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.

Rates of evolutionary change in viruses: patterns and determinants

Key Points

  • Viral mutation rates vary over five orders of magnitude, whereas viral substitution rates vary over six orders of magnitude.

  • Instead of simplifying the differences by stating that RNA viruses mutate faster than DNA viruses owing to differences in polymerase fidelity, it seems more likely that small viruses mutate faster than large viruses irrespective of the nucleic acid of their genome.

  • Single-stranded DNA (ssDNA) and double-stranded DNA (dsDNA) viruses have non-overlapping ranges of mutation and substitution rates, with ssDNA viruses behaving more like RNA viruses. However, there are currently no good estimates of substitution rate for dsRNA viruses.

  • There are several sources of mutation in addition to polymerase error. These include host antiviral enzymes, spontaneous chemical reactions and environmental mutagens such as ultraviolet irradiation.

  • Various processes shape the evolution of mutation rates in viruses, although more research is needed to determine their precise contribution, and whether and how natural selection has acted to optimize these rates.

  • Coalescent methods that use serially sampled data represent a powerful way to estimate substitution rates from rapidly evolving RNA and ssDNA viruses.

Abstract

Understanding the factors that determine the rate at which genomes generate and fix mutations provides important insights into key evolutionary mechanisms. We review our current knowledge of the rates of mutation and substitution, as well as their determinants, in RNA viruses, DNA viruses and retroviruses. We show that the high rate of nucleotide substitution in RNA viruses is matched by some DNA viruses, suggesting that evolutionary rates in viruses are explained by diverse aspects of viral biology, such as genomic architecture and replication speed, and not simply by polymerase fidelity.

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: Average rates of spontaneous mutation in viruses, adjusted to the rate per genome replication.
Figure 2: Comparison between viral mutation and substitution rates.
Figure 3: Factors influencing mutation and substitution rates in viruses.

Similar content being viewed by others

References

  1. Baer, C. F., Miyamoto, M. M. & Denver, D. R. Mutation rate variation in multicellular eukaryotes: causes and consequences. Nature Rev. Genet. 8, 619–631 (2007).

    Article  CAS  PubMed  Google Scholar 

  2. Drake, J. W., Charlesworth, B., Charlesworth, D. & Crow, J. F. Rates of spontaneous mutation. Genetics. 148, 1667–1686 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Li, W.-H. Molecular Evolution (Sinauer Associates, Inc., Sunderland, Massachusetts, 1997).

    Google Scholar 

  4. Posada, D., Crandall, K. A. & Holmes, E. C. Recombination in evolutionary genomics. Annu. Rev. Genet. 36, 75–97 (2002).

    Article  CAS  PubMed  Google Scholar 

  5. Philippe, H. & Douady, C. J. Horizontal gene transfer and phylogenetics. Curr. Opin. Microbiol. 6, 498–505 (2003).

    Article  CAS  PubMed  Google Scholar 

  6. Clark, A. G. et al. Evolution of genes and genomes on the Drosophila phylogeny. Nature 450, 203–218 (2007).

    Article  PubMed  CAS  Google Scholar 

  7. Drummond, A. J., Pybus, O. G., Rambaut, A., Forsberg, R. & Rodrigo, A. G. Measurably evolving populations. Trends Ecol. Evol. 18, 481–488 (2003).

    Article  Google Scholar 

  8. Gibbs, A. J., Calisher, C. H. & García-Arenal, F. Molecular Basis of Virus Evolution (Cambridge University Press, New York, 1995).

    Book  Google Scholar 

  9. Domingo, E. & Holland, J. J. RNA virus mutations and fitness for survival. Annu. Rev. Microbiol. 51, 151–178 (1997).

    Article  CAS  PubMed  Google Scholar 

  10. Drake, J. W. Rates of spontaneous mutation among RNA viruses. Proc. Natl Acad. Sci. USA 90, 4171–4175 (1993).This paper was the first comprehensive survey of rates of mutation in RNA viruses.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Drake, J. W. & Hwang, C. B. C. On the mutation rate of herpes simplex virus type 1. Genetics. 170, 969–970 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Raney, J. L., Delongchamp, R. R. & Valentine, C. R. Spontaneous mutant frequency and mutation spectrum for gene A of φX174 grown in E. coli. Environ. Mol. Mutag. 44, 119–127 (2004).

    Article  CAS  Google Scholar 

  13. Flint, S. J., Enquist, L. W., Racaniello, V. R. & Skalka, A. M. Principles of Virology Molecular Biology, Pathogenesis and Control of Animal Viruses (ASM, Washington, 2004).

    Google Scholar 

  14. Garcia-Diaz, M. & Bebenek, K. Multiple functions of DNA polymerases. Crit. Rev. Plant Sci. 26, 105–122 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Jung, A. et al. Multiply infected spleen cells in HIV patients. Nature 418, 144 (2002).

    Article  CAS  PubMed  Google Scholar 

  16. Bonhoeffer, S., Holmes, E. C. & Nowak, M. A. Causes of HIV diversity. Nature 376, 125 (1995).

    Article  CAS  PubMed  Google Scholar 

  17. Walsh, C. P. & Xu, G. L. Cytosine methylation and DNA repair. Curr. Topics Microbiol. Immunol. 301, 283–315 (2006).

    CAS  Google Scholar 

  18. Mangeat, B. et al. Broad antiretroviral defence by human APOBEC3G through lethal editing of nascent reverse transcripts. Nature 424, 99–103 (2003).

    Article  CAS  PubMed  Google Scholar 

  19. Caride, E. et al. Sexual transmission of HIV-1 isolate showing G–A hypermutation. J. Clin. Virol. 23, 179–189 (2002).

    Article  CAS  PubMed  Google Scholar 

  20. Frederico, L. A., Kunkel, T. A. & Shaw, B. R. A sensitive genetic assay for the detection of cytosine deamination — determination of rate constants and the activation energy. Biochemistry. 29, 2532–2537 (1990).

    Article  CAS  PubMed  Google Scholar 

  21. Xia, X. & Yuen, K. Y. Differential selection and mutation between dsDNA and ssDNA phages shape the evolution of their genomic at percentage. BMC Genet. 6, 20 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Pita, J. S., De Miranda, J. R., Schneider, W. L. & Roossinck, M. J. Environment determines fidelity for an RNA virus replicase. J. Virol. 81, 9072–9077 (2007). This paper describes an experimental test of in vivo rates of mutation caused by insertions and deletions. The study comprises a powerful comparison of RdRp fidelity of viruses that infect pepper and tobacco plants.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Schrag, S. J., Rota, P. A. & Bellini, W. J. Spontaneous mutation rate of measles virus: direct estimation based on mutations conferring monoclonal antibody resistance. J. Virol. 73, 51–54 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Bull, J. J., Sanjuan, R. & Wilke, C. O. Theory of lethal mutagenesis for viruses. J. Virol. 81, 2930–2939 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Drake, J. W., Bebenek, A., Kissling, G. E. & Peddada, S. Clusters of mutations from transient hypermutability. Proc. Natl Acad. Sci. USA 102, 12849–12854 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Drake, J. W. Mutations in clusters and showers. Proc. Natl Acad. Sci. USA 104, 8203–8204 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Drake, J. W. Too many mutants with multiple mutations. Crit. Rev. Biochem. Mol. Biol. 42, 247–258 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pathak, V. K. & Temin, H. M. 5-azacytidine and RNA secondary structure increase the retrovirus mutation rate. J. Virol. 66, 3093–3100 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Hudson, R. E., Bergthorsson, U., Roth, J. R. & Ochman, H. Effect of chromosome location on bacterial mutation rates. Mol. Biol. Evol. 19, 85–92 (2002).

    Article  CAS  PubMed  Google Scholar 

  30. Chao, L., Rang, C. U. & Wong, L. E. Distribution of spontaneous mutants and inferences about the replication mode of the RNA bacteriophage φ6. J. Virol. 76, 3276–3281 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. French, R. & Stenger, D. C. Evolution of wheat streak mosaic virus: dynamics of population growth within plants may explain limited variation. Annu. Rev. Phytopathol. 41, 199–214 (2003).

    Article  CAS  PubMed  Google Scholar 

  32. Jenkins, G. M., Rambaut, A., Pybus, O. G. & Holmes, E. C. Rates of molecular evolution in RNA viruses: a quantitative phylogenetic analysis. J. Mol. Evol. 54, 156–165 (2002). This paper describes a large-scale comparative analysis of the rates of nucleotide substitution in RNA viruses, with the caveat that this study pre-dates the use of more sophisticated analytical techniques.

    Article  CAS  PubMed  Google Scholar 

  33. Hanada, K., Suzuki, Y. & Gojobori, T. A large variation in the rates of synonymous substitution for RNA viruses and its relationship to a diversity of viral infection and transmission modes. Mol. Biol. Evol. 21, 1074–1080 (2004).

    Article  CAS  PubMed  Google Scholar 

  34. Chare, E. R. & Holmes, E. C. Selection pressures in the capsid genes of plant RNA viruses reflect mode of transmission. J. Gen. Virol. 85, 3149–3157 (2004).

    Article  CAS  PubMed  Google Scholar 

  35. Holmes, E. C. Error thresholds and the constraints to RNA virus evolution. Trends Microbiol. 11, 543–546 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Woelk, C. H. & Holmes, E. C. Reduced positive selection in vector-borne RNA viruses. Mol. Biol. Evol. 19, 2333–2336 (2002).

    Article  CAS  PubMed  Google Scholar 

  37. Weaver, S. C. & Barrett, A. D. Transmission cycles, host range, evolution and emergence of arboviral disease. Nature Rev. Microbiol. 2, 789–801 (2004).

    Article  CAS  Google Scholar 

  38. Berry, I. M. et al. Unequal evolutionary rates in the human immunodeficiency virus type 1 (HIV-1) pandemic: the evolutionary rate of HIV-1 slows down when the epidemic rate increases. J. Virol. 81, 10625–10635 (2007).

    Article  CAS  PubMed Central  Google Scholar 

  39. Nielsen, R. & Yang, Z. Likelihood models for detecting positively selected amino acid sites and applications to the HIV-1 envelope gene. Genetics. 148, 929–936 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Li, B. et al. Rapid reversion of sequence polymorphisms dominates early human immunodeficiency virus type 1 evolution. J. Virol. 81, 193–201 (2007).

    Article  CAS  PubMed  Google Scholar 

  41. Switzer, W. M. et al. Ancient co-speciation of simian foamy viruses and primates. Nature 434, 376–380 (2005).This study revealed that SFVs co-diverged with their primate hosts over many millions of years, leading to the inference that SFV evolves more slowly than any other RNA virus.

    Article  CAS  PubMed  Google Scholar 

  42. Salemi, M. et al. Different population dynamics of human T-cell lymphotropic virus type II in intravenous drug users compared with endemically infected tribes. Proc. Natl Acad. Sci. USA 96, 13253–13258 (1999). This paper is an important example of how different transmission routes (for example, needle-sharing versus mother-to child transmission) can markedly affect the rate of nucleotide substitution.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Vandamme, A. M., Bertazzoni, U. & Salemi, M. Evolutionary strategies of human T-cell lymphotropic virus type II. Gene 261, 171–180 (2000).

    Article  CAS  PubMed  Google Scholar 

  44. Lemey, P., Pybus, O. G., Van Dooren, S. & Vandamme, A. M. A Bayesian statistical analysis of human T-cell lymphotropic virus evolutionary rates. Infect. Genet. Evol. 5, 291–298 (2005).

    Article  CAS  PubMed  Google Scholar 

  45. Suzuki, Y. et al. Slow evolutionary rate of GB virus C/hepatitis G virus. J. Mol. Evol. 48, 383–389 (1999).

    Article  CAS  PubMed  Google Scholar 

  46. Plyusnin, A. & Morzunov, S. P. Virus evolution and genetic diversity of hantaviruses and their rodent hosts. Curr. Topics Microbiol. Immunol. 256, 47–75 (2001).

    CAS  Google Scholar 

  47. García-Arenal, F., Fraile, A. & Malpica, J. M. Variability and genetic structure of plant virus populations. Annu. Rev. Phytopathol. 39, 157–186 (2001).

    Article  PubMed  Google Scholar 

  48. Charrel, R. N., De Micco, P. & De Lamballerie, X. Phylogenetic analysis of GB viruses A and C: Evidence for cospeciation between virus isolates and their primate hosts. J. Gen. Virol. 80, 2329–2335 (1999).

    Article  CAS  PubMed  Google Scholar 

  49. Hughes, A. L. & Friedman, R. Evolutionary diversification of protein-coding genes of hantaviruses. Mol. Biol. Evol. 17, 1558–1568 (2000).

    Article  CAS  PubMed  Google Scholar 

  50. Charleston, M. A. & Robertson, D. L. Preferential host switching by primate lentiviruses can account for phylogenetic similarity with the primate phylogeny. Syst. Biol. 51, 528–535 (2002).

    Article  CAS  PubMed  Google Scholar 

  51. Rodríguez-Cerezo, E., Elena, S. F., Moya, A. & García-Arenal, F. High genetic stability in natural populations of the plant RNA virus tobacco mild green mosaic virus. J. Mol. Evol. 32, 328–332 (1991).

    Article  Google Scholar 

  52. Kim, T. et al. Molecular analysis of quasispecies of Kyuri green mottle mosaic virus. Virus Res. 110, 161–167 (2005).

    Article  CAS  PubMed  Google Scholar 

  53. Marco, C. F. & Aranda, M. A. Genetic diversity of a natural population of Cucurbit yellow stunting disorder virus. J. Gen. Virol. 86, 815–822 (2005).

    Article  CAS  PubMed  Google Scholar 

  54. Fraile, A. et al. A century of tobamovirus evolution in an Australian population of Nicotiana glauca. J. Virol. 71, 8316–8320 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Li, H. Y. & Roossinck, M. J. Genetic bottlenecks reduce population variation in an experimental RNA virus population. J. Virol. 78, 10582–10587 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Fargette, D. et al. Rice yellow mottle virus, an RNA plant virus, evolves as rapidly as most RNA animal viruses. J. Virol. 16 Jan 2008 (doi:10.1128/JVI.02506-07).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. McGeoch, D. J. & Gatherer, D. Integrating reptilian herpesviruses into the family Herpesviridae. J. Virol. 79, 725–731 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Bernard, H. U. Coevolution of papillomaviruses with human populations. Trends Microbiol. 2, 140–143 (1994).

    Article  CAS  PubMed  Google Scholar 

  59. Pavesi, A. Utility of JC polyomavirus in tracing the pattern of human migrations dating to prehistoric times. J. Gen. Virol. 86, 1315–1326 (2005).

    Article  CAS  PubMed  Google Scholar 

  60. Hatwell, J. N. & Sharp, P. M. Evolution of human polyomavirus JC. J. Gen. Virol. 81, 1191–1200 (2000).

    Article  CAS  PubMed  Google Scholar 

  61. Shackelton, L. A. & Holmes, E. C. Phylogenetic evidence for the rapid evolution of human B19 erythrovirus. J. Virol. 80, 3666–3669 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Ge, L. M., Zhang, J. T., Zhou, X. P. & Li, H. Y. Genetic structure and population variability of tomato yellow leaf curl China virus. J. Virol. 81, 5902–5907 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Lopez-Bueno, A., Villarreal, L. P. & Almendral, J. M. Parvovirus variation for disease: a difference with RNA viruses? Curr. Topics Microbiol. Immunol. 299, 349–370 (2006).

    CAS  Google Scholar 

  64. Isnard, M., Granier, M., Frutos, R., Reynaud, B. & Peterschmitt, M. Quasispecies nature of three maize streak virus isolates obtained through different modes of selection from a population used to assess response to infection of maize cultivars. J. Gen. Virol. 79, 3091–3099 (1998).

    Article  CAS  PubMed  Google Scholar 

  65. Drake, J. W. A constant rate of spontaneous mutation in DNA-based microbes. Proc. Natl Acad. Sci. USA 88, 7160–7164 (1991). This study was the first comprehensive survey of rates of mutation among DNA viruses. It predicts that smaller DNA viruses should have higher mutation rates per nucleotide.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Harrison, B. D. & Robinson, D. J. Natural genomic and antigenic variation in whitefly-transmitted geminiviruses (begomoviruses). Annu. Rev. Phytopathol. 37, 369–398 (1999).

    Article  CAS  PubMed  Google Scholar 

  67. Hino, S. & Miyata, H. Torque teno virus (TTV): current status. Rev. Med. Virol. 17, 45–57 (2007).

    Article  CAS  PubMed  Google Scholar 

  68. Shackelton, L. A., Parrish, C. R., Truyen, U. & Holmes, E. C. High rate of viral evolution associated with the emergence of carnivore parvovirus. Proc. Natl Acad. Sci. USA 102, 379–384 (2005).

    Article  CAS  PubMed  Google Scholar 

  69. Umemura, T., Tanaka, Y., Kiyosawa, K., Alter, H. J. & Shih, J. W.-K. Observation of positive selection within hypervariable regions of a newly identified DNA virus (SEN virus). FEBS Lett. 510, 171–174 (2002).

    Article  CAS  PubMed  Google Scholar 

  70. Duffy, S. & Holmes, E. C. Phylogenetic evidence for rapid rates of molecular evolution in the single-stranded DNA begomovirus Tomato yellow leaf curl virus (TYLCV). J. Virol. 82, 957–965 (2008).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  72. Sanjuan, R., Moya, A. & Elena, S. F. The distribution of fitness effects caused by single-nucleotide substitutions in an RNA virus. Proc. Natl Acad. Sci. USA 101, 8396–8401 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Sasaki, A. & Nowak, M. A. Mutation landscapes. J. Theor. Biol. 224, 241–247 (2003).

    Article  PubMed  Google Scholar 

  74. Kamp, C., Wilke, C. O., Adami, C. & Bornholdt, S. Viral evolution under the pressure of an adaptive immune system: optimal mutation rates for viral escape. Complexity. 8, 28–33 (2003).

    Article  Google Scholar 

  75. Drake, J. W. Spontaneous mutation. Annu. Rev. Genet. 25, 125–146 (1991).

    Article  CAS  PubMed  Google Scholar 

  76. Anderson, J. P., Daifuku, R. & Loeb, L. A. Viral error catastrophe by mutagenic nucleosides. Annu. Rev. Microbiol. 58, 183–205 (2004).

    Article  CAS  PubMed  Google Scholar 

  77. Mansky, L. M. & Cunningham, K. S. Virus mutators and antimutators: roles in evolution, pathogenesis and emergence. Trends Genet. 16, 512–517 (2000).

    Article  CAS  PubMed  Google Scholar 

  78. Pfeiffer, J. K. & Kirkegaard, K. A single mutation in poliovirus RNA-dependent RNA polymerase confers resistance to mutagenic nucleotide analogs via increased fidelity. Proc. Natl Acad. Sci. USA 100, 7289–7294 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Sniegowski, P. D., Gerrish, P. J., Johnson, T. & Shaver, A. The evolution of mutation rates: separating causes from consequences. Bioessays. 22, 1057–1066 (2000).

    Article  CAS  PubMed  Google Scholar 

  80. Vignuzzi, M., Stone, J. K. & Andino, R. Ribavirin and lethal mutagenesis of poliovirus: molecular mechanisms, resistance and biological implications. Virus Res. 107, 173–181 (2005).

    Article  CAS  PubMed  Google Scholar 

  81. Furio, V., Moya, A. & Sanjuan, R. The cost of replication fidelity in an RNA virus. Proc. Natl Acad. Sci. USA 102, 10233–10237 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Graci, J. D. et al. Lethal mutagenesis of poliovirus mediated by a mutagenic pyrimidine analogue. J. Virol. 81, 11256–11266 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Elena, S. F., Carrasco, P., Daros, J. A. & Sanjuan, R. Mechanisms of genetic robustness in RNA viruses. EMBO Rep. 7, 168–173 (2006). This paper is a useful discussion of how genetic robustness is defined, measured and achieved in RNA viruses.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Ewald, P. W. Evolution of mutation rate and virulence among human retroviruses. Philos. Trans. R. Soc. Lond., B, Biol. Sci. 346, 333–341 (1994).

    Article  CAS  Google Scholar 

  85. Crotty, S., Cameron, C. E. & Andino, R. RNA virus error catastrophe: direct molecular test by using ribavirin. Proc. Natl Acad. Sci. USA 98, 6895–6900 (2001). This study was an important test of the hypothesis that increasing the mutation rate of an RNA virus (with the application of an antiviral drug) can result in viral extinction, probably owing to error catastrophe and/or lethal mutagenesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Mansky, L. M. & Bernard, L. C. 3′-azido-3′-deoxythymidine (AZT) and AZT-resistant reverse transcriptase can increase the in vivo mutation rate of human immunodeficiency virus type 1. J. Virol. 74, 9532–9539 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Summers, J. & Litwin, S. Examining the theory of error catastrophe. J. Virol. 80, 20–26 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Eigen, M. Steps Towards Life (Oxford University Press, Oxford, 1996).

    Google Scholar 

  89. Codoner, F. M., Daros, J. A., Sole, R. V. & Elena, S. F. The fittest versus the flattest: experimental confirmation of the quasispecies effect with subviral pathogens. PLoS Pathog. 2, e136 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  90. Elena, S. F. & Sanjuán, R. Adaptive value of high mutation rates of RNA viruses: separating causes from consequences. J. Virol. 79, 11555–11558 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Kuiken, T. et al. Host species barriers to influenza virus infections. Science 312, 394–397 (2006).

    Article  CAS  PubMed  Google Scholar 

  92. Kunkel, T. A. DNA replication fidelity. J. Biol. Chem. 279, 16895–16898 (2004).

    Article  CAS  PubMed  Google Scholar 

  93. Arguello-Astorga, G. R., Ascencia-Ibáñez, J. T., Dallas, M. B., Orozco, B. M. & Hanley-Bowdoin, L. High-frequency reversion of geminivirus replication protein mutants during infection. J. Virol. 81, 11005–11015 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Gutierrez, C. Geminivirus DNA replication. Cell. Mol. Life Sci. 56, 313–329 (1999).

    Article  CAS  PubMed  Google Scholar 

  95. Chen, Y. et al. Analysis of 15 novel full-length BK virus sequences from three individuals: Evidence of a high intra-strain genetic diversity. J. Gen. Virol. 85, 2651–2663 (2004).

    Article  CAS  PubMed  Google Scholar 

  96. Shackelton, L. A., Rambaut, A., Pybus, O. G. & Holmes, E. C. JC virus evolution and its association with human populations. J. Virol. 80, 9928–9933 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Ghedin, E. et al. Large-scale sequencing of human influenza reveals the dynamic nature of viral genome evolution. Nature 437, 1162–1166 (2005).

    Article  CAS  PubMed  Google Scholar 

  98. Edwards, R. A. & Rohwer, F. Viral metagenomics. Nature Rev. Microbiol. 3, 504–510 (2005).

    Article  CAS  Google Scholar 

  99. Lukashov, V. V. & Goudsmit, J. Recent evolutionary history of human immunodeficiency virus type 1 subtype B: reconstruction of epidemic onset based on sequence distances to the common ancestor. J. Mol. Evol. 54, 680–691 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Rambaut, A. Estimating the rate of molecular evolution: incorporating non-contemporaneous sequences into maximum likelihood phylogenies. Bioinformatics. 16, 395–399 (2000).

    Article  CAS  PubMed  Google Scholar 

  101. Drummond, A. J. & Rambaut, A. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol. Biol. 7, 214 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  102. Drummond, A. J., Ho, S. Y. W., Phillips, M. J. & Rambaut, A. Relaxed phylogenetics and dating with confidence. PLoS Biol. 4, 699–710 (2006). This is an important methodological paper showing how rates of nucleotide substitution and times to common ancestry can be estimated from serially sampled data sets such as those from RNA viruses and ssDNA viruses.

    Article  CAS  Google Scholar 

  103. Malpica, J. M. et al. The rate and character of spontaneous mutation in an RNA virus. Genetics. 162, 1505–1511 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Aulicino, P. C., Holmes, E. C., Rocco, C., Mangano, A. & Sen, L. Extremely rapid spread of human immunodeficiency virus type 1 BF recombinants in Argentina. J. Virol. 81, 427–429 (2007).

    Article  CAS  PubMed  Google Scholar 

  105. Davis, P. L. et al. Phylogeography, population dynamics, and molecular evolution of European bat lyssaviruses. J. Virol. 79, 10487–10497 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by a National Science Foundation grant, DBI 06-03070, to S.D. We thank A. Rambaut for his comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Edward C. Holmes.

Related links

Related links

DATABASES

Entrez Genome

http://www.ncbi.nlm.nih.gov/sites/entrez?db=genome

bacteriophage RB69

canine parvovirus

GB virus C

hepatitis B virus

herpes simplex virus type 1

HTLV-I

human immunodeficiency virus type-1

human T-cell lymphotropic virus type II

JC polyomavirus

parvovirus B19

phage φX174

RNA phage Qβ

simian foamy virus

Tomato yellow leaf curl virus

vesicular stomatitis virus

FURTHER INFORMATION

Edward Holmes homepage

BEAST software

Glossary

Generation time

The time between rounds of production of viral progeny, including any time required for virions to seek a susceptible host cell, followed by adsorption and infection of the susceptible cell, then viral replication and release.

Coalescent

A population genetic theory that links the divergence times of a phylogenetic tree of individuals sampled from the same population with the demographic history (that is, rates of population growth and decline) of that population.

Positive selection

The fixation of advantageous alleles as a result of differential reproductive success.

Hypermutation

The long stretches of nucleotide transitions observed in RNA virus sequences (first noticed in human immunodeficiency virus with G-to-A transitions). This term can also be used to describe an elevated mutation rate of any kind, not necessarily in a run of adjacent nucleotides.

Amber mutaton reversion

The change of an amber stop codon (UAG) within a gene to a codon for an amino acid. This typically restores protein function in a gene that had been purposefully selected to contain an amber nonsense mutation as a genetic marker.

Stamping machine

Linear stamping machine replication is when the single virus that initiates an infection is the direct parent of all progeny genomes. That is, the parental genome (or its single complement) is the template for the semi-conservative replication of all the genomes that are produced in an infected cell. As there is only one template within the cell, progeny genomes accumulate linearly over time.

Geometric genome replication

A mode of viral replication in which the progeny genomes that are replicated early during infection can become templates for further genome replication. As the infection progresses, the number of templates for semi-conservative replication increases, and progeny genomes can be produced at an exponential, or geometric, rate.

Effective population size

The smallest theoretical population size that can evolve in the same way as the actual population under study. It is strongly influenced by population bottlenecks, such as those that occur during transmission of viruses between hosts, and therefore is often smaller than the total population size.

Linear regression

The estimation of a first-order relationship between two variables (for example, number of nucleotide substitutions and time), which involves fitting the best straight line to the data.

Constant molecular clock

The idea that nucleotide substitutions accumulate at a fixed (constant) rate over time, and that this can be used to estimate divergence times between sequences.

Maximum likelihood

A statistical method that selects the hypothesis (for example, the phylogenetic tree) that has the highest probability of explaining the data, under a specific model.

Bayesian Markov chain Monte Carlo

(MCMC). Bayesian methods incorporate prior information in assessing the probability of model parameters. Because the prior distribution (the users' belief about the probabilities of different parameter values before the data have been analysed) can have a large affect on the posterior distribution (the results) it must be chosen carefully. MCMC methods allow sampling from the posterior distribution to get an estimate of the distribution.

Relaxed molecular clock

A form of molecular clock in which rates of nucleotide substitution are allowed to vary among lineages.

Purifying selection

The purging of deleterious alleles as a result of differential reproductive success.

Cytotoxic T-lymphocyte

An antigen-specific T-cell of the vertebrate immune system that recognizes and destroys virus-infected cells.

Co-divergence

The parallel diversification or speciation of a parasite and its host, which is inferred when there is strong congruence between the phylogenetic trees of the host and parasite, and similar divergence times of corresponding nodes on the phylogenies.

Population bottleneck

The smallest size of a viral population at any point in viral propagation. A common bottleneck for viral populations occurs during transmission between hosts, when the population size can be as small as one virus or as large as several million virions.

Error threshold

The theoretical limit to the mutation rate of viruses, beyond which too many errors accumulate and populations of the virus become extinct. It is used to explain why it is difficult to generate RNA viruses with much higher mutation rates than those observed in natural isolates, and why RNA viruses have constrained genome sizes.

Burst size

The (average) number of progeny viruses produced from a single infected cell. This is more straightforward to measure for obligately lytic viruses than for viruses that can integrate into their host genomes.

Robustness

The constancy of a phenotype in the face of changing environments or changing genetics (mutations). Current research indicates that robustness is a trait that is under selection in viruses, and changes in viral robustness can be observed in laboratory experimental evolution.

Fitness landscape

A metaphorical contour map of the varied fitness values that are experienced by different genotypes of an organism. As a genotype moves through genotype space, it can climb to a higher fitness peak, or stumble down to a less-fit genotype.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Duffy, S., Shackelton, L. & Holmes, E. Rates of evolutionary change in viruses: patterns and determinants. Nat Rev Genet 9, 267–276 (2008). https://doi.org/10.1038/nrg2323

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg2323

This article is cited by

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