Skip to main content
Log in

On heteroclinic cycles of competitive maps via carrying simplices

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

We concentrate on the effects of heteroclinic cycles and the interplay of heteroclinic attractors or repellers on the boundary of the carrying simplices for low-dimensional discrete-time competitive systems. Based on the existence of the carrying simplex for the competitive mapping, we provide the criteria on stability of the heteroclinic cycle. This result can be seen as a discrete counterpart of that for the continuous-time systems. Several concrete discrete-time competition models are further analyzed, which do admit heteroclinic cycles. The criteria on the stability of the heteroclinic cycle for each model are also given, which are comparable with the corresponding continuous-time models.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  • Afraimovich VS, Rabinovich MI, Varona P (2004) Heteroclinic contours in neural ensembles and the winnerless competition principle. Int J Bifurc Chaos 14:1195–1208

    Article  MathSciNet  MATH  Google Scholar 

  • Allen LJS, Allen EJ, Atkinson DN (1999) Integrodifference equations applied to plant dispersal, competition, and control. In: Ruan S, Wolkowicz G, Wu J (eds) Differential equations with applications to biology fields institute communications. American Mathematical Society, RI, pp 15–30

    Google Scholar 

  • Baigent S (2013) Geometry of carrying simplices of 3-species competitive Lotka–Volterra systems. Nonlinearity 26:1001–1029

    Article  MathSciNet  MATH  Google Scholar 

  • Chi CW, Hsu SB, Wu LI (1998) On the asymmetric May–Leonard model of three competing species. SIAM J Appl Math 58:211–226

    Article  MathSciNet  MATH  Google Scholar 

  • Chow SN, Hale JK (1982) Methods of bifurcation theory. Springer, New York

    Book  MATH  Google Scholar 

  • Cushing JM, Levarge S, Chitnis N, Henson SM (2004) Some discrete competition models and the competitive exclusion principle. J Differ Equ Appl 10:1139–1151

    Article  MathSciNet  MATH  Google Scholar 

  • Diekmann O, Wang Y, Yan P (2008) Carrying simplices in discrete competitive systems and age-structured semelparous populations. Discret Contin Dyn Syst 20:37–52

    MathSciNet  MATH  Google Scholar 

  • Dold A (1972) Lectures on algebraic topology. Springer, Berlin

    Book  MATH  Google Scholar 

  • Feng B (1998) The heteroclinic cycle in the model of competition between \(n\) species and its stability. Acta Math Appl Sin 14:404–413

    Article  MATH  Google Scholar 

  • Garay BM, Hofbauer J (2003) Robust permanence for ecological differential equations, minimax, and discretizations. SIAM J Math Anal 34:1007–1039

    Article  MathSciNet  MATH  Google Scholar 

  • Gyllenberg M, Hanski I, Lindström T (1997) Continuous versus discrete single species population models with adjustable reproductive strategies. Bull Math Biol 59:679–705

    Article  MATH  Google Scholar 

  • Gyllenberg M, Yan P, Wang Y (2006) A 3D competitive Lotka–Volterra system with three limit cycles: a falsification of a conjecture by Hofbauer and So. Appl Math Lett 19:1–7

    Article  MathSciNet  MATH  Google Scholar 

  • Gyllenberg M, Yan P (2009) Four limit cycles for a three-dimensional competitive Lotka–Volterra system with a heteroclinic cycle. Comput Math Appl. 58:649–669

    Article  MathSciNet  MATH  Google Scholar 

  • Gyllenberg M, Yan P (2009) On the number of limit cycles for three dimensional Lotka–Volterra systems. Discret Contin Dyn Syst Ser B 11:347–352

    Article  MathSciNet  MATH  Google Scholar 

  • Hale JK (1988) Asymptotic behavior of dissipative systems. American Mathematical Society, Providence

    MATH  Google Scholar 

  • Hirsch MW (1988) Systems of differential equations which are competitive or cooperative: III. Competing species. Nonlinearity 1:51–71

    Article  MathSciNet  MATH  Google Scholar 

  • Hirsch MW (2008) On existence and uniqueness of the carrying simplex for competitive dynamical systems. J Biol Dyn 2:169–179

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Hutson V, Jansen W (1987) Coexistence for systems governed by difference equations of Lotka–Volterra type. J Math Biol 25:553–570

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer J (1994) Heteroclinic cycles in ecological differential equations. Tatra Mt Math Publ 4:105–116

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Sigmund K (1988) The theory of evolution and dynamical systems. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Hofbauer J, So JW-H (1994) Multiple limit cycles for three dimensional Lotka–Volterra equations. Appl Math Lett 7:65–70

    Article  MathSciNet  MATH  Google Scholar 

  • Hou Z, Baigent S (2013) Heteroclinic limit cycles in competitive Kolmogorov systems. Discret Contin Dyn Syst 33:4071–4093

    Article  MathSciNet  MATH  Google Scholar 

  • Hutson V, Moran W (1982) Persistence of species obeying difference equations. J Math Biol 15:203–213

    Article  MathSciNet  MATH  Google Scholar 

  • Jiang J, Mierczyński J, Wang Y (2009) Smoothness of the carrying simplex for discrete-time competitive dynamical systems: a characterization of neat embedding. J Differ Equ 246:1623–1672

    Article  MATH  Google Scholar 

  • Jiang J, Niu L, Zhu D (2014) On the complete classification of nullcline stable competitive three-dimensional Gompertz models. Nonlinear Anal Real World Appl 20:21–35

    Article  MathSciNet  MATH  Google Scholar 

  • Jiang J, Niu L (2016) On the equivalent classification of three-dimensional competitive Atkinson/Allen models relative to the boundary fixed points. Discret Contin Dyn Syst 36:217–244

    MathSciNet  Google Scholar 

  • Kon R, Takeuchi Y (2003) Permanence of 2-host 1-parasitoid systems. Dyn Contin Discret Impuls Syst Ser B Appl Algorithms 10:389–402

    MathSciNet  MATH  Google Scholar 

  • Leslie PH, Gower JC (1958) The properties of a stochastic model for two competing species. Biometrika 45:316–330

    Article  MathSciNet  MATH  Google Scholar 

  • Lu Z, Luo Y (2003) Three limit cycles for a three-dimensional Lotka–Volterra competitive system with a heteroclinic cycle. Comput Math Appl 46:231–238

    Article  MathSciNet  MATH  Google Scholar 

  • May RM, Leonard WJ (1975) Nonlinear aspects of competition between three species. SIAM J Appl Math 29:243–253

    Article  MathSciNet  MATH  Google Scholar 

  • May RM, Oster GF (1976) Bifurcations and dynamic complexity in simple ecological models. Am Nat 110:573–599

    Article  Google Scholar 

  • Mickens RE (1994) Nonstandard finite difference models of differential equations. World Scientific, Singapore

    MATH  Google Scholar 

  • Mierczyński J (1994) The \(C^1\)-property of carrying simplices for a class of competitive systems of ODEs. J Differ Equ 111:385–409

    Article  MATH  Google Scholar 

  • Mierczyński J (1999) On smoothness of carrying simplices. Proc Am Math Soc 127:543–551

    Article  MATH  Google Scholar 

  • Ricker WE (1954) Stock and recruitment. J Fish Res Board Can 5:559–623

    Article  Google Scholar 

  • Roeger L-IW (2005) Discrete May–Leonard competition models II. Discret Contin Dyn Syst Ser B 5:841–860

    Article  MathSciNet  MATH  Google Scholar 

  • Roeger L-IW, Allen LJS (2004) Discrete May–Leonard competition models I. J Differ Equ Appl 10:77–98

    Article  MathSciNet  MATH  Google Scholar 

  • Ruiz-Herrera A (2013) Exclusion and dominance in discrete population models via the carrying simplex. J Differ Equ Appl 19:96–113

    Article  MathSciNet  MATH  Google Scholar 

  • Schuster P, Sigmund K, Wolff R (1979) On \(\omega \)-limits for competition between three species. SIAM J Appl Math 37:49–54

    Article  MathSciNet  MATH  Google Scholar 

  • Shen W, Wang Y (2008) Carrying simplices in nonautonomous and random competitive Kolmogorov systems. J Differ Equ 245:1–29

    Article  MATH  Google Scholar 

  • Smith HL (1986) Periodic competitive differential equations and the discrete dynamics of competitive maps. J Differ Equ 64:165–194

    Article  MATH  Google Scholar 

  • Smith HL (1998) Planar competitive and cooperative difference equations. J Differ Equ Appl 3:335–357

    Article  MATH  Google Scholar 

  • Van Den Driessche P, Zeeman ML (1998) Three-dimensional competitive Lotka–Volterra systems with no periodic orbits. SIAM J Appl Math 58:227–234

    Article  MathSciNet  MATH  Google Scholar 

  • Wang Y, Jiang J (2001) The general properties of discrete-time competitive dynamical systems. J Differ Equ 176:470–493

    Article  MATH  Google Scholar 

  • Wang Y, Jiang J (2002) Uniqueness and attractivity of the carrying simplex for discrete-time competitive dynamical systems. J Differ Equ 186:611–632

    Article  MATH  Google Scholar 

  • Wang R, Xiao D (2010) Bifurcations and chaotic dynamics in a 4-dimensional competitive Lotka–Volterra system. Nonlinear Dyn 59:411–422

    Article  MathSciNet  MATH  Google Scholar 

  • Xiao D, Li W (2000) Limit cycles for the competitive three dimensional Lotka–Volterra systems. J Differ Equ 164:1–15

    Article  MathSciNet  MATH  Google Scholar 

  • Zeeman ML (1993) Hopf bifurcations in competitive three-dimensional Lotka–Volterra systems. Dyn Stab Syst 8:189–217

    MathSciNet  MATH  Google Scholar 

  • Zeeman EC, Zeeman ML (2002) An n-dimensional competitive Lotka–Volterra system is generically determined by the edges of its carrying simplex. Nonlinearity 15:2019–2032

    Article  MathSciNet  MATH  Google Scholar 

  • Zeeman EC, Zeeman ML (2002) From local to global behavior in competitive Lotka–Volterra systems. Trans Am Math Soc 355:713–734

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgments

The authors are greatly indebted to several anonymous referees for very careful reading, pointing out some errors in the original manuscript and providing lots of other very inspiring and helpful comments and suggestions which led to much improvement of the earlier version of this paper. The authors are also very grateful to Professor O. Diekmann for his valuable suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lei Niu.

Additional information

Dedicated to Professor Mats Gyllenberg on the occasion of his 60th birthday.

This work is supported by the National Natural Science Foundation of China (NSFC) under Grant No. 11371252, Research and Innovation Project of Shanghai Education Committee under Grant No. 14zz120, and the Program of Shanghai Normal University (DZL121). Y.W. is partially supported by NSF of China Nos. 11371338, 11471305 and the Fundamental Research Funds for the Central Universities.

Appendix: Existence of the carrying simplex

Appendix: Existence of the carrying simplex

Let \(W\subseteq \mathbf {R}^n\) be an open neighborhood containing \(\mathbf {R}^n_+\), and let \(T:W(\subseteq \mathbf {R}^n)\rightarrow \mathbf {R}^n\) be a \(C^1\)-map given by the form (1) so that \(T\mathbf {R}^n_+\subseteq \mathbf {R}^n_+\) and \(T^{-1}(\mathbf {R}^n_+)\subseteq \mathbf {R}^n_+\). For any subset \(A\subseteq \mathbf {R}^n_+\), we denote by \(\mathrm {Int}A\) (resp. \(\partial A\)) the interior (resp. boundary) of A relative to \(\mathbf {R}^n_+\) for short.

We further assume the following properties of T:

\(({{\varUpsilon }}\mathbf{0})\) :

\(G_i(0)>1\) for any \(i\in N\); and moreover, \(G_i(x)>0\) whenever \(x=(x_1,\ldots ,x_n)\in \mathbf {R}^n_+\) with \(x_i>0\).

\(({{\varUpsilon }}\mathbf{1})\) :

For each i, \(T|_{\dot{H}_{\{i\}}^+}\) has a globally attracting fixed point \(q_{\{i\}}=q_ie_{\{i\}}\) (called an axial fixed point), and set \(q=\sum q_{\{i\}}=(q_1,\ldots ,q_n)\).

\(({{\varUpsilon }}\mathbf{2})\) :

There exists a lower order convex subset S with \([0,q]\subseteq S \subseteq \mathbf {R}_+^n,\) which satisfies that \(T(\mathbf {R}_+^n)\subseteq S\) and

$$\begin{aligned} Tx<_I Ty \Longrightarrow x\ll _I y,\quad \text {whenever }x\in S^+_I, y\in \dot{S}^+_I\quad \text {and}\quad \emptyset \ne I\subseteq N. \end{aligned}$$
\(({{\varUpsilon }}\mathbf{3})\) :

For any \(x,y\in [0,q]\) with \(y=Tx\), \([0,y]\subseteq T[0,x]\).

\(({{\varUpsilon }}\mathbf{4})\) :

For any \(x\in [0,q]\cap H_I^+\), the matrix \([\partial G_i/\partial x_j(x)]_{i,j\in I}\) has strictly negative entries.

Our main result in the Appendix is the following:

Theorem 6

(Carrying Simplex) Assume that \(({\varUpsilon }0\))–\(({\varUpsilon }4)\) hold. Then \(T|_{\mathbf {R}^n_+}\) admits a compact global attractor \({\varGamma }\), which satisfies the following property:

  1. (i)

    \({\varGamma }\) is lower order convex, i.e., \([0,x]\subseteq {\varGamma }\) whenever \(x\in {\varGamma };\)

  2. (ii)

    \(\partial {\varGamma }\) is a positively invariant compact set and is homeomorphic to the \((n-1)\)-dimensional standard probability simplex \({\varDelta }^{n-1}\) via radial projection; 

  3. (iii)

    \(\mathrm{Int}{\varGamma }\subseteq \{y\in {\varGamma }:\alpha (y)=\{0\}\), for any negative orbit in \({\varGamma }\) through \(y\};\)

  4. (iv)

    For any \(x\in \mathbf {R}^n_+{\setminus }\{0\}\), \(\omega (x)\subseteq \partial {\varGamma }\).

Assume further that T is locally injective in an open neighborhood V of [0, q] relative to W (hence relative to \(\mathbf {R}^n\)). Then the above statements (i)–(iv) remain true even if \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) and

\(({{\varUpsilon }}\mathbf{4'})\) :

For any \(x\in [0,q]\cap H_I^+\), the diagonal entries of \([\partial G_i/\partial x_j(x)]_{i,j\in I}\) are strictly negative, while the off-diagonal ones are non-positive

hold. Moreover, \(\partial {\varGamma }\) is invariant,

$$\begin{aligned} \mathrm{Int}{\varGamma }=\{y\in {\varGamma }:\alpha (y)=\{0\},\quad \text {for any negative orbit in }{\varGamma }\text { through }y \} \end{aligned}$$

and for any \(x\in \mathbf {R}^n_+{\setminus } \{0\}\), there exists some \(y\in \partial {\varGamma }\) such that

$$\begin{aligned} \Vert T^kx-T^ky\Vert \rightarrow 0\quad \text { as }\,k\rightarrow +\infty . \end{aligned}$$
(17)

We will break the proof of the above theorem into several Lemmas.

Lemma 2

If conditions \(({\varUpsilon }1)\) and \(({\varUpsilon }2)\) hold, then \(T[0,q]\subseteq [0,q]\). Moreover, for any bounded set \(U\subseteq \mathbf {R}^n_+\), the positive orbit \(O^+(U)\) is bounded and \(\omega (U)\subseteq [0,q]\).

Proof

The proof is very similar to the arguments in Smith (1986, Lemma 3.2 and Proposition 3.4). For the sake of completeness, we provide the details here. Given any positive number \(a>0\) and \(i\in N\) with \(ae_{\{i\}}\in S\), we claim that, for any \(y=(y_1,\ldots ,y_n)\in S\), if \(y_i\leqslant a\), then \(T_i(y) \le T_i(ae_{\{i\}})\). Indeed, if such \(y\in H_{\{i\}}^+\), then clearly \(T_i(y)\le T_i(ae_{\{i\}})\) (Otherwise, one has \(T_i(y)>T_i(ae_{\{i\}})\). By (\({\varUpsilon }2\)), we obtain that \(a<y_i\), a contradiction). If such y satisfies that its ith coordinate \(y_i=0\), then obviously \(T_i(y)=0<T_i(ae_{\{i\}})\). As a consequence, one can now assume that such y satisfies \(y_i>0\) and \(y_j>0\), for some \(j\ne i\). Let \(I\subseteq N\) be the nonempty index set such that \(y\in \dot{S}^+_I\). Then \(ae_{\{i\}}\in S^+_I\). Suppose that \(T_i(y)> T_i(ae_{\{i\}})\). Then it is easy to see that \(Ty>_I T(ae_{\{i\}})\), so by (\({\varUpsilon }\)2) we have \(y\gg _I ae_{\{i\}}\), and hence \(a<y_i\), which contradicts \(y_i\leqslant a\). Thus we have proved the claim.

Now, for any \(y\in [0,q]\), one can use the claim above to obtain that \(T_i(y)\leqslant q_i\) because \(y_i\leqslant q_i\). So \(T[0,q]\subseteq [0,q]\).

Moreover, for any bounded set \(U\subseteq \mathbf {R}_+^n\), let \(c_i=\sup \{(Ty)_i:y\in U\}<+\infty \) for each i. Without loss of generality, we assume that \(c_i>0\) for all i. Recall that \(T(\mathbf {R}_+^n)\subseteq S\) and S is lower order convex, so \(c_i e_{\{i\}}\in S\). Then one can repeatedly utilize the claim above to obtain that

$$\begin{aligned} T^{k+1}_i(y)\leqslant T^{k}_i(c_i e_{\{i\}})\quad \text { for any }y\in U\quad \text {and}\quad k=0,1,\ldots . \end{aligned}$$
(18)

Recall that \(T^k_i(c_i e_{\{i\}})\rightarrow q_i\) as \(k\rightarrow \infty \). Then there is an integer \(k_0>0\) such that \(T^{k+1}_i(y)\le q_i+1\) for any \(k\ge k_0\) and \(y\in U\). This implies that the positive orbit \(O^+(U)\) is bounded. Furthermore, by (18) and \(T^k_i(c_i e_{\{i\}})\rightarrow q_i\) as \(k\rightarrow \infty \) again, we obtain that \(\omega (U)\subseteq [0,q]\).

Proposition 1

Let \(({\varUpsilon }1)\)\(({\varUpsilon }2)\) hold. Let also \({\varGamma }:=\bigcap ^{\infty }_{k=1}T^k[0,q]\subseteq [0,q]\subseteq S\). Then we have

  1. (i)

    \({\varGamma }\) is a non-empty compact invariant set which contains all the axial fixed points \(q_{\{i\}}, 1\leqslant i\leqslant n\). Moreover, \(\omega (U)\subseteq {\varGamma }\) for any bounded set \(U\subseteq \mathbf {R}^n_+\), and \({\varGamma }\) is a global attractor for \(T|_{\mathbf {R}^n_+};\)

  2. (ii)

    If additionally \(({\varUpsilon }3)\) holds, then \({\varGamma }\) is lower order convex, i.e., \([0,x]\subseteq {\varGamma }\) whenever \(x\in {\varGamma };\)

  3. (iii)

    If \(({\varUpsilon }1)\)\(({\varUpsilon }3)\) hold, then \(\partial {\varGamma }\) is a compact positively invariant set. Furthermore, \(\partial {\varGamma }\) is weakly unordered in the sense that, for any \(\emptyset \ne I\subseteq N\), there do not exist two points \(x_1,x_2\in \partial {\varGamma }\cap H_I^+\) such that \(x_1\ll _I x_2\).

Proof

(i) By Lemma 2, \(T[0,q]\subseteq [0,q]\). Since \(T^k[0,q]\) is compact and \(0\in {\varGamma }\ne \emptyset \), \({\varGamma }\) is a nonempty compact set. Note also that

$$\begin{aligned} T{\varGamma }=T(\cap _kT^k[0,q])\subseteq \cap _k T(T^k[0,q])={\varGamma }. \end{aligned}$$

Then \({\varGamma }\) is positively invariant under T. On the other hand, given any \(x\in {\varGamma }\) and each integer \(l\ge 1\), there exists \(x_l\in [0,q]\), such that \(x=T^lx_l=T(T^{l-1}x_l)\). Because \(T^{l-1}x_l\in [0,q]\) for any l, we may assume without loss of generality that \(T^{l-1}x_l\rightarrow y\in [0,q]\). Moreover, for a fixed k, \(T^{l-1}x_l\in T^k[0,q]\) for any \(l\ge k+2\), which implies that \(y\in T^k[0,q]\). Consequently, \(y\in {\varGamma }\) and \(x=Ty\), that is, \(T{\varGamma }={\varGamma }\). Furthermore, for any bounded \(U\subseteq \mathbf {R}_+^n\), it follows from \(\omega (U)\subseteq [0,q]\) (by Lemma 2) and the total invariance of \(\omega (U)\) that \(\omega (U)\subseteq \cap _kT^k[0,q]\), i.e., \(\omega (U)\subseteq {\varGamma }\).

Since \(T:\mathbf {R}_+^n\rightarrow \mathbf {R}_+^n\) is continuous and point-dissipative, it follows from Hale (1988, Theorem 2.4.6) that there is a global attractor \({\varGamma }_1\) of T in \(\mathbf {R}_+^n\). We will prove that \({\varGamma }\) is the global attractor by showing that \({\varGamma }={\varGamma }_1\). In fact, for the bounded set \({\varGamma }\) obtained above, one has \(\omega ({\varGamma })\subseteq {\varGamma }_1\). Recall that \(T{\varGamma }={\varGamma }\). Then \({\varGamma }=\omega ({\varGamma })\subseteq {\varGamma }_1\). On the other hand, one also has \(\omega ({\varGamma }_1)\subseteq {\varGamma }\) by the above result, which implies that \({\varGamma }_1=\omega ({\varGamma }_1)\subseteq {\varGamma }\). Thus, \({\varGamma }={\varGamma }_1\), which implies that \({\varGamma }\) is the global attractor for \(T|_{\mathbf {R}^n_+}\).

(ii) For any \(x\in {\varGamma }\) and integer \(k\ge 0\), the definition of \({\varGamma }\) implies that there exists \(y_k\in [0,q]\) such that \(T^k y_k=x\). If additionally (\({\varUpsilon }\)3) holds, then

$$\begin{aligned} {[0,x]}=[0,T^k y_k]\subseteq T^k[0,y_k]\subseteq T^k[0,q], \end{aligned}$$

which implies that \([0,x]\subseteq {\varGamma }\), i.e., \({\varGamma }\) is lower order convex.

(iii) Clearly, \({\varGamma }\) is compact. We now show that \(\partial {\varGamma }\) is positively invariant. Without loss of generality, we assume that the interior \(\mathrm{Int}{\varGamma }\ne \emptyset \) (otherwise, \(\partial {\varGamma }={\varGamma }\) is invariant). Suppose that there is an \(x\in \partial {\varGamma }\) such that \(Tx\in \mathrm{Int}{\varGamma }\). Then one can find some \(y\in {\varGamma }\cap \dot{\mathbf {R}}^n_+\), nearby Tx, such that \(Tx\ll y\). Since \({\varGamma }\) is invariant, we choose some \(w\in {\varGamma }\) with \(Tw=y\). By (\({\varUpsilon }\)2), it follows that \(x\ll w\). Then the lower order convexity of \({\varGamma }\) implies that \(x\in \mathrm{Int}{\varGamma }\), a contradiction. Thus, \(\partial {\varGamma }\) is positively invariant.

Finally, by the lower order convexity of \({\varGamma }\), it is clear that there do not exist two points \(x_1,x_2\in \partial {\varGamma }\) such that \(x_1\ll x_2\). Moreover, for each \(\emptyset \ne I\subseteq N\), note that the hypotheses (\({\varUpsilon }\)2)–(\({\varUpsilon }\)3) still hold with \(\mathbf {R}_+^n\) and q replaced by \(H_I^+\) and \(q_I\), respectively. Here the i-th coordinate \((q_I)_i\) of \(q_I\) satisfies that \((q_I)_i=q_i\) if \(i\in I\) and \((q_I)_i=0\) otherwise. Then one can repeat the arguments above to obtain the compact invariant set \({\varGamma }_I\) satisfying (i)–(iii) in \(H_I^+\). Furthermore, it is easy to see that \({\varGamma }_I={\varGamma }\cap H_I^+\) and \(\partial _{H_I^+}{\varGamma }_I=\partial {\varGamma }\cap H_I^+\). In particular, there do not exist two points \(x_1,x_2\in \partial {\varGamma }\cap H_I^+\) such that \(x_1\ll _I x_2\).

Remark 6

By the definition of \({\varGamma }\), it is clear that \(p=(p_1,\ldots ,p_n)\in {\varGamma }\) implies that \(p_i\le q_i\) for any \(i\in N\). Furthermore, we claim that if the coordinate \(p_i=q_i\) for some i, then \(p=q_{\{i\}}\). (Otherwise, let \(I\subseteq N\) be the index set such that \(p\in \dot{H}_I^+\). Then \(p>_Iq_{\{i\}}\). Since \({\varGamma }\) is invariant, there is a \(p^*\in {\varGamma }\) such that \(Tp^*=p>_Iq_{\{i\}}=Tq_{\{i\}}\). By virtue of (\({\varUpsilon }\)2), \(p^*\gg _I q_{\{i\}}\), which entails that \(p^*\notin [0,q]\), a contradiction to \({\varGamma }\subseteq [0,q]\).)

By virtue of this claim, one can obtain that \(\partial {\varGamma }\) is equal to the boundary \(\partial _{[0,q]}{\varGamma }\) of \({\varGamma }\) relative to [0, q], that is, \(\partial {\varGamma }=\partial _{[0,q]}{\varGamma }\). (Indeed, it is clear that \(\partial _{[0,q]}{\varGamma }\subseteq \partial {\varGamma }\). Suppose that there is some \(z\in \partial {\varGamma }{\setminus } \partial _{[0,q]}{\varGamma }\). Then \(z\ne q_{\{i\}}\) for any i, because \(q_{\{i\}}\in \partial _{[0,q]}{\varGamma }\). Noticing also that \(z\in \partial {\varGamma }\subseteq {\varGamma }\), the claim above then yields that \(z\ll q\). So any small neighborhood of z relative to \(\mathbf {R}^n_+\) is also a neighborhood of z relative to [0, q]. Thus, \(z\in \partial {\varGamma }\) implies that \(z\in \partial _{[0,q]}{\varGamma }\), contradicting to \(z\notin \partial _{[0,q]}{\varGamma }\).) Therefore, one also has \(\mathrm {Int}{\varGamma }=\mathrm {Int}_{[0,q]}{\varGamma }\). As a consequence, we hereafter always write \(\mathrm {Int}{\varGamma }\) and \(\partial {\varGamma }\) relative to [0, q] without any confusion.

The following remark is about the assumption (\({\varUpsilon }\)0).

Remark 7

The first statement of (\({\varUpsilon }\)0) guarantees that, for any index set \(\emptyset \ne I\subseteq N\), \(x\ll _I Tx\) whenever \(x\in \dot{H}_I^+\) is sufficiently close to 0. As a consequence, we hereafter always fix some \(0\ll x_0\in [0,q]\) nearby 0, such that, \(w\ll _I Tw\) whenever \(w\in [0,x_0]\cap \dot{H}_I^+\) for any \(I\ne \emptyset \). In particular, the origin 0 is the unique fixed point of T in \([0,x_0]\).

Meanwhile, the second statement of (\({\varUpsilon }\)0) implies that \(T^{-1}(H_I^+)\subseteq H_I^+\) for any \(I\subseteq N\). In particular, \(T^{-1}(\{0\})=\{0\}\).

Lemma 3

Let \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) hold. Then \(\mathrm{Int}{\varGamma }\ne \emptyset \). Moreover, for any \(x\in \dot{H}^+_I\) with \(I\ne \emptyset \), \(x\in \mathrm{Int}{\varGamma }\) if and only if \(x\ll _Iw_x\), where \(w_x\in \partial {\varGamma }\) is the (unique) point of intersection of the ray \(L_{x}:=\{tx:t\ge 0\}\) with \(\partial {\varGamma }\).

Proof

Let \(x_0\gg 0\) be in Remark 7. Clearly, \(x_0\ll Tx_0\) and \(x_0\le q\). Combining with (\({\varUpsilon }\)3), we obtain that \([0,x_0]\subseteq [0,Tx_0]\subseteq T[0,x_0]\subseteq T[0,q]\). So,

$$\begin{aligned}{}[0,x_0]\subseteq T[0,x_0]\subseteq \cdots \subseteq T^m[0,x_0]\subseteq T^m[0,q] \end{aligned}$$
(19)

for any \(m\ge 0\). Hence, \([0,x_0]\subseteq {\varGamma }\), which implies that

$$\begin{aligned} \left[ 0,\frac{x_0}{2}\right] \subseteq \mathrm{Int}{\varGamma }. \end{aligned}$$
(20)

Thus, \(\mathrm{Int}{\varGamma }\ne \emptyset \).

Let \(x\in \dot{H}^+_I\) with \(I\ne \emptyset \). Note that \({\varGamma }\) is bounded and \(tx\in [0,x_0]\subseteq {\varGamma }\) for small \(t\ge 0\). Then there exists at least one point \(w_x\in \partial {\varGamma }\cap L_x\). By the non-ordering property of \(\partial {\varGamma }\) in Proposition 1(iii), \(w_x\) is unique. If \(x\in \mathrm{Int}{\varGamma }\), then it is easy to see that \(x\ll _I w_x\) (otherwise, \(w_x\ll _I x\). The lower order convexity of \({\varGamma }\) in Proposition 1(ii) then implies that \(w_x\in \mathrm{Int}{\varGamma }\), a contradiction). On the other hand, if \(x\ll _I w_x\), then we will show \(x\in \mathrm{Int}{\varGamma }\). Suppose \(x\notin \mathrm{Int}{\varGamma }\). Then the weakly unordered property of \(\partial {\varGamma }\) implies that \(x\notin {\varGamma }\), and hence \(w_x\notin {\varGamma }\) by the lower order convexity of \({\varGamma }\), contradicting that \(w_x\in \partial {\varGamma }\). Thus, we have proved that \(x\in \mathrm{Int}{\varGamma }\) if and only if \(x\ll _Iw_x\).

Lemma 4

Let \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) hold. Let also \(B_0:=\bigcup _{k\ge 0}T^k[0,\frac{x_0}{2}]\). Then

  1. (i)

    \(B_0\) is an invariant and lower order convex subset of \({\varGamma };\)

  2. (ii)

    \(B_0= \{y\in {\varGamma }\): there exists a negative orbit in \({\varGamma }\) of y such that \(\alpha (y)=\{0\} \};\)

  3. (iii)

    \(\partial B_0\) is compact, positively invariant and weakly unordered, i.e., for any \(\emptyset \ne I\subseteq N\), there do not exist two points \(x_1,x_2\in \partial B_0\cap H_I^+\) such that \(x_1\ll _I x_2\).

Proof

(i) Clearly, \(B_0\) is positively invariant. By Remark 7 and (\({\varUpsilon }\)3), one has that \([0,x_0/2]\subseteq T[0,x_0/2]\subseteq \cdots \subseteq T^m[0,x_0/2]\) for any \(m\ge 1\), which implies that \(B_0\) is negatively invariant, hence \(B_0\) is invariant. Moreover, \(B_0\subseteq {\varGamma }\) is immediate from (20) and the invariance of \({\varGamma }\). For any \(y\in B_0\), there exist some \(k\ge 1\), \(v\in [0,\frac{x_0}{2}]\) such that \(y=T^kv\). So, by (\({\varUpsilon }\)3), \([0,y]\subseteq [0,T^kv]\subseteq T^k[0,v]\subseteq T^k[0,\frac{x_0}{2}]\subseteq B_0,\) which implies that \(B_0\) is lower order convex.

(ii) Set \(M=\{y\in {\varGamma }\): there exists a negative orbit in \({\varGamma }\) of y such that \(\alpha (y)=\{0\} \}\). Given any \(y\in M\), let \(\{y(-m),m\ge 0\}\) be a negative orbit in \({\varGamma }\) through y such that \(\alpha (y)=\{0\}\). Then \(y(-m)\in [0,\frac{x_0}{2}]\) for some m sufficiently large. So, \(y=T^my(-m)\in T^m[0,\frac{x_0}{2}]\), which implies that \(M\subseteq B_0\). On the other hand, for any \(v \in [0,x_0]\) and any negative orbit \(\{v(-m),m\ge 0\}\) in \({\varGamma }\) through v, it follows from (\({\varUpsilon }\)2) that \(\{v(-m):m\ge 1\}\) is a strictly-decreasing sequence in \([0,x_0]\). By Remark 7, it entails that \(v(-m)\rightarrow 0\) as \(m\rightarrow \infty \) and hence, \([0,x_0]\subseteq M\). Now for any \(y\in B_0\), there exist some \(k\ge 1\) and \(v\in [0,\frac{x_0}{2}]\) such that \(T^kv=y\). So, \(\{v(-m):m\ge 0\}\cup \{Tv,T^2v,\ldots ,T^{k-1}v,y\}\) forms a negative orbit in \({\varGamma }\) through y, where \(\{v(-m)\}\) is a negative orbit in \({\varGamma }\) through v. Since \(v(-m)\rightarrow 0\), such a negative orbit through y possesses a trivial \(\alpha (y)=\{0\}\). Thus, \(B_0\subseteq M\), and we have proved that \(M=B_0\).

(iii) Clearly, \(\overline{B_0}\) is also invariant by continuity of T and \(\overline{B_0}\subseteq {\varGamma }\). We now show that \(\partial B_0\) is positively invariant. Note that \(\mathrm{Int}B_0\ne \emptyset \). Suppose that there is an \(x\in \partial B_0\) such that \(Tx\in \mathrm{Int}B_0\). Then one can find some \(y\in B_0\cap \dot{\mathbf {R}}^n_+\) so close to Tx that \(Tx\ll y\). Since \(B_0\) is invariant, we choose some \(w\in B_0\) with \(Tw=y\). Then \({\varUpsilon }\)2) implies that \(x\ll w\). Hence \(x\in \mathrm{Int}B_0\) by the lower order convexity of \(B_0\), a contradiction. Thus, \(\partial B_0\) is positively invariant. The proof of the remaining statement is similar as in Proposition 1(iii). We omit it here.

Lemma 5

Assume that \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) hold. Then for any \(y\in {\varGamma }{\setminus }\{0\}\) and any negative-orbit \(\{y(-k),k\ge 0\}\) through y, the \(\alpha \)-limit set \(\alpha (y)\) associated to \(\{y(-k)\}\) satisfies: either (i) \(\alpha (y)\cap [0,\frac{x_0}{2}]=\emptyset \), or (ii) \(\alpha (y)=\{0\}\).

Proof

Choose some \(\emptyset \ne I\subseteq N\) such that \(y\in \dot{H}_I^+\). By the second statement of (\({\varUpsilon }\)0) and form (1), we can write such a negative-orbit as \(\{y(-k):k\ge 0\}\subseteq \dot{H}_I^+\) such that \(y(0)=y\) and \(Ty(-k-1)=y(-k)\). Suppose that \(z\in \alpha (y)\cap [0,\frac{x_0}{2}]\ne \emptyset \). Then there exists a sequence \(k_m\rightarrow \infty \) such that \(y(-k_m)\in [0,x_0]\subseteq [0,q]\) for all \(k_m\) sufficiently large. So, it follows from the first statement of Lemma 2 that \(\{y(-k):k\ge 0\}\subseteq [0,q]\), and hence, \(\{y(-k):k\ge 0\}\subseteq \dot{S}_I^+\).

Choose some \(k_*\in \{k_m\}\) such that \(y(-k_*)\in [0,x_0]\) (and hence, by Remark 7, \(y(-k_*)\ll _I Ty(-k_*)=y(-k_*+1)\). So for any \(k\ge k_*\), one has \(T^{k-k_*}y(-k)=y(-k_*)\ll _I Ty(-k_*)=T^{k-k_*}y(-k+1)\). Since \(\{y(-k):k\ge 0\}\subseteq \dot{S}_I^+\), (\({\varUpsilon }2\)) yields that \(y(-k)\ll _I y(-k+1)\). Thus \(\{y(-k):k\ge 0\}\) is an eventually decreasing sequence, which implies that \(\alpha (y)\) is a fixed point of T in \([0,x_0]\). So \(\alpha (y)=\{0\}\).

Lemma 6

Assume that \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) hold. Then for any \(y\in \mathbf {R}^n_+{\setminus }\{0\}\), \(\omega (y)\cap [0,\frac{x_0}{2}]=\emptyset \).

Proof

By Proposition 1, we already know that \(\omega (y)\subseteq {\varGamma }\) for any \(y\in \mathbf {R}^n_+{\setminus }\{0\}\). Suppose that there are some \(y_0\in \dot{H}_I^+(I\ne \emptyset )\) and a sequence \(k_m\rightarrow \infty \) such that \(T^{k_m}y_0\rightarrow z\in \omega (y_0)\cap [0,\frac{x_0}{2}]\ne \emptyset \). By Remark 7 again, \(T^{k_m}y_0\ll _I T^{k_m+1}y_0\) for all \(k_m\) sufficiently large. So, for each \(k\ge 1\), one can find a \(k_m\ge k\) such that \(T^{k_m-k}(T^ky_0)=T^{k_m}y_0\ll _I T^{k_m+1}y_0=T^{k_m-k}(T^{k+1}y_0)\). It then follows from (\({\varUpsilon }\)2) that \(T^ky_0\ll _I T^{k+1}y_0\) (Since \(T^k y_0\in S\) as \(k \ge 1\)). Thus, \(\{T^ky_0:k\ge 0\}\) is an eventually increasing sequence in \({\varGamma }\), which implies that \(T^ky_0\rightarrow z\gg _I0\) as \(k\rightarrow \infty \). So z is a non-zero fixed point in \([0,\frac{x_0}{2}]\), contradicting Remark 7. We have proved the lemma.

Remark 8

Up to now, we would like to point out here that Proposition 1 (i.e., the existence of \({\varGamma }\), with \(\partial {\varGamma }\) being positively invariant) and Lemmas 36 were obtained only under the assumptions (\({\varUpsilon }0\))–(\({\varUpsilon }3\)). In order to show that \(\partial {\varGamma }\) attracts all non-trivial orbits, we need the further assumption (\({\varUpsilon }\)4) or (\({\varUpsilon }\)4’), by which one can obtain the following two technical lemmas, motivated by Shen and Wang (2008), which are crucial for the proof of the attractivity of \(\partial {\varGamma }\).

Lemma 7

Assume that \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) and \(({\varUpsilon }4')\) hold. Given \(\emptyset \ne I\subseteq N\), let \(x,y\in \dot{H}_I^+\cap [0,q]\) satisfying \(x \ll _I y\). Let also \(\{x(-m):m\ge 0\}\) and \(\{y(-m):m\ge 0\}\) be negative orbits of x and y in [0, q], respectively. If \(\liminf _{m\rightarrow \infty }||y(-m)||>0\), then

$$\begin{aligned} \liminf _{m\rightarrow +\infty }\Vert x(-m)-y(-m)\Vert >0. \end{aligned}$$
(21)

Proof

Since \(0\in H_I^+\), \(x,y\in \dot{H}_I^+\) and \(0\ll _I x\ll _I y\), by (\({\varUpsilon }\)0) and (\({\varUpsilon }\)2) we get that \(x(-m),y(-m)\in \dot{H}_I^+\) and \(0\ll _I x(-m)\ll _I y(-m)\) for any \(m\ge 1\).

For each \(i\in I\), let \(K_i(-m)=\dfrac{x_i(-m)}{y_i(-m)}\) for \(m=0,1,\ldots \). Then

$$\begin{aligned} K_i(-m)-K_i(-m+1)=K_i(-m)\cdot \frac{G_i(y(-m))-G_i(x(-m))}{G_i(y(-m))}, \end{aligned}$$

for \(m\ge 1\). Here the denominator is nonzero due to \(0\ll _I y(-m)\) and the second statement of (\({\varUpsilon }\)0). Recall that \(\{x(-m):m\ge 0\},\{y(-m):m\ge 0\}\subseteq [0,q]\) (where q is from (\({\varUpsilon }\)1)) and \([\partial G_i/\partial x_j]_{i,j\in I}\le 0\) on \([0,q]\cap H_I^+\) (see (\({\varUpsilon }\)4’)). Then we obtain that \(K_i(-m)\le K_i(-m+1)\). Consequently,

$$\begin{aligned} K_i(-m)\le \cdots \le K_i(-1)\le K_i(0)=\dfrac{x_i}{y_i}<1 \end{aligned}$$

for any \(m\ge 1\). Set \(K_*=\min _{i\in I}|K_i(0)-1|>0\). Then

$$\begin{aligned} \Vert x(-m)-y(-m)\Vert&=\sum _{i\in I}|x_i(-m)-y_i(-m)| \\&=\sum _{i\in I}|K_i(-m)-1|\cdot |y_i(-m)|\\&\ge K_*\sum _{i\in I}|y_i(-m)|=K_*\Vert y(-m)\Vert . \end{aligned}$$

So if \(\liminf _{m\rightarrow +\infty }\Vert y(-m)\Vert >0\), then we obtain that \(\liminf _{m\rightarrow +\infty }\Vert x(-m)-y(-m)\Vert >0\).

Lemma 8

Assume that \(({\varUpsilon }0)\)\(({\varUpsilon }4)\) hold. Given \(\emptyset \ne I\subseteq N\), let \(x,y\in \dot{H}_I^+\cap [0,q]\) satisfying \(x \ll _I y\). Let also \(\{x(-m):m\ge 0\}\) and \(\{y(-m):m\ge 0\}\) be negative orbits of x and y in [0, q], respectively. Assume further that there exists a sequence \(m_k\rightarrow \infty \) such that \(x(-m_k)\rightarrow x_*\) and \(y(-m_k)\rightarrow y_*\) as \(k\rightarrow \infty \). Then one has \(x_*=0\).

Proof

As the proof of Lemma 7, we have \(x(-m),y(-m)\in \dot{H}_I^+\) and \(0\ll _I x(-m)\ll _I y(-m)\in [0,q]\) for any \(m\ge 0\). Suppose that \(x_*\ne 0\). Then one has \(y_*\ge x_* \ne 0\). By Lemma 5, \(\alpha (y)\cap [0,\frac{x_0}{2}]=\emptyset \), hence \(\liminf _{m\rightarrow \infty }||y(-m)||>0\). Now, Lemma 7 entails that (21) holds, i.e., there is a real number \(d>0\) such that

$$\begin{aligned} \liminf _{m\rightarrow +\infty }\Vert x(-m)-y(-m)\Vert =d>0. \end{aligned}$$
(22)

As a consequence, \(x_*,y_*\in H_I^+\) and \(x_*<y_*\) with \(||x_*-y_*||\ge d>0\).

Thus, one can find nonempty index subsets \(I_*\subseteq I_1\subseteq I\) such that \(x_*\in \dot{H}_{I_*}^+\) and \(y_*\in \dot{H}_{I_1}^+\). Now, for each \(l\ge 1\), it follows from (\({\varUpsilon }\)0) that \(T^l x_*\in \dot{H}_{I_*}^+\), \(T^l y_*\in \dot{H}_{I_1}^+\). Moreover, we have \(T^l x_*=\lim _{k\rightarrow +\infty } T^l x(-m_k)=\lim _{k\rightarrow +\infty } x(l-m_k)\). Similarly, one has \(T^l y_*=\lim _{k\rightarrow +\infty } y(l-m_k).\) Note that \(x(l-m_k)\ll _I y(l-m_k)\in [0,q]\) for all \(m_k\) sufficiently large. Then it holds that \(T^l x_*\le T^l y_*\le q\). Again, by virtue of (22), \(T^l x_*< T^l y_*\) and \(\Vert T^l x_* - T^l y_*\Vert \ge d > 0\). Therefore, together with (\({\varUpsilon }\)2), we have obtained that \(0\ll _{I_*} T^l x_*\ll _{I_1} T^l y_*\le q\), for any \(l\ge 0\).

For each \( i\in I_1\) and \(l\ge 0\), define \(K_i(l):=\dfrac{(T^l x_*)_i}{(T^l y_*)_i}\). Obviously, \(0<K_i(l)<1\) for \(i\in I_*\), and \(K_i(l)=0\) for \(i\in I_1{\setminus } I_*\). Moreover,

$$\begin{aligned} K_i(l+1)-K_i(l)=K_i(l)\cdot \frac{G_i(T^{l}x_*)-G_i(T^{l}y_*)}{G_i(T^{l}y_*)}. \end{aligned}$$
(23)

Since \([\frac{\partial G_i}{\partial x_j}]_{i,j\in I_1}\le 0\) on \([0,q]\cap H_{I_1}^+\), one has \(K_i(l+1)\ge K_i(l)\). In particular, for any \(i\in I_*\ne \emptyset \), we have

$$\begin{aligned} K_i(l+1)-K_i(l)&=\frac{K_i(l)}{G_i(T^{l}y_*)}\sum ^n_{j=1}\int ^1_0\frac{\partial G_i}{\partial x_j}(sT^{l}x_*+(1-s)T^{l}y_*)ds\cdot (T^{l}x_*-T^{l}y_*)_j\\&\ge \frac{\delta _i\cdot K_i(0)}{G_i(T^{l}y_*)}\cdot \sum _{j\in I_1}(T^{l}y_*-T^{l}x_*)_j\\&= \frac{\delta _i\cdot K_i(0)}{G_i(T^{l}y_*)}\cdot ||T^ly_*-T^lx_*||\ge \frac{\delta _i\cdot K_i(0)\cdot d}{G_i(T^{l}y_*)}, \end{aligned}$$

where \(K_i(0)>0\) and \(\delta _i=\min \nolimits _{\begin{array}{c} j\in I_1\\ z\in [0,q] \end{array}}{|\frac{\partial G_i}{\partial x_j}(z)|}.\) By virtue of (\({\varUpsilon }\)4), it is clear that \(\delta _i>0\). Hence,

$$\begin{aligned} K_i(l)-K_i(0)=\sum ^{l-1}_{j=0}[K_i(j+1)-K_i(j)]\ge \dfrac{\delta _i\cdot K_i(0)\cdot d}{G_*}\cdot l, \end{aligned}$$

where \(G_*=\max _{x\in [0,q]}(\sum _{i\in I^*}G_i(x))>0\). Then

$$\begin{aligned} K_i(l)\ge K_i(0)+\dfrac{\delta _i\cdot K_i(0)\cdot d}{G_*}\cdot l\rightarrow +\infty \end{aligned}$$

as \(l\rightarrow \infty \), a contradiction to the fact that \(K_i(l)\le 1\) for any \(l\ge 0\). Thus, we have shown \(x_*=0\), which completes our proof.

Proposition 2

Assume that \(({\varUpsilon }0)\)\(({\varUpsilon }4)\) hold. Then we have the following two statements:

  1. (i)

    \(\mathrm{Int}{\varGamma }\subseteq \{y\in {\varGamma }:\alpha (y)=\{0\},\) for any negative orbit of y in \({\varGamma }\}\subseteq B_0\);

  2. (ii)

    For any \(x\in \mathbf {R}^n_+{\setminus }\{0\}\), \(\omega (x)\subseteq \partial {\varGamma }\).

Proof

(i) It suffices to show that any non-zero element of \(\mathrm{Int}{\varGamma }\) is contained in \(\{y\in {\varGamma }:\alpha (y)=\{0\},\text { for any negative orbit of }y \text { in }{\varGamma }\}\). For any such non-zero \(y\in \mathrm{Int}{\varGamma }\), let I be the nonempty index subset such that \(y\in \dot{H}_I^+\). Then it follows from Lemma 3 that \(y\ll _I w_y\), where \(w_y=\partial {\varGamma }\cap L_y\). Suppose that there is a negative orbit \(\{y(-k):k\ge 0\}\) in \({\varGamma }\) through y such that \(\alpha (y)\ne \{0\}\). Then Lemma 5 ensures that \(\alpha (y)\cap [0,\frac{x_0}{2}]=\emptyset \).

Note also that \(w_y\in \partial {\varGamma }\subseteq {\varGamma }\), the invariance of \({\varGamma }\) implies that there is a negative orbit \(\{w_y(-k):k\ge 0\}\subseteq {\varGamma }\) through \(w_y\). By (\({\varUpsilon }\)2), the second statement of (\({\varUpsilon }\)0) and form (1), we obtain that \(0\ll _Iy(-k)\ll _I w_y(-k)\) for any \(k\ge 0\), where \(\{w_y(-k),k\ge 0\}\) is the negative orbit contained in \({\varGamma }\) through \(w_y\). By virtue of Lemma 8, we obtain that \(0\in \alpha (y)\), a contradiction to \(\alpha (y)\cap [0,\frac{x_0}{2}]=\emptyset \). Thus, we have proved \(\mathrm{Int}{\varGamma }\subseteq \{y\in {\varGamma }:\alpha (y)=\{0\},\text { for any negative orbit of }y \text { in }{\varGamma }\}\). The last inclusion in (i) follows from Lemma 4(ii).

(ii) Again by Proposition 1, one has \(\omega (x)\subseteq {\varGamma }\). Suppose that there is an \(x\in \mathbf {R}^n_+{\setminus }\{0\}\) such that \(\omega (x)\nsubseteq \partial {\varGamma }\). Then we choose a \(z\in \omega (x)\cap \mathrm{Int}{\varGamma }\). Since \(0\notin \omega (x)\) by Lemma 6, one can find a nonempty index set I such that \(z\in \dot{H}_I^+\). By Lemma 3, we have \(z\ll _Iw_z\), where \(\{w_z\}=\partial {\varGamma }\cap L_z\). By the invariance of \(\omega (x)\) and \({\varGamma }\), there is a negative orbit \(\{z(-k):k\ge 0\}\subseteq \omega (x)\) (resp. a negative orbit \(\{w(-k):k\ge 0\}\subseteq {\varGamma }\)) through z (resp. \(w_z\)). Moreover, (\({\varUpsilon }\)2) also implies that \(z(-k)\ll _I w(-k)\) for any \(k\ge 0\). Without loss of generality, we assume that \(z(-k)\rightarrow z_*\in \omega (x)\). Again by Lemma 6, \(z_*\notin [0,\frac{x_0}{2}]\). On the other hand, it follows from Lemma 8 that \(z_*=0\), a contradiction. We have completed the proof.

Remark 9

We here emphasize that all the previous Propositions and Lemmas are obtained under (\({\varUpsilon }0\))–(\({\varUpsilon }4\)), and without the local injectivity assumption. In the following lemma, we need to impose the additional condition that T is locally injective in an open subset V of \(\mathbf {R}^n\) with \([0,q]\subseteq V\subseteq W\). As we will see in the following Lemma 9, the local injectivity of T in V will guarantee the invariance of \(\partial {\varGamma }\) and \(\partial B_0\). Without such assumption, we can only obtain the positive invariance of them.

Lemma 9

Let \(({\varUpsilon }0)\)\(({\varUpsilon }3)\) hold. Assume further that T is locally injective in an open subset V of \(\mathbf {R}^n\) with \([0,q]\subseteq V\subseteq W\). Then we have

  1. (i)

    \(\partial {\varGamma }\) is invariant and unordered, i.e., any two points in \(\partial {\varGamma }\) cannot be related by “\(\le \)”.

  2. (ii)

    \(B_0\) is an open subset in \({\varGamma }\), and hence, \(B_0\subseteq \mathrm{Int}{\varGamma }\). Moreover,

    $$\begin{aligned} B_0=\{y\in {\varGamma }:\alpha (y)=\{0\},\text { for any negative orbit in } {\varGamma }\text { of }y \}, \end{aligned}$$

    i.e., \(B_0\) is the basin of repulsion.

  3. (iii)

    \(\partial B_0\) is invariant and unordered, i.e., any two points in it cannot be related by “\(\le \)”.

Proof

(i) In order to show the invariance of \(\partial {\varGamma }\), by Proposition 1(iii), one only needs to prove the negative invariance of \(\partial {\varGamma }\). Suppose that there is some \(z\in \partial {\varGamma }\) such that its pre-image \(T^{-1}(\{z\})\cap \mathrm{Int}{\varGamma }\ne \emptyset \). Then choose a \(z_1\in T^{-1}(\{z\})\cap \mathrm{Int}{\varGamma }\). Since T is locally injective in the open subset V of \(\mathbf {R}^n\), the domain invariance theorem (Dold 1972) implies that T is an open mapping in V (It deserves to point out that the domain invariance Theorem is not valid if one only assumes that T is locally injective in [0, q], instead of V). This means that, for any \(x_0\in V\) and any neighborhood \(B(x_0)\) of \(x_0\) in V, \(TB(x_0)\) will contain an open neighborhood \(B(Tx_0)\) of \(Tx_0\) in \(\mathbf {R}^n\).

Since \(z_1\in \mathrm{Int}{\varGamma }\subseteq [0,q]\subseteq V\), we choose a neighborhood \(B(z_1)\) of \(z_1\) in V so small that \(B(z_1)\cap \mathbf {R}^n_+\subseteq {\varGamma }\). Moreover, \(TB(z_1)\) contains an open neighborhood B(z) of z in \(\mathbf {R}^n\). Because B(z) is open in \(\mathbf {R}^n\), one can find some \(y\in B(z)\) such that \(z\ll y\). Note also that \(B(z)\subseteq TB(z_1)\), there is a \(y_1\in B(z_1)\) such that \(Ty_1=y\). Since \(y \in \mathbf {R}^n_+\) and \(T^{-1}(\mathbf {R}^n_+)\subseteq \mathbf {R}^n_+\), one has \(y_1\in B(z_1)\cap \mathbf {R}^n_+\subseteq {\varGamma }\). Therefore, \(y=Ty_1\in {\varGamma }\). Recall that \(z\ll y\), it then implies that \(z\in \mathrm{Int}{\varGamma }\), a contradiction to \(z\in \partial {\varGamma }\). Thus, we have proved that \(\partial {\varGamma }\) is invariant. Moreover, together with (\({\varUpsilon }\)2), the invariance of \(\partial {\varGamma }\) clearly implies that it is unordered.

(ii) For any \(z\in B_0\), there exist some integer \(m\ge 0\) and some \(0\le y\ll x_0/2\) such that \(T^my=z\). Since T is an open mapping in V (hence \(T^m\)), now, one can choose a small neighborhood \(V_z\) of z in \(\mathbf {R}^n\) such that \(({V_z\cap \mathbf {R}^n_+})\subseteq T^m(U_y\cap \mathbf {R}^n_+)\subseteq T^m[0,x_0/2]\subseteq B_0\), where \(U_y\) is a small neighborhood of y in \(\mathbf {R}^n\) with \((U_y\cap \mathbf {R}^n_+)\subseteq [0,x_0/2]\). Thus, \(B_0\) is an open subset in \({\varGamma }\), and hence, \(B_0\subseteq \mathrm{Int}{\varGamma }\).

Next, we will prove \(B_0= \{y\in {\varGamma }:\alpha (y)=\{0\}\), for any negative orbit in \({\varGamma }\) of \(y \}\). To this end, by Lemma 4(ii), it suffices to prove that for any \(y\in \mathrm{Int}{\varGamma }\), there is a unique \(y_1\in {\varGamma }\) such that \(Ty_1=y\). Suppose that one can find two distinct \(y_1,y_2\in {\varGamma }\) s.t. \(Ty_1=y=Ty_2\). By the invariance of \(\partial {\varGamma }\), we have \(y_1,y_2\in \mathrm{Int}{\varGamma }\). Since \(y_1,y_2,y\in \mathrm{Int}{\varGamma }\subseteq V\) and T is a local homeomorphism in V, there exist open neighborhoods \(V_i\subseteq V\) of \(y_i\), \(V_*\subseteq V\) of y s.t. \(V_i\cap \mathbf {R}^n_+\subseteq [0,q]\) and \(T:V_i\rightarrow V_*,i=1,2,\) are homeomorphisms. Now, consider the homeomorphism \(T:V_1\rightarrow V_*\), choose \(z_n\gg y\) with \(z_n\rightarrow y\) (hence \(z_n\in \mathbf {R}^n_+\)). Recall that \(T^{-1}(\mathbf {R}^n_+)\subseteq \mathbf {R}^n_+\) and \(T:V_1\rightarrow V_*\) is homeomorphic. Then there is a sequence \(x_n\in \mathbf {R}_+^n\cap V_1\subseteq [0,q]\) such that \(Tx_n=z_n\) and \(x_n\rightarrow y_1\). Consequently, \(Tx_n\gg y=Ty_2\). By (\({\varUpsilon }\)2), one has \(x_n\ge y_2\), which implies that \(y_1\ge y_2.\) Similarly, one can obtain \(y_2\ge y_1\). So \(y_1=y_2\), a contradiction. Thus, we have proved the uniqueness of \(y_1\). This imples that for any \(y\in B_0\subseteq \mathrm{Int}{\varGamma }\), there is a unique negative orbit \(\{y(-m):m\ge 0\}\subseteq {\varGamma }\) through y. So, Lemma 4 directly entails that

$$\begin{aligned} B_0= \{y\in {\varGamma }:\alpha (y)=\{0\},\text { for any negative orbit in } {\varGamma }\text { of }y \}. \end{aligned}$$

(iii) By Lemma 4(iii), it suffices to prove the negative invariance of \(\partial B_0\). Note that \(\overline{B_0}\) and \(B_0\) are invariant and \(B_0\) is open. Then this can be done by the similar arguments in (i).

Now we are ready to prove Theorem 6.

Proof of Theorem 6

Under the assumptions (\({\varUpsilon }\)0)–(\({\varUpsilon }\)4), the statements of (i)-(iv) are obtained directly from Propositions 1 and 2. Moreover, by Proposition 1(iii), the map \(x\mapsto w_x\) from \(\{x\in \mathbf {R}_+^n:||x||=1\}\) to \(\partial {\varGamma }\) is well defined, where \(w_x\) is the point of intersection of the ray \(L_{x}=\{tx:t\ge 0\}\) with \(\partial {\varGamma }\), and it is also injective and onto \(\partial {\varGamma }\) (by (20)). The inverse mapping from \(\partial {\varGamma }\) onto \(\{x\in \mathbf {R}_+^n:||x||=1\}\) is clearly continuous, hence it is a homeomorphism by the compactness of the two sets. Thus, \(\partial {\varGamma }\) is homeomorphic to the unit sphere in \(\mathbf {R}_+^n\), which in turn, is homeomorphic to \({\varDelta }^{n-1}\).

Hereafter, we always assume that T is locally injective in an open subset V of \(\mathbf {R}^n\) with \([0,q]\subseteq V\subseteq W\). By Proposition 2(i) and Lemma 9, (\({\varUpsilon }\)0)–(\({\varUpsilon }\)4) immediately imply that \(\partial {\varGamma }\) is invariant and

$$\begin{aligned} \mathrm{Int}{\varGamma }=\{y\in {\varGamma }:\alpha (y)=\{0\},\text { for any negative orbit in }{\varGamma }\text { through }y \}=B_0. \end{aligned}$$
(24)

Moreover, we claim that (24), as well as (i)–(iv), can still hold even under the weaker assumption (\({\varUpsilon }4\)’), instead of (\({\varUpsilon }4\)), when T is locally injective in V. Indeed, under (\({\varUpsilon }\)0)–(\({\varUpsilon }\)3), Lemma 9 implies that \(\partial {\varGamma }\) and \(\partial B_0\) are invariant and unordered. In the following paragraph, we will prove that \(\partial B_0=\partial {\varGamma }\). But, before doing it, we first show how it implies the claim. In fact, for any \(y\in \mathrm{Int}{\varGamma }\), let I be the nonempty index subset such that \(y\in \dot{H}_I^+\). Then it follows from Lemma 3 that \(y\ll _I w_y\in \partial {\varGamma }=\partial B_0\). Hence, one can find some \(z\in B_0\) so close to \(w_y\) that \(y\ll z\). Lemma 4(i) then entails that \(y\in B_0\). So, \(\mathrm {Int}{\varGamma }\subseteq B_0\). Thus, by Lemma 9(ii), one can also obtain (24). Now by Proposition 1, Lemma 9 and (24), (i)–(iii) still hold under (\({\varUpsilon }\)4’). As for (iv), suppose that there is an \(x\in \mathbf {R}^n_+{\setminus }\{0\}\) such that \(\omega (x)\nsubseteq \partial {\varGamma }\). Then we choose a \(y\in \omega (x)\cap \mathrm{Int}{\varGamma }=\omega (x)\cap B_0\). By the invariance of \(\omega (x)\), there exists a negative orbit \(\{y(-k):k\ge 0\}\) through y such that \(\{y(-k)\}\subseteq \omega (x)\subseteq {\varGamma }\) for all \(k\ge 0\). Since \(y\in B_0\), (24) entails that \(0\in \omega (x)\), which contradicts Lemma 6. Thus, we have proved (iv), that is, \(\omega (x)\subseteq \partial {\varGamma }\) for all \(x\in \mathbf {R}^n_+{\setminus }\{0\}\). This confirms our claim above.

So, in order to prove the claim, it suffices to show \(\partial B_0=\partial {\varGamma }\) under (\({\varUpsilon }\)4’), which will be done by induction on n. If \(n=1\), then it is clear that \(\partial B_0=\partial {\varGamma }=\{q_{1}\}\). From now on, we assume \(n > 1\). Then the induction hypothesis is that \(\partial B_0=\partial {\varGamma }\) holds for systems in \(\mathbf {R}^m_+\) for any \(m < n\). Therefore, fix \(n>1\), the induction hypothesis implies that for any \(\phi \ne I\subseteq N\), we have \(\partial B_0\cap H^+_I=\partial {\varGamma }\cap H^+_I\). Hence, \(\partial B_0\cap \partial \mathbf {R}^n_+ =\partial {\varGamma }\cap \partial \mathbf {R}^n_+\). So, one only needs to prove that \(\partial B_0\cap \dot{\mathbf {R}}^n_+=\partial {\varGamma }\cap \dot{\mathbf {R}}^n_+\). Note that \(\forall u\in \dot{\mathbf {R}}^n_+, |u|=1\), the ray \(L_{u}=\{tu:t\ge 0\}\) meets \(\partial B_0\) and \(\partial {\varGamma }\) in points x and y. If there exist \(x\ne y\) such that \(0\ll x\ll y\), then by the invariance of \(\partial B_0\) and \(\partial {\varGamma }\) there exist negative orbits \(\{x(-k):k\ge 0\}\subseteq \partial B_0, \{y(-k):k\ge 0\}\subseteq \partial {\varGamma }\) such that \(x(-k)\ll y(-k)\), \(\forall k\ge 1\), and \(\liminf _{k\rightarrow +\infty }\Vert x(-k)-y(-k)\Vert =d>0\) by Lemma 7. Choose a sequence of integers \(m_k\rightarrow +\infty \) such that \(x_*=\lim _{k\rightarrow +\infty }x(-m_k)\), and \(y_*=\lim _{k\rightarrow +\infty }y(-m_k)\), so \(x_*<y_*\). Note that \(x_*\in \partial B_0\) and \(y_*\in \partial {\varGamma }\). Suppose \(x_*\in \partial \mathbf {R}^n_+\), then \(x_*\in \partial B_0\cap \partial \mathbf {R}^n_+=\partial {\varGamma }\cap \partial \mathbf {R}^n_+\), hence \(x_*\in \partial {\varGamma }\), which contradicts that \(\partial {\varGamma }\) is unordered. Consequently, one may assume that \(0\ll x_*<y_*\). Furthermore, one can obtain that \(T^l x_*< T^l y_*\) and \(\Vert T^l x_* - T^l y_*\Vert \ge d > 0\) for any \(l\ge 0\). For \( i\in N, l\ge 0\), set \(K_i(l):=\frac{(T^{l} x_*)_i}{(T^l y_*)_i}\). Obviously, \(0<K_i(l)<1\) for any \(i\in N\). Then it follows from (23) that \(K_i(l)\ge K_i(l-1)\) for any \(i\in N\); and moreover, by (\({\varUpsilon }\)4’), we have

$$\begin{aligned} K_i(l+1)-K_i(l)&=\frac{K_i(l)}{G_i(T^{l}y_*)}\sum ^n_{j=1}\int ^1_0\frac{\partial G_i}{\partial x_j}(sT^{l}x_*+(1-s)T^{l}y_*)ds\cdot (T^{l}x_*-T^{l}y_*)_j\\&\ge \frac{\delta \cdot K_i(0)}{G_i(T^{l}y_*)}\cdot (T^{l}y_*-T^{l}x_*)_i \ge \frac{\delta K_*}{G_*}\cdot (T^{l} y_*-T^{l} x_*)_i, \end{aligned}$$

where \(\delta =\min _{\begin{array}{c} i\in N\\ x\in [0,q] \end{array}}{|\frac{\partial G_i}{\partial x_i}(x)|}>0\) (see (\({\varUpsilon }4\)’)), \(K_*=\min _{i\in N}K_i(0)>0\), and \(G_*=\max _{x\in [0,q]}(\sum _{i=1}^n G_i(x))>0\). Consequently,

$$\begin{aligned} \sum ^n_{i=1}K_i(l) \ge \sum ^n_{i=1}K_i(0)+\frac{\delta K_*}{G_*}\sum ^{l-1}_{j=0}\Vert T^j y_*-T^j x_*\Vert \ge \sum ^n_{i=1}K_i(0)+\frac{\delta K_*}{G_*}dl\rightarrow +\infty \end{aligned}$$

as \(l\rightarrow +\infty \), which contradicts \(\sum ^n_{i=1}K_i(l)\le n\). Therefore, it yields that \(x=y\). Thus, we have proved \(\partial B_0=\partial {\varGamma }\).

Finally, we will show (17). Given any \(x\in {\mathbf {R}^n_+}{\setminus }\{0\}\), one can find some index set \(\emptyset \ne I\subseteq N\) such that \(x\in \dot{H}^+_I\). By Proposition 1(iii), the ray from the origin through \(T^kx\) will meet \(\partial {\varGamma }\) at the unique point \(w_k\in \dot{H}^+_I\), \(\forall k\ge 0\). So one of the following three alternatives will occur:

  1. (a)

    There exists some \(k_0\ge 0\), such that \(T^{k_0}x=w_{k_0}\);

  2. (b)

    There exists some \(k_0\ge 0\), such that \(T^{k_0}x\ll _I w_{k_0}\);

  3. (c)

    \(T^{k}x\gg _I w_k\) for any \(k\ge 0\).

For case (a), since \(\partial {\varGamma }\) is invariant, one can choose \(y\in T^{-k_0}(w_{k_0})\cap \partial {\varGamma }\), and hence, (17) holds immediately. For case (b), Lemma 3 implies that \(T^{k_0}x\in \mathrm{Int}{\varGamma }\). Recall that \(\mathrm{Int}{\varGamma }\) is invariant. Then it entails that \(T^kx\in \mathrm{Int}{\varGamma }\) for all \(k\ge k_0\). Without loss of generality, we hereafter assume that \(k_0=0\).

Now, for each \(k\ge 0\), define \(M_+(T^kx)=\{y\in \partial {\varGamma }:y\ge T^kx\}\). Clearly, \(M_+(T^kx)\subseteq \partial {\varGamma }\) is a compact set, and moreover, \(M_+(T^kx)\ne \emptyset \) by Lemma 3. Note that \(T^kx\notin \partial {\varGamma }\). Then \(M_+(T^kx)=\{y\in \partial {\varGamma }:y> T^kx\}\). Together with the invariance of \(\partial {\varGamma }\), (\({\varUpsilon }\)2) then implies that

$$\begin{aligned} \emptyset \ne T^{-1}(M_+(T^{k+1}x))\cap \partial {\varGamma }\subseteq M_+(T^{k}x),\quad \text {for all } k\ge 0. \end{aligned}$$
(25)

Now we define

$$\begin{aligned} D_1(T^kx):=T^{-1}(M_+(T^{k}x))\cap \partial {\varGamma },\quad \text {for } k\ge 0, \end{aligned}$$

and

$$\begin{aligned} D_{j+1}(T^{k}x):=T^{-1}(D_{j}(T^{k+1}x))\cap \partial {\varGamma }, \quad \text {for } j\ge 1,k\ge 0. \end{aligned}$$

By the invariance of \(\partial {\varGamma }\), \(D_j(T^kx)\) is a nonempty compact subset in \(\partial {\varGamma }\) for any \(j\ge 1\) and \(k\ge 0\). Then one has the following properties for \(D_j(T^kx)\):

  1. (P1)

    \(D_{j+1}(T^kx)\subseteq D_j(T^kx)\subseteq \cdots \subseteq D_1(T^kx)\subseteq M_+(T^{k-1}x)\), for any \(j,k\ge 1\);

  2. (P2)

    If \(z\in D_{j}(Tx)\), then \(T^kz\in D_{j-k}(T^{k+1}x)\) for any \(j\ge 1\) and \(0\le k<j\).

As a matter of fact, the last inclusion in (P1) is directly from the definition of \(D_1(T^kx)\) and (25). In order to prove \(D_{j+1}(T^kx)\subseteq D_j(T^kx)\) in (P1), by the iterated definition of \(D_{j}(T^{k}x)\), it suffices to show that \(D_2(T^kx)\subseteq D_1(T^kx)\) for any \(k\ge 0\), and this can be easily verified by (25). Moreover, (P2) can be checked directly by the definition of \(D_{j}(T^kx)\).

Now we can prove (17) for some \(y\in \partial {\varGamma }\). Indeed, by virtue of (P1), we define \(D_x=\bigcap _{j\ge 1}^\infty D_j(Tx)\). Clearly, \(D_x\) is a nonempty compact subset in \(\partial {\varGamma }\) satisfying \(D_x\subseteq M_+(x)\). Choose a \(y\in D_x\). Then \(y\in D_j(Tx)\) for any \(j\ge 1\). By virtue of (P1)–(P2), it follows that \(T^ky\in D_{j-k}(T^{k+1}x)\subseteq M_+(T^{k}x)\) for any \(k<j\). As a consequence, we obtain that \(T^kx<T^ky\) for any \(k\ge 0\). Suppose that \(\Vert T^ky-T^kx\Vert \nrightarrow 0\), then one can choose a subsequence \(\{k_i\}\) such that \(T^{k_i}y\rightarrow y_0\in \omega (y)\subseteq \partial {\varGamma }\), \(T^{k_i}x \rightarrow x_0\in \omega (x)\subseteq \partial {\varGamma }\) and \(y_0>x_0\), contradicting that \(\partial {\varGamma }\) is unordered. Thus, we have proved (17).

For case (c), a similar argument as in the last part of the previous paragraph shows the result. We have completed the proof of Theorem 6. \(\square \)

Remark 10

It deserves to mention that the assumption \(({\varUpsilon }4)\) is needed (see in Lemma 8), if we do not have the local injectivity of T in V (see also in Remark 9). However, a careful examination of the arguments after (24) yields that, if the local injectivity of T is assumed, then \(\partial {\varGamma }\) and \(\partial B_0\) are invariant and unordered with respect to “\(\le \)”. Based on this, one can replace \(({\varUpsilon }4)\) by the weaker assumption (\({\varUpsilon }4\)’) to obtain (i)–(iv) in Theorem 6.

In the following, we will give a simple example, in which \([\partial G_i/\partial x_j(x)]_{i,j\in I}\) has non-positive off-diagonal entries [i.e., (\({\varUpsilon }4\)’) holds], which admits a carrying simplex.

Example

Consider the following two-dimensional map:

$$\begin{aligned} T(x)=\left( \frac{2x_1}{1+x_1+b_{12}x_2^2},\frac{2x_2}{1+b_{21}x_1+x_2}\right) ,\quad b_{12}>0,\;0<b_{21}\le \frac{1}{2},\;x\in \mathbf {R}^2_+. \end{aligned}$$
(26)

Let \(G_1(x)=\frac{2}{1+x_1+b_{12}x_2^2}\), \(G_2(x)=\frac{2}{1+b_{21}x_1+x_2}\) and \(G(x)=(G_1(x),G_2(x))\). It is clear that \(G(x)\gg 0\) in a small open neighborhood \(W\subseteq \mathbf {R}^2\) of \(\mathbf {R}^2_+\). Now \(T_i(x)=x_iG_i(x)\) can be defined on W with \(T(\mathbf {R}^2_+)\subseteq \mathbf {R}^2_+\) and \(T^{-1}(\mathbf {R}^2_+)\subseteq \mathbf {R}^2_+\). Note that \(\partial G_1(x)/\partial x_2=-\frac{4b_{12}x_2}{(1+x_1+b_{12}x_2^2)^2}\le 0\) for any \(x\in \mathbf {R}^2_+\) and \(\partial G_1(x)/\partial x_2=0\) for \(x\in H_{\{1\}}^+\). On the other hand, \(\forall x\in \mathbf {R}^2_+\), \(\partial G_i(x)/\partial x_j<0\) for \(i\ne 1,j\ne 2\). So, (\({\varUpsilon }\)4’) holds [(\({\varUpsilon }\)4) does not]. Clearly, \(G_i(0)>1\) and \(T|_{\dot{H}_{\{i\}}^+}\) has a globally attracting fixed point \(q_{\{i\}}=e_{\{i\}}\). Hence, (\({\varUpsilon }\)0) and (\({\varUpsilon }\)1) hold for T. Set \(q=(1,1)\) and \(S=[0, 2)\times [0,2)\). Then \(T(\mathbf {R}^2_+)\subseteq S\) and \([0,q]\subseteq S\).

A routine computation yields that for any \(x\in S\),

$$\begin{aligned} \det DT(x)=\frac{4(1+b_{21}x_1+b_{12}x_2^2(1-b_{21}x_1))}{(1+x_1+b_{12}x^2_2)^2(1+b_{21}x_1+x_2)^2} >0. \end{aligned}$$

Hence, \(DT(x)^{-1}\) exists for every \(x \in S\) and T is locally homeomorphic on S. Together with \(T^{-1}(\{0\})=\{0\}\), Lemma 3.4 in Chow and Hale (1982, p. 27) implies that T is also homeomorphic on S. On the other hand, it is not difficult to calculate that

$$\begin{aligned} D(T|_{H^+_I})(x)^{-1}=(DT(x)^{-1})_I\gg 0, \quad \forall \emptyset \ne I\subseteq N, x\in \dot{S}^+_I. \end{aligned}$$

Hence (\({\varUpsilon }\)5) holds on S. Now Corollary 2 implies that T admits a carrying simplex \({\varSigma }\).

Postscript After the first version of this manuscript had been submitted, we noticed the detailed work in Ruiz-Herrera (2013, Appendix) containing result similar to the one we present in the Appendix here. We thank one of the referees to remind us for this. Comparing with (Ruiz-Herrera 2013), we first show that T can still admit a global attractor \({\varGamma }\) only under assumptions (\({\varUpsilon }\)1)–(\({\varUpsilon }\)2). We further study the geometrical structure of \({\varGamma }\) and \(\partial {\varGamma }\), as well as the global attractivity of \(\partial {\varGamma }\), under assumptions (\({\varUpsilon }\)0)–(\({\varUpsilon }\)4) without the local injectivity assumption. The local injectivity assumption is only needed for the proof of the negative invariance of \(\partial {\varGamma }\) and the property (17). Moreover, we have replaced the positivity of G(x) in the whole domain \(\mathbf {R}^n_+\) by (\({\varUpsilon }\)0), and negativity of the entries of DG(x) in the whole domain \(\mathbf {R}^n_+\) by the weaker condition (\({\varUpsilon }\)4) (or (\({\varUpsilon }\)4’)).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jiang, J., Niu, L. & Wang, Y. On heteroclinic cycles of competitive maps via carrying simplices. J. Math. Biol. 72, 939–972 (2016). https://doi.org/10.1007/s00285-015-0920-1

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-015-0920-1

Keywords

Mathematics Subject Classification

Navigation