Competitive and Cooperative Differential Equations

Similar documents
ẋ = f(x, y), ẏ = g(x, y), (x, y) D, can only have periodic solutions if (f,g) changes sign in D or if (f,g)=0in D.

In particular, if A is a square matrix and λ is one of its eigenvalues, then we can find a non-zero column vector X with

Research Article Global Dynamics of a Competitive System of Rational Difference Equations in the Plane

Half of Final Exam Name: Practice Problems October 28, 2014

Equilibria with a nontrivial nodal set and the dynamics of parabolic equations on symmetric domains

CHAPTER 2: CONVEX SETS AND CONCAVE FUNCTIONS. W. Erwin Diewert January 31, 2008.

(x k ) sequence in F, lim x k = x x F. If F : R n R is a function, level sets and sublevel sets of F are any sets of the form (respectively);

Set, functions and Euclidean space. Seungjin Han

Feedback control for a chemostat with two organisms

2 Sequences, Continuity, and Limits

2.10 Saddles, Nodes, Foci and Centers

1 Lyapunov theory of stability

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

(x 1, y 1 ) = (x 2, y 2 ) if and only if x 1 = x 2 and y 1 = y 2.

A note on the monotonicity of matrix Riccati equations

On John type ellipsoids

GENERALIZED CONVEXITY AND OPTIMALITY CONDITIONS IN SCALAR AND VECTOR OPTIMIZATION

THE ROSENZWEIG-MACARTHUR PREDATOR-PREY MODEL

CHAPTER 7. Connectedness

Topological properties

Course 212: Academic Year Section 1: Metric Spaces

A NICE PROOF OF FARKAS LEMMA

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

1 The Observability Canonical Form

Optimality Conditions for Nonsmooth Convex Optimization

1. Bounded linear maps. A linear map T : E F of real Banach

DYNAMICAL CUBES AND A CRITERIA FOR SYSTEMS HAVING PRODUCT EXTENSIONS

7 Planar systems of linear ODE

(c) For each α R \ {0}, the mapping x αx is a homeomorphism of X.

Metric Spaces and Topology

p,q H (X), H (Y ) ), where the index p has the same meaning as the

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

Stable and Unstable Manifolds for Planar Dynamical Systems

CALCULUS ON MANIFOLDS

Global Qualitative Analysis for a Ratio-Dependent Predator Prey Model with Delay 1

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

Asymptotic Stability by Linearization

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

Stability Implications of Bendixson s Criterion

B. Appendix B. Topological vector spaces

Iowa State University. Instructor: Alex Roitershtein Summer Homework #5. Solutions

Stereographic projection and inverse geometry

Research Article Bifurcation and Global Dynamics of a Leslie-Gower Type Competitive System of Rational Difference Equations with Quadratic Terms

The uniformization theorem

Linear ODEs. Existence of solutions to linear IVPs. Resolvent matrix. Autonomous linear systems

1.7. Stability and attractors. Consider the autonomous differential equation. (7.1) ẋ = f(x),

Part III. 10 Topological Space Basics. Topological Spaces

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Analysis-3 lecture schemes

An introduction to some aspects of functional analysis

In English, this means that if we travel on a straight line between any two points in C, then we never leave C.

A REMARK ON THE GLOBAL DYNAMICS OF COMPETITIVE SYSTEMS ON ORDERED BANACH SPACES

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

Chapter 7. Extremal Problems. 7.1 Extrema and Local Extrema

Extreme points of compact convex sets

USING FUNCTIONAL ANALYSIS AND SOBOLEV SPACES TO SOLVE POISSON S EQUATION

1 Directional Derivatives and Differentiability

Introduction to Real Analysis Alternative Chapter 1

ORBITAL SHADOWING, INTERNAL CHAIN TRANSITIVITY

Lecture 4. Chapter 4: Lyapunov Stability. Eugenio Schuster. Mechanical Engineering and Mechanics Lehigh University.

NORMS ON SPACE OF MATRICES

Notes on Ordered Sets

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

DO NOT OPEN THIS QUESTION BOOKLET UNTIL YOU ARE TOLD TO DO SO

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

Discrete dynamics on the real line

ADDING MACHINES, ENDPOINTS, AND INVERSE LIMIT SPACES. 1. Introduction

Dynamical Systems and Ergodic Theory PhD Exam Spring Topics: Topological Dynamics Definitions and basic results about the following for maps and

An Undergraduate s Guide to the Hartman-Grobman and Poincaré-Bendixon Theorems

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

The Hurewicz Theorem

Exercises: Brunn, Minkowski and convex pie

Some notes on Coxeter groups

Congurations of periodic orbits for equations with delayed positive feedback

Problem List MATH 5173 Spring, 2014

Math 117: Topology of the Real Numbers

Mathematical Economics. Lecture Notes (in extracts)

Convex Analysis and Economic Theory AY Elementary properties of convex functions

Chapter III. Stability of Linear Systems

Dynamical Systems & Lyapunov Stability

LINEAR CHAOS? Nathan S. Feldman

NOTES ON LINEAR ODES

COMPLEX VARIETIES AND THE ANALYTIC TOPOLOGY

Chapter 2 Metric Spaces

THE JORDAN-BROUWER SEPARATION THEOREM

Tree sets. Reinhard Diestel

Linear Algebra I. Ronald van Luijk, 2015

Observations on the Stability Properties of Cooperative Systems

FOURTH ORDER CONSERVATIVE TWIST SYSTEMS: SIMPLE CLOSED CHARACTERISTICS

Continued fractions for complex numbers and values of binary quadratic forms

7. Homotopy and the Fundamental Group

April 13, We now extend the structure of the horseshoe to more general kinds of invariant. x (v) λ n v.

27. Topological classification of complex linear foliations

An introduction to Mathematical Theory of Control

INVERSE FUNCTION THEOREM and SURFACES IN R n

Convex Functions and Optimization

Detailed Proof of The PerronFrobenius Theorem

Summary of topics relevant for the final. p. 1

THE INVERSE FUNCTION THEOREM

Transcription:

Chapter 3 Competitive and Cooperative Differential Equations 0. Introduction This chapter and the next one focus on ordinary differential equations in IR n. A natural partial ordering on IR n is generated by the cone of vectors with nonnegative components but we shall see that the other orthants of IR n can equally serve to generate useful partial orderings. An ordinary differential equation can be solved both forward in time and backward in time. Roughly speaking, we will say that a system of ordinary differential equations is cooperative if it generates a monotone semiflow in the forward time direction and competitive if it generates a monotone semiflow in the backward time direction. The chapter begins with the classical Kamke condition for an autonomous ordinary differential equation to generate a monotone semiflow in the forward time direction. Section two contains a useful result on positively invariant sets and monotone solutions. The main results of the chapter are Theorem 3.4 and Theorem 4.1. The former asserts that a compact limit set of a competitive or cooperative system in IR n is topologically equivalent to a flow on a compact invariant set of a Lipschitz system of differential equations in IR n 1. In other words, the long term dynamics of an n-dimensional competitive or cooperative system can be no more badly behaved than that of a general system of one less dimension. Since our understanding of one and two dimensional (general) systems is fairly complete, we can expect to apply this result effectively to two and three dimensional competitive and cooperative systems. 1

Theorem 4.1 establishes that the Poincaré-Bendixson Theorem holds for three dimensional competitive and cooperative systems. Theorem 5.2 shows that every bounded solution of a planar competitive or cooperative system converges to equilibrium. Competitive systems in IR 3 can have attracting periodic orbits (cooperative systems cannot!) so additional attention is focused on these systems. It can be difficult to apply Theorem 4.1 even in the simple case where the system has only a single unstable equilibrium because of the potential existence of homoclinic orbits. In Theorem 4.2, sufficient conditions are given for all solutions that start off the (one-dimensional) stable manifold of the equilibrium to approach a nontrivial periodic orbit. Every periodic orbit of a three dimensional competitive or cooperative system contains an equilibrium inside it in a sense made precise in Proposition 4.3. These ideas are applied to the classical Field-Noyes model of the Belousov-Zhabotinsky reaction in section 6. The strong order preserving property plays no role here. As a result, this chapter is essentially independent of the previous ones. In the following chapter, the irreducibility assumption on the Jacobian of the vector field is introduced in order to establish the strong order preserving property. This assumption is avoided in the present chapter. 1. The Kamke Condition Consider the autonomous system of ordinary differential equations x = f(x) (1.1) where f is continuously differentiable on an open subset D IR n. We change our notation slightly to conform to more traditional notation for ordinary differential equations. For example, let φ t (x) denote the solution of (1.1) that starts at the point x at t = 0. φ t will be referred to as the flow corresponding to (1.1). We sometimes refer to f as the vector field generating the flow φ t. If φ t (x) is defined for all t 0 (t 0), then we write γ + (x) = {φ t (x) : t 0} (γ (x) = {φ t (x) : t 0}) for the positive orbit (negative 2

orbit) through the point x. If γ (x) has compact closure in D, then the alpha limit set, α(x), is defined in a similar way as the omega limit set: α(x) = t 0 τ t φ τ (x). The nonnegative cone in IR n, denoted by IR n +, is the set of all n-tuples with nonnegative coordinates. It gives rise to a partial order on IR n by y x if x y IR n +. Less formally, this is true if and only if y i x i for all i. We write x < y if x y and x i < y i for some i and we write x y if x i < y i for all i. Our immediate goal is to provide sufficient conditions for φ to be a monotone dynamical system with respect to the ordering. f is said to be of type K in D if for each i, f i (a) f i (b) for any two points a and b in D satisfying a b and a i = b i. The following classical result asserts that the type K condition is sufficient for the order preserving property to hold. It is also a necessary condition. Proposition 1.1. Let f be type K on D and x 0, y 0 D. Let < r denote any one of the relations, < or. If x 0 < r y 0, t > 0 and φ t (x 0 ) and φ t (y 0 ) are defined, then φ t (x 0 ) < r φ t (y 0 ). Proof: For m = 1, 2,..., let φ m t (x) be the flow corresponding to x = f(x) + (1/m)e where e = (1, 1,..., 1). Suppose that x 0 y 0, t > 0 and both φ t (x 0 ) and φ t (y 0 ) are defined. Then (Hale(1980), Chapt.1, Lemma 3.1), φ m s (y 0 + e/m) is defined on 0 s t for all large m, say m > M, and φ m s (y 0 + e/m) φ s (y 0 ) as m, uniformly in s [0, t]. We claim that φ s (x 0 ) φ m s (y 0 + e/m) (1.2) 3

holds for 0 s t, for all m > M. Fix m > M. Since x 0 = φ 0 (x 0 ) y 0 + e/m = φ m 0 (y 0 + e/m), the claim holds for small s by continuity. If the claim were false, then there exists t 0, satisfying 0 < t 0 t, φ s (x 0 ) φ m s (y 0 + e/m) on 0 s < t 0, and an index i such that φ t0 (x 0 ) i = φ m t 0 (y 0 + e/m) i, where the subscript i denotes the i-th component. Also, it follows that d ds s=t 0 φ s (x 0 ) i d ds s=t 0 φ m s (y 0 + e/m) i. However, as φ t0 (x 0 ) j φ m t 0 (y 0 + e/m) j for j i, the type K condition implies that f i (φ t0 (x 0 )) f i (φ m t 0 (y 0 + e/m)) < f i (φ m t 0 (y 0 + e/m)) + 1/m, or d ds s=t 0 φ s (x 0 ) i < d ds s=t 0 φ m s (y 0 + e/m) i. This contradiction proves the claim. Taking the limit m in (1.2) results in φ t (x 0 ) φ t (y 0 ). If x 0 < y 0 then φ t (x 0 ) φ t (y 0 ) and equality does not hold since φ t is one-to-one. Therefore, φ t (x 0 ) < φ t (y 0 ). If x 0 y 0 then φ t maps the open set [x 0, y 0 ] D into the set [φ t (x 0 ), φ t (y 0 )] by the first part of the proof. Since φ t is a homeomorphism on D, the latter set must be open. This holds if and only if φ t (x 0 ) φ t (y 0 ). The type K condition is most easily identifiable from the sign structure of the Jacobian matrix of the vector field on suitable domains. The next remark describes this structure. Remark 1.1 The type K condition can be expressed in terms of the partial derivatives of f on suitable domains. We say that D is p-convex if tx+(1 t)y D for all t [0, 1] whenever x, y D and x y. If D is a convex set then it is also p-convex. If D is a 4

p-convex subset of IR n and f i x j (x) 0, i j, x D, (1.3) holds then the fundamental theorem of calculus, implies that f is of type K in D. In fact, if a b and a i = b i, then by (1.3). f i (b) f i (a) = 1 0 j i f i x j (a + r(b a))(b j a j )dr 0, Some extensions of Proposition 1.1 are described in the following remarks. Remark 1.2 Proposition 1.1 has an obvious analog for nonautonomous systems. For example, suppose that f(t, x) and f x (t, x) are continuous in IR+ D where D is an open subset of IR n and for each t 0, f(t, ) satisfies the type K condition. If x(t) and y(t) are two solutions of x = f(t, x) on the interval [a, b] satisfying x(a) < r y(a), then x(b) < r y(b). Here, < r stands for one of the three relations, <,. The proof of this remark is very similar to the proof of Proposition 1.1 so we do not give it here. Remark 1.3 The previous remark can be applied to the linear system y = Df(x(t))y where x(t) is a solution of (1.1) defined on IR + and Df(x) is the Jacobian matrix of f at x. That is, the function g(t, y) = Df(x(t))y satisfies the hypotheses of Remark 1.2. Consequently, if y(t) is a solution of the linear system satisfying 0 < r y(0), then 0 < r y(t) for t > 0. 5

Remark 1.4 In the applications it frequently occurs that the natural domain for (1.1) is a closed set D and we would like for the conclusions of Proposition 1.1 to hold on D. Suppose that f is continuously differentiable on some neighborhood of D and that D is the closure of an open set G on which f is type K. Assume that D is positively invariant for (1.1). Suppose further that whenever x, y D satisfy x < y, there exists sequences x n, y n G satisfying x n < y n and x n x, y n y as n. Then Proposition 1.1 holds on D. The proof follows from continuity of solutions with respect to initial conditions and the fact that Proposition 1.1 holds on G. The system (1.1) is said to be a cooperative system if (1.3) holds on the p-convex domain D. It is called a competitive system on D if D is p-convex and the inequalities (1.3) are reversed: f i x j (x) 0, i j, x D. Observe that if (1.1) is a competitive system with flow φ t then x = f(x) is a cooperative system with flow ψ t where ψ t (x) = φ t (x), and conversely. Therefore, by time reversal, a competitive system becomes a cooperative system and vice-versa. This fact will be exploited repeatedly below. A cooperative system generates a monotone dynamical system; the order relation is preserved by the forward flow according to Proposition 1.1. A competitive system has the property that the time-reversed flow is monotone; if x y and t < 0 then φ t (x) φ t (y). In particular, if x and y are not related, that is, neither x y nor y x holds, and if t > 0 then φ t (x) and φ t (y) are not related (for if they were, then application of φ t would lead to x and y being related). Therefore, the forward flow of a competitive system preserves the property of two points being unrelated. 2. Positively Invariant Sets and Monotone Solutions 6

Cooperative and Competitive systems have certain canonical positively invariant sets that are identified in the next result. Proposition 2.1. If (1.1) is cooperative and < r stands for one of the relations, <,, then P + = {x D : 0 < r f(x)} and P = {x D : f(x) < r 0} are positively invariant. If x P + (x P ), then φ t (x) is nondecreasing (nonincreasing) for t 0. If, in addition, γ + (x) has compact closure in D, then ω(x) is an equilibrium. If (1.1) is competitive, then U + = {x D : f i (x) > 0 for some i} and U = {x D : f i (x) < 0 for some i} are positively invariant. Also, V + = {x D : f i (x) 0 for some i} and V = {x D : f i (x) 0 for some i} are positively invariant. V + V is closed, positively invariant and contains any compact invariant set that contains no equilibria. Proof: When (1.1) is cooperative, the positive invariance of the sets P +, P follows directly from Remark 1.3 concerning the linear system y = Df(x(t))y where x(t) is a solution of (1.1). Indeed, if x(t) is a solution of (1.1), then y(t) = f(x(t)) is a solution of the linear system. Since (1.3) holds, IntIR n +, IR n + \ {0} and IR n + are positively invariant for the linear system. If K denotes any one of these sets then it follows that K is also positively invariant for the linear system. Now suppose that (1.1) is competitive, x 0 U + and there is a s > 0 such that φ s (x 0 ) / U +. Then y 0 = φ s (x 0 ) G = {z D : f(z) 0}. G is positively invariant for the time-reversed cooperative system x = f(x) by the paragraph above. The flow for the time-reversed system is φ t (x) so x 0 = φ s (y 0 ) G. This contradiction (to x 0 U + ) proves that U + is positively invariant for the competitive system. Similar arguments show the positive invariance of U, V +, V. Consequently, V + V is positively invariant. If x does not belong to V + then f(x) 0 so f(φ t (x)) 0 for t > 0 by the first paragraph of the proof. Therefore, d dt φ t(x) = f(φ t (x)) 0 for 7

t < 0 so φ t (x) is strictly decreasing on t < 0. If, in addition, x A, where A is a compact invariant set, then it follows that α(x) is an equilibrium in A. Since A is assumed to have no equilibria, it follows that A V +. A similar argument shows that A V. According to Proposition 2.1, a solution φ t (x) of a competitive system which never meets V + V must be strictly increasing or strictly decreasing in t. If such a solution generates a positive orbit with compact closure in D, then it converges to equilibrium. Furthermore, any periodic orbit must be contained in V + V. The most important assertion of Proposition 2.1 is the first one for cooperative systems. The positive invariance of the sets P + and P means that any bounded solution starting in one of these sets is monotone and therefore must converge to equilibrium. Points of P + where < r = are sub-equilibria and points of P + where < r =< are strict sub-equilibria. Similarly, P are super-equilibria or strict superequilibria. See section 5 of Chapter 2. Figure 2.1 depicts the positively invariant set V + V for the two dimensional competitive Lotka Volterra system. 3. Main Results A key result of this chapter is that a compact limit set of a competitive or cooperative system cannot contain two points x 1, x 2 satisfying x 1 x 2. If the system is cooperative and the limit set is an omega limit set, then this result follows from the Nonordering of Limit Sets from Chapter 1. The proof of the more general result is based on the next lemma. First, some notation is required. Let x(t) be a solution of the system (1.1) on an interval I. A subinterval [a, b] of I is called a rising interval if x(a) < x(b) and a falling interval if x(b) < x(a). Lemma 3.1. Let x(t) be a solution of the cooperative system (1.1) on the interval I. 8

Then x(t) cannot have a rising interval and a falling interval that are disjoint. Proof: The most important observation is that if [a, b] is a rising (falling) interval contained in I and if s > 0 is such that [a + s, b + s] is contained in I, then it is also a rising (falling) interval in I. In fact, if x(a) < x(b), then by Proposition 1.1, x(a + s) = φ s (x(a)) < φ s (x(b)) = x(b + s). Consequently, rising and falling intervals remain such under right translation. Suppose that I contains the falling interval [a, r] and the rising interval [s, b] and suppose that a < r < s < b. The other cases can be treated similarly. Let A = {t [s, b] : x(t) x(s)} and s = sup A. Then s s < b and [s, b] is a rising interval that contains no falling interval [s, r] for any r (s, b]. Redefine s = s so that the rising interval [s, b] has the above property. A contradiction will be reached in each of the two cases r a b s and r a > b s. If r a b s then [s, s + r a] is a translate to the right of [a, r], which is contained in I, and therefore is a falling interval. As s + r a b, this contradicts that [s, b] contains no such falling interval. If r a > b s then a < a+b r < s < b so [a+b r, b] is a translate to the right of [a, r], contained in I, so it is a falling interval. It follows that x(s) < x(b) < x(a+b r). Let c = sup{t [a + b r, s] : x(b) x(t)}. Then c < s < b and x(b) x(c) so [c, s] is a falling interval adjacent to the rising interval [s, b]. If s c b s then [s, 2s c] is a translate to the right of the falling interval [c, s], contained in [s, b], and therefore is a falling interval. But this contradicts the earlier argument that [s, b] contains no such interval. If s c > b s then c < c + b s < s and [c + b s, b] is a right translate of the falling interval [c, s] so it is a falling interval. Therefore x(b) < x(c + b s) and this contradicts the definition of c. The next result is the key tool for the main results of this chapter. Theorem 3.2. A compact limit set of a competitive or cooperative system cannot 9

contain two points related by. Proof: By time reversal if necessary, we can suppose that (1.1) is cooperative. The proof will be given in case that the limit set L is an alpha limit set, α(x 0 ). The reader should have no problem adapting the argument given in this case to the case that L is an omega limit set. Suppose that L = α(x 0 ) contains points x 1 and x 2 satisfying x 1 x 2. Let x(t) = φ t (x 0 ) for t 0. Since {x : x 1 x} is an open neighborhood of x 2 and the latter belongs to L, there exists t 1 < 0 such that x 1 x(t 1 ). As {x : x x(t 1 )} is an open neighborhood of x 1 and the latter belongs to L, there exists t 2 satisfying t 2 < t 1 and x(t 2 ) x(t 1 ). Continuing in this way, we find t 3 < t 2 such that x(t 3 ) x 2 and t 4 < t 3 such that x(t 3 ) x(t 4 ). Therefore, the interval I = [t 4, t 1 ] contains the falling interval [t 4, t 3 ] and the rising interval [t 2, t 1 ] and these intervals are disjoint. This contradicts Lemma 3.1, proving the Theorem in this case. A periodic orbit γ of a competitive or cooperative system is a compact limit set and consequently it cannot contain two points related by. Notice that this fact precludes the existence of periodic orbits of competitive or cooperative systems in IR 2 since any Jordan curve in the plane necessarily contains two points related by. The following sharper result concerning periodic orbits will be required later. Proposition 3.3. Let γ be a nontrivial periodic orbit of a competitive or cooperative system. Then γ cannot contain two points that are related by <. Proof: We can assume that the system is cooperative. Suppose that y i γ, i = 1, 2 satisfy y 1 < y 2. Let T > 0 be the minimal period of the solution φ t (y 1 ). There exists τ (0, T ) such that φ τ (y 1 ) = y 2 > y 1. By the Convergence Criterion of Chapter 1, the omega limit set ω(y 1 ) is a periodic orbit of period τ. But ω(y 1 ) = γ and this contradicts that T is the minimal period. 10

The next result is the main result of this Chapter. The following definitions will make it easier to state. Let A be an invariant set for (1.1) with flow φ t and let B be an invariant set for the system y = F (y) with flow ψ t. We say that the flow φ t on A is topologically equivalent to the flow ψ t on B provided there is a homeomorphism Q : A B such that Q(φ t (x)) = ψ t (Q(x)) for all x A and all t IR. The relationship of topological equivalence is one of several equivalence relations on the set of all flows that say, roughly, that the dynamics of the two flows are the same. A system of differential equations y = F (y), defined on IR k, is said to be Lipschitz if F is Lipschitz. That is, there exists K > 0 such that F (y 1 ) F (y 2 ) K y 1 y 2 for all y 1, y 2 IR k. Theorem 3.4. The flow on a compact limit set of a competitive or cooperative system in IR n is topologically equivalent to a flow on a compact invariant set of a Lipschitz system of differential equations in IR n 1. Proof: Let L be the limit set. Let v be a unit vector satisfying 0 v and let H v be the hyperplane orthogonal to v. H v consists of vectors x such that x v = 0, where is the standard dot (or scalar product) in IR n. Let Q be the orthogonal projection onto H v, that is Qx = x (x v)v. See Figure 3.1. By Theorem 3.2, Q is one-to-one on L (this could fail only if L contains two points that are related by ). Therefore, Q L, the restriction of Q to L, is a Lipschitz homeomorphism of L onto a compact subset of H v. We argue by contradiction to establish the existence of m > 0 such that Q L x 1 Q L x 2 m x 1 x 2 whenever x 1 x 2 are points of L. If this were false, then there exists sequences x n, y n L, x n y n such that Q(x n ) Q(y n ) x n y n 0 11

as n. Equivalently, (x n y n ) v[v (x n y n )] x n y n 0, or, w n v(v w n ) 0 as n where w n = x n y n x n y n. We can assume that w n w as n where w = 1. Then, w = v(v w) and therefore, (v w) 2 = 1 so w = ±v. But then x n y n x n y n ±v as n and this implies that x n y n or y n x n for all large n, contradicting Theorem 3.2. Therefore, Q 1 L is Lipschitz on Q(L). Since L is a limit set, it is an invariant set for (1.1). It follows that the dynamical system restricted to L can be modeled on a dynamical system in H v. In fact, if y Q(L) then y = Q L (x) for a unique x L and ψ t (y) Q L (φ t (x)) is a dynamical system on Q(L) generated by the vector field F (y) = Q L (f(q 1 L (y))) on Q(L). According to McShane(1934), a Lipschitz vector field on an arbitrary subset of H v can be extended to a Lipschitz vector field on all of H v, preserving the Lipschitz constant. It follows that F can be extended to all of H v as a Lipschitz vector field. It is easy to see that Q(L) is an invariant set for the latter vector field. We have established the topological equivalence of the flow φ t on L with the flow ψ t on Q(L). Q(L) is a compact invariant set for the (n 1)-dimensional dynamical system on H v generated by the extended vector field. 12

As a consequence of Theorem 3.4, the flow on a compact limit set, L, of a competitive or cooperative system shares common dynamical properties with the flow of a system of differential equations in one less dimension, restricted to a compact invariant set, Q(L). However, notice that Q(L) need not be a limit set of any orbit of the Lipschitz vector field. Obviously, Q(L) is connected since L is connected. A dynamical property possessed by all limit sets (or, more accurately, the flow restricted to the limit set) is that of chain recurrence Conley(1972,1978). The definition is as follows. Let A be a compact invariant set for the flow ψ t. Given two points z and y in A and positive numbers ɛ and t, an (ɛ, t)-chain from z to y in A is an ordered set {z = x 1, x 2,..., x n+1 = y; t 1, t 2,..., t n } (3.1) of points x i A and times t i t such that ψ ti (x i ) x i+1 < ɛ, i = 1, 2,..., n. (3.2) Figure 3.2 depicts such a chain. A is said to be chain recurrent if for every z A and for every ɛ > 0 and t > 0, there is an (ɛ, t)-chain from z to z in A. In Theorem 1 of the appendix, we prove that alpha and omega limit sets have the property of chain recurrence. It is easy to see that the flow ψ t on the compact invariant space Q(L) is chain recurrent since it is topologically equivalent to the flow φ t on the chain recurrent limit set L. 4. Three Dimensional Systems The Poincaré-Bendixson Theorem for three dimensional cooperative and competitive systems, proved below, is the most notable consequence of Theorem 3.4. However, it is harder to apply than its planar analog because it is more difficult to exclude that no equilibrium belongs to the limit set for a three dimensional system. Theorem 4.2 of 13

this section gives more easily checked conditions for most solutions to approach a nontrivial periodic orbit when there is only a single unstable equilibrium. The remainder of the section is devoted to describing the sense in which the existence of a periodic orbit implies the existence of a corresponding equilibrium inside it. Essentially, there is a topological 2-sphere, containing the periodic orbit as an equator, that bounds a positively invariant ball. This ball must contain at least one equilibrium. Theorem 4.1. A compact limit set of a competitive or cooperative system in IR 3 that contains no equilibrium points is a periodic orbit. Proof: Let L be the limit set. By Theorem 3.4, the flow φ t on L is topologically equivalent to the flow ψ t, generated by a Lipschitz planar vector field, restricted to the compact, connected, chain recurrent invariant set Q(L). Since L contains no equilibria neither does Q(L). The Poincaré-Bendixson Theorem implies that Q(L) consists of periodic orbits and, possibly, entire orbits whose omega and alpha limit sets are periodic orbits contained in Q(L). The chain recurrence of the flow ψ t on Q(L) will be exploited to show that Q(L) consists entirely of periodic orbits. Let z Q(L) and suppose that z does not belong to a periodic orbit. Then ω(z) and α(z) are distinct periodic orbits in Q(L). Let ω(z) = γ and suppose for definiteness that z belongs to the interior component, D, of IR 2 \ γ so that ψ t (z) spirals toward γ in D. The other case is treated similarly. Then γ is asymptotically stable relative to D. Standard arguments using transversals (see Figure 4.1) imply the existence of compact, positively invariant neighborhoods U 1 and U 2 of γ in D such that U 2 Int D U 1, z / U 1 and there exists t 0 > 0 for which ψ t (U 1 ) U 2 for t t 0. Let ɛ > 0 be such that the 2ɛ-neighborhood of U 2 in D is contained in U 1. Choose t 0 larger if necessary such that ψ t (z) U 2 for t t 0. This can be done since ω(z) = γ. Then any (ɛ, t 0 )-chain (3.1) in Q(L) beginning at x 1 = z satisfies ψ t1 (x 1 ) U 2 and, by (3.2) and the fact that the 2ɛ-neighborhood of U 2 is contained in U 1, x 2 U 1. As t 2 > t 0, 14

it then follows that ψ t2 (x 2 ) U 2 and (3.2) again implies that x 3 U 1. Continuing this argument, it is evident that the (ɛ, t 0 )-chain cannot return to z. There can be no (ɛ, t 0 )-chain in Q(L) from z to z and therefore we have contradicted that Q(L) is chain recurrent. Consequently, every orbit of Q(L) is periodic. Since Q(L) is connected, it is either a single periodic orbit or an annulus consisting of periodic orbits. It follows that L is either a single periodic orbit or a cylinder of periodic orbits. To complete the proof we must rule out the possibility that Q(L) consists of an annulus of periodic orbits. We can assume that the system is cooperative. The argument will be separated into two cases: L = ω(x) or L = α(x). If L = ω(x) consists of more than one periodic orbit then Q(L) is an annulus of periodic orbits in the plane containing an open subset O. Then there exists t 0 > 0 such that Q(φ t0 (x)) O. Let y be the unique point of L such that Q(y) = Q(φ t0 (x)). y = φ t0 (x) cannot hold since this would imply that L is a single periodic orbit so it follows that either y φ t0 (x) or φ t0 (x) y. Suppose that the latter holds, the argument is similar in the other case. Then there exists t 1 > t 0 such that φ t1 (x) is so near y that φ t0 (x) φ t1 (x). But then the Convergence Criterion from Chapter 1 implies that φ t (x) converges to equilibrium, a contradiction to our assumption that L contains no equilibria. This proves the theorem in this case. If L = α(x) and Q(L) consists of an annulus of periodic orbits, let C L be a periodic orbit such that Q(L) contains C in its interior. Q(C) separates Q(L) into two components. Fix a and b in L \ C such that Q(a) and Q(b) belong to different components of Q(L)\Q(C). Since φ t (x) repeatedly visits every neighborhood of a and b as t, Q(φ t (x)) must cross Q(C) at a sequence of times t k. Therefore, there exist z k C such that Q(z k ) = Q(φ tk (x)) and consequently, as in the previous case, either z k φ tk (x) or φ tk (x) z k holds for each k. Passing to a subsequence, we can assume that either z k φ tk (x) holds for all k or φ tk (x) z k holds for all k. Assume the latter, the argument is essentially the same in the other case. We claim 15

that for every s < 0 there is a point w C such that φ s (x) > w. For if t k < s then φ s (x) = φ s tk φ tk (x) < φ s tk (z k ) C. If y L then φ sn (x) y for some sequence s n. By the claim, there exists w n C such that φ sn (x) > w n. Passing to a subsequence if necessary, we can assume that w n w C and y w. Therefore, every point of L is related by to some point of C. The same reasoning applies to every periodic orbit C L for which Q(C ) belongs to the interior of Q(L): either every point of L is some point of C or every point of L is some point of C. Since there are three different periodic orbits in L whose projections are contained in the interior of Q(L), there will be two of them for which the same inequality holds between points of L and points of the orbit. Consider the case that there are two periodic orbits C 1 and C 2 such that every point of L is some point of C 1 and some point of C 2. The case that the opposite relations hold is treated similarly. If u C 1 then it belongs to L so we can find w C 2 such that u < w (equality can t hold since the points belong to different periodic orbits). But w L so we can find z C 1 such that w < z. Consequently, u, z C 1 satisfy u < z, a contradiction to Proposition 3.3. This completes the proof. A cooperative system of differential equations cannot have an attracting periodic orbit by Theorem 1.2.2 but a competitive system can. The next result gives sufficient conditions for most orbits of a three dimensional competitive system to be asymptotic to a periodic orbit. Recall that an equilibrium p is said to be hyperbolic if the Jacobian matrix of f at p has no purely imaginary eigenvalues. We use the notation W s (p) for the stable manifold of p (See e.g. Hale(1980)). Theorem 4.2. Let (1.1) be a competitive system in D IR 3 and suppose that D contains a unique equilibrium point p which is hyperbolic. Suppose further that W s (p) 16

is one-dimensional and tangent at p to a vector v 0. If q D \ W s (p) and γ + (q) has compact closure in D then ω(q) is a nontrivial periodic orbit. Proof: The result will follow from Theorem 4.1 provided p / ω(q). Clearly, ω(q) p since q / W s (p). If p ω(q) then the Butler-McGehee Lemma (Butler and Waltman (1986)) implies that ω(q) contains a point y on W s (p) different from p. By the invariance of ω(q) we may assume that y is a point of the local stable manifold, arbitrarily close to p. Since W s (p) is tangent at p to a positive vector, it can be assumed that either y p or p y. In either case we have contradicted Theorem 3.2 since both p and y belong to ω(q). The requirement that W s (p) be tangent to a positive vector can be verified by showing that the Jacobian matrix of f at p is irreducible. This follows from the Perron-Frobenius Theorem (Theorem 4.3.1). Indeed, the Jacobian matrix of f at p, Df(p), has the property that A = Df(p) + si 0 for all large values of s, where I denotes the identity matrix and the inequality signifies that each entry of the matrix is nonnegative. If Df(p) is irreducible then so is A and consequently, by the Perron- Frobenius Theorem, A has a positive eigenvector v corresponding to the eigenvalue r > 0 with maximum modulus. Then, v is an eigenvector for Df(p) corresponding to the eigenvalue s r. The eigenvalues λ of Df(p) give rise to the eigenvalues s λ of A, so, as r is the eigenvalue of A with maximal real part, s r is the eigenvalue of Df(p) of minimal real part. The hypotheses of Theorem 4.2 imply that s r < 0 and the corresponding eigenvector is v. See Corollary 4.3.2 for a similar argument. A remarkable fact about three dimensional competitive or cooperative systems on suitable domains is that the existence of a periodic orbit implies the existence of an equilibrium point inside a certain closed ball having the periodic orbit on its boundary. The next result makes this assertion precise. The proof is quite long and tedious, so it might be skipped on first reading. We can assume the system is competitive. Let γ 17

denote the periodic orbit and assume that there exist p, q with p q such that γ [p, q] D. (4.1) Define K = {x IR 3 : x is not related to any point y γ} = (γ + IR 3 +) c (γ IR 3 +) c. Here we use the notation B c for the complement of the subset B in IR 3. Another way to define K is as K c = y γ [(y + IR 3 +) (y IR 3 +)]. Proposition 4.3. Let γ be a non-trivial periodic orbit of a competitive system in D IR 3 and suppose that (4.1) holds. Then K is an open subset of IR 3 consisting of two connected components, one bounded and one unbounded. The bounded component, K(γ), is homeomorphic to the open unit ball in IR 3. K(γ) [p, q], is positively invariant and its closure contains an equilibrium. Proof: That K is open is a consequence of the fact that γ + IR 3 + and γ IR 3 + are closed. We show that K D is positively invariant. If x K D, y γ and t > 0 then φ t (y) γ so x is not related to it. Since the forward flow of a competitive system preserves the property of being unrelated, φ t (x) is unrelated to y. Therefore, φ t (x) K D. As in the proof of Theorem 3.4, let v 0 be a unit vector, H v be the hyperplane orthogonal to v and Q be the orthogonal projection onto H v along v. Q is one-to-one on γ so Q(γ) is a Jordan curve in H v. Let H i and H e denote the interior and exterior components of H v \ Q(γ). If x Q 1 (Q(γ)) then Q(x) = Q(y) for some y γ and therefore either x = y, x y or y x. In any case, x / K. Hence, K = (K Q 1 (H i )) (K Q 1 (H e )). 18

Set K(γ) = K Q 1 (H i ). There exists M > 0 such that for all z H i, z + tv γ for all t > M and z + tv γ for all t < M since v 0. It follows that {t IR : z + tv y for some y γ} = (, t (z)] where t (z) M and there exists y (z) γ such that z + t (z)v y (z). (4.2) The maximality of t (z) implies that y (z) and z + t (z)v share at least one common component. Thus M t (z). Similarly, {t IR : z + tv y for some y γ} = [t + (z), ) where t + (z) M and there exists y + (z) γ such that z + t + (z)v y + (z). (4.3) The minimality of t + (z) implies that y + (z) and z + t + (z)v share at least one common component. Thus t + (z) M. The inequalities (4.2) and (4.3) lead to (t + (z) t (z))v y + (z) y (z) or y (z) y + (z) + (t (z) t + (z))v. If t (z) > t + (z) then y (z) y + (z) which contradicts Proposition 3.2. If t (z) = t + (z) then y (z) y + (z) and by Proposition 3.3, it must be the case that y (z) = y + (z). Therefore, y + (z) z + t + (z)v = z + t (z)v y (z) and equality must hold so z + t + (z)v = y + (z) or z Q(γ). This contradiction proves that t (z) < t + (z). It follows that z + tv K for t (z) < t < t + (z). Hence K(γ) = {z + tv : z H i, t (z) < t < t + (z)}. Now we show that the maps z z + t (z)v and z z + t + (z)v are continuous on H i. Suppose z n, z H i and z n z as n. As M t (z n ) M, we can select a convergent subsequence t k = t (z nk ) such that t k t. By passing to a subsequence if necessary, we can also suppose that the corresponding subsequence y k = y (z nk ) converges to ȳ γ. Since z n + t (z n )v y (z n ) it follows that z + tv ȳ. But z + t (z)v y (z) and by the maximality of t (z) it follows that t t (z). If strict inequality holds, then, since v 0, z nk + t (z nk )v y (z) for all large k, 19

contradicting the maximality of t (z nk ) for these k. Consequently, t = t (z) and every convergent subsequence of t (z n ) must converge to t (z) and so t (z n ) t (z) as n. This establishes the continuity of the map z z + t (z)v on H i. Similar arguments show that t + (z) is continuous. Finally, suppose that w Q(γ). There is a unique y γ such that Q(y) = w or, equivalently, y = w + sv for s = y v. Given ɛ > 0 there is a neighborhood U of w in H v such that z + (s ɛ)v y z + (s + ɛ)v for all z U. Hence, s ɛ t (z) < t + (z) s + ɛ for all z U H i and therefore, t + (z) s and t (z) s as z w in H i. Since H i is homeomorphic to the open unit ball in IR 2, these arguments imply that K(γ) is homeomorphic to the open unit ball in IR 3. Consequently, it is connected. Similar arguments can be applied to H e to show that K Q 1 (H e ) = {z + tv : z H e, t (z) < t < t + (z)} and that the maps z z+t (z)v, z+t + (z)v are continuous. This gives the unbounded component of K. As K(γ) is a connected component of the positively invariant set K, it follows that it is positively invariant. Consequently, its closure is a positively invariant set homeomorphic to the closed unit ball in IR 3. It must contain an equilibrium by a standard argument using the Brouwer Fixed Point Theorem (e.g. Hale(1980), Thm.I.8.2). In order to show that K(γ) [p, q], it is convenient to use the same arguments as above only with the projection Q replaced by the projection onto the (x 1, x 2 )-plane. Let π : IR 3 IR 2 be defined by π(x 1, x 2, x 3 ) = (x 1, x 2, 0) and set e 3 = (0, 0, 1). Then π is one-to-one on γ by Proposition 3.3 so the image of γ under π is a Jordan curve, π(γ), in the plane. Let H i and H e denote the interior and exterior components of IR 2 \ π(γ). Then K = (K π 1 (H i )) (K π 1 (H e )) since every point of π 1 (π(γ)) is related to some point of γ and therefore cannot 20

belong to K. Now, the unbounded component of K contains points with arbitrarily large x 1 and x 2 components and it is mapped by π onto an open, connected subset of H i H e having π(γ) as part of its boundary. It follows that this component is projected into H e by π. On the other hand, π(k(γ)) is an open, bounded, connected subset of H i H e and π(γ) belongs to its boundary as well. Since H i π(k), as shown below, it follows that π(k(γ)) = H i. Similarly, H e π(k) so the projection of the unbounded component of K is H e. Therefore, K(γ) = K π 1 (H i ). We now show that H i π(k). Let z = (x 1, x 2, 0) H i. Then, there are points z, z π(γ) such that z z z. See Figure 4.2. Let z = π( x) and z = π( x) where x and x belong to γ. Consider the set A(z) = {t : z + te 3 y, for some y γ}. It is bounded above since max{y 3 : y γ} is bounded. If t x 3 then z + te 3 x so t A(z). It follows that A(z) = (, t (z)] where t (z) sup A(z) and there exists y (z) γ such that z + t (z)e 3 y (z). Similar arguments show that {t : z + te 3 y for some y γ} = [t + (z), ) and there exists y + (z) γ such that z + t + (z)e 3 y + (z). Observe that z π(y (z)), π(y + (z)) since the latter belong to π(γ). Also, t (z) = y (z) 3 and t + (z) = y + (z) 3. If t + (z) < t (z) then, as z+t (z)e 3 y (z) and z+t + (z)e 3 y + (z), we have y (z) y + (z) + (t (z) t + (z))e 3. Thus, y (z) > y + (z), contradicting Proposition 3.3. If t (z) = t + (z) then y + (z) z + t + (z)e 3 = z + t (z)e 3 y (z) so, by Proposition 3.3, we conclude that y (z) = y + (z) = z + t (z)e 3. Consequently, z = π(y + (z)) = π(y (z)), contradicting the observation above. Therefore, t (z) < t + (z) and z + te 3 K(γ) for t (z) < t < t + (z). We conclude that K(γ) = {z + te 3 : z H i, t (t (z), t + (z))}. Now we show that K(γ) [p, q]. If x K(γ) then x = z+te 3 for some z H i and some t (t (z), t + (z)). Since γ [p, q] it follows that π(γ) π[p, q] = [π(p), π(q)] and therefore, H i [π(p), π(q)]. Hence, p i x i = z i q i, i = 1, 2 for every x K(γ). Now, x 3 = t < t + (z) = y + (z) 3 and y + (z) γ implies y + (z) 3 q 3 by (3.6), so x 3 q 3. Similarly, x 3 > t (z) = y (z) 3 p 3. Therefore, p x q and K(γ) [p, q] D. 21

As the geometry of the set K(γ) is difficult to visualize, we offer an (artificial) example where at least it is possible to give an explicit description of K(γ). Let γ consist of the three line segments forming the boundary of the standard simplex S 2 = {x IR 3 + : x 1 + x 2 + x 3 = 1}, oriented as in Figure 4.3. Although γ cannot be a periodic orbit, it is a Jordan curve with the property that x < y does not hold for any two points x, y γ. These are the only properties that were used in the proof of Proposition 4.3. Then, K(γ) = {x : x i > 0, x 1 + x 2 < 1, x 1 + x 3 < 1, x 2 + x 3 < 1}. This is easy to see. If x y for some y γ, then it follows that x i + x j 1 for some i j, so x / K(γ). If x y for some y γ then x i 0 for some i, so x / K(γ). Therefore, no point of the set above contains a point that is related to a point of γ. On the other hand, if x belongs to any one of the six planes bounding K(γ), then it is related to a point of γ. For example, if x 1 + x 3 = 1 but each of the five other inequalities are satisfied, then x y (x 1, 0, x 3 ) γ. 5. Alternative Cones It is advantageous to consider other cones besides IR n + since by doing so, we can enlarge the class of competitive and cooperative systems. This is the primary motivation for this section. We restrict attention here to the case that the cone is one of the other orthants of IR n. While it is a straightforward exercise to translate the Kamke condition to analogous conditions for other partial orders, the problem of identifying a monotone system, with respect to some partial order, from the vector field becomes more difficult. That is, given a vector field, how are we to determine if it is a monotone system and what is the appropriate partial order. Much of this section is devoted to the identification problem. A planar competitive or cooperative system is both competitive and cooperative although the partial order is different in each case. Taking advantage of this fact and 22

the fact that these systems cannot have a nontrivial periodic orbit, we give a simple proof that all bounded solutions converge in Theorem 5.2. We begin with some notation identifying the various orthants. Let m = (m 1, m 2,..., m n ) where m i {0, 1} and K m = {x IR n : ( 1) m i x i 0, 1 i n}. K m is a cone in IR n and as such, it generates a partial order m defined by x m y if and only if y x K m. Equivalently, x i y i for those i for which m i = 0 and y i x i for those i for which m i = 1. We write x < m y when x m y and x y, and x m y when y x IntK m or, equivalently, when x i < y i for those i for which m i = 0 and y i < x i for those i for which m i = 1. Let P be the diagonal matrix defined by P = diag[( 1) m 1, ( 1) m 2,..., ( 1) m n ]. Observe that P = P 1 is an order isomorphism, that is x m y if and only if P x P y. The domain D is said to be p m -convex if tx + (1 t)y D whenever x, y D and x m y. (1.1) is cooperative with respect to K m provided D is p m -convex and ( 1) m i+m j f i x j (x) 0, i j, x D. (5.1) It is competitive with respect to K m if the reverse inequality holds ( 1) m i+m j f i x j (x) 0, i j, x D. (5.2) If m = 0 then we recover the earlier definitions. Proposition 5.1. Let D be p m -convex and f be a continuously differentiable vector field on D such that (5.1) holds. Let < r denote any one of the relations m, < m, m. If x < r y, t > 0 and φ t (x) and φ t (y) are defined, then φ t (x) < r φ t (y). If (5.2) holds, then similar conclusions are valid for t < 0. Proof: P D is p-convex and g : P D IR n, defined by g(y) = P f(p y), satisfies (1.3) if (5.1) holds. The vector field g generates a flow ψ t defined by ψ t (y) = P φ t (P y). If 23

x m y, it follows that P x P y and therefore, by Proposition 1.1, ψ t (P x) ψ t (P y). This implies that φ t (x) m φ t (y), as asserted. The other assertions follow similarly. Another way to express Proposition 5.1 is to say that (1.1) is cooperative (competitive) with respect to K m if and only if, in the variable y = P x, the resulting system is cooperative (competitive), that is, y = g(y), g(y) P f(p y) is cooperative (competitive), as defined at the beginning of this chapter. It is clear that all the results of this chapter that hold for cooperative (competitive) systems also hold for systems that are cooperative (competitive) with respect to K m for some m. Proposition 5.1 suggests an algorithm for determining whether a given system (1.1) is cooperative or competitive in a domain D with respect to one of the cones K m. It clearly suffices to consider only the cooperative case since the competitive case can be determined by time reversal. The first test is to check that the off-diagonal elements of the Jacobian matrix are sign-stable in D. This means that for each i j, either (a) f i x j (x) 0 for all x D, or (b) f i x j (x) 0 for all x D. Assuming this test is passed, then the Jacobian matrix must be tested for sign-symmetry: f i x j (x) f j x i (y) 0 for all i j, and all x, y D. If this test is satisfied, then for each i < j set s ij = 0 if f i x j (x) + f j x i (x) > 0 for some x D, s ij = 1 if f i x j (x) + f j x i (x) < 0 for some x D, and let s ij {0, 1} be arbitrary if f i x j (x) + f j x i (x) vanishes identically in D. Now, consider the system of n(n 1)/2 linear equations in the n unknowns m i {0, 1}, given by m i + m j = s ij (mod 2), i < j, (5.3) Those equations among (5.3) for which s ij is arbitrary may be deleted since they can be satisfied by appropriate choice of s ij. If the remaining equations can be solved for m then (1.1) is cooperative with respect to K m. It can be seen that if m solves (5.3) then so does m + 1. 24

There is a relatively simple graph theoretic test for determining if a given system is cooperative or competitive in a domain D with respect to some K m. Assume that the system is sign-stable and sign-symmetric in the domain D. Consider the graph G with vertices {1, 2,..., n} where an undirected edge connects vertices i and j if at least one of the partial derivatives f i x j (x) or f j x i (x) does not vanish identically in D. Attach a sign + or to the edge depending on the sign of one of the partial derivatives at a point where it is nonzero. Then (1.1) is cooperative in D with respect to some K m if and only if for every closed loop in G the number of edges with signs is even; it is competitive if every closed loop in G has an odd number of edges with signs. If the test indicates that the system is cooperative, then the appropriate cone K m must then be determined by solving (5.3). If the test indicates that the system is competitive, then the time-reversed system is cooperative and the appropriate cone is determined by solving (5.3) after all entries of the Jacobian have been multiplied by 1 to account for time-reversal. A canonical form for a system that is cooperative with respect to K m for some m can be obtained by permuting equations and variables in the same way so that m consists of k ones followed by n k minus ones. This leads to a system for x = (x 1, x 2 ) IR k IR n k where x 1 IR k and x 2 IR n k of the form x 1 = f 1 (x 1, x 2 ) x 2 = f 2 (x 1, x 2 ). The k-by-k ((n k)-by-(n k)) Jacobian f 1 / x 1 ( f 2 / x 2 ) has nonnegative offdiagonal entries, while the k-by-(n k) ((n k)-by-k) Jacobian f 1 / x 2 ( f 2 / x 1 ) has nonpositive entries. A planar competitive system is cooperative with respect to K m where m = (0, 1) provided that D is p m -convex. That is, the forward flow is monotone with respect to the partial order generated by the cone K m = {(x 1, x 2 ) IR 2 : x 1 0, x 2 0}. This observation leads to a simple proof of the following Theorem. 25

Theorem 5.2. Let (1.1) be a competitive or cooperative system on a convex domain D in IR 2. If x(t) is a solution defined for all t 0 (t 0), then there exists a T 0 such that for each i = 1, 2, x i (t) is monotone on t T (t T ). In particular, if γ + (x(0)) (γ (x(0))) has compact closure in D, then ω(x(0)) (α(x(0))) is a single equilibrium. Proof: It suffices to consider only the case that x(t) is defined for t 0 since otherwise the corresponding solution of the time-reversed system is so defined and this system is also competitive or cooperative. Furthermore, we can assume that the system is cooperative since if it is competitive, then the transformation y = P x, where P = diag[1, 1], results in a cooperative system. By Proposition 2.1, P + = {x D : 0 f(x)} and P = {x D : f(x) 0} are positively invariant for (1.1) and if x(t 0 ) P + (x(t 0 ) P ) then x i (t) is nondecreasing (nonincreasing) for t t 0, i = 1, 2. Therefore, the result is proved if x(t) P + P for some t 0. If this doesn t occur, that is if x(t) P + P for all t 0, then for each t 0 either 0 m f(x(t)) or f(x(t)) m 0, where m = (0, 1). In other words, the vector f(x(t)) must always point into the open second or fourth quadrant of the plane. But {f(x(t)) : t 0} is connected and therefore, either 0 m f(x(t)) holds for all t 0 or f(x(t)) m 0 holds for all t 0. In the first case, x 1 (t) is strictly increasing and x 2 (t) is strictly decreasing and in the second case the reverse holds. Thus we have established that x i (t) is eventually monotone for i = 1, 2. The domain D was assumed to be convex because in the case that, after a possible time-reversal, (1.1) is competitive, then p-convexity of D is implied but in order for the transformed system in the y-variable (defined on P (D)) to be cooperative, P (D) must be also be p-convex. 6. The Field-Noyes Model 26

The following example illustratives the use of Theorem 4.2. The differential equations below are a scaled version of the Field-Noyes model of the Belousov-Zhabotinski reaction, due to Murray(1989). ɛx = y xy + x(1 qx) y = y xy + 2fz (6.1) z = δ(x z). The parameters ɛ, q, f, δ are positive and, in the interesting case, 0 < q < 1. The cube D = {(x, y, z) : 1 < x < q 1, 2fq 1 + q < y < f q, 1 < z < q 1 } is positively invariant since the vector field points strictly inward on the boundary of D. The Jacobian matrix of the vector field at a point (x, y, z) is given by J = 1 y 2qx ɛ 1 x ɛ 0 y 1 x 2f δ 0 δ It is immediately apparent that the off-diagonal entries of J are sign-stable and signsymmetric in D. The incidence graph of J on the vertices x, y, z is given in Figure 6.1 below. It consists of a single loop with one minus sign on the edge connecting x to y. Therefore, the system is competitive in D. In order to determine the appropriate partial order relation for the time-reversed cooperative system, which has Jacobian J, we need to solve (5.3) where the s ij are determined from J. This leads to the. system m 1 + m 2 = 0 m 1 + m 3 = 1. m 2 + m 3 = 1 A solution is m 1 = m 2 = 0, m 3 = 1. This corresponds to the cone of positive vectors given by K m = {(x, y, z) IR 3 : x 0, y 0, z 0}. 27

The reader may have guessed immediately that changing z to z in (6.1) results in a system satisfying (1.3). The system (6.1) has a unique equilibrium p = (x, y, z ) in D given by x = z, y = 2fx 1 + x, q(x ) 2 + (2f + q 1)x (2f + 1) = 0, taking the positive root of the quadratic. The determinant of J at p is Det(J) = δ ɛ (2y + qx + q(x ) 2 ), where we used the quadratic equation determining x to simplify the expression. As the determinant is negative, there are two possibilities if we consider only the case that p is hyperbolic. Either p is asymptotically stable or it is unstable with a one dimensional stable manifold. Both cases can occur, depending on the values of the parameters (see Murray(1989)). Theorem 4.2 can be applied in the second case and leads to the following. Proposition 6.1. Suppose that p is hyperbolic and unstable for (6.1). Then the stable manifold of p, W s (p), is one dimensional and ω(q) is a nontrivial periodic orbit in D for every point q D \ W s (p). Proof: By Theorem 4.2, it is only necessary to check that the one dimensional stable manifold is tangent at p to a positive vector, that is, to a vector in K m. Since J at p is irreducible, this follows from the discussion following Theorem 4.2, taking into account the order relation generated by K m. Figure 6.2 illustrates the Proposition. 7. Remarks and Discussion Proposition 1.1 is due to Kamke((1932) and Muller(1926). See Coppel(1965) or Walter(1970) for more general results. Competitive and cooperative systems were 28