This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Selection-recombination-mutation dynamics: Gradient, limit cycle, and closed invariant curve

Suman Chakraborty [email protected] Department of Physics, Indian Institute of Technology Kanpur, Uttar Pradesh 208016, India    Sagar Chakraborty [email protected] Department of Physics, Indian Institute of Technology Kanpur, Uttar Pradesh 208016, India
Abstract

In this paper, the replicator dynamics of the two-locus two-allele system under weak mutation and weak selection is investigated in a generation-wise non-overlapping unstructured population of individuals mating at random. Our main finding is that the dynamics is gradient-like when the point mutations at the two loci are independent. This is in stark contrast to the case of one-locus multi-allele where the existence gradient behaviour is contingent on a specific relationship between the mutation rates. When the mutations are not independent in the two-locus two-allele system, there is the possibility of non-convergent outcomes, like asymptotically stable oscillations, through either the Hopf bifurcation or the Neimark–Sacker bifurcation depending on the strength of the weak selection. The results can be straightforwardly extended for multi-locus two-allele systems.

Replicator dynamics, Recombination, Mutation, Gradient system, Hopf bifurcation
preprint: APS/123-QED

I introduction

How a living being appears and behaves is inherently determined by the variety of genes located on the loci of chromosomes. The entire genotype to phenotype map is very vast and complex, and is further complicated by the environmental feedback effects. While the evolutionary dynamics—e.g., replicator dynamics Taylor and Jonker (1978); Schuster and Sigmund (1983); Cressman and Tao (2014); Mukhopadhyay and Chakraborty (2021)—of simplified one-locus many-allele systems exhibits quite rich dynamical behaviour, the phenomenon of recombination during meiosis presents one with richer extension of the evolutionary dynamics in a population of sexually mating organisms. In the latter, the simplest non-trivial mathematical setup is that of a two-locus two-allele system. While two different loci can determine two different phenotypic traits (e.g., seed colour and seed colour in the revolutionary dihybrid cross experiments of Mendel O’Brien (1993); Mendel and and (1866)), there are plethora of examples where two loci control the same trait and their interaction has considerable effect on the phenotype. Some such examples that quickly comes to mind are comb-types on the head of chicken Bateson et al. (1909), flower colour in peas Dooner et al. (1991), wheat kernel color Carver (2009), blood group in humans Dean and Dean (2005), age-related hearing loss resistance in the Japanese wild-derived inbred MSM/Ms mice Yasuda et al. (2022), and eye colour of humans White and Rabago-Smith (2010).

Haldane Haldane (1931) and Wright Wright (1935) were among the firsts to mathematically explore the evolutionary outcome when selection acts on more than one locus assumed statistically independent, i.e., system is in linkage equilibrium. Later Lewontin and Kojima Lewontin and Kojima (1960) worked on general 2L2A model with both selection and recombination included, and showed that strong epistasis together with linkage disequilibrium can lead to significantly different outcomes. Subsequently, many extensions Bodmer and Felsenstein (1967); Karlin et al. (1970); Feldman and Libermann (1979); Hastings (1985); Ewens (1968) of the model were studied—starting from prediction and calculation of all possible fixed points and determination of stability condition for some special cases in simplified fitness model to the occurrence of cyclic motion. For continuous time model, Akin Akin (1982) and for discrete time model, Hastings Hastings (1981), Hofbauer and Iooss Hofbauer and Iooss (1984) showed that periodic orbits can be a possible outcome when system is in linkage disequilibrium; when the 2L2A model is in linkage equilibrium such complex behaviour can not occur and the system become gradient-like as shown by Nagylaki Nagylaki (1989). Nagylaki also showed that for multi-locus system under weak selection—selection strength sufficiently smaller compared to recombination—linkage disequilibrium decays close to zero within a few generations and the dynamics of the entire multilocus system is governed by the dynamical outcomes of time-continuous multi-locus system Nagylaki (1993); Nagylaki et al. (1999a); Nagylaki (2013a). Very recently Pontz et al. (2018), Pontz and coworkers studied the 2L2A model in detail under weak selection limit and classified all possible equilibrium structure and phase portraits for different payoff structures.

Mutation is the most important and indispensable ingredient of the evolutionary processes. It is important to note that the relationship between mutations at different loci is often complex and can be influenced by a variety of factors. It is possible for a mutation at one locus to affect the mutation rate at another locus on the same chromosome Nowak et al. (2004); Li et al. (2013) directly or indirectly. For example, a mutation in a gene that codes for a repair enzyme involved in maintaining the integrity of DNA could alter the mutation rate of other genes by affecting the efficiency of DNA repair Steele et al. (1998). Similarly, a mutation in a gene that regulates the transcription or translation of other genes can affect the rate of mutations in those genes. Mutation may not be always completely random: It can depend on environmental changes Galhardo et al. (2007) and fitness Baer (2008), and organisms can have evolved mechanisms which can influence the timing or genomic location of mutation Rando and Verstrepen (2007). In view of the above, in this paper, we distinguish between the cases of mutations that occur randomly (independent of what mutations occur at the other loci) from the cases where mutations can be treated to have occurred independently and randomly at a locus—we call the latter independent mutations.

The effects of mutation was studied in one-locus two-allele model for both overlapping and non-overlapping generations, and the results are really interesting: Mutation can stop the extinction of cooperators by producing cyclic or chaotic outcomes and even can produce coexistence regions where both the cooperators and defectors can survive simultaneously Mobilia (2010); Toupo and Strogatz (2015); Mittal et al. (2020); Mukhopadhyay et al. (2021). Evolutionary dynamics were studied for the case of the 2L2A with mutation as well Park and Krug (2011); Qu et al. (2020); Bengtsson and Christiansen (1983); Jain and Seetharaman (2021); Bürger (1989); Christiansen and Frydenberg (1977); Karlin and McGregor (1971). Effect of weak selection in 2L2A model, in the absence of mutation, was studied in detail by Pontz et al. Pontz et al. (2018). If the selection is sufficiently weak then the linkage disequilibrium disappears after a few generations and dynamics become quite straightforward to find its equilibrium structure for the general fitness matrix. The system behaves like a gradient system and if all the fixed point are hyperbolic then for any initial condition dynamics ultimately approaches fixed points: No periodic or chaotic outcome can arise. On the other hand, the effect of both weak mutation and weak selection was studied for one-locus many-allele model by Hofbauer and Sigmund Hofbauer and Sigmund (1998). They have shown that the corresponding system becomes gradient system when the mutation matrix satisfies certain condition: When mutation probability dependents on the target gene only Hofbauer (1985), the system behaves like a gradient system, and no periodic or chaotic orbit is possible. In general, however, the system need not be a gradient system, and hence periodic and chaotic orbits can be possible outcome.

In such contexts, the study of the 2L2A model in presence of weak selection and weak mutation is still, to the best of our knowledge, remains to be reported. Specifically, we are interested in whether the 2L2A model is always gradient-like or there is a specific condition on mutation for that to happen. Does the answer depend on whether the mutations at the two locus are independent? Furthermore, if and when the epistasis induced dynamics in 2L2A system be altered due to presence of mutation is also a question of interest because epistasis or interaction between different loci has important effects on dynamics, like occurrence of cyclic motion Hastings (1981) and sustaining polymorphism Pontz et al. (2018).

Without further ado, we introduce the model in the next section (Sec. II) and then we present the gradient 2L2A system in Sec. III. Subsequently, we discuss the non-gradient dynamics in Sec. III before concluding in Sec. IV.

II The 2L2A Model

We consider a generation-wise non-overlapping, unstructured population of randomly mating diploid individuals. The population is assumed to evolve under the action of viability selection that acts on two diallelic, recombining loci. We, thus, have the standard two-locus two-allele (2L2A) model with viability selection Pontz et al. (2018); Karlin (1975); Karlin and Carmelli (1975); Sacker and von Bremen (2003); Hallgrímsdóttir and Yuster (2008); Lewontin and Feldman (1988); Franklin and Feldman (1977); Bodmer and Felsenstein (1967). Let A1A_{1} and A2A_{2} be the alleles at locus AA, and B1B_{1} and B2B_{2} be the alleles at locus BB. Let G1G_{1}, G2G_{2}, G3G_{3}, and G4G_{4} represent the four possible gametes, viz., A1B1A_{1}B_{1}, A1B2A_{1}B_{2}, A2B1A_{2}B_{1}, and A2B2A_{2}B_{2} respectively. Let the frequency of the GiG_{i} gametes be xix_{i} (obviously, i=14xi=1\sum_{i=1}^{4}x_{i}=1). In these notations, the frequencies of allele A1A_{1} and B1B_{1} are respectively x1+x2x_{1}+x_{2} and x1+x3x_{1}+x_{3} that we henceforth denote as pp and qq respectively. Defining Dx1x4x2x3D\equiv x_{1}x_{4}-x_{2}x_{3}, the measure of linkage disequilibrium, following identities follow:

x1=pq+D,\displaystyle x_{1}=pq+D, (1a)
x2=p(1q)D,\displaystyle x_{2}=p(1-q)-D, (1b)
x3=(1p)qD,\displaystyle x_{3}=(1-p)q-D, (1c)
x4=(1p)(1q)+D.\displaystyle x_{4}=(1-p)(1-q)+D. (1d)

Furthermore, let the fitness of the GiGjG_{i}G_{j} genotype be wijw_{ij}. For simplicity, we assume that the fitness is independent of the fact which gamete is from mother and which one is from father, i.e., wij=wjiw_{ij}=w_{ji}. We also assume that w14=w23w_{14}=w_{23}, the last assumption means that the corresponding fitnesses are the same regardless of whether A1A_{1} and B1B_{1} are on the same chromosome or opposite ones. Consequently, the following matrix suffices for fully representing the fitnesses of genotypes:

A1A1A1A2A2A2[w11B1B1w12B1B2w22B2B2w13w14w24w33w34w44].\begin{array}[]{@{}c@{}}\mbox{\scriptsize$A_{1}A_{1}$}\\ \mbox{\scriptsize$A_{1}A_{2}$}\\ \mbox{\scriptsize$A_{2}A_{2}$}\end{array}\mathop{\left[\begin{array}[]{ *{5}{c} }\displaystyle\smash{\mathop{w_{11}}^{\raisebox{6.0pt}{\scriptsize$B_{1}B_{1}$}}}&\displaystyle\smash{\mathop{w_{12}}^{\raisebox{6.0pt}{\scriptsize$B_{1}B_{2}$}}}&\displaystyle\smash{\mathop{w_{22}}^{\raisebox{6.0pt}{\scriptsize$B_{2}B_{2}$}}}\\ w_{13}&w_{14}&w_{24}\\ w_{33}&w_{34}&w_{44}\end{array}\right].}^{\begin{array}[]{@{}c@{}}\\ \mathstrut\end{array}}

We include two further phenomena in the model: recombination and mutation. While we let rr denote the probability of recombination, the mutation from one gamete to another gamete is specified by the row stochastic matrix 𝖰{\sf Q} whose (i,ji,j)th element QijQ_{ij} is the probability that an offspring with gamete GjG_{j} is born to a parent with gamete GiG_{i}. With this multiplicative Toupo and Strogatz (2015); Hofbauer and Sigmund (1998) mutation that takes place during DNA replication process, the evolution of gamete frequencies is given by the following equation Bengtsson and Christiansen (1983); Pritchett-Ewing (1981a); Christiansen and Frydenberg (1977); Karlin and McGregor (1971):

xi=j[xjwjθjrDw14]Qjiw¯,x^{\prime}_{i}=\frac{\sum_{j}\left[x_{j}w_{j}-\theta_{j}rDw_{14}\right]Q_{ji}}{\bar{w}}, (2)

where wijwijxjw_{i}\equiv\sum_{j}w_{ij}x_{j}, w¯jwjxj\bar{w}\equiv\sum_{j}w_{j}x_{j}, jQij=1\sum_{j}Q_{ij}=1, and θ1=θ2=θ3=θ4=1\theta_{1}=-\theta_{2}=-\theta_{3}=\theta_{4}=1. Here, prime is the tag for immediately succeeding generation. We suppose that the mutation happens at the gametic stage Bengtsson and Christiansen (1983); Bergero et al. (2021).

As elaborated in the introduction to this paper, since our goal is to observe effect of both weak selection and weak mutation on the dynamics of the 2L2A selection-recombination model, we introduce a small parameter ss such that the fitnesses and mutations can be recast as

wij=1+smij,\displaystyle w_{ij}=1+sm_{ij}, (3a)
Qij=sϵijforij.\displaystyle Q_{ij}=s\epsilon_{ij}\ \ \ {\rm for}\,i\neq j. (3b)

Thus, the limit of small ss renders the dynamics of Eq. (2) to be that under weak selection and weak mutation.

It is well-known Hofbauer and Sigmund (1998) that recombination drives 2L2A model to Wright manifold in the absence of selection and mutation: The Wright manifold is a linkage equilibrium manifold (Λ0\Lambda_{0}), where D=0D=0, and a population in linkage equilibrium always remains in equilibrium, i.e. it is an invariant manifold. In the presence of weak selection (sr)(s\ll r) and weak mutation, on using Eq. (3) in Eq. (2), we obtain

xi=xiθirD+𝒪(s).x^{\prime}_{i}=x_{i}-\theta_{i}rD+\mathcal{O}(s). (4)

Interestingly, we find that the effect of mutation at this order is completely absent. Hence, the known results Nagylaki (2013b); Nagylaki et al. (1999b) for the case of the models without mutation (but with selection) holds in our case as well: Close to Λ0\Lambda_{0}, there exists a smooth invariant manifold, Λs\Lambda_{s}, that is globally attracting for Eq. (2); for any initial values the linkage disequilibrium D(t)D(t) becomes 𝒪(s)\mathcal{O}(s) asymptotically in time.

Thus, in the weak selection and weak mutation limit, the linkage disequilibrium D0D\to 0, and consequently, the dynamical equation is further simplified because of the fact that it can now be specified using only two variables pp and qq, for tt1t\geq t_{1}. Specifically, in the set of Eqs. (1), DD is replaced by 𝒪(s)\mathcal{O}(s) terms. Rescaling time t(=0,1,2)t(=0,1,2...) tagging generations as τ=st\tau=st, it is easy to see that as s0s\to 0, Eq. (2) approaches the following differential equations:

p˙\displaystyle\dot{p} =\displaystyle= q(1p)ν1+(1q)(1p)ν2qpν3(1q)pν4\displaystyle{q}{(1-p)}\nu_{1}+{(1-q)}{(1-p)}\nu_{2}-{q}{p}\nu_{3}-{(1-q)}{p}\nu_{4} (5a)
+p(1p)12m¯p,\displaystyle+p(1-p)\frac{1}{2}\frac{\partial\bar{m}}{\partial p},
q˙\displaystyle\dot{q} =\displaystyle= p(1q)ν5+(1p)(1q)ν6pqν7(1p)qν8\displaystyle{p}{(1-q)}\nu_{5}+{(1-p)}{(1-q)}\nu_{6}-{p}{q}\nu_{7}-{(1-p)}{q}\nu_{8} (5b)
+q(1q)12m¯q.\displaystyle+q(1-q)\frac{1}{2}\frac{\partial\bar{m}}{\partial q}.

where, ν1(ϵ31+ϵ32),ν2(ϵ41+ϵ42),ν3(ϵ13+ϵ14),ν4(ϵ23+ϵ24),ν5(ϵ21+ϵ23),ν6(ϵ41+ϵ43),ν7(ϵ12+ϵ14),ν8(ϵ32+ϵ34)\nu_{1}\equiv{(\epsilon_{31}+\epsilon_{32})},~{}\nu_{2}\equiv{(\epsilon_{41}+\epsilon_{42})},~{}\nu_{3}\equiv{(\epsilon_{13}+\epsilon_{14})},~{}\nu_{4}\equiv{(\epsilon_{23}+\epsilon_{24})},~{}\nu_{5}\equiv{(\epsilon_{21}+\epsilon_{23})},~{}\nu_{6}\equiv{(\epsilon_{41}+\epsilon_{43})},~{}\nu_{7}\equiv{(\epsilon_{12}+\epsilon_{14})},~{}\nu_{8}\equiv{(\epsilon_{32}+\epsilon_{34})}, and

m¯m1pq+m2p(1q)+m3(1p)q+m4(1p)(1q)\bar{m}\equiv m_{1}pq+m_{2}p(1-q)+m_{3}(1-p)q+m_{4}(1-p)(1-q) (6)

is the average fitness of the population wherein the marginal fitness of the gamete ii,

mimi1pq+mi2p(1q)+mi3(1p)q+mi4(1p)(1q).m_{i}\equiv m_{i1}pq+m_{i2}p(1-q)+m_{i3}(1-p)q+m_{i4}(1-p)(1-q). (7)

Since, at p=0p=0 and p=1p=1, p˙0\dot{p}\geq 0 and p˙0\dot{p}\leq 0 respectively; and at q=0q=0 and q=1q=1, q˙0\dot{q}\geq 0 and q˙0\dot{q}\leq 0 respectively, Eq. (5) is forward invariant in the pp-qq phase space—a closed unit square. It may also be noted that the dynamical equation remain invariant under addition (but not multiplication) of a constant matrix with the matrix 𝖬{\sf M} whose elements are mijm_{ij}’s. It may be remarked that had the mutation been taken as additive rather than multiplicative Mobilia (2010); Toupo and Strogatz (2015), the form of the final dynamical equation, in the limit of weak selection and weak mutation, remains same as Eqs. (5).

III 2L2A Gradient System

If a dynamical system is gradient-like, then the non-convergent solutions like periodic or chaotic solutions can be ruled out the system. So it is very useful to find the condition on the mutation rates for which the 2L2A model becomes gradient-like.

It is of notational convenience to represent Eqs. (5) in the form:

z˙i=f~i(𝒛)zi(1zi)fi(𝒛),i=1,2,3,4;\displaystyle\dot{z}_{i}=\tilde{f}_{i}({\bm{z}})\equiv z_{i}(1-z_{i})f_{i}({\bm{z}}),\,\,i=1,2,3,4; (8)

here, z1pz_{1}\equiv p, z2qz_{2}\equiv q, z3(1p)z_{3}\equiv(1-p) and z4(1q)z_{4}\equiv(1-q) correspond to the four coordinates in the allele frequency space. The explicit expression of fif_{i}’s is obvious when compared with Eqs. (5). The phase space \mathbb{P}, hence, is Σ2×Σ2{\Sigma}^{2}\times\Sigma^{2} embedded in 4\mathbb{R}^{4}; here, Σ2\Sigma^{2} is one dimensional simplex. We need to consider the interior of \mathbb{P}, int{\rm int}\mathbb{P} (say); at any point 𝒛int{\bm{z}}\in{\rm int}\mathbb{P}, let the tangent space be denoted by 𝕋𝒛\mathbb{T}_{\bm{z}}\mathbb{P}.

Now for 𝒛int{\bm{z}}\in{\rm int}\mathbb{P} and 𝜼𝕋𝒛\bm{\eta}\in\mathbb{T}_{\bm{z}}\mathbb{P}, the inner product

𝒛˙,𝜼𝒛=𝒇~(𝒛),𝜼𝒛=gij(𝒛)f~i(𝒛)ηj\displaystyle\langle\dot{\bm{z}},\bm{\eta}\rangle_{\bm{z}}=\langle\bm{\tilde{f}}(\bm{z}),\bm{\eta}\rangle_{\bm{z}}=g_{ij}(\bm{z})\tilde{f}_{i}(\bm{z})\eta_{j} (9)

where (and henceforth) sum over repeated indices—Einstein’s summation convention—has been imposed. gijg_{ij} is the metric in int{\rm int}\mathbb{P}. Considering the metric gij=δij/[zi(1zi)]g_{ij}={\delta_{ij}}/[z_{i}(1-z_{i})]—reminiscent of the Shahshahani gradient Hofbauer and Sigmund (1998); Shahshahani (1979); Sigmund (1985)—and recalling Eq. (8), it is obvious that

𝒛˙,𝜼𝒛=Uziηi,\displaystyle\langle\dot{\bm{z}},\bm{\eta}\rangle_{\bm{z}}=\frac{\partial U}{\partial z_{i}}\eta_{i}, (10)

if there exists a continuous and differentiable scalar function U(𝒛)U({\bm{z}}) such that fi(𝒛)=U(𝒛)/zif_{i}(\bm{z})={\partial U({\bm{z}})}/{\partial z_{i}}. Actually, since i=14ηi=0\sum_{i=1}^{4}\eta_{i}=0 because 𝜼𝕋𝒛\bm{\eta}\in\mathbb{T}_{\bm{z}}\mathbb{P}, even if a more general condition, viz., fi(𝒛)=U(𝒛)/zi+ϕ(𝒛)f_{i}(\bm{z})={\partial U({\bm{z}})}/{\partial z_{i}}+\phi({\bm{z}}) (ϕ\phi is a scalar function) is satisfied, Eq. (10) holds good (see Appendix A).

In summary, what we have found is that if z˙i=fi(𝒛)\dot{z}_{i}={f}_{i}({\bm{z}}) is gradient system, then so is z˙i=f~i(𝒛)\dot{z}_{i}=\tilde{f}_{i}({\bm{z}}). Consequently, for Eqs. (5) to be gradient system, one requires

f1q=f2p.\frac{\partial f_{1}}{\partial q}=\frac{\partial f_{2}}{\partial p}. (11)

This condition, on putting the expression of fif_{i}’s, ultimately takes the form

1p(ν1ν2)+1(1p)(ν3ν4)=1q(ν5ν6)+1(1q)(ν7ν8).\frac{1}{p}(\nu_{1}-\nu_{2})+\frac{1}{(1-p)}(\nu_{3}-\nu_{4})=\frac{1}{q}(\nu_{5}-\nu_{6})+\frac{1}{(1-q)}(\nu_{7}-\nu_{8}). (12)

As pp and qq are independent variables, the system is gradient system for all possible allowed values of pp and qq if

ν1=ν2,ν3=ν4,ν5=ν6,andν7=ν8.\displaystyle\nu_{1}=\nu_{2},\,\nu_{3}=\nu_{4},\,\nu_{5}=\nu_{6},\,{\rm and}\,\nu_{7}=\nu_{8}. (13)

Evidently, the condition on mutation for 2L2A model to be gradient-like is completely different from the one-locus four-allele model Hofbauer and Sigmund (1998).

In the light of condition (13), Eqs. (5) take the form

p˙=p(1p)12m¯p+ν1(1p)ν3p,\displaystyle\dot{p}=p(1-p)\frac{1}{2}\frac{\partial\bar{m}}{\partial p}+\nu_{1}(1-p)-\nu_{3}p, (14a)
q˙=q(1q)12m¯q+ν5(1q)ν7q.\displaystyle\dot{q}=q(1-q)\frac{1}{2}\frac{\partial\bar{m}}{\partial q}+\nu_{5}(1-q)-\nu_{7}q. (14b)

It is interesting to note that with V(p,q)exp(m¯)p2ν1(1p)2ν3q2ν5(1q)2ν7V(p,q)\equiv\exp(\bar{m})p^{2\nu_{1}}(1-p)^{2\nu_{3}}q^{2\nu_{5}}(1-q)^{2\nu_{7}} , Eq. (14) can be recast as:

p˙=p(1p)2VVp,\displaystyle\dot{p}=\frac{p(1-p)}{2V}\frac{\partial V}{\partial p}, (15a)
q˙=q(1q)2VVq;\displaystyle\dot{q}=\frac{q(1-q)}{2V}\frac{\partial V}{\partial q}; (15b)

the interior fixed points (p,q)(p^{*},q^{*}) correspond to V/p=V/q=0{\partial V}/{\partial p}={\partial V}/{\partial q}=0. In fact, it can be verified (see Appendix B) that V(p,q)V(p,q) is local Lyapunov function for the gradient system (14) because

V˙=2Vp(1p)p˙2+2Vq(1q)q˙2,\displaystyle\dot{V}=\frac{2V}{p(1-p)}\dot{p}^{2}+\frac{2V}{q(1-q)}\dot{q}^{2}, (16)

implying V˙=0\dot{V}=0 only at any (p,q)(0,1)2(p^{*},q^{*})\in(0,1)^{2} and V˙>0\dot{V}>0 always for all possible (p,q)(0,1)2(p,q)\in(0,1)^{2} other than the fixed points.

The existence of the Lyapunov function leads to the conclusion for Eq. (2): For suffciently small ss and if all the equilibria of Eqs. (15) are hyperbolic, every solution of Eq. (2) converges to a fixed point. This is essentially as extension of a theorem due to Nagyalaki Nagylaki et al. (1999a) and can be succinctly understood as follows. Substituting wij=1+smijw_{ij}=1+sm_{ij} and Qij=sϵijQ_{ij}=s\epsilon_{ij} for all iji\neq j in equation (2), and assuming linkage equilibrium (D=0D=0) and condition (13), we arrive at

Δp=sw¯[p(1p)2m¯p+ν1(1p)ν3p],\displaystyle\Delta p=\frac{s}{\bar{w}}\left[\frac{p(1-p)}{2}\frac{\partial\bar{m}}{\partial p}+\nu_{1}(1-p)-\nu_{3}p\right], (17a)
Δq=sw¯[q(1q)2m¯q+ν5(1q)ν7q],\displaystyle\Delta q=\frac{s}{\bar{w}}\left[\frac{q(1-q)}{2}\frac{\partial\bar{m}}{\partial q}+\nu_{5}(1-q)-\nu_{7}q\right], (17b)

which reduce to Eqs. (15) as s0s\rightarrow 0. Here, w¯=1+sm¯\bar{w}=1+s\bar{m}. It is straightforward to note that the fixed points of Eqs. (15) and Eqs. (17), and their stability properties for small enough ss are same. This means that if the fixed points of Eqs. (15) are hyperbolic, then due to the system being a gradient system, the initial conditions will be attracted to some stable fixed points; and so is going to be the fate of the initial conditions under the map given by Eq. (2) whose corresponding fixed points are in Λs\Lambda_{s} manifold.

The search for which kind of replicator-mutator 2L2A system can be a gradient system interestingly stops at the most simple case of random point mutation, happening independent of what is happening elsewhere (e.g., at another locus): We call such mutations independent. It can be expressed in our setup as follows: Let the probability of mutation from AiA_{i} allele to AjA_{j} allele be given by μijA\mu_{ij}^{A} and the probability of mutation from BiB_{i} allele to BjB_{j} allele is given by μijB\mu_{ij}^{B}. Naturally μiiA\mu_{ii}^{A} and μiiB\mu_{ii}^{B} correspond to the probabilities of no mutation from AiA_{i} and BiB_{i} alleles respectively. So the corresponding mutation matrix 𝖰{\sf Q} representing the probability of mutation from one gamete to another gamete, when the mutation at the two locus are independent, can be represented by the following matrix:

A1B1A1B2A2B1A2B2[μ11Aμ11BA1B1μ11Aμ12BA1B2μ12Aμ11BA2B1μ12Aμ12BA2B2μ11Aμ21Bμ11Aμ22Bμ12Aμ21Bμ12Aμ22Bμ21Aμ11Bμ21Aμ12Bμ22Aμ11Bμ22Aμ12Bμ21Aμ21Bμ21Aμ22Bμ22Aμ21Bμ22Aμ22B].\begin{array}[]{@{}c@{}}\mbox{\scriptsize$A_{1}B_{1}$}\\ \mbox{\scriptsize$A_{1}B_{2}$}\\ \mbox{\scriptsize$A_{2}B_{1}$}\\ \mbox{\scriptsize$A_{2}B_{2}$}\end{array}\mathop{\left[\begin{array}[]{ *{5}{c} }\displaystyle\smash{\mathop{\mu_{11}^{A}\mu_{11}^{B}}^{\raisebox{6.0pt}{\scriptsize$A_{1}B_{1}$}}}&\displaystyle\smash{\mathop{\mu_{11}^{A}\mu_{12}^{B}}^{\raisebox{6.0pt}{\scriptsize$A_{1}B_{2}$}}}&\displaystyle\smash{\mathop{\mu_{12}^{A}\mu_{11}^{B}}^{\raisebox{6.0pt}{\scriptsize$A_{2}B_{1}$}}}&\displaystyle\smash{\mathop{\mu_{12}^{A}\mu_{12}^{B}}^{\raisebox{6.0pt}{\scriptsize$A_{2}B_{2}$}}}\\ \mu_{11}^{A}\mu_{21}^{B}&\mu_{11}^{A}\mu_{22}^{B}&\mu_{12}^{A}\mu_{21}^{B}&\mu_{12}^{A}\mu_{22}^{B}\\ \mu_{21}^{A}\mu_{11}^{B}&\mu_{21}^{A}\mu_{12}^{B}&\mu_{22}^{A}\mu_{11}^{B}&\mu_{22}^{A}\mu_{12}^{B}\\ \mu_{21}^{A}\mu_{21}^{B}&\mu_{21}^{A}\mu_{22}^{B}&\mu_{22}^{A}\mu_{21}^{B}&\mu_{22}^{A}\mu_{22}^{B}\end{array}\right]}^{\begin{array}[]{@{}c@{}}\\ \mathstrut\end{array}}.

Now note that condition (13) for the gradient system in more convenient notions is

ν1=ν2ϵ31+ϵ32=ϵ41+ϵ42,\displaystyle\nu_{1}=\nu_{2}\implies\epsilon_{31}+\epsilon_{32}=\epsilon_{41}+\epsilon_{42}, (18a)
ν3=ν4ϵ13+ϵ14=ϵ23+ϵ24,\displaystyle\nu_{3}=\nu_{4}\implies\epsilon_{13}+\epsilon_{14}=\epsilon_{23}+\epsilon_{24}, (18b)
ν5=ν6ϵ21+ϵ23=ϵ41+ϵ43,\displaystyle\nu_{5}=\nu_{6}\implies\epsilon_{21}+\epsilon_{23}=\epsilon_{41}+\epsilon_{43}, (18c)
ν7=ν8ϵ12+ϵ14=ϵ32+ϵ34.\displaystyle\nu_{7}=\nu_{8}\implies\epsilon_{12}+\epsilon_{14}=\epsilon_{32}+\epsilon_{34}. (18d)

This is trivially satisfied by the aforementioned mutation matrix for independent mutation because of the fact that j=12μijA=j=12μijB=1\sum_{j=1}^{2}\mu^{A}_{ij}=\sum_{j=1}^{2}\mu^{B}_{ij}=1—the normalization condition of probability. In conclusion, in the presence of independent point mutations at the two different loci, the 2L2A system always behaves as a gradient system in the weak selection limit and no complicated dynamics like oscillatory or chaotic dynamics can appear. It may be noted, however, that converse of this result need not be true: Mathematically speaking, mutations that are not independent (mentioned in Sec. I) may or may not satisfy the gradient condition.

IV 2L2A Non-gradient system

Now let us see some of the example where non-convergent dynamical outcomes are possible in 2L2A model in the presence of mutation in the weak selection limit. As is clear from the preceding discussions, we have to consider non-independent mutations as they can result in non-gradient dynamics, and hence oscillatory outcomes.

Let us consider a simple fitness matrix well studied in literature Lewontin and Kojima (1960); Hastings (1981):

𝖬=[𝗄𝖫𝟣𝗄𝖫𝟤𝖪𝖫𝟤𝗄𝖫𝟣𝗄].\sf{M}=\begin{bmatrix}k&L_{1}&k\\ L_{2}&K&L_{2}\\ k&L_{1}&k\end{bmatrix}. (19)

It basically says that a genotype homozygous at both locus has the fitness kk, a genotype homozygous at AA (BB) locus and heterozygous at BB (AA) locus has the fitness L1L_{1} (L2L_{2}), and a genotype heterozygous at both locus has the fitness KK. Our aim is to illustrate the non-fixed point type dynamics that can appear in such systems in the presence of mutation.

To this end, let us take a mutation matrix that does not satisfies condition (13), and hence the corresponding dynamics given by Eqs. (5) is not a gradient system. Specifically, let us take ϵ31=ϵ24=ϵ1\epsilon_{31}=\epsilon_{24}=\epsilon_{1}, ϵ12=ϵ43=ϵ2\epsilon_{12}=\epsilon_{43}=\epsilon_{2}, and other mutation probabilities as zero. Furthermore, in order to achieve analytical tractability, let us choose ν1+ν2ν3ν4=0\nu_{1}+\nu_{2}-\nu_{3}-\nu_{4}=0 and ν5+ν6ν7ν8=0\nu_{5}+\nu_{6}-\nu_{7}-\nu_{8}=0 (note that our choice of 𝖰{\sf Q} satisfies these) so that the internal fixed point (p,q)(p^{*},q^{*}) is conveniently located at (1/2,1/2)(1/2,1/2).

While performing linear stability analysis of Eqs. (5) about fixed point (1/2,1/2)(1/2,1/2), the Jacobian comes out to be

J=[αβγδ],J=\begin{bmatrix}\alpha&\beta\\ \gamma&\delta\\ \end{bmatrix}, (20)

where α=(kK)/4(ν1+ν2+ν3+ν4)/2,β=(ν1ν2ν3+ν4)/2,γ=(ν5ν6ν7+ν8)/2\alpha={(k-K)}/{4}-(\nu_{1}+\nu_{2}+\nu_{3}+\nu_{4})/{2},\beta=(\nu_{1}-\nu_{2}-\nu_{3}+\nu_{4})/2,\gamma=(\nu_{5}-\nu_{6}-\nu_{7}+\nu_{8})/2, and δ=(kK)/4(ν5+ν6+ν7+ν8)/2\delta={(k-K)}/{4}-(\nu_{5}+\nu_{6}+\nu_{7}+\nu_{8})/2; and the corresponding eigenvalues are

λ=(α+δ)±(αδ)2+4βγ2.\lambda=\frac{(\alpha+\delta)\pm\sqrt{(\alpha-\delta)^{2}+4\beta\gamma}}{2}\,. (21)

For the Hopf bifurcation to occur one must have 4βγ<(αδ)24\beta\gamma<-(\alpha-\delta)^{2} and it happens at α+δ=0\alpha+\delta=0 which can be recast to read (kK)=i=18νi(k-K)=\sum_{i=1}^{8}\nu_{i}. For illustrative purpose, if we now fix ϵ2=0.2\epsilon_{2}=0.2 and (kK)=1(k-K)=1, and work with ϵ1\epsilon_{1} as a variable parameter then at ϵ1=0.3\epsilon_{1}=0.3, the Hopf bifurcation should occurs (see Fig. 1).

We observe that in the setting we are working, the condition of the Hopf bifurcation is independent of any constraints on L1L_{1} and L2L_{2}. This has an interesting implication that we now point out: The case K>L1,L2K>L_{1},L_{2} and k>L1,L2k>L_{1},L_{2} corresponds to strong epistasis in literature Hastings (1981) where it was shown that for oscillatory behaviour to appear in 2L2A model (without mutation), strong epistasis is a necessary condition. However, when non-independent mutation is in action, the limit cycle can appear even in the absence of the strong epistasis because the corresponding Hopf bifurcation condition is dependent only on the values of kk and KK.

Refer to caption
Figure 1: (Colour online) Mutation drives oscillatory outcomes in non-gradient 2L2A system: As discussed in Sec. IV, for illustrative purpose, we take ν1=ν4=0.3,ν6=ν7=0.2\nu_{1}=\nu_{4}=0.3,\nu_{6}=\nu_{7}=0.2, ν2=ν3=ν5=ν8=0.0\nu_{2}=\nu_{3}=\nu_{5}=\nu_{8}=0.0, and kK=1.0k-K=1.0. The Hopf bifurcation and the Neimark–Sacker bifurcation (in the time discrete version) occur at ϵ=0.3\epsilon=0.3. Subplots (a) and (b) depict the appearance of limit cycle (blue closed curve) via Hopf bifurcation as ϵ1\epsilon_{1} decreases. whereas subplots (c) and (d) exhibit birth of closed invariant cycle (blue closed curve) via Neimark–Sacker bifurcation (we take s=0.01s=0.01). The red and the green trajectories depict evolution of two separate initial conditions. (In the lower panel, the red and the green curves are aid for eyes to follow the discrete orbit.)

In this context, we recall that the 2L2A gradient system closely approximates the corresponding discrete-time dynamics of 2L2A model in the weak selection limit: Each initial condition approaches a stable fixed point. Somewhat analogous effect is seen corresponding to the limit cycles in discrete-time 2L2A non-gradient system: For the same parameter values and sufficiently small ss, an invariant closed curve is generated. However, in general, the orbits on the invariant curve need not necessarily be periodic with finite number of periodic points as another possibility is that the orbit on the curve can be everywhere dense (may be due to quasiperiodcity or chaos). As long as ss is sufficiently small (we can ignore 𝒪(s2)\mathcal{O}(s^{2}) and higher order terms), it is easily seen that the time-discrete equation (which is effectively the Euler forward discretization of the continuous-time equation) under the same parameter values chosen above (while discussing the Hopf bifurcation), undergoes the Neimark–Sacker bifurcation. We do not present the calculations for the sake of avoiding trivial repetition. A numerical evidence of this is showcased in Fig. 1.

V Conclusion

In conclusion, we have mathematically investigated the two-locus two-allele selection-recombination-mutation dynamics through the replicator equations in the limit of weak selection and weak mutation. We note that when the selection and mutation are sufficiently weak, one can ignore linkage disequilibrium completely and the system becomes two dimensional continuous time differential equation which faithfully approximates the evolutionary dynamics when the dynamics is gradient-like. Our search for the cases of mutation for which the dynamics is gradient-like has led to the result that whenever the point mutation rates at one locus is not affected by what happens at the other, the dynamics is always gradient-like. When the dynamics is not gradient-like, we found that stable limit cycle or stable closed invariant curve can appear. The dynamics on the invariant curve can, in principle, sustain non-convergent solutions like periodic, quasiperiodic, and chaotic orbits. We have also discussed how the dynamical manifestation of epistasis can completely changed in presence of mutation.

We are tempted to claim that our results imply that, within the paradigm of replicator dynamics of evolutionary process in the weak selection and mutation limit, an observation of oscillatory gametic or allelic frequencies is indirect indication of the existence of mutations that are not independent. An immediate future work that builds on the present paper is to find how the aforementioned results manifest themselves in a finite population where stochastic effects can no longer be ignored. Furthermore, many characteristics like our body size Reed et al. (2003), height Lettre et al. (2008), and the shape of different organs like the ear Adhikari et al. (2015) are controlled by more than two loci of a chromosome. So understanding the evolutionary dynamical outcome of multi-locus systems Zhivotovsky and Pylkov (1998); Wray and Goddard (2010), in the presence of mutation, is still another natural direction to investigate. While our main result—occurrence of gradient system in the presence of independent mutation—is valid for multi-locus two-allele systems (see Appendix A), the condition for general multi-locus multi-allele systems’ becoming gradient-like remains an open problem.

Acknowledgements.
The authors thank Yash Varshney for discussions during the initial stages of this work. Sagar Chakraborty acknowledges the support from SERB (DST, govt. of India) through project no. MTR/2021/000119.

Appendix A Multi-locus two-allele gradient system

Let there be a total n+1n+1 loci and at llth (l=0,1,2,,nl=0,1,2,\cdots,n) locus two possible alleles be denoted by A0(l)A_{0}^{(l)} and A1(l)A_{1}^{(l)}. The gametes—each represented by a sequence, Aς0(0)Aς1(1)Aς2(2)Aς3(3)Aςj(j)Aςn(n)A_{\varsigma_{0}}^{(0)}A_{\varsigma_{1}}^{(1)}A_{\varsigma_{2}}^{(2)}A_{\varsigma_{3}}^{(3)}\cdots A_{\varsigma_{j}}^{(j)}\cdots A_{\varsigma_{n}}^{(n)}, where ςj{0,1}\varsigma_{j}\in\{0,1\} j\forall j—are mathematically sorted in a fashion such that iith gamete (GiG_{i}) is listed at position i=(ς0×20)+(ς1×21)+(ς2×22)++(ςn×2n)i=(\varsigma_{0}\times 2^{0})+(\varsigma_{1}\times 2^{1})+(\varsigma_{2}\times 2^{2})+\cdots+(\varsigma_{n}\times 2^{n}), thereby helping us to directly associate a gamete number with a gamete. Let the frequency of GiG_{i} be xix_{i}.

Furthermore, let the fitness of the GiGjG_{i}G_{j} type of genotype be wijw_{ij}. Assuming that the fitness is independent of which gamete is contributed by which parent, the equality, wij=wjiw_{ij}=w_{ji}, follows. The fitness of the iith gamete and average fitness of the population are, thus, given by wijwijxjw_{i}\equiv\sum_{j}w_{ij}x_{j} and w¯jwjxj\bar{w}\equiv\sum_{j}w_{j}x_{j}, respectively.

Now for multi-locus system the selection-recombination evolutionary equation Nagylaki et al. (1999a); Nagylaki (1993) for the gametes’ frequency in the presence of multiplicative mutation which occur during gametic stage Pritchett-Ewing (1981b), takes the form

xi=1w¯j=02n1(xjwjDj)Qji,x_{i}^{\prime}=\frac{1}{\bar{w}}\sum_{j=0}^{2^{n}-1}(x_{j}w_{j}-D_{j})Q_{ji}, (22)

where QjiQ_{ji} corresponds to the probability of mutation from the jjth gamete to the iith gamete. The symbol DiD_{i} is understood as follows:

Let us decompose the set of loci l={0,1,2,,n}l=\{0,1,2,\cdots,n\} into two disjoint sets, II and JJ. Let the probability of recombination with the loci in II inherited from one parent with the loci in JJ inherited from the other parent be rIr_{I}. Now DijIrI(wijxixjwiIjJ,jIiJxiIjJxjIiJ)D_{i}\equiv\sum_{j}\sum_{I}r_{I}(w_{ij}x_{i}x_{j}-w_{i_{I}j_{J},j_{I}i_{J}}x_{i_{I}j_{J}}x_{j_{I}i_{J}}) represents the measure of linkage disequilibrium for iith gamete. xiIjJx_{i_{I}j_{J}} denotes the frequency of gametes consisting of the genes iIi_{I} located in II and the genes jJj_{J} located in JJ inherited from two different parents, where iIi_{I} (or jJj_{J}) corresponds to a vector with component ili_{l} (or jlj_{l}) for every lIl\in I (or lJl\in J). In passing, note that, as it should DiD_{i} ultimately boils down to θiw14rD\theta_{i}w_{14}rD for two-locus two-allele scenario, when w14=w23w_{14}=w_{23}, as used in the main text. In consistent notation, wiIjJ,jIiJw_{i_{I}j_{J},j_{I}i_{J}} is the fitness of the genotype formed by the two gametes—one whose frequency is xiIjJx_{i_{I}j_{J}} and another one whose frequency is xjIiJx_{j_{I}i_{J}}.

Here again we are interested in evolutionary outcome under weak selection and weak mutation. So let us take the form of the fitness and mutation as we have taken for two-locus case (3). Under weak mutation, à la Nagylaki Nagylaki (1993), we can show that under weak selection linkage disequilibrium in Eq. (22) goes to quasi-linkage equilibrium within a few generations (t>t1t>t_{1}). So we can write Eq. (22) after t>t1t>t_{1} as

xi=1w¯j=02(n+1)1Qjixjwj.x^{\prime}_{i}=\frac{1}{\bar{w}}\sum_{j=0}^{2^{(n+1)}-1}Q_{ji}x_{j}w_{j}. (23)

Rescaling time t(=0,1,2)t(=0,1,2...) of generations as τ=st\tau=st and taking the limit s0s\to 0, we ultimately come-up with the following differential equation Hofbauer and Sigmund (1998):

x˙i=xi[mim¯]+j=02(n+1)1(ϵjixjϵijxi).\dot{x}_{i}=x_{i}\left[m_{i}-\bar{m}\right]+\sum_{j=0}^{2^{(n+1)}-1}(\epsilon_{ji}x_{j}-\epsilon_{ij}x_{i}). (24)

We now want to require this equation in terms of the frequency of the alleles.

We denote the frequency of the allele A0(l)A_{0}^{(l)}, i.e., frequency of the 0th allele at llth locus by zlz_{l} and the frequency of the 11st allele at the llth locus by zn+lz_{n+l}. Since each locus contains two alleles, it is obvious that zn+l=1zlz_{n+l}=1-z_{l}. At the linkage equilibrium manifold, the gamete frequency is given by just the product of the corresponding allele frequencies Nagylaki et al. (1999a),

xi=l=0nzl+ςlnx_{i}=\prod_{l=0}^{n}z_{l+\varsigma_{l}n} (25)

where i=Σl=0nςl2li=\Sigma_{l=0}^{n}\varsigma_{l}2^{l}, in our notations. zlz_{l} can be easily calculated by taking the sum over all possible gametes in which the llth locus contains the 0th allele. Again, by virtue of the numbering convention of the gametes introduced in the beginning of this section, one can write the allele frequency in terms of gamete frequency explicitly as follows:

zl=x0+x1++x2l1+x2.(2l)++x3.(2l)1+x4.(2l)+.z_{l}=x_{0}+x_{1}+\cdots+x_{2^{l}-1}+x_{2.{(2^{l})}}+\cdots+x_{3.{(2^{l})}-1}+x_{4.{(2^{l})}}+\cdots. (26)

Taking the time derivative of Eq. (26) and using Eq. (24), we ultimately get the evolution equation for allele frequency:

z˙lf~l(𝒛)zl(1zl)[12m¯zl+νl0x0zl(1zl)+νl1x1zl(1zl)++νl2nx2nzl(1zl)]zl(1zl)fl,\dot{z}_{l}\equiv\tilde{f}_{l}(\bm{z})\equiv z_{l}(1-z_{l})\left[\frac{1}{2}\frac{\partial\bar{m}}{\partial z_{l}}+\nu_{{l}_{0}}\frac{x_{0}}{z_{l}(1-z_{l})}+\nu_{{l}_{1}}\frac{x_{1}}{z_{l}(1-z_{l})}+...+\nu_{{l}_{2^{n}}}\frac{x_{2^{n}}}{z_{l}(1-z_{l})}\right]\equiv z_{l}(1-z_{l})f_{l}, (27)

where, m¯=1+sw¯\bar{m}=1+s\bar{w},
and     νl0[ϵ0,(2l)++ϵ0,2(2l)1+ϵ0,3(2l)++ϵ0,4(2l)1+]\nu_{l_{0}}\equiv-\left[\epsilon_{0,(2^{l})}+\cdots+\epsilon_{0,2(2^{l})-1}+\epsilon_{0,3(2^{l})}+\cdots+\epsilon_{0,4(2^{l})-1}+\cdots\right],
          νl1[ϵ1,(2l)++ϵ1,2(2l)1+ϵ1,3(2l)++ϵ1,4(2l)1+]\nu_{l_{1}}\equiv-\left[\epsilon_{1,(2^{l})}+\cdots+\epsilon_{1,2(2^{l})-1}+\epsilon_{1,3(2^{l})}+\cdots+\epsilon_{1,4(2^{l})-1}+\cdots\right],
          \cdots\equiv\cdots
     νl(2l1)[ϵ(2l1),(2l)++ϵ(2l1),2(2l)1+ϵ(2l1),3(2l)++ϵ(2l1),4(2l1)+],\nu_{l_{(2^{l}-1)}}\equiv-\left[\epsilon_{{(2^{l}-1)},(2^{l})}+\cdots+\epsilon_{{(2^{l}-1)},2(2^{l})-1}+\epsilon_{{(2^{l}-1)},3(2^{l})}+\cdots+\epsilon_{{(2^{l}-1)},4(2^{l}-1)}+\cdots\right],
       νl(2l)+[ϵ(2l),0++ϵ(2l),(2l)1+ϵ(2l),2(2l)++ϵ(2l),3(2l)1+]\nu_{l_{(2^{l})}}\equiv+\left[\epsilon_{(2^{l}),0}+\cdots+\epsilon_{(2^{l}),(2^{l})-1}+\epsilon_{(2^{l}),2(2^{l})}+\cdots+\epsilon_{(2^{l}),3(2^{l})-1}+\cdots\right],
          \cdots\equiv\cdots Here we have put a comma in the subscript of ϵij\epsilon_{ij} merely for the visual clarity in reading the subscripts.

We know the Shahshahani metric in the gametic space (a simplex of dimension 2n+112^{n+1}-1) is gij(x)=δij/xig_{ij}(x)={\delta_{ij}}/{x_{i}}. Now we are interested to calculate the corresponding metric (g¯ij\bar{g}_{ij}, say) in the allelic space, (z1,z2,,z2n)(z_{1},z_{2},\cdots,z_{2n}). Transformation of the coordinates yields:

g¯ij(z)=xρzixσzjgρσ(x).\bar{g}_{ij}(z)=\frac{\partial x^{\rho}}{\partial z_{i}}\frac{\partial x^{\sigma}}{\partial z_{j}}g_{\rho\sigma}(x). (28)

Evidently, for iji\neq j, the metric elements g¯ij(z)\bar{g}_{ij}(z) is zero and for i=ji=j the metric is non-zero such that the metric can eventually be written as

g¯ij(z)=δijzi(1zi).\displaystyle\bar{g}_{ij}(z)=\frac{\delta_{ij}}{z_{i}(1-z_{i})}. (29a)

Recall that the allelic phase space \mathbb{P} is (n+1)(n+1)-dimensional space, Σ2×Σ2××Σ2{\Sigma}^{2}\times\Sigma^{2}\times\cdots\times\Sigma^{2}, embedded in 2(n+1)\mathbb{R}^{2(n+1)}; here, Σ2\Sigma^{2} is one dimensional simplex.

We need to consider the interior of \mathbb{P}, int{\rm int}\mathbb{P} (say); at any point 𝒛int{\bm{z}}\in{\rm int}\mathbb{P}, let the tangent space be denoted by 𝕋𝒛\mathbb{T}_{\bm{z}}\mathbb{P}. Now for 𝒛int{\bm{z}}\in{\rm int}\mathbb{P} and 𝜼𝕋𝒛\bm{\eta}\in\mathbb{T}_{\bm{z}}\mathbb{P}, the inner product

𝒛˙,𝜼𝒛=𝒇~(𝒛),𝜼𝒛=g¯ij(𝒛)f~i(𝒛)ηj.\displaystyle\langle\dot{\bm{z}},\bm{\eta}\rangle_{\bm{z}}=\langle\bm{\tilde{f}}(\bm{z}),\bm{\eta}\rangle_{\bm{z}}=\bar{g}_{ij}(\bm{z})\tilde{f}_{i}(\bm{z})\eta_{j}. (30)

Considering the metric g¯ij\bar{g}_{ij}, it is obvious that

𝒛˙,𝜼𝒛=Uziηi,\displaystyle\langle\dot{\bm{z}},\bm{\eta}\rangle_{\bm{z}}=\frac{\partial U}{\partial z_{i}}\eta_{i}, (31)

if there exists a continuous and differentiable scalar functions U(𝒛)U({\bm{z}}) and ϕ(𝒛)\phi(\bm{z}) such that fi(𝒛)=U(𝒛)/zi+ϕ(𝒛)f_{i}(\bm{z})={\partial U({\bm{z}})}/{\partial z_{i}}+\phi(\bm{z}). In other words, this form of fif_{i} is the condition for 𝒛˙=𝒇~\dot{\bm{z}}=\tilde{\bm{f}} to be a gradient system.

Now, consider the bilinear form 𝑯𝒛𝒇~\bm{H_{z}\tilde{f}}:

𝑯𝒛𝒇~(𝝃,𝜼)i,j=12n1zi(1zi)f~izjξiηj,\bm{H_{z}\tilde{f}}{(\bm{\xi},\bm{\eta})}\equiv\sum_{i,j=1}^{2n}\frac{1}{z_{i}(1-z_{i})}\frac{\partial\tilde{f}_{i}}{\partial z_{j}}\xi_{i}\eta_{j}, (32)

where 𝝃,𝜼𝕋𝒛\bm{\xi,\eta}\in\mathbb{T}_{\bm{z}}\mathbb{P}. We find that

f~izj=δij(12zi)fi+zi(1zi)fizj.\frac{\partial\tilde{f}_{i}}{\partial z_{j}}=\delta_{ij}(1-2z_{i})f_{i}+z_{i}(1-z_{i})\frac{\partial f_{i}}{\partial z_{j}}. (33)

Putting this relation in Eq.  (32) and imposing fi(𝒛)=U(𝒛)/zi+ϕ(𝒛)f_{i}(\bm{z})={\partial U({\bm{z}})}/{\partial z_{i}}+\phi(\bm{z}), we find that 𝑯𝒛𝒇~(𝝃,𝜼)\bm{H_{z}\tilde{f}}{(\bm{\xi},\bm{\eta})} is symmetric which, in turn, implies

i,j=12n(fizjfjzi)ξiηj=0\sum_{i,j=1}^{2n}\left(\frac{\partial f_{i}}{\partial z_{j}}-\frac{\partial f_{j}}{\partial z_{i}}\right)\xi_{i}\eta_{j}=0 (34)

for all possible tangent vectors 𝝃,𝜼𝕋𝒛\bm{\xi,\eta}\in\mathbb{T}_{\bm{z}}\mathbb{P}. With 𝝃\bm{\xi}=𝐞𝐢𝐞𝐤\mathbf{e_{i}-e_{k}} and 𝜼\bm{\eta}=𝐞𝐣𝐞𝐤\mathbf{e_{j}-e_{k}} (𝐞{\bf e}’s are the unit basis vectors of 2(n+1)\mathbb{R}^{2(n+1)}), we obtain

fizj+fjzk+fkzi=fizk+fkzj+fjzi.\frac{\partial f_{i}}{\partial z_{j}}+\frac{\partial f_{j}}{\partial z_{k}}+\frac{\partial f_{k}}{\partial z_{i}}=\frac{\partial f_{i}}{\partial z_{k}}+\frac{\partial f_{k}}{\partial z_{j}}+\frac{\partial f_{j}}{\partial z_{i}}. (35)

As an aside, one notes that this general condition for 2L2A reduces to the simpler condition, f0/z1=f1/z0\partial f_{0}/\partial z_{1}=\partial f_{1}/\partial z_{0} as observed in the main text making the presence of ϕ(𝒛)\phi(\bm{z}) redundant in the condition.

We now show that just like the fact that independent point mutation renders the 2L2A model a gradient system, multi-locus two-allele model in presence of independent point mutation is also a gradient system. To this end, it suffices to ignore ϕ(𝒛)\phi(\bm{z}), meaning we have as a special case of Eq. (35):

fizj=fjzi,j{0,1,2,,n},\frac{\partial f_{i}}{\partial z_{j}}=\frac{\partial f_{j}}{\partial z_{i}},\quad\forall j\in\{0,1,2,\cdots,n\}, (36)

at all values of 𝒛{\bm{z}}. Thus, we explicitly have

fizj=[(νi2jνi0)(x0+x2j)+(νi2j+1νi1)(x0+x2j+1)++(νi2iνi2j+2i)(x2i+x2j+2i)+]1zi(1zi)\displaystyle\frac{\partial f_{i}}{\partial z_{j}}=\left[(\nu_{i_{2^{j}}}-\nu_{i_{0}})(x_{0}+x_{2^{j}})+(\nu_{i_{2^{j}+1}}-\nu_{i_{1}})(x_{0}+x_{2^{j}+1})+\cdots+(\nu_{i_{2^{i}}}-\nu_{i_{2^{j}+2^{i}}})(x_{2^{i}}+x_{2^{j}+2^{i}})+\cdots\right]\frac{1}{z_{i}(1-z_{i})}
+122m¯zjzi,\displaystyle+\frac{1}{2}\frac{\partial^{2}\bar{m}}{\partial z_{j}\partial z_{i}}, (37a)
fjzi=[(νj2iνj0)(x0+x2i)+(νj2i+1νj1)(x0+x2i+1)++(νj2jνj2i+2j)(x2j+x2i+2j)+]1zj(1zj)\displaystyle\frac{\partial f_{j}}{\partial z_{i}}=\left[(\nu_{j_{2^{i}}}-\nu_{j_{0}})(x_{0}+x_{2^{i}})+(\nu_{j_{2^{i}+1}}-\nu_{j_{1}})(x_{0}+x_{2^{i}+1})+\cdots+(\nu_{j_{2^{j}}}-\nu_{j_{2^{i}+2^{j}}})(x_{2^{j}}+x_{2^{i}+2^{j}})+\cdots\right]\frac{1}{z_{j}(1-z_{j})}
+122m¯zizj.\displaystyle+\frac{1}{2}\frac{\partial^{2}\bar{m}}{\partial z_{i}\partial z_{j}}. (37b)

Owing to the independence of xix_{i}’s, from Eq. (37a) and (37b), one can easily say that the condition (36) is satisfied when

νi2j=νi0,νi2j+1=νi1,,νi2i=νi2j+2i,\nu_{i_{2^{j}}}=\nu_{i_{0}},~{}~{}\nu_{i_{2^{j}+1}}=\nu_{i_{1}},\cdots,~{}~{}\nu_{i_{2^{i}}}=\nu_{i_{2^{j}+2^{i}}},\cdots (38)

and

νj2i=νj0,νj2i+1=νj1,,νj2j=νj2i+2j,.\nu_{j_{2^{i}}}=\nu_{j_{0}},~{}~{}\nu_{j_{2^{i}+1}}=\nu_{j_{1}},\cdots,~{}~{}\nu_{j_{2^{j}}}=\nu_{j_{2^{i}+2^{j}}},\cdots. (39)

Now let the probability of mutation from Ai(l)A_{i}^{(l)} allele to Aj(l)A_{j}^{(l)} (i,j{0,1}i,j\in\{0,1\} and iji\neq j) allele at locus ll be μijl\mu^{l}_{ij}. In independent point mutations, the probability of mutation from one gamete to another gamete is just the product of the corresponding allelic mutation probabilities. So we can write down the explicit expressions of νi0\nu_{i_{0}} and νi2j\nu_{i_{2^{j}}} (j<i\forall j<i):

νi0=[(μ000μ001μ01iμ00n1μ00n)++(μ010μ011μ01iμ00n1μ00n)+(μ000μ001μ01iμ01i+1μ00n)+],\displaystyle\nu_{i_{0}}=-\left[\left(\mu^{0}_{00}\mu^{1}_{00}\cdots\mu^{i}_{01}\cdots\mu^{n-1}_{00}\mu^{n}_{00}\right)+\cdots+\left(\mu^{0}_{01}\mu^{1}_{01}\cdots\mu^{i}_{01}\cdots\mu^{n-1}_{00}\mu^{n}_{00}\right)+\left(\mu^{0}_{00}\mu^{1}_{00}\cdots\mu^{i}_{01}\mu^{i+1}_{01}\cdots\mu^{n}_{00}\right)+\cdots\right],~{}~{}~{}~{}~{}~{}~{}~{}~{} (40a)
νi2j=[(μ000μ10jμ01iμ00n)++(μ010μ11jμ01iμ00n)+(μ000μ10jμ01iμ01i+1μ00n)+],\displaystyle\nu_{i_{2^{j}}}=-\left[\left(\mu^{0}_{00}\cdots\mu^{j}_{10}\cdots\mu^{i}_{01}\cdots\mu^{n}_{00}\right)+\cdots+\left(\mu^{0}_{01}\cdots\mu^{j}_{11}\cdots\mu^{i}_{01}\cdots\mu^{n}_{00}\right)+\left(\mu^{0}_{00}\cdots\mu^{j}_{10}\cdots\mu^{i}_{01}\mu^{i+1}_{01}\cdots\mu^{n}_{00}\right)+\cdots\right], (40b)

and similarly all the other ν\nu’s in terms of the multiplication of mutation probabilities at each locus. One can check from the expressions above that both νi0\nu_{i_{0}} and νi2j\nu_{i_{2^{j}}} ultimately boil down to μ01i-\mu^{i}_{01}. An intuition behind it may be gained by noting that νi0\nu_{i_{0}} (and νi2j\nu_{i_{2^{j}}}) contains terms which are (apart from an overall negative sign) addition of the probabilities of mutation from 0th (and 2j{2^{j}}th) gamete to the gametes which do not contribute to the allele frequency, ziz_{i}. Actually, the sum of the latter gametes’ frequencies give 1zi1-z_{i}. Consequently, we can effectively consider the sum under question as the probability of mutation from A0(i)A_{0}^{(i)} allele to A1(i)A_{1}^{(i)} allele at the iith locus. Likewise, other conditions [(38) and (39)] are also satisfied trivially in the presence of independent mutations. So we can conclude that in the presence of independent point mutations, the multi-locus two-allele system is gradient-like.

Appendix B Local Lyapunov function

Here we show that the function, V(p,q)exp(m¯)p2ν1(1p)2ν3q2ν5(1q)2ν7V(p,q)\equiv\exp(\bar{m})p^{2\nu_{1}}(1-p)^{2\nu_{3}}q^{2\nu_{5}}(1-q)^{2\nu_{7}}, is a local Lyapunov function for the gradient system (14). First we note that the value of the function V(p,q)V(p,q) is always positive for all values of pp and qq in int{\rm int}\mathbb{P}. Next we show below that the function has local maxima (minima) at the stable (unstable) fixed points of Eqs. (14).

Considering the alternate form of Eqs. (14), viz., Eqs. (15), we see that every fixed points (p,q)int(p^{*},q^{*})\in{\rm int}\mathbb{P} corresponds to V/p=V/q=0{\partial V}/{\partial p}={\partial V}/{\partial q}=0. For later use, we write below the explicit expressions of the second derivative of V(p,q)V(p,q):

2Vp2|(p,q)=2V[mAν1ν3]p(1p),\displaystyle\frac{\partial^{2}V}{\partial p^{2}}\bigg{|}_{(p^{*},q^{*})}=\frac{2V\left[m_{A}-\nu_{1}-\nu_{3}\right]}{p^{*}(1-p^{*})}, (41a)
2Vq2|(p,q)=2V[mBν5ν7]q(1q),\displaystyle\frac{\partial^{2}V}{\partial q^{2}}\bigg{|}_{(p^{*},q^{*})}=\frac{2V\left[m_{B}-\nu_{5}-\nu_{7}\right]}{q^{*}(1-q^{*})}, (41b)
2Vpq|(p,q)=Vm¯;\displaystyle\frac{\partial^{2}V}{\partial p\partial q}\bigg{|}_{(p^{*},q^{*})}=V\bar{m}; (41c)

where mAp[p(1p)2m¯p]|(p,q)m_{A}\equiv\frac{\partial}{\partial p}\left[\frac{p(1-p)}{2}\frac{\partial\bar{m}}{\partial p}\right]\bigg{|}_{(p^{*},q^{*})} and mBq[q(1q)2m¯q]|(p,q)m_{B}\equiv\frac{\partial}{\partial q}\left[\frac{q(1-q)}{2}\frac{\partial\bar{m}}{\partial q}\right]\bigg{|}_{(p^{*},q^{*})}.

Now, in the process of the linear stability analysis, the Jacobian of the linearized system at (p,qp^{*},q^{*}) is given by

J1[mAν1ν32p(1p)m¯2q(1q)m¯mBν5ν7][abcc]J_{1}\equiv\begin{bmatrix}m_{A}-\nu_{1}-\nu_{3}&2p^{*}(1-p^{*})\bar{m}\\ 2q^{*}(1-q^{*})\bar{m}&m_{B}-\nu_{5}-\nu_{7}\\ \end{bmatrix}\equiv\begin{bmatrix}a&b\\ c&c\\ \end{bmatrix} (42)

and the corresponding eigenvalues are

λ=12[(a+d)±(ad)2+4bc].\lambda=\frac{1}{2}\left[(a+d)\pm\sqrt{(a-d)^{2}+4bc}\right]. (43)

Obviously, a necessary condition for the fixed point (p,q)(p^{*},q^{*}) to be stable is (a+d)<0(a+d)<0, which implies

eitherν1+ν3>mA\displaystyle\text{either}~{}~{}\nu_{1}+\nu_{3}>m_{A} (44a)
orν5+ν7>mB.\displaystyle~{}~{}\text{or}~{}~{}~{}~{}~{}\nu_{5}+\nu_{7}>m_{B}. (44b)

This condition in the light of Eqs. (41) means that

either2Vp2|(p,q)<0\displaystyle\text{either}~{}~{}\frac{\partial^{2}V}{\partial p^{2}}\bigg{|}_{(p^{*},q^{*})}<0 (45a)
or2Vq2|(p,q)<0.\displaystyle~{}~{}\text{or}~{}~{}~{}~{}~{}\frac{\partial^{2}V}{\partial q^{2}}\bigg{|}_{(p^{*},q^{*})}<0. (45b)

Finally, the sufficient condition for the stability of the fixed point is (a+d)2>(ad)2+4bc(a+d)^{2}>(a-d)^{2}+4bc, which boils down to

[mAν1ν3][mBν5ν7]>4pq(1p)(1q)m¯2,\displaystyle\left[m_{A}-\nu_{1}-\nu_{3}\right]\left[m_{B}-\nu_{5}-\nu_{7}\right]>4p^{*}q^{*}(1-p^{*})(1-q^{*})\bar{m}^{2},

which in the light of Eqs. (41) implies that

[2Vp2|(p,q)][2Vq2|(p,q)]>[2Vpq|(p,q)]2.\displaystyle\left[\frac{\partial^{2}V}{\partial p^{2}}\bigg{|}_{(p^{*},q^{*})}\right]\left[\frac{\partial^{2}V}{\partial q^{2}}\bigg{|}_{(p^{*},q^{*})}\right]>\left[\frac{\partial^{2}V}{\partial p\partial q}\bigg{|}_{(p^{*},q^{*})}\right]^{2}.\quad (47)

Conditions (45) and (47) together imply that the stable fixed point corresponds to local maximum of the function V(p,q)V(p,q). Similar calculation shows that the unstable fixed point corresponds to local minimum and saddle fixed point corresponds to saddle point of the function V(p,q)V(p,q). Furthermore, we already know from Eq. (16) that V˙>0\dot{V}>0 except at the fixed points where it vanishes. Hence, it qualifies for the local Lyapunov function.

Actually, in the more conventional sense Jordan and Smith (2007), the local Lyapunov function near stable a fixed point, (p,q)(p^{*},q^{*}), is V(p,q)V(p,q)V(p^{*},q^{*})-V(p,q) and near unstable a fixed point, (p,q)(p^{*},q^{*}), is V(p,q)V(p,q)V(p,q)-V(p^{*},q^{*}).

References

  • Taylor and Jonker (1978) P. D. Taylor and L. B. Jonker, Mathematical Biosciences 40, 145 (1978).
  • Schuster and Sigmund (1983) P. Schuster and K. Sigmund, Journal of Theoretical Biology 100, 533 (1983), URL https://doi.org/10.1016/0022-5193(83)90445-9.
  • Cressman and Tao (2014) R. Cressman and Y. Tao, Proc. Natl. Acad. Sci. 111, 10810 (2014), URL https://doi.org/10.1073/pnas.1400823111.
  • Mukhopadhyay and Chakraborty (2021) A. Mukhopadhyay and S. Chakraborty, Chaos: An Interdisciplinary Journal of Nonlinear Science 31, 023123 (2021), URL https://doi.org/10.1063/5.0032311.
  • O’Brien (1993) S. J. O’Brien, Genetic maps: locus maps of complex genomes (Cold Spring Harbor Laboratory Press, 1993).
  • Mendel and and (1866) G. Mendel and R. C. P. and, Versuche über Pflanzen-Hybriden / (Im Verlage des Vereines,, 1866).
  • Bateson et al. (1909) W. Bateson, E. R. Waunders, and R. C. Punnett, Zeitschrift für Induktive Abstammungs- und Vererbungslehre 2, 17 (1909).
  • Dooner et al. (1991) H. K. Dooner, T. P. Robbins, and R. A. Jorgensen, Annual Review of Genetics 25, 173 (1991), URL https://doi.org/10.1146/annurev.ge.25.120191.001133.
  • Carver (2009) B. F. Carver, ed., Wheat Science and Trade (Wiley-Blackwell, 2009), URL https://doi.org/10.1002/9780813818832.
  • Dean and Dean (2005) L. Dean and L. Dean, Blood groups and red cell antigens, vol. 2 (NCBI Bethesda, 2005).
  • Yasuda et al. (2022) S. P. Yasuda, Y. Miyasaka, X. Hou, Y. Obara, H. Shitara, Y. Seki, K. Matsuoka, A. Takahashi, E. Wakai, H. Hibino, et al., Biomedicines 10, 2221 (2022), URL https://doi.org/10.3390/biomedicines10092221.
  • White and Rabago-Smith (2010) D. White and M. Rabago-Smith, J. Hum. Genet. 56, 5 (2010), URL https://doi.org/10.1038/jhg.2010.126.
  • Haldane (1931) J. B. S. Haldane, 27, 137 (1931).
  • Wright (1935) S. Wright, Journal of Genetics 30, 257 (1935).
  • Lewontin and Kojima (1960) R. Lewontin and K. Kojima, Evolution pp. 458–472 (1960).
  • Bodmer and Felsenstein (1967) W. F. Bodmer and J. Felsenstein, Genetics 57, 237 (1967).
  • Karlin et al. (1970) S. Karlin, M. W. Feldman, et al., Theoretical population biology 1, 39 (1970).
  • Feldman and Libermann (1979) M. W. Feldman and U. Libermann, Genetics 92, 1355 (1979).
  • Hastings (1985) A. Hastings, Genetics 109, 255 (1985).
  • Ewens (1968) W. Ewens, Theoretical and Applied Genetics 38, 140 (1968).
  • Akin (1982) E. Akin, Journal of Mathematical Biology 13, 305 (1982).
  • Hastings (1981) A. Hastings, PNAS 78, 7224 (1981), URL https://doi.org/10.1073/pnas.78.11.7224.
  • Hofbauer and Iooss (1984) J. Hofbauer and G. Iooss, Monatshefte für Mathematik 98, 99 (1984).
  • Nagylaki (1989) T. Nagylaki, Genetics 122, 235 (1989).
  • Nagylaki (1993) T. Nagylaki, Genetics 134, 627 (1993).
  • Nagylaki et al. (1999a) T. Nagylaki, J. Hofbauer, and P. Brunovskỳ, Journal of mathematical biology 38, 103 (1999a).
  • Nagylaki (2013a) T. Nagylaki, Introduction to theoretical population genetics, vol. 21 (Springer Science & Business Media, 2013a).
  • Pontz et al. (2018) M. Pontz, J. Hofbauer, and R. Bürger, Journal of mathematical biology 76, 151 (2018).
  • Nowak et al. (2004) M. A. Nowak, F. Michor, N. L. Komarova, and Y. Iwasa, PNAS 101, 10635 (2004), URL https://doi.org/10.1073/pnas.0400747101.
  • Li et al. (2013) W. Li, F. Teng, T. Li, and Q. Zhou, Nat. Biotechnol. 31, 684 (2013), URL https://doi.org/10.1038/nbt.2652.
  • Steele et al. (1998) R. J. C. Steele, A. M. Thompson, P. A. Hall, and D. P. Lane, British Journal of Surgery 85, 1460 (1998), URL https://doi.org/10.1046/j.1365-2168.1998.00910.x.
  • Galhardo et al. (2007) R. S. Galhardo, P. J. Hastings, and S. M. Rosenberg, Crit. Rev. Biochem. Mol. Biol. 42, 399 (2007), URL https://doi.org/10.1080/10409230701648502.
  • Baer (2008) C. F. Baer, PLoS Biology 6, e52 (2008), URL https://doi.org/10.1371/journal.pbio.0060052.
  • Rando and Verstrepen (2007) O. J. Rando and K. J. Verstrepen, Cell 128, 655 (2007), URL https://doi.org/10.1016/j.cell.2007.01.023.
  • Mobilia (2010) M. Mobilia, Journal of Theoretical Biology 264, 1 (2010), URL https://doi.org/10.1016/j.jtbi.2010.01.008.
  • Toupo and Strogatz (2015) D. F. Toupo and S. H. Strogatz, Physical Review E 91, 052907 (2015).
  • Mittal et al. (2020) S. Mittal, A. Mukhopadhyay, and S. Chakraborty, Phys. Rev. E 101, 042410 (2020).
  • Mukhopadhyay et al. (2021) A. Mukhopadhyay, S. Chakraborty, and S. Chakraborty, Journal of Physics: Complexity 2, 035005 (2021).
  • Park and Krug (2011) S.-C. Park and J. Krug, Journal of mathematical biology 62, 763 (2011).
  • Qu et al. (2020) J. Qu, S. D. Kachman, D. Garrick, R. L. Fernando, and H. Cheng, Frontiers in genetics 11, 362 (2020).
  • Bengtsson and Christiansen (1983) B. O. Bengtsson and F. B. Christiansen, Theoretical Population Biology 24, 59 (1983).
  • Jain and Seetharaman (2021) K. Jain and S. Seetharaman, Journal of Nonlinear Mathematical Physics 18, 321 (2021), URL https://doi.org/10.1142/s1402925111001556.
  • Bürger (1989) R. Bürger, Genetics 121, 175 (1989), URL https://doi.org/10.1093/genetics/121.1.175.
  • Christiansen and Frydenberg (1977) F. B. Christiansen and O. Frydenberg, American Journal of Human Genetics 29, 195 (1977).
  • Karlin and McGregor (1971) S. Karlin and J. McGregor, Theoretical Population Biology 2, 60 (1971), URL https://doi.org/10.1016/0040-5809(71)90005-0.
  • Hofbauer and Sigmund (1998) J. Hofbauer and K. Sigmund, Evolutionary Games and Population Dynamics (Cambridge University Press, Cambridge, 1998).
  • Hofbauer (1985) J. Hofbauer, J. Math. Biol. 23, 41 (1985), URL https://doi.org/10.1007/bf00276557.
  • Karlin (1975) S. Karlin, Theoretical Population Biology 7, 364 (1975), URL https://doi.org/10.1016/0040-5809(75)90025-8.
  • Karlin and Carmelli (1975) S. Karlin and D. Carmelli, Theoretical Population Biology 7, 399 (1975), URL https://doi.org/10.1016/0040-5809(75)90026-x.
  • Sacker and von Bremen (2003) R. J. Sacker and H. F. von Bremen, Journal of Difference Equations and Applications 9, 441 (2003), URL https://doi.org/10.1080/1023619031000076551.
  • Hallgrímsdóttir and Yuster (2008) I. B. Hallgrímsdóttir and D. S. Yuster, BMC Genetics 9 (2008), URL https://doi.org/10.1186/1471-2156-9-17.
  • Lewontin and Feldman (1988) R. Lewontin and M. W. Feldman, Theoretical Population Biology 34, 177 (1988), URL https://doi.org/10.1016/0040-5809(88)90041-x.
  • Franklin and Feldman (1977) I. R. Franklin and M. W. Feldman, Theoretical Population Biology 12, 95 (1977), URL https://doi.org/10.1016/0040-5809(77)90037-5.
  • Pritchett-Ewing (1981a) E. Pritchett-Ewing, Genetics 98, 409 (1981a), URL https://doi.org/10.1093/genetics/98.2.409.
  • Bergero et al. (2021) R. Bergero, P. Ellis, W. Haerty, L. Larcombe, I. Macaulay, T. Mehta, M. Mogensen, D. Murray, W. Nash, M. J. Neale, et al., Biological Reviews 96, 822 (2021), URL https://doi.org/10.1111/brv.12680.
  • Nagylaki (2013b) T. Nagylaki, Introduction to theoretical population genetics, vol. 21 (Springer Science & Business Media, 2013b).
  • Nagylaki et al. (1999b) T. Nagylaki, J. Hofbauer, and P. Brunovskỳ, Journal of mathematical biology 38, 103 (1999b).
  • Shahshahani (1979) S. Shahshahani, A new mathematical framework for the study of linkage and selection (American Mathematical Soc., 1979).
  • Sigmund (1985) K. Sigmund, The maximum priciple for replicator equations, in: Lotka-Volterra-approach to Cooperation and Competition in Dynamic Systems, eds. W. Ebeling and M. Peschel, vol. 23 (Akademie-Verlag, 1985).
  • Reed et al. (2003) D. R. Reed, X. Li, A. H. McDaniel, K. Lu, S. Li, M. G. Tordoff, A. R. Price, and A. A. Bachmanov, Mammalian Genome 14, 302 (2003), URL https://doi.org/10.1007%2Fs00335-002-2170-y.
  • Lettre et al. (2008) G. Lettre, , A. U. Jackson, C. Gieger, F. R. Schumacher, S. I. Berndt, S. Sanna, S. Eyheramendy, B. F. Voight, J. L. Butler, et al., Nature Genetics 40, 584 (2008), URL https://doi.org/10.1038%2Fng.125.
  • Adhikari et al. (2015) K. Adhikari, G. Reales, A. J. P. Smith, E. Konka, J. Palmen, M. Quinto-Sanchez, V. Acuña-Alonzo, C. Jaramillo, W. Arias, M. Fuentes, et al., Nature Communications 6 (2015), URL https://doi.org/10.1038/ncomms8500.
  • Zhivotovsky and Pylkov (1998) L. A. Zhivotovsky and K. V. Pylkov, Journal of Genetics 77, 115 (1998), URL https://doi.org/10.1007/bf02966597.
  • Wray and Goddard (2010) N. R. Wray and M. E. Goddard, Genome Medicine 2, 10 (2010), URL https://doi.org/10.1186/gm131.
  • Pritchett-Ewing (1981b) E. Pritchett-Ewing, Genetics 98, 409 (1981b), URL https://doi.org/10.1093/genetics/98.2.409.
  • Jordan and Smith (2007) D. Jordan and P. Smith, Nonlinear ordinary differential equations: an introduction for scientists and engineers (OUP Oxford, 2007).