• 沒有找到結果。

Melnikov function and Transversality of Manifolds

3.3 Proof of the Existence Results

3.3.1 Melnikov function and Transversality of Manifolds

To prove the main results in the last subsection , we first detect the transversality of invariant manifolds for connecting orbits by investigating the Melnikov function.

First, we recall the results for the formula of Melnikov function, [5, 20, 27, 34].

Lemma 3.8 Consider the plane system:

y0= R0(y) + ¯εR1(y, ¯ε), (3.23)

where ¯ε≥ 0 and R0, R1 ∈ Cr with r≥ 2. Suppose that y01 and y20 are two different hyperbolic saddle points of (3.23)|ε=0¯ , and there exists a heteroclinic orbit y0(t) of (3.23)|¯ε=0connecting from y10 to y02. Then the Melnikov function of (3.23) is

M(y0) = Z

−∞

eR0tσ(s)dsD(t)dt (3.24)

where σ(t) = tr∂R∂y0(y0(t)) and D(t) = R0(y0(t))∧ R1(y0(t), 0).

According to formula (3.24), we can compute the Melnikov function of system (3.17) in the fol-lowing.

Lemma 3.9 Suppose, for some c0 and w0, system (3.17) has a heteroclinic orbit Γ := r(t; w0) = (u0(t; w0), v0(t; w0)).

(1) For fixed w = w0 and varying c, the Melnikov function with respect to the heteroclinic orbit Γ is given by

M(c0) = Z

−∞

e−c0tv02(t; w0)dt.

In particular,M(c0)6= 0.

(2) For fixed c = c0 and varying w, the Melnikov function is given by M(w0) = a0(w0)

Z

−∞

e−c0tv0(t; w0)u0(t; w0)(1− u0(t; w0))dt. (3.25)

Proof. (1) For fixed w = w0, let F (u, v; c) be the vector field of system (3.17), i.e., F (u, v; c) =

Base on the results of Lemma 3.9, we now compute the first order Melnikov function of system (3.17) when w varies in different parameter regions.

Lemma 3.10 Assume a0(w0) 6= 0, then M(w0) 6= 0 for any c0 ∈ (0, 1/√

2c0∈ (1, 2). According to Remark 2.1, the heteroclinic orbit (u0(t; w0), v0(t; w0)) can be represented explicitly in the following:

M(w0)

where B(x, y) and Γ(x) are the Beta function and the Gamma function respectively. Note that B(x, y) = Γ(x)Γ(y)/Γ(x + y).

(3) If w0∈ H3(c0) then a(w0) = 2 +√

2c0∈ (2, 3). Similar to the proof of part (2), the heteroclinic orbit (u0(t; w0), v0(t; w0)) satisfies

2c0 ∈ (−2, −1) and the heteroclinic orbit (u0(t; w0), v0(t; w0)) satisfies

where−1 < γ < 0. Then we have

M(w0) = a0(w0)a(w0)(1− a(w0))2Γ(1 + γ)Γ(2− γ)/Γ(4) < 0.

The proof is complete.

However, if a0(w0) = 0 in Lemma 3.10 then M(w0) = 0, and there give no information about the transversality of the invariant manifolds. Therefore we need to compute the higher order term of Melnikov function to detect the transversality of the invariant manifolds. The following we only investigate the second-order term of Melnikov functionM2(w0).

Lemma 3.11 Suppose, for some small c0 and w0, a0(w0) = 0 and system (3.17) has a heteroclinic

Proof. Without lost of generality, we may assume w0= 0. For such fixed w near w0, let’s write the system (3.17) in the following vector form

d

As w = 0, system (3.17) has a heteroclinic orbit r(t; w) connecting two equilibria, E01 and E02. Let L be a line segment transversal to r(t; 0) at r(0; 0). For sufficiently small w, there exists a unique bounded solution ru(t; w) for t≤ 0 such that ru(t; w) in the unstable manifold of one equilibrium Ew1 and ru(0; w)∈ L. For t ≤ 0, let’s define

zu(t) := ∂

∂wru(t; w)|w=0, 4u(t) := zu(t)∧ R0(ru(t; 0)),

2

Differentiating equation (3.26) with respect to w, we have

Similarly, there exists a unique bounded solution rs(t; w) for t≥ 0 such that rs(t; w) in the unstable manifold of the other equilibrium Ew2 and rs(0; w)∈ L. For t ≥ 0, let’s define

By the proof of Lemma 3.10 and Lemma 3.11, we have the following corollary.

Corollary 3.12 Let’s assume the same assumptions as stated in Lemma 3.11. If a00(w0)6= 0 then M2(w0)6= 0 for any c0∈ (0, 1/√

2) and w0∈ Hi(c0), i = 1.· · · 6.

For more higher order terms of Melnikov function, the computation is similar but more complicated.

The following we only illustrate the general result, and skip the proof.

Lemma 3.13 Suppose, for some small c0 and w0, a(i)(w0) = 0 for all 1≤ i < k, where k is a given positive integer, and system (3.17) has a heteroclinic orbit Γ0:= (u0(t; w0), v0(t; w0)). For fixed c = c0 and varying parameter w, the kth order Melnikov function is given by

Mk(w0) = a(k)(w0) Z

−∞

e−c0tv0(t; w0)u0(t; w0)(1− u0(t; w0))dt.

As a consequence of previous lemmas and corollary, we have the following conclusions for the transver-sality of invariant manifolds.

Lemma 3.14 Let M t N be in the sense that manifolds M and N intersect transversally, and Cδ(c) := (c− δ, c + δ).

(1) Consider system (3.17) with c = c0∈ (0, 1/√

2). The transversality of various invariant mani-folds of (3.17) along Γ(w) are illustrated in Table 2.

Region of w transversality of manifolds along Γ(w) w∈ H1(c0) W0u(M0δ(w)) t W0c(M1δ(w)) w∈ H2(c0) W0u(M1δ(w)) t W0c(M0δ(w)) w∈ H3(c0) W0u(Maδ(w)) t W0c(M0δ(w)) w∈ H4(c0) W0u(M1δ(w)) t W0c(Maδ(w)) w∈ H5(c0) W0u(Maδ(w)) t W0c(M0δ(w)) w∈ H6(c0) W0u(Maδ(w)) t W0c(M1δ(w)) w∈ G1(c0) W0cu(0, 0, w) t W0c(M1δ(w)) w∈ G2(c0) W0cu(1, 0, w) t W0c(M0δ(w)) w∈ G3(c0) W0cu(0, 0, w) t W0c(Maδ(w)) w∈ G4(c0) W0cu(1, 0, w) t W0c(Maδ(w))

Table 2: Transversalities of manifolds W0c(M0,1,aδ (w)), W0u(M0,1,aδ (w)), W0cu(0, 0, w) and W0cu(1, 0, w).

(2) Consider (2.10) with c = c0 ∈ (0, 1/√

2). The transversality of various invariant manifolds of

Region of w transversality of manifolds along Γ(w)× {c0} w∈ H1(c0) W0u(M0δ(w))× {c0} t W0c(M1δ(w)× Cδ(c0)) w∈ H2(c0) W0u(M1δ(w))× {c0} t W0c(M0δ(w)× Cδ(c0)) w∈ H3(c0) W0u(M0δ(w)× {c0}) t W0c(Maδ(w)× Cδ(c0)) w∈ H4(c0) W0u(M1δ(w))× {c0} t W0c(Maδ(w)× Cδ(c0)) w∈ H5(c0) W0u(Maδ(w)× {c0}) t W0c(M0δ(w)× Cδ(c0)) w∈ H6(c0) W0u(Maδ(w))× {c0} t W0c(M1δ(w)× Cδ(c0)) w∈ G1(c0) W0cu(0, 0, w, c0) t W0c(M1δ(w)× Cδ(c0)) w∈ G2(c0) W0cu(1, 0, w, c0) t W0c(M0δ(w)× Cδ(c0)) w∈ G3(c0) W0cu(0, 0, w, c0) t W0c(Maδ(w)× Cδ(c0)) w∈ G4(c0) W0cu(1, 0, w, c0) t W0c(Maδ(w)× Cδ(c0))

Table 3: Transversalities of manifolds W0c(M0,1,aδ (w)× Cδ(c0)), W0u(M0,1,aδ (w))× {c0}, W0cu(0, 0, w, c0) and W0cu(1, 0, w, c0).

Now we begin the proof of the main theorems.

Proof of Theorem 3.5.

We only prove the first part of the theorem. The proof for the second part of the theorem is the same as the proof of Theorem 2.3 of [33] and omitted.

Assume w = (w1, ..., wn) is s-admissible with respect to c, where s = (s1, ..., sn+1). We first claim that the singular local orbit 0−−−→ 1w=0 −−−−→ sw=w1 2 is weakly shadowed by a true local orbit.

According to Table 1 and the assumption a(0) = a1(c), we know that 0∈ H1(c) and there exists a singular local orbit 0−→ 1 at w = 0. For the following local path 1 → s2, the admissible conditions lead to s2= 0 or a. Therefore, there exists a singular local orbit 1−→ s2at w = w1if w1∈ H2(c)∪ G2(c) or H4(c)∪ G4(c) (in fact, w1∈ H2(c) or H4(c)). Take M0= W0u((0, 0, 0)× Cδ(c)). According to Lemma 3.14, we have

M0t W0c(M1δ(0)× Cδ(c)).

Let N0 be their intersection. Since the phase space of system (2.10) is R4and dimensions of M0and W0c(M1δ(0)×Cδ(c)) are 2 and 3 respectively, then dimN0= 2 + 3−4 = 1. We now apply the exchange lemma 3.2 to the vicinity of the slow manifold M1× Cδ(c) along the slow orbit from (1, 0, 0, c) to (1, 0, w1, c). Taking Mε = Wεu((0, 0, 0)× Cδ(c)) be such that Mε → M0 as ε → 0. By condition (A1) and part (1) of Lemma 3.2, a portion Mpε

0 of Mεwill proceed near the singular orbit and leave

the vicinity of slow orbit close to Wεu(M1δ(w1)× Cδ(c)). Note that dimW0u(M1δ(w1)× Cδ(c)) = 3.

Thus, the singular orbit 0−−−→ 1w=0 −−−−→ sw=w1 2 is weakly shadowed by a true orbit.

Next, we claim that the singular local orbit 0−−−→ 1w=0 −−−−→ sw=w1 2−−−−→ sw=w2 3 is weakly shadowed by a true local orbit. Two cases for s2 = 0 or a are considered. For the case s2 = 0, a portion Mpε

1 of Wεu((a(w1), 0, w1)× Cδ(c)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((s3, 0)× (w2− δ, w2+ δ)× Cδ(c)) or Wεcu((s3, 0)× {w2} × {c}). The details of proof can be found in [33] by using the Exchange Lemma with turning point (Lemma 3.2) and omitted. For the case s2= a, the admissible conditions imply s3= 1 or 0. Thus there exists a singular local orbit s2−→ s3 at w = w2 if w2∈ H5(c) or H6(c). From the admissible condition (A4), there is no turning point lying between w1 and w2. It’s also easy to see that W0u((1, 0)× (w1− δ, w1+ δ)× (Cδ(c)) and W0c((a(w1), 0)× (w1− δ, w1+ δ)× Cδ(c)) intersect transversally. Then, by Lemma 3.3, a portion Mpε1 of Wεu((a(w1), 0)× (w1− δ, w1+ δ)× Cδ(c)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((s3, 0)× (w2− δ, w2+ δ)× Cδ(c)). Note that dimW0u(M1δ(w1)× Cδ(c)) =dimW0cu((s3, 0)× {w2} × Cδ(c)) = 3.

According to the above discussions, we can conclude inductively that the singular local orbit 0−−−→ 1w=0 −−−−→ · · ·w=w1 −−−−−−→ sw=wn−1 nis weakly shadowed by a true local orbit. Generally, for 2 < i < (n+1), we consider the following two cases.

(1) Assume there exists a turning point lying between wi−1 and wi. By Lemma 3.2, a portion Mpεi−1 of Wεu((si, 0, wi−1)× Cδ(c)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((si+1, 0)× (wi− δ, wi+ δ)× Cδ(c)) or Wεcu((si+1, 0, wi)× Cδ(c)).

(2) Assume there is no turning point lying between wi−1and wi. By Lemma 3.3, a portion Mpεi−1 of Wεu((si, 0, wi−1)× Cδ(c)) will leave the vicinity of slow orbit close to Wεu((si+1, 0)× (wi− δ, wi+ δ)× Cδ(c)).

Furthermore, we have dimW0u((si+1, 0)× (wi− δ, wi+ δ)× Cδ(c)) = 3 and dimW0cu((si+1, 0, wi)× Cδ(c)) = 3.

Finally, we prove that the true orbits obtained by above arguments are C1 O(ε)-close to the unstable manifold W0u((sn+1, 0, wn)× Cδ(c)). Since sn+1 = 1, the admissible conditions lead to wn ∈ H1(c)∪ G1(c) or H6(c). Thus, there exists a singular local orbit sn −→ 1 at w = wn. By condition (A1), we have a(w) < 1 for all w∈ [wn, 1] and the singular orbit will approach to (1, 0, 1) as time goes infinity. Moreover, W0u((sn, 0, wn)× {c}) intersects W0c(M1δ(wn)× Cδ(c)) transversally.

As a result, the true orbit will approach a neighborhood of (1, 0, 1, c), near the singular orbit and C1

O(ε)-close to the unstable manifold W0u((sn+1, 0, wn)×Cδ(c)). The proof of the theorem is complete.

Proof of Theorem 3.6.

The proof of Theorem 3.6 is similar to that of Theorem 3.5. Let w = (w1, ..., wn) is s-admissible for c = c and s = (s1, ..., sn+1). We first claim that the singular local orbit 0 −−−→ 1w=0 −−−−→ sw=w1 2 is strongly shadowed by a true orbit.

According to Table 1 and the assumption a(0) ≤ a6(c), we know that 0 ∈ G1(c) and there exists a singular local orbit 0 −→ 1 at w = 0. For the following local path 1 → s2, the admissible conditions lead to s2 = 0 or a. Therefore, there exists a singular local orbit 1 −→ s2 at w = w1 if w1 ∈ H2(c)∪ G2(c) or H4(c)∪ G4(c) (in fact, w1∈ H2(c) or H4(c)). Take M0= W0u(0, 0, 0).

According to Lemma 3.14, we have

M0t W0c{M1δ(0)}.

Let N0 be their intersection. Since the phase space of system (2.3) is R3 and both dimensions of M0 and W0c(M1δ(0)) are 2, then dimN0 = 2 + 2− 3 = 1. We now apply the exchange lemma 3.2 to the vicinity of the slow manifold M1 along the slow orbit from (1, 0, 0) to (1, 0, w1). Taking Mε= Wεu((0, 0, 0)) be such that Mε→ M0 as ε→ 0. By condition (A1) and part (1) of Lemma 3.2, a portion Mpε0 of Mε will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu(M1δ(w1)). Note that dimW0u(M1δ(w1)) = 2. Thus, the singular orbit 0 −−−→ 1w=0 −−−−→ sw=w1 2 is strongly shadowed by a true orbit.

Next, we claim that the singular local orbit 0−−−→ 1w=0 −−−−→ sw=w1 2−−−−→ sw=w2 3 is strongly shadowed by a true local orbit. Two cases for s2 = 0 or a are considered. For the case s2 = 0, a portion Mpε1 of Wεu((a(w1), 0, w1)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((s3, 0)×(w2−δ, w2+ δ)) or Wεcu((s3, 0)×{w2}). The details of proof can be found in [33] by using the exchange lemma with turning point (Lemma 3.2) and omitted. For the case s2= a, the admissible conditions imply s3= 1 or 0. Thus there exists a singular local orbit s2−→ s3at w = w2if w2∈ H5(c) or H6(c). From the admissible condition (A4), there is no turning point lying between w1 and w2. It’s also easy to see that W0u((1, 0)× (w1− δ, w1+ δ)) and W0c((a(w1), 0)× (w1− δ, w1+ δ)) intersect transversally. Then, by Lemma 3.3, a portion Mpε1 of Wεu((a(w1), 0)× (w1− δ, w1+ δ)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((s3, 0)× (w2− δ, w2+ δ)). Note that dimW0u(M1δ(w1)) =dimW0cu((s3, 0)× {w2}) = 2.

According to the above discussions, we can conclude inductively that the singular local orbit 0−−−→w=0 1−−−−→ · · ·w=w1 −−−−−−→ sw=wn−1 n is strongly shadowed by a true local orbit. Generally, for 2 < i < (n + 1), we consider the following two cases.

(1) Assume there exists a turning point lying between wi−1 and wi. By Lemma 3.2, a portion Mpεi−1 of Wεu((si, 0, wi−1)) will proceed near the singular orbit and leave the vicinity of slow orbit close to Wεu((si+1, 0)× (wi− δ, wi+ δ)) or Wεcu((si+1, 0, wi)).

(2) Assume there is no turning point lying between wi−1and wi. By Lemma 3.3, a portion Mpε

i−1

of Wεu((si, 0, wi−1)) will leave the vicinity of slow orbit close to Wεu((si+1, 0)× (wi− δ, wi+ δ)).

Furthermore, we have dimW0u((si+1, 0)× (wi− δ, wi+ δ)) = dimW0cu((si+1, 0, wi)) = 2.

Finally, we prove that the true orbits obtained by above arguments are C1O(ε)-close to the unsta-ble manifold W0u((sn+1, 0, wn)). Since sn+1= 1, the admissible conditions lead to wn ∈ H1(c)∪G1(c) or H6(c). Thus, there exists a singular local orbit sn −→ 1 at w = wn. By condition (A1), we have a(w) < 1 for all w ∈ [wn, 1] and the singular orbit will approach to (1, 0, 1) as time goes infinity.

Moreover, W0u((sn, 0, wn)) intersects W0c(M1δ(wn)) transversally. As a result, the true orbit will ap-proach a neighborhood of (1, 0, 1), near the singular orbit and C1O(ε)-close to the unstable manifold W0u((sn+1, 0, wn)). The proof of the theorem is complete.

The results of Theorem 3.7 can also be proved in the same way and omitted.

4 Wazewski Theorem and Lasalle Invariance Principle

4.1 Wazewski theorem

We now recall a variant of the Wazewski Theorem which is a formalization and extension of the shooting method in higher dimension (see [9], p. 563). Let usconsider the differential equation:

y0(s) = f (y(s)) (4.27)

where f : Rn → Rn is a Lipschitz continuous function. Suppose y(s; y0) is the unique solution of (4.27) with initial value y(0) = y0. For convenience, set y(s; y0) = y0· s and let Y · S be the set of points y· s, where y ∈ Y ⊂ Rn and s∈ S ⊂ R. Given W ⊆ Rn, define the immediate exit set W of W by

W={y0∈ W : ∀s > 0, y0· [0, s) * W }.

Given Σ⊆ W , let

Σ0={y0∈ Σ : ∃s0> 0 such that y0· s0∈ W }./ For y0∈ Σ, define the exit time T (y0) of y0by

T (y0) = sup{s : y0· [0, s] ⊂ W }.

Note that y0· T (y0)∈ W and T (y0) = 0 if and only if y0∈ W. The Wazewkski Theorem is then stated as the following.

Theorem 4.1 Consider (4.27) and suppose that

(i) if y0∈ Σ and y0· [0, s] ⊆ c`(W ), then y0· [0, s] ⊆ W ;

(ii) if y0∈ Σ, y0· s ∈ W , y0· s /∈ W, then there is an open set Vsabout y0· s disjoint from W; (iii) Σ = Σ0, Σ is a compact set and intersects a trajectory of y0 = f (y) only once.

Then the mapping F (y0) = y0·T (y0) is a homeomorphism from Σ to its image on W. A set W ⊆ Rn satisfying the conditions (i) and (ii) is called a Wazewski set

相關文件