Skip to main content

Advertisement

Log in

Virus Replication as a Phenotypic Version of Polynucleotide Evolution

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

In this paper, we revisit and adapt to viral evolution an approach based on the theory of branching process advanced by Demetrius et al. (Bull. Math. Biol. 46:239–262, 1985), in their study of polynucleotide evolution. By taking into account beneficial effects, we obtain a non-trivial multivariate generalization of their single-type branching process model. Perturbative techniques allows us to obtain analytical asymptotic expressions for the main global parameters of the model, which lead to the following rigorous results: (i) a new criterion for “no sure extinction”, (ii) a generalization and proof, for this particular class of models, of the lethal mutagenesis criterion proposed by Bull et al. (J. Virol. 18:2930–2939, 2007), (iii) a new proposal for the notion of relaxation time with a quantitative prescription for its evaluation, (iv) the quantitative description of the evolution of the expected values in four distinct “stages”: extinction threshold, lethal mutagenesis, stationary “equilibrium”, and transient. Finally, based on these quantitative results, we are able to draw some qualitative conclusions.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  • Aguirre, J., & Manrubia, S. C. (2008). Effects of spatial competition on the diversity of a quasispecies. Phys. Rev. Lett., 100(3), 038106.

    Article  Google Scholar 

  • Aguirre, J., Lázaro, E., & Manrubia, S. C. (2009). A trade-off between neutrality and adaptability limits the optimization of viral quasispecies. J. Theor. Biol., 261, 148–155.

    Article  Google Scholar 

  • Antoneli, F., Dias, A. P. S., Golubitsky, M., & Wang, Y. (2005). Patterns of synchrony in lattice dynamical systems. Nonlinearity, 18(5), 2193–2209. doi:10.1088/0951-7715/18/5/016.

    Article  MathSciNet  MATH  Google Scholar 

  • Antoneli, F., Bosco, F., Castro, D., & Janini, L. M. (2013a). Viral evolution and adaptation as a multivariate branching process. In R. P. Mondaini (Ed.), BIOMAT 2012—proceedings of the international symposium on mathematical and computational biology. Singapore: World Scientific. http://arxiv.org/abs/1110.3368.

    Google Scholar 

  • Antoneli, F., Bosco, F., & Janini, L. M. (2013b, in preparation). Fluctuations and correlations of a phenotypic quasi-species.

  • Athreya, K. B., & Ney, P. E. (1972). Branching processes. Berlin: Springer.

    Book  MATH  Google Scholar 

  • Bull, J. J., Meyers, L. A., & Lachmann, M. (2005). Quasispecies made simple. PLoS Comput. Biol., 1(6), 450–460.

    Article  Google Scholar 

  • Bull, J. J., Sanjuán, R., & Wilke, C. O. (2007). Theory of lethal mutagenesis for viruses. J. Virol., 18(6), 2930–2939. doi:10.1128/JVI.01624-06.

    Article  Google Scholar 

  • Bull, J. J., Sanjuán, R., & Wilke, C. O. (2008). Lethal mutagenesis. In E. Domingo, C. R. Parrish, & J. J. Holland (Eds.), Origin and evolution of viruses (2nd ed., pp. 207–218). London: Academic Press. Chap. 9. doi:10.1016/B978-0-12-374153-0.00009-6.

    Chapter  Google Scholar 

  • Capitán, J. A., Cuesta, J. A., Manrubia, S. C., & Aguirre, J. (2011). Severe hindrance of viral infection propagation in spatially extended hosts. PLoS ONE, 6(8), 358. doi:10.1371/journal.pone.0023358. e23.

    Article  Google Scholar 

  • Carrasco, P., de la Iglesia, F., & Elena, S. F. (2007). Distribution of fitness and virulence effects caused by single-nucleotide substitutions in tobacco etch virus. J. Virol., 18(23), 12979–12984.

    Article  Google Scholar 

  • Castro, D. (2011). Simulação computacional e análise de um modelo fenotípico de evolução viral. Master’s thesis, Universidade Federal de São Paulo—UNIFESP, São Paulo.

  • Castro, D., Antoneli, F., Bosco, F., & Janini, L. M. (2011, in preparation). Computational simulation of a model for viral evolution.

  • Crotty, S., Cameron, C. E., & Andino, R. (2001). RNA virus error catastrophe: direct molecular test by using ribavirin. Proc. Natl. Acad. Sci. USA, 98, 6895–6900.

    Article  Google Scholar 

  • Cuesta, J. A. (2011). Huge progeny production during transient of a quasi-species model of viral infection, reproduction and mutation. Math. Comput. Model., 54, 1676–1681.

    Article  MathSciNet  MATH  Google Scholar 

  • Cuesta, J. A., Aguirre, J., Capitán, J. A., & Manrubia, S. C. (2011). Struggle for space: viral extinction through competition for cells. Phys. Rev. Lett., 106(2), 028104.

    Article  Google Scholar 

  • Delbrück, M. (1945). The burst size distribution in the growth of bacterial viruses (bacteriophages). J. Bacteriol., 50(2), 131–135.

    Google Scholar 

  • Demetrius, L., & Legendre, S. (2013). Evolutionary entropy predicts the outcome of selection: competition for resources that vary in abundance and diversity. Theor. Popul. Biol., 83, 39–54.

    Article  Google Scholar 

  • Demetrius, L., Schuster, P., & Sigmund, K. (1985). Polynucleotide evolution and branching processes. Bull. Math. Biol., 47(2), 239–262. http://www.springerlink.com/content/h362q6007842h042/.

    MathSciNet  MATH  Google Scholar 

  • Dietz, K. (2005). Darwinian fitness, evolutionary entropy and directionality theory. BioEssays, 27, 1097–1101.

    Article  Google Scholar 

  • Domingo, E., et al. (1985). The quasispecies (extremely heterogeneous) nature of viral RNA genome populations: biological relevance—a review. Gene, 40(1), 1–8.

    Article  Google Scholar 

  • Domingo, E., Baranowski, E., Ruiz-Jarabo, C. M., Martín-Hernández, A. M., Sáiz, J. C., & Escarmís, C. (1998). Quasispecies structure and persistence of RNA viruses. Emerg. Infect. Dis., 4(4), 521–527.

    Article  Google Scholar 

  • Domingo-Calap, P., Cuevas, J. M., & Sanjuán, R. (2009). The fitness effects of random mutations in single-stranded DNA and RNA bacteriophages. PLoS Genet., 5(11), e1000742.

    Article  Google Scholar 

  • Eigen, M. (1971). Selforganization of matter and the evolution of biological macromolecules. Naturwissenschaften, 58, 465–523.

    Article  Google Scholar 

  • Eigen, M., & Schuster, P. (1979). The hypercycle. a principle of natural self-organization. Berlin: Springer.

    Google Scholar 

  • Elena, S. F., & Sanjuán, R. (2005). Adaptive value of high mutation rates of RNA viruses: separating causes from consequences. J. Virol., 79(18), 11555–11558. doi:10.1128/JVI.79.18.11555-11558.2005.

    Article  Google Scholar 

  • Elena, S. F., Wilke, C. O., Ofria, C., & Lenski, R. E. (2007). Effects of population size and mutation rate on the evolution of mutational robustness. Evolution, 61(3), 666–674. doi:10.1111/j.1558-5646.2007.00064.x.

    Article  Google Scholar 

  • Escarmís, C., Lázaro, E., & Manrubia, S. C. (2006). Population bottlenecks in quasispecies dynamics. Curr. Top. Microbiol. Immunol., 299, 141–170.

    Article  Google Scholar 

  • Eyre-Walker, A., & Keightley, P. D. (2007). The distribution of fitness effects of new mutations. Nat. Rev. Genet., 8(8), 610–618. doi:10.1038/nrg2146.

    Article  Google Scholar 

  • Feller, W. (1968). An introduction to probability theory and its applications (3rd ed., Vol. 1). New York: Wiley.

    MATH  Google Scholar 

  • Forster, R., Adami, C., & Wilke, C. O. (2006). Selection for mutational robustness in finite populations. J. Theor. Biol., 243(2), 181–190. doi:10.1016/j.jtbi.2006.06.020.

    Article  MathSciNet  Google Scholar 

  • Gantmacher, F. (2005). Applications of the theory of matrices. New York: Dover.

    Google Scholar 

  • Gray, R. M. (2006). Toeplitz and circulant matrices: a review. Hanover: Now Publishers.

    MATH  Google Scholar 

  • Grenander, U., & Szegö, G. (1958). Toeplitz forms and their applications. Berkeley: University of California Press.

    MATH  Google Scholar 

  • Harris, T. (1963). The theory of branching processes. Berlin: Springer.

    Book  MATH  Google Scholar 

  • Hermisson, J., Redner, O., Wagner, H., & Baake, E. (2002). Mutation-selection balance: ancestry, load, and maximum principle. Theor. Popul. Biol., 62(1), 9–46. doi:10.1006/tpbi.2002.1582.

    Article  MATH  Google Scholar 

  • Imhof, M., & Schlotterer, C. (2001). Fitness effects of advantageous mutations in evolving escherichia coli populations. Proc. Natl. Acad. Sci. USA, 98(3), 1113–1117. doi:10.1073/pnas.98.3.1113.

    Article  Google Scholar 

  • Jagers, P., Klebaner, F. C., & Sagitov, S. (2007). On the path to extinction. Proc. Natl. Acad. Sci. USA, 104(15), 6107–6111.

    Article  MathSciNet  Google Scholar 

  • Jiřina, M. (1970). A simplified proof of the Sevastyanov theorem on branching processes. Ann. Inst. Henri Poincaré, B Calc. Probab. Stat., 6(1), 1–7.

    MATH  Google Scholar 

  • Keightley, P. D., & Lynch, M. (2003). Toward a realistic model of mutations affecting fitness. Evolution, 57(3), 683–685; discussion 686–689.

    Google Scholar 

  • Kimmel, M., & Axelrod, D. (2002). Branching processes in biology. New York: Springer.

    MATH  Google Scholar 

  • Kowald, A., & Demetrius, L. (2005). Directionality theory: a computational study of an entropic principle. Proc. R. Soc. B, 272, 741–749.

    Article  Google Scholar 

  • Kurtz, T. G., Lyons, R., Pemantle, R., & Peres, Y. (1994). A conceptual proof of the Kesten-Stigum theorem for multi-type branching processes. In K. Athreya & P. Jagers (Eds.), IMA math. appl.: Vol. 84. Classical and modern branching processes (pp. 181–185). New York: Springer.

    Chapter  Google Scholar 

  • Lázaro, E., Escarmís, C., Domingo, E., & Manrubia, S. C. (2002). Modeling viral genome fitness evolution associated with serial bottleneck events: evidence of stationary states of fitness. J. Virol., 76(17), 8675–8681.

    Article  Google Scholar 

  • Leuthäusser, I. (1987). Statistical mechanics of Eigen’s evolution model. J. Stat. Phys., 48(1/2), 343–360.

    Article  Google Scholar 

  • Manrubia, S. C., Lázaro, E., Pérez-Mercader, J., Escarmís, C., & Domingo, E. (2003). Fitness distributions in exponentially growing asexual populations. Phys. Rev. Lett., 90(18), 188102.

    Article  Google Scholar 

  • Meyer, C. (2000). Matrix analysis and applied linear algebra. Philadelphia: SIAM.

    Book  MATH  Google Scholar 

  • Miralles, R., Gerrish, P. J., Moya, A., & Elena, S. F. (1999). Clonal interference and the evolution of RNA viruses. Science, 285(5434), 1745–1747.

    Article  Google Scholar 

  • Ojosnegros, S., Beerenwinkel, N., Antal, T., Nowak, M. A., Escarmís, C., & Domingo, E. (2010a). Competition-colonization dynamics in an RNA virus. Proc. Natl. Acad. Sci. USA, 107(5), 2108–2112. doi:10.1073/pnas.0909787107.

    Article  Google Scholar 

  • Ojosnegros, S., Beerenwinkel, N., & Domingo, E. (2010b). Competition-colonization dynamics: an ecology approach to quasispecies dynamics and virulence evolution in RNA viruses. Commun. Integr. Biol., 3(4), 333–336.

    Article  Google Scholar 

  • Orr, H. A. (2003). The distribution of fitness effects among beneficial mutations. Genetics, 163(4), 1519–1526.

    Google Scholar 

  • Parera, M., Fernàndez, G., Clotet, B., & Martínez, M. A. (2007). HIV-1 protease catalytic efficiency effects caused by random single amino acid substitutions. Mol. Biol. Evol., 24(2), 382–387. doi:10.1093/molbev/msl168.

    Article  Google Scholar 

  • Rokyta, D. R., Beisel, C. J., Joyce, P., Ferris, M. T., Burch, C. L., & Wichman, H. A. (2008). Beneficial fitness effects are not exponential for two viruses. J. Mol. Evol., 67(4), 368–376. doi:10.1007/s00239-008-9153-x.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Steinhauer, D. A., Domingo, E., & Holland, J. J. (1992). Lack of evidence for proofreading mechanisms associated with an RNA virus polymerase. Gene, 122(2), 281–288.

    Article  Google Scholar 

  • Swetina, J., & Schuster, P. (1982). Self-replication with errors: a model for polynucleotide replication. Biophys. Chem., 16(4), 329–345. doi:10.1016/0301-4622(82)87037-3.

    Article  Google Scholar 

  • Thompson, C. J., & McBride, J. L. (1974). On Eigen’s theory of the self-organization of matter and the evolution of biological macromolecules. Math. Biosci., 21(1–2), 127–142. doi:10.1016/0025-5564(74)90110-2.

    Article  MathSciNet  MATH  Google Scholar 

  • Timm, A., & Yin, J. (2012). Kinetics of virus production from single cells. Virology, 424(1), 11–17. doi:10.1016/j.virol.2011.12.005.

    Article  Google Scholar 

  • Watson, H. W., & Galton, F. (1874). On the probability of the extinction of families. J. Anthropol. Inst. G. B. Irel., 4, 138–144.

    Google Scholar 

  • Wilke, C. O. (2003). Probability of fixation of an advantageous mutant in a viral quasispecies. Genetics, 163(2), 467–474.

    Google Scholar 

  • Wilke, C. O., Wang, J. L., Ofria, C., Lenski, R., & Adami, C. (2001). Evolution of digital organisms at high mutation rates leads to survival of the flattest. Nature, 412(6844), 331–333. doi:10.1038/35085569.

    Article  Google Scholar 

  • Wilkinson, J. H. (1965). Algebraic eigenvalue problem. Oxford: Oxford University Press.

    MATH  Google Scholar 

  • Zhu, Y., Yongky, A., & Yin, J. (2009). Growth of an RNA virus in single cells reveals a broad fitness distribution. Virology, 385(1), 39–46. doi:10.1016/j.virol.2008.10.031.

    Article  Google Scholar 

  • Ziehe, M., & Demetrius, L. (2005). Directionality theory: an empirical study of an entropic principle in life-history evolution. Proc. R. Soc. B, 272, 1185–1194.

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank C.O. Wilke for some pertinent comments on a previous version of this work. FA wishes to acknowledge the support of CNPq through the grant PQ-313224/2009-9. FB received support from the Brazilian agency FAPESP. DC received financial support from the Brazilian agency CAPES.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fernando Antoneli.

Additional information

Dedicated to the memory of our dear collaborator and friend Francisco de Assis Ribas Bosco (1955–2012).

Appendix: Spectral Perturbation Theory

Appendix: Spectral Perturbation Theory

In order to carry out the perturbative calculations, we need to solve the spectral problem for the matrix

The eigenvalues \(\lambda_{r}^{0}\) of the mean matrix M 0 are

$$\lambda_r^0=rc=r(1-d) \quad r=0,\ldots,R. $$

In particular, the Malthusian parameter is the largest positive eigenvalue

$$m^0=\lambda_R^0=Rc=R(1-d). $$

The normalized left and right eigenvectors, \(\boldsymbol{v}_{r}^{0}\) and \(\boldsymbol {u}_{r}^{0}\), associated to the eigenvalue \(\lambda_{r}^{0}\), satisfy

$$\bigl(\boldsymbol{v}_s^0 \bigr)^{\mathrm{t}} \boldsymbol{u}_s^0=1 \quad\text{and}\quad \boldsymbol{1}^{\mathrm{t}} \boldsymbol{u}_s^0=1. $$

Writing \(\boldsymbol{v}_{r}^{0}\) and \(\boldsymbol{u}_{r}^{0}\) in components as

one has that

$$v_r^0(k)= \begin{cases} 0 & \mbox{for}\ k=0,\ldots,r-1, \\ \frac{-d(r+1)}{(1-d)^{r+1}} & \mbox{for}\ k=r+1,\\ \frac{((-1)d)^{k-r}}{(1-d)^{k}} & \mbox{for}\ k=r,r+2,\ldots,R \end{cases} $$

and

$$u_r^0(k)= \begin{cases} \binom{r}{k} (1-d)^k d^{r-k} & \mbox{for}\ k=0,\ldots,r, \\ 0 & \mbox{for}\ k=r+1,\ldots,R. \end{cases} $$

Now we write the eigenvalue λ r of M as a function of the parameter b, expanded as a power series

$$\lambda_r = \lambda_r^0 + \lambda_r^1 b + \lambda_r^2 b^2 + \cdots, $$

where \(\lambda_{r}^{0}\) is the corresponding eigenvalue of M 0. Clearly, \(\lambda_{r}(b)\to\lambda_{r}^{0}\) as b→0. The higher order perturbation coefficients \(\lambda_{r}^{i}\), (i=1,2,3,…) are written in terms of all the eigenvalues \(\lambda_{s}^{0}\) (s=0,…,R) of M 0 and their associated left and right normalized eigenvectors \(\boldsymbol{v}_{s}^{0}\) and \(\boldsymbol{u}_{s}^{0}\) and the perturbation matrix

Note that the operator norm of P is ∥P∥=2(R−1) and so the magnitude of the perturbation is 2b(R−1).

The general expressions for the first and second order coefficients of the perturbation expansion of an eigenvalue are

We want to use these formulas to compute a perturbation approximation (for b around 0) of the Malthusian parameter m(b) of ( M). Recall that, m(0)=m 0=R(1−d) and the corresponding left and right eigenvectors are given by

with u R (k)=binom(k;R,1−d). The first-order coefficient of the Malthusian parameter is

$$m^1 = \bigl(\boldsymbol{v}_R^0 \bigr)^{\mathrm {t}} \boldsymbol{P}\boldsymbol{u}_R^0 = \dfrac {(R-1)u_R(R-1)}{(1-d)^R} = \dfrac{R(R-1)d}{1-d}. $$

For the second-order coefficient, we also need

$$\boldsymbol{v}_{R-1}^0=1/(1-d)^{R-1} \bigl(0, \ldots,0,1,-Rd/(1-d) \bigr). $$

The second-order coefficient is

Analogous perturbative formulas exist for the perturbation expansion of the left and right eigenvector corresponding to an eigenvalue λ r . The left and right eigenvector v r and u r of the matrix M associated with the eigenvalue λ r can be written as a function of the parameter b, expanded as a vector-valued power series:

where \(\boldsymbol{v}_{r}^{0}\) and \(\boldsymbol{u}_{r}^{0}\) are, respectively, the left and right eigenvectors of the matrix M 0 associated to the eigenvalue \(\lambda_{r}^{0}\). The perturbation terms \(A_{r,s}^{i}\) and \(B_{r,s}^{i}\) with i=1,2,3,… , can be written in terms of the eigenvalues of M 0 and their associated left and right eigenvectors and the perturbation matrix P. Observe that when b→0 we get the normalized left and right eigenvectors \(\boldsymbol{v}_{R}^{0}\) and \(\boldsymbol{u}_{R}^{0}\) associated to the dominant eigenvalue m 0=(1−d)R of the mean matrix M 0.

We are interested in the normalized eigenvectors v R and u R associated to the Malthusian parameter m=(1−d)R

$$\boldsymbol{v}_R=\boldsymbol{v}_R^0 + b \sum _{k=0}^{R} \alpha_k \boldsymbol{v}_k^0 + \mathit {O}\bigl(b^2 \bigr), $$

with \(\alpha_{k}=A_{R,k}^{1}\), (0⩽kR−1) and \(\alpha_{R}=-\sum_{k=0}^{R-1} \alpha_{k}\),

$$\boldsymbol{u}_R=\boldsymbol{u}_R^0 + b \sum _{k=0}^{R} \beta_k \boldsymbol{u}_k^0 + \mathit {O}\bigl(b^2 \bigr), $$

with \(\beta_{k}=B_{R,k}^{1}\), (0⩽kR−1) and \(\beta_{R}=-\sum_{k=0}^{R-1} \beta_{k}\).

The first-order coefficients are given by

and the second-order terms are given by

for s=0,…,R−1.

One may use the formula of v R up to second order

$$\boldsymbol{v}_R = \boldsymbol{v}_R^0 + b \dfrac{(R-1)}{(1-d)^2}\boldsymbol{v}_{R-1}^0 + \mathit {O}\bigl(b^2 \bigr) $$

and the first-order expansion \(\boldsymbol{u}_{R} = \boldsymbol{u}_{R}^{0} + \mathit {O}(b)\) in order to estimate the product of coefficients u R (r)v R (r) (up to their leading order). For instance, it is easy to find that

In general, using a higher order expansion formula for v R , one concludes that

$$u_R(R-r)v_R(R-r) = \mathit {O}\bigl(b^r \bigr). $$

This follows form the fact that the coefficient of order b r in the perturbation expansion of v R is a linear combination of the left eigenvectors \(\boldsymbol{v}_{R-1}^{0},\ldots,\boldsymbol{v}_{r-R}^{0}\), all of which have zeros in the first r−1 components.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Antoneli, F., Bosco, F., Castro, D. et al. Virus Replication as a Phenotypic Version of Polynucleotide Evolution. Bull Math Biol 75, 602–628 (2013). https://doi.org/10.1007/s11538-013-9822-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9822-9

Keywords

Navigation