B, Wimmer E, Mueller S 2008. Virus attenuation by genome-scale changes in codon pair bias. Science 320:1784–87
    [Google Scholar]
  73. 73. 
    Li P, Ke X, Wang T, Tan Z, Luo D et al. 2018. Zika virus attenuation by codon pair deoptimization induces sterilizing immunity in mouse models. J. Virol.92:e00701–18
    [Google Scholar]
  74. 74. 
    Atkinson NJ, Witteveldt J, Evans DJ, Simmonds P 2014. The influence of CpG and UpA dinucleotide frequencies on RNA virus replication and characterization of the innate cellular pathways underlying virus attenuation and enhanced replication. Nucleic Acids Res 42:4527–45
    [Google Scholar]
  75. 75. 
    Akashi H. 1994. Synonymous codon usage in Drosophila melanogaster: natural selection and translational accuracy. Genetics 136:927–35
    [Google Scholar]
  76. 76. 
    Rodnina MV, Fischer N, Maracci C, Stark H 2017. Ribosome dynamics during decoding. Philos. Trans. R. Soc. B372:20160182
    [Google Scholar]
  77. 77. 
    Precup J, Parker J. 1987. Missense misreading of asparagine codons as a function of codon identity and context. J. Biol. Chem. 262:11351–55
    [Google Scholar]
  78. 78. 
    Stansfield I, Jones KM, Herbert P, Lewendon A, Shaw WV, Tuite MF 1998. Missense translation errors in Saccharomyces cerevisiae. J. Mol. Biol. 282:13–24
    [Google Scholar]
  79. 79. 
    Ribas de Pouplana L, Santos MA, Zhu JH, Farabaugh PJ, Javid B 2014. Protein mistranslation: friend or foe?. Trends Biochem. Sci. 39:355–62
    [Google Scholar]
  80. 80. 
    Stoletzki N, Eyre-Walker A. 2007. Synonymous codon usage in Escherichia coli: selection for translational accuracy. Mol. Biol. Evol. 24:374–81
    [Google Scholar]
  81. 81. 
    Gilchrist MA, Shah P, Zaretzki R 2009. Measuring and detecting molecular adaptation in codon usage against nonsense errors during protein translation. Genetics 183:1493–505
    [Google Scholar]
  82. 82. 
    Drummond DA, Wilke CO. 2008. Mistranslation-induced protein misfolding as a dominant constraint on coding-sequence evolution. Cell 134:341–52
    [Google Scholar]
  83. 83. 
    Kramer EB, Farabaugh PJ. 2007. The frequency of translational misreading errors in E. coli is largely determined by tRNA competition. RNA 13:87–96
    [Google Scholar]
  84. 84. 
    Punde N, Kooken J, Leary D, Legler PM, Angov E 2019. Codon harmonization reduces amino acid misincorporation in bacterially expressed P. falciparum proteins and improves their immunogenicity. AMB Express 9:167
    [Google Scholar]
  85. 85. 
    Wohlgemuth I, Pohl C, Rodnina MV 2010. Optimization of speed and accuracy of decoding in translation. EMBO J 29:3701–9
    [Google Scholar]
  86. 86. 
    Johansson M, Zhang J, Ehrenberg M 2012. Genetic code translation displays a linear trade-off between efficiency and accuracy of tRNA selection. PNAS 109:131–36
    [Google Scholar]
  87. 87. 
    Mordret E, Dahan O, Asraf O, Rak R, Yehonadav A et al. 2019. Systematic detection of amino acid substitutions in proteomes reveals mechanistic basis of ribosome errors and selection for translation fidelity. Mol. Cell 75:42741.e5
    [Google Scholar]
  88. 88. 
    Huang Y, Koonin EV, Lipman DJ, Przytycka TM 2009. Selection for minimization of translational frameshifting errors as a factor in the evolution of codon usage. Nucleic Acids Res 37:6799–810
    [Google Scholar]
  89. 89. 
    Korniy N, Goyal A, Hoffmann M, Samatova E, Peske F et al. 2019. Modulation of HIV-1 Gag/Gag-Pol frameshifting by tRNA abundance. Nucleic Acids Res 47:5210–22
    [Google Scholar]
  90. 90. 
    Caliskan N, Peske F, Rodnina MV 2015. Changed in translation: mRNA recoding by −1 programmed ribosomal frameshifting. Trends Biochem. Sci. 40:265–74
    [Google Scholar]
  91. 91. 
    Koutmou KS, Schuller AP, Brunelle JL, Radhakrishnan A, Djuranovic S, Green R 2015. Ribosomes slide on lysine-encoding homopolymeric A stretches. eLife 4:e05534
    [Google Scholar]
  92. 92. 
    Ikemura T. 1981. 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
    [Google Scholar]
  93. 93. 
    Moriyama EN, Powell JR. 1997. Codon usage bias and tRNA abundance in Drosophila. J. Mol. Evol. 45:514–23
    [Google Scholar]
  94. 94. 
    dos Reis M, Savva R, Wernisch L 2004. Solving the riddle of codon usage preferences: a test for translational selection. Nucleic Acids Res 32:5036–44
    [Google Scholar]
  95. 95. 
    Pechmann S, Frydman J. 2013. Evolutionary conservation of codon optimality reveals hidden signatures of cotranslational folding. Nat. Struct. Mol. Biol. 20:237–43
    [Google Scholar]
  96. 96. 
    Bulmer M. 1987. Coevolution of codon usage and transfer RNA abundance. Nature 325:728–30
    [Google Scholar]
  97. 97. 
    Kirchner S, Cai Z, Rauscher R, Kastelic N, Anding M et al. 2017. Alteration of protein function by a silent polymorphism linked to tRNA abundance. PLOS Biol 15:e2000779
    [Google Scholar]
  98. 98. 
    Letzring DP, Dean KM, Grayhack EJ 2010. Control of translation efficiency in yeast by codon-anticodon interactions. RNA 16:2516–28
    [Google Scholar]
  99. 99. 
    Frumkin I, Lajoie MJ, Gregg CJ, Hornung G, Church GM, Pilpel Y 2018. Codon usage of highly expressed genes affects proteome-wide translation efficiency. PNAS 115:E4940–49
    [Google Scholar]
  100. 100. 
    Dong H, Nilsson L, Kurland CG 1996. Co-variation of tRNA abundance and codon usage in Escherichia coli at different growth rates. J. Mol. Biol. 260:649–63
    [Google Scholar]
  101. 101. 
    Berg OG, Kurland CG. 1997. Growth rate-optimised tRNA abundance and codon usage. J. Mol. Biol. 270:544–50
    [Google Scholar]
  102. 102. 
    Avcilar-Kucukgoze I, Bartholomaus A, Cordero Varela JA, Kaml RF, Neubauer P et al. 2016. Discharging tRNAs: a tug of war between translation and detoxification in Escherichia coli. Nucleic Acids Res 44:8324–34
    [Google Scholar]
  103. 103. 
    Wohlgemuth SE, Gorochowski TE, Roubos JA 2013. Translational sensitivity of the Escherichia coli genome to fluctuating tRNA availability. Nucleic Acids Res 41:8021–33
    [Google Scholar]
  104. 104. 
    Pang YL, Abo R, Levine SS, Dedon PC 2014. Diverse cell stresses induce unique patterns of tRNA up- and down-regulation: tRNA-seq for quantifying changes in tRNA copy number. Nucleic Acids Res 42:e170
    [Google Scholar]
  105. 105. 
    Torrent M, Chalancon G, de Groot NS, Wuster A, Madan Babu M 2018. Cells alter their tRNA abundance to selectively regulate protein synthesis during stress conditions. Sci. Signal.11:eaat6409
    [Google Scholar]
  106. 106. 
    Sagi D, Rak R, Gingold H, Adir I, Maayan G et al. 2016. Tissue- and time-specific expression of otherwise identical tRNA genes. PLOS Genet 12:e1006264
    [Google Scholar]
  107. 107. 
    Gingold H, Tehler D, Christoffersen NR, Nielsen MM, Asmar F et al. 2014. A dual program for translation regulation in cellular proliferation and differentiation. Cell 158:1281–92
    [Google Scholar]
  108. 108. 
    Hernandez-Alias X, Benisty H, Schaefer MH, Serrano L 2020. Translational efficiency across healthy and tumor tissues is proliferation-related. Mol. Syst. Biol. 16:e9275
    [Google Scholar]
  109. 109. 
    Rudolph KL, Schmitt BM, Villar D, White RJ, Marioni JC et al. 2016. Codon-driven translational efficiency is stable across diverse mammalian cell states. PLOS Genet 12:e1006024
    [Google Scholar]
  110. 110. 
    Saikia M, Wang X, Mao Y, Wan J, Pan T, Qian SB 2016. Codon optimality controls differential mRNA translation during amino acid starvation. RNA 22:1719–27
    [Google Scholar]
  111. 111. 
    Zaborske JM, DuMont VL, Wallace EW, Pan T, Aquadro CF, Drummond DA 2014. A nutrient-driven tRNA modification alters translational fidelity and genome-wide protein coding across an animal genus. PLOS Biol 12:e1002015
    [Google Scholar]
  112. 112. 
    Deng W, Babu IR, Su D, Yin S, Begley TJ, Dedon PC 2015. Trm9-catalyzed tRNA modifications regulate global protein expression by codon-biased translation. PLOS Genet. 11:e1005706
    [Google Scholar]
  113. 113. 
    Chan C, Pham P, Dedon PC, Begley TJ 2018. Lifestyle modifications: coordinating the tRNA epitranscriptome with codon bias to adapt translation during stress responses. Genome Biol 19:228
    [Google Scholar]
  114. 114. 
    Gomez-Roman N, Grandori C, Eisenman RN, White RJ 2003. Direct activation of RNA polymerase III transcription by c-Myc. Nature 421:290–94
    [Google Scholar]
  115. 115. 
    Yang J, Smith DK, Ni H, Wu K, Huang D et al. 2020. SOX4-mediated repression of specific tRNAs inhibits proliferation of human glioblastoma cells. PNAS 117:5782–90
    [Google Scholar]
  116. 116. 
    Dittmar KA, Sorensen MA, Elf J, Ehrenberg M, Pan T 2005. Selective charging of tRNA isoacceptors induced by amino-acid starvation. EMBO Rep 6:151–57
    [Google Scholar]
  117. 117. 
    Elf J, Nilsson D, Tenson T, Ehrenberg M 2003. Selective charging of tRNA isoacceptors explains patterns of codon usage. Science 300:1718–22
    [Google Scholar]
  118. 118. 
    Subramaniam AR, Pan T, Cluzel P 2013. Environmental perturbations lift the degeneracy of the genetic code to regulate protein levels in bacteria. PNAS 110:2419–24
    [Google Scholar]
  119. 119. 
    Darnell AM, Subramaniam AR, O'Shea EK 2018. Translational control through differential ribosome pausing during amino acid limitation in mammalian cells. Mol. Cell 71:22943.e11
    [Google Scholar]
  120. 120. 
    Kothe U, Rodnina MV. 2007. Codon reading by tRNAAla with modified uridine in the wobble position. Mol. Cell 25:167–74
    [Google Scholar]
  121. 121. 
    Ran W, Higgs PG. 2010. The influence of anticodon-codon interactions and modified bases on codon usage bias in bacteria. Mol. Biol. Evol. 27:2129–40
    [Google Scholar]
  122. 122. 
    Novoa EM, Pavon-Eternod M, Pan T, Ribas de Pouplana L 2012. A role for tRNA modifications in genome structure and codon usage. Cell 149:202–13
    [Google Scholar]
  123. 123. 
    Bornelov S, Selmi T, Flad S, Dietmann S, Frye M 2019. Codon usage optimization in pluripotent embryonic stem cells. Genome Biol 20:119
    [Google Scholar]
  124. 124. 
    Nedialkova DD, Leidel SA. 2015. Optimization of codon translation rates via tRNA modifications maintains proteome integrity. Cell 161:1606–18
    [Google Scholar]
  125. 125. 
    Zinshteyn B, Gilbert WV. 2013. Loss of a conserved tRNA anticodon modification perturbs cellular signaling. PLOS Genet 9:e1003675
    [Google Scholar]
  126. 126. 
    Thommen M, Holtkamp W, Rodnina MV 2017. Co-translational protein folding: progress and methods. Curr. Opin. Struct. Biol. 42:83–89
    [Google Scholar]
  127. 127. 
    Komar AA. 2018. Unraveling co-translational protein folding: concepts and methods. Methods 137:71–81
    [Google Scholar]
  128. 128. 
    Holtkamp W, Kokic G, Jager M, Mittelstaet J, Komar AA, Rodnina MV 2015. Cotranslational protein folding on the ribosome monitored in real time. Science 350:1104–7
    [Google Scholar]
  129. 129. 
    Shiber A, Doring K, Friedrich U, Klann K, Merker D et al. 2018. Cotranslational assembly of protein complexes in eukaryotes revealed by ribosome profiling. Nature 561:268–72
    [Google Scholar]
  130. 130. 
    Purvis IJ, Bettany AJ, Santiago TC, Coggins JR, Duncan K et al. 1987. The efficiency of folding of some proteins is increased by controlled rates of translation in vivo. A hypothesis. J. Mol. Biol. 193:413–17
    [Google Scholar]
  131. 131. 
    O'Brien EP, Vendruscolo M, Dobson CM 2012. Prediction of variable translation rate effects on cotranslational protein folding. Nat. Commun. 3:868
    [Google Scholar]
  132. 132. 
    Sharma AK, O'Brien EP. 2018. Non-equilibrium coupling of protein structure and function to translation-elongation kinetics. Curr. Opin. Struct. Biol. 49:94–103
    [Google Scholar]
  133. 133. 
    Komar AA, Lesnik T, Reiss C 1999. Synonymous codon substitutions affect ribosome traffic and protein folding during in vitro translation. FEBS Lett 462:387–91
    [Google Scholar]
  134. 134. 
    Cortazzo P, Cervenansky C, Marin M, Reiss C, Ehrlich R, Deana A 2002. Silent mutations affect in vivo protein folding in Escherichia coli. Biochem. Biophys. Res. Commun. 293:537–41
    [Google Scholar]
  135. 135. 
    Sander IM, Chaney JL, Clark PL 2014. Expanding Anfinsen's principle: contributions of synonymous codon selection to rational protein design. J. Am. Chem. Soc. 136:858–61
    [Google Scholar]
  136. 136. 
    Buhr F, Jha S, Thommen M, Mittelstaet J, Kutz F et al. 2016. Synonymous codons direct cotranslational folding toward different protein conformations. Mol. Cell 61:341–51
    [Google Scholar]
  137. 137. 
    Walsh IM, Bowman MA, Soto Santarriaga IF, Rodriguez A, Clark PL 2020. Synonymous codon substitutions perturb cotranslational protein folding in vivo and impair cell fitness. PNAS 117:3528–34
    [Google Scholar]
  138. 138. 
    Kimchi-Sarfaty C, Oh JM, Kim IW, Sauna ZE, Calcagno AM et al. 2007. A “silent” polymorphism in the MDR1 gene changes substrate specificity. Science 315:525–28
    [Google Scholar]
  139. 139. 
    Fu J, Murphy KA, Zhou M, Li YH, Lam VH et al. 2016. Codon usage affects the structure and function of the Drosophila circadian clock protein PERIOD. Genes Dev 30:1761–75
    [Google Scholar]
  140. 140. 
    Kim SJ, Yoon JS, Shishido H, Yang Z, Rooney LA et al. 2015. Translational tuning optimizes nascent protein folding in cells. Science 348:444–48
    [Google Scholar]
  141. 141. 
    Wan Makhtar WR, Browne G, Karountzos A, Stevens C, Alghamdi Y et al. 2017. Short stretches of rare codons regulate translation of the transcription factor ZEB2 in cancer cells. Oncogene 36:6640–48
    [Google Scholar]
  142. 142. 
    Alexaki A, Hettiarachchi GK, Athey JC, Katneni UK, Simhadri V et al. 2019. Effects of codon optimization on coagulation factor IX translation and structure: implications for protein and gene therapies. Sci. Rep. 9:15449
    [Google Scholar]
  143. 143. 
    Hunt R, Hettiarachchi G, Katneni U, Hernandez N, Holcomb D et al. 2019. A single synonymous variant (c.354G>A [p.P118P]) in ADAMTS13 confers enhanced specific activity. Int. J. Mol. Sci.20:5734
    [Google Scholar]
  144. 144. 
    Thanaraj TA, Argos P. 1996. Protein secondary structural types are differentially coded on messenger RNA. Protein Sci 5:1973–83
    [Google Scholar]
  145. 145. 
    Oresic M, Shalloway D. 1998. Specific correlations between relative synonymous codon usage and protein secondary structure. J. Mol. Biol. 281:31–48
    [Google Scholar]
  146. 146. 
    Gu W, Zhou T, Ma J, Sun X, Lu Z 2004. The relationship between synonymous codon usage and protein structure in Escherichia coli and Homo sapiens. Biosystems 73:89–97
    [Google Scholar]
  147. 147. 
    Saunders R, Deane CM. 2010. Synonymous codon usage influences the local protein structure observed. Nucleic Acids Res 38:6719–28
    [Google Scholar]
  148. 148. 
    Chaney JL, Steele A, Carmichael R, Rodriguez A, Specht AT et al. 2017. Widespread position-specific conservation of synonymous rare codons within coding sequences. PLOS Comput. Biol. 13:e1005531
    [Google Scholar]
  149. 149. 
    Zhang G, Ignatova Z. 2009. Generic algorithm to predict the speed of translational elongation: implications for protein biogenesis. PLOS ONE 4:e5036
    [Google Scholar]
  150. 150. 
    Zhou T, Weems M, Wilke CO 2009. Translationally optimal codons associate with structurally sensitive sites in proteins. Mol. Biol. Evol. 26:1571–80
    [Google Scholar]
  151. 151. 
    Jacobs WM, Shakhnovich EI. 2017. Evidence of evolutionary selection for cotranslational folding. PNAS 114:11434–39
    [Google Scholar]
  152. 152. 
    Zhang F, Saha S, Shabalina SA, Kashina A 2010. Differential arginylation of actin isoforms is regulated by coding sequence-dependent degradation. Science 329:1534–37
    [Google Scholar]
  153. 153. 
    Clarke TF 4th, Clark PL 2010. Increased incidence of rare codon clusters at 5′ and 3′ gene termini: implications for function. BMC Genom 11:118
    [Google Scholar]
  154. 154. 
    Pechmann S, Chartron JW, Frydman J 2014. Local slowdown of translation by nonoptimal codons promotes nascent-chain recognition by SRP in vivo. Nat. Struct. Mol. Biol. 21:1100–5
    [Google Scholar]
  155. 155. 
    Sanguinetti M, Iriarte A, Amillis S, Marin M, Musto H, Ramon A 2019. A pair of non-optimal codons are necessary for the correct biosynthesis of the Aspergillus nidulans urea transporter, UreA. R. Soc. Open Sci. 6:190773
    [Google Scholar]
  156. 156. 
    Hoekema A, Kastelein RA, Vasser M, de Boer HA 1987. Codon replacement in the PGK1 gene of Saccharomyces cerevisiae: experimental approach to study the role of biased codon usage in gene expression. Mol. Cell. Biol. 7:2914–24
    [Google Scholar]
  157. 157. 
    Koda A, Bogaki T, Minetoki T, Hirotsune M 2005. High expression of a synthetic gene encoding potato α-glucan phosphorylase in Aspergillus niger. J. Biosci. Bioeng. 100:531–37
    [Google Scholar]
  158. 158. 
    Coghlan A, Wolfe KH. 2000. Relationship of codon bias to mRNA concentration and protein length in Saccharomyces cerevisiae. Yeast 16:1131–45
    [Google Scholar]
  159. 159. 
    dos Reis M, Wernisch L, Savva R 2003. Unexpected correlations between gene expression and codon usage bias from microarray data for the whole Escherichia coli K-12 genome. Nucleic Acids Res 31:6976–85
    [Google Scholar]
  160. 160. 
    Stimac E, Groppi VE, Coffino P 1983. Increased histone mRNA levels during inhibition of protein synthesis. Biochem. Biophys. Res. Commun. 114:131–37
    [Google Scholar]
  161. 161. 
    Roy B, Jacobson A. 2013. The intimate relationships of mRNA decay and translation. Trends Genet 29:691–99
    [Google Scholar]
  162. 162. 
    Presnyak V, Alhusaini N, Chen YH, Martin S, Morris N et al. 2015. Codon optimality is a major determinant of mRNA stability. Cell 160:1111–24
    [Google Scholar]
  163. 163. 
    Boel G, Letso R, Neely H, Price WN, Wong KH et al. 2016. Codon influence on protein expression in E. coli correlates with mRNA levels. Nature 529:358–63
    [Google Scholar]
  164. 164. 
    Bazzini AA, Del Viso F, Moreno-Mateos MA, Johnstone TG, Vejnar CE et al. 2016. Codon identity regulates mRNA stability and translation efficiency during the maternal-to-zygotic transition. EMBO J. 35:2087–103
    [Google Scholar]
  165. 165. 
    Mishima Y, Tomari Y. 2016. Codon usage and 3′ UTR length determine maternal mRNA stability in zebrafish. Mol. Cell 61:874–85
    [Google Scholar]
  166. 166. 
    Narula A, Ellis J, Taliaferro JM, Rissland OS 2019. Coding regions affect mRNA stability in human cells. RNA 25:1751–64
    [Google Scholar]
  167. 167. 
    Hia F, Yang SF, Shichino Y, Yoshinaga M, Murakawa Y et al. 2019. Codon bias confers stability to human mRNAs. EMBO Rep 20:e48220
    [Google Scholar]
  168. 168. 
    Wu Q, Medina SG, Kushawah G, DeVore ML, Castellano LA et al. 2019. Translation affects mRNA stability in a codon-dependent manner in human cells. eLife8:e45396
    [Google Scholar]
  169. 169. 
    Burow DA, Martin S, Quail JF, Alhusaini N, Coller J, Cleary MD 2018. Attenuated codon optimality contributes to neural-specific mRNA decay in Drosophila. Cell Rep 24:1704–12
    [Google Scholar]
  170. 170. 
    Carlini DB. 2005. Context-dependent codon bias and messenger RNA longevity in the yeast transcriptome. Mol. Biol. Evol. 22:1403–11
    [Google Scholar]
  171. 171. 
    Radhakrishnan A, Chen Y-H, Martin S, Alhusaini N, Green R, Coller J 2016. The DEAD-box protein Dhh1p couples mRNA decay and translation by monitoring codon optimality. Cell 167:122–32e9
    [Google Scholar]
  172. 172. 
    Webster MW, Chen YH, Stowell JAW, Alhusaini N, Sweet T et al. 2018. mRNA deadenylation is coupled to translation rates by the differential activities of Ccr4-Not nucleases. Mol. Cell 70:1089100.e8
    [Google Scholar]
  173. 173. 
    Buschauer R, Matsuo Y, Sugiyama T, Chen YH, Alhusaini N et al. 2020. The Ccr4-Not complex monitors the translating ribosome for codon optimality. Science368:eaay6912
    [Google Scholar]
  174. 174. 
    Kudla G, Lipinski L, Caffin F, Helwak A, Zylicz M 2006. High guanine and cytosine content increases mRNA levels in mammalian cells. PLOS Biol 4:e180
    [Google Scholar]
  175. 175. 
    Newman ZR, Young JM, Ingolia NT, Barton GM 2016. Differences in codon bias and GC content contribute to the balanced expression of TLR7 and TLR9. PNAS 113:E1362–71
    [Google Scholar]
  176. 176. 
    Harigaya Y, Parker R. 2016. Analysis of the association between codon optimality and mRNA stability in Schizosaccharomyces pombe. BMC Genom 17:895
    [Google Scholar]
  177. 177. 
    Tillo D, Hughes TR. 2009. G+C content dominates intrinsic nucleosome occupancy. BMC Bioinform. 10:442
    [Google Scholar]
  178. 178. 
    Chamary JV, Parmley JL, Hurst LD 2006. Hearing silence: non-neutral evolution at synonymous sites in mammals. Nat. Rev. Genet. 7:98–108
    [Google Scholar]
  179. 179. 
    Cartegni L, Chew SL, Krainer AR 2002. Listening to silence and understanding nonsense: exonic mutations that affect splicing. Nat. Rev. Genet. 3:285–98
    [Google Scholar]
  180. 180. 
    Zhou Z, Dang Y, Zhou M, Yuan H, Liu Y 2018. Codon usage biases co-evolve with transcription termination machinery to suppress premature cleavage and polyadenylation. eLife7:e33569
    [Google Scholar]
  181. 181. 
    Mordstein C, Savisaar R, Young RS, Bazile J, Talmane L et al. 2020. Codon usage and splicing jointly influence mRNA localization. Cell Syst 10:351–62.e8
    [Google Scholar]
  182. 182. 
    Courel M, Clement Y, Bossevain C, Foretek D, Vidal Cruchez O et al. 2019. GC content shapes mRNA storage and decay in human cells. eLife8:e49708
    [Google Scholar]
  183. 183. 
    Mittal P, Brindle J, Stephen J, Plotkin JB, Kudla G 2018. Codon usage influences fitness through RNA toxicity. PNAS 115:8639–44
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-071320-112701
Loading
/content/journals/10.1146/annurev-biochem-071320-112701
Loading

Data & Media loading...

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