87
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Convexity of non-compact carrying simplices in logarithmic coordinates

&
Received 20 Jun 2023, Accepted 07 Jun 2024, Published online: 26 Jun 2024

Abstract

We study a non-compact version of the carrying simplex for the planar Leslie–Gower and planar Ricker maps when they are written in logarithmic variables. We show that for both of these models there is a convex (unbounded) invariant set X, and all orbits are attracted to X. For the Leslie–Gower map, which is injective, the boundary of X globally attracts all orbits and we identify it with a non-compact carrying simplex. As the Ricker map is not invertible, the boundary of X may not be invariant. We establish conditions on the parameters of the Ricker map which guarantee that there is a convex non-compact carrying simplex when r, s<1 which maps into a compact carrying simplex in the standard untransformed coordinates.

2020 Mathematics Subject Classifications:

1. Introduction

Let R+=[0,), R++=(0,) and F:R+2R+2 be a continuous function. Consider the following planar difference equation: (1) x0R+2,xn+1:=F(xn)=Fn+1(x0),nN:={0,1,2,}.(1) In this article we are interested in a special invariant curve of (Equation1) known as the carrying simplex. Hirsch's definition [Citation9] of a carrying simplex, when applied to the above system, is as follows.

Definition 1.1

We call ΣR+2{0} a carrying simplex if

(CS1)

Σ is compact and invariant.

(CS2)

For any xR+2{0} there exists yΣ such that limn|Fn(x)Fn(y)|=0. (asymptotic completeness)

(CS3)

Σ is unordered. (i.e. for (x1,y1),(x2,y2)Σ, if x1<x2, then y2<y1, and if y1<y2, then x2<x1)

When it exists, the carrying simplex Σ is thus a compact and invariant manifold for (Equation1) that attracts R+2{0} and that has the special property that Σ is the graph of a decreasing and continuous function. To date, and to the best of the authors' knowledge, planar carrying simplices for discrete dynamics have been studied exclusively in the context of retrotone systems (e.g. [Citation9,Citation11,Citation17,Citation18]).

This paper explores the pros and cons of working in alternative coordinates where compactness of the carrying simplex is lost.

Definition 1.2

A map F=(F1,F2):R+2R+2 is retrotone (e.g. [Citation9,Citation18]) in a subset DR+2 if for x,yD such that F1(x)F1(y) and F2(x)F2(y) but F(x)F(y) we have x1>y1 provided y1>0 and x2>y2 provided y2>0.

A retrotone map is sometimes also called a competitive map (see, for example, [Citation19]). In the planar case a map satisfying Definition 1.2 has the special property that it maps the graph of a decreasing function on D to the graph of a new decreasing function on D [Citation2,Citation5].

The Leslie–Gower map from ecology [Citation15] that we study in Section 3.1 is retrotone for all biologically realistic parameter values, and it is well-known that it has a unique carrying simplex [Citation13]. On the other hand, the Ricker map is not retrotone everywhere in R+2 (see for example [Citation9,Citation11,Citation18]), and so existence of a carrying simplex in the standard coordinates of population densities, by means of retronicity, is only known for a limited set of parameter values.

Here we will extend the notion of the carrying simplex applied to planar systems to allow it to be non-compact, and we will call a set ΣR2 a non-compact carrying simplex if it satisfies (CS1) without compactness and (CS3), but (CS2) is replaced by the lesser requirement that Σ globally attracts R2. The issue of asymptotic completeness will be addressed elsewhere.

In working with non-compact carrying simplices we may work in alternative coordinate systems for which the systems (Equation1) that we consider here have at most one (finite) fixed point, but in so doing we lose compactness of the global attractor and asymptotic completeness. We have found that by using logarithmically transformed coordinates, we are sometimes able to obtain stronger geometrical properties for the non-compact carrying simplex, namely that it is the graph of a concave decreasing function. While the corresponding compact carrying simplices are also known to be graphs of decreasing functions, whether or not those functions are convex or concave is not generally known (for results on convexity of carrying simplices see [Citation1,Citation3,Citation4,Citation21]). Here, we will also discuss the convexity of the boundary of the basin of repulsion of infinity in the logarithmically scaled Leslie–Gower and Ricker models. When all the parameters are positive, then the maps in the logarithmically scaled versions of both models are concave (i.e. each component of the map is a concave function [Citation14]). We take advantage of this fact to prove that the basin of repulsion of infinity is an invariant convex set. We establish a relationship between the convexity and the strict decreasingness of the members of a sequence of sets that converges to the boundary of the basin of repulsion of infinity. Then, it becomes straightforward to show that this boundary satisfies (CS3).

2. Preliminary results

In this section, we prove three lemmas that play pivotal roles. The first lemma will enable us to prove that the boundary of each of the sets we are discussing is the graph of a continuous strictly decreasing function. The second lemma shows that for given a set in a certain class of subsets of R2 whose members have boundary that is the graph of a continuous strictly decreasing function, that set must be convex.

Lemma 2.1

Let XR2 and a,bR be given. Suppose there exist two continuous functions A:(,a)R and B:(,b)R such that (2) {x|(x,c)X}={(,A(c)],c(,a),otherwise(2) (3) {y|(d,y)X}={(,B(d)],d(,b),otherwise.(3) Then X(,b)×(,a), both A and B are strictly decreasing functions, and (4) ∂X={(A(c),c)|c(,a)}(4) (5) ={(d,B(d))|d(,b)}.(5) In other words, the boundary of X is the graph of a strictly decreasing function and X is the set of all points on or under the graph of that function.

Proof.

It is clear that we have {(A(c),c)|c(,a)}∂X{(d,B(d))|d(,b)}∂X.To prove (Equation4), we observe that for each (x,y)∂X there exist {(xn,yn)}n=1X and {(xn,yn)}n=1(,b)×(,a)X such that limn(xn,yn)=limn(xn,yn)=(x,y).For each nN we have xnA(yn) and xn>A(yn). It follows that since A is continuous we have x=limnxnlimnA(yn)=A(y)=limnA(yn)limnxn=x.Hence, x=A(y) and (x,y){(A(c),c)|c(,a)}. This proves (Equation4). Proving (Equation5) is similar. It is clear that (Equation4) and (Equation5) imply that A and B are inverse of each other and they are both bijective. Hence, by using the fact that they are continuous functions, we deduce that A and B are strictly decreasing functions (These functions cannot be strictly increasing since X(,b)×(,a) and the boundary of X is equal to each of the graphs of A and B).

Before stating Lemma 2.2, we have to define the relation ‘≪’ between some members of R2. Let (x1,y1),(x2,y2)R2, we write (x1,y1)(x2,y2) if x1<x2 and y1<y2.

Lemma 2.2

Let X(,b)×(,a) be the set of points on or under the graph of the continuous strictly decreasing function B:(,b)(,a). Assume that for every x,yX and 0<λ<1 there exists at least one zX such that (6) λx+(1λ)yz.(6) Then X is convex.

Proof.

Assume that X is not convex. Then there exist x,yX and 0<λ<1 such that λx+(1λ)yX.Assume that z is as stated in the theorem. Since z1>λx1+(1λ)y1 and B is strictly decreasing, we have B(z1)<B(λx1+(1λ)y1)and since λx+(1λ)yX and zX, we have z2B(z1)B(λx1+(1λ)y1)<λx2+(1λ)y2.Hence, z2<λx2+(1λ)y2.which contradicts (Equation6).

Lemma 2.3

Let x0R be given and suppose that p:R×RR, q:R×RR and K:(,x0)R are continuous functions and p satisfies (7) limxsup{p(x,y)|yR}=.(7) Suppose also that there exists G:(,y)R defined by G(c):=supp({(x,q(x,c))|x(,K(c)]}).Then G is continuous and Ωc:=p({(x,q(x,c))|x(,K(c)]})=(,G(c)],c(,y).

Proof.

Fix c(,y). Since the continuous image of a connected set is connected, we deduce that Ωc is connected. Equation (Equation7) implies that limxp(x,q(x,c))=,and hence Ωc is unbounded below and there exists L(,K(c)) such that for every x<L we have p(x,q(x,c))<G(c)1. Therefore, by compactness of Yc:=[L,K(c)] we deduce that (8) G(c)=supΩcp({(x,q(x,c))|xYc})Ωc.(8) Connectedness of Ωc along with the fact that it is unbounded below and G(c)=supΩcΩc proves Ωc=(,G(c)].

We now prove continuity of G by contradiction. Suppose that G is not continuous at some c0(,x0). Then there exist a sequence {an} which converges to c0 and ϵ>0 such that for every nN we have |G(c0)G(an)|ϵ. By (Equation8) we know that G(c0)Ωc0. Hence there exists wc0Yc0 such that p(wc0,q(wc0,c0))=G(c0). Similarly, for every nN there exists wanYan such that p(wan,q(wan,an))=G(an). Since K is continuous, we can find a sequence {vn} which converges to wc0 and for every nN we have vn(,K(an)]. By continuity of p and q we have limnp(vn,q(vn,an))=p(wc0,q(wc0,c0))=G(c0).For every nN we have p(vn,q(vn,an))supp({(x,q(x,an))|x(,K(an)]})=G(an).Thus (9) G(c0)lim infnG(an).(9) Inequality (Equation9) along with |G(c0)G(an)|ϵ implies that there exists M>0 such that for every n>M we have G(c0)+ϵ2<G(an)=p(wan,q(wan,an)). From (Equation7), there exists x1R such that for every x<x1 and yR we have p(x,y)<G(c0)+ϵ2. Hence for every n>M we have x1wanK(an). This along with the fact that K is continuous, implies that there exists M>0 such that for every n>M we have x1wanK(c0)+1. Thus {wan} is bounded and has a convergent subsequence {wamn}. By the continuity of p and q we have (10) G(c0)+ϵ2limnG(amn)=limnp(wamn,q(wamn,amn))=p(b,q(b,c0)),(10) where b=limnwamn. But it is clear that we also have p(b,q(b,c0))supp({(x,q(x,c0))|x(,K(c0)]})=G(c0)which contradicts (Equation10). Therefore, G is continuous at c0. Since c0(,x0) is arbitrary we see that G is continuous on (,x0).

We will now combine Lemmas 2.1, 2.2 and 2.3 to show that two well-known maps from theoretical ecology have globally attracting and invariant 1-dimensional manifolds, and also determine when they are the invariant boundary of an invariant convex set.

3. Applications to ecological models

In this section, we use the above theory to prove the convexity of a unique non-compact carrying simplex in logarithmically scaled versions of the Leslie–Gower Model and Ricker models from theoretical ecology.

3.1. The Leslie–Gower model

The planar Leslie–Gower model [Citation7,Citation15] is defined by the Leslie–Gower map (11) F(u,v):=(ru1+u+αv,sv1+v+βu).(11) When r, s<1 and α,β>0, then (0,0) is globally asymptotically stable on R+2 (see [Citation7]). Hence, the system has no carrying simplex when r, s<1 and α,β>0 since no ΣR+2{0} can satisfy (CS2).

When r, s>1 the Leslie–Gower map has fixed points (12) (0,0),(r1,0),(0,s1)and, if positive, (α(s1)r+1αβ1,β(r1)s+1αβ1).(12) A number of authors [Citation1,Citation9,Citation11,Citation12] have shown that for r, s>1 and α,β>0, the model (Equation11) has a unique carrying simplex. In our approach, we use an alternative set of coordinates to those in (Equation11): We scale (Equation11) as follows (13) u=ex,v=ey,(13) to obtain the following log-scaled version of the model: (14) f(x,y):=(ln(r)+xln(1+ex+αey),ln(s)+yln(1+ey+βex)).(14) The only finite fixed point of the log-scale Leslie–Gower map is (15) (ln(α(s1)r+1αβ1),ln(β(r1)s+1αβ1)),(15) when the expressions are real.

We wish to study the invariant subsets of f:R2R2, and to this end we define X0:=R2,Xn:=f(Xn1)¯,n=1,2,3,and finally (16) X:=n=0Xn.(16) We recall that for a non-empty set ARd, the ω-limit set of A for a (continuous) map T:RdRd is defined to be ωT(A):=n=0m=nTm(A)¯, (e.g. [Citation20]). In this context X is actually ωf(R2) for f given by (Equation14), but we do not have compactness of any fn(R2) to apply standard results (e.g. [Citation20, Theorem 2.11]) for ω-limit sets to conclude that X is non-empty and invariant. Instead we prove these facts directly in a series of lemmas below.

The first lemma is needed because standard theorems on non-empty intersections of decreasing sequences of compact sets cannot be applied here as our sets Xn are not compact.

Lemma 3.1

When r, s>1 we have X.

Proof.

Suppose that ζx,y and ηx,y are defined as follows: ζx,y:=1αyysrx+αβxyys,ηx,y:=1βxxrsy+αβxyxr.Since lim(x,y)(0,0)(ζx,y,ηx,y)=(1r,1s)(1,1), there exists x,yR++ such that for every (0,0)(x,y)(x,y) we have ζx,y<1, ηx,y<1. Hence if S=(0,x]×(0,y], for every (x,y)(x,y) we have (xζx,y,yηx,y)S. It can also be easily verified that F(xζx,y,yηx,y)=(x,y). Thus SF(S) and if we define S:={(ln(x),ln(y))|(x,y)S}, then Sf(S). It means that for n=0,1,2, we have SXn. Therefore we have SX.

In the following, we rely strongly on the fact that f in (Equation14) is invertible.

Lemma 3.2

For the log Leslie–Gower map (Equation14) the sets X=ωf(R2) and X are non-empty and invariant.

Proof.

It is well-known (e.g. [Citation20]) that, for a given AR2 and a continuous map f:AR2, the omega-limit set ω(A) is closed and forward-invariant under f. Therefore, f(X)=f(ωf(R2))ωf(R2)=X.

To prove Xf(X), suppose for the sake of contradiction that there exists xX such that xf(X). For every m=0,1,2, we have xXm+1=f(Xm)¯ so that for every m=0,1,2, there exists sequence {yn(m)}Xm such that limnf(yn(m))=x.

It can be easily proven that X1=f(R2)¯(,ln(r))×(,ln(s)). Hence, for every m=1,2, we have {yn(m)}Xm(,ln(r))×(,ln(s)).

There must be a,bR such that {yn(m)}(a,ln(r))×(b,ln(s)), because otherwise {f(yn(m))} would not be bounded below which is a contradiction to the fact that {f(yn(m))} is convergent.

Now from the fact that {yn(m)} is bounded, we deduce that it has a limit point y. Since {yn(m)}Xm and Xm is closed, we have yXm. And from limnf(yn(m))=x, we deduce f(y)=x. Now, since f is invertible, y is the only point whose value of f is equal to x. It means that for every m=1,2, we would have the same y for that x. Therefore, for every m=1,2,, we have yXm, thus yX and x=f(y)f(X). This proves Xf(X). And along with the fact that X is forward invariant, we have X=f(X) and X is invariant.

As f is a diffeomorphism, both f and f1 map the interior of X into itself. Hence the interior of X is invariant. As X is also invariant, X must be invariant.

Lemma 3.3

For any r,s,α,β>0, x,yR2, xy and 0<λ<1 we have (17) f(λx+(1λ)y)λf(x)+(1λ)f(y).(17)

Proof.

Define ξ1:[0,1]R as follows ξ1(λ):=f1(λx+(1λ)y)If x=(x1,x2) and y=(y1,y2), then we have ξ1(λ)=(x1y1)2eλx1+(1λ)y1+α(x2y2)2eλx2+(1λ)y2(1+eλx1+(1λ)y1+αeλx2+(1λ)y2)2α(x1y1x2+y2)2eλx1+(1λ)y1eλx2+(1λ)y2(1+eλx1+(1λ)y1+αeλx2+(1λ)y2)2<0.Hence ξ1 is strictly concave. Similarly ξ2:[0,1]R defined by ξ2(λ):=f2(λx+(1λ)y) is also strictly concave. The inequality (Equation17) is now a direct result of the strict concavity of ξ1 and ξ2 and the following facts: ξi(0)=fi(y),ξi(1)=fi(x),i=1,2.

Lemma 3.4

(a) Let XR2 be the set of all points on or under the graph of a continuous strictly decreasing function B:(,b)R. Then for the log-scaled Leslie–Gower map f=(f1,f2) in (Equation14) we have {x|(x,c)f(X)}={f1({(x,g(x)+h(c))|x(,K(c))}),c(,y),otherwisewhere y=sup{f2(x,y)|(x,y)f(X)}, and g(x)=ln(1+βex)h(c)=cln(sec)K(c)=H1(cln(sec))and H is the invertible continuous function defined by H(x)=B(x)ln(1+βex).(b) We have {x|(x,c)f(R2)}={(,ln(r1+ecαβsec)),c(,ln(s)),otherwisewhere h and g are as defined in part (a).

Proof.

(a) For c(,y) we have {x|(x,c)f(X)}=f1(X{(x,y)R2|f2(x,y)=c})=f1(X{(x,y)R2|y=cln(sec)+ln(1+βex)})=f1({(x,y)R2|y=cln(sec)+ln(1+βex),yB(x)})=f1({(x,ln(1+βex)+cln(sec))|cln(sec)+ln(1+βex)B(x)})=f1({(x,g(x)+h(c))|cln(sec)B(x)ln(1+βex)})=f1({(x,g(x)+h(c))|cln(sec)H(x)}).Since B is strictly decreasing, H is strictly decreasing and invertible. Hence, {x|(x,c)f(X)}=f1({(x,g(x)+h(c))|x(,H1(cln(sec)))}).(b) For c(,ln(s)) we have {x|(x,c)f(R2)}=f1({(x,y)R2|f2(x,y)=c})=f1({(x,y)R2|y=cln(sec)+ln(1+βex)})=f1({(x,ln(1+βex)+cln(sec))|xR})=f1({(x,g(x)+h(c))|xR}),={f1(x,g(x)+h(c))|xR},={ln(r)+xln(1+ex+αeg(x)+h(c))|xR},={ln(r)+xln(1+ex+αeln(1+βex)+cln(sec))|xR},={ln(r)+xln(1+αecsec+ex(1+ecαβsec))|xR}.Since c(,ln(s)), sec is positive. Hence, 1+αecsec>0 and 1+ecαβsec>0 which implies that the derivative of the function w(x)=ln(r)+xln(1+αecsec+ex(1+ecαβsec)) is always positive. Thus {ln(r)+xln(1+αecsec+ex(1+ecαβsec))|xR}=(,limx+w(x))=(,ln(r1+ecαβsec)).

Combining the previous lemmas together we obtain.

Theorem 3.1

For any r, s>1 and α,β>0, for the log-scaled Leslie–Gower map (Equation14), the set X defined by (Equation16) is convex and invariant. Moreover, X is invariant and attracts R2.

Proof.

It is clear that X0 is convex. We use induction to prove that for n=1,2,, Xn is convex, from which it follows that their intersection X is convex. To prove convexity of X1, first we observe that by Lemma 3.4(b) for every c(,ln(s)) we have {x|(x,c)X1}={x|(x,c)f(R2)}¯=(,ln(r1+ecαβsec)].Now A(c):=ln(r1+ecαβsec) and X=X1 satisfy the conditions stated for A in Lemma 2.1.

So far, we have proven the existence of A which satisfies the conditions of Lemma 2.1 for X=X1. But to apply Lemma 1 we also need to prove the existence of the second function B of that lemma. Indeed it is easy to check that B is given by B(c)=A1(c)=ln(s1+ecαβrec). Hence by Lemma 2.1, X1 is the set of all points on or under the graph of a continuous strictly decreasing function. It is obvious that for every x,yX1 and 0<λ<1 we have f(λx+(1λ)y)f(R2)=X1. Moreover by Lemma 3.3 we have λf(x)+(1λ)f(y)f(λx+(1λ)y). Therefore, for every x,yX1 and 0<λ<1 there exists z=f(λx+(1λ)y)X1 such that λx+(1λ)yz. Now since X1 satisfies the conditions of Lemma 2.2, we deduce that X1 is convex.

Assume that for n1, Xn is convex and that it is the set of all points on or under the graph of a continuous strictly decreasing function B:(,b)R. By Lemma 3.4(a), for every c(,y) we have {x|(x,c)Xn+1}={x|(x,c)f(Xn)}=f1({(x,g(x)+h(c))|x(,K(c))}),where g, h and K are as defined in Lemma 3.4. The functions p(x,y):=f1(x,y), q(x,y):=g(x)+h(y) and K satisfy the conditions of Lemma 2.3, so that for every c(,y) we have Ωc=(,G(c)], where G:(,y)R is continuous. Thus Xn+1=c(,y)Ωc=c(,y)(,G(c)].A: = G and X:=Xn+1 satisfy the conditions stated for A in Lemma 2.1. Owing to the symmetric structure of the definition of the log-scaled Leslie–Gower map we can prove the existence of B which satisfies the conditions of Lemma 2.1 for X=Xn+1. Therefore, Lemma 2.1 shows that Xn+1 is the set of all points on or under the graph of a continuous strictly decreasing function.

Suppose that x,yXn+1. Since the sequence {Xn} is decreasing, we have x,yXn. Now since Xn is convex, for every 0<λ<1 we have λx+(1λ)yXn, thus f(λx+(1λ)y)f(Xn)=Xn+1. By Lemma 3.3 we have λf(x)+(1λ)f(y)f(λx+(1λ)y). Therefore, for every x,yXn+1 and 0<λ<1 there exists z=f(λx+(1λ)y)Xn+1 such that λx+(1λ)yz. Since Xn+1 satisfies the conditions of Lemma 2.2, we deduce that Xn+1 is convex. We conclude that X is convex.

X is invariant by Lemma 3.2. We know that X is the graph of a concave and strictly decreasing function A. By invariance, f1(A1(y),y)=A1(f2(A1(y),y)), i.e. A1(y)+lnrln(1+αeA1(y)+ey)=A1(y+lnsln(1+ey+βeA1(y))). As A1 is strictly decreasing and bounded above, limyA1(y)=x and invariance yields x+lnrln(1+ex)=x, so x=ln(r1). A similar argument shows that limxA(x)=ln(s1).

To show that X attracts R2, first we show that any finite fixed point of (Equation14) must belong to X. As there can be at most one finite fixed point, if P=(P1,P2) is a finite fixed point not in X then X contains no finite fixed point and dynamics on X is monotone. On X we may consider the one-dimensional dynamics xn+1=f1(xn,A1(xn)) or yn+1=f2(A(yn),yn). Suppose xn when n: 0>xn+1xn=lnrln(1+exn+αeA1(xn))lnrln(1+α(s1))so we need r1<α(s1). On the other hand, 0<yn+1yn=lnsln(1+eyn+βeA(yn))lnsln(1+β(r1))so we also need s1>β(r1). The pair of conditions r1<α(s1) and s1>β(r1) are incompatible with existence of a finite fixed point (Equation15) of (Equation14). A similar contradiction is obtained when the dynamics is monotone increasing in xn. Hence we conclude that whenever a finite fixed point exists for (Equation14) it must belong to X.

Next we recall (e.g. [Citation7]) that all non-trivial dynamics for the unscaled Leslie–Gower map converge to a fixed point which can be (r1,0), (0,s1), or (α(s1)r+1αβ1,β(r1)s+1αβ1) when it is positive. Hence any orbit of (Equation11) not convergent to the positive fixed point must converge to (r1,0) or (0,s1).

Now consider an orbit (xn,yn) of (Equation14) that does not converge to a finite fixed point. Then by convergence of Leslie–Gower orbits, (exn,eyn) tends to (r1,0) or (0,s1) as n. Suppose (exn,eyn) tends to (r1,0) as n. Then xnln(r1), yn as n. On the other hand, A1(yn)ln(r1) as n. Thus (xn,yn)(A1(yn),yn)=|xnA1(yn)|0 as n. Hence in this case we have (xn,yn)X as n. The case (exn,eyn) tends to (0,s1) is similar.

Finally for the case that an orbit (xn,yn) of (Equation14) converges a positive fixed point P, as we showed in the previous paragraph PX.

Thus we conclude that X is attracting.

3.2. The Ricker model

The planar Ricker model is defined by the non-invertible map (18) F(u,v):=(ueruαv,vesvβu),(u,v)R+2(18) where α,β,r,s>0.

With the coordinates stated in (Equation13), we have the following log-scaled version of the model: (19) f(x,y):=(x+rexαey,y+seyβex).(19) For the time being we work with the Ricker map in these standard coordinates to see when we can expect a carrying simplex to be unique when it exists. Log coordinates will be introduced later. The following points are always fixed points of F: (20) c1:=(0,0),c2:=(r,0),c3:=(0,s).(20) When αβ1 and c4 defined in (Equation21) below is a member of R++2, then F has exactly four fixed points and the fourth fixed point is: (21) c4:=(rαβ1,rβsαβ1).(21) We will find it more convenient to now use the log-scaled version (Equation19). We define Y0:=R2,X0:=Q3,where Q3={(x,y):x0 and y0} is the third quadrant. We let Yn:=f(Yn1),Xn:=f(Xn1)¯,n=1,2,and define the sets (22) Y:=n=0Yn=n=0fn(R2),X:=n=0Xn=n=0fn(Q3)¯.(22)

Lemma 3.5

If s, r<1 then Y=X.

Proof.

It is obvious that XY. To prove that we also have YX, it is sufficient to prove that for any (x,y)R2, there exists nN such that fn(x,y)Q3.

If x>0 then we have f1(x,y)x=rexαey<r1<0,and if y>0 then f2(x,y)y=seyβey<s1<0.Therefore, if we define n as follows, n={0x, y01+x1rx>0, y01+y1sy>0, x01+max{x1r,y1s}x, y>0,then fn(x,y)Q3.

Lemma 3.6

When r, s>0 we have X.

Proof.

x,y<0 can be found such that for every (x1,y1)(x,y) the following equations have at least one solution with (x,y)(x1,y1): x+rexαey=x1,y+seyβex=y1.

Hence if S=(,x]×(,y] then Sf(S). Therefore for n=0,1,2, we have SXn and SX.

Now we will show that X is invariant. Since the log-scaled Ricker map f is not invertible, to show invariance of X we need a property weaker than invertibility. We use the fact that the log-scaled Ricker map f is a proper map (i.e. for every compact set XR2, f1(X) is compact). To see this, note that if {xn} is a sequence such that |xn|, then, according to the terms in (Equation19), |f(xn)|. Since f is continuous, we conclude that for every closed and bounded set X, f1(X) is closed and bounded. Therefore, for every compact set X, f1(X) is compact and we conclude that the log-scaled Ricker map f is proper.

Lemma 3.7

When 0<r, s<1 and f is the log-scaled Ricker map, X defined by (Equation22) is invariant.

Proof.

If xf(X) then xf(Xn)f(Xn)¯=Xn+1 for nN. Hence xn=1Xn and since X0=Q3, we have xn=1Xn=n=0Xn=X. This proves f(X)X.

We prove that f(Xn)¯=f(Xn) for nN. It is sufficient to show that f(Xn) is closed if Xn is closed. Suppose that x is a limit point of f(Xn). There exists a sequence {ymn}Xn such that limmf(ymn)=x. {ymn} is bounded since, being proper, f maps unbounded subsets of X0 into unbounded sets whereas {f(ymn)} is convergent, and hence bounded. Boundedness of {ymn}Xn and the fact that Xn is closed, imply that {ymn} has a limit point yXn. Continuity of f implies f(y)=x. Thus xf(Xn), which proves that f(Xn) is closed and hence f(Xn)¯=f(Xn).

Now if xX, then xXn+1=f(Xn)¯=f(Xn) for nN. Hence for nN there exists znXn such that f(zn)=x. Since f is proper, f1({x}) is compact. Therefore, since {zn}f1({x}), {zn} is bounded and has at least one limit point. Let z be such a limit point. z is also a limit point of f1({x}), and since f1({x}) is closed, we have zf1(x). Thus f(z)=x and since z is a limit point of {zn}, and for nN we have znXn where {Xn} is a decreasing sequence of closed sets, we conclude that zX. This proves Xf(X).

Lemma 3.8

Lemma 3.3 also holds for the log-scaled Ricker map f defined in (Equation19).

Proof.

We define χ1:[0,1]R and χ2:[0,1]R as follows χ1(λ):=f1(λx+(1λ)y)χ2(λ):=f2(λx+(1λ)y).It is easy to show that both χ1 and χ2 are strictly concave (being the sum of linear minus exponential terms), and the rest of the proof is straightforward.

Lemma 3.9

(a) Let bR and XQ3 be the set of all points on or under the graph of a continuous strictly decreasing function B:(,b)R. Let f=(f1,f2) denote the log-scaled Ricker map. Then we have {x|(x,c)f(X)}={f1({(x,P(cs+βex))|x(,K(c)]}),c(,y),otherwisewhere y=sup{f2(x,y)|(x,y)f(X)}, P(x):=xW0(ex),W0 is the principal branch of the Lambert W function (see, for example, [Citation16]), K(c):=G1(cs)and G is the continuous invertible function defined by G(x):=B(x)eB(x)βex.(b) We have {x|(x,c)f(Q3)}={f1({(x,P(cs+βex))|x(,K(c)]}),c(,y),otherwisewhere P is the function defined in part (a) and y=sup{f2(x,y)|(x,y)f(Q3)},K(c)=min{0,ln(1c+sβ)}.

Proof.

(a) For c(,y) we have {x|(x,c)f(X)}=f1(X{(x,y)Q3|f2(x,y)=c})=f1(X{(x,y)Q3|y+seyβex=c})=f1(X{(x,y)Q3|yey=cs+βex})When t<0, tet is strictly increasing. Hence, for every lR, tet=l has at most one solution for t0. Since {tet|t0}=(,1], tet=l has a unique solution for t0 and l(,1] and no solution when t0 and l>−1. We claim that t=P(l) is the unique solution for tet=l when t0 and l(,1]. To prove that, we have (23) P(l)eP(l)=lW0(el)elW0(el).(23) By the properties of the Lambert W function W0 we have W0(el)eW0(el)=el. Hence elW0(el)=W0(el). This along with (Equation23) implies P(l)eP(l)=l (since t=P(l) must satisfy t<0, only the principal branch of the Lambert W function can be used to provide a solution). Now with l=cs+βex and t = y we have f1(X{(x,y)Q3|yey=cs+βex})=f1(X{(x,y)Q3|y=P(cs+βex)})=f1({(x,y)Q3|y=P(cs+βex),yB(x)})=f1({(x,P(cs+βex))|P(cs+βex)B(x)}).Since P(cs+βex)0, B(x)0 in the above sets, and since when t<0, tet is strictly increasing, we have P(cs+βex)B(x)cs+βex=P(cs+βex)eP(cs+βex)B(x)eB(x).Thus f1({(x,P(cs+βex))|P(cs+βex)B(x)})=f1({(x,P(cs+βex))|cs+βexB(x)eB(x)})=f1({(x,P(cs+βex))|csB(x)eB(x)βex})=f1({(x,P(cs+βex))|csG(x)}).(b) For c(,s1) we have {x|(x,c)f(Q3)}=f1({(x,y)Q3|f2(x,y)=c})=f1({(x,y)Q3|y+seyβex=c})=f1({(x,y)Q3|yey=cs+βex})=f1({(x,y)Q3|y=P(cs+βex)})=f1({(x,P(cs+βex))|x(,0],P(cs+βex)(,0]})=f1({(x,P(cs+βex))|x(,0],P(cs+βex)eP(cs+βex)1})=f1({(x,P(cs+βex))|x0,cs+βex1})=f1({(x,P(cs+βex))|x0,xln(1c+sβ)})=f1({(x,P(cs+βex))|xmin{0,ln(1c+sβ)}}).

Lemma 3.10

For any 0<r, s<1 and α,β>0, and f is the log-scaled Ricker map, X defined in (Equation22) is invariant and convex.

Notice that Lemma 3.10 is not a complete analogue of Theorem 3.1 since we are not claiming that X is necessarily invariant. We will address this after proving Lemma 3.10.

Lemma 3.10. Convexity of X can be proved with a similar argument to that used for the scaled Leslie–Gower model. So we explain the argument more briefly. It is obvious that X0 is convex, and by using induction we can prove convexity of Xn for n=1,2,, and that implies convexity of X.

To prove convexity of X1, by Lemma 3.9(b) for c(,y) we have {x|(x,c)f(X1)}={x|(x,c)f(Q3)}=f1({(x,P(cs+βex))|x(,K(c)]})where P, K and y are defined in that part of the lemma. Now it is easy to verify that p=f1 and q:R2R defined by q(x,y)=P(ys+βex) and K satisfy the conditions of Lemma 2.3. By Lemma 2.3 for every c(,y) we have Ωc=(,G(c)], where G:(,y)R is continuous. Thus X1=c(,y)Ωc=c(,y)(,G(c)].Now A:=G,X:=X1 satisfy the conditions stated for A in Lemma 2.1.

Owing to the symmetric structure of the definition of the log-scaled Ricker map, we can state similar lemmas to prove that there exists B such that it satisfies the conditions of Lemma 2.1 for X=X1. Now since, by Lemma 2.1, X1 is the set of all points on or under the graph of a continuous strictly decreasing function, we can use Lemma 2.2 and Lemma 3.8 with a similar argument to that used in Theorem 1 to prove that X1 is convex.

Assume that for n1, Xn is convex and it is the set of all points on or under the graph of a continuous strictly decreasing function B:(,b)R. By Lemma 3.9(a) for every c(,y) we have {x|(x,c)Xn+1}={x|(x,c)f(Xn)}=f1({(x,P(cs+βex))|x(,K(c))}).Then p:=f1 and q:R2R defined by q(x,y):=P(ys+βex) satisfy the conditions of Lemma 2.3 and G defined in that lemma is continuous. So A: = G and X:=Xn+1 satisfy the conditions stated for A in Lemma 2.1. Again, owing to the symmetric structure of the definition of the log-scaled Ricker map we can prove the existence of B which satisfies the conditions of Lemma 2.1 for X=Xn+1. Therefore, by that lemma, Xn+1 is the set of all points on or under the graph of a continuous strictly decreasing function. Now we can use Lemmas 2.2 and 3.8 and a similar argument to that used in Theorem 3.1 to prove that Xn+1 is convex.

According to Lemma 3.7 X is invariant, and as the intersection of convex sets it is convex.

As the log-scaled Ricker map f is not invertible we cannot conclude that X is also invariant. In order to prove that XX1 is invariant, it is sufficient to show that the restriction f|X1f(X1) is invertible. If the Jacobian of f is non-vanishing throughout X1 then f is locally invertible at any point of X1. But as is well known, locally invertibility does not always imply global invertibility. Ho [Citation10] proved that a local homeomorphism between a pathwise connected Hausdorff space and a simply connected Hausdorff space is a global homeomorphism if and only if that map is proper. We have already established that the log-scaled Ricker map f is proper (in the paragraph preceding Lemma 3.7). Hence, if we prove that the Jacobian of f does not vanish anywhere in X1 for a given range of parameters, then we can deduce that XX1 is invariant for that same range of parameters.

Thus now we consider where the Jacobian vanishes.

Using [Citation6] (which studies the unscaled Ricker map (Equation18)) the Jacobian of the log-scaled Ricker map f only vanishes on LC1 defined by (24) LC1:={(x,y)Q3:1ex=ey(1(1αβ)ex)}.(24) Set q(t)=1et(αβ1)et+1.When αβ1, then q(t)>0 if and only if t<0. If αβ<1, then q(t)>0 if and only if t(,0)(ln(1αβ),+). In this case, LC1 is the union of two connected curves. By Lemma 3.5, if r, s<1, then X=YQ3. So in this case we only need to consider LC11:={(ln(q(t)),t)|t<0},and investigate whether or not the Jacobian vanishes at some points on X.

According to [Citation6], Y1:=f(R2) is bounded by the set of points on or under LC01 defined by (25) LC01:={(ln(q(t))+rq(t)αet,t+setβq(t))|t<0}.(25) Since XY1, X is a subset of the set of points on or under LC01. Hence, if r, s<1 and LC11 does not intersect that space, then the Jacobian of f does not vanish anywhere in X since LC1X=.

Lemma 3.11

If r, s<1 then LC11 does not intersect the set of points on or under LC01. Hence, if r, s<1 then X is invariant.

Proof.

It is sufficient to show that if (x1,y)LC11 and (x0,y)LC01, then x1>x0. Since (x0,y)LC01, for some t<0 we have (x0,y)=(ln(q(t))+rq(t)αet,t+setβq(t)). Since x1=ln(q(y)) and y=t+setβq(t), we have x1=ln(q(t+setβq(t))).We define R(t):=x1x0=ln(q(t+setβq(t)))(ln(q(t))+rq(t)αet). We have (26) q(t)(eR(t)1)=q(t+setβq(t))er+q(t)+αetq(t).(26) It is easy to show that when h<0, then q(h)>0. We use this fact multiple times in this proof. From t<0 we have tet<1. This along with s<1 and q(t)>0 implies t+setβq(t)<0. Thus q(t+setβq(t))>0. Now since er+q(t)+αet1r+q(t)+αet, we have (27) q(t+setβq(t))er+q(t)+αetq(t+setβq(t))(1r+q(t)+αet).(27) Now (Equation26) and (Equation27) imply (28) q(t)(eR(t)1)q(t+setβq(t))(1r+q(t)+αet)q(t).(28) For the sake of expressing equations in a simpler way, let T:=et. We have 0<T<1, q(t)=q(ln(T))=1T(αβ1)T+1 and we can rewrite (Equation28) as follows (29) q(t)(eR(t)1)q(ln(T)+sTβq(ln(T)))(1r+q(ln(T))+αT)q(ln(T)).(29) We have (30) q(ln(T)+sTβq(ln(T)))=1eln(T)+sTβq(ln(T))(αβ1)eln(T)+sTβq(ln(T))+1=1TesTβ1T(αβ1)T+1(αβ1)TesTβ1T(αβ1)T+1+1=1αβTesTβ1T(αβ1)T+1(αβ1)TesTβ1T(αβ1)T+1+1=1αβT(αβ1)T+es+T+β1T(αβ1)T+1.(30) As we mentioned before, ln(T)+sTβq(ln(T))=t+setβq(t)<0. Hence, from αβ1>1, we deduce (αβ1)TesTβ1T(αβ1)T+1+1=(αβ1)eln(T)+sTβq(ln(T))+1>0,thus (31) (αβ1)T+es+T+β1T(αβ1)T+1=es+T+β1T(αβ1)T+1((αβ1)TesTβ1T(αβ1)T+1+1)>0.(31) From T>0, s<1 and q(t)>0 we have (32) (αβ1)T+1+(s+T+β1T(αβ1)T+1)=αβT+1s+βq(t)>0.(32) Since eh>1+h for hR, we have (33) (αβ1)T+es+T+β1T(αβ1)T+1(αβ1)T+1+(s+T+β1T(αβ1)T+1).(33) Now (Equation31), (Equation32) and (Equation33) imply 1αβT(αβ1)T+es+T+β1T(αβ1)T+11αβTαβT+1s+β1T(αβ1)T+1.This along with (Equation29) and (Equation30) implies q(t)(eR(t)1)(1αβTαβT+1s+β1T(αβ1)T+1)×(1r+q(ln(T))+αT)q(ln(T))=(1αβT((αβ1)T+1)(αβT+1s)((αβ1)T+1)+β(1T))×(1r+1T(αβ1)T+1+αT)1T(αβ1)T+1=(1s)((αβ1)T+1)+β(1T)(αβT+1s)((αβ1)T+1)+β(1T)×(1r+1T(αβ1)T+1+αT)1T(αβ1)T+1=E(T)(αβ1)T+1,where E(T):=(1s)((αβ1)T+1)+β(1T)(αβT+1s)((αβ1)T+1)+β(1T)×((1r+αT)((αβ1)T+1)+1T)(1T)=(1s)((αβ1)T+1)+β(1T)(αβT+1s)((αβ1)T+1)+β(1T)×((1r+αT)((αβ1)T+1)+1T)(αβT+1s)((αβ1)T+1)(1T)+β(1T)2(αβT+1s)((αβ1)T+1)+β(1T)=(1s)((αβ1)T+1)(1r+αT)((αβ1)T+1)+β(1T)(1r+αT)((αβ1)T+1)(αβT+1s)((αβ1)T+1)+β(1T)+(1s)((αβ1)T+1)(1T)(αβT+1s)((αβ1)T+1)(1T)(αβT+1s)((αβ1)T+1)+β(1T).Now by the above inequalities we have q(t)(eR(t)1)(1s)(1r+αT)((αβ1)T+1)+β(1T)(1r+αT)(αβT+1s)((αβ1)T+1)+β(1T)+(1s)(1T)(αβT+1s)(1T)(αβT+1s)((αβ1)T+1)+β(1T)=(1s)(1r+αT)((αβ1)T+1)+β(1T)(1r)(αβT+1s)((αβ1)T+1)+β(1T).From s<1, r<1, αβ1>1 and 0<T<1 we can deduce that (1s)(1r+αT)((αβ1)T+1)+β(1T)(1r)(αβT+1s)((αβ1)T+1)+β(1T)>0.Therefore, q(t)(eR(t)1)>0. Now since q(t)>0, we deduce that eR(t)1>0, which implies x1x0=R(t)>0. This proves x1>x0. Now since X is a subset of the set of points on or under LC01, LC11 does not intersect X and by the argument we stated before, we deduce that X is invariant.

We may now put together Lemmas 3.10 and 3.11 to obtain the analogue of Theorem 3.1 for the Ricker map:

Theorem 3.2

For any 0<r, s<1 and α,β>0 and f is the log-scaled Ricker map, X defined in (Equation22) is invariant and convex, X is invariant and attracts R2.

Proof.

All that is left to do is show that X is attracting. It is proven that if r, s<2, then every non-trivial orbit converges to one of the non-zero fixed points (see [Citation5]). The possible non-zero fixed points on the x or y axis are the same as for the Leslie–Gower model with r−1 replaced by r and s−1 replaced by s. So we may use the same method as used for the log-scaled Leslie–Gower map to show attraction to X.

From this we obtain the following improvement on the known conditions r+s<1+rs(1αβ)<2,(e.g. [Citation8,Citation9,Citation17,Citation18]) for the existence of a carrying simplex for the Ricker model. These inequalities fail for some α,β>0, when r, s<1, namely those α,β that satisfy r, s<1 and αβ(11/r)(11/s). However, Lemma 3.11 also shows that the unscaled Ricker map is retrotone on [0,r]×[0,s] when r, s<1 so we obtain

Corollary 3.1

When r, s<1 and α,β>0 the Ricker map (x,y)R+2(xerxαy,yesyβx) has a (compact) carrying simplex.

Proof.

In the absence of asymptotic completeness in Theorem 3.2, we apply standard results on retrotone systems (e.g. [Citation9,Citation11,Citation17,Citation18]).

Disclosure statement

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

References

  • S. Baigent, Convexity of the carrying simplex for discrete-time planar competitive Kolmogorov systems, J. Differ. Eq. Appl. 22(5) (2016), pp. 609–620.
  • S. Baigent, Convexity of the carrying simplex for discrete-time planar competitive Kolmogorov systems, J. Differ. Eq. Appl. 22(5) (2016), pp. 1–14.
  • S. Baigent, Carrying simplices for competitive maps, in Difference Equations, Discrete Dynamical Systems and Applications. ICDEA 2017, S. Elaydi, C. Pötzsche, and A. Sasu, eds., Springer Proceedings in Mathematics and Statistics, Vol. 287. Springer, Cham.
  • S. Baigent, Convex geometry of the carrying simplex for the May–Leonard map, Discrete Contin. Dyn. Syst. B 24(4) (2019), pp. 1697–1723.
  • S. Baigent, Z. Hou, S. Elaydi, E.C. Balreira, and R. Luís, A global picture for the planar Ricker map: Convergence to fixed points and identification of the stable/unstable manifolds, J. Differ. Eq. Appl. 29(5) (2023), pp. 575–591.
  • E.C. Balreira, S. Elaydi, and R. Luís, Local stability implies global stability for the planar Ricker competition model, Discrete Contin. Dyn. Syst. Ser. B 19(2) (2014), pp. 323–351.
  • J.M. Cushing, S. Levarge, N. Chitnis, and S.M. Henson, Some discrete competition models and the competitive exclusion principle, J. Differ. Eq. Appl. 10 (2004), pp. 1139–1151.
  • M. Gyllenbeg, J. Jiang , L. Niu, and P. Yan, On the dynamics of multi-species Ricker models admitting a carrying simplex, J. Differ. Eq. Appl. 25(11) (2019), pp. 1489–1530.
  • M.W. Hirsch, On existence and uniqueness of the carrying simplex for competitive dynamical systems, J. Biol. Dyn. 2(2) (2008), pp. 169–179.
  • C.W. Ho, A note on proper maps, Proc. Am. Math. Soc. 51 (1975), pp. 237–241.
  • Z. Hou, On existence and uniqueness of a modified carrying simplex for discrete Kolmogorov systems, J. Differ, Eq. Appl. 27(2) (2021), pp. 284–315.
  • J. Jiang and L. Niu, The theorem of the carrying simplex for competitive system defined on the n-rectangle and its application to a three-dimensional system, Int. J. Biomath. 07(06) (2014), pp. 1450063–1450012.
  • J. Jiang and L. Niu, On the equivalent classification of three-dimensional competitive Leslie–Gower models via the boundary dynamics on the carrying simplex, J. Math. Biol. 74(5) (2016), pp. 1–39.
  • U. Krause, Positive Dynamical Systems in Discrete Time: Theory, Models, and Applications, De Gruyter, Germany, 2015.
  • P.H. Leslie and J.C. Gower, The properties of a stochastic model for two competing species, Biometrika45(3/4) (1958), pp. 316–330.
  • I. Mezo, The Lambert W Function: Its Generalizations and Applications, CRC Press, Oxford, 2022.
  • J. Mierczyński and S. Baigent, Existence of the carrying simplex for a retrotone map, J. Differ. Equ. Appl. 10(3) (2024), pp. 287–319. doi:10.1080/10236198.2023.2285394
  • A. Ruiz-Herrera, Exclusion and dominance in discrete population models via the carrying simplex, J. Differ. Eq. Appl. 19(1) (2013), pp. 96–113.
  • H.L. Smith, Planar competitive and cooperative difference equations, J. Differ. Eq. Appl. 3(5-6) (1998), pp. 335–357.
  • H.L. Smith and H.R. Thieme, Dynamical Systems and Population Persistence, Graduate Studies in Mathematics Vol. 118, American Mathematical Society, Providence, RI, 2011.
  • M.L. Zeeman, Hopf bifurcations in competitive three dimensional Lotka–Volterra systems, Dyn. Stab. Syst. 8 (1993), pp. 189–217.