Skip to main content
Log in

The Common Ancestor Process Revisited

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

Abstract

We consider the Moran model in continuous time with two types, mutation, and selection. We concentrate on the ancestral line and its stationary type distribution. Building on work by Fearnhead (J. Appl. Probab. 39 (2002), 38–54) and Taylor (Electron. J. Probab. 12 (2007), 808–847), we characterise this distribution via the fixation probability of the offspring of all individuals of favourable type (regardless of the offspring’s types). We concentrate on a finite population and stay with the resulting discrete setting all the way through. This way, we extend previous results and gain new insight into the underlying particle picture.

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
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. Note the missing factor of 1/n in his Eq. (28).

References

  • Baake, E., & Bialowons, R. (2008). Ancestral processes with selection: branching and Moran models. Banach Cent. Publ., 80, 33–52.

    Article  MathSciNet  MATH  Google Scholar 

  • Barton, N. H., Etheridge, A. M., & Sturm, A. K. (2004). Coalescence in a random background. Ann. Appl. Probab., 14, 754–785.

    Article  MathSciNet  MATH  Google Scholar 

  • Birkhoff, G., & Rota, G. (1969). Ordinary differential equations (2nd ed.). Lexington: Xerox College Publ.

    MATH  Google Scholar 

  • Durrett, R. (2002). Probability models for DNA sequence evolution. New York: Springer.

    Book  MATH  Google Scholar 

  • Durrett, R. (2008). Probability models for DNA sequence evolution (2nd ed.). New York: Springer.

    Book  MATH  Google Scholar 

  • Etheridge, A. M., & Griffiths, R. C. (2009). A coalescent dual process in a Moran model with genic selection. Theor. Popul. Biol., 75, 320–330.

    Article  MATH  Google Scholar 

  • Etheridge, A. M., Griffiths, R. C., & Taylor, J. E. (2010). A coalescent dual process in a Moran model with genic selection, and the Lambda coalescent limit. Theor. Popul. Biol., 78, 77–92.

    Article  Google Scholar 

  • Ewens, W. J. (2004). Mathematical population genetics. I. Theoretical introduction (2nd ed.). New York: Springer.

    Book  MATH  Google Scholar 

  • Fearnhead, P. (2002). The common ancestor at a nonneutral locus. J. Appl. Probab., 39, 38–54.

    Article  MathSciNet  MATH  Google Scholar 

  • Ford, L. R. (1955). Differential equations (2nd ed.). New York: McGraw-Hill.

    MATH  Google Scholar 

  • Karlin, S., & Taylor, H. M. (1981). A second course in stochastic processes. San Diego: Academic Press.

    MATH  Google Scholar 

  • Kingman, J. F. C. (1982a). The coalescent. Stoch. Process. Appl., 13, 235–248.

    Article  MathSciNet  MATH  Google Scholar 

  • Kingman, J. F. C. (1982b). On the genealogy of large populations. J. Appl. Probab., 19A, 27–43.

    Article  MathSciNet  MATH  Google Scholar 

  • Krone, S. M., & Neuhauser, C. (1997). Ancestral processes with selection. Theor. Popul. Biol., 51, 210–237.

    Article  MATH  Google Scholar 

  • Mano, S. (2009). Duality, ancestral and diffusion processes in models with selection. Theor. Popul. Biol., 75, 164–175.

    Article  MATH  Google Scholar 

  • Neuhauser, C., & Krone, S. M. (1997). The genealogy of samples in models with selection. Genetics, 145, 519–534.

    Google Scholar 

  • Norris, J. R. (1999). Markov chains. Cambridge: Cambridge University Press.

    MATH  Google Scholar 

  • Pokalyuk, C., & Pfaffelhuber, P. (2013). The ancestral selection graph under strong directional selection. Theor. Popul. Biol. 87, 25–33.

    Article  MATH  Google Scholar 

  • Stephens, M., & Donnelly, P. (2003). Ancestral inference in population genetics models with selection. Aust. N. Z. J. Stat., 45, 901–931.

    Article  MathSciNet  MATH  Google Scholar 

  • Taylor, J. E. (2007). The common ancestor process for a Wright-Fisher diffusion. Electron. J. Probab., 12, 808–847.

    Article  MathSciNet  MATH  Google Scholar 

  • Vogl, C., & Clemente, F. (2012). The allele-frequency spectrum in a decoupled Moran model with mutation, drift, and directional selection, assuming small mutation rates. Theor. Popul. Biol., 81, 197–209.

    Article  MATH  Google Scholar 

  • Wakeley, J. (2008). Conditional gene genealogies under strong purifying selection. Mol. Biol. Evol., 25, 2615–2626.

    Article  Google Scholar 

  • Wakeley, J., & Sargsyan, O. (2009). The conditional ancestral selection graph with strong balancing selection. Theor. Popul. Biol., 75, 355–364.

    Article  MATH  Google Scholar 

Download references

Acknowledgements

It is our pleasure to thank Anton Wakolbinger for enlightening discussions, and for Fig. 2. We are grateful to Jay Taylor for valuable comments on the manuscript, and to Barbara Gentz for pointing out a gap in an argument at an earlier stage of the work. This project received financial support by Deutsche Forschungsgemeinschaft (DFG-SPP 1590), Grant no. BA2469/5-1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ellen Baake.

Appendix

Appendix

In Sect. 4.1, we have presented an alternative derivation of the boundary value problem for the conditional probability h. It remains to prove that \(\lim_{N \to \infty} h^{N}_{k_{N}} = h(x)\), with x∈[0,1], 0<k N <N, \(\lim_{N \to \infty} \frac{k_{N}}{N} = x\) and h as given as in (7).

Since \(h_{k}^{N} = \frac{k}{N} + \psi^{N}_{k}\) and h(x)=x+ψ(x), respectively, it suffices to show the corresponding convergence of the \(\psi^{N}_{k}\). For ease of exposition, we assume here that the process is stationary.

Lemma 3

Let \(\tilde{x}\) be as in (8). Then

$$ \lim_{N \to \infty} N \psi^N_{N-1} = \frac{\sigma}{1 + \theta \nu^{}_1}(1-\tilde{x}). $$

Proof

Since the stationary distribution \(\pi^{N}_{Z}\) of \(( Z^{N}_{t} )_{t \geqslant 0}\) (cf. (2)) satisfies

$$ \prod_{i=1}^{n-1} \frac{\lambda_i^{N}}{\mu_i^{N}} = \frac{\pi^N_Z (n)}{C_N} \frac{\mu^N_n}{\lambda^N_0}, $$
(47)

for 1⩽nN, Eq. (30) leads to

$$\begin{aligned} N \psi^N_{N-1} &= \frac{Ns^{}_N}{1 + N u^{}_N \nu^{}_1} \frac{\sum_{n=1}^{N} \pi^N_Z(n) \mu^N_{n} \frac{N-n}{N}}{\sum_{n=1}^{N}\pi^N_Z(n) \mu^N_{n}} \\ &=\frac{Ns^{}_N}{1 + N u^{}_N \nu^{}_1} \frac{\sum_{n=1}^{N} \pi^N_Z(n) \frac{n(N-n)^2}{N^3} \bigl( 1 + \frac{N u^{}_N \nu^{}_1}{N-n} \bigr)}{\sum_{n=1}^{N}\pi^N_Z(n) \frac{n(N-n)}{N^2} \bigl(1 + \frac{N u^{}_N \nu^{}_1}{N-n} \bigr)} , \end{aligned}$$

where we have used (1) in the last step. The stationary distribution of the rescaled process \(( X^{N}_{t} )_{t \geqslant 0}\) is given by \(( \pi^{N}_{X} ( \frac{i}{N} ) )_{0 \leqslant i \leqslant N}\), where \(\pi^{N}_{X} ( \frac{i}{N} ) = \pi^{N}_{Z} (i)\). Besides, the sequence of processes \((X^{N}_{t} )_{t \geqslant 0}\) converges to (X t ) t⩾0 in distribution, hence

$$\begin{aligned} \lim_{N \to \infty} N \psi^N_{N-1} &= \lim_{N \to \infty} \frac{N s^{}_N}{1 + N u^{}_N \nu^{}_1} \frac{\mathbb{E}_{\pi^N_X}\bigl( X^N ( 1- X^N )^2 ( 1 + \frac{u^{}_N \nu^{}_1}{1-X^N} ) \bigr)}{\mathbb{E}_{\pi^N_X}\bigl( X^N ( 1- X^N ) ( 1 + \frac{u^{}_N \nu^{}_1}{1-X^N} ) \bigr)} \\ &= \frac{\sigma}{1 + \theta \nu^{}_1} \frac{\mathbb{E}_{\pi^{}_X} ( X (1-X)^2 ) }{\mathbb{E}_{\pi^{}_X} ( X (1-X))} = \frac{\sigma}{1 + \theta \nu^{}_1}(1-\tilde{x}), \end{aligned}$$

as claimed. □

Remark 2

The proof gives an alternative way to obtain the initial value \(a^{}_{1}\) (cf. (18)) of recursion (15).

Theorem 3

For a given x∈[0,1], let \(( k_{N} )_{N \in \mathbb{N}}\) be a sequence with 0<k N <N and \(\lim_{N \to \infty} \frac{k_{N}}{N} = x\). Then

$$ \lim_{N \to \infty} \psi_{k_N}^N = \psi(x), $$

where ψ is the solution of the boundary value problem (11).

Proof

Using first Theorem 1, then (47), and finally (1), we obtain

$$\begin{aligned} \psi_k^N &= \frac{k(N-k)}{\mu_k^N} \sum _{n=1}^{N-k} \Biggl( \prod _{i=k+1}^{N-n} \frac{\lambda^N_i}{\mu^N_i} \Biggr) \biggl( \frac{\mu_{N-1}^N}{N-1} \psi^N_{N-1} - \frac{s^{}_N (n-1)}{N^2} \biggr) \\ &= \frac{k(N-k)}{\mu_k^N} \bigl( \mu^N_{k+1} \pi^N_Z (k+1) \bigr)^{-1} \\ &\quad {}\times \sum_{n=0}^{N-k-1} \mu^N_{N-n} \pi^N_Z (N-n) \biggl( \frac{\mu_{N-1}^N}{N-1} \psi^N_{N-1} - \frac{s^{}_N n}{N^2} \biggr) \\ &= \biggl( 1 + \mathcal{O} \biggl( \frac{1}{N} \biggr) \biggr) \biggl( \frac{k+1}{N} \frac{N-k-1}{N} \pi^N_Z (k+1) \biggr)^{-1} \\ & \hphantom{=} \times \frac{1}{N} \sum_{n=0}^{N-k-1} \pi^N_Z (N-n) \frac{N-n}{N} \frac{n}{N} \biggl( 1 + \frac{Nu^{}_N \nu^{}_1}{n} \biggr) \\ & \hphantom{= \times \frac{1}{N} \sum_{n=0}^{N-k-1}} \times \biggl( (1 + N u^{}_N \nu^{}_1) N \psi^N_{N-1} - N s^{}_N \frac{n}{N} \biggr). \end{aligned}$$

In order to analyse the convergence of this expression, define

$$\begin{aligned} S_1^N (k) &:= \frac{k+1}{N} \frac{N-k-1}{N} \pi^N_Z (k+1), \\ S_2^N (k) &:= \frac{1}{N} \sum _{n=0}^{N-k-1} \pi^N_Z (N-n) \frac{N-n}{N} \frac{n}{N} \biggl( (1 + N u^{}_N \nu^{}_1) N \psi^N_{N-1} - N s^{}_N \frac{n}{N} \biggr) \\ & \hphantom{:} = \int_0^1 T^N_k (y) \,dy, \\ S_3^N (k) &:= \frac{1}{N} \sum _{n=0}^{N-k-1} \pi^N_Z (N-n) \frac{N-n}{N} u^{}_N \nu^{}_1 \biggl( (1 + N u^{}_N \nu^{}_1) N \psi^N_{N-1} - N s^{}_N \frac{n}{N} \biggr) \\ & \hphantom{:} = \int_0^1 \tilde{T}_k^N (y) \,dy, \end{aligned}$$

with step functions \(T^{N}_{k} : [0,1] \rightarrow \mathbb{R}\), \(\tilde{T}^{N}_{k} : [0,1] \rightarrow \mathbb{R}\) given by

Consider now a sequence \(( k_{N} )_{N \in \mathbb{N}}\) as in the assumptions. Then \(\lim_{N \to \infty} \pi^{N}_{Z} (k_{N}) = \pi^{}_{X}(x)\) (cf. Durrett, 2008, p. 319), and due to Lemma 3

Since \(T^{N}_{k}\) and \(\tilde{T}^{N}_{k}\) are bounded, we have

$$\begin{aligned} \lim_{N \to \infty} S_2^N (k_N) &= \int_0^{1-x} \pi^{}_X (1-y) (1-y)y \bigl(\sigma (1- \tilde{x})- \sigma y \bigr) \,dy, \\ \lim_{N \to \infty} S_3^N (k_N) &=0, \end{aligned}$$

thus

$$ \lim_{N \to \infty} \psi^N_{k_N} = \bigl( x (1-x) \pi^{}_X (x) \bigr)^{-1} \int_0^{1-x} \pi^{}_X (1-y) (1-y)y \bigl(\sigma (1- \tilde{x})- \sigma y \bigr) \,dy. $$

Substituting on the right-hand side yields

$$\begin{aligned} \lim_{N \to \infty} \psi^N_{k_N} &= \bigl( x (1-x) \pi^{}_X (x) \bigr)^{-1} \sigma \int_x^1 \pi^{}_X(y)y(1-y) (y - \tilde{x}) \,dy \\ &= \bigl( x (1-x) \pi^{}_X (x) \bigr)^{-1} \sigma \biggl[ \int_0^1 \pi^{}_X (y) y (1-y) (y - \tilde{x})\,dy \\ &\quad {} + \int_0^x \pi^{}_X (y)y (1-y) ( \tilde{x}-y)\,dy \biggr] \\ &= \bigl( x (1-x) \pi^{}_X (x) \bigr)^{-1} \sigma \int_0^x \pi^{}_X (y) y (1-y) (\tilde{x}-y)\,dy = \psi(x) , \end{aligned}$$

where the second-last equality goes back to the definition of \(\tilde{x}\) in (8), and the last is due to (7), (10), and (5). □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Kluth, S., Hustedt, T. & Baake, E. The Common Ancestor Process Revisited. Bull Math Biol 75, 2003–2027 (2013). https://doi.org/10.1007/s11538-013-9871-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9871-0

Keywords

Navigation