345
Views
1
CrossRef citations to date
0
Altmetric
Special Issue in Memory of Abdul-Aziz Yakubu

Adaptive delayed reproduction in a 2-dimensional discrete-time competition model

Article: 2248171 | Received 16 Mar 2023, Accepted 10 Aug 2023, Published online: 17 Aug 2023

Abstract

This paper studies a 2-dimensional discrete-time competition model of Ricker type with reproductive delay. The model is examined under the assumption that species 1 and 2 have the same properties except that a fraction η of species 1 individuals delays the initiation of reproduction. This assumption ensures that species 1 is dominated by species 2 in the sense that species 2 is increasing whenever species 1 is increasing. It is shown that, even under this assumption, delayed reproduction can be adaptive, i.e. species 1 can invade the monoculture system of species 2 while species 2 cannot invade the monoculture system of species 1, if the population is fluctuating. The result is obtained by analytically examining the species invasibility at boundary 2-cycles, whose coordinates can be estimated by assuming η0.

1. Introduction

This paper studies the following discrete-time competition model: {x1(n+1)=bexp[α{(1η)sx1(n)+sx2(n)}](1η)sx1(n)+ηsx1(n)x2(n+1)=bexp[α{(1η)sx1(n)+sx2(n)}]sx2(n),where x1 and x2 denote the number of juveniles of species 1 and 2, respectively. It is assumed that both species are semelparous, i.e. individuals reproduce only once in their lifetime and die immediately after reproduction. The parameter s denotes the survival probability of juveniles. The most important assumption is that a fraction η of survived juveniles of species 1 postpones maturity and reproduction while all survived juveniles of species 2 mature and reproduce in unit time. The density dependence is assumed to act on reproduction, so the factor bexp[α{(1η)sx1(n)+sx2(n)}] represents the number of offspring produced by an adult individual. Here b>0 and α>0. It is intuitive that delayed reproduction is maladaptive since delayed reproduction increases the possibility of death before reproduction. However, as shown in several papers (e.g. see [Citation18]), delayed reproduction can be adaptive in variable environments. Although these studies assume external factors modelling variable environments, we show that delayed reproduction can be adaptive even in a simple nonlinear deterministic model that exhibits population fluctuation.

By writing x~1(n)=α(1η)sx1(n), x~2(n)=αsx2, and λ=ln(bs), the above model equation becomes (1) {x1(n+1)=x1(n){(1η)exp(λx1(n)x2(n))+ηs}x2(n+1)=x2(n)exp(λx1(n)x2(n)),(1) where the tilde is dropped for convenience. This system shall be studied in this paper under the condition that λ>0, s,η(0,1). It is clear that this system is a special case of the following system (2) {x1(n+1)=x1(n)exp(r1k1(x1(n)+x2(n)))+βx1(n)x2(n+1)=x2(n)exp(r2k2(x1(n)+x2(n))),(2) where r1, r2, k1, k2, and β are positive. This competition system was studied by Yakubu [Citation19], who considered that the first terms of the right-hand sides of both equations represent the density dependent self-reproduction while the term βx1(n) indicates that species 1 is planted (or stocked) at a fixed rate and found several important properties of system (Equation2). To understand these properties, we need to recall the dynamics of the following competition model of Ricker type: (3) {x1(n+1)=x1(n)exp(r1a11x1(n)a12x2(n))x2(n+1)=x2(n)exp(r2a21x1(n)a22x2(n)),(3) where r1, r2, a11, a12, a21, and a22 are positive. System (Equation2) with β=0 is a special case of system (Equation3). This system has been studied in many papers (e.g. see [Citation5,Citation10,Citation11,Citation15]) since Hassell and Comins [Citation8] to understand competition between two species whose reproductions occur discretely in time. Franke and Yakubu [Citation6, Theorem 3.1] studied system (Equation3) as a special case of (4) {x1(n+1)=x1(n)g1(a11x1(n)+a12x2(n))x2(n+1)=x2(n)g2(a21x1(n)+a22x2(n))(4) and showed that extinction of species i occurs in system (Equation3) irrespective of initial conditions if species i is dominated by the other species in the sense that the other species is increasing whenever species i is increasing. More precisely, species i is said to be dominated by species j if Di is a proper subset of Dj, where Di:={(x1,x2)R+2:gi(ai1x1+ai2x2)1} and R+2:={(x1,x2)R2:x10,x20}. In the case of system (Equation2) with β=0, except for the special case r1k1=r2k2, either species 1 or 2 is dominated by the other. Thus, under the assumption that species 1 is dominated by species 2 in system (Equation2) with β=0, Yakubu [Citation19] examined the influence of planting of species 1 on its persistence and revealed that two species of system (Equation2) can coexist at a stable positive 2-cycle. The general result of Franke and Yakubu [Citation5, Theorem 6.1] on system (Equation4) shows that if each single-species dynamics within the coordinate axes has a globally attractive positive fixed point, then the dominated species goes extinct irrespective of initial conditions. Therefore, for species coexistence in system (Equation2) oscillation of single-species population is crucial. This result is fascinating since coexistence can occur even though system (Equation2) has no positive fixed points.

After the paper [Citation19], Yakubu [Citation20] posed open problems concerning the global attractivity of a positive 2-cycle of (Equation2) and the dynamics of system (Equation2) and its variant have been studied in several papers. By applying a general result, Elaydi and Yakubu [Citation4] showed that the positive 2-cycle of (Equation2) cannot be globally attracting in the interior of R+2. Kon [Citation14] showed that species 2 goes extinct in system (Equation2) irrespective of initial conditions if species 2 is dominated by species 1. Kang and Smith [Citation13] studied a variant of system (Equation2) and showed that the system has a positive 2-cycle that attracts all positive points except Lebesgue measure zero set (see also [Citation12]). Although these studies advanced our understanding of system (Equation2), its dynamics when species 1 is dominated by species 2 is still not well understood. In this paper, by focussing on the stability of boundary 2-cycles of system (Equation1), which is a special case of system (Equation2), we reveal its dynamics and give mathematical evidence that delayed reproduction can be adaptive when the population is fluctuating.

The rest of this paper is organized as follows. Section 2 studies the dynamics of system (Equation1) on the boundary of R+2. We are concerned with boundary 2-cycles. Since the dynamics on the x2-axis obeys the Ricker map, we review the analytical expression of its 2-cycles. By using the fact that the dynamics on the x1-axis can be seen as a perturbation of the Ricker map, we obtain an approximate expression of 2-cycles on the x1-axis by assuming η0. Section 3 reviews the fixed-point stability of system (Equation1) and confirms that species 1 is dominated by species 2. Section 4 examines the stability of the boundary 2-cycles constructed in Section 2 and analytically obtain the stability criterion in terms of parameters. By the result of Section 4, we find that delayed reproduction can be adaptive. Furthermore, we find that it is unlikely that two species can either mutually invade or mutually prevent invasion of each other at boundary 2-cycles if η is small. However, we numerically show that the parameter region for mutual invasion at boundary 2-cycles gets larger as η increases. In such a parameter region, we find that two species can coexist. Section 5 examines the global dynamics of (Equation1) both analytically and numerically. We analytically show that it is unlikely that system (Equation1) has a positive 2-cycle if η is small. Numerical simulations suggest that if system (Equation1) has a stable boundary 2-cycle, then it attracts almost all positive points as long as η0. Section 6 includes some concluding remarks.

2. Boundary dynamics and boundary 2-cycles

On the x1- and x2-axes, system (Equation1) is reduced to the one-dimensional maps f1(x):=(1η)xexp(λx)+ηsxandf2(x):=xexp(λx),respectively. Note that f1(x)=f2(x) for all x if η=0.

The map f2 is the well-known Ricker map (e.g. see [Citation16]). This map has two fixed points 0 and X2:=λ. The inequality f2(0)>1 always holds and |f2(X2)|<1 holds if and only if λ(0,2). Thus, the origin is always unstable and the positive fixed point X2 is asymptotically stable if λ(0,2) and is unstable if λ>2. It is known that, at the critical value of λ=2, a period-doubling bifurcation occurs resulting in the birth of an asymptotically stable 2-cycle. Let P2 and P2 (0<P2<P2) be 2-periodic points of the 2-cycle. Then they have the following analytical expression [Citation17]: (5) P2=2ξ1eξ=λ+ξ,P2=2ξeξ1eξ=λξ,(5) where ξ=h1(λ) and h is the strictly increasing function defined by (6) h(ξ):=ξ1+eξ1eξ=ξtanh(ξ2)(6) for ξ>0. Note that h(ξ)2 as ξ0+. Since P2P2=2h1(λ) is an increasing function of λ, the amplitude of the 2-cycle {P2,P2} increases as λ increases. The 2-cycle {P2,P2} is asymptotically stable if |f2(P2)f2(P2)|<1, which is equivalent to 2<λ<λ2, where λ22.5264 (e.g. see [Citation3]).

The map f1 has two fixed points 0 and X1:=λ+ln1η1sη. The inequality f1(0)>1 always holds and |f1(X1)|<1 holds if and only if λ(0,21η+ln1ηs1η). Concerning a 2-cycle of f1, we obtain the following lemma.

Lemma 2.1

The map f1 has a positive 2-cycle {P1,P1} for all η>0 sufficiently small if λ>2. The 2-cycle {P1,P1} is asymptotically stable (resp. unstable) for all η>0 sufficiently small if 2<λ<λ2 (resp. λ>λ2). The 2-periodic points have the following expression: (7) P1=P2+2+s(P2P2+P2P2)+P2sP2P2P2(1P2)+1η+O(η2)P1=P2+2+s(P2P2+P2P2)+P2sP2P2P2(1P2)+1η+O(η2).(7)

Proof.

Let P1 and P1 (0<P1<P1) be 2-periodic points of the map f1. Then they are given by solving the following equation for x: ϕ(x,η):={(1η)exp(λx)+ηs}{(1η)exp(λx)+ηs}=1,where x={(1η)exp(λx)+ηs}x. Since f1 is reduced to the Ricker map f2 when η=0, both ϕ(P2,0)=1 and ϕ(P2,0)=1 hold. Thus, by the implicit function theorem, the equation ϕ(x,η)=1 has the solutions (8) P1=P2ηϕηϕx|(x,η)=(P2,0)+O(η2)P1=P2ηϕηϕx|(x,η)=(P2,0)+O(η2)(8) in neighbourhoods of (P2,0) and (P2,0), respectively. It is straightforward to show that Equation (Equation8) is equivalent to (Equation7). Since the 2-cycle {P2,P2} of the map f2 exists if λ>2, the 2-cycle {P1,P1} of the map f1 exists for all η>0 sufficiently small if λ>2. Furthermore, the 2-cycle {P1,P1} of the map f1 is asymptotically stable (resp. unstable) for all η>0 sufficiently small if λ(2,λ2) (resp. λ>λ2) since f1, P1, and P1 continuously depend on η in a neighbourhood of η=0 and f1(P1)f1(P1)=f2(P2)f2(P2) holds if η=0.

3. Stability of fixed points

The fixed-point stability of system (Equation2) is studied by Yakubu [Citation19] (see also [Citation4]). In this section, we briefly review the fixed-point stability of a special case of system (Equation2), i.e. (Equation1). The fixed points of system (Equation1) are given by solving the system of equations (9) {x1=x1{(1η)exp(λ(x1+x2))+ηs}x2=x2exp(λ(x1+x2)).(9) It is clear that system (Equation1) has the three boundary fixed points (0,0), (X1,0), and (0,X2), where X1=λ+ln1η1sη and X2=λ as defined in the previous section. Since η,s(0,1), the inequality X2>X1 always holds. If both x1 and x2 are not zero, the first and second equations of (Equation9) are reduced to the incompatible equations x1+x2=X1 and x1+x2=X2, respectively. Thus, there are no positive fixed points. Furthermore, X2>X1 implies that D1 is a proper subset of D2, i.e. species 1 is dominated by species 2 (see Figure ).

Figure 1. The phase plane (x1,x2) of system (Equation1). The set D1 is a proper subset of D2. Species 1 is dominated by species 2. The unstable manifold of (X1,0) and the stable manifold of (0,X2) have nonempty intersections with the interior of R+2.

Figure 1. The phase plane (x1,x2) of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The set D1 is a proper subset of D2. Species 1 is dominated by species 2. The unstable manifold of (X1,0) and the stable manifold of (0,X2) have nonempty intersections with the interior of R+2.

Let J(x1,x2) be the Jacobi matrix of system (Equation1) evaluated at (x1,x2). Then J(0,0) has the eigenvalues (1η)eλ+ηs and eλ, both of which are larger than one for all η>0 sufficiently small since λ>0. Thus, the origin is a source for all η>0 sufficiently small. J(X1,0) has the eigenvalues 1(1sη)X1 and exp(λX1)=exp(X2X1)>1. The former eigenvalue determines the stability of the fixed point X1 of the map f1 defined in the previous section. Note that X1 is destabilized at λ=21η+ln1ηs1η as λ is increased. The latter eigenvalue is transversal in the sense that it has an eigenvector transversal to the x1-axis. Thus the unstable manifold of (X1,0) always has a nonempty intersection with the interior of R+2. J(0,X2) has the eigenvalues 1X2 and (1η)exp(λX2)+ηs=1η(1s)<1. The former eigenvalue determines the stability of the fixed point X2 of the map f2 defined in the previous section. Note that X2 is destabilized at λ=2 as λ is increased. The latter eigenvalue is transversal. Thus the stable manifold of (0,X2) always has a nonempty intersection with the interior of R+2. Figure summarizes some of the above results.

4. Stability of boundary 2-cycles

Concerning the stability of boundary 2-cycles of system (Equation1), we obtain the following theorem.

Theorem 4.1

If λ>2, then system (Equation1) with η>0 sufficiently small has the 2-cycles P1:={(P1,0),(P1,0)} and P2:={(0,P2),(0,P2)}.

(a)

If 2<λ<λ2 and λ<h(cosh11s), then P1 is a saddle with an unstable manifold that intersects the interior of R+2 and P2 is a sink whenever η>0 is sufficiently small.

(b)

If 2<λ<λ2 and λ>h(cosh11s), then P1 is a sink and P2 is a saddle with an unstable manifold that intersects the interior of R+2 whenever η>0 is sufficiently small.

(c)

If λ>2 and λ<h(cosh11s), then P1 is a source and P2 is a saddle with a stable manifold that intersects the interior of R+2 whenever η>0 is sufficiently small.

(d)

If λ>2 and λ>h(cosh11s), then P1 is a saddle with a stable manifold that intersects the interior of R+2 and P2 is a source whenever η>0 is sufficiently small.

Proof.

Since system (Equation1) is reduced to the Ricker map on the x2-axis, system (Equation1) has the 2-cycle P2={(0,P2),(0,P2)} for all η>0 if λ>2. Recall that J(x1,x2) denotes the Jacobi matrix of system (Equation1) evaluated at (x1,x2). Then the Jacobi matrix of the second iterate of system (Equation1) evaluated at P2 is given by J(0,P2)J(0,P2). This matrix has the eigenvalues (10) f2(P2)f2(P2)and{(1η)exp(λP2)+ηs}{(1η)exp(λP2)+ηs}.(10) As mentioned in Section 2, the absolute value of the former eigenvalue is less than one if λ(2,λ2). The latter eigenvalue takes the form 1+[2+s{exp(λP2)+exp(λP2)}]η+O(η2).Using (Equation5), we find that the absolute value of this eigenvalue is less (resp. larger) than one for all η>0 sufficiently small if s<1coshξ (resp. s>1coshξ). Therefore, the case where the 2-cycle P2 is hyperbolic can be divided into four cases: (a) 2<λ<λ2 and s<1coshξ, (b) 2<λ<λ2 and s>1coshξ, (c) λ>λ2 and s<1coshξ, and (d) λ>λ2 and s>1coshξ. The 2-cycle P2 is a sink in case (a), a saddle in cases (b) and (c), and a source in case (d). Furthermore, we find that P2 has an unstable manifold that intersects the interior of R+2 in case (b) since the latter eigenvalue is transversal. Similarly, we find that P2 has a stable manifold that intersects the interior of R+2 in case (c).

By Lemma 2.1, system (Equation1) has the 2-cycle P1={(P1,0),(P1,0)} on the x1-axis for all η>0 sufficiently small if λ>2. The Jacobi matrix of the second iterate of system (Equation1) evaluated at P1 is given by J(P1,0)J(P1,0). This matrix has the eigenvalues (11) f1(P1)f1(P1)andexp(λP1)exp(λP1).(11) As shown in Lemma 2.1, the absolute value of the former eigenvalue is less than one for all η>0 sufficiently small if λ(2,λ2). Equation (Equation5) reduces the latter eigenvalue into exp(2λP2P22(scoshξ1)η+O(η2)),whose absolute value is less (resp. larger) than one for all η>0 sufficiently small if s>1coshξ (reps. s<1coshξ) since 2λ=P2+P2. Therefore, the case where the 2-cycle P1 is hyperbolic can be divided into four cases (a), (b), (c), and (d) as above. The 2-cycle P1 is a saddle in case (a), a sink in cases (b), a source in case (c), and a saddle in case (d). Furthermore, we find that P1 has an unstable manifold that intersects the interior of R+2 in case (a) since the latter eigenvalue is transversal. Similarly, we find that P1 has a stable manifold that intersects the interior of R+2 in case (d).

The inequalities s<1coshξ and s>1coshξ can be rewritten as follows: (12) s<1coshξλ<h(cosh11s)ands>1coshξλ>h(cosh11s).(12) In fact, since the function h(x) defined by (Equation6) is strictly increasing for x>0, λ=h(ξ)<h(cosh11s) if and only if ξ<cosh11s and λ=h(ξ)>h(cosh11s) if and only if ξ>cosh11s. Thus we obtain (Equation12).

Note that P1 is asymptotically stable for all η>0 sufficiently small if 2<λ<λ2 and λ>h(cosh11s) and is unstable for all η>0 sufficiently small if either λ>λ2 or λ<h(cosh11s) and P2 is asymptotically stable for all η>0 sufficiently small if 2<λ<λ2 and λ<h(cosh11s) and is unstable for all η>0 sufficiently small if either λ>λ2 or λ>h(cosh11s).

The above result is summarized in Table . The table shows that each boundary fixed point or 2-cycle has the indicated stability if they satisfy the indicated parameter conditions. Note, however, that the fixed point (X1,0) and the 2-cycles P1 and P2 have the indicated stability if η>0 is sufficiently small. Since cosh1x=ln(x+x21), we have the following expression: h(cosh11s)=h(ln(1s+1s21)).Figure  shows the parameter plane demarcated with the stability of the boundary fixed points and the boundary 2-cycles of system (Equation1) with η0.

Table 1. Stability of the boundary fixed points and the boundary 2-cycles of system (Equation1).

Figure 2. The parameter plane (λ,s) of system (Equation1) with η0. In regions A, B, and C, (0,X2), P2, and P1 are asymptotically stable, respectively. There are no stable fixed points nor stable 2-cycles on the boundary of R+2 if λ>λ2.

Figure 2. The parameter plane (λ,s) of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ) with η≈0. In regions A, B, and C, (0,X2), P2, and P1 are asymptotically stable, respectively. There are no stable fixed points nor stable 2-cycles on the boundary of R+2 if λ>λ2.

Figure 3. The parameter plane (λ,s) of system (Equation1). The green and red regions correspond to the region B and C in Figure , respectively. In the yellow region, both P1 and P2 have unstable transversal eigenvalues, i.e. both species can mutually invade at the boundary 2-cycles. On the left and right panels, η=0.01 and η=0.1, respectively.

Figure 3. The parameter plane (λ,s) of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The green and red regions correspond to the region B and C in Figure 2, respectively. In the yellow region, both P1 and P2 have unstable transversal eigenvalues, i.e. both species can mutually invade at the boundary 2-cycles. On the left and right panels, η=0.01 and η=0.1, respectively.

From Table , we find that if (λ,s) belongs to the region C of Figure and η0, then the monoculture system of species 2 is settled in P2 and cannot prevent invasion of species 1 while the monoculture system of species 1 is settled in P1 and can prevent invasion of species 2. This implies that delayed reproduction can be adaptive if a fraction η of juveniles delays the initiation of reproduction and both the juvenile survival probability and the amplitude of population fluctuation are sufficiently large. Although we do not have a mathematical estimate how small η should be, the numerical result shown in Figure  suggests that the region C persists even if η=0.1. In Figure , the eigenvalues shown in (Equation10) and (Equation11) are numerically calculated. In the yellow region of Figure , two species can mutually invade each other at P1 and P2 and coexist. We see that the yellow region gets larger as η increases.

5. Global dynamics

In this section, we examine the global dynamics of system (Equation1) analytically and numerically. First, we show that it is unlikely that system (Equation1) has a positive 2-cycle as long as η0.

Theorem 5.1

Suppose that either λ2 or λ>2 and λh(cosh11s). Then system (Equation1) has no positive 2-cycles for all η>0 sufficiently small.

Proof.

Let {(x1,x2),(x1,x2)} be a positive 2-cycle of system (Equation1). Then it satisfies the system of equations (13) {1={(1η)exp(λx1x2)+ηs}{(1η)exp(λx1x2)+ηs}2λ=x1+x2+x1+x2,(13) where (14) x1=x1{(1η)exp(λx1x2)+ηs}andx2=x2exp(λx1s2).(14) By using the second equation of (Equation13), we can remove x1+x2 from the first equation of (Equation13) as follows: (15) (1η)sY2(2ηηs2)Y+(1η)s=0,(15) where (16) Y=exp(λ(x1+x2)).(16) Since s,η(0,1), the quadratic Equation (Equation15) has the two positive distinct real roots Y±(s,η):=2ηηs2±(1s2){η2(1s2)+4(1η)}2(1η)s.By using (Equation14) and (Equation16), we can remove x1+x2 from the second equation of (Equation13) as follows: 2λ=(λlnY)+x1{(1η)Y+ηs}+x2Y.Thus, x1 and x2 must satisfy the following system of linear equations: (17) {λlnY±(s,η)=x1+x2,λ+lnY±(s,η)={(1η)Y±(s,η)+ηs}x1+Y±(s,η)x2.(17) Since the two equations of (Equation17) represent parallel lines on the plane (x1,x2) if η=0 and have negative slopes, (Equation17) does not have positive solutions (x1,x2) for all η>0 sufficiently small if λlnY±(s,0)λ+lnY±(s,0)Y±(s,0).Since Y±(s,0)=1s±1s211 for s(0,1), the above condition is equivalent to λ(lnY±(s,0))1+1Y±(s,0)11Y±(s,0)=(lnY+(s,0))1+1Y+(s,0)11Y+(s,0)=h(lnY+(s,0))=h(cosh11s),where Y+(s,0)Y(s,0)=1 is used to obtain the first equality.

Since h(cosh11s) is strictly decreasing for s(0,1), we have h(cosh11s)>lims1h(cosh11s)=2, which implies that λh(cosh11s) holds for all s(0,1) if λ2.

Since system (Equation1) has no positive fixed points, any orbit of system (Equation1) cannot converge to a positive point. Furthermore, by the above theorem, we see that it is unlikely that an orbit staring at a positive point converges to a positive 2-cycle if η0. Although system (Equation1) might have more complicated positive attractors, numerical simulations suggest that any orbit of system (Equation1) starting at a positive point cannot remain in a compact set bounded away from the coordinate axes as long as η0 and (λ,s) belongs to the region B or C of Figure (see Figure ). Figure (a) shows the basin of attraction of P1 of system (Equation1) whose (λ,s) belongs to the region C of Figure and η=0.1. The points in the green and yellow regions are attracted by P1. The two colours indicate whether a point is attracted by (P1,0) or (P1,0) when looking at the orbit every other time. It is unlikely that the points on the boundaries of regions with different colours are attracted by P1. At least it is clear that the fixed point (0,X2) has a stable manifold that intersects with the interior of R+2. Therefore, the points on the boundaries of two different colours seems to be attracted by (0,X2). Figure  shows a single forward orbit of system (Equation1) with the same parameters as in Figure (a). The orbit starts in a neighbourhood of (X1,0). Since the fixed point (X1,0) is unstable, the orbit leaves its neighbourhood and approaches P1 after passing by the fixed point (0,X2). This behaviour suggests that there is a heteroclinic orbit from (X1,0) to (0,X2). Figure (b) shows the basin of attraction of P2 of system (Equation1) whose (λ,s) belongs to the region B of Figure and η=0.1.

Figure 4. The basins of attraction of the boundary 2-cycles of system (Equation1). The vertical and horizontal axes are x1 and x2, respectively. (a) The points in the green and yellow regions are attracted by (P1,0) and (P1,0), respectively, under the second iterate of system (Equation1). The parameters are λ=2.5, s = 0.6, and η=0.1. (b) The points in the black and red regions are attracted by (0,P2) and (0,P2), respectively, under the second iterate of system (Equation1). The parameters are λ=2.5, s = 0.1, and η=0.1.

Figure 4. The basins of attraction of the boundary 2-cycles of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The vertical and horizontal axes are x1 and x2, respectively. (a) The points in the green and yellow regions are attracted by (P1,0) and (P1∗,0), respectively, under the second iterate of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The parameters are λ=2.5, s = 0.6, and η=0.1. (b) The points in the black and red regions are attracted by (0,P2) and (0,P2∗), respectively, under the second iterate of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The parameters are λ=2.5, s = 0.1, and η=0.1.

Figure 5. A single forward orbit of system (Equation1). The points of the orbit are connected with black and red lines if they are points of 2nth and 2n + 1th iteration, respectively. The parameters are the same as in Figure (a).

Figure 5. A single forward orbit of system (Equation1(1) {x1(n+1)=x1(n){(1−η)exp⁡(λ−x1(n)−x2(n))+ηs}x2(n+1)=x2(n)exp⁡(λ−x1(n)−x2(n)),(1) ). The points of the orbit are connected with black and red lines if they are points of 2nth and 2n + 1th iteration, respectively. The parameters are the same as in Figure 4 (a).

6. Concluding remarks

We studied the dynamics of system (Equation1), which describes competition between two semelparous species. One species is assumed to have a fraction η of individuals who delay the initiation of reproduction while the other species is assumed to reproduce every unit time. All other properties of these two species are assumed to be identical. Under the assumption η0, we analytically obtained a parameter plane (λ,s) demarcated with the advantage of these two species (see Figure ). The parameter plane shows that delayed reproduction can be adaptive if both juvenile survival probability and amplitude of population fluctuation are sufficiently large as long as a fraction of juveniles delays reproduction.

System (Equation1) is a special case of system (Equation2) studied by Yakubu [Citation19], who considers competition between two semelparous species and assumes that one species is endangered and planted (or stocked) with a constant rate. In this interpretation, system (Equation1) considers the case where the endangered species whose fertility is reduced by the proportion η in compared with the competitor is planted (or stocked) at a rate ηs (0<s<1). Our result shows that if the population is fluctuating, then the endangered species can be saved by poor planting not covering the reduction of fertility. Note that a similar interpretation is possible if species 1 is considered iteroparous and species 2 semelparous (see [Citation1,Citation7]).

By applying the general result of Franke and Yakubu [Citation5, Theorem 6.1] to system (Equation1), we can show that if the fixed points (X1,0) and (0,X2) are global attractors within each positive coordinate axis, then (0,X2) attracts all points in the interior of R+2 since species 1 is dominated by species 2. However, as we showed, if the coordinate axes have the 2-cycles P1 and P2, then both P1 and P2 can be asymptotically stable in system (Equation1) with η0. Note, however, that a stable boundary 2-cycle cannot attract all points in the interior of R+2 since (0,X2) always has a stable manifold that intersects with the interior of R+2. Our numerical simulation suggests that heteroclinic orbits from (X1,0) to (0,X2) belong to the stable manifold. Furthermore, as shown by Kang and Smith [Citation13], who examines a variant of system (Equation2), it is likely that all the pre-images of the closure of the union of all the heteroclinic orbits also cannot be attracted by P1 (see also [Citation12]). It is a future problem to reveal the basin of attraction of P1 when it is asymptotically stable. It is worth noting that riddled basins of attraction are observed in a slightly more general system that includes system (Equation2) as a special case [Citation2] and intermingled basins of attraction are constructed in a certain discrete-time competition model [Citation9].

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by JSPS KAKENHI [grant number 20K03735], Japan.

References

  • M. Bulmer, Theoretical Evolutionary Ecology, Sinauer Associates, Sunderland, Massachusetts, 1994.
  • O. de Feo and R. Ferriere, Bifurcation analysis of population invasion: On-off intermittency and basin riddling, Internat. J. Bifur. Chaos Appl. Sci. Engrg. 10(2) (2000), pp. 443–452.
  • S. Elaydi and R.J. Sacker, Basin of attraction of periodic orbits of maps on the real line, J. Differ. Equ. Appl. 10(10) (2004), pp. 881–888.
  • S. Elaydi and A.-A. Yakubu, Global stability of cycles: Lotka-Volterra competition model with stocking, J. Differ. Equ. Appl. 8(6) (2002), pp. 537–549.
  • J.E. Franke and A.-A. Yakubu, Global attractors in competitive systems, Nonlinear Anal. 16(2) (1991), pp. 111–129.
  • J.E. Franke and A.-A. Yakubu, Mutual exclusion versus coexistence for discrete competitive systems, J. Math. Bio. 30(2) (1991), pp. 161–168.
  • M. Gatto, The evolutionary optimality of oscillatory and chaotic dynamics in simple population models, Theor. Popul. Biol. 43(3) (1993), pp. 310–336.
  • M. Hassell and H.N. Comins, Discrete time models for two-species competition, Theor. Popul. Biol. 9(2) (1976), pp. 202–221.
  • F. Hofbauer, J. Hofbauer, P. Raith, and T. Steinberger, Intermingled basins in a two species system, J. Math. Bio. 49(3) (2004), pp. 293–309.
  • J. Hofbauer, V. Hutson, and W. Jansen, Coexistence for systems governed by difference equations of lotka-volterra type, J. Math. Bio. 25(5) (1987), pp. 553–570.
  • H. Jiang and T.D. Rogers, The discrete dynamics of symmetric competition in the plane, J. Math. Bio. 25(6) (1987), pp. 573–596.
  • Y. Kang, Pre-images of invariant sets of a discrete-time two-species competition model, J. Differ. Equ. Appl. 18(10) (2011), pp. 1709–1733.
  • Y. Kang and H. Smith, Global dynamics of a discrete two-species Lottery-Ricker competition model, J. Biol. Dyn. 6(2) (2012), pp. 358–376.
  • R. Kon, Convex dominates concave: An exclusion principle in discrete-time Kolmogorov systems, Proc. Amer. Math. Soc. 134(10) (2006), pp. 3025–3034.
  • R. Luís, S. Elaydi, and H. Oliveira, Stability of a Ricker-type competition model and the competitive exclusion principle, J. Biol. Dyn. 5(6) (2011), pp. 636–660.
  • R.M. May and G.F. Oster, Bifurcations and dynamic complexity in simple ecological models, Am. Nat. 110(974) (1976), pp. 573–599.
  • S.D. Mylius and O. Diekmann, The resident strikes back: Invader-induced switching of resident attractor, J. Theoret. Biol. 211(4) (2001), pp. 297–311.
  • S. Tuljapurkar, Delayed reproduction and fitness in variable environments, Proc. Natl. Acad. Sci. USA 87(3) (1990), pp. 1139–1143.
  • A.-A. Yakubu, The effects of planting and harvesting on endangered species in discrete competitive systems, Math. Biosc. 126(1) (1995), pp. 1–20.
  • A.-A. Yakubu, A discrete competition system with planting, J. Differ. Equ. Appl. 4(2) (1998), pp. 213–214.