Skip to main content
Log in

A Stochastic Model for Reproductive Isolation Under Asymmetrical Mating Preferences

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

Abstract

More and more evidence shows that mating preference is a mechanism that may lead to a reproductive isolation event. In this paper, a haploid population living on two patches linked by migration is considered. Individuals are ecologically and demographically neutral on the space and differ only on a trait, a or A, affecting both mating success and migration rate. The special feature of this paper is to assume that the strengths of the mating preference and the migration depend on the trait carried. Indeed, patterns of mating preferences are generally asymmetrical between the subspecies of a population. I prove that mating preference interacting with frequency-dependent migration behavior can lead to a reproductive isolation. Then, I describe the time before reproductive isolation occurs. To reach this result, I use an original method to study the limiting dynamical system, analyzing first the system without migration and adding migration with a perturbation method. Finally, I study how the time before reproductive isolation is influenced by the parameters of migration and of mating preferences, highlighting that large migration rates tend to favor types with weak mating preferences.

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
Fig. 5

Similar content being viewed by others

References

  • Athreya KB, Ney PE (1972) Branching processes. Springer, Berlin

    Book  MATH  Google Scholar 

  • Berger M, Gostiaux B (1992) Géométrie différentielle: variétés, courbes et surfaces. Presses Universitaires de France

  • Blaustein AR (1983) Kin recognition mechanisms: phenotypic matching or recognition alleles? Am Nat 121(5):749–754

    Article  Google Scholar 

  • Bolker B, Pacala SW (1997) Using moment equations to understand stochastically driven spatial pattern formation in ecological systems. Theor Popul Biol 52(3):179–197

    Article  MATH  Google Scholar 

  • Boughman JW (2001) Divergent sexual selection enhances reproductive isolation in sticklebacks. Nature 411(6840):944–948

    Article  Google Scholar 

  • Brown JS, Pavlovic NB (1992) Evolution in heterogeneous environments: effects of migration on habitat specialization. Evolut Ecol 6(5):360–382

    Article  Google Scholar 

  • Bürger R, Schneider KA (2006) Intraspecific competitive divergence and convergence under assortative mating. Am Natur 167(2):190–205

    Article  Google Scholar 

  • Clobert J, Danchin E, Dhondt AA, Nichols JD (2001) Dispersal. Oxford University Press, Oxford

    Google Scholar 

  • Collet P, Méléard S, Metz JA (2011) A rigorous model study of the adaptive dynamics of mendelian diploids. J Math Biol 67(3):569–607

    Article  MathSciNet  MATH  Google Scholar 

  • Coron C, Costa M, Leman H, Smadi C (2018) A stochastic model for speciation by mating preferences. J Math Biol 76(6):1421–1463

    Article  MathSciNet  MATH  Google Scholar 

  • Cuevas J, Moya A, Elena S (2003) Evolution of rna virus in spatially structured heterogeneous environments. J Evolut Biol 16(3):456–466

    Article  Google Scholar 

  • Dieckmann U, Law R (2000) Relaxation projections and the method of moments. In: Dieckmann U, Law R, Metz JAJ (eds) The geometry of ecological interactions: symplifying spatial complexity. Cambridge University Press, Cambridge, pp 412–455

    Chapter  Google Scholar 

  • Elena SF, Agudelo-Romero P, Lalić J (2009) The evolution of viruses in multi-host fitness landscapes. Open Virol J 3:1–6

    Article  Google Scholar 

  • Ethier SN, Kurtz TG (1986) Markov processes: characterization and convergence. Wiley, New York

  • Fournier N, Méléard S (2004) A microscopic probabilistic description of a locally regulated population and macroscopic approximations. Ann Appl Probab 14(4):1880–1919

    Article  MathSciNet  MATH  Google Scholar 

  • Gavrilets S (2004) Fitness landscapes and the origin of species (MPB-41), vol 41. Princeton University Press, Princeton

    Google Scholar 

  • Gavrilets S (2014) Models of speciation: Where are we now? J Hered 105(S1):743–755

    Article  Google Scholar 

  • Gavrilets S, Boake CR (1998) On the evolution of premating isolation after a founder event. Am Natur 152(5):706–716

    Article  Google Scholar 

  • Gregorius HR (1989) Characterization and analysis of mating systems. Ekopan Verlag, Witzenhausen

  • Hartl DL, Clark AG, Clark AG (1997) Principles of population genetics, vol 116. Sinauer Associates, Sunderland

    Google Scholar 

  • Heinzmann D (2009) Extinction times in multitype Markov branching processes. J Appl Probab 46(1):296–307

    Article  MathSciNet  MATH  Google Scholar 

  • Hollocher H, Ting CT, Pollack F, Wu CI (1997) Incipient speciation by sexual isolation in drosophila melanogaster: variation in mating preference and correlation between sexes. Evolution 51:1175–1181

    Google Scholar 

  • Hoppensteadt FC (1966) Singular perturbations on the infinite interval. Trans Am Math Soc 123:521–535

    Article  MathSciNet  MATH  Google Scholar 

  • Jones A, Ratterman N (2009) Mate choice and sexual selection: What have we learned since darwin? PNAS 106(1):10,001–10,008

    Article  Google Scholar 

  • Jugovic J, Crne M, Luznik M (2017) Movement, demography and behaviour of a highly mobile species: a case study of the black-veined white, aporia crataegi (lepidoptera: Pieridae). Eur J Entomol 114:113

    Article  Google Scholar 

  • Karlin S, McGregor J (1972) Polymorphisms for genetic and ecological systems with weak coupling. Theor Popul Biol 3(2):210–238

    Article  MathSciNet  MATH  Google Scholar 

  • Kirkpatrick M (1982) Sexual selection and the evolution of female choice. Evolution 41:1–12

    Article  Google Scholar 

  • Kondrashov AS, Shpak M (1998) On the origin of species by means of assortative mating. Proc R Soc Lond B Biol Sci 265(1412):2273–2278

    Article  Google Scholar 

  • Lande R (1981) Models of speciation by sexual selection on polygenic traits. Proc Nat Acad Sci 78(6):3721–3725

    Article  MathSciNet  Google Scholar 

  • Lande R, Kirkpatrick M (1988) Ecological speciation by sexual selection. J Theor Biol 133(1):85–98

    Article  MathSciNet  Google Scholar 

  • LaSalle JP (1960) Some extensions of liapunov’s second method. Trans Circuit Theory IRE 7(4):520–527

    Article  MathSciNet  Google Scholar 

  • Merrill RM, Wallbank RWR, Bull V, Salazar PCA, Mallet J, Stevens M, Jiggins CD (2012) Disruptive ecological selection on a mating cue. Proc R Soc Lond B Biol Sci 279(1749):4907–4913

    Article  Google Scholar 

  • M’Gonigle LK, Mazzucco R, Otto SP, Dieckmann U (2012) Sexual selection enables long-term coexistence despite ecological equivalence. Nature 484(7395):506–509

    Article  Google Scholar 

  • O’Donald P (1960) Assortive mating in a population in which two alíeles are segregating. Heredity 15:389–396

    Article  Google Scholar 

  • Panhuis TM, Butlin R, Zuk M, Tregenza T (2001) Sexual selection and speciation. Trends Ecol Evol 16(7):364–371

    Article  Google Scholar 

  • Payne R, Krakauer D (1997) Sexual selection, space, and speciation. Evolution 51(1):1–9

    Article  Google Scholar 

  • Rudnicki R, Zwoleński P (2015) Model of phenotypic evolution in hermaphroditic populations. J Math Biol 70(6):1295–1321

    Article  MathSciNet  MATH  Google Scholar 

  • Ruelle D (1989) Elements of differentiable dynamics and bifurcation theory. Academic Press Inc, Boston

    MATH  Google Scholar 

  • Servedio M, Bürger R (2014) The counterintuitive role of sexual selection in species maintenance and speciation. Proc Nat Acad Sci 111(22):8113–8118

    Article  Google Scholar 

  • Servedio M, Bürger R (2015) The effects of sexual selection on trait divergence in a peripheral population with gene flow. Evolution 69(10):2648–2661

    Article  Google Scholar 

  • Servedio MR (2011) Limits to the evolution of assortative mating by female choice under restricted gene flow. Proc R Soc Lond B Biol Sci 278(1703):179–187

    Article  Google Scholar 

  • Smadi C, Leman H, Llaurens V (2018) Looking for the right mate in diploid species: How does genetic dominance affect the spatial differentiation of a sexual trait? J Theor Biol 447:154–170

    Article  MathSciNet  MATH  Google Scholar 

  • Smadja C, Ganem G (2005) Asymmetrical reproductive character displacement in the house mouse. J Evolut Biol 18(6):1485–1493

    Article  Google Scholar 

  • Smadja C, Catalan J, Ganem G (2004) Strong premating divergence in a unimodal hybrid zone between two subspecies of the house mouse. J Evolut Biol 17(1):165–176

    Article  Google Scholar 

  • Uy JAC, Patricelli GL, Borgia G (2001) Complex mate searching in the satin bowerbird ptilonorhynchus violaceus. Am Nat 158(5):530–542

    Article  Google Scholar 

  • Van Doorn G, Noest A, Hogeweg P (1998) Sympatric speciation and extinction driven by environment dependent sexual selection. Proc R Soc Lond B Biol Sci 265(1408):1915–1919

    Article  Google Scholar 

  • Wu CI, Hollocher H, Begun DJ, Aquadro CF, Xu Y, Wu ML (1995) Sexual isolation in drosophila melanogaster: a possible case of incipient speciation. Proc Nat Acad Sci 92(7):2519–2523

    Article  Google Scholar 

Download references

Acknowledgements

I would like to thank Pierre Collet for his help on the theory of dynamical systems. I also thank two anonymous reviewers for their constructive comments, which helped me to improve the manuscript. This work was partially funded by the Chair “Modélisation Mathématique et Biodiversité” of VEOLIA-Ecole Polytechnique-MNHN-F.X.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hélène Leman.

Appendices

Appendix A: Dynamical System Without Migration

In this appendix, we will prove the results of Sect. 3.2, which is related to the case without migration. To this aim, we use the two following weighted quantities

$$\begin{aligned}&\varOmega (t):=(\beta _A-1)z_A(t)-(\beta _a-1)z_a(t), \\&\varSigma (t):=(\beta _A-1)z_A(t)+(\beta _a-1)z_a(t). \end{aligned}$$

From (12), we find that

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\varOmega (t)= & {} \varOmega \left[ b\dfrac{\beta _Az_A+\beta _a z_a}{z_A+z_a}-d-c(z_A+z_a) \right] , \end{aligned}$$
(23)
$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\varSigma (t)= & {} \varSigma \left[ b\dfrac{\beta _Az_A+\beta _a z_a}{z_A+z_a}-d-c(z_A+z_a) \right] \nonumber \\&-\,2b(\beta _A-1)(\beta _a-1)\dfrac{z_az_A}{z_A+z_a}. \end{aligned}$$
(24)

Proof of Lemma 3

We start by studying the stability of equilibrium (0, 0). Assume that \(\varSigma (0)>0\). From (24), we derive

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\varSigma \ge \varSigma \left[ b-d +b\left( \frac{\varSigma }{z_A+z_a}-2(\beta _A-1)(\beta _a-1)\frac{z_az_A}{(z_A+z_a)\varSigma } \right) -c(z_A+z_a)\right] . \end{aligned}$$

Since \(\varSigma ^2-2(\beta _A-1)(\beta _a-1)z_az_A \ge 0\) and \(-(\beta _{\min }-1)(z_A+z_a) \ge -\varSigma \), we deduce that

$$\begin{aligned} \frac{{\hbox {d}}}{{\hbox {d}}t}\varSigma \ge \varSigma \left[ b-d -c\frac{\varSigma }{(\beta _{\min }-1)}\right] . \end{aligned}$$

Hence, as long as \(\varSigma \in ]0,(b-d)(\beta _{\min }-1)/c[\), \(\varSigma (t)\) is increasing. Thus (0, 0) is an unstable equilibrium.

The stability of the three other equilibria, \((\zeta _A,0)\), \((0,\zeta _a)\) and \((\chi _A,\chi _a)\), can be deduce by a direct computation of Jacobian matrices at these points, which we do not detail.

Finally, let us study the long-time behavior of any solution. Equation (23) implies that the sign of \(\varOmega (t)\) is equal at all time and, that \({\mathcal {D}}^A_0\) is a positively invariant set under dynamical system (12). Moreover, there exists only a stable equilibrium that belongs to the set \({\mathcal {D}}^A_0\), which is \((\zeta _A,0)\).

We consider the function \(W: {\mathcal {D}}^A_0 \rightarrow \mathbb {R}\):

$$\begin{aligned} W(z_A,z_a):=\ln \left( \dfrac{\varSigma }{\varOmega } \right) =\ln \left( \dfrac{(\beta _A-1)z_A+(\beta _a-1)z_a}{(\beta _A-1)z_A-(\beta _a-1)z_a}\right) \ge 0 . \end{aligned}$$
(25)

From (23) and (24), we deduce that

$$\begin{aligned} \dfrac{{\hbox {d}}W(z_A(t),z_a(t))}{{\hbox {d}}t}= -2b(\beta _A-1)(\beta _a-1)\dfrac{z_az_A}{(z_A+z_a)\varSigma } \le 0 . \end{aligned}$$

Moreover for any \((z_A,z_a)\in {\mathcal {D}}^A_0\), \(W(z_A,z_a)=0\) if and only if \(z_a=0\). \(W(z_A,z_a)\) converges to \(+\,\infty \) when \((\beta _A-1)z_A-(\beta _a-1)z_a\) converges to 0 and \(\frac{dW}{dt}\) is non-positive on \({\mathcal {D}}^A_0\) and is equal to zero if and only if \(z_a=0\). It ensures that W is a Lyapunov function for (12) on the set \({\mathcal {D}}^A_0\) which cancels only on \({\mathcal {D}}^A_0 \cap \{z_a=0\}\). Furthermore, a simple computation gives that the largest invariant set in \({\mathcal {D}}^A_0 \cap \{z_a=0\}\) is \(\{(\zeta _A,0)\}\). Theorem 1 of LaSalle (1960) is thus sufficient to conclude that any solution of (12) with initial condition in \({\mathcal {D}}^A_0\) converges to \((\zeta _A,0)\) when t tends to \(+\,\infty \). Similarly, we prove that any solution with initial condition in \( {\mathcal {D}}^a_0 \) converges to \((0,\zeta _a)\).

Finally, assume that \(\varOmega (0)=0\). Then, \(\varOmega (t)=0\) for all \(t\ge 0\) according to (23) and, in addition with (12), we derive for all \(\alpha \in \mathcal {A}\),

$$\begin{aligned} \dfrac{{\hbox {d}}}{{\hbox {d}}t}z_\alpha =z_\alpha \left[ b\frac{\beta _A\beta _a-1}{\beta _A+\beta _a-2}-d-c\frac{\beta _A+\beta _a-2}{\beta _{\bar{\alpha }}-1}z_{\alpha } \right] . \end{aligned}$$

We deduce the last point of Lemma 3 easily. \(\square \)

Appendix B: Extinction Time

This subsection is devoted to the proof of Theorem 2 following ideas similar to the ones of the proof of Theorem 3 and Proposition 4.1 in Coron et al. (2018). Hence, we do not give all details, but explain only parts that are different.

Assume that \(m_A\le m_0\), \(m_a\le m_0\) and that \(\mathbf{{Z}}^K(0)\) converges in probability to a deterministic vector \(\mathbf{{z}^0}\) belonging to \({\mathcal {D}}^{A,a}_{m_A,m_a}\), Lemma 3 and Theorem 1 ensure that \((\mathbf{{Z}}^K(t),t\ge 0)\) reaches a neighborhood of the equilibrium \((\zeta _A,0,0,\zeta _a)\) after a finite time independent from K. Indeed, the process dynamics is close to the one of the limiting deterministic system (4).

To prove Theorem 2, it remains to estimate the time before all a-individuals in patch 1 and all A-individuals in patch 2 disappear. We denote it by

$$\begin{aligned} T^K_0=\inf \{t\ge 0, Z^K_{a,1}(t)+Z^K_{A,2}(t)=0\}, \end{aligned}$$
(26)

and we assume that the process is initially close to equilibrium \((\zeta _A,0,0,\zeta _a)\). The estimation is deduced from the following Lemma.

Lemma 4

There exist two positive constants \(\varepsilon _0\) and \(C_0\) such that for any \(\varepsilon \le \varepsilon _0\), if there exists \(\eta \in ]0,1/2[\) that satisfies \(\max (|z_{A,1}^0-\zeta _A|,|z_{a,2}^0-\zeta _a|) \le \varepsilon \) and \(\eta \varepsilon /2 \le z_{a,1}^0,z_{A,2}^0 \le \varepsilon /2\), then

$$\begin{aligned} \begin{aligned}&\text {for all } C>(\omega (A,a))^{-1}+C_0\varepsilon ,&\quad \mathbb {P}(T_0^K\le C \log (K)) \underset{K\rightarrow +\infty }{\rightarrow } 1,\\&\text {for all } 0\le C <(\omega (A,a))^{-1}-C_0\varepsilon ,&\quad \mathbb {P}(T_0^K\le C \log (K)) \underset{K\rightarrow +\infty }{\rightarrow } 0. \end{aligned} \end{aligned}$$

Proof

Following the first step of Proposition 4.1’s proof given by Coron et al. (2018), we prove that as long as the population processes \(Z_{a,1}^K(t)\) and \(Z_{A,2}^K(t)\) have small values, the processes \(Z_{A,1}^K(t)\) and \(Z_{a,2}^K(t)\) stay close to \(\zeta _A\) and \(\zeta _a\) respectively.

Then, by bounding death rates, birth rates and migration rates of \((Z^K_{a,1}(t),t\ge 0)\) and \((Z^K_{A,2}(t),t\ge 0)\), we are able to compare the dynamics of these two processes with the ones of

$$\begin{aligned} \left( \frac{{\mathcal {N}}_a(t)}{K},\frac{{\mathcal {N}}_A(t)}{K},\quad t\ge 0 \right) , \end{aligned}$$

where \(({\mathcal {N}}_a(t),{\mathcal {N}}_A(t)) \in \mathbb {N}^{\{a,A\}}\) is a two types of branching process with types a and A and for which

  • any \(\alpha \)-individual gives birth to a \(\alpha \)-individual at rate b,

  • any \(\alpha \)-individual gives birth to a \(\bar{\alpha }\)-individual at rate \(m_{\bar{\alpha }}\),

  • any \(\alpha \)-individual dies at rate \(b\beta _{\bar{\alpha }}+m_\alpha \).

The goal is thus to estimate the extinction time of such a subcritical two types of branching process. Let M(t) be the mean matrix of the multitype process, that is,

$$\begin{aligned} M(t)= \begin{pmatrix} \mathbb {E}\Big [\mathbb {E}\Big [{\mathcal {N}}_a(t)\Big |({\mathcal {N}}_a(0),{\mathcal {N}}_A(0))=(1,0)\Big ]\Big ] &{} &{} \mathbb {E}\Big [\mathbb {E}\Big [{\mathcal {N}}_A(t)\Big |({\mathcal {N}}_a(0),{\mathcal {N}}_A(0))=(1,0)\Big ]\Big ]\\ &{} &{} \\ \mathbb {E}\Big [\mathbb {E}\Big [{\mathcal {N}}_a(t)\Big |({\mathcal {N}}_a(0),{\mathcal {N}}_A(0))=(0,1)\Big ]\Big ] &{} &{} \mathbb {E}\Big [\mathbb {E}\Big [{\mathcal {N}}_A(t)\Big |({\mathcal {N}}_a(0),{\mathcal {N}}_A(0))=(0,1)\Big ]\Big ] \end{pmatrix}, \end{aligned}$$

and let G be the infinitesimal generator of the semigroup \(\{M(t),t\ge 0\}\). From the book of Athreya and Ney (1972) p.202, we deduce a formula of G which is

$$\begin{aligned} G=\begin{pmatrix} -b(\beta _A-1)-m_a &{}\quad m_A \\ m_a &{}\quad -\,b(\beta _a-1)-m_A \end{pmatrix}. \end{aligned}$$

Applying Theorem 3.1 of Heinzmann (2009), we find that

$$\begin{aligned}&\mathbb {P}\Big (({\mathcal {N}}_a(t),{\mathcal {N}}_A(t))=(0,0) \Big |({\mathcal {N}}_a(0),{\mathcal {N}}_A(0))=(z_{a,1}^0 K,z_{A,2}^0 K)\Big )\nonumber \\&\quad = (1-c_a e^{rt})^{z_{a,1}^0 K}(1-c_A e^{rt})^{z_{A,2}^0 K}, \end{aligned}$$
(27)

where \(c_a, c_A\) are two positive constants and r is the largest eigenvalue of the matrix G. With a simple computation, we find that \(r=-\omega (A,a)\). From (27), we deduce that the extinction time is of order \(\omega (A,a)^{-1}\log K\) when K tends to \(+\,\infty \) by arguing as in step 2 of Proposition 4.1’s proof of Coron et al. (2018). This concludes the proof of Lemma 4. \(\square \)

Finally, this gives all elements to induce Theorem 2.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Leman, H. A Stochastic Model for Reproductive Isolation Under Asymmetrical Mating Preferences. Bull Math Biol 80, 2502–2525 (2018). https://doi.org/10.1007/s11538-018-0479-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-018-0479-2

Keywords

Mathematics Subject Classification

Navigation