Generic behavior of differentially positive systems
on a globally
orderable Riemannian manifold
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 -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 assigns to each point of a manifold a closed convex cone in the tangent space of .
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 -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 naturally induces a “conal order relation” as follows: two points are conal ordered, denoted by , if there exists a conal curve on beginning at and ending at . 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 “” induced by a cone field on is certainly reflexive and transitive, but not always antisymmetric. In fact, there are examples of 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 “” is not necessarily closed, i.e., the set is not a closed subset in (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 “”. 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 a canonical defined vector field , called the Perron-Frobenius vector field, such that the vector lines in the interior of the cone at each point . By appealing to the Perron-Frobenius vector field , 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 , the -limit set 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 . 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 :
-
(H1)
is globally orderable equipped with a continuous solid cone field.
(H1) means that the conal order “” on 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 , the -limit set is a singleton.
In the present paper, we will prove conjecture A under the following two reasonable assumptions:
-
(H2)
The conal order “” is quasi-closed.
-
(H3)
Both the cone field and the Riemannian metric on are -invariant.
More precisely, our main result is the following
Theorem A.
Assume that (H1)-(H3) hold. Then, for almost every , the -limit set 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 “” in (H2) means that whenever and as and for all integer (see Definition 2.3). Here, we write 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 beginning at and ending at (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 is derived from the timelike curve in general relativity (see e.g., [43]). In fact, one can prove that the conal order is indeed quasi-closed in many causally continuous space-times (see Theorem A.1 in the Appendix).
As for (H3), a cone field is called -invariant if there exists a linear invertible mapping for all , such that . While, the Riemannian metric on is -invariant if for all and . Actually, -invariance is motivated by the homogeneous structure of the manifolds. A well-known example admitting (H3) is a homogeneous space of a connected Lie group assigned to each point a closed convex cone in the tangent space such that the cone field is invariant under the action of (see e.g., [26, 35, 17]); and a homogeneous Riemannian metric on such homogeneous space is naturally -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 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 , 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 . 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 . 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, will be reserved as a smooth manifold of dimension . The tangent bundle is denoted by and the tangent space at a point by . is endowed with a Riemannian metric tensor, represented by a inner product on the tangent space . We set for any (sometimes we omit the subscripts in ) and the Riemannian metric endows the manifold with the Riemannian distance . We assume that is a complete metric space.
Let be a dimensional real linear space. A nonempty closed subset of the linear space is called a closed convex cone if , for all , and . A convex cone is solid if its interior . A convex cone in is said to surround a cone if .
A cone field on a manifold is a map , such that is a closed convex cone in for each . A manifold equipped with a cone field is called a conal manifold. A cone field is solid if each convex cone is solid.
A cone field is called a -invariant cone field if there exists a linear invertible mapping for each , such that . Moreover, is continuous with respect to and for all , where is the identity map. We say that the Riemannian metric is -invariant under the linear invertible mapping if for all and . In this case, holds for all and .
Pick a smooth chart , where is an open set containing and is an open set. Let denote the representation of the cone field in the chart, i.e., . Following [26], a cone field on is upper semicontinuous at if given a smooth chart at and a convex cone surrounding , there exists a neighborhood of such that for all . The cone field is lower semicontinuous at if given any open set such that is non-empty, there exists a neighborhood of such that is non-empty for all . A cone field is continuous at if it is both lower and upper semicontinuous at and continuous if it is continuous at every .
Definition 2.1.
A continuous piecewise smooth curve defined on into a conal manifold is called a conal curve if whenever , 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, is called a strictly conal curve if for .
For two points , we say and are ordered, denoted by , if there exists a conal curve such that and . This relation is an order on . In fact, it is always reflexive (i.e., for all ) and transitive (i.e., and implies ). The order “” is referred to as the conal order. We write if there exists a so-called strictly conal curve with and . Clearly, the relation “” is always transitive. Let be the subsets of . We write (resp., ) if (resp., ) for any and .
The following proposition implies that the relation “” is open.
Proposition 2.1.
If , then there exist neighborhoods of , of such that .
Proof.
We only prove the existence of , i.e., if , there exists a neighborhood of such that for each . Since , there is a strictly conal curve such that and . Let , . Then , and .
Let be a chart with such that is convex. We set for and . Then . So, we can choose a compact convex neighborhood of in the interior of and set . Since the cone field is continuous, there is a neighborhood of such that for all and is convex.
On the other hand, . Then there is a such that and . Let be an open convex neighborhood of in .
Let , then is an open neighborhood of . For each , we set , . Then , and .
Thus, there is a point and an open neighborhood of such that for all . Since , we obtain for all . ∎
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 “” is a partial order which is locally order convex.
The conal order “” is a partial order relation if it is additionally antisymmetric (i.e., and implies ). A partial order “” is locally order convex if it has a basis of neighborhoods at each point that are order convex ( implies ). See [26, 35].
Remark 2.1.
In the terminology of general relativity, each point in a globally orderable manifold is called a strong point (see Lawson [26, Section 5]), by which means that every neighborhood of contains a smaller neighborhood of such that every conal curve that begins in and leaves terminates outside of (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 “” is quasi-closed if whenever and as and for all .
Remark 2.3.
Now, we consider a system generated by a smooth vector field on equipped with a continuous solid cone field . The induced flow by system is denoted by . We write as the tangent map from to . Let , the positive semiorbit (resp., negative semiorbit) of , be denoted by (resp., ). The full orbit of is denoted by . An equilibrium is a point such that . Let be the set of all the equilibria of . The -limit set of is defined by . A point if and only if there exists a sequence , , such that as . If is precompact, then is nonempty, compact, connected, and invariant (i.e., for any ). A point is called a convergent point if consists of a single point of . The set of all convergent points is denoted by .
Throughout the paper, we always assume that all orbits are forward complete, which means is well defined for all , and the orbit of has compact closure for each .
Definition 2.4.
The system is said to be differentially positive (DP) with respect to if
And the differentially positive system is said to be strongly differentially positive (SDP) with respect to if
See Fig.1(b).
In the present paper, we focus on the strongly differentially positive system on a Riemannian manifold . We first give the following two useful preliminary results.
Proposition 2.2.
If , then for . Moreover, for all .
Proof.
If , there exists a conal curve such that and . Since is SDP and , then for . So, is a strictly conal curve. ∎
Proposition 2.3.
If , then for , there exist neighborhoods of , of such that for any .
Proof.
Since , for . One can take neighborhoods of and of such that by Proposition 2.1. By the continuity of , there are neighborhoods of , of such that and . ∎


Fig 1: Conal curves and strongly differentially positive flows on manifold .
As mentioned in the introduction, we hereafter impose the following hypotheses:
-
(H1)
is a globally orderable conal manifold equipped with a continuous solid cone field .
-
(H2)
The conal order “” is quasi-closed.
-
(H3)
Both the cone field and the Riemannian metric on are -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
-
(a)
The -limit set cannot contain two points and with ;
-
(b)
If , then either , or .
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 under the hypotheses (H1)-(H3).
Theorem 3.1.
(Generic Convergence) Assume that (H1)-(H3) hold. If the cone field admits a -section and satisfies the compactness condition (P), then is dense in .
Recall that denotes the set of all convergent points. Here, we say that a cone field admits a -section if there is a vector field such that for all . We also formulate the compactness condition (P): For each , has compact closure contained in , where with (or ) for .
Remark 3.1.
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 on an invariant compact set is called a Perron-Frobenius vector field on , if for each ,
(i) with ;
(ii) for ;
(iii) for all ,
where is the Hilbert Metric induced by (see [9, Section VI]).
Lemma 3.1.
Let and be the Perron-Frobenius vector field on . If is not a singleton, then there exist and such that
Moreover, there exists such that for all .
Proof.
Fix . Since , then for all . So, for some .
We then pick a smooth chart , where is the coordinate neighborhood of , is an open set in . Using this coordinate map , we write , and . Let . Then and . So, . Since system is SDP and , then . Moreover, , since . By the Perron-Frobenius Theorem, we obtain that , where is the spectral radius of .
Since is not a singleton, we obtain that there exists a sequence such that and . Let . Then , where as . Let , then , where as . If as , we obtain that . Hence, .
If , then . Hence, there exists a neighborhood of in the interior of such that there is a convex neighborhood of satisfies that for any , . Thus, there is a such that and for all . Let for , then . Then, is a conal curve in connecting and . Thus, . Since , it is a contradiction to Lemma 2.1(a). Thus, we obtain that . The conclusion of the lemma follows from the compactness of and the continuity of the spectral radius (see e.g., [24, 28]). ∎
Proposition 3.1.
If , then either , or for some .
Proof.
We just need to prove the case that for some by Lemma 2.1(b). Suppose that . Let be the one in Lemma 3.1. Let , for . Since , there exists a conal curve such that , and . Let , then is a strictly conal curve connecting and such that . Suppose that contains more than a single element.
By passing to a subsequence if necessary, we assume that , as , where . Thus, . Otherwise, with , which is a contradiction to Lemma 2.1(a). The length of is . We assert that as .
is a globally orderable conal manifold, then every point in is a strong point (see [26, Proposition 5.3]). We claim that for any open neighborhood of , there exists a such that for all . In fact, suppose there exist a neighborhood of and a subsequence such that leaves for all . Since is a strong point, there exists open containing such that every conal curve that begins in and leaves terminates outsides of (see [26, Lemma 5.2]). On the other hand, and , then there exists a such that for , and are in , which is a contradiction. Thus, we have the claim. Since is locally compact, we can find an open neighborhood of such that is compact. By the previous claim, there is a such that for all . Let . Clearly, . If there is a with , then there exist neighborhoods of and of such that since is a Hausdorff space. implies that there is a subsequence such that , where . Then there is a such that for , with , which contradicts the previous claim for neighborhood of . Hence, . If there exist a and a subsequence such that , then , which is a contradiction. So, as .
Since is attracted to , one can choose such that as . On the other hand, as . Hence, as . Then there exist a and a coordinate neighborhood of such that for all . Since there is a smooth coordinate chart , all notations in the following have coordinate representations. With the diffeomorphism , we treat the following notations both in manifold and .
For each , we have
(3.1) |
where is defined as above.
Since as , and is continuous with for any , then for any , there is a such that for , . Since is smooth, then for any , there is a such that for , . Let and , then for , we can obtain that and . If there is a such that for all , then for , we have that
(3.2) |
Next, we deal with the term in inequation (3.1). Due to Lemma 3.1, we have that , where and is the Perron-Frobenius vector field. And we obtain that as . Since preserves the metric, . Since and are continuous, we have as . Hence, we obtain that as . Since is bounded for , let , then as and we have that
Thus,
(3.3) |
Since as and by the Lemma 3.1, we can choose small enough such that there exist a and satisfying for all , . Thus, we obtain that . So, , which is a contradiction to as . Thus, is a singleton. ∎
Lemma 3.2.
Suppose that the cone field on admits a -section , then for each , there exist and a conal curve such that and for .
Proof.
See [27, Proposition 9.2]. In fact, is the integral curve of . ∎
By the Lemma 3.2, we obtain that for any , there exists a conal curve passing through . Thus, we say that can be approximated from below (resp. above) in (i.e., there exists a sequence in such that (resp. ) for and as ).
Lemma 3.3.
If satisfies that there exists a sequence such that for and , then one of the following alternatives must occur :
-
There exists such that for all .
-
for .
-
There exists such that for all .
Proof.
Let have the property that it is approximated from below in by a sequence such that for and . By Proposition 3.1, either there exists a such that for all , or there is a subsequence such that for all . Thus, we assume that either for all , or , where .
Suppose that for and as . We first obtain that for all . In fact, if there exists such that , then for all , which is a contradiction. Let . By the compactness condition , the set belogs to a compact set . Furthermore, is nonempty. If and are two subsequences of such that and as . Since , then for each , there exists such that . Thus, . A similar argument shows that , then . So, and . Thus, is nonempty. Suppose that have the property that there exist two sequences , with and , as . Since (resp. ), we obtain that (resp. ). Then . Thus, is a singleton. On the other hand, since each is invariant, we obtain that is positively invariant. So, . It is easy to see from the arbitrariness of in that . Since , then . If , then by Lemma 2.1(a). Thus, is a convergent point and holds. If , we will get a contradiction. Since and is invariant, we obtain that . For each , there exist , a neighborhood of and a neighborhood of such that for all . Since is an open cover of , we obtain that , where . Meanwhile, is a neighborhood of . Let , then for . On the other hand, there exists such that . Since , there is a such that . By , we have that for . Thus, . By the definition of , for all . Thus, and for all , which is a contradiction. Hence, we have proved the case .
Suppose that for all . By Proposition 3.1, for all . Since , then or . So, and hold. ∎
Remark 3.2.
An analogous result holds if is approximated from above.
Now, we are ready to prove Theorem 3.1.
Proof of Theorem 3.1.
Suppose that , then we will prove that . In fact, there is a sequence such that . Since the cone field admits a -section, then all can be approximated from below and above in . Without loss of generality, we assume that for each , there exists a sequence such that for and as . Since for all , then the case of Lemma 3.3 holds for each sequence .
We claim that for each . If the claim holds, then since and is a closed set. Thus, we obtain the theorem.
Now, it suffices to prove the claim. For each , as and , then there exists a such that for all . For , there exist a neighborhood of and such that for . Since is an open cover of , we obtain that , where . Let , then for . On the other hand, there is a such that . Furthermore, one can choose a neighborhood of such that . If with , then there exist a neighborhood of , , a neighborhood of and such that for . We can choose a such that , then we have for . Since , and for , we obtain that for . Thus, for . Since for all , then for all . Hence, . Since the cone field admits a -section, then there is a sequence in such that and as . By the previous proof, . Thus, . 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 for some , then as .
Proof.
If and is SDP, then for , there exist neighborhoods of , of such that by Proposition 2.3.
By the continuity, there exists such that for . We first claim that is a -periodic orbit for any . We just prove the case of and a similar argument for . If , then for . Thus, there exists such that as . In fact, if there exist two sequences and satisfying and , then for each , there is a such that . Thus, by (H2). A similar argument shows that . So, by (H1). Consider the orbit of , for all . Thus, is a -periodic orbit. Suppose that there exist and such that and as . For each , , , , then . Thus, as where such that as . So, .
Since is a -periodic orbit for all and , then for all (i.e., is -periodic). Let be the set of all periods of . It is easy to see that and for all and . Thus, . We next prove that . In fact, let and , then . Thus, . Then .
For each , , where , and . . Thus, and . ∎
Now, we prove Lemma 2.1(a).
Proof of Lemma 2.1(a).
In order to prove Lemma 2.1(b), we further need several propositions.
Proposition 4.2.
If , , and as , then .
Proof.
Since is SDP, there exist a neighborhood of , a neighborhood of for such that by Proposition 2.3. Let be small such that and . Then for any . Thus, for large enough. Let and , we obtain for all . As a similar argument, we obtain for all . So, for all . For any , we write , where and . So, . Thus, . ∎
Proposition 4.3.
If , then .
Proof.
Lemma 4.1.
Let , , and . If (or ) is an equilibrium, then .
Proof.
Assume that . Since and , there exists a sequence such that and hence, there exists a such that . Thus, for all . Thus, . By Lemma 2.1(a), . Moreover, . In fact, since is invariant, for any , there exists such that and . Thus, .
Since , there exists a sequence such that . Since , there exists a such that . Thus, for all . Since is compact and invariant, we obtain that . By Lemma 2.1(a), . Since are compact and invariant, then . In fact, let and , there is a such that and with , then .
A similar argument is used if . ∎
Lemma 4.2.
Assume that is a compact set in which the flow is SDP. Then there exists with the following property. Let denote the flow of a vector field such that , where is a -neighborhood of in space of vector fields on with the topology. Then there exists such that if is positively invariant under the flow of , then is SDP for all .
Proof.
We first assume that , where is fixed. Since is SDP, we obtain that for all with . With a coordinate chart, we treat the following notations both in manifold and . By the continuity of the cone field, there exist neighborhoods of and of such that for all . So, there exists such that for , and (see e.g., [40, 41]). Thus, for , .
For , let us write , where and . Define , . It is clear that if is positively invariant. Then . By the preceding proof, for .
Thus, we have proved that is SDP for all . ∎
Lemma 4.3.
Let , , and . If (or ) belongs to a periodic orbit, then .
Proof.
Assume that , is a periodic orbit and is not an equilibrium (the other case is similar).
If , then . Thus, by Proposition 4.3, which is a contradiction. Thus, . Since , then the orbit closure of is in the compact set .
By the Closing Lemma (see e.g., [20, 47]), there is a vector field whose flow has a closed orbit passing through . Moreover, can be chosen to approximate as closely as desired and to coincides with outside a given neighborhood of the orbit closure of with respect to such that . Thus, is eventually SDP by Lemma 4.2 and is also a closed orbit of . In the following, we write (resp. ) as the orbit generated by (resp. ). So, .
For the system generated by vector field , and are two periodic orbits and , with . Then .
In fact, if , then there is a such that for all . Let be the periods of and and . If is a rational number, let , , then . Define . Suppose that . Then . Since , one has . Thus, there exists such that for all , which is a contradiction. Thus, for all . So, . If is an irrational number and , the set is dense in for any fixed . For any , there exists a sequence such that and as . Thus, . So, .
For system , we obtain that . Since , then there is a sequence such that as . Then there exists such that . Thus, we obtain that for all . So, . On the other hand, and are invariant, then . Since , in a similar way we can obtain that . ∎
Proposition 4.4.
Let , , and . Then .
Proof.
If , then by Lemma 4.1. In the following, we assume that is not an equilibrium. Thus, .
By the Closing Lemma (see e.g., [20, 47]), there exists such that for any , we can choose a vector field such that outside the -neighborhood of the set . Moreover, with respect to the vector field has a closed orbit passing through and is eventually SDP. For small , . It is easy to see that there exists a such that and , where . In fact, since , there is a neighborhood of such that and . Then there is a such that . Let , then and since outside with . Since , and , then by Lemma 4.3. Thus, . Since , there exists a sequence such that . Then there exists such that . Thus, we obtain that for all . So, . Since and are nonordering invariant sets and is SDP, then and . ∎
Now, we are ready to prove Lemma 2.1(b).
Appendix A Appendix
A.1 Order structures on space-times
A Lorentz metric for a smooth manifold of dimension four is a smooth nondegenerate symmetric tensor field of type on such that for each , by suitable choice of the basis, has the matrix diag. A space-time is a connected Hausdorff manifold of dimension four with a Lorentz metric .
Remark A.1.
Let be a space-time, a vector is said to be timelike, null, spacelike according to whether is negative, zero, or positive, respectively. The non-spacelike (i.e., timelike and null) vectors in form two so-called Lorentzian cones and (see e.g., [12]). Furthermore, the timelike vectors form the interior of the Lorentzian cones. See Fig.A.1.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/b35d96d8-52af-49b8-82d6-ed458acc7f8b/x3.png)
Fig A.1: The timelike vectors, null vectors and spacelike vectors in .
A space-time is said to be time-orientable if admits a continuous Lorentzian cone field , which is generated by Lorentz metric .
In the remained of this subsection, we assume that space-time is time-orientable. Thus, is a conal manifold with respect to a continuous Lorentzian cone field . In such situation, the non-spacelike vectors, which belong to cone field , 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 “” (resp., “”) on 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 corresponds to an event. And a signal can be sent from to if there is a (future directed) causal curve from to . Thus, closed causal curves generate paradoxes involving causality (i.e., violate causality). As a result, we assume that the space-time is causal, i.e., contain no closed non-spacelike (conal) curves (see e.g., [2, 12, 13, 43]).
For a given point , the chronological future , chronological past , causal future , and causal past of are defined as follows:
-
•
; ;
-
•
; .
Remark A.2.
In some articles, the set would be called forward set or reachable set from and be written as ; the set would be called backward set or controllable set from and be written as .
Remark A.3.
For any , are open by Proposition 2.1.
is said to be inner continuous at if each compact set , there exists a neighborhood of such that for each . is said to be outer continuous at if each compact set , there exists a neighborhood of such that for each . The inner and outer continuity of are defined dually. See [13].
Proposition A.1.
For any , are inner continuous.
Proof.
Suppose is compact. For any , i.e., , there is a strictly conal curve such that and . Let , then , i.e., and . Since is open, then is an open neighborhood of . Thus, is an open covering of . Since is a compact set, we choose to determine a finite subcovering. On the other hand, implies that , . So, is an open neighborhood of . For any , , . Then . For any , since and , then , i.e., . So, for . Thus, . So, we have proved that is inner continuous.
A similar argument is used for . ∎
Proposition A.2.
-
, implies ,
-
, implies .
Remark A.4.
Proposition A.3.
.
Theorem A.1.
For any , if are closed and are outer continuous, then the conal order “” is quasi-closed.
Proof.
If for all and and as , we just need to prove that by Proposition A.3. Suppose that , i.e., . Then there exists a compact set containing such that . Since as , there is a such that for all . On the other hand, is outer continuous at , then there exists a neighborhood of such that for any , . Since as , there is a such that and for all . Let . If , then and with , which is a contradiction. Thus, . ∎
Proposition A.4.
The following conditions are equivalent.
-
(A)
For all and in , if and only if ;
-
(B)
For all and in , if and only if .
Proof.
Proposition A.5.
If any one of the equivalent conditions in Proposition A.4 holds, then are outer continuous for any .
Proof.
We first assert that for , there exists such that . Before proving this assertion, we will show how it implies this Proposition. Let be compact. This assertion implies that is an open covering of . Choose a finite subcovering determined by and . Then is a neighborhood of such that for any , . In fact, since , then , . So, . For any , for . Then for . Thus, for . So, . Then is outer continuous.
It remains to prove the assertion. Suppose that for any , then we will get a contradiction. Since , then by Proposition A.3. Since condition of Proposition A.4 holds, then for all . On the other hand, for any , i.e., , there is a strictly conal curve such that and . Let , then . Thus, and . Since for all , then . So, for any . Then . By Proposition A.3, . Thus, by Proposition A.4, which is a contradiction with . ∎
Remark A.5.
Lemma A.1.
If are closed for all , then are outer continuous.
Proof.
Corollary A.1.
If are closed for all , then the conal order “” is quasi-closed.
A.2 Globally orderable homogeneous spaces
Let be a connected Lie group and be a smooth manifold. A left action of Lie group on manifold is a smooth map satisfying and for all , , where is the identity element in . We write or for . The action is said to be transitive if for every pair of points , there exists such that . For each , the isotropy group of , denoted by , is the set .
A smooth manifold endowed with a transitive smooth action by a Lie group is called a homogeneous G-space (or a homogeneous space).
Let be a Lie group and be a closed subgroup. The left coset space is a smooth manifold of dimension (), and the left action of on is given by . Hence, is a homogeneous space (see [27, Theorem 21.17]).
Let be a Lie group and be a homogeneous G-space. If is any point of , then the isotropy group is a cloed subgroup of , and the defined by 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 , where is a Lie group and is a closed subgroup of .
Let be a Lie group and fix . Define the left translation map by . The left translation map is diffeomorphism since it is smooth with smooth inverse. The inverse of is clearly the map . The diffeomorphism induces a vector space isomorphism , where is the Lie algebra of and is the identity element in .
Let be a homogeneous space and the natural projection , be a submersion. For each , define the left translation by . Then the left translations are related to the left translations on the Lie group by for each . Let be the Lie algebra of and . The differential is a vector space homomorphism with , we obtain that , where is the set of cosets for (see [16, P488] or [1, P71]).
A wedge 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 on a manifold assigns to each point a wedge in the tangent space .
Let be any left group action on such that each of the maps defined by forms a diffeomorphism of , Then a wedge field is said to be G-invariant or homogeneous if for all and .
Lemma A.2.
Let be a closed subgroup of a Lie group and a wedge in such that (i) , and (ii) , where is the adjoint representation of . Define and by
where , is the identity element in , is the base point in , , and is the convex cone in obtained as the projection of onto . Then, is an invariant wedge field on and is a well-defined homogeneous or G-invariant cone field on . Moreover, for each ,
where is the canonical projection .
Proof.
See [16, Lemma VI.1.5]. ∎
Proposition A.6.
The homogeneous cone field on a homogeneous space is continuous and admits sections.
Proof.
See [26, Proposition 4.6]. ∎
Theorem A.2.
Let be a homogeneous cone field on as described in Lemma A.2. If , then and is globally orderable with respect to if and only if , where .
Proof.
Let be a homogeneous space with base-point . A Riemannian metric , , is said to be G-invariant or homogeneous, if it satisfies for each and .
Proposition A.7.
Let be a Lie group, a closed subgroup, then the space is complete in any G-invariant metric.
Proof.
See [14, p148]. ∎
A homogeneous space is called reductive if there exists a subspace of such that and for all . Hence, (see [1]). The next Proposition gives a simple description of G-invariant Riemannian metrics on a homogeneous space.
Proposition A.8.
Let be a reductive homogeneous space. Then there is a one-to-one correspondence between G-invariant Riemannian metrics on and -invariant inner products on ; that is, for all , .
Proof.
See [1, Proposition 5.1]. The homomorphism such that is the isotropy representation of the homogeneous space . ∎
A.3 Differential positivity in flat spaces
Let be the -dimensional Eucliean space and is a closed convex cone of . There is a partial order “” on generated by cone ( if and only if ). 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 defined by . Such cone field is said to be a contant cone field. Thus, we can define the order “” on with respect to cone field ( if and only if there exists a conal curve such that , and ).
It is easy to see that the contant cone field on Eucliean space satisfies the smoothness conditions and invariance condition in the previous sections.
The next proposition implies that the order “” generated by contant cone feild agrees with the partial order “” on generated by cone (see [26, Proposition 1.10]).
Proposition A.9.
Let be a -dimensional Eucliean space and is a convex cone in such that forms a constant cone field . Then for , if and only if .
Proof.
If , i.e., , then we choose a curve , where . Thus, , , and . So, is a conal curve and .
If , then there is a conal curve such that , and . For any , where is the dual cone of , . On the other hand, . Since , then . So, . Thus, we obtain that , i.e., . ∎
And a similar result can be obtained for the order “” with “”, where if and only if .
Let be a dynamical system in with the flow . The system is said to be monotone with respect to partial order “” if whenever and and stongly monotone if whenever , and (see [22] and [50]).
The following Proposition shows that in , a monotone system is differentially positive (see [9, Theorem 1]).
Proposition A.10.
Let be the -dimensional Eucliead space and is a convex cone in such that forms a constant cone field . Then a system is monotone with respect to partial order “” generated by cone if and only if this system is differentially positive with respect to cone field .
Proof.
By Proposition A.9, order “” and order “” are equivalent.
Suppose that the system is differentially positive. For , there exists a conal curve such that , and . Then, we obtain that is also a conal curve for each . In fact, . Since is differentially positive and , then . So, for . And hence, is monotone with respect to “”.
If the system is monotone. For and , then there is a conal curve , , such that , and with . Since is monotone, then for each , , i.e., . Thus, . So, we obtain that . 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, 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 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 -smooth strongly monotone discrete dynamical systems, J. Differ. Equ. 269(2020), 9804-9818.
- [56] Y. Wang and J. Yao, Sharpened dynamics alternative and its -robustness for strongly monotone discrete dynamical systems, J. Funct. Anal. 283(2022), 109538.