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

Generic behavior of differentially positive systems
on a globally orderable Riemannian manifold

Lin Niu
School of Mathematics and Physics
University of Science and Technology Beijing
Beijing, 100083, P. R. China
Yi Wang
School of Mathematical Sciences
University of Science and Technology of China
Hefei, Anhui, 230026, P. R. China
Supported by the National Natural Science Foundation of China No.12201034 and 12090012.Supported by the National Natural Science Foundation of China No.12331006.
Abstract

Differentially positive systems are the nonlinear systems whose linearization along trajectories preserves a cone field on a smooth Riemannian manifold. One of the embryonic forms for cone fields in reality is originated from the general relativity. By utilizing the Perron-Frobenius vector fields and the Γ\Gamma-invariance of cone fields, we show that generic (i.e.,“almost all” in the topological sense) orbits are convergent to certain single equilibrium. This solved a reduced version of Forni-Sepulchre’s conjecture in 2016 for globally orderable manifolds.

Keywords: Differential positivity, Causal order, Generic convergence, Homogeneous space, Cone field

AMS Subject Classification (2020): 37C20, 37C65, 53C30, 22F30

1 Introduction

Differential analysis provides a general framework for investigating a nonlinear dynamical system by analyzing the linearization of the system at every point in the state space. The motivation is that the local behavior of a system often has strong implications for the global nonlinear behavior. The current series of papers focus on the nonlinear dynamical system whose linearization along trajectories preserves a cone field. Here, a cone field CMC_{M} assigns to each point xx of a manifold MM a closed convex cone CM(x)C_{M}(x) in the tangent space TxMT_{x}M of xx.

One of the embryonic forms for cone fields in reality is originated from the general relativity, for which the time-orientable space-time (i.e., a connected 4-dimensional Hausdorff CC^{\infty}-manifold endowed with a Lorentz metric) naturally generates at each point a Lorentzian cone in tangent space (see, e.g. [2, 12, 43] or Appendix A.1). The structure of a cone field in the time-orientable space-time turns out to be one of the crucial mathematical tools in the study of general relativity including causality theory, singularity theory and black holes, etc.. Among others, the well-known non-spacelike curve (also called causal curve) in these theories (c.f. Hawking and Ellis [12] and Penrose [43]) is actually one whose tangent vector at each point falls in the Lorentzian cone field. Moreover, the causally related points in a space-time should be joined by a non-spacelike curve.

From this point of view, a cone field on the manifold MM naturally induces a “conal order relation” as follows: two points x,yMx,y\in M are conal ordered, denoted by xMyx\leq_{M}y, if there exists a conal curve on MM beginning at xx and ending at yy. Here, a conal curve is a piecewise smooth curve whose tangent vector lies in the cone at every point along the curve wherever it is defined (see Definition 2.1 and Fig.1(a)). In particular, in the setting of space-times (see [48, 11, 29] and [16, 42, 54]), the causal curves are actually the conal curves; while, the causally related points are the ones that are conal ordered. Besides, conal order also has various applications in hyperbolic partial differential equations and harmonic analysis (see [5, 6]), as well as in the theory of Wiener-Hopf operators on symmetric spaces [18].

It deserves to point out that, unlike the standard order relation induced by a single closed convex cone in a topological vector space, the conal order relation “M\leq_{M}” induced by a cone field on MM is certainly reflexive and transitive, but not always antisymmetric. In fact, there are examples of MM that contain the closed conal curves (e.g., the closed timelike curves in time-orientable space-times [12, Chapter 5]), which reveals that the anti-symmetry fails. Moreover, the relation “M\leq_{M}” is not necessarily closed, i.e., the set {(x,y)M×M:xMy}\{(x,y)\in M\times M:x\leq_{M}y\} is not a closed subset in M×MM\times M (see [26, p.299] or [35, p.470]). For instance, such set is not necessarily closed in Minkowski space (see [12, p.183] or [43, p.12]).

In many recent works [8, 7, 9, 34, 33], the nonlinear system whose linearization along trajectories preserves a convex cone field is also referred as differentially positive system, the flow of which naturally keeps the conal order “M\leq_{M}”. To the best of our knowledge, Forni and Sepulchre [8, 7, 9] first studied the dynamics of differentially positive systems. Among others, they constructed on MM a canonical defined vector field 𝒲\mathcal{W}, called the Perron-Frobenius vector field, such that the vector 𝒲(x)\mathcal{W}(x) lines in the interior of the cone CM(x)C_{M}(x) at each point xMx\in M. By appealing to the Perron-Frobenius vector field 𝒲\mathcal{W}, a dichotomy characterization of limit sets for differentially positive systems was provided. Based on this, Forni and Sepulchre [9, p.353] further posed the following

Conjecture ([Forni and Sepulchre [9]). For almost every xMx\in M, the ω\omega-limit set ω(x)\omega(x) is given by either a fixed point, or a limit cycle, or fixed points and connecting arcs.

This conjecture predicts what typically happens to orbits, as time increases to ++\infty. The description of typical properties of orbits usually means those shared by “most” orbits.

In our present work, we will tackle this conjecture and made an attempt to establish the asymptotic behavior of “most” orbits for differentially positive systems. For this purpose, we first introduce the following assumption for MM:

  1. (H1)

    MM is globally orderable equipped with a continuous solid cone field.

(H1) means that the conal order “M\leq_{M}” on MM is a partial order (see [26, 16, 17, 35] and Definition 2.2); and hence, it is actually anti-symmetric. This occurs naturally in various situations. One of the well-known examples is a homogeneous space of positive definite matrices with the affine-invariant cone field (see, e.g, [32, Section 3]). As a matter of fact, (H1) excludes the occurrence of the closed conal curves (see [26, Section 5]). In the setting of space-times, such closed conal (causal) curves would seem to generate paradoxes involving causality and are said to “violate causality” ([2, 12, 13, 29, 43]).

As a consequence, by (H1), the possibility of a limit cycle, or fixed points and connecting arcs (which are all closed conal curves) can be ruled out. In other words, Forni-Sepulchre’s Conjecture is naturally reduced to the following

Conjecture A. For almost every xMx\in M, the ω\omega-limit set ω(x)\omega(x) is a singleton.

In the present paper, we will prove conjecture A under the following two reasonable assumptions:

  1. (H2)

    The conal order “M\leq_{M}” is quasi-closed.

  2. (H3)

    Both the cone field CMC_{M} and the Riemannian metric on MM are Γ\Gamma-invariant.

More precisely, our main result is the following

Theorem A.

Assume that (H1)-(H3) hold. Then, for almost every xMx\in M, the ω\omega-limit set ω(x)\omega(x) is a singleton.

This theorem, in a detailed version, will be proved in Section 3 (see Theorem 3.1). It concludes that generic (i.e.,“almost all” in the topological sense) orbits are convergent to a single equilibrium.

The quasi-closedness of “M\leq_{M}” in (H2) means that xMyx\leq_{M}y whenever xnxx_{n}\to x and ynyy_{n}\to y as nn\to\infty and xnMynx_{n}\ll_{M}y_{n} for all integer n0n\geq 0 (see Definition 2.3). Here, we write xMyx\ll_{M}y if there exists a so-called strictly conal curve (whose tangent vector lies in the interior of cone at every point along the curve) on MM beginning at xx and ending at yy (see Definition 2.1). We point out that (H2) is motivated by the causal continuity of space-times in Hawking and Sachs [13], while the notation ``M"``\ll_{M}" is derived from the timelike curve in general relativity (see e.g., [43]). In fact, one can prove that the conal order ``M"``\leq_{M}" is indeed quasi-closed in many causally continuous space-times (see Theorem A.1 in the Appendix).

As for (H3), a cone field CMC_{M} is called Γ\Gamma-invariant if there exists a linear invertible mapping Γ(x1,x2):Tx1MTx2M\Gamma(x_{1},x_{2}):T_{x_{1}}M\to T_{x_{2}}M for all x1,x2Mx_{1},x_{2}\in M, such that Γ(x1,x2)CM(x1)=CM(x2)\Gamma(x_{1},x_{2})C_{M}(x_{1})=C_{M}(x_{2}). While, the Riemannian metric (,)x(\cdot,\cdot)_{x} on MM is Γ\Gamma-invariant if (u,v)x1=(Γ(x1,x2)u,Γ(x1,x2)v)x2(u,v)_{x_{1}}=(\Gamma(x_{1},x_{2})u,\Gamma(x_{1},x_{2})v)_{x_{2}} for all x1,x2Mx_{1},x_{2}\in M and u,vTx1Mu,v\in T_{x_{1}}M. Actually, Γ\Gamma-invariance is motivated by the homogeneous structure of the manifolds. A well-known example admitting (H3) is a homogeneous space M=G/HM=G/H of a connected Lie group GG assigned to each point xMx\in M a closed convex cone in the tangent space TxMT_{x}M such that the cone field is invariant under the action of GG (see e.g., [26, 35, 17]); and a homogeneous Riemannian metric on such homogeneous space is naturally Γ\Gamma-invariant (see e.g., [1, 14]). In the Appendix A.2, more detailed structures of the globally orderable homogeneous spaces are presented. We further mention here that the homogeneous cone field comes from the Lie theory. It is shown that cone fields arise as quotients these so-called Lie wedge fields (see [16, Lemma VI.1.5]). This provides a crucial link between the Lie theory of semigroups and that of cone fields on manifolds (see [16, 17, 19, 26, 35]). Very recently, the works in [36, 37, 38, 31, 39] form an ongoing project aiming at the connections between causal structures on homogeneous spaces, algebraic quantum field theory, modular theory of operator algebras and unitary representations of Lie groups.

Needless to say, differential positivity in Forni and Sepulchre [9] with respect to a constant cone field on a flat space MM is reduced to the classical monotonicity ([21, 22, 50, 51, 4, 23, 49, 45, 44, 52, 15, 30, 46, 53, 55, 56]) with respect to a closed convex cone. From this point of view, differentially positive systems can be regarded as a natural generalization of the so-called classical monotone systems to nonlinear manifolds. We pointed out that the order introduced by a closed convex cone is a closed partial order (see the detail in Appendix A.3). In such a flat space, Theorem A automatically implies the celebrated Hirsch’s Generic Convergence Theorem [21].

The paper is organized as follows. In Section 22, we introduce some notations and definitions and summarize some preliminary results. Theorem A (i.e., Theorem 3.1) with its proof will be given in Section 33. Several fundamental tools and critical lemmas, which turns out be useful in the proof of Theorem A (or Theorem 3.1), will be postponed in Section 44. Finally, in the Appendix, we will present the order structures on space-times, the globally orderable homogeneous spaces with homogeneous cone fields, as well as the differential positivity in the flat spaces.

2 Notations and preliminary results

Throughout this paper, MM will be reserved as a smooth manifold of dimension nn. The tangent bundle is denoted by TMTM and the tangent space at a point xMx\in M by TxMT_{x}M. MM is endowed with a Riemannian metric tensor, represented by a inner product (,)x(\cdot,\cdot)_{x} on the tangent space TxMT_{x}M. We set |v|x:=(v,v)x\lvert v\rvert_{x}:=\sqrt{(v,v)_{x}} for any vTxMv\in T_{x}M (sometimes we omit the subscripts xx in ||x\lvert\cdot\rvert_{x}) and the Riemannian metric endows the manifold with the Riemannian distance dd. We assume that (M,d)(M,d) is a complete metric space.

Let EE be a nn dimensional real linear space. A nonempty closed subset CC of the linear space EE is called a closed convex cone if C+CCC+C\subset C, αCC\alpha C\subset C for all α0\alpha\geq 0, and C(C)={0}C\cap(-C)=\{0\}. A convex cone CC is solid if its interior IntC\text{Int}C\neq\emptyset. A convex cone CC^{\prime} in EE is said to surround a cone CC if C{0}IntCC\setminus\{0\}\subset\text{Int}C^{\prime}.

A cone field on a manifold MM is a map xCM(x)x\mapsto C_{M}(x), such that CM(x)C_{M}(x) is a closed convex cone in TxMT_{x}M for each xMx\in M. A manifold equipped with a cone field is called a conal manifold. A cone field is solid if each convex cone CM(x)C_{M}(x) is solid.

A cone field CMC_{M} is called a Γ\Gamma-invariant cone field if there exists a linear invertible mapping Γ(x1,x2):Tx1MTx2M\Gamma(x_{1},x_{2}):T_{x_{1}}M\to T_{x_{2}}M for each x1,x2Mx_{1},x_{2}\in M, such that Γ(x1,x2)CM(x1)=CM(x2)\Gamma(x_{1},x_{2})C_{M}(x_{1})=C_{M}(x_{2}). Moreover, Γ\Gamma is continuous with respect to (x1,x2)(x_{1},x_{2}) and Γ(x,x)=Idx\Gamma(x,x)=\text{Id}_{x} for all xMx\in M, where Idx:TxMTxM\text{Id}_{x}:T_{x}M\to T_{x}M is the identity map. We say that the Riemannian metric is Γ\Gamma-invariant under the linear invertible mapping Γ\Gamma if (u,v)x1=(Γ(x1,x2)u,Γ(x1,x2)v)x2(u,v)_{x_{1}}=(\Gamma(x_{1},x_{2})u,\Gamma(x_{1},x_{2})v)_{x_{2}} for all x1,x2Mx_{1},x_{2}\in M and u,vTx1Mu,v\in T_{x_{1}}M. In this case, |v|x1=|Γ(x1,x2)v|x2\lvert v\rvert_{x_{1}}=\lvert\Gamma(x_{1},x_{2})v\rvert_{x_{2}} holds for all x1,x2Mx_{1},x_{2}\in M and vTx1Mv\in T_{x_{1}}M.

Pick a smooth chart Φ:UV\Phi:U\to V, where UMU\subset M is an open set containing xx and VnV\subset\mathbb{R}^{n} is an open set. Let CMΦ(y)nC_{M}^{\Phi}(y)\subset\mathbb{R}^{n} denote the representation of the cone field in the chart, i.e., dΦ(y)(CM(y))={Φ(y)}×CMΦ(y)d\Phi(y)(C_{M}(y))=\{\Phi(y)\}\times C_{M}^{\Phi}(y). Following [26], a cone field xCM(x)x\mapsto C_{M}(x) on MM is upper semicontinuous at xMx\in M if given a smooth chart Φ:Un\Phi:U\to\mathbb{R}^{n} at xx and a convex cone CC^{\prime} surrounding CMΦ(x)C_{M}^{\Phi}(x), there exists a neighborhood WW of xx such that CMΦ(y)CC_{M}^{\Phi}(y)\subset C^{\prime} for all yWy\in W. The cone field is lower semicontinuous at xx if given any open set NN such that NCMΦ(x)N\cap C_{M}^{\Phi}(x) is non-empty, there exists a neighborhood WW of xx such that NCMΦ(y)N\cap C_{M}^{\Phi}(y) is non-empty for all yWy\in W. A cone field is continuous at xx if it is both lower and upper semicontinuous at xx and continuous if it is continuous at every xx.

Definition 2.1.

A continuous piecewise smooth curve tγ(t)t\mapsto\gamma(t) defined on [t0,t1][t_{0},t_{1}] into a conal manifold MM is called a conal curve if γ(t)CM(γ(t))\gamma^{\prime}(t)\in C_{M}(\gamma(t)) whenever t0t<t1t_{0}\leq t<t_{1}, in which the derivative is the right hand derivative at those finitely many points where the derivative is not continuous. (See Fig.1(a)). Moreover, γ\gamma is called a strictly conal curve if γ(t)IntCM(γ(t))\gamma^{\prime}(t)\in\text{Int}C_{M}(\gamma(t)) for t0t<t1t_{0}\leq t<t_{1}.

For two points x,yMx,y\in M, we say xx and yy are ordered, denoted by xMyx\leq_{M}y, if there exists a conal curve α:[t0,t1]M\alpha:[t_{0},t_{1}]\subset\mathbb{R}\to M such that α(t0)=x\alpha(t_{0})=x and α(t1)=y\alpha(t_{1})=y. This relation is an order on MM. In fact, it is always reflexive (i.e., xMxx\leq_{M}x for all xMx\in M) and transitive (i.e., xMyx\leq_{M}y and yMzy\leq_{M}z implies xMzx\leq_{M}z). The order “M\leq_{M}” is referred to as the conal order. We write xMyx\ll_{M}y if there exists a so-called strictly conal curve γ\gamma with γ(t0)=x\gamma(t_{0})=x and γ(t1)=y\gamma(t_{1})=y. Clearly, the relation “M\ll_{M}” is always transitive. Let U,VU,V be the subsets of MM. We write UMVU\leq_{M}V (resp., UMVU\ll_{M}V) if xMyx\leq_{M}y (resp., xMyx\ll_{M}y) for any xUx\in U and yVy\in V.

The following proposition implies that the relation “M\ll_{M}” is open.

Proposition 2.1.

If xMyx\ll_{M}y, then there exist neighborhoods UU of xx, VV of yy such that UMVU\ll_{M}V.

Proof.

We only prove the existence of VV, i.e., if xMyx\ll_{M}y, there exists a neighborhood VV of yy such that xMzx\ll_{M}z for each zVz\in V. Since xMyx\ll_{M}y, there is a strictly conal curve γ1:[t0,t1]M\gamma_{1}:[t_{0},t_{1}]\subset\mathbb{R}\to M such that γ1(t0)=x\gamma_{1}(t_{0})=x and γ1(t1)=y\gamma_{1}(t_{1})=y. Let γ(t)=γ1(t1t)\gamma(t)=\gamma_{1}(t_{1}-t), t[0,t1t0]t\in[0,t_{1}-t_{0}]. Then γ(0)=y\gamma(0)=y, γ(t1t0)=x\gamma(t_{1}-t_{0})=x and γ(0)Int{CM(γ(0))}\gamma^{\prime}(0)\in\text{Int}\{-C_{M}(\gamma(0))\}.

Let Φ:UΦ(U)n\Phi:U\to\Phi(U)\subset\mathbb{R}^{n} be a chart with Φ(γ(0))=0\Phi(\gamma(0))=0 such that Φ(U)\Phi(U) is convex. We set C~(Φ(z))=dΦ(z)(CM(z))-\widetilde{C}(\Phi(z))=d\Phi(z)(-C_{M}(z)) for zUz\in U and α=Φγ\alpha=\Phi\circ\gamma. Then α(0)Int{C~(0)}\alpha^{\prime}(0)\in\text{Int}\{-\widetilde{C}(0)\}. So, we can choose a compact convex neighborhood BB of α(0)\alpha^{\prime}(0) in the interior of C~(0)-\widetilde{C}(0) and set W=+B:={λB:λ>0}C~(0)W=\mathbb{R}^{+}B:=\{\lambda B:\lambda>0\}\subset-\widetilde{C}(0). Since the cone field is continuous, there is a neighborhood VMV\subset M of γ(0)\gamma(0) such that WC~(z)W\subset-\widetilde{C}(z) for all zU=Φ(V)z\in U^{\prime}=\Phi(V) and UU^{\prime} is convex.

On the other hand, α(0)=limt0+1tα(t)IntW\alpha^{\prime}(0)=\lim\limits_{t\to 0^{+}}\frac{1}{t}\alpha(t)\in\text{Int}W. Then there is a s0>0s_{0}>0 such that α((0,s0])IntW\alpha((0,s_{0}])\subset\text{Int}W and α([0,s0])U\alpha([0,s_{0}])\subset U^{\prime}. Let BB^{\prime} be an open convex neighborhood of α(s0)\alpha(s_{0}) in WUW\cap U^{\prime}.

Let B′′={α(s0)u:uB}UB^{\prime\prime}=\{\alpha(s_{0})-u:u\in B^{\prime}\}\cap U^{\prime}, then B′′B^{\prime\prime} is an open neighborhood of α(0)\alpha(0). For each vB′′v\in B^{\prime\prime}, we set αv(t)=α(s0)+t(vα(s0))\alpha_{v}(t)=\alpha(s_{0})+t(v-\alpha(s_{0})), t[0,1]t\in[0,1]. Then αv(0)=α(s0)\alpha_{v}(0)=\alpha(s_{0}), αv(1)=v\alpha_{v}(1)=v and αv(t)=vα(s0)BC~(αv(t))\alpha_{v}^{{}^{\prime}}(t)=v-\alpha(s_{0})\in-B^{\prime}\subset\widetilde{C}(\alpha_{v}(t)).

Thus, there is a point pγ1p\in\gamma_{1} and an open neighborhood DD of yy such that pMzp\ll_{M}z for all zDz\in D. Since xMpx\ll_{M}p, we obtain xMzx\ll_{M}z for all zDz\in D. ∎

According to our standing assumption (H1) in the introduction, we now give the following crucial definitions:

Definition 2.2.

A conal manifold is said to be globally orderable if the order “M\leq_{M}” is a partial order which is locally order convex.

The conal order “M\leq_{M}” is a partial order relation if it is additionally antisymmetric (i.e., xMyx\leq_{M}y and yMxy\leq_{M}x implies x=yx=y). A partial order “\leq” is locally order convex if it has a basis of neighborhoods at each point that are order convex (z,xUz,x\in U implies {y:zyz}U\{y:z\leq y\leq z\}\subset U). See [26, 35].

Remark 2.1.

In the terminology of general relativity, each point pp in a globally orderable manifold MM is called a strong point (see Lawson [26, Section 5]), by which means that every neighborhood UU of pp contains a smaller neighborhood VV of pp such that every conal curve that begins in VV and leaves UU terminates outside of VV (see e.g., [12, p.192], [43, p.28] or [2, p.59]).

Remark 2.2.

The globality of the conal order, i.e., whether the conal order associated with a cone field can be extended to a partial order on the global manifold, is a central problem in [26]. The equivalence between globality of the conal order in a homogeneous manifold and globality of the Lie wedge in the Lie group has been shown in [26, Section 5] (see also [35, Theorem 1.6] and [17, Section 4.3]). Globality of a conal order occurs naturally in various situations; for example, the conal order induced by the affine-invariant cone field on the homogeneous space of positive definite matrices, as we mentioned above, is a partial order (see [32, Theorem 2]). Besides, if the manifold is globally orderable, one can exclude the occurrence of the closed conal curves (see [26, Section 5]).

Definition 2.3.

The order “M\leq_{M}” is quasi-closed if xMyx\leq_{M}y whenever xnxx_{n}\to x and ynyy_{n}\to y as nn\to\infty and xnMynx_{n}\ll_{M}y_{n} for all nn.

Remark 2.3.

In general, the conal order is not a closed order (xMyx\leq_{M}y whenever xnxx_{n}\to x and ynyy_{n}\to y as nn\to\infty and xnMynx_{n}\leq_{M}y_{n}), see [26, p.299] and [35, p.470]. The quasi-closed order relationship here is inspired by the causal continuity of space-times studied by Hawking and Sachs in [13]. One may refer to Appendix A.1 for more details.

Now, we consider a system Σ\Sigma generated by a smooth vector field ff on MM equipped with a continuous solid cone field CMC_{M}. The induced flow by system Σ\Sigma is denoted by φt\varphi_{t}. We write dφt(x)d\varphi_{t}(x) as the tangent map from TxMT_{x}M to Tφt(x)MT_{\varphi_{t}(x)}M. Let xMx\in M, the positive semiorbit (resp., negative semiorbit) of xx, be denoted by O+(x)={φt(x):t0}O^{+}(x)=\{\varphi_{t}(x):t\geq 0\} (resp., O(x)={φt(x):t0}O^{-}(x)=\{\varphi_{t}(x):t\leq 0\}). The full orbit of xx is denoted by O(x)=O+(x)O(x)O(x)=O^{+}(x)\cup O^{-}(x). An equilibrium is a point xx such that O(x)={x}O(x)=\{x\}. Let EE be the set of all the equilibria of φt\varphi_{t}. The ω\omega-limit set ω(x)\omega(x) of xx is defined by ω(x)=s0tsφt(x)¯\omega(x)=\cap_{s\geq 0}\overline{\cup_{t\geq s}\varphi_{t}(x)}. A point zω(x)z\in\omega(x) if and only if there exists a sequence {ti}\{t_{i}\}, tit_{i}\to\infty, such that φti(x)z\varphi_{t_{i}}(x)\to z as ii\to\infty. If O+(x)O^{+}(x) is precompact, then ω(x)\omega(x) is nonempty, compact, connected, and invariant (i.e., φt(ω(x))=ω(x)\varphi_{t}(\omega(x))=\omega(x) for any tt\in\mathbb{R}). A point xx is called a convergent point if ω(x)\omega(x) consists of a single point of EE. The set of all convergent points is denoted by CC.

Throughout the paper, we always assume that all orbits are forward complete, which means φt(x)\varphi_{t}(x) is well defined for all t0t\geq 0, and the orbit of xx has compact closure for each xMx\in M.

Definition 2.4.

The system Σ\Sigma is said to be differentially positive (DP) with respect to CMC_{M} if

dφt(x)CM(x)CM(φt(x)),xM,t0.d\varphi_{t}(x)C_{M}(x)\subseteq C_{M}(\varphi_{t}(x)),\ \ \forall x\in M,\ \ \forall t\geq 0.

And the differentially positive system Σ\Sigma is said to be strongly differentially positive (SDP) with respect to CMC_{M} if

dφt(x){CM(x)\{0}}IntCM(φt(x)),xM,t>0.d\varphi_{t}(x)\{C_{M}(x)\backslash\{0\}\}\subseteq\text{Int}C_{M}(\varphi_{t}(x)),\ \ \forall x\in M,\ \ \forall t>0.

See Fig.1(b).

In the present paper, we focus on the strongly differentially positive system Σ\Sigma on a Riemannian manifold MM. We first give the following two useful preliminary results.

Proposition 2.2.

If xMyx\leq_{M}y, then φt(x)Mφt(y)\varphi_{t}(x)\leq_{M}\varphi_{t}(y) for t0t\geq 0. Moreover, φt(x)Mφt(y)\varphi_{t}(x)\ll_{M}\varphi_{t}(y) for all t>0t>0.

Proof.

If xMyx\leq_{M}y, there exists a conal curve γ(s)\gamma(s) such that γ(0)=x\gamma(0)=x and γ(1)=y\gamma(1)=y. Since Σ\Sigma is SDP and ddsγ(s)CM(γ(s))\{0}\frac{d}{ds}\gamma(s)\in C_{M}(\gamma(s))\backslash\{0\}, then ddsφt(γ(s))=dφt(γ(s))ddsγ(s)IntCM(φt(γ(s)))\frac{d}{ds}\varphi_{t}(\gamma(s))=d\varphi_{t}(\gamma(s))\frac{d}{ds}\gamma(s)\in\text{Int}C_{M}(\varphi_{t}(\gamma(s))) for t>0t>0. So, φt(γ(s))\varphi_{t}(\gamma(s)) is a strictly conal curve. ∎

Proposition 2.3.

If xMyx\leq_{M}y, then for t0>0t_{0}>0, there exist neighborhoods UU of xx, VV of yy such that φt(U)Mφt(V)\varphi_{t}(U)\ll_{M}\varphi_{t}(V) for any tt0t\geq t_{0}.

Proof.

Since xMyx\leq_{M}y, φt0(x)Mφt0(y)\varphi_{t_{0}}(x)\ll_{M}\varphi_{t_{0}}(y) for t0>0t_{0}>0. One can take neighborhoods U¯\bar{U} of φt0(x)\varphi_{t_{0}}(x) and V¯\bar{V} of φt0(y)\varphi_{t_{0}}(y) such that U¯MV¯\bar{U}\ll_{M}\bar{V} by Proposition 2.1. By the continuity of φt0\varphi_{t_{0}}, there are neighborhoods UU of xx, VV of yy such that φt0(U)U¯\varphi_{t_{0}}(U)\subset\bar{U} and φt0(V)V¯\varphi_{t_{0}}(V)\subset\bar{V}. ∎

Refer to caption
(a) A conal curve γ\gamma on manifold MM.
Refer to caption
(b) Strongly differentially positive system on manifold MM

Fig 1: Conal curves and strongly differentially positive flows on manifold MM.

As mentioned in the introduction, we hereafter impose the following hypotheses:

  1. (H1)

    MM is a globally orderable conal manifold equipped with a continuous solid cone field CMC_{M}.

  2. (H2)

    The conal order “M\leq_{M}” is quasi-closed.

  3. (H3)

    Both the cone field CMC_{M} and the Riemannian metric on MM are Γ\Gamma-invariant.

Before ending this section, we give the following critical lemma, which turns out to be important for the proof of our main results in the forthcoming sections.

Lemma 2.1.

Assume that (H1)-(H2) hold. Then

  1. (a)

    The ω\omega-limit set cannot contain two points xx and yy with xMyx\leq_{M}y;

  2. (b)

    If xMyx\leq_{M}y, then either ω(x)Mω(y)\omega(x)\ll_{M}\omega(y), or ω(x)=ω(y)E\omega(x)=\omega(y)\subset E.

For the sake of the completeness, we will postpone its proof in Section 4.

3 Proof of Theorem A

In this section, we will focus on the generic behavior of the SDP flow φt\varphi_{t} under the hypotheses (H1)-(H3).

Theorem 3.1.

(Generic Convergence) Assume that (H1)-(H3) hold. If the cone field admits a CC^{\infty}-section and φt\varphi_{t} satisfies the compactness condition (P), then IntC{\rm Int}C is dense in MM.

Recall that CC denotes the set of all convergent points. Here, we say that a cone field admits a CC^{\infty}-section if there is a CC^{\infty} vector field XX such that X(x)CM(x)X(x)\in C_{M}(x) for all xMx\in M. We also formulate the compactness condition (P): For each x0Mx_{0}\in M, n0ω(xn)\cup_{n\geq 0}\omega(x_{n}) has compact closure contained in MM, where {xn}x0\{x_{n}\}\to x_{0} with xnMxn+1Mx0x_{n}\leq_{M}x_{n+1}\leq_{M}x_{0} (or x0Mxn+1Mxnx_{0}\leq_{M}x_{n+1}\leq_{M}x_{n}) for n1n\geq 1.

Remark 3.1.

We point out that the homogeneous cone field on a homogeneous space naturally admits CC^{\infty}-sections (see [26] or Proposition A.6 in the Appendix). While the compactness condition (P) is a relatively weak compactness requirement which is frequently satisfied (see e.g., [50, 51]).

The proof of Theorem 3.1 will be divided into several steps. To proceed it, we recall the so-called Perron-Frobenius vector field, which was first introduced by Forni and Sepulchre [9]. A continuous vector field vv on an invariant compact set ΩM\Omega\subset M is called a Perron-Frobenius vector field on Ω\Omega, if for each xΩx\in\Omega,

(i) v(x)IntCM(x)v(x)\in\text{Int}C_{M}(x) with |v(x)|x=1|v(x)|_{x}=1;

(ii) v(φt(x))=dφt(x)v(x)|dφt(x)v(x)|v(\varphi_{t}(x))=\dfrac{d\varphi_{t}(x)v(x)}{\lvert d\varphi_{t}(x)v(x)\rvert} for t0t\geq 0;

(iii) limtdφt(x)(dφt(x)u,v(φt(x)))=0\lim\limits_{t\to\infty}d_{\varphi_{t}(x)}(d\varphi_{t}(x)u,v(\varphi_{t}(x)))=0 for all uCM(x)\{0}u\in C_{M}(x)\backslash\{0\},

where dφt(x)(,)d_{\varphi_{t}(x)}(\cdot,\cdot) is the Hilbert Metric induced by CM(φt(x))C_{M}(\varphi_{t}(x)) (see [9, Section VI]).

Lemma 3.1.

Let ω(x)E\omega(x)\subset E and vv be the Perron-Frobenius vector field on ω(x)\omega(x). If ω(x)\omega(x) is not a singleton, then there exist τ>0\tau>0 and ρ(e)>1\rho(e)>1 such that

dφτ(e)v(e)=ρ(e)v(e), for eω(x).d\varphi_{\tau}(e)v(e)=\rho(e)v(e),\quad\textnormal{ for }e\in\omega(x).

Moreover, there exists ρ>1\rho>1 such that ρ(e)>ρ\rho(e)>\rho for all eω(x)e\in\omega(x).

Proof.

Fix τ>0\tau>0. Since eω(x)Ee\in\omega(x)\subset E, then φt(e)=φτ(e)=e\varphi_{t}(e)=\varphi_{\tau}(e)=e for all t0t\geq 0. So, dφτ(e)v(e)=ρ(e)v(e)d\varphi_{\tau}(e)v(e)=\rho(e)v(e) for some ρ(e)>0\rho(e)>0.

We then pick a smooth chart Φ:UW\Phi:U\to W, where UMU\subset M is the coordinate neighborhood of ee, WW is an open set in n\mathbb{R}^{n}. Using this coordinate map Φ\Phi, we write e~=Φ(e)\widetilde{e}=\Phi(e), v~(e~)=dΦ(e)v(e)\widetilde{v}(\widetilde{e})=d\Phi(e)v(e) and C~(e~)=dΦ(e)CM(e)\widetilde{C}(\widetilde{e})=d\Phi(e)C_{M}(e). Let A=ΦφτΦ1A=\Phi\circ\varphi_{\tau}\circ\Phi^{-1}. Then A(e~)=e~A(\widetilde{e})=\widetilde{e} and dA(e~)=dΦ(e)dφτ(e)dΦ1(e~):Te~nTe~ndA(\widetilde{e})=d\Phi(e)\circ d\varphi_{\tau}(e)\circ d\Phi^{-1}(\widetilde{e}):T_{\widetilde{e}}\mathbb{R}^{n}\to T_{\widetilde{e}}\mathbb{R}^{n}. So, dA(e~)v~(e~)=dΦ(e)dφτ(e)dΦ1(e~)v~(e~)=dΦ(e)dφτ(e)v(e)=dΦ(e)ρ(e)v(e)=ρ(e)dΦ(e)v(e)=ρ(e)v~(e~)dA(\widetilde{e})\widetilde{v}(\widetilde{e})=d\Phi(e)\circ d\varphi_{\tau}(e)\circ d\Phi^{-1}(\widetilde{e})\widetilde{v}(\widetilde{e})=d\Phi(e)\circ d\varphi_{\tau}(e)v(e)=d\Phi(e)\rho(e)v(e)=\rho(e)d\Phi(e)v(e)=\rho(e)\widetilde{v}(\widetilde{e}). Since system is SDP and C~(e~)=dΦ(e)CM(e)\widetilde{C}(\widetilde{e})=d\Phi(e)C_{M}(e), then dA(e~)(C~(e~)\{0})IntC~(e~)dA(\widetilde{e})(\widetilde{C}(\widetilde{e})\backslash\{0\})\subset\text{Int}\widetilde{C}(\widetilde{e}). Moreover, v~(e~)IntC~(e~)\widetilde{v}(\widetilde{e})\in\text{Int}\widetilde{C}(\widetilde{e}), since v(e)IntCM(e)v(e)\in\text{Int}C_{M}(e). By the Perron-Frobenius Theorem, we obtain that ρ(e)=σ(dA(e~))\rho(e)=\sigma(dA(\widetilde{e})), where σ(dA(e~))\sigma(dA(\widetilde{e})) is the spectral radius of dA(e~)dA(\widetilde{e}).

Since ω(x)\omega(x) is not a singleton, we obtain that there exists a sequence enω(x)Ue_{n}\in\omega(x)\cap U such that enee_{n}\neq e and enee_{n}\to e. Let e~n=Φ(en)\widetilde{e}_{n}=\Phi(e_{n}). Then e~e~n=A(e~)A(e~n)=dA(e~)(e~e~n)+o(|e~e~n|)\widetilde{e}-\widetilde{e}_{n}=A(\widetilde{e})-A(\widetilde{e}_{n})=dA(\widetilde{e})(\widetilde{e}-\widetilde{e}_{n})+o(\lvert\widetilde{e}-\widetilde{e}_{n}\rvert), where o(|e~e~n|)/|e~e~n|0o(\lvert\widetilde{e}-\widetilde{e}_{n}\rvert)/\lvert\widetilde{e}-\widetilde{e}_{n}\rvert\to 0 as nn\to\infty. Let wn=e~e~n/|e~e~n|w_{n}=\widetilde{e}-\widetilde{e}_{n}/\lvert\widetilde{e}-\widetilde{e}_{n}\rvert, then wn=dA(e~)wn+rnw_{n}=dA(\widetilde{e})w_{n}+r_{n}, where rn0r_{n}\to 0 as nn\to\infty. If wnww_{n}\to w as nn\to\infty, we obtain that w=dA(e~)ww=dA(\widetilde{e})w. Hence, σ(dA(e~))1\sigma(dA(\widetilde{e}))\geq 1.

If σ(dA(e~))=1\sigma(dA(\widetilde{e}))=1, then wIntC~(e~)w\in\text{Int}\widetilde{C}(\widetilde{e}). Hence, there exists a neighborhood N~\widetilde{N} of ww in the interior of C~(e~)\widetilde{C}(\widetilde{e}) such that there is a convex neighborhood W~\widetilde{W} of e~\widetilde{e} satisfies that for any z~W~\widetilde{z}\in\widetilde{W}, N~C~(z~)\widetilde{N}\subset\widetilde{C}(\widetilde{z}). Thus, there is a N>0N>0 such that e~nW~\widetilde{e}_{n}\in\widetilde{W} and wnN~w_{n}\in\widetilde{N} for all nNn\geq N. Let β~(s)=se~+(1s)e~n\widetilde{\beta}(s)=s\widetilde{e}+(1-s)\widetilde{e}_{n} for s[0,1]s\in[0,1], then ddsβ~(s)=e~e~nIntC~(β~(s))\frac{d}{ds}\widetilde{\beta}(s)=\widetilde{e}-\widetilde{e}_{n}\in\text{Int}\widetilde{C}(\widetilde{\beta}(s)). Then, Φ1(β~(s))\Phi^{-1}(\widetilde{\beta}(s)) is a conal curve in MM connecting ene_{n} and ee. Thus, enMee_{n}\leq_{M}e. Since en,eω(x)e_{n},e\in\omega(x), it is a contradiction to Lemma 2.1(a). Thus, we obtain that ρ(e)>1\rho(e)>1. The conclusion of the lemma follows from the compactness of ω(x)\omega(x) and the continuity of the spectral radius (see e.g., [24, 28]). ∎

Proposition 3.1.

If xMyx\leq_{M}y, then either ω(x)Mω(y)\omega(x)\ll_{M}\omega(y), or ω(x)=ω(y)={e}\omega(x)=\omega(y)=\{e\} for some eEe\in E.

Proof.

We just need to prove the case that ω(x)=ω(y)={e}\omega(x)=\omega(y)=\{e\} for some eEe\in E by Lemma 2.1(b). Suppose that ω(x)=ω(y)=KE\omega(x)=\omega(y)=K\subset E. Let τ>0\tau>0 be the one in Lemma 3.1. Let xn=φnτ(x)x_{n}=\varphi_{n\tau}(x), yn=φnτ(y)y_{n}=\varphi_{n\tau}(y) for n1n\geq 1. Since xMyx\leq_{M}y, there exists a conal curve γ:[0,1]M\gamma:[0,1]\to M such that γ(0)=x\gamma(0)=x, γ(1)=y\gamma(1)=y and ddsγ(s)CM(γ(s))\frac{d}{ds}\gamma(s)\in C_{M}(\gamma(s)). Let γn(s)=φnτ(γ(s))\gamma_{n}(s)=\varphi_{n\tau}(\gamma(s)), then γn(s)\gamma_{n}(s) is a strictly conal curve connecting xnx_{n} and yny_{n} such that xnMynx_{n}\ll_{M}y_{n}. Suppose that KK contains more than a single element.

By passing to a subsequence if necessary, we assume that xnpx_{n}\to p, ynqy_{n}\to q as nn\to\infty, where p,qKp,q\in K. Thus, p=qp=q. Otherwise, pqp\neq q with pMqp\leq_{M}q, which is a contradiction to Lemma 2.1(a). The length of γn\gamma_{n} is L(γn)=01|ddsγn(s)|γn(s)𝑑sL(\gamma_{n})=\int_{0}^{1}\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}ds. We assert that L(γn)0L(\gamma_{n})\to 0 as nn\to\infty.

MM is a globally orderable conal manifold, then every point in MM is a strong point (see [26, Proposition 5.3]). We claim that for any open neighborhood UU of pp, there exists a N>0N>0 such that γnU\gamma_{n}\subset U for all nNn\geq N. In fact, suppose there exist a neighborhood U1U_{1} of pp and a subsequence {γnk}\{\gamma_{n_{k}}\} such that γnk\gamma_{n_{k}} leaves U1U_{1} for all kk. Since pp is a strong point, there exists V1U1V_{1}\subset U_{1} open containing pp such that every conal curve that begins in V1V_{1} and leaves U1U_{1} terminates outsides of V1V_{1} (see [26, Lemma 5.2]). On the other hand, γnk(0)p\gamma_{n_{k}}(0)\to p and γnk(1)p\gamma_{n_{k}}(1)\to p, then there exists a k0>0k_{0}>0 such that for kk0k\geq k_{0}, γnk(0)\gamma_{n_{k}}(0) and γnk(1)\gamma_{n_{k}}(1) are in V1V_{1}, which is a contradiction. Thus, we have the claim. Since MM is locally compact, we can find an open neighborhood UU of pp such that U¯\bar{U} is compact. By the previous claim, there is a NN such that γnU\gamma_{n}\subset U for all nNn\geq N. Let Ω={zM:z is a limit point of a sequence {zn},znγn}\Omega=\{z\in M:z\text{ is a limit point of a sequence }\{z_{n}\},z_{n}\in\gamma_{n}\}. Clearly, pΩp\in\Omega. If there is a qΩq\in\Omega with qpq\neq p, then there exist neighborhoods W1W_{1} of pp and W2W_{2} of qq such that W1W2=W_{1}\cap W_{2}=\emptyset since MM is a Hausdorff space. qΩq\in\Omega implies that there is a subsequence {zni}\{z_{n_{i}}\} such that zniqz_{n_{i}}\to q, where zniγniz_{n_{i}}\in\gamma_{n_{i}}. Then there is a I1>0I_{1}>0 such that for iI1i\geq I_{1}, zniW2z_{n_{i}}\in W_{2} with zniW1z_{n_{i}}\notin W_{1}, which contradicts the previous claim for neighborhood W1W_{1} of pp. Hence, Ω=p\Omega=p. If there exist a α>0\alpha>0 and a subsequence njn_{j} such that L(γnj)α>0L(\gamma_{n_{j}})\geq\alpha>0, then Ω1={zM:z is a limit point of a sequence {znj},znjγnj}=p\Omega_{1}=\{z\in M:z\text{ is a limit point of a sequence }\{z_{n_{j}}\},z_{n_{j}}\in\gamma_{n_{j}}\}=p, which is a contradiction. So, L(γn)0L(\gamma_{n})\to 0 as nn\to\infty.

Since xnx_{n} is attracted to KK, one can choose enKe_{n}\in K such that d(xn,en)0d(x_{n},e_{n})\to 0 as nn\to\infty. On the other hand, L(γn)0L(\gamma_{n})\to 0 as nn\to\infty. Hence, max0s1d(γn(s),en)0\underset{0\leq s\leq 1}{\max}d(\gamma_{n}(s),e_{n})\to 0 as nn\to\infty. Then there exist a N¯>0\bar{N}>0 and a coordinate neighborhood UU of pp such that γn,enU\gamma_{n},e_{n}\in U for all nN¯n\geq\bar{N}. Since there is a smooth coordinate chart Φ:UWn\Phi:U\to W\subset\mathbb{R}^{n}, all notations in the following have coordinate representations. With the diffeomorphism Φ\Phi, we treat the following notations both in manifold MM and n\mathbb{R}^{n}.

For each s[0,1]s\in[0,1], we have

|ddsγn+1(s)|γn+1(s)=|dφτ(γn(s))ddsγn(s)|φτ(γn(s))=|Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))ddsγn(s)|φτ(en)|Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))ddsγn(s)dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)+|dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en),\begin{split}\lvert\frac{d}{ds}\gamma_{n+1}(s)\rvert_{\gamma_{n+1}(s)}&=\lvert d\varphi_{\tau}(\gamma_{n}(s))\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(\gamma_{n}(s))}\\ &=\lvert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ &\geq-\lvert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ &\quad+\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})},\end{split} (3.1)

where Γ\Gamma is defined as above.

Since d(γn(s),en)0d(\gamma_{n}(s),e_{n})\to 0 as nn\to\infty, and Γ\Gamma is continuous with Γ(x,x)=Idx\Gamma(x,x)=\text{Id}_{x} for any xMx\in M, then for any δ1>0\delta_{1}>0, there is a N1N¯N_{1}\geq\bar{N} such that for nN1n\geq N_{1}, Γ(γn(s),en)Γ(en,en)δ1\lVert\Gamma(\gamma_{n}(s),e_{n})-\Gamma(e_{n},e_{n})\rVert\leq\delta_{1}. Since φt\varphi_{t} is smooth, then for any δ2>0\delta_{2}>0, there is a N2N¯N_{2}\geq\bar{N} such that for nN2n\geq N_{2}, Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))Γ(φτ(en),φτ(en))dφτ(en)δ2\lVert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))-\Gamma(\varphi_{\tau}(e_{n}),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(e_{n})\rVert\leq\delta_{2}. Let δ=min{δ1,δ2}\delta=\min\{\delta_{1},\delta_{2}\} and N=max{N1,N2}N=\max\{N_{1},N_{2}\}, then for nNn\geq N, we can obtain that Γ(γn(s),en)Γ(en,en)δ\lVert\Gamma(\gamma_{n}(s),e_{n})-\Gamma(e_{n},e_{n})\rVert\leq\delta and Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))Γ(φτ(en),φτ(en))dφτ(en)δ\lVert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))-\Gamma(\varphi_{\tau}(e_{n}),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(e_{n})\rVert\leq\delta. If there is a L>0L>0 such that dφτ(x)L\lVert d\varphi_{\tau}(x)\rVert\leq L for all xKx\in K, then for nNn\geq N, we have that

|Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))ddsγn(s)dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))dφτ(en)Γ(γn(s),en)|ddsγn(s)|γn(s)(Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))Γ(φτ(en),φτ(en))dφτ(en)+dφτ(en)Γ(γn(s),en)dφτ(en)Γ(en,en))|ddsγn(s)|γn(s)(δ+δdφτ(en))|ddsγn(s)|γn(s)δ(1+L)|ddsγn(s)|γn(s).\begin{split}&\lvert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ \leq&\lVert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))-d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\rVert\cdot\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}\\ \leq&\,(\lVert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))-\Gamma(\varphi_{\tau}(e_{n}),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(e_{n})\rVert\\ &+\lVert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})-d\varphi_{\tau}(e_{n})\Gamma(e_{n},e_{n})\rVert)\cdot\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}\\ \leq&\,(\delta+\delta\lVert d\varphi_{\tau}(e_{n})\rVert)\cdot\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}\\ \leq&\,\delta(1+L)\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}.\end{split} (3.2)

Next, we deal with the term |dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})} in inequation (3.1). Due to Lemma 3.1, we have that dφτ(en)v(en)=ρ(en)v(en)d\varphi_{\tau}(e_{n})v(e_{n})=\rho(e_{n})v(e_{n}), where ρ(en)>ρ>1\rho(e_{n})>\rho>1 and vv is the Perron-Frobenius vector field. And we obtain that |1|ddsγn(s)|ddsγn(s)v(γn(s))|γn(s)0\lvert\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-v(\gamma_{n}(s))\rvert_{\gamma_{n}(s)}\to 0 as nn\to\infty. Since Γ\Gamma preserves the metric, |Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)Γ(γn(s),en)v(γn(s))|en0\lvert\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-\Gamma(\gamma_{n}(s),e_{n})v(\gamma_{n}(s))\rvert_{e_{n}}\to 0. Since Γ(,)\Gamma(\cdot,\cdot) and v()v(\cdot) are continuous, we have |Γ(γn(s),en)v(γn(s))Γ(en,en)v(en)|en0\lvert\Gamma(\gamma_{n}(s),e_{n})v(\gamma_{n}(s))-\Gamma(e_{n},e_{n})v(e_{n})\rvert_{e_{n}}\to 0 as d(γn(s),en)0d(\gamma_{n}(s),e_{n})\to 0. Hence, we obtain that |Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)v(en)|en0\lvert\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-v(e_{n})\rvert_{e_{n}}\to 0 as nn\to\infty. Since dφτ(en)\lVert d\varphi_{\tau}(e_{n})\rVert is bounded for enKe_{n}\in K, let ϵn=|dφτ(en)Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)dφτ(en)v(en)|φτ(en)\epsilon_{n}=\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})v(e_{n})\rvert_{\varphi_{\tau}(e_{n})}, then ϵn0\epsilon_{n}\to 0 as nn\to\infty and we have that

|dφτ(en)Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)|φτ(en)=|dφτ(en)Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)dφτ(en)v(en)+dφτ(en)v(en)|φτ(en)|dφτ(en)v(en)|φτ(en)|dφτ(en)Γ(γn(s),en)1|ddsγn(s)|ddsγn(s)dφτ(en)v(en)|φτ(en)=ρ(en)|v(en)|enϵn>ρϵn,\begin{split}&\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ =&\,\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})v(e_{n})+d\varphi_{\tau}(e_{n})v(e_{n})\rvert_{\varphi_{\tau}(e_{n})}\\ \geq&\,\lvert d\varphi_{\tau}(e_{n})v(e_{n})\rvert_{\varphi_{\tau}(e_{n})}-\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{1}{\lvert\frac{d}{ds}\gamma_{n}(s)\rvert}\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})v(e_{n})\rvert_{\varphi_{\tau}(e_{n})}\\ =&\,\rho(e_{n})\lvert v(e_{n})\rvert_{e_{n}}-\epsilon_{n}\\ >&\,\rho-\epsilon_{n},\end{split}

Thus,

|dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)>(ρϵn)|ddsγn(s)|γn(s).\begin{split}\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}>(\rho-\epsilon_{n})\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}.\end{split} (3.3)

It follows from (3.1), (LABEL:inequ1) and (3.3) that

|ddsγn+1(s)|γn+1(s)|Γ(φτ(γn(s)),φτ(en))dφτ(γn(s))ddsγn(s)dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)+|dφτ(en)Γ(γn(s),en)ddsγn(s)|φτ(en)δ(1+L)|ddsγn(s)|γn(s)+(ρϵn)|ddsγn(s)|γn(s)=(ρϵnδ(1+L))|ddsγn(s)|γn(s).\begin{split}\lvert\frac{d}{ds}\gamma_{n+1}(s)\rvert_{\gamma_{n+1}(s)}&\geq-\lvert\Gamma(\varphi_{\tau}(\gamma_{n}(s)),\varphi_{\tau}(e_{n}))d\varphi_{\tau}(\gamma_{n}(s))\frac{d}{ds}\gamma_{n}(s)-d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ &\quad+\lvert d\varphi_{\tau}(e_{n})\Gamma(\gamma_{n}(s),e_{n})\frac{d}{ds}\gamma_{n}(s)\rvert_{\varphi_{\tau}(e_{n})}\\ &\geq-\delta(1+L)\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}+(\rho-\epsilon_{n})\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}\\ &=(\rho-\epsilon_{n}-\delta(1+L))\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}.\end{split}

Since ϵn0\epsilon_{n}\to 0 as nn\to\infty and ρ>1\rho>1 by the Lemma 3.1, we can choose δ\delta small enough such that there exist a N~>N\tilde{N}>N and l>1l>1 satisfying for all nN~n\geq\tilde{N}, ρϵnδ(1+L)l>1\rho-\epsilon_{n}-\delta(1+L)\geq l>1. Thus, we obtain that |ddsγn+1(s)|γn+1(s)l|ddsγn(s)|γn(s)\lvert\frac{d}{ds}\gamma_{n+1}(s)\rvert_{\gamma_{n+1}(s)}\geq l\lvert\frac{d}{ds}\gamma_{n}(s)\rvert_{\gamma_{n}(s)}. So, L(γn+1)lL(γn)L(\gamma_{n+1})\geq lL(\gamma_{n}), which is a contradiction to L(γn)0L(\gamma_{n})\to 0 as nn\to\infty. Thus, KK is a singleton. ∎

Lemma 3.2.

Suppose that the cone field on MM admits a CC^{\infty}-section VV, then for each xMx\in M, there exist ϵx>0\epsilon_{x}>0 and a conal curve γx:(ϵx,ϵx)M\gamma_{x}:(-\epsilon_{x},\epsilon_{x})\to M such that γx(0)=x\gamma_{x}(0)=x and ddsγx(s)=V(γx(s))\frac{d}{ds}\gamma_{x}(s)=V(\gamma_{x}(s)) for s(ϵx,ϵx)s\in(-\epsilon_{x},\epsilon_{x}).

Proof.

See [27, Proposition 9.2]. In fact, γx\gamma_{x} is the integral curve of VV. ∎

By the Lemma 3.2, we obtain that for any xMx\in M, there exists a conal curve passing through xx. Thus, we say that xx can be approximated from below (resp. above) in MM (i.e., there exists a sequence {xn}\{x_{n}\} in MM such that xnMxn+1Mxx_{n}\leq_{M}x_{n+1}\leq_{M}x (resp. xMxn+1Mxnx\leq_{M}x_{n+1}\leq_{M}x_{n}) for n1n\geq 1 and xnxx_{n}\to x as nn\to\infty).

Lemma 3.3.

If x0Mx_{0}\in M satisfies that there exists a sequence {xn}\{x_{n}\} such that xnMxn+1Mx0x_{n}\leq_{M}x_{n+1}\leq_{M}x_{0} for n1n\geq 1 and xnx0x_{n}\to x_{0}, then one of the following alternatives must occur :

  1. (1)(1)

    There exists pEp\in E such that ω(xn)Mω(xn+1)Mp=ω(x0)\omega(x_{n})\ll_{M}\omega(x_{n+1})\ll_{M}p=\omega(x_{0}) for all n1n\geq 1.

  2. (2)(2)

    ω(xn)=ω(x0)=pE\omega(x_{n})=\omega(x_{0})=p\in E for n1n\geq 1.

  3. (3)(3)

    There exists pEp\in E such that ω(xn)=pMω(x0)\omega(x_{n})=p\ll_{M}\omega(x_{0}) for all n1n\geq 1.

Proof.

Let x0Mx_{0}\in M have the property that it is approximated from below in MM by a sequence {xn}\{x_{n}\} such that xnMxn+1Mx0x_{n}\leq_{M}x_{n+1}\leq_{M}x_{0} for n1n\geq 1 and xnx0x_{n}\to x_{0}. By Proposition 3.1, either there exists a N>0N>0 such that ω(xn)=ω(xm)\omega(x_{n})=\omega(x_{m}) for all n,mNn,m\geq N, or there is a subsequence {xni}\{x_{n_{i}}\} such that ω(xni)Mω(xni+1)\omega(x_{n_{i}})\ll_{M}\omega(x_{n_{i+1}}) for all i1i\geq 1. Thus, we assume that either ω(xn)=ω(xm)\omega(x_{n})=\omega(x_{m}) for all n,mn,m, or ω(xn)Mω(xn+1)\omega(x_{n})\ll_{M}\omega(x_{n+1}), where xnx0x_{n}\to x_{0}.

Suppose that ω(xn)Mω(xn+1)\omega(x_{n})\ll_{M}\omega(x_{n+1}) for n1n\geq 1 and xnx0x_{n}\to x_{0} as nn\to\infty. We first obtain that ω(xn)Mω(x0)\omega(x_{n})\ll_{M}\omega(x_{0}) for all n1n\geq 1. In fact, if there exists n0>0n_{0}>0 such that ω(xn0)=ω(x0)\omega(x_{n_{0}})=\omega(x_{0}), then ω(xn)=ω(x0)\omega(x_{n})=\omega(x_{0}) for all nn0n\geq n_{0}, which is a contradiction. Let Ω={y:y=limnyn,ynω(xn)}\Omega=\{y:y=\lim\limits_{n\to\infty}y_{n},y_{n}\in\omega(x_{n})\}. By the compactness condition (P)(P), the set Ω\Omega belogs to a compact set n0ω(xn)¯\overline{\underset{n\geq 0}{\bigcup}\omega(x_{n})}. Furthermore, Ω\Omega is nonempty. If {ynk}\{y_{n_{k}}\} and {ymk}\{y_{m_{k}}\} are two subsequences of {yn}\{y_{n}\} such that ynkpy_{n_{k}}\to p and ymkqy_{m_{k}}\to q as kk\to\infty. Since ynω(xn)y_{n}\in\omega(x_{n}), then for each kk, there exists l(k)l(k) such that ynkMyml(k)y_{n_{k}}\ll_{M}y_{m_{l(k)}}. Thus, pMqp\leq_{M}q. A similar argument shows that qMpq\leq_{M}p, then p=qp=q. So, limnyn=p\lim\limits_{n\to\infty}y_{n}=p and pΩp\in\Omega. Thus, Ω\Omega is nonempty. Suppose that u,wΩu,w\in\Omega have the property that there exist two sequences {un}\{u_{n}\}, {wn}\{w_{n}\} with un,wnω(xn)u_{n},w_{n}\in\omega(x_{n}) and unuu_{n}\to u, wnww_{n}\to w as nn\to\infty. Since unMwn+1u_{n}\ll_{M}w_{n+1} (resp. wnMun+1w_{n}\ll_{M}u_{n+1}), we obtain that uMwu\leq_{M}w (resp. wMuw\leq_{M}u). Then u=wu=w. Thus, Ω\Omega is a singleton. On the other hand, since each ω(xn)\omega(x_{n}) is invariant, we obtain that Ω\Omega is positively invariant. So, Ω={p}E\Omega=\{p\}\in E. It is easy to see from the arbitrariness of yny_{n} in ω(xn)\omega(x_{n}) that limndist(ω(xn),p)=0\lim\limits_{n\to\infty}\text{dist}(\omega(x_{n}),p)=0. Since ω(xn)Mω(x0)\omega(x_{n})\ll_{M}\omega(x_{0}), then pMω(x0)p\leq_{M}\omega(x_{0}). If pω(x0)p\in\omega(x_{0}), then ω(x0)=p\omega(x_{0})=p by Lemma 2.1(a). Thus, x0x_{0} is a convergent point and (1)(1) holds. If pω(x0)p\notin\omega(x_{0}), we will get a contradiction. Since pEp\in E and ω(x0)\omega(x_{0}) is invariant, we obtain that pMω(x0)p\ll_{M}\omega(x_{0}). For each zω(x0)z\in\omega(x_{0}), there exist tz>0t_{z}>0, a neighborhood UzU_{z} of pp and a neighborhood VzV_{z} of zz such that φt(Uz)Mφt(Vz)\varphi_{t}(U_{z})\ll_{M}\varphi_{t}(V_{z}) for all ttzt\geq t_{z}. Since {Vz}\{V_{z}\} is an open cover of ω(x0)\omega(x_{0}), we obtain that ω(x0)i=1nVzi=V\omega(x_{0})\subset\bigcup_{i=1}^{n}V_{z_{i}}=V, where ziω(x0)z_{i}\in\omega(x_{0}). Meanwhile, U=i=1nUziU=\bigcap_{i=1}^{n}U_{z_{i}} is a neighborhood of pp. Let t0=max1in{tzi}t_{0}=\underset{1\leq i\leq n}{\max}\{t_{z_{i}}\}, then φt(U)Mφt(V)\varphi_{t}(U)\ll_{M}\varphi_{t}(V) for tt0t\geq t_{0}. On the other hand, there exists t1>0t_{1}>0 such that φt1(x0)V\varphi_{t_{1}}(x_{0})\in V. Since xnx0x_{n}\to x_{0}, there is a N1>0N_{1}>0 such that φt1(xN1)V\varphi_{t_{1}}(x_{N_{1}})\in V. By pEp\in E, we have that pMφt(xN1)p\ll_{M}\varphi_{t}(x_{N_{1}}) for t>t0+t1t>t_{0}+t_{1}. Thus, pMω(xN1)p\leq_{M}\omega(x_{N_{1}}). By the definition of Ω\Omega, ω(xn)Mω(xn+1)Mp\omega(x_{n})\ll_{M}\omega(x_{n+1})\leq_{M}p for all n1n\geq 1. Thus, ω(xN1)=p\omega(x_{N_{1}})=p and ω(xn)=p\omega(x_{n})=p for all n>N1n>N_{1}, which is a contradiction. Hence, we have proved the case (1)(1).

Suppose that ω(xn)=ω(xm)\omega(x_{n})=\omega(x_{m}) for all n,mn,m. By Proposition 3.1, ω(xn)=pE\omega(x_{n})=p\in E for all n>1n>1. Since xnMx0x_{n}\leq_{M}x_{0}, then ω(xn)=ω(x0)\omega(x_{n})=\omega(x_{0}) or ω(xn)Mω(x0)\omega(x_{n})\ll_{M}\omega(x_{0}). So, (2)(2) and (3)(3) hold. ∎

Remark 3.2.

An analogous result holds if x0x_{0} is approximated from above.


Now, we are ready to prove Theorem 3.1.

Proof of Theorem 3.1.

Suppose that x0M\IntCx_{0}\in M\backslash\text{Int}C, then we will prove that x0IntC¯x_{0}\in\overline{\text{Int}C}. In fact, there is a sequence {xn}M\C\{x_{n}\}\subset M\backslash C such that xnx0x_{n}\to x_{0}. Since the cone field admits a CC^{\infty}-section, then all xnx_{n} can be approximated from below and above in MM. Without loss of generality, we assume that for each xnx_{n}, there exists a sequence {zmn}\{z_{m}^{n}\} such that zmnMzm+1nMxnz_{m}^{n}\leq_{M}z_{m+1}^{n}\leq_{M}x_{n} for m1m\geq 1 and zmnxnz_{m}^{n}\to x_{n} as mm\to\infty. Since {xn}M\C\{x_{n}\}\subset M\backslash C for all nn, then the case (3)(3) of Lemma 3.3 holds for each sequence {zmn}\{z_{m}^{n}\}.

We claim that xnIntC¯x_{n}\in\overline{\text{Int}C} for each nn. If the claim holds, then x0IntC¯x_{0}\in\overline{\text{Int}C} since xnx0x_{n}\to x_{0} and IntC¯\overline{\text{Int}C} is a closed set. Thus, we obtain the theorem.

Now, it suffices to prove the claim. For each nn, zmnxnz_{m}^{n}\to x_{n} as mm\to\infty and {xn}M\C\{x_{n}\}\subset M\backslash C, then there exists a pEp\in E such that ω(zmn)=pMω(xn)\omega(z_{m}^{n})=p\ll_{M}\omega(x_{n}) for all m1m\geq 1. For yω(xn)y\in\omega(x_{n}), there exist a neighborhood WyW_{y} of yy and ty>0t_{y}>0 such that pMφt(Wy)p\ll_{M}\varphi_{t}(W_{y}) for ttyt\geq t_{y}. Since {Wy}\{W_{y}\} is an open cover of ω(xn)\omega(x_{n}), we obtain that ω(xn)i=1kWyi=W\omega(x_{n})\subset\bigcup_{i=1}^{k}W_{y_{i}}=W, where yiω(xn)y_{i}\in\omega(x_{n}). Let t0=max1ik{tyi}t_{0}=\underset{1\leq i\leq k}{\max}\{t_{y_{i}}\}, then pMφt(W)p\ll_{M}\varphi_{t}(W) for tt0t\geq t_{0}. On the other hand, there is a t1>0t_{1}>0 such that φt1(xn)W\varphi_{t_{1}}(x_{n})\in W. Furthermore, one can choose a neighborhood UU of xnx_{n} such that φt1(U)W\varphi_{t_{1}}(U)\subset W. If xUx\in U with xMxnx\leq_{M}x_{n}, then there exist a neighborhood VV of xx, VUV\subset U, a neighborhood OO of xnx_{n} and t2>0t_{2}>0 such that φt(V)Mφt(O)\varphi_{t}(V)\ll_{M}\varphi_{t}(O) for tt2t\geq t_{2}. We can choose a L>0L>0 such that zLnOz_{L}^{n}\in O, then we have φt(V)Mφt(zLn)\varphi_{t}(V)\ll_{M}\varphi_{t}(z_{L}^{n}) for tt2t\geq t_{2}. Since VUV\subset U, φt1(U)W\varphi_{t_{1}}(U)\subset W and pMφt(W)p\ll_{M}\varphi_{t}(W) for tt0t\geq t_{0}, we obtain that pMφt+t1(V)p\ll_{M}\varphi_{t+t_{1}}(V) for tt0t\geq t_{0}. Thus, pMφt(V)Mφt(zLn)p\ll_{M}\varphi_{t}(V)\ll_{M}\varphi_{t}(z_{L}^{n}) for tt0+t1+t2t\geq t_{0}+t_{1}+t_{2}. Since ω(zmn)=p\omega(z_{m}^{n})=p for all m1m\geq 1, then ω(s)=pE\omega(s)=p\in E for all sVs\in V. Hence, xIntCx\in\text{Int}C. Since the cone field admits a CC^{\infty}-section, then there is a sequence {umn}\{u_{m}^{n}\} in UU such that umnMxnu_{m}^{n}\leq_{M}x_{n} and umnxnu_{m}^{n}\to x_{n} as mm\to\infty. By the previous proof, umIntCu_{m}\in\text{Int}C. Thus, xnIntC¯x_{n}\in\overline{\text{Int}C}. And hence, we have proved the claim. ∎

4 Proof of Lemma 2.1

In this section, we focus on proving the critical Lemma 2.1. We first need the following proposition.

Proposition 4.1.

If xMφT(x)x\leq_{M}\varphi_{T}(x) for some T>0T>0, then φt(x)pE\varphi_{t}(x)\to p\in E as tt\to\infty.

Proof.

If xMφT(x)x\leq_{M}\varphi_{T}(x) and Σ\Sigma is SDP, then for t0>0t_{0}>0, there exist neighborhoods UU of xx, VV of φT(x)\varphi_{T}(x) such that φt0(U)Mφt0(V)\varphi_{t_{0}}(U)\ll_{M}\varphi_{t_{0}}(V) by Proposition 2.3.

By the continuity, there exists 0<ϵ0<t00<\epsilon_{0}<t_{0} such that φt0(x)Mφt0+T+ϵ(x)\varphi_{t_{0}}(x)\ll_{M}\varphi_{t_{0}+T+\epsilon}(x) for ϵ(ϵ0,ϵ0)\epsilon\in(-\epsilon_{0},\epsilon_{0}). We first claim that ω(x)\omega(x) is a τ\tau-periodic orbit for any τ(Tϵ0,T+ϵ0)\tau\in(T-\epsilon_{0},T+\epsilon_{0}). We just prove the case of τ=T\tau=T and a similar argument for τ(Tϵ0,T+ϵ0)\tau\in(T-\epsilon_{0},T+\epsilon_{0}). If xMφT(x)x\leq_{M}\varphi_{T}(x), then φnT(x)Mφ(n+1)T(x)\varphi_{nT}(x)\ll_{M}\varphi_{(n+1)T}(x) for n=1,2,n=1,2,\cdots. Thus, there exists pMp\in M such that φnT(x)p\varphi_{nT}(x)\to p as nn\to\infty. In fact, if there exist two sequences φnkT(x)\varphi_{n_{k}T}(x) and φnlT(x)\varphi_{n_{l}T}(x) satisfying φnkT(x)p\varphi_{n_{k}T}(x)\to p and φnlT(x)q\varphi_{n_{l}T}(x)\to q, then for each kk, there is a l(k)l(k) such that φnkT(x)Mφnl(k)T(x)\varphi_{n_{k}T}(x)\ll_{M}\varphi_{n_{l(k)}T}(x). Thus, pMqp\leq_{M}q by (H2). A similar argument shows that qMpq\leq_{M}p. So, p=qp=q by (H1). Consider the orbit of pp, φt+T(p)=φt+T(limnφnT(x))=limnφ(n+1)T+t(x)=φt(p)\varphi_{t+T}(p)=\varphi_{t+T}(\lim\limits_{n\to\infty}\varphi_{nT}(x))=\lim\limits_{n\to\infty}\varphi_{(n+1)T+t}(x)=\varphi_{t}(p) for all t0t\geq 0. Thus, O(p)O(p) is a TT-periodic orbit. Suppose that there exist qMq\in M and {tj}\{t_{j}\} such that tjt_{j}\to\infty and φtj(x)q\varphi_{t_{j}}(x)\to q as jj\to\infty. For each jj, tj=njT+sjt_{j}=n_{j}T+s_{j}, nj{0,1,2,}n_{j}\in\{0,1,2,\cdots\}, 0sj<T0\leq s_{j}<T, then φtj(x)=φnjT+sj(x)=φsj(φnjT(x))\varphi_{t_{j}}(x)=\varphi_{n_{j}T+s_{j}}(x)=\varphi_{s_{j}}(\varphi_{n_{j}T}(x)). Thus, φtj(x)=φsj(φnjT(x))φs(p)\varphi_{t_{j}}(x)=\varphi_{s_{j}}(\varphi_{n_{j}T}(x))\to\varphi_{s}(p) as jj\to\infty where s[0,T]s\in[0,T] such that sjss_{j}\to s as jj\to\infty. So, ω(x)=O(p)\omega(x)=O(p).

Since ω(x)\omega(x) is a τ\tau-periodic orbit for all τ(Tϵ0,T+ϵ0)\tau\in(T-\epsilon_{0},T+\epsilon_{0}) and ω(x)=O(p)\omega(x)=O(p), then φt+τ(p)=φt(p)\varphi_{t+\tau}(p)=\varphi_{t}(p) for all t0t\geq 0 (i.e., O(p)O(p) is τ\tau-periodic). Let PP be the set of all periods of φt(p)\varphi_{t}(p). It is easy to see that (Tϵ0,T+ϵ0)P(T-\epsilon_{0},T+\epsilon_{0})\subset P and φt+s(p)=φt(φs(p))=φt(φs+T(p))=φt(p)\varphi_{t+s}(p)=\varphi_{t}(\varphi_{s}(p))=\varphi_{t}(\varphi_{s+T}(p))=\varphi_{t}(p) for all s[0,ϵ0)s\in[0,\epsilon_{0}) and t0t\geq 0. Thus, [0,ϵ0)P[0,\epsilon_{0})\subset P. We next prove that T+ϵ0PT+\epsilon_{0}\in P. In fact, let ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0}) and t=ϵ0ϵ(0,ϵ0)t=\epsilon_{0}-\epsilon\in(0,\epsilon_{0}), then φT+ϵ0(p)=φT+ϵ+(ϵ0ϵ)(p)=φT+ϵ(φt(p))=φT+ϵ(p)=p\varphi_{T+\epsilon_{0}}(p)=\varphi_{T+\epsilon+(\epsilon_{0}-\epsilon)}(p)=\varphi_{T+\epsilon}(\varphi_{t}(p))=\varphi_{T+\epsilon}(p)=p. Thus, T+ϵ0PT+\epsilon_{0}\in P. Then [0,ϵ0]P[0,\epsilon_{0}]\subset P.

For each t>0t>0, t=nϵ0+τt=n\epsilon_{0}+\tau, where n{0,1,2,}n\in\{0,1,2,\cdots\}, and 0τ<ϵ00\leq\tau<\epsilon_{0}. φt(p)=φnϵ0+τ(p)=p\varphi_{t}(p)=\varphi_{n\epsilon_{0}+\tau}(p)=p. Thus, pEp\in E and ω(x)=p\omega(x)=p. ∎

Now, we prove Lemma 2.1(a).

Proof of Lemma 2.1(a).

If there exist two points x,yω(z)x,y\in\omega(z) with xMyx\leq_{M}y, then we can find a neighborhood UU of xx, a neighborhood VV of yy for t0>0t_{0}>0 such that φt0(U)Mφt0(V)\varphi_{t_{0}}(U)\ll_{M}\varphi_{t_{0}}(V) by Proposition 2.3. Since x,yω(z)x,y\in\omega(z), there exist 0<t1<t20<t_{1}<t_{2} such that φt1(z)U\varphi_{t_{1}}(z)\in U and φt2(z)V\varphi_{t_{2}}(z)\in V. Then φt0+t1(z)Mφt0+t2(z)=φt2t1(φt0+t1(z))\varphi_{t_{0}+t_{1}}(z)\ll_{M}\varphi_{t_{0}+t_{2}}(z)=\varphi_{t_{2}-t_{1}}(\varphi_{t_{0}+t_{1}}(z)). By Proposition 4.1, φt(z)pE\varphi_{t}(z)\to p\in E. So, ω(z)=p\omega(z)=p, which is a contradiction. ∎

In order to prove Lemma 2.1(b), we further need several propositions.

Proposition 4.2.

If xMyx\leq_{M}y, tkt_{k}\to\infty, φtk(x)p\varphi_{t_{k}}(x)\to p and φtk(y)p\varphi_{t_{k}}(y)\to p as kk\to\infty, then pEp\in E.

Proof.

Since Σ\Sigma is SDP, there exist a neighborhood UU of xx, a neighborhood VV of yy for t0>0t_{0}>0 such that φt0(U)Mφt0(V)\varphi_{t_{0}}(U)\ll_{M}\varphi_{t_{0}}(V) by Proposition 2.3. Let δ>0\delta>0 be small such that {φl(x):0lδ}U\{\varphi_{l}(x):0\leq l\leq\delta\}\subset U and {φl(y):0lδ}V\{\varphi_{l}(y):0\leq l\leq\delta\}\subset V. Then φs(x)=φt0(φst0(x))Mφt0(y)\varphi_{s}(x)=\varphi_{t_{0}}(\varphi_{s-t_{0}}(x))\ll_{M}\varphi_{t_{0}}(y) for any t0st0+δt_{0}\leq s\leq t_{0}+\delta. Thus, φtkt0(φs(x))Mφtkt0(φt0(y))\varphi_{t_{k}-t_{0}}(\varphi_{s}(x))\ll_{M}\varphi_{t_{k}-t_{0}}(\varphi_{t_{0}}(y)) for kk large enough. Let r=st0r=s-t_{0} and kk\to\infty, we obtain φr(p)Mp\varphi_{r}(p)\leq_{M}p for all r[0,δ]r\in[0,\delta]. As a similar argument, we obtain pMφr(p)p\leq_{M}\varphi_{r}(p) for all r[0,δ]r\in[0,\delta]. So, p=φr(p)p=\varphi_{r}(p) for all r[0,δ]r\in[0,\delta]. For any t>0t>0, we write t=nδ+τt=n\delta+\tau, where τ[0,δ)\tau\in[0,\delta) and n{0,1,2,}n\in\{0,1,2,\cdots\}. So, φt(p)=φnδ+τ(p)=p\varphi_{t}(p)=\varphi_{n\delta+\tau}(p)=p. Thus, pEp\in E. ∎

Proposition 4.3.

If xMyx\leq_{M}y, then ω(x)ω(y)E\omega(x)\cap\omega(y)\subset E.

Proof.

Let pω(x)ω(y)p\in\omega(x)\cap\omega(y), there exists a sequence {tk}\{t_{k}\} such that tkt_{k}\to\infty and φtk(x)p\varphi_{t_{k}}(x)\to p as kk\to\infty. By a subsequence, we assume that φtk(y)q\varphi_{t_{k}}(y)\to q as kk\to\infty. Since xMyx\leq_{M}y and Σ\Sigma is SDP, we obtain pMqp\leq_{M}q and p,qω(y)p,q\in\omega(y). The Lemma 2.1(a) means p=qp=q and Proposition 4.2 implies pEp\in E. ∎

Lemma 4.1.

Let xMyx\leq_{M}y, pω(x)p\in\omega(x), qω(y)q\in\omega(y) and pMqp\ll_{M}q. If pp (or qq) is an equilibrium, then ω(x)Mω(y)\omega(x)\ll_{M}\omega(y).

Proof.

Assume that pEp\in E. Since qω(y)q\in\omega(y) and pMqp\ll_{M}q, there exists a sequence {tk}\{t_{k}\}\to\infty such that φtk(y)q\varphi_{t_{k}}(y)\to q and hence, there exists a k(q)>0k(q)>0 such that pMφtk(q)(y)p\ll_{M}\varphi_{t_{k}(q)}(y). Thus, p=φt(p)Mφt(φtk(q)(y))p=\varphi_{t}(p)\ll_{M}\varphi_{t}(\varphi_{t_{k}(q)}(y)) for all t>0t>0. Thus, pMω(y)p\leq_{M}\omega(y). By Lemma 2.1(a), pω(y)p\notin\omega(y). Moreover, pMω(y)p\ll_{M}\omega(y). In fact, since ω(y)\omega(y) is invariant, for any zω(y)z\in\omega(y), there exists T>0T>0 such that φT(z)ω(y)\varphi_{-T}(z)\in\omega(y) and pMφT(z)p\leq_{M}\varphi_{-T}(z). Thus, pMzp\ll_{M}z.

Since pω(x)p\in\omega(x), there exists a sequence {tl}\{t_{l}\}\to\infty such that φtl(x)p\varphi_{t_{l}}(x)\to p. Since pMω(y)p\ll_{M}\omega(y), there exists a l(p)l(p) such that φtl(p)(x)Mω(y)\varphi_{t_{l}(p)}(x)\ll_{M}\omega(y). Thus, φt(φtl(p)(x))Mφt(ω(y))\varphi_{t}(\varphi_{t_{l}(p)}(x))\ll_{M}\varphi_{t}(\omega(y)) for all t>0t>0. Since ω(y)\omega(y) is compact and invariant, we obtain that ω(x)Mω(y)\omega(x)\leq_{M}\omega(y). By Lemma 2.1(a), ω(x)ω(y)=\omega(x)\cap\omega(y)=\emptyset. Since ω(x),ω(y)\omega(x),\omega(y) are compact and invariant, then ω(x)Mω(y)\omega(x)\ll_{M}\omega(y). In fact, let aω(x)a\in\omega(x) and bω(y)b\in\omega(y), there is a T1>0T_{1}>0 such that φT1(a)ω(x)\varphi_{-T_{1}}(a)\in\omega(x) and φT1(b)ω(y)\varphi_{-T_{1}}(b)\in\omega(y) with φT1(a)MφT1(b)\varphi_{-T_{1}}(a)\leq_{M}\varphi_{-T_{1}}(b), then aMba\ll_{M}b.

A similar argument is used if qEq\in E. ∎

Lemma 4.2.

Assume that KMK\subset M is a compact set in which the flow φt\varphi_{t} is SDP. Then there exists δ>0\delta>0 with the following property. Let ψt\psi_{t} denote the flow of a C1C^{1} vector field gg such that gUδ(f)g\in U_{\delta}(f), where Uδ(f)U_{\delta}(f) is a δ\delta-neighborhood of ff in space of C1C^{1} vector fields on MM with the C1C^{1} topology. Then there exists t0>0t_{0}>0 such that if KK is positively invariant under the flow of ψt\psi_{t}, then ψt\psi_{t} is SDP for all tt0t\geq t_{0}.

Proof.

We first assume that t[t0,2t0]t\in[t_{0},2t_{0}], where t0>0t_{0}>0 is fixed. Since Σ\Sigma is SDP, we obtain that dφt(x)vIntCM(φt(x))d\varphi_{t}(x)v\in\text{Int}C_{M}(\varphi_{t}(x)) for all vCM(x)v\in C_{M}(x) with |v|x=1|v|_{x}=1. With a coordinate chart, we treat the following notations both in manifold and n\mathbb{R}^{n}. By the continuity of the cone field, there exist neighborhoods UU of φt(x)\varphi_{t}(x) and VV of dφt(x)vd\varphi_{t}(x)v such that VCM(y)V\subset C_{M}(y) for all yUy\in U. So, there exists δ>0\delta>0 such that for gUδ(f)g\in U_{\delta}(f), ψt(x)U\psi_{t}(x)\in U and dψt(x)vVd\psi_{t}(x)v\in V (see e.g., [40, 41]). Thus, dψt(x)vIntCM(ψt(x))d\psi_{t}(x)v\in\text{Int}C_{M}(\psi_{t}(x)) for t[t0,2t0]t\in[t_{0},2t_{0}], vCM(x)\{0}v\in C_{M}(x)\backslash\{0\}.

For t2t0t\geq 2t_{0}, let us write t=r+kt0t=r+kt_{0}, where r[t0,2t0)r\in[t_{0},2t_{0}) and k{1,2,}k\in\{1,2,\cdots\}. Define xj=ψjt0(x)x_{j}=\psi_{jt_{0}}(x), j=1,2,,kj=1,2,\cdots,k. It is clear that xjKx_{j}\in K if KK is positively invariant. Then dψt(x)v=dψr(xk)dψt0(xk1)dψt0(x)vd\psi_{t}(x)v=d\psi_{r}(x_{k})d\psi_{t_{0}}(x_{k-1})\cdots d\psi_{t_{0}}(x)v. By the preceding proof, dψt(x)vIntCM(ψt(x))d\psi_{t}(x)v\in\text{Int}C_{M}(\psi_{t}(x)) for t>2t0t>2t_{0}.

Thus, we have proved that ψt\psi_{t} is SDP for all tt0t\geq t_{0}. ∎

Lemma 4.3.

Let xMyx\leq_{M}y, pω(x)p\in\omega(x), qω(y)q\in\omega(y) and pMqp\ll_{M}q. If pp (or qq) belongs to a periodic orbit, then ω(x)Mω(y)\omega(x)\ll_{M}\omega(y).

Proof.

Assume that pγp\in\gamma, γ\gamma is a periodic orbit and qq is not an equilibrium (the other case is similar).

If γω(y)\gamma\cap\omega(y)\neq\emptyset, then pω(y)p\subset\omega(y). Thus, pEp\in E by Proposition 4.3, which is a contradiction. Thus, ω(y)M\γ\omega(y)\subset M\backslash\gamma. Since qω(y)q\in\omega(y), then the orbit closure of qq is in the compact set ω(y)M\γ\omega(y)\subset M\backslash\gamma.

By the Closing Lemma (see e.g., [20, 47]), there is a C1C^{1} vector field gg whose flow ψt\psi_{t} has a closed orbit βg\beta_{g} passing through qq. Moreover, gg can be chosen to C1C^{1} approximate ff as closely as desired and to coincides with ff outside a given neighborhood UU of the orbit closure of qq with respect to ff such that Uγ=U\cap\gamma=\emptyset. Thus, ψt\psi_{t} is eventually SDP by Lemma 4.2 and γ\gamma is also a closed orbit of ψt\psi_{t}. In the following, we write γf\gamma_{f} (resp. γg\gamma_{g}) as the orbit generated by ff (resp. gg). So, γ=γf=γg\gamma=\gamma_{f}=\gamma_{g}.

For the system generated by vector field gg, γg\gamma_{g} and βg\beta_{g} are two periodic orbits and pγgp\in\gamma_{g}, qβgq\in\beta_{g} with pMqp\ll_{M}q. Then γgMq\gamma_{g}\ll_{M}q.

In fact, if pMqp\ll_{M}q, then there is a ϵ0>0\epsilon_{0}>0 such that ψϵ(p)Mq\psi_{\epsilon}(p)\ll_{M}q for all ϵ[0,ϵ0]\epsilon\in[0,\epsilon_{0}]. Let T1,T2>0T_{1},T_{2}>0 be the periods of γg\gamma_{g} and βg\beta_{g} and a=T2T1a=\frac{T_{2}}{T_{1}}. If aa is a rational number, let a=n1n2a=\frac{n_{1}}{n_{2}}, n1,n2{1,2,}n_{1},n_{2}\in\{1,2,\cdots\}, then n1T1=n2T2n_{1}T_{1}=n_{2}T_{2}. Define ϵ1=sup{ϵ>0:ψτ(p)Mq for all τ[0,ϵ]}\epsilon_{1}=\sup\{\epsilon>0:\psi_{\tau}(p)\ll_{M}q\text{ for all }\tau\in[0,\epsilon]\}. Suppose that ϵ1<\epsilon_{1}<\infty. Then ψϵ1(p)Mq\psi_{\epsilon_{1}}(p)\leq_{M}q. Since ψϵ1(p)=ψn1T1(ψϵ1(p))=ψn2T2(ψϵ1(p))Mψn2T2(q)=q\psi_{\epsilon_{1}}(p)=\psi_{n_{1}T_{1}}(\psi_{\epsilon_{1}}(p))=\psi_{n_{2}T_{2}}(\psi_{\epsilon_{1}}(p))\ll_{M}\psi_{n_{2}T_{2}}(q)=q, one has ψϵ1(p)Mq\psi_{\epsilon_{1}}(p)\ll_{M}q. Thus, there exists δ>0\delta>0 such that ψϵ1+ϵ(p)Mq\psi_{\epsilon_{1}+\epsilon}(p)\ll_{M}q for all ϵ[0,δ]\epsilon\in[0,\delta], which is a contradiction. Thus, ψϵ(p)Mq\psi_{\epsilon}(p)\ll_{M}q for all ϵ0\epsilon\geq 0. So, γgMq\gamma_{g}\ll_{M}q. If aa is an irrational number and T2=aT1T_{2}=aT_{1}, the set W={ψnT2(ψϵ(p)):n=1,2,}W=\{\psi_{nT_{2}}(\psi_{\epsilon}(p)):n=1,2,\cdots\} is dense in γg\gamma_{g} for any fixed ϵ[0,ϵ0]\epsilon\in[0,\epsilon_{0}]. For any zγgz\in\gamma_{g}, there exists a sequence {zi}W\{z_{i}\}\subset W such that ziMqz_{i}\ll_{M}q and zizz_{i}\to z as ii\to\infty. Thus, γgMq\gamma_{g}\leq_{M}q. So, γg=ψT2(γg)Mq\gamma_{g}=\psi_{T_{2}}(\gamma_{g})\ll_{M}q.

For system Σ\Sigma, we obtain that γMq\gamma\ll_{M}q. Since qω(y)q\in\omega(y), then there is a sequence {tk}\{t_{k}\}\to\infty such that φtk(y)q\varphi_{t_{k}}(y)\to q as kk\to\infty. Then there exists k(q)>0k(q)>0 such that γMφtk(q)(y)\gamma\ll_{M}\varphi_{t_{k(q)}}(y). Thus, we obtain that γMφt(φtk(q)(y))\gamma\ll_{M}\varphi_{t}(\varphi_{t_{k}(q)}(y)) for all t>0t>0. So, γMω(y)\gamma\leq_{M}\omega(y). On the other hand, γω(y)=\gamma\cap\omega(y)=\emptyset and γ,ω(y)\gamma,\omega(y) are invariant, then γMω(y)\gamma\ll_{M}\omega(y). Since γω(x)\gamma\subset\omega(x), in a similar way we can obtain that ω(x)Mω(y)\omega(x)\ll_{M}\omega(y). ∎

Proposition 4.4.

Let xMyx\leq_{M}y, pω(x)p\in\omega(x), qω(y)q\in\omega(y) and pMqp\ll_{M}q. Then ω(x)Mω(y)\omega(x)\ll_{M}\omega(y).

Proof.

If pEp\in E, then ω(x)Mω(y)\omega(x)\ll_{M}\omega(y) by Lemma 4.1. In the following, we assume that pp is not an equilibrium. Thus, pω(y)p\notin\omega(y).

By the Closing Lemma (see e.g., [20, 47]), there exists T>0T>0 such that for any ϵ>0\epsilon>0, we can choose a vector field gg such that g=fg=f outside the ϵ\epsilon-neighborhood NϵN_{\epsilon} of the set {φt(p):|t|T}\{\varphi_{t}(p):|t|\leq T\}. Moreover, ψt\psi_{t} with respect to the vector field gg has a closed orbit γg\gamma_{g} passing through pp and ψt\psi_{t} is eventually SDP. For small ϵ\epsilon, ω(y)M\Nϵ\omega(y)\subset M\backslash N_{\epsilon}. It is easy to see that there exists a y0y_{0} such that pMy0p\leq_{M}y_{0} and ωg(y0)=ω(y)\omega_{g}(y_{0})=\omega(y), where ωg(y0)={zM:there exists a sequence tk such that ψtk(y0)z}\omega_{g}(y_{0})=\{z\in M:\text{there exists a sequence }t_{k}\to\infty\text{ such that }\psi_{t_{k}}(y_{0})\to z\}. In fact, since pMqp\ll_{M}q, there is a neighborhood UqU_{q} of qq such that UqNϵ=U_{q}\cap N_{\epsilon}=\emptyset and pMUqp\ll_{M}U_{q}. Then there is a tq>0t_{q}>0 such that φtq(y)Uq\varphi_{t_{q}}(y)\in U_{q}. Let y0=φtq(y)y_{0}=\varphi_{t_{q}}(y), then pMy0p\ll_{M}y_{0} and ωg(y0)=ω(y)\omega_{g}(y_{0})=\omega(y) since g=fg=f outside NϵN_{\epsilon} with ω(y)M\Nϵ\omega(y)\subset M\backslash N_{\epsilon}. Since pγgp\in\gamma_{g}, qωg(y0)q\in\omega_{g}(y_{0}) and pMqp\ll_{M}q, then γg=ωg(p)Mωg(y0)\gamma_{g}=\omega_{g}(p)\ll_{M}\omega_{g}(y_{0}) by Lemma 4.3. Thus, pMω(y)p\ll_{M}\omega(y). Since pω(x)p\in\omega(x), there exists a sequence {tk}\{t_{k}\}\to\infty such that φtk(x)p\varphi_{t_{k}}(x)\to p. Then there exists k(p)>0k(p)>0 such that φtk(p)(x)Mω(y)\varphi_{t_{k(p)}}(x)\ll_{M}\omega(y). Thus, we obtain that φt(φtk(p)(x))Mφt(ω(y))\varphi_{t}(\varphi_{t_{k(p)}}(x))\ll_{M}\varphi_{t}(\omega(y)) for all t>0t>0. So, ω(x)Mω(y)\omega(x)\leq_{M}\omega(y). Since ω(x)\omega(x) and ω(y)\omega(y) are nonordering invariant sets and Σ\Sigma is SDP, then ω(x)ω(y)=\omega(x)\cap\omega(y)=\emptyset and ω(x)Mω(y)\omega(x)\ll_{M}\omega(y). ∎

Now, we are ready to prove Lemma 2.1(b).

Proof of Lemma 2.1(b).

If pω(x)p\in\omega(x), then there exists a sequence tkt_{k}\to\infty such that φtk(x)p\varphi_{t_{k}}(x)\to p. By passing to a subsequence if necessary, we assume that φtk(y)qω(y)\varphi_{t_{k}}(y)\to q\in\omega(y). Since Σ\Sigma is SDP and xMyx\leq_{M}y, we obtain that pMqp\leq_{M}q. If pqp\neq q, then φt(p)Mφt(q)\varphi_{t}(p)\ll_{M}\varphi_{t}(q) for any t>0t>0. Thus, ω(x)ω(y)=\omega(x)\cap\omega(y)=\emptyset and ω(x)Mω(y)\omega(x)\ll_{M}\omega(y) by Proposition 4.4.

If p=qp=q, then p=qEp=q\in E by Proposition 4.2. If ω(x)=ω(y)\omega(x)=\omega(y), then ω(x)=ω(y)E\omega(x)=\omega(y)\subset E by Proposition 4.3. If ω(x)ω(y)\omega(x)\neq\omega(y), then there exists p1ω(x)\ω(y)p_{1}\in\omega(x)\backslash\omega(y) such that there is a sequence tlt_{l}\to\infty with φtl(x)p1\varphi_{t_{l}}(x)\to p_{1}. Therefore, there is q1ω(y)q_{1}\in\omega(y) such that φtl(y)q1\varphi_{t_{l}}(y)\to q_{1} by a subsequence if necessary. Thus, p1q1p_{1}\neq q_{1} and p1Mq1p_{1}\leq_{M}q_{1}. So, φt(p1)Mφt(q1)\varphi_{t}(p_{1})\ll_{M}\varphi_{t}(q_{1}) for any t>0t>0 and hence, we obtain that ω(x)ω(y)=\omega(x)\cap\omega(y)=\emptyset and ω(x)Mω(y)\omega(x)\ll_{M}\omega(y) by Proposition 4.4, which is a contradiction. ∎

Appendix A Appendix

A.1 Order structures on space-times

A Lorentz metric gg for a smooth manifold MM of dimension four is a smooth nondegenerate symmetric tensor field of type (0,2)(0,2) on MM such that for each pMp\in M, by suitable choice of the basis, g|pg|_{p} has the matrix diag(+1,+1,+1,1)(+1,+1,+1,-1). A space-time (M,g)(M,g) is a connected CC^{\infty} Hausdorff manifold MM of dimension four with a Lorentz metric gg.

Remark A.1.

Although the arguments refer to 44-dimensional space-times, the results can be extended to a space-time of n(2)n(\geq 2)-dimensions. See [2, 12, 43].

Let (M,g)(M,g) be a space-time, a vector vTpMv\in T_{p}M is said to be timelike, null, spacelike according to whether g(v,v)g(v,v) is negative, zero, or positive, respectively. The non-spacelike (i.e., timelike and null) vectors in TpMT_{p}M form two so-called Lorentzian cones CC and C-C (see e.g., [12]). Furthermore, the timelike vectors form the interior of the Lorentzian cones. See Fig.A.1.

[Uncaptioned image]

Fig A.1: The timelike vectors, null vectors and spacelike vectors in TpMT_{p}M.

A space-time (M,g)(M,g) is said to be time-orientable if MM admits a continuous Lorentzian cone field CMC_{M}, which is generated by Lorentz metric gg.

In the remained of this subsection, we assume that space-time (M,g)(M,g) is time-orientable. Thus, MM is a conal manifold with respect to a continuous Lorentzian cone field CMC_{M}. In such situation, the non-spacelike vectors, which belong to cone field CMC_{M}, are called future directed.

A non-spacelike curve is a continuous piecewise smooth curve whose tangent vector is future directed non-spacelike. A timelike curve is a continuous piecewise smooth curve whose tangent vector is future directed timelike. Thus, in a time-orientable space-time, a non-spacelike (resp., timelike) curve is a conal (resp., strictly conal) curve and vice versa. The order “M\leq_{M}” (resp., “M\ll_{M}”) on MM is well-defined by the non-spacelike (resp., timelike) curves.

A non-spacelike curve is also known as causal curve. In general relativity, each point of manifold MM corresponds to an event. And a signal can be sent from pp to qq if there is a (future directed) causal curve from pp to qq. Thus, closed causal curves generate paradoxes involving causality (i.e., violate causality). As a result, we assume that the space-time (M,g)(M,g) is causal, i.e., (M,g)(M,g) contain no closed non-spacelike (conal) curves (see e.g., [2, 12, 13, 43]).

For a given point pMp\in M, the chronological future I+(p)I^{+}(p), chronological past I(p)I^{-}(p), causal future J+(p)J^{+}(p), and causal past J(p)J^{-}(p) of pp are defined as follows:

  • I+(p)={qM:pMq}I^{+}(p)=\{q\in M:p\ll_{M}q\};    I(p)={qM:qMp}I^{-}(p)=\{q\in M:q\ll_{M}p\};

  • J+(p)={qM:pMq}J^{+}(p)=\{q\in M:p\leq_{M}q\};    J(p)={qM:qMp}J^{-}(p)=\{q\in M:q\leq_{M}p\}.

Remark A.2.

In some articles, the set J+(p)J^{+}(p) would be called forward set or reachable set from pp and be written as p\uparrow p; the set J(p)J^{-}(p) would be called backward set or controllable set from pp and be written as p\downarrow p.

Remark A.3.

For any pMp\in M, I±(p)I^{\pm}(p) are open by Proposition 2.1.

I+I^{+} is said to be inner continuous at pMp\in M if each compact set KI+(p)K\subset I^{+}(p), there exists a neighborhood U(p)U(p) of pp such that KI+(q)K\subset I^{+}(q) for each qU(p)q\in U(p). I+I^{+} is said to be outer continuous at pMp\in M if each compact set KM\I+(p)¯K\subset M\backslash\overline{I^{+}(p)}, there exists a neighborhood U(p)U(p) of pp such that KM\I+(q)¯K\subset M\backslash\overline{I^{+}(q)} for each qU(p)q\in U(p). The inner and outer continuity of II^{-} are defined dually. See [13].

Proposition A.1.

For any pMp\in M, I±(p)I^{\pm}(p) are inner continuous.

Proof.

Suppose KI(p)K\subset I^{-}(p) is compact. For any xI(p)x\in I^{-}(p), i.e., xMpx\ll_{M}p, there is a strictly conal curve γ\gamma such that γ(0)=x\gamma(0)=x and γ(1)=p\gamma(1)=p. Let w=γ(12)w=\gamma(\frac{1}{2}), then xMwMpx\ll_{M}w\ll_{M}p, i.e., wI(p)w\in I^{-}(p) and xI(w)x\in I^{-}(w). Since I(w)I^{-}(w) is open, then I(w)I^{-}(w) is an open neighborhood of xx. Thus, {I(w):wI(p)}\{I^{-}(w):w\in I^{-}(p)\} is an open covering of KK. Since KK is a compact set, we choose w1,w2,,wnw_{1},w_{2},\cdots,w_{n} to determine a finite subcovering. On the other hand, wiI(p)w_{i}\in I^{-}(p) implies that pI+(wi)p\in I^{+}(w_{i}), i=1,2,,ni=1,2,\cdots,n. So, U=i=1nI+(wi)U=\bigcap^{n}_{i=1}I^{+}(w_{i}) is an open neighborhood of pp. For any uUu\in U, uI+(wi)u\in I^{+}(w_{i}), i=1,2,,ni=1,2,\cdots,n. Then wiI(u)w_{i}\in I^{-}(u). For any yI(wi)y\in I^{-}(w_{i}), since yMwiy\ll_{M}w_{i} and wiMuw_{i}\ll_{M}u, then yMuy\ll_{M}u, i.e., yI(u)y\in I^{-}(u). So, I(wi)I(u)I^{-}(w_{i})\subset I^{-}(u) for i=1,2,,ni=1,2,\cdots,n. Thus, Ki=1nI(wi)I(u)K\subset\bigcup^{n}_{i=1}I^{-}(w_{i})\subset I^{-}(u). So, we have proved that I(p)I^{-}(p) is inner continuous.

A similar argument is used for I+(p)I^{+}(p). ∎

Proposition A.2.
  1. (1)(1)

    xMyx\ll_{M}y, yMzy\leq_{M}z implies xMzx\ll_{M}z,

  2. (2)(2)

    xMyx\leq_{M}y, yMzy\ll_{M}z implies xMzx\ll_{M}z.

Proof.

See [12, p183] and [43, Proposition 2.18]. ∎

Remark A.4.

In [3, 10], Proposition A.2 is called push-up lemma. The push-up lemma fails in the case where the spacetime metric (i.e., Lorentz metric) is continuous, see [3, Example 1.11] or [10].

Proposition A.3.

I±(p)J±(p)I±(p)¯={qM:I±(q)I±(p)}I^{\pm}(p)\subset J^{\pm}(p)\subset\overline{I^{\pm}(p)}=\{q\in M:I^{\pm}(q)\subset I^{\pm}(p)\}.

Proof.

It is clear that I+(p)J+(p)I^{+}(p)\subset J^{+}(p). See [43, Proposition 3.3] for I+(p)¯={qM:I+(q)I+(p)}\overline{I^{+}(p)}=\{q\in M:I^{+}(q)\subset I^{+}(p)\} and [43, Proposition 3.9] for J+(p)I+(p)¯J^{+}(p)\subset\overline{I^{+}(p)}. ∎

Theorem A.1.

For any pMp\in M, if J±(p)J^{\pm}(p) are closed and I±(p)I^{\pm}(p) are outer continuous, then the conal order “M\leq_{M}” is quasi-closed.

Proof.

If xnMynx_{n}\ll_{M}y_{n} for all nn and xnxx_{n}\to x and ynyy_{n}\to y as nn\to\infty, we just need to prove that yJ+(x)=I+(x)¯y\in J^{+}(x)=\overline{I^{+}(x)} by Proposition A.3. Suppose that yI+(x)¯y\notin\overline{I^{+}(x)}, i.e., yM\I+(x)¯y\in M\backslash\overline{I^{+}(x)}. Then there exists a compact set KK containing yy such that KM\I+(x)¯K\subset M\backslash\overline{I^{+}(x)}. Since ynyy_{n}\to y as nn\to\infty, there is a N1>0N_{1}>0 such that ynKy_{n}\in K for all n>N1n>N_{1}. On the other hand, I+I^{+} is outer continuous at xx, then there exists a neighborhood UU of xx such that for any zUz\in U, KM\I+(z)¯K\subset M\backslash\overline{I^{+}(z)}. Since xnxx_{n}\to x as nn\to\infty, there is a N2>0N_{2}>0 such that xnUx_{n}\in U and KM\I+(xn)¯K\subset M\backslash\overline{I^{+}(x_{n})} for all n>N2n>N_{2}. Let N=max{N1,N2}N=\max\{N_{1},N_{2}\}. If n>Nn>N, then xnUx_{n}\in U and ynKy_{n}\in K with ynI+(xn)y_{n}\in I^{+}(x_{n}), which is a contradiction. Thus, yJ+(x)=I+(x)¯y\in J^{+}(x)=\overline{I^{+}(x)}. ∎

Proposition A.4.

The following conditions are equivalent.

  1. (A)

    For all pp and qq in MM, I+(p)I+(q)I^{+}(p)\subset I^{+}(q) if and only if I(q)I(p)I^{-}(q)\subset I^{-}(p);

  2. (B)

    For all pp and qq in MM, pJ+(q)¯p\in\overline{J^{+}(q)} if and only if qJ(p)¯q\in\overline{J^{-}(p)}.

Proof.

The conclusion can be immediately obtained with the Proposition A.3. See also [13, Proposition 1.3]. ∎

Proposition A.5.

If any one of the equivalent conditions in Proposition A.4 holds, then I±(p)I^{\pm}(p) are outer continuous for any pMp\in M.

Proof.

We first assert that for vM\I(p)¯v\in M\backslash\overline{I^{-}(p)}, there exists wI+(p)w\in I^{+}(p) such that vM\I(w)¯v\in M\backslash\overline{I^{-}(w)}. Before proving this assertion, we will show how it implies this Proposition. Let KM\I(p)¯K\subset M\backslash\overline{I^{-}(p)} be compact. This assertion implies that {M\I(w)¯:wI+(p)}\{M\backslash\overline{I^{-}(w)}:w\in I^{+}(p)\} is an open covering of KM\I(p)¯K\subset M\backslash\overline{I^{-}(p)}. Choose a finite subcovering {M\I(wi)¯:wiI+(p),i=1,2,,n}\{M\backslash\overline{I^{-}(w_{i})}:w_{i}\in I^{+}(p),i=1,2,\cdots,n\} determined by w1,w2,,wnw_{1},w_{2},\cdots,w_{n} and U=i=1nI(wi)U=\bigcap^{n}_{i=1}I^{-}(w_{i}). Then UU is a neighborhood of pp such that for any uUu\in U, KM\I(u)¯K\subset M\backslash\overline{I^{-}(u)}. In fact, since wiI+(p)w_{i}\in I^{+}(p), then pI(wi)p\in I^{-}(w_{i}), i=1,2,,ni=1,2,\cdots,n. So, pi=1nI(wi)=Up\in\bigcap^{n}_{i=1}I^{-}(w_{i})=U. For any uUu\in U, uI(wi)u\in I^{-}(w_{i}) for i=1,2,,ni=1,2,\cdots,n. Then I(u)I(wi)I^{-}(u)\subset I^{-}(w_{i}) for i=1,2,,ni=1,2,\cdots,n. Thus, M\I(wi)¯M\I(u)¯M\backslash\overline{I^{-}(w_{i})}\subset M\backslash\overline{I^{-}(u)} for i=1,2,,ni=1,2,\cdots,n. So, KM\I(u)¯K\subset M\backslash\overline{I^{-}(u)}. Then I(p)I^{-}(p) is outer continuous.

It remains to prove the assertion. Suppose that vI(w)¯v\in\overline{I^{-}(w)} for any wI+(p)w\in I^{+}(p), then we will get a contradiction. Since vI(w)¯v\in\overline{I^{-}(w)}, then I(v)I(w)I^{-}(v)\subset I^{-}(w) by Proposition A.3. Since condition (A)(A) of Proposition A.4 holds, then I+(w)I+(v)I^{+}(w)\subset I^{+}(v) for all wI+(p)w\in I^{+}(p). On the other hand, for any yI+(p)y\in I^{+}(p), i.e., pMyp\ll_{M}y, there is a strictly conal curve γ:[0,1]M\gamma:[0,1]\mapsto M such that γ(0)=p\gamma(0)=p and γ(1)=y\gamma(1)=y. Let x=γ(12)x=\gamma(\frac{1}{2}), then pMxMyp\ll_{M}x\ll_{M}y. Thus, yI+(x)y\in I^{+}(x) and xI+(p)x\in I^{+}(p). Since I+(w)I+(v)I^{+}(w)\subset I^{+}(v) for all wI+(p)w\in I^{+}(p), then I+(x)I+(v)I^{+}(x)\subset I^{+}(v). So, yI+(v)y\in I^{+}(v) for any yI+(p)y\in I^{+}(p). Then I+(p)I+(v)I^{+}(p)\subset I^{+}(v). By Proposition A.3, pI+(v)¯p\in\overline{I^{+}(v)}. Thus, vI(p)¯v\in\overline{I^{-}(p)} by Proposition A.4, which is a contradiction with vM\I(p)¯v\in M\backslash\overline{I^{-}(p)}. ∎

Remark A.5.

If for all pp and qq in MM, either I+(p)=I+(q)I^{+}(p)=I^{+}(q) or I(p)=I(q)I^{-}(p)=I^{-}(q) implies p=qp=q, then that I±I^{\pm} are outer continuous is equivalent to any one of the conditions in Proposition A.4. See [13, Theorem 2.1].

Lemma A.1.

If J±(p)J^{\pm}(p) are closed for all pMp\in M, then I±(p)I^{\pm}(p) are outer continuous.

Proof.

If J±(p)J^{\pm}(p) are closed for all pMp\in M, then condition (B) of Proposition A.4 holds. So, I±(p)I^{\pm}(p) are outer continuous by Proposition A.5. ∎

Corollary A.1.

If J±(p)J^{\pm}(p) are closed for all pMp\in M, then the conal order “M\leq_{M}” is quasi-closed.

Proof.

If J±(p)J^{\pm}(p) are closed for all pMp\in M, then I±(p)I^{\pm}(p) are outer continuous by Lemma A.1. Thus, the conal order “M\leq_{M}” is quasi-closed by Theorem A.1. ∎

A.2 Globally orderable homogeneous spaces

Let GG be a connected Lie group and MM be a smooth manifold. A left action of Lie group GG on manifold MM is a smooth map θ:G×MM\theta:G\times M\to M satisfying θ(e,x)=x\theta(e,x)=x and θ(g1g2,x)=θ(g1,θ(g2,x))\theta(g_{1}g_{2},x)=\theta(g_{1},\theta(g_{2},x)) for all g1,g2Gg_{1},g_{2}\in G, xMx\in M, where ee is the identity element in GG. We write gxg\cdot x or gxgx for θ(g,x)\theta(g,x). The action is said to be transitive if for every pair of points x,yMx,y\in M, there exists gGg\in G such that gx=yg\cdot x=y. For each xMx\in M, the isotropy group of xx, denoted by GxG_{x}, is the set Gx={gG:gx=x}G_{x}=\{g\in G:g\cdot x=x\}.

A smooth manifold endowed with a transitive smooth action by a Lie group GG is called a homogeneous G-space (or a homogeneous space).

Let GG be a Lie group and HGH\subset G be a closed subgroup. The left coset space G/H={gH:gG}G/H=\{gH:g\in G\} is a smooth manifold of dimension (dimGdimH\text{dim}G-\text{dim}H), and the left action of GG on G/HG/H is given by g1(g2H)=(g1g2)Hg_{1}\cdot(g_{2}H)=(g_{1}g_{2})H. Hence, G/HG/H is a homogeneous space (see [27, Theorem 21.17]).

Let GG be a Lie group and MM be a homogeneous G-space. If pp is any point of MM, then the isotropy group GpG_{p} is a cloed subgroup of GG, and the F:G/GpMF:G/G_{p}\to M defined by F(gGp)=gpF(gG_{p})=g\cdot p is an equivariant diffeomorphism (see [27, Theorem 21.18]). Because of this equivariant diffeomorphism, we can define a homogeneous space to be a coset space of the form G/HG/H, where GG is a Lie group and HH is a closed subgroup of GG.

Let GG be a Lie group and fix aGa\in G. Define the left translation map La:GGL_{a}:G\to G by La(g)=agL_{a}(g)=ag. The left translation map is diffeomorphism since it is smooth with smooth inverse. The inverse of LaL_{a} is clearly the map La1L_{a^{-1}}. The diffeomorphism LgL_{g} induces a vector space isomorphism dLg|e:𝔤=TeGTgGdL_{g}|_{e}:\mathfrak{g}=T_{e}G\to T_{g}G, where 𝔤\mathfrak{g} is the Lie algebra of GG and ee is the identity element in GG.

Let M=G/HM=G/H be a homogeneous space and the natural projection π:GG/H\pi:G\to G/H, π(g)=gH\pi(g)=gH be a submersion. For each aGa\in G, define the left translation τa:G/HG/H\tau_{a}:G/H\to G/H by τa(gH)=agH\tau_{a}(gH)=agH. Then the left translations τg\tau_{g} are related to the left translations LgL_{g} on the Lie group GG by πLg=τgπ\pi\circ L_{g}=\tau_{g}\circ\pi for each gGg\in G. Let 𝔥\mathfrak{h} be the Lie algebra of HH and o=π(e)=eHo=\pi(e)=eH. The differential dπ|e:TeGToMd\pi|_{e}:T_{e}G\to T_{o}M is a vector space homomorphism with kerdπ|e=𝔥\ker d\pi|_{e}=\mathfrak{h}, we obtain that 𝔤/𝔥ToM\mathfrak{g}/\mathfrak{h}\cong T_{o}M, where 𝔤/𝔥\mathfrak{g}/\mathfrak{h} is the set of cosets X+𝔥={X+Y:Y𝔥}X+\mathfrak{h}=\{X+Y:Y\in\mathfrak{h}\} for X𝔤X\in\mathfrak{g} (see [16, P488] or [1, P71]).

A wedge WW is a closed and convex subset of a vector space that is invariant by scaling with real positive numbers (see e.g., [16] and [25]). Thus, a convex cone is a wedge in a vector space. A wedge field WMW_{M} on a manifold MM assigns to each point xMx\in M a wedge WM(x)W_{M}(x) in the tangent space TxMT_{x}M.

Let Φ:G×MM\Phi:G\times M\to M be any left group action on MM such that each of the maps τg:MM\tau_{g}:M\to M defined by τg(x)=Φ(g,x)=gx\tau_{g}(x)=\Phi(g,x)=g\cdot x forms a diffeomorphism of MM, Then a wedge field WMW_{M} is said to be G-invariant or homogeneous if dτg|x(WM(x))=WM(gx)d\tau_{g}|_{x}(W_{M}(x))=W_{M}(g\cdot x) for all gGg\in G and xMx\in M.

Lemma A.2.

Let HH be a closed subgroup of a Lie group GG and WW a wedge in 𝔤\mathfrak{g} such that (i) WW=𝔥W\cap-W=\mathfrak{h}, and (ii) Ad(H)(W)=W\mathrm{Ad}(H)(W)=W, where Ad\mathrm{Ad} is the adjoint representation of GG. Define WGW_{G} and CMC_{M} by

WG(g)=dLg|eW,CM(x)=dτg|oC,W_{G}(g)=dL_{g}|_{e}W,\ \ C_{M}(x)=d\tau_{g}|_{o}C,

where M=G/HM=G/H, ee is the identity element in GG, o=eHo=eH is the base point in MM, x=gHx=gH, and CC is the convex cone in ToMT_{o}M obtained as the projection of WW onto 𝔤/𝔥\mathfrak{g}/\mathfrak{h}. Then, WGW_{G} is an invariant wedge field on GG and CMC_{M} is a well-defined homogeneous or G-invariant cone field on MM. Moreover, for each gGg\in G,

dπ|g(WG)=CM(π(g)),d\pi|_{g}(W_{G})=C_{M}(\pi(g)),

where π:GM\pi:G\to M is the canonical projection π(g)=gH\pi(g)=gH.

Proof.

See [16, Lemma VI.1.5]. ∎

Proposition A.6.

The homogeneous cone field on a homogeneous space is continuous and admits CC^{\infty} sections.

Proof.

See [26, Proposition 4.6]. ∎

Theorem A.2.

Let CMC_{M} be a homogeneous cone field on M=G/HM=G/H as described in Lemma A.2. If S=<expW>H¯GS=\overline{<\exp W>H}\subset G, then S=π1({xM:oMx}¯)S=\pi^{-1}(\overline{\{x\in M:o\leq_{M}x\}}) and G/HG/H is globally orderable with respect to CMC_{M} if and only if W=L(S)W=\textbf{L}(S), where L(S)={Z𝔤:exp(+Z)S}\textbf{L}(S)=\{Z\in\mathfrak{g}:\exp(\mathbb{R}^{+}Z)\subset S\}.

Proof.

This Theorem is derived from [35, Theorem 1.6] or [17, Theorem 4.21]. Where exp:𝔤G\exp:\mathfrak{g}\to G is the Lie group exponential map, exp(tZ)\exp(tZ) with Z𝔤,tZ\in\mathfrak{g},t\in\mathbb{R} is the one-parameter subgroup on GG (see e.g., [27]) and M\leq_{M} is the order generated by cone field CMC_{M}. ∎

Let M=G/HM=G/H be a homogeneous space with base-point o=eHo=eH. A Riemannian metric (,)p(\cdot,\cdot)_{p}, pMp\in M, is said to be G-invariant or homogeneous, if it satisfies (X,Y)o=(dτg(X),dτg(Y))go(X,Y)_{o}=(d\tau_{g}(X),d\tau_{g}(Y))_{g\cdot o} for each gGg\in G and X,YToMX,Y\in T_{o}M.

Proposition A.7.

Let GG be a Lie group, HH a closed subgroup, then the space G/HG/H is complete in any G-invariant metric.

Proof.

See [14, p148]. ∎

A homogeneous space M=G/HM=G/H is called reductive if there exists a subspace 𝔪\mathfrak{m} of 𝔤\mathfrak{g} such that 𝔤=𝔥𝔪\mathfrak{g}=\mathfrak{h}\oplus\mathfrak{m} and Ad(h)𝔪𝔪\text{Ad}(h)\mathfrak{m}\subset\mathfrak{m} for all hHh\in H. Hence, 𝔪ToM\mathfrak{m}\cong T_{o}M (see [1]). The next Proposition gives a simple description of G-invariant Riemannian metrics on a homogeneous space.

Proposition A.8.

Let M=G/HM=G/H be a reductive homogeneous space. Then there is a one-to-one correspondence between G-invariant Riemannian metrics on M=G/HM=G/H and AdG/H\mathrm{Ad}^{G/H}-invariant inner products ,\langle\cdot,\cdot\rangle on 𝔪ToM\mathfrak{m}\cong T_{o}M; that is, X,Y=AdG/H(h)X,AdG/H(h)Y\langle X,Y\rangle=\langle\mathrm{Ad}^{G/H}(h)X,\mathrm{Ad}^{G/H}(h)Y\rangle for all X,Y𝔪X,Y\in\mathfrak{m}, hHh\in H.

Proof.

See [1, Proposition 5.1]. The homomorphism AdG/H:HGL(ToM)\mathrm{Ad}^{G/H}:H\to GL(T_{o}M) such that AdG/H(h)=dτh|o\mathrm{Ad}^{G/H}(h)=d\tau_{h}|_{o} is the isotropy representation of the homogeneous space M=G/HM=G/H. ∎

A.3 Differential positivity in flat spaces

Let MM be the nn-dimensional Eucliean space n\mathbb{R}^{n} and CC is a closed convex cone of MM. There is a partial order “\leq” on MM generated by cone CC (xyx\leq y if and only if yxCy-x\in C). It should be pointed out that the order introduced by a closed convex cone is a closed partial order.

We consider the cone field on MM defined by CM(x)={x}×CTM=n×nC_{M}(x)=\{x\}\times C\subset TM=\mathbb{R}^{n}\times\mathbb{R}^{n}. Such cone field is said to be a contant cone field. Thus, we can define the order “M\leq_{M}” on MM with respect to cone field CMC_{M} (xMyx\leq_{M}y if and only if there exists a conal curve γ:[0,1]M\gamma:[0,1]\to M such that γ(0)=x\gamma(0)=x, γ(1)=y\gamma(1)=y and ddsγ(s)CM(γ(s))=C\frac{d}{ds}\gamma(s)\in C_{M}(\gamma(s))=C).

It is easy to see that the contant cone field on Eucliean space n\mathbb{R}^{n} satisfies the smoothness conditions and invariance condition in the previous sections.

The next proposition implies that the order “M\leq_{M}” generated by contant cone feild CMC_{M} agrees with the partial order “\leq” on MM generated by cone CC (see [26, Proposition 1.10]).

Proposition A.9.

Let MM be a nn-dimensional Eucliean space and CC is a convex cone in MM such that CC forms a constant cone field CMC_{M}. Then for x,yMx,y\in M, xyx\leq y if and only if xMyx\leq_{M}y.

Proof.

If xyx\leq y, i.e., yxCy-x\in C, then we choose a curve α(s)=sy+(1s)x\alpha(s)=sy+(1-s)x, where s[0,1]s\in[0,1]. Thus, α(0)=x\alpha(0)=x, α(1)=y\alpha(1)=y, and ddsα(s)=yxC\frac{d}{ds}\alpha(s)=y-x\in C. So, α(s)\alpha(s) is a conal curve and xMyx\leq_{M}y.

If xMyx\leq_{M}y, then there is a conal curve α:[0,1]M\alpha:[0,1]\to M such that α(0)=x\alpha(0)=x, α(1)=y\alpha(1)=y and ddsα(s)C\frac{d}{ds}\alpha(s)\in C. For any λC\{0}\lambda\in C^{*}\backslash\{0\}, where CC^{*} is the dual cone of CC, λ(ddsα(s))0\lambda(\frac{d}{ds}\alpha(s))\geq 0. On the other hand, λ(yx)=λ(α(1)α(0))=λ(α(1))λ(α(0))\lambda(y-x)=\lambda(\alpha(1)-\alpha(0))=\lambda(\alpha(1))-\lambda(\alpha(0)). Since ddsλ(α(s))=λ(ddsα(s))0\frac{d}{ds}\lambda(\alpha(s))=\lambda(\frac{d}{ds}\alpha(s))\geq 0, then λ(α(1))λ(α(0))\lambda(\alpha(1))\geq\lambda(\alpha(0)). So, λ(yx)0\lambda(y-x)\geq 0. Thus, we obtain that yxCy-x\in C, i.e., xyx\leq y. ∎

And a similar result can be obtained for the order “\ll” with “M\ll_{M}”, where xyx\ll y if and only if yxIntCy-x\in\text{Int}C.

Let dxds=f(x)\frac{dx}{ds}=f(x) be a dynamical system in n\mathbb{R}^{n} with the flow φt\varphi_{t}. The system is said to be monotone with respect to partial order “\leq” if φt(x)φt(y)\varphi_{t}(x)\leq\varphi_{t}(y) whenever xyx\leq y and t0t\geq 0 and stongly monotone if φt(x)φt(y)\varphi_{t}(x)\ll\varphi_{t}(y) whenever xyx\leq y, xyx\neq y and t>0t>0 (see [22] and [50]).

The following Proposition shows that in n\mathbb{R}^{n}, a monotone system is differentially positive (see [9, Theorem 1]).

Proposition A.10.

Let MM be the nn-dimensional Eucliead space and CC is a convex cone in MM such that CC forms a constant cone field CMC_{M}. Then a system is monotone with respect to partial order “\leq” generated by cone CC if and only if this system is differentially positive with respect to cone field CMC_{M}.

Proof.

By Proposition A.9, order “\leq” and order “M\leq_{M}” are equivalent.

Suppose that the system is differentially positive. For xyx\leq y, there exists a conal curve γ:[0,1]M\gamma:[0,1]\to M such that γ(0)=x\gamma(0)=x, γ(1)=y\gamma(1)=y and ddsγ(s)C\frac{d}{ds}\gamma(s)\in C. Then, we obtain that φt(γ())\varphi_{t}(\gamma(\cdot)) is also a conal curve for each t0t\geq 0. In fact, ddsφt(γ(s))=dφt(γ(s))ddsγ(s)\frac{d}{ds}\varphi_{t}(\gamma(s))=d\varphi_{t}(\gamma(s))\frac{d}{ds}\gamma(s). Since φt\varphi_{t} is differentially positive and ddsγ(s)C\frac{d}{ds}\gamma(s)\in C, then ddsφt(γ(s))C\frac{d}{ds}\varphi_{t}(\gamma(s))\in C. So, φt(x)φt(y)\varphi_{t}(x)\leq\varphi_{t}(y) for t0t\geq 0. And hence, φt\varphi_{t} is monotone with respect to “\leq”.

If the system is monotone. For xMx\in M and vCv\in C, then there is a conal curve γ(s)\gamma(s), sIs\in I, such that γ(0)=x\gamma(0)=x, ddsγ(s)|s=0=v\frac{d}{ds}\gamma(s)|_{s=0}=v and xγ(s)x\leq\gamma(s) with s0s\geq 0. Since φt\varphi_{t} is monotone, then for each t0t\geq 0, φt(x)φt(γ(s))\varphi_{t}(x)\leq\varphi_{t}(\gamma(s)), i.e., φt(γ(s))φt(γ(0))C\varphi_{t}(\gamma(s))-\varphi_{t}(\gamma(0))\in C. Thus, ddsφt(γ(s))|s=0=limΔs0+φt(γ(Δs))φt(γ(0))ΔsC\frac{d}{ds}\varphi_{t}(\gamma(s))|_{s=0}=\lim\limits_{\Delta s\to 0^{+}}\frac{\varphi_{t}(\gamma(\Delta s))-\varphi_{t}(\gamma(0))}{\Delta s}\in C. So, we obtain that dφt(x)v=ddsφt(γ(s))|s=0Cd\varphi_{t}(x)v=\frac{d}{ds}\varphi_{t}(\gamma(s))|_{s=0}\in C. And hence, the system is differentially positive. ∎

References

  • [1] A. Arvanitoyeorgos, An introduction to Lie groups and the geometry of homogeneous spaces, vol 22. American Mathematical Society, Providence, 2003.
  • [2] J. K. Beem and P. E. Ehrlich, Global Lorentzian Geometry, Monographs and Textbooks in Pure and Applied Mathematics, 67. Marcel Dekker, Inc., New York, 1981.
  • [3] P. T. Chruściel and J. D. E. Grant, On Lorentzian causality with continuous metrics, Class. Quantum Grav. 29(2012), 145001.
  • [4] I. Chueshov, Monotone Random Systems Theory and Applications, Lecture Notes in Mathematics, vol. 1779, Springer-Verlag, Berlin, 2002.
  • [5] J. Faraut, Algèbres de Volterra et transformation de Laplace sphérique sur certains espaces symétriques ordonnés, Symp. Math. 29(1987), 183-196.
  • [6] J. Faraut, Espaces symétriques ordonnés et algèbres de Volterra, J. Math. Soc. Japan, 43(1991), 133-147.
  • [7] F. Forni and R. Sepulchre, Differential analysis of nonlinear systems: Revisiting the pendulum example, in 53rd IEEE Conference of Decision and Control, Los Angeles, IEEE, Piscataway, NJ, 2014, 3848-3859.
  • [8] F. Forni, Differential positivity on compact sets, in 2015 54th IEEE Conference on Decision and Control (CDC), IEEE, Piscataway, NJ, 2015, 6355-6360.
  • [9] F. Forni and R. Sepulchre, Differentially positive systems, IEEE Trans. Automat. Control, 61(2016), 346-359.
  • [10] L. García-Heveling, Causality theory of spacetimes with continuous Lorentzian metrics revisited, Class. Quantum Grav. 38(2021), 145028.
  • [11] A. K. Guts and A. V. Levichev, On the Foundations of Relativity Theory, Sov. Math. Dokl. 30(1984), 253-257.
  • [12] S. H. Hawking and G. F. R. Ellis, The large scale structure of space-time, Cambridge University Press, Cambridge, 1973.
  • [13] S. H. Hawking and R. K. Sachs, Causally continuous space-times, Commun. Math. Phys. 35(1974), 287-296.
  • [14] S. Helgason, Differential Geometry, Lie Groups, and Symmetric Spaces, Amer. Math. Soc., 2001.
  • [15] P. Hess and P. Poláčik, Boundedness of prime periods of stable cycles and convergence to fixed points in discrete monotone dynamical systems, SIAM J. Math. Anal. 24(1993), 1312-1330.
  • [16] J. Hilgert, K. H. Hofmann and J. Lawson, Lie groups, convex cones, and semigroups, Oxford University Press, Oxford, 1989.
  • [17] J. Hilgert and K.-H. Neeb, Lie semigroups and their applications, Lecture Notes in Math. 1552, Springer, 1993.
  • [18] J. Hilgert and K.-H. Neeb, Wiener-Hopf operators on ordered homogeneous spaces, I, J. Funct. Anal. 132(1995), 86-118.
  • [19] J. Hilgert and G. Ólafsson, Causal Symmetric Spaces. Geometry and Harmonic Analysis, Perspectives in Mathematics, vol. 18, Academic Press, 1997.
  • [20] M. W. Hirsch, Systems of differential equations which are competitive or cooperative II: convergence almost everywhere, SIAM J. Math. Anal. 16(1985), 423-439.
  • [21] M. W. Hirsch, Stability and convergence in strongly monotone dynamical systems, J. Reine Angew. Math. 383(1988), 1-53.
  • [22] M. W. Hirsch and H. L. Smith, Monotone dynamical systems, Handbook of Differential Equations: Ordinary Differential Equations, vol. 2, Elsevier, Amsterdam 2005.
  • [23] J. Jiang and X. Zhao, Convergence in monotone and uniformly stable skew-product semiflows with applications, J. Reine Angew. Math. 589(2005), 21-55.
  • [24] T. Kato, Perturbation theory for linear operators, Springer-Verlag, Berlin, 1976.
  • [25] M. A. Krasnosel’skij, Je. A. Lifshits and A. V. Sobolev, Positive linear systems, the method of positive operators, Heldermann Verlag, Berlin, 1989.
  • [26] J. D. Lawson, Ordered manifolds, invariant cone fields, and semigroups, Forum Math. 1(1989), 273-308.
  • [27] J. M. Lee, Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics, Springer, 22 edition, 2013.
  • [28] B. Lemmens and R. D. Nussbaum, Nonlinear Perron-Frobenius Theoty, Cambridge, Cambridge Univ. Press, 2012.
  • [29] A. Levichev, Sufficient conditions for the nonexistence of closed causal curves in homogeneous space-times, Izv. Phys. 10(1985), 118-119.
  • [30] J. Mierczyński and W. Shen, Spectral Theory for Random and Nonautonomous Parabolic Equations and Applications, Chapman Hall/CRC Monogr. Surv. Pure Appl. Math., vol. 139, CRC Press, Boca Raton, FL, 2008.
  • [31] V. Morinelli and K.-H. Neeb, Covariant homogeneous nets of standard subspaces, Commun. Math. Phys. 386(2021), 305-385.
  • [32] C. Mostajeran and R. Sepulchre, Ordering positive definite matrices, Inf. Geom. 1(2018), 287-313.
  • [33] C. Mostajeran and R. Sepulchre, Monotonicity on homogeneous spaces, Math. Control Signals Systems, 30(2018), 1-25.
  • [34] C. Mostajeran and R. Sepulchre, Positivity, monotonicity, and consensus on Lie groups, SIAM J. Control Optim. 56(2018), 2436-2461.
  • [35] K.-H. Neeb, Conal orders on homogeneous spaces, Inventiones mathematicae 104(1991), 467-496.
  • [36] K.-H. Neeb and G. Ólafsson, Nets of standard subspaces on Lie groups, Advances in Math. 384(2021), 107715.
  • [37] K.-H. Neeb and G. Ólafsson, Wedge domains in compactly causal symmetric spaces, Int. Math. Res. Not. IMRN 12(2023), 10209-10312.
  • [38] K.-H. Neeb and G. Ólafsson, Wedge domains in non-compactly causal symmetric spaces, Geometriae Dedicata 217(2023) 30.
  • [39] K.-H. Neeb, B. Ørsted and G. Ólafsson, Standard subspaces of Hilbert spaces of holomorphic functions on tube domains, Commun. Math. Phys. 386(2021), 1437-1487.
  • [40] L. Niu, Eventually competitive systems generated by perturbations, Electronic Journal of Differential Equations 121(2019) 1-12.
  • [41] L. Niu and X. Xie, Generic Poincaré-Bendixson theorem for singularly perturbed monotone systems with respect to cones of rank-2, Proc. Amer. Math. Soc. 151 (2023) 4199-4212.
  • [42] G. I. Olshanskii, Convex cones in symmetric Lie algebras, Lie semigroups and invariant causal (order) structures on pseudo-Riemann symmetric spaces, Soviet Math. Dokl., Amer. Math. Soc. translation 26(1982) 97-101.
  • [43] R. Penrose, Techniques of differential topology in relativity, Regional Conference Series in Applied Math. vol. 7, SIAM, Philadelphia, 1972.
  • [44] P. Poláčik, Convergence in smooth strongly monotone flows defined by semilinear parabolic equations, J. Differ. Equ. 79(1989) 89-110.
  • [45] P. Poláčik, Parabolic equations: asymptotic behavior and dynamics on invariant manifolds, in: Handbook on Dynamical systems, vol. 2, Elsevier, Amsterdam, 2002, 835-883.
  • [46] P. Poláčik and I. Tereščák, Convergence to cycles as a typical asymptotic behavior in smooth strongly monotone discrete-time dynamical systems, Arch. Ration. Mech. Anal. 116(1992) 339-360.
  • [47] C. Pugh, An improved Closing Lemma and general density theorem, Amer. J. Math. 89(1967) 1010-1021.
  • [48] I. E. Segal, Mathematical Cosmology and Extragalactic Astronomy, Academic Press, New York, 1976.
  • [49] W. Shen and Y. Yi, Almost automorphic and almost periodic dynamics in skew-product semiflows, Mem. Am. Math. Soc. 136(647) 1998.
  • [50] H. L. Smith, Monotone Dynamical Systems, an introduction to the theory of competitive and cooperative systems, Math. Surveys and Monographs, 41, Amer. Math. Soc., Providence, Rhode Island 1995.
  • [51] H. L. Smith, Monotone dynamical systems: Reflections on new advances and applications, Discrete Contin. Dyn. Syst. 37(2017), 485-504.
  • [52] H. L. Smith and H. R. Thieme, Convergence for strongly ordered preserving semiflows, SIAM J. Math. Anal. 22(1991), 1081-1101.
  • [53] I. Tereščák, Dynamics of C1C^{1} smooth strongly monotone discrete-time dynamical systems, preprint, Comenius University, Bratislava, 1994.
  • [54] E. B. Vinberg, Invariant cones and orderings in Lie groups, Funct. Anal. Appl. 14(1980), 1-13.
  • [55] Y. Wang and J. Yao, Dynamics alternatives and generic convergence for C1C^{1}-smooth strongly monotone discrete dynamical systems, J. Differ. Equ. 269(2020), 9804-9818.
  • [56] Y. Wang and J. Yao, Sharpened dynamics alternative and its C1C^{1}-robustness for strongly monotone discrete dynamical systems, J. Funct. Anal. 283(2022), 109538.