´ ON THE ESTIMATION OF POINCARE MAPS OF THREE-DIMENSIONAL VECTOR FIELDS NEAR A HYPERBOLIC CRITICAL POINT By Yuting Zou A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Mathematics 2011 ABSTRACT ´ ON THE ESTIMATION OF POINCARE MAPS OF THREE-DIMENSIONAL VECTOR FIELDS NEAR A HYPERBOLIC CRITICAL POINT By Yuting Zou We study the estimation of Poincar´ maps of three-dimensional vector fields e near a hyperbolic critical point, which involves linearization problems. Standard linearization theorems have several defects in applications. They usually require complicated non-resonance conditions on the eigenvalues of the vector field at the critical point. Even when one has these non-resonance conditions, as one gets close to a resonance, the size of the neighborhood where the C 1 linearization exists typically gets too small for practical uses. We seek for a linearization theorem that overcomes these shortcomings and may have broad practical applications. We have proved a partial linearization theorem that gives a C 1 linearization h near a hyperbolic critical point p on a two-dimensional invariant surface Σ of a three-dimensional vector field X. Let the eigenvalues of DX(p) be a, b and c, where a > 0 > b > c. Essentially our theorem only requires that 2b > c to obtain h in some neighborhood U of p in Σ. In addition, the explicit size of U is found, which depends on the C 2 information of X, as well as the C 0 and C 1 sizes of h. Based on our partial linearization theorem, we obtain desired estimation of Poincar´ maps from e some transversal curve to the stable manifold of p to another transversal surface to the unstable manifold of p. Our estimation of such Poincar´ maps will have many applications, including an e in-depth study of the famous Lorenz equations. For example, it seems likely that we will be able to substantially improve results of Tucker on the existence of the Lorenz strange attractor, and obtain rigorous results on the existence of chaos near the first homoclinic bifurcation as numerically investigated in the well-known book of Colin Sparrow. ACKNOWLEDGMENT I would like to express my deepest and sincerest gratitude to my advisor, Dr. Sheldon Newhouse, for his exceptional support and unfailing patience. Without his guidance, this thesis would not have been possible. It is a valuable opportunity for me to learn from such a great mathematician and a wonderful person. He has a very broad perspective, sharp and quick mathematical mind, good sense of humor, and a variety of interests besides mathematics. More than I could express my appreciation, he is a strikingly philosophical, inspiring and open-minded person. Not only have I learned abundantly from him, but also his unique personality has a life-long influence on me. I am very thankful for his help that has pulled me through the most challenging times of my doctoral program. It has been an enormously enjoyable experience to work with him. My gratitude goes to my defense committee members, Dr. Martin Berz, Dr. Kyoko Makino, Dr. Yang Wang and Dr. T.Y. Li, for their expertise and precious time. Special thanks to Dr. Martin Berz, Dr. Kyoko Makino and Dr. Alexander Wittig for their intense interests in our work, which gives me a strong encouragement to carry on my work to its finest details. Their tool, COSY VI-integrator, that made verified numerical integration possible, is an important source of inspiration for this thesis. Finally I would like to thank Dr. Zhangfang Zhou for his unceasing care and support from the beginning of my doctoral program to the end. He is one of the kindest and most generous people I have ever met. Whenever I have difficulties, he iv whole-heartedly offers his helps. It is hard to imagine completing my Ph.D. track without his presence. v TABLE OF CONTENTS List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . vii 1 Introduction 1 2 Linearization theorems 6 3 A quantitative partial linearization theorem in R3 3.1 A sketch of the proof . . . . . . . . . . . . . . . . . . . . . . . 3.2 Linear estimates in dimension two . . . . . . . . . . . . . . . . 3.3 Conjugacy restricted to the stable manifold . . . . . . . . . . . 3.4 Local C 1 linearization for two dimensional diffeomorphisms . 3.5 Three dimensional vector fields . . . . . . . . . . . . . . . . . 3.6 Straightening Invariant manifolds of a three dimensional map . . . . . . . . . . . . . . . . . . . Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . vi . . . . . . 12 15 17 18 25 52 89 97 LIST OF FIGURES 3.1 Intervals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.2 D0 . For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation 31 3.3 F 1 (y) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.4 F 2 (y) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 vii Chapter 1 Introduction Let M be a smooth (C ∞ ) three dimensional Riemannian manifold with associated topological metric d, and let X be a C 2 vector field on M with a hyperbolic critical point p0 ∈ M with real distinct eigenvalues a, b, c satisfying a > 0 > b > c. (1.1) Assume that X is forward complete in the sense that, if φ(t, x) denotes the local flow of X, then, φ(t, x) is defined for all t ≥ 0 and all x ∈ M. Let W s (p0 ), W ss (p0 ), W u (p0 ) denote, respectively, the stable, strong stable, and unstable manifolds of p0 . These are defined as follows. W s (p0 ) = {x ∈ M : φ(t, x) exists for all t ≥ 0 and lim d(φ(t, x), p0 ) = 0} t→∞ 1 W ss (p0 ) = {x ∈ M : φ(t, x) exists for all t ≥ 0 1 and lim sup log d(φ(t, x), p0 ) < b} t→∞ t W u (p0 ) = {x ∈ M : φ(t, x) exists for all t ≤ 0 and lim d(φ(t, x), p0 ) = 0}. t→−∞ It is well-known that W s (p0 ), W ss (p0 ) and W u (p0 ) are C 2 injectively immersed copies of the plane and line, respectively, in M. Let us use the notation Comp(E, x) for the connected component of the set E containing the point x. For a point p ∈ W s (p0 ) and a neighborhood U of p in M, let W s (p, U) = Comp(W s (p0 ) U, p). One can choose arbitrarily small neighborhoods U so that W s (p, U) is a topological 2-disk and U \ W s (p, U) consists of a disjoint pair of open 3-balls. Let p ∈ W s (p0 ) \ {p0 }, q ∈ W u (p0 ) \ {p0 } and let N be a smoothly embedded 2-disk in M, such that N is transversal to W u (p0 ) at q. An application of the Grobman-Hartman theorem (C 0 local linearization theorem) shows that there is a small neighborhood U of p in M and a connected component U0 of U \ W s (p, U) such that there is a well-defined Poincar´ map P from U0 into N. This map simply e takes a point in U0 to the first point on its positive orbit which hits N. In many applications (e.g. Section 2.2 in [7], [9], and the books [6], [1]) one encounters the following problem. Let p ∈ W s (p0 ) \ W ss (p0 ) and let γ be a C 2 embedded arc in U meeting 2 W s (p0 ) transversely at p and assume that U is small enough so that for each y ∈ γ, d(y, p) is boundedly related to the arc length of the subarc of γ joining p to y. Let ′ U0 , and, for a point y ∈ γ0 , let γ0 (y) denote the unit tangent vector to ′ γ0 at y. One seeks an estimate of the derivative P ′ (y) = DPy (γ0 (y)). γ0 = γ If the vector field X is C 1 linearizable near p0 , then it is well-known and easy to prove that one can find constants 0 < C1 < C2 such that b C1 |P ′ (y)| ≤ d(y, p)| a |−1 ≤ C2 |P ′ (y)|. (1.2) In most applications using estimates of the type (1.2), one merely assumes that X is, in fact, at least C 1 linearizable near p0 . In view of the Sternberg linearization theorem, a generic C ∞ vector field is C ∞ linearizable near a hyperbolic critical point, so it does not seem to be a big assumption to assume linearity. However, it should be noted that there are several defects to the application of the Sternberg theorem even to get C 1 linearizations. Among these defects are the following. • The theorem requires so-called non-resonance conditions on the eigenvalues a, b, c; • Even in the case that X is C ∞ , as one gets close to a resonance, the size of the neighborhood on which the C 1 linearization exists typically gets very small; • The amount of smoothness required on X depends on the eigenvalues a, b, c. c Our main result shows that, with the mild additional assumption that b > 2 , we obtain the estimate (1.2) above for general C 2 vector fields in dimension three. 3 Theorem 1.0.1. Let X, p0 , p, γ0 and P be as above. Also, as above, for y ∈ γ0 , let ′ P ′(y) = DPy (γ0 (y)) ′ be the derivative of P on the unit tangent vector γ0 (y) to γ0 at y. Assume that the eigenvalues a, b, c of X at the critical point p0 satisfy a > 0 > 2b > c. (1.3) Then, there are constants 0 < C1 < C2 such that b C1 |P ′ (y)| ≤ d(y, p)| a |−1 ≤ C2 |P ′ (y)|. (1.4) The main tools we use are modifications of techniques in so-called normal hyperbolicity theory to get a C 2 invariant surface Σ which contains both γ0 and p0 and is tangent at p0 to the sum of the eigenspaces of a, b. Since, by a theorem of P. Hartman [2], two dimensional C 2 vector fields with a hyperbolic saddle point p0 are C 1 linearizable near p0 , one can get the required estimate. Since we obtain a linearization of X restricted to an invariant two-dimensional surface through p0 and not in a full neighborhood of p0 , we will refer to our result as a partial linearization theorem. Remarks. • It should be noted that the general tools of normal hyperbolicity in the literature give some two dimensional surfaces tangent at p0 to the sum of the 4 eigenspaces of a, b. However, these surfaces are far from unique, and are usually constructed by means of globalization techniques. Since we need such surfaces which contain an a priori given curve γ0 , the globalization technique is not applicable as far as we know. The existence of C 2 invariant surfaces through p0 which contain an arbitrary given transverse curve γ0 seems to be a new result, and our proof avoids the use of globalizations. • For various applications, it is important to know the sizes of the constants C1 , C2 in (1.2). We take care to give explicit estimates for these in our work below. Using appropriate local coordinates near p0 and replacing p and q by points in their orbits, it suffices to prove our results for vector fields in R3 with a hyperbolic critical point at the origin. The precise statement needed is in Theorem 3.0.6. We first review some of the history of linearization theorems. 5 Chapter 2 Linearization theorems The first general linearization theorem was given by Poincar´ in his thesis. He e obtained a theorem that can be rephrased as follows: Theorem 2.0.2. Let dxi = λi xi + fi (x1 , . . . , xn ), 1 ≤ i ≤ n. dt (2.1) be a system of differential equations such that λi is a complex number, and each fi is a complex analytic function in a neighborhood of the origin which vanishes together with its first order partial derivatives at the origin. Suppose all the λi lie in the same open half-plane about the origin and n λi = mj λj (2.2) j=1 for any non-negative integers mj with 2 ≤ 6 k m . j=1 j Then, there exists a complex analytic change of coordinates yi = ψi (x1 , x2 , · · · , xn ) transforming (2.1) into the linear system dyi/dt = λi yi , 1 ≤ i ≤ n The conditions (2.2) are called non-resonance conditions and, without them, there are examples even of polynomials fi where the theorem fails. Let us recall the definitions of local linearizations in both the flow and diffeomorphism contexts. Let f : Rn → Rn be a C r diffeomorphism, r ≥ 1, from a neighborhood U of 0 in Rn into Rn such that f (0) = 0, Df (0) = L. Then, L is, of course, a linear automorphism of Rn . A local C k linearization of f near 0 is a C k homeomorphism h from a neighborhood V of 0 in U to another neighborhood V ′ of 0 such that Lh(x) = hf (x) for x ∈ V f −1 (V ) . Next, consider a C r local flow near 0 in Rn with a fixed point (or critical point) at 0. This is a C r map φ(t, x) from a neighborhood (−ǫ, ǫ) × U into Rn where ǫ is a positive real number such that 7 1. φ(0, x) = x for all x ∈ U, 2. φ(t, 0) = 0 for all t ∈ (−ǫ, ǫ), and 3. φ(t + s, x) = φ(t, φ(s, x)) for t, s ∈ (−ǫ, ǫ) such that t + s ∈ (−ǫ, ǫ). We often use φ or φt to denote the local flow where φt (x) = φ(t, x). Sometimes we reverse the t and x and write φ(x, t). A C r vector field X near 0 in Rn is a C r map X : U0 → Rn from a neighborhood U0 of 0. If X is a C r vector field near 0 such that X(0) = 0 (with r ≤ 1), then, for every T > 0 there is a neighborhood U of 0 in Rn and a local flow φ(t, x) = φX (t, x) defined on (−T, T ) × U such that ∂φ(t, x) = X(φ(t, x)) ∂t for all t ∈ (−T, T ) and x ∈ U. The local flow φ(t, x) is called the flow of X near 0. The local flow φ(t, x) is called linear if, in addition to the usual flow properties, we have x → φ(t, x) is a linear automorphism of Rn for each t. Let φt and ψt , |t| < ǫ be two local flows near 0 with fixed points at 0. Let k ≥ 0 be a positive integer. A local C k conjugacy near 0 between φt and ψt is a C k homeomorphism h from a neighborhood V of 0 to another neighborhood V ′ of 0 such that, there is a positive real number ǫ1 such that 8 1. ψt is defined on V ′ for all |t| ≤ ǫ1 , 2. h ◦ φt is defined for all |t| ≤ ǫ1 , and 3. ψt ◦ h = h ◦ φt for all |t| ≤ ǫ1 . If ψt is linear, then a local C k conjugacy between φt and ψt near 0 is called a local linearization of φt near 0. If X is a C 1 local vector field near 0 with associated local flow φt and derivative L = DX(0), then a local C k linearization of X near 0 is a local C k conjugacy between φt and the linear flow given by L. In this case, we may assume that the local flow φt is defined for |t| ≤ T where T is arbitrary. For large T , we simply have to shrink the neighborhood on which the linearization is defined. We are now in a position to state the well-known Grobman-Hartman Theorem [6], [3]. Theorem 2.0.3. (Grobman-Hartman). 1. Let f be a C 1 local diffeomorphism of Rn with a hyperbolic fixed point at 0 (i.e., the eigenvalues of Df (0) have norm different from 1). Then, f has a local C 0 linearization near 0. 2. Let X be a C 1 local vector field near 0 with a hyperbolic critical point at 0 (i.e. X(0) = 0 and the eigenvalues of DX(0) have non-zero real parts). Let L = DX(0) and let ψt be the flow of L. Then, for any T > 0, there is a local C 0 conjugacy h between φt and ψt for all |t| ≤ T . 9 This is a beautiful and powerful theorem. One could ask for smooth local linearizations for smoother f and/or X. In this direction there is another well-known theorem due to Sternberg. See the appendix starting on page 256 in [3] for a proof of this theorem as well as related results, and further references. Theorem 2.0.4. (Sternberg) 1. Let f be a C ∞ local diffeomorphism of Rn with a hyperbolic fixed point at 0 (i.e., the eigenvalues of Df (0) have norm different from 1). Let λ1 , λ2 , . . . , λn be the eigenvalues of Df (0). Suppose that, for every positive 1 ≤ i ≤ n,and any n-tuple of non-negative n m , we have j=1 j integers (m1 , m2 , . . . , mn ) with 2 ≤ n λi = j=1 mj λj . (2.3) Then, f has a local C ∞ linearization near 0. 2. Let X be a C ∞ local vector field near 0 with a hyperbolic critical point at 0 (i.e. X(0) = 0 and the eigenvalues of DX(0) have non-zero real parts). Let L = DX(0) and let ψt be the flow of L. Suppose that, for every positive 1 ≤ i ≤ n,and any n-tuple of non-negative n m , we have j=1 j integers (m1 , m2 , . . . , mn ) with 2 ≤ n λi = mj λj . j=1 10 (2.4) Then, for any T > 0, there is a local C ∞ conjugacy h between φt and ψt for all |t| ≤ T . Again the conditions (2.3) and (2.4) are called non-resonance conditions. Without them, even a real-analytic system may not be C 1 linearizable. Sternberg’s theorem is very important, and has many applications. However, as we mentioned in our introduction, it does have some shortcomings. Finally, we mention the two dimensional C 1 linearization theorems of Hartman [2] which is of major relevance to this thesis. Here, fortunately, there are no nonresonance conditions to worry about. Theorem 2.0.5. 1. Let f be a C 2 local diffeomorphism of R2 with a hyperbolic saddle fixed point at 0 (i.e., a pair of real eigenvalues λ1 , λ2 of Df (0) satisfy 0 < |λ2 | < 1 < |λ1 |). Then, f has a local C 1 linearization near 0. 2. Let X be a C 2 local vector field near 0 in R2 with a hyperbolic saddle critical point at 0 (i.e. X(0) = 0 and the eigenvalues λ1 , λ2 of L = DX(0) satisfy λ2 < 0 < λ1 ). Let ψt be the flow of L. Then, for any T > 0, there is a local C 1 conjugacy h between φt and ψt for all |t| ≤ T . As previously mentioned, we will give a new proof of Theorem 2.0.5 with explicit estimates for the C 0 and C 1 sizes of the linearization. 11 Chapter 3 A quantitative partial linearization theorem in R3 Let x = (x1 , x2 , x3 ) ∈ R3 , and X(x) be a forward complete C 2 vector field which vanishes at the origin and can be written as ¯ ¯ ¯ X(x) = (X1 (x), X2 (x), X3 (x)) = (ax1 + X1 (x), bx2 + X2 (x), cx3 + X3 (x)) ¯ where c < 2b < 0 < a, and Xi,x (0, 0, 0) = 0 for i, j = 1, 2, 3. j For small positive ε0 , let B1 = B1ε = {x ∈ R3 ||x2 | ≤ ε0 , |x1 | ≤ ε0 , |x3 | ≤ ε0 } 0 and let | · | = supx∈B | · | denote the maximum norm of various quantities in 1 ¯ B1 . Then, there is a finite constant M > 0, such that |Xi,x ,x (x)| ≤ M for j k 12 i, j, k = 1, 2, 3, and, using the mean value theorem, we have some small constants ¯ ¯ ǫ1 > 0 and ǫ2 > 0, such that |Xi,x (x)| ≤ M · ε0 = ǫ2 , and |Xi (x)| ≤ ǫ2 · ε0 = ǫ1 j for i, j = 1, 2, 3. Let B1top = {x ∈ B1 |x2 = ε0 }, B1bot = {x ∈ B1 |x2 = −ε0 }, + B1 = {x ∈ B1 |x1 = ε0 }, and − B1 = {x ∈ B1 |x1 = −ε0 }. and let ϕ(x, t) = ϕt (x) be the local flow of X, so we may write ϕ(x, t) = (ϕ1 , ϕ2 , ϕ3 ) = (eat x1 + ϕ1 , ebt x2 + ϕ2 , ect x3 + ϕ3 ) ¯ ¯ ¯ where ϕi for i = 1, 2, 3 are the higher order terms. Set ψt (x) to be the flow of the ¯ corresponding linear vector field of X, i.e, ψt (x) = (eat x1 , ebt x2 , ect x3 ) 13 Let us denote the origin (0, 0, 0) by 0, and let W s (0) denote its stable manifold. s s Let Wloc = Wloc (0) = Comp(W s (0) of the W s (0) B1 , 0) denote the connected component B1 containing 0. In the following we choose B1 small enough so that the indicated statements are valid. s Let Γ = Wloc B1top . Then Γ divides B1top into two connected open sets each of whose closures is a topological 2-disk. By the Grobman-Hartman theorem, − + if x ∈ B1top \ Γ, then there exists τφ (x) > 0, such that ϕ(x, τφ (x)) ∈ B1 ∪ B1 , − + and for 0 ≤ s < τ (x), ϕ(x, s) is not in B1 ∪ B1 . Let + + B1top = {x ∈ B1top |ϕ(x, τφ (x)) ∈ B1 }, and − − B1top = {x ∈ B1top |ϕ(x, τφ (x)) ∈ B1 }. + − Then, the map P (x) = ϕ(x, τφ (x)) is called a Poincar´ map from B1top ∪ B1top \ Γ e − + to the sides B1 ∪ B1 . Recall that we use the maximum norm |(v1 , v2 , v3 )| = max(|v1 |, |v2 |, |v3 |) on vectors v = (v1 , v2 , v3 ) ∈ R3 and its induced norm on various other quantities. For instance, the arclength of curves, lengths of tangent vectors, and norms of linear 14 maps are all defined relative to this maximum norm. Consider a C 2 curve γ in B1top which is transverse to Γ at the point p ∈ Γ. ± Let γ ± = γ ∩ B1top . Assume that γ is parametrized by arclength and, for y ∈ γ, let ℓy denote the arclength of the subarc of γ from p to y. Assume that there are constants A2 > A1 > 0 such that A1 ℓy ≤ |y − p| ≤ A2 ℓy . Let vy be the unit tangent vector to γ ± at the point y ∈ γ ± , and let P ′ (y) = DPy (vy ). The main result needed to prove Theorem 1.0.1 is then the following Theorem 3.0.6. There are constants C2 > C1 > 0 such that b C1 |P ′ (y)| ≤ |y − p|| a |−1 ≤ C2 |P ′(y)|. 3.1 A sketch of the proof The proof of Theorem 3.0.6, which we call the quantitative partial linearization theorem, is divided into several parts, shown step by step in the subsequent sections. s u First, we straighten Wloc (0) and Wloc (0) and then obtain a C 2 surface Σ in B1 which meets B1top in the curve γ and is invariant under the flow φ = φ(x, t) on orbit pieces which stay in B1 . This allows us to transform the problem to an analogous problem for a flow η = η(x, t) in R2 with a hyperbolic critical point at (0, 0) with a 15 transverse curve γ through a point p in the stable manifold of η. Now, a well-known ˜ ˜ theorem of Hartman [2] gives a C 1 linearization of η in some neighborhood of (0, 0). Using this linearization, we can complete the proof of the Theorem 3.0.6. General comment. We emphasize that, while many of the general ideas in our construction have been available in the dynamics literature, we believe that our results are new in at least two ways. • the construction of the invariant C 2 surface Σ containing an arbitrary transverse curve γ. • the detailed estimates of the C 1 sizes in our linearization. Comment on the two dimensional local linearization. Following Sternberg [8] to linearize a flow, one first linearizes the time T map for some T > 0 and uses the so-called integral technique. Thus the problem reduces to finding a local C 1 linearization near a hyperbolic fixed point for a C 2 planar diffeomorphism f . This theorem is also proved in Hartman [3]. A new proof for the diffeomorphism case was given by Palis and Takens in [5]. Their method consists of first C 1 linearizing along the stable and unstable manifolds, and then contructing two transverse C 1 foliations containing these manifolds which are used to define the coordinates of the required C 1 linearization. Their proof relies on a clever use of the stable manifold theorem for the derivative map Df on the tangent bundle, and, hence involves a four dimensional local diffeomorphism. While their proof can be modified to work in our case, it seems somewhat complicated to obtain estimates of the size of the domain of the linearization and 16 its C 1 size from that approach. Since these types of estimates are important for applications, we proceed to construct our own proof of the Hartman theorem. The only thing that we borrow from Palis and Takens is the idea of transverse foliations. Our construction of the foliations and the ensuing estimates are new. 3.2 Linear estimates in dimension two Here let us derive some properties of a linear flow in R2 : ψt (x) = (ψ1 , ψ2 ) = (eat x1 , ebt x2 ) where a > 0 > b. Here in R2 , we only consider B1 = {x = (x1 , x2 ) ∈ R2 |0 ≤ x1 ≤ ε0 , 0 ≤ x2 ≤ ε0 }, with B1top = {(x1 , ε0 ) : 0 ≤ x1 ≤ ζε0 , 0 < ζ < 1}, B1r = {x ∈ B1 |x1 = ε0 }. Since other quadrants can be done similarly, we do not show them here. Let the point p = (0, ε0 ), as is stated above, for any initial condition x = (x1 , ε0 ) ∈ B1top \ p, there is a finite time τ (x) > 0, such that ψ(x, τ (x)) = (ε0 , y2 ) ∈ B1r . Let us name such map Pl (x1 , ε0 ) = (ε0 , y2 ), where x1 = 0. We want to find the size of DPl , so we proceed as follows. We have Pl (x1 , ε0 ) = (ε0 , y2 ) = (eat x1 , ebt ε0 ), where x1 = 0. By comparing the first component, we have ε 1 et = ( 0 ) a . x1 17 Plugging to the second component, we have ε b y2 = ε0 ( 0 ) a . x1 So we can write ε b Pl (x1 , ε0 ) = (ε0 , ε0 ( 0 ) a ). x1 Since −ε0 ε b b ε b b ε b ∂(ε0 ( 0 ) a )/∂x1 = ε0 · · ( 0 ) a −1 = − ( 0 ) a +1 x1 a x1 a x1 (x1 )2 we have     0  0 0  0 = DPl =  .  ∂y  b b ( ε0 ) a +1 0 2 0 −a x ∂x1 1 So 3.3     0 1   DPl ·   =  . b +1  b ( ε0 ) a 0 −a x 1 Conjugacy restricted to the stable manifold In this section, we will adapt the proof of Sternberg’s linearization theorem [1956] into the following quantitative version, with explicit conditions on the neighborhood and sizes of the linearization worked out. ′ Theorem 3.3.1. Let f : R → R be a C 2 function such that f (0) = 0 and f (0) = a ′′ where |a| ∈ (0, 1). If in Bε0 (0) = {y ∈ R : |y| ≤ ε0 }, we have |f | ≤ M and β = |a| + Mε0 /2 < 1, then there is a unique C 1 map u : Bε0 → R, such that 18 u′ (0) = 0, and if h(y) = (id + u)(y) then h(f (y)) = ah(y). Moreover, |h| ≤ ε0 exp( C ) 1−β and ′ 2C |h | ≤ exp( ) 1−β where C = Mε0 /(2|a|). Proof: Using Taylor’s theorem, in a small neighborhood of the origin Bε , there is a real number θy ∈ (0, y) such that ′′ f (θy ) 2 y . f (y) = ay + 2 For k ≥ 0, let fk (y) = f ◦ f ◦ · · · ◦ f (y) k so that k fk (y) = y fi (y)/fi−1 (y) i=1 and k fk (y) =y fi (y)/(afi−1 (y)). ak i=1 Since ′′ f (θi ) fi (y) = f (fi−1 (y)) = afi−1 (y) + (fi−1 (y))2 2 19 (3.1) we have ′′ f (θi ) f (y), fi (y)/(afi−1 (y)) = 1 + 2a i−1 and, hence, fk (y) =y ak k k fi (y)/(afi−1 (y)) = y i=1 i=1 ′′ f (θi ) [1 + f (y)]. 2a i−1 Let us define a function h(y) by f (y) h(y) = lim k . k→∞ ak Assuming this, notice that it provides a linearization since f (y) f (y) f (f (y)) h(f (y)) = lim k = lim k+1 = lim a k+1 = ah(y). k→∞ k→∞ ak+1 k→∞ ak ak Let us proceed to prove that h(y) is well defined; i.e., that the limit exists. We have f (y) h(y) = lim k = lim y k→∞ ak k→∞ k i=1 ′′ f (θi ) [1 + f (y)] 2a i−1 (3.2) so, it suffices to show that ∞ i=1 ′′ f (θi ) f (y) |< ∞. | 2a i−1 20 (3.3) Let |·| = sup | · |. y∈Bε0 Now, ′′ ′′ f (θ) f (θ) 2 y = (a + y)y f (y) = ay + 2 2 ′′ and, since |f | ≤ M and |a| ∈ (0, 1), we have |f (y)| ≤ (|a| + M ε )|y|. 2 0 By induction, for i ≥ 1 we then get M M |fi−1 (y)| ≤ (|a| + ε0 )i−1 |y| ≤ (|a| + ε0 )i−1 ε0 . 2 2 Hence ′′ f (θi ) M M fi−1 (y)| ≤ (|a| + ε0 )i−1 ε0 = Cβ i−1 | 2a 2|a| 2 (3.4) ε M where the constants C = 0 and β = |a| + M ε0 ∈ (0, 1) are those given in the 2 2|a| theorem. The above inequality implies that the infinite product in (3.2) converges uniformly, and thus, the linearization h(y) exists. As the uniform limit of continuous functions is continuous, we conclude that h(y) is continuous. Using log(1 + x) ≤ x for x > 0, we have, for any positive real numbers pn 21 ∞ ∞ (1 + pn ) = exp(log( n=1 ∞ ∞ (1 + pn ))) = exp( n=1 ≤ exp( log(1 + pn )) n=1 pn ). n=1 Using that f (y) h(y) = lim k k→∞ ak where fk (y) =y ak k k fi (y)/(afi−1 (y)) = y i=1 i=1 ′′ f (θi ) [1 + f (y)] 2a i−1 we get  ′′ ∞  f (θi ) | h(y) | = | y  [1 + f (y)] | 2a i−1 i=1 ∞ C ≤ ε0 exp( Cβ i−1 ) ≤ ε0 exp( ). 1−β i=1 Further, for each y ∈ (0, ε0 ) there is a τi ∈ (0, ε0 ) such that f ′ (y) = a + f ′′ (τi )y. Since, for each k > 0 we have (with a possibly different τi ) 22 ′ fk (y) = ak k k a + f ′′ (τ )f f ′ (fi−1 (y)) i i−1 (y) ) ( )= ( a a i=1 i=1 k f ′′ (τi ) = (1 + fi−1 (y)) a i=1 (3.5) and, as above, | f ′′ (τi ) fi−1 (y)) |≤ 2C β i−1 a we get h′ (y) exists and f ′ (y) h′ (y) = lim k . k ak In addition, with C as above, we get 2C | h′ (y) |≤ exp( ). 1−β Notice that, from the last expression in (3.5), we have that ′ fk (y) =1 ak for y = 0. Thus, h(y) = y + u(u) where u is a C 1 function such that u′ (0) = 0. By now we have shown the existence of a C 1 linearization h described in the theorem. It remains to show such h is unique. Let L(y) = ay. Suppose there is 23 another function g : Bε0 → R of the same properties as h; i.e., g −1 f g = L, g(0) = 0, and g ′ (0) = 1. Then f = gLg −1 and so h−1 gLg −1 h(y) = ay. Let ρ = h−1 g, we have ρLρ−1 = L, ρ(0) = 0, and ρ′ (0) = 1. So we can write ρ(y) = y + v(y) with v(y) ∈ C 1 , v(0) = 0 and v ′ (0) = 0. Then let us show for such ρ, we have ρ(y) = y for all y ∈ Bε0 , i.e., v(y) = 0 for all y ∈ Bε0 . By ρLρ−1 (y) = ay, we have ρL(y) = aρ(y), which is ay + v(ay) = a(y + v(y)) so v(ay) = av(y), for all y ∈ Bε0 . Taking derivatives, we get av ′ (y) = av ′ (ay), i.e., v ′ (y) = v ′ (ay) for all y ∈ Bε0 . We claim that v ′ (y) = 0 for all y ∈ Bε0 . Suppose there is z ∈ Bε0 such that v ′ (z) = 0, then v ′ (z) = v ′ (az) = 0. Since |a| < 1, ai z ∈ Bε0 for integer i ≥ 0. We have v ′ (an z) = · · · = v ′ (az) = v ′ (z) = 0. But limn→∞ an z = 0, by the continuity of v ′ we have v ′ (z) = v ′ (0) = 0 which is a contradiction. So the claim is proved, which says v ′ (y) = 0 for all y ∈ Bε0 , i.e, v(y) is a constant function. Since v(0) = 0, we have v(y) = 0, giving ρ(y) = y or thus h = g. The uniqueness is proved, and so the proof of Theorem 3.3.1 is complete. Now consider a local C 2 flow φt , |t| ≤ 1, on R, with a hyperbolic critical point at the origin, i.e., if f = φ1 , then f satisfies the assumptions of Theorem 3.3.1. Thus, ˜ ˜ f (x) = ax + f (x), where 0 < a < 1, x ∈ R, and f (x) stands for the higher order terms. Using the uniqueness of h, which is derived from a standard technique as in 24 [ref Newhouse Notes] show that h also linearizes {φt }|t|≤1 . That is, if ψt (x) = eat x, then hφt (x) = ψt h(x) for x ∈ Bε0 . 3.4 Local C 1 linearization for two dimensional diffeomorphisms Remark: It is well-known that linearizations for flows follow from consideration of associated time−1 maps. Hence we restrict here to the case of diffeomorphisms (see p. 817 in [8]). In this section, we consider the C 2 diffeomorphism f : R2 → R2 with a hyperbolic fixed point at (0, 0). We assume the local unstable and stable manifolds of f , denoted by W u (0), W s (0), are contained in the coordinate lines {(x1 , x2 ) : x2 = 0}, {(x1 , x2 ) : x1 = 0}, respectively. Let ε0 > 0 and consider the square Bε0 = {(x1 , x2 ) : |x1 | ≤ ε0 , |x2 | ≤ ε0 }. For x = (x1 , x2 ) ∈ Bε0 , we assume that ˜ ˜ f (x1 , x2 ) = (λ1 x1 + f1 (x1 , x2 ), λ2 x2 + f2 (x1 , x2 )) = (f1 , f2 ) ˜ where fi , i = 1, 2, are the nonlinear terms, and λ1 > 1, 0 < λ2 < 1. 25 (3.6) Let us write the Jacobian matrix of Dfx as   Ax Bx  Dfx =  . C x Dx In addition, we assume that there are a small ǫ > 0 and a constant K > 0 such that ˜ ˜ fi (0, 0) = 0 and the first partial derivatives fix (0, 0) = 0, for i, j = 1, 2 j (3.7) |B| ≤ ǫ, |C| ≤ ǫ, 0 < λ2 − ǫ ≤ D ≤ λ2 + ǫ < 1, 1 < λ1 − ǫ ≤ A ≤ λ1 + ǫ, (3.8) and, |dA−1 | ≤ K, |dB| ≤ K, |dC| ≤ K, |dD| ≤ K. (3.9) Using the previous section, suppose we have found C 1 linearizations hs : U → R ′ of f |W s (0) and hu : U → R of f |W u (0) . Then in this section, we will proceed to linearize f : R2 → R2 in some neighborhood B1 of W s (0) W u (0) in Bε0 . This will be done via a modification of ideas in [[5]]. We will construct the neighborhood B1 and C 1 submersions πu : B1 → W s (0) and πs : B1 → W u (0) which commute with f , and, together with the linearizations hs and hu restricted to the local stable 26 and unstable manifolds, will give a linearization h in B1 as h(q) = (h1 , h2 ) = (hu πs (q), hs πu (q)) = q ¯ −1 −1 q q h−1 (¯) = πu (hs )−1 (¯y ) ∩ πs (hu )−1 (¯x ). q The inverse images of the submersions πs , πu will define two transverse folations which are invariant by f . These foliations, after adjustments using hu , hs , become the local coordinate curves of the linearization h. We actually only deal with the construction of πu , leaving the analogous construction of πs to the reader. Thus, we prove the following theorem. Theorem 3.4.1. Let f : R2 → R2 be a C 2 diffeomorphism satisfying (3.6)-(3.9) above. Define α= λ + 2ǫ M(λ2 + 2ǫ) + ǫ 1 , M= 1 , µ= . (λ1 − ǫ)(λ2 − ǫ) λ1 − ǫ ǫ2 ) α(1 − α (3.10) Assume that ǫ is small enough so that λ2 + 3ǫ < 1 and µ < 1. λ1 − ǫ Then, there is a neighborhood B1 of W s (0) sion πu : B1 → W s (0) such that 27 (3.11) W u (0) in Bε0 and a C 1 submer- πu f = f πu on B1 f −1 (B1 ). (3.12) Moreover, one has the estimates |πu (x1 , x2 ) − x2 | ≤ ε0 , (3.13) and ε F ∂πu F | ≤ ε0 (1 + ) · exp( 0 ) ∂x1 1−µ 1−µ ε ·F ε ·F ∂πu − 1| ≤ ( 0 ) exp( 0 ) | ∂x2 1−µ 1−µ | where F = M · K( 3 + λ2 + 3ǫ). λ1 − ǫ Let us start proving the theorem. We write f (x1 , x2 ) = (f1 (x1 , x2 ), f2 (x1 , x2 )). − + Consider the sets Bε and Bε defined by 0 0 + Bε = {(x1 , x2 ) ∈ Bε0 |x2 ≥ 0}, 0 (3.14) − Bε = {(x1 , x2 ) ∈ Bε0 |x2 ≤ 0} 0 (3.15) 28 + − We will construct πu on Bε and leave the analogous construction on Bε to 0 0 the reader. We choose 0 < y0 < ε0 , and a small δ > 0, such that + f −1 (0, y0 + δ) ∈ Bε . 0 If we write p = (0, y0 ), and p± = (0, y0 ± δ), then the above condition becomes + f −1 (p+ ) ∈ Bε . 0 We write I1 to be the interval between p− and p+ on the y-axis, and similarly write I2 to be the interval between f (p− ) and f (p+ ), and I0 to be the interval between f (p− ) and p+ (see figure 3.1). In addition, we use I−1 to denote the set of the second coordinates of f −1 (0 × I0 ) (see figure 3.1). Then we use l0 to represent the + horizontal line segment that passes through p inside Bε . And we take D0 as the 0 + region bounded above and below by l0 and f (l0 ), which stays inside Bε , including 0 its boundaries (see figure 3.2). ∞ (f n (D ) ∩ B + ) that ε0 0 n=0 contains p. Now let us proceed to define a C 2 horizontal foliation F u over the base We define B1 to be the connected component of I0 , which can then be extended to the whole B1 by the iteration of f . First, we may choose a smooth bump function ρ(x2 ) over I0 such that, ρ(x2 ) = 1 over I1 ; ρ(x2 ) = 0 over I2 ; and over I0 \ (I1 ∪ I2 ), ρ(x2 ) is monotonically increasing w.r.t. x2 with 0 < ρ(x2 ) < 1. 29 Figure 3.1: Intervals Second, we define the first horizontal foliation F 1 over the base I0 ∪ I−1 (see figure 3.3). For every given x2 ∈ I0 ∪ I−1 , define the leaf through (0, x2 ) as the set F 1 (x2 ) = {(x1 , x2 )| |x1 | ≤ ε0 } which can also be rewritten as F 1 (x2 ) = {(x1 , η1 (x1 )| |x1 | ≤ ε0 } where η1 (x1 ) = x2 for every given x2 ∈ I0 ∪ I−1 , i.e., η1 is a constant function. Next, let us define our second horizontal foliation F 2 (see figure 3.4) over the 30 Figure 3.2: D0 . For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation + base I0 . We write f −1 (0, x2 ) = (0, x2 ) ∈ Bε . And for every given x2 ∈ I0 , define ˜ 0 the leaf through (0, x2 ) to be + F 2 (x2 ) = f (F 1 (˜2 )) ∩ Bε . x 0 We may also write it as F 2 (x2 ) = {(x1 , η2 (x1 )| |x1 | ≤ ε0 } where η2 (0) = x2 . Finally, our desired horizontal foliation F u over I0 is a combination of F 1 and F 2 via the bump function ρ(x2 ). As for every given x2 ∈ I0 , we define the leaf of 31 Figure 3.3: F 1 (y) F u through (0, x2 ) to be: F u (x2 ) = {(x1 , ρ(x2 )η1 (x1 ) + (1 − ρ(x2 ))η2 (x1 )) : |x1 | ≤ ε0 }. Observe that, by our definition of ρ(x2 ), F u (x2 ) agrees with F 1 (x2 ) for x2 ∈ I1 , and agrees with F 2 (x2 ) for x2 ∈ I2 . For every given x2 , the tangent vectors along the leaf F 1 (x2 ) equals 0; and the tangent vectors along the leaf F 2 (x2 ) can be written as        1  f1x1 f1x2  1 f1x1  rescaled   = −− − − f − − −→      2x1  f2x f2x 0 f2x f1x 1 2 1 1 where we know f1x = 0. So the tangent vectors to the leaf F u (x2 ) can be repre1 32 Figure 3.4: F 2 (y) sented as       1   1   1 = ρ(x2 )   + (1 − ρ(x2 ))  f f2x  .  2x1   1 (1 − ρ(x2 )) 0 f1x f1x 1 1 (3.16) The computation above also implies that our leafs of F u (x2 ) for different x2 ’s do not intersect, i.e, the tangent vector at every point to a given leaf is uniquely defined. Moreover, in the next paragraph we show the slope of each tangent vector is bounded by 1, which implies our foliation F u (x2 ) is well defined. Let     1  Xu,1   Xu =  =   f2x  . 1 (1 − ρ(x2 )) Xu,2 f1x 1 We will need the following estimate, which says the absolute value of the slope of 33 each tangent vector above is less than 1. Recall our assumption in (3.8), we know f1x ≥ λ1 − ǫ > 1 and |f2x | ≤ ǫ, so we have 1 1 |Xu,2 | = |(1 − ρ(x2 )) |Xu,1 | f2x 1| f1x 1 f2x 1| f1x 1 ǫ ≤ < 1. λ1 − ǫ ≤ | We now extend Xu to B1 . For each x ∈ B1 \ B1bot , there is a least integer n(x), such that f −n(x) (x) ∈ D0 . Let n(x) Xu (x) = Df · Xu (f −n(x) (x)). −n(x) (x) f A λ-lemma (inclination lemma) argument shows that the absolute value of the slope |Xu,2 (x)/Xu,1 (x)| of Xu (x) remains bounded by 1 for all x ∈ B1 and converges to 0 as x approaches B1bot . Let Px = Xu,2 (x) Xu,1 (x) be this slope. For each x we define   1 Yu =   Px for x ∈ B1 \ B1bot , and   1 Yu =   0 34 (3.17) for x ∈ B1bot . This Px is a continuous function from B1 to R, such that |Px | ≤ 1 for all x ∈ B1 . Our main result of this section is the following. Theorem 3.4.2. The vector field Yu constructed above is C 1 in B1 . Equivalently, the function x → Px is C 1 in B1 . As we indicated above the integral curves of Yu will give our desired foliation F u . The proof that Yu is C 1 , or equivalently, the function x → Px is C 1 , will require several steps. The proof will be obtained by finding a sequence g1 , g2 , · · · , of C 1 functions, such that 1. gn converges to P uniformly on B1 , 2. Dgn converges uniformly on B1 . Following this, it is standard that P is C 1 and D P = limn→∞ Dgn . We will use the technique of the Fiber Contraction Theorem familiar from the Invariant Manifold Theory as in [4]. This involves definitions of certain function spaces and associated mappings. Let us proceed to define our first function space G. Consider the space C 0 (B1 , R) of bounded continuous functions from the box B1 to R with supremum norm supx∈B | · | that makes it a Banach space. Let 1 35 0 C1 (B1 , R) be the closed subset of C 0 (B1 , R) with each of its element bounded by 0 1. Then C1 (B1 , R) is a complete metric space with the metric induced by this norm. Let 0 ˜ G = {P |P ∈ C1 (B1 , R), P (x) = P (x) f or x ∈ D0 ∪ B1bot }. G is also a complete metric space equipped with the same norm supx∈B | · |. 1 The first step is to show that Px is the unique fixed point of a contraction map Γ on G, which we now define. Let f = φX,T , by construction we have     1  c  Dfy ·   =   Py c · Px (3.18) where y = f −1 (x) and c is a scalar. Using the notation at the beginning of this section, we write        Ay By   1   c   · = . C y Dy Py c · Px By solving the above equations for P , we have Ay + By Py = c C y + Dy P y = c · P x . So Px = C y + Dy P y Ay + By Py 36 (3.19) which is equivalent to Ay Px + Py By Px = Cy + Py Dy . So we have Px = A−1 Cy + A−1 Dy Py − A−1 By Py Px . y y y Given P ∈ G, we define Γ(P ) as follows. For x ∈ D0 ∪ B1bot , set Γ(P )x = Px . For x ∈ B1 \ (D0 ∪ B1bot ), let y = f −1 (x), and Γ(P )x = A−1 Cy + A−1 Dy Py − A−1 By Px Py . y y y (3.20) Clearly, Γ(P )x is continuous in x. To show that Γ(P ) ∈ G, it suffices to show that sup |Γ(P )x | ≤ 1. x∈B1 For each x ∈ B1 and y = f −1 (x), we have |Γ(P )x| ≤ |A−1 Cy | + |A−1 Dy Py | + |A−1 By Px Py | y y y ≤ λ +ǫ ǫ ǫ + 2 + . λ1 − ǫ λ1 − ǫ λ1 − ǫ 37 If the size of B1 is properly chosen, we will have λ2 + 3ǫ < 1. λ1 − ǫ So |Γ(P )| < 1 which means Γ maps G into itself. Moreover, the Lipschitz constant λ of Γ can be computed as follows |Γ(P1 )x − Γ(P2 )x | ≤ |A−1 Cy + A−1 Dy P1 (y) − A−1 By P1 (x)P1 (y) y y y −A−1 Cy − A−1 Dy P2 (y) + A−1 By P2 (x)P2 (y)| y y y ≤ |A−1 Dy ||P1 (y) − P2 (y)| y +|A−1 By ||P1 (x)P1 (y) − P2 (x)P2 (y)| y where |P1 (x)P1 (y) − P2 (x)P2 (y)| ≤ |P1 (x)P1 (y) − P1 (x)P2 (y) +P1 (x)P2 (y) − P2 (x)P2 (y)| ≤ |P1 (x)||P1 (y) − P2 (y)| + |P2 (y)||P1(x) − P2 (x)|. Since |P1 | ≤ 1 and |P2 | ≤ 1, |P1 (x)P1 (y) − P2 (x)P2 (y)| ≤ 2|P1 − P2 | 38 we have |Γ(P1 ) − Γ(P2 )| ≤ (|A−1 D| + 2|A−1 B|)|P1 − P2 | Let λ + 3ǫ λ +ǫ ǫ +2 = 2 λ = |A−1 D| + 2|A−1 B| ≤ 2 λ1 − ǫ λ1 − ǫ λ1 − ǫ If in B1 , we have λ < 1, then Γ is a contraction mapping with contraction constant λ. That means, for any initial choice of P ∈ G, Γn (P ) converges uniformly to a unique fixed point. Since Γ(P ) = P , we know P is the unique attracting fixed point. So we may choose our converging sequence as follows. To get our initial P0 , we may use a smooth bump function ρ : R2 → R such that ρ(x) = 1 on a neighborhood of D0 and ρ(x) = 0 on a neighborhood of B1bot . Then we define a C 1 function g0 , such that g0 |B ∈ G: 1 ˜ ˜ g0 = ρ(x)P + (1 − ρ(x)) · 0 = ρ(x)P ˜ for all points in a neighborhood U0 of B1 . Let P0 = g0 |B , we see that P0 |D = P , 1 0 P0 | B = 0 and is C 1 on B1 . Now for y = f −1 (x), let 1bot g1 (x) = Γ(g0 ) = A−1 Cy + A−1 Dy g0 (y) − A−1 By g0 (x)g0 (y) y y y which is C 1 on a neighborhood of B1 . Inductively, set gn (x) = Γ(gn−1 ) = A−1 Cy + A−1 Dy gn−1 (y) − A−1 By gn−1 (x)gn−1 (y) y y y 39 This gives us a sequence of C 1 functions gn with the following properties: 1. gn is defined and is C 1 on a neighborhood Un of B1 , 2. gn |B ∈ G, 1 3. Γ(gn ) converges uniformly to P on B1 . Let Pn = gn |B for n ≥ 0, we have Γn (P0 ) = Pn , and we can find the C 0 size 1 of P , that is |P | ≤ λn |P − P0 | + |Pn |. 1−λ 1 (3.21) Also it can be shown that |P | ≤ λ |P − P0 |. 1−λ 1 We wish to show that DPn converges uniformly on B1 . For this purpose we will use the Fiber Contraction Theorem on a suitable space G × H which we proceed to define. First of all, let us briefly review some definitions and the Fiber Contraction Theorem. Definition 3.4.3. Let (G, dG ) and (H, dH ) be metric spaces. The map Λ : G ×H → G × H of the form Λ(P, Q) = (Γ(P ), Ψ(P, Q)) is called a bundle map on G × H over the base Γ : G → G with principle part Ψ : G × H → H. Definition 3.4.4. The bundle map is called a fiber contraction on G × H if there is a k ∈ [0, 1), such that, for every P ∈ G, the map Q → Ψ(P, Q) is a contraction mapping with contraction constant k. 40 Theorem 3.4.5. [Fiber contraction theorem ] Let (G, dG ) and (H, dH ) be metric spaces and Λ a continuous fiber contraction on G × H over the base Γ : G → G with principle part Ψ : G × H → H. If P and Q are unique attracting fixed points of Γ and Q → Ψ(P , Q) respectively, then (P , Q) is a unique attracting fixed point of Λ. Now let us define the set of candidates of DP where P is C 1 , and then construct an operator which has a unique fixed point as DP in this set. Let H = C 0 (B1 , L(R2 , R)) be the space of bounded continuous functions from B1 to the linear maps from R2 to R. Let Q be any element in H, H is a Banach space equipped with the norm |Q| = sup |Qx · v|. 2 ,|v|=1 x∈B1 ,v∈R We now define an operator Q → Ψ(P, Q) for Q ∈ H, P ∈ G. For a C 1 function P , we have −1 dΓ(P )x = d(A−1 −1 (x) Cf −1 (x) ) + d(Af −1 (x) Df −1 (x) Pf −1 (x) ) f −d(A−1 B Px P −1 ). f (x) f −1 (x) f −1 (x) Let y = f −1 (x) as before, dΓ(P )x = {dA−1 df −1 (x)Cy + A−1 dCy df −1 (x) + d(A−1 Dy )df −1 (x)Py y y y +A−1 Dy dPy df −1 (x) − d(A−1 By )df −1 (x)(Px Py )} y y −(A−1 By )(dPx Py + Px dPy df −1 (x)). y 41 In this form, we define Ψ(P, Q)(x) = {dA−1 df −1 (x)Cy + A−1 dCy df −1 (x) y y (3.22) +d(A−1 Dy )df −1 (x)Py + A−1 Dy Qy df −1 (x) y y −d(A−1 By )df −1 (x)(Px Py )} y −(A−1 By )(Qx Py + Px Qy df −1 (x)). y Consider the bundle map Λ(P, Q) : G × H → G × H defined by Λ(P, Q) = (Γ(P ), Ψ(P, Q)). Observe that, since Γn (P0 ) = Pn , we have Λ(P0 , dP0 ) = (Γ(P0 ), Ψ(P0 , dP0 )) = (Γ(P0 ), dΓ(P0 )) = (P1 , dP1 ). Suppose Λn−1 (P0 , dP0 ) = (Pn−1 , dPn−1 ) Inductively, Λn (P0 , dP0 ) = (Γ(Pn−1 ), Ψ(Pn−1 , dPn−1 )) = (Γ(Pn−1 ), dΓ(Pn−1 )) = (Pn , dPn ). which converges uniformly if Λ is a fiber contraction, by the Fiber Contraction Theorem. And thus dPn converges uniformly. Now let us show that Λ is a fiber contraction. 42 First, we prove that for any given P ∈ G, the map Q → Ψ(P, Q) maps H to itself. It is easy to see that Ψ(P, Q), for a given P , is continuous with respect to x ∈ B1 , since f is C 2 . We will need to show Ψ(P, Q) is bounded. We proceed as follows |Ψ(P, Q)| ≤ {|dA−1 ||C| + |A−1 ||dC| + |d(A−1 D)||P | − |d(A−1 B)|P |2 }|df −1 | +|A−1 D · Q||df −1 | + |A−1 B||Q · P + P · Q · df −1 | ≤ {|dA−1 ||C| + |A−1 ||dC| + |d(A−1 D)| + |d(A−1 B)|}|df −1 | +|A−1 D||Q||df −1| + |A−1 B|(|Q| + |Q||df −1 |) ≤ {|dA−1 ||C| + |A−1 ||dC| + |d(A−1 D)| + |d(A−1 B)|}|df −1 | (3.23) +{|A−1 D||df −1| + |A−1 B|(1 + |df −1 |)}|Q| where we have used |P | ≤ 1. Now, we need to compute |df −1 | to proceed. For    A B  df =   C D we know df −1 =   1  D −B   . AD − BC −C A Recall the definitions of α and M from (3.10), we have α= 1 (λ1 − ǫ)(λ2 − ǫ) and 43 (3.24) M= λ1 + 2ǫ λ1 1 ≈ = , λ1 λ2 λ2 ǫ2 ) α(1 − α (3.25) We claim that |df −1 | ≤ M. (3.26) 2 Indeed, since |AD| > α and | BC | < ǫ , we have α AD |df −1 | ≤ |   1  D −B  ||  | AD − BC −C A 1 |(|A| + |C|) AD − BC 1 ≤ (|A| + |C|) |AD||1 − BC | AD 1 ≤ (λ1 + ǫ + ǫ) |AD||1 − BC | AD 1 (λ1 + 2ǫ) = M. ≤ 2 α(1 − ǫ ) α ≤ | In addition, from (3.9) we have {|dA−1 ||C| + |A−1 ||dC| + |d(A−1 D)| + |d(A−1 B)|}|df −1 | K K K + K(λ2 + ǫ) + + Kǫ + ) λ1 − ǫ λ1 − ǫ λ1 − ǫ 3 = M · K( + λ2 + 3ǫ). λ1 − ǫ ≤ M(ǫK + 44 To simplify the expression, let us write F = M · K( 3 + λ2 + 3ǫ). λ1 − ǫ (3.27) Now, going back to (3.23), we have λ +ǫ ǫ |Ψ(P, Q)| ≤ F + (M 2 + (1 + M))|Q|. λ1 − ǫ λ1 − ǫ Since Q ∈ H is also bounded, we have Ψ(P, Q) is bounded. So we have proved the map Q → Ψ(P, Q) maps H to itself. Next, we will show that the map Q → Ψ(P, Q) is a contraction mapping. For any given P ∈ G and any Q, Q ∈ H, we know Ψ(P, Q)x − Ψ(P, Q)x = A−1 Dy Qy df −1 (x) − A−1 Dy Qy df −1 (x) y y −(A−1 By ) · (Qx · Py + Px · Qy · df −1 (x)) y +(A−1 By ) · (Qx · Py + Px · Qy · df −1 (x)) y = A−1 Dy (Qy − Qy )df −1 (x) − A−1 By (Qx − Qy )Py y y −A−1 By Px (Qy − Qy )df −1 (x). y 45 Since |P | ≤ 1, we have λ +ǫ ǫ |Ψ(P, Q) − Ψ(P, Q)| ≤ M 2 |Q − Q| + |Q − Q| λ1 − ǫ λ1 − ǫ ǫ +M |Q − Q| λ1 − ǫ M(λ2 + 2ǫ) + ǫ |Q − Q|. = λ1 − ǫ µ Plugging in M ≈ 1 from (3.25), we have λ2 0<µ≈ 1 < 1. λ1 (3.28) Assuming the contraction µ, we have proved that, for any given P ∈ G, the map Q → Ψ(P, Q) is a contraction mapping. This implies that, for P ∈ G, the map Q → Ψ(P , Q) has a unique fixed point Q ∈ H. If we choose Λ(P, Q) = (Γ(P ), Ψ(P, Q)) as the bundle map on G × H. Then, by definition, Λ(P, Q) is a fiber contraction on G × H. Since it has been proved that P is the unique attracting fixed point of Γ, then by the Fiber Contraction Theorem, (P , Q) is the unique globally attracting fixed point of Λ. It remains to show that Q = dP , so that P ∈ C 1 . We may start with (P0 , dP0 ), where P0 ∈ C 1 . Let Γn (P0 ) = Pn , we have Λ(P0 , dP0 ) = (Γ(P0 ), Ψ(P0 , dP0 )) = (Γ(P0 ), dΓ(P0 )) = (P1 , dP1 ). Suppose Λn−1 (P0 , dP0 ) = (Pn−1 , dPn−1 ). 46 Then Λn (P0 , dP0 ) = (Γ(Pn−1 ), Ψ(Pn−1 , dPn−1 )) = (Γ(Pn−1 ), dΓ(Pn−1 )) = (Pn , dPn ). We know Pn converges to P uniformly, and dPn converges to Q uniformly. It is standard that Q = dP , and thus, P is C 1 . Now we proceed to find the size of dP . Since dΓ(P )x = {dA−1 df −1 (x)Cy + A−1 dCy df −1 (x) + d(A−1 Dy )df −1 (x)Py y y y −1 −d(A−1 By )df −1 (x)Px Py } + Ay Dy dPy df −1 (x) y −(A−1 By ) · (dPx · Py + Px · dPy · df −1 (x)). y In addition, we have |P | ≤ 1 and |df −1 | ≤ M, so we can write the form of the size of |dΓ(P )| for any P ∈ G as follows F |dΓ(P )| ≤ {|dA−1 ||C| + |A−1 ||dC| + |d(A−1 D)||P | + |d(A−1 B)||P |2 }M +M|A−1 D||dP | + |A−1 B|(|dP | + M|dP |) ≤ F + (M|A−1 D| + (1 + M)|A−1 B|)|dP | ≤ F + (τ1 τ2 M + (1 + M)τ1ε )|dP | 0 = F + µ|dP | 47 where 3 + λ2 + 3ǫ) ≤ M · K(4 + 3ǫ). λ1 − ǫ (3.29) = |dΓ(Pm−1 )| (3.30) ≤ F + µ|dPm−1 | ≤ F + µ(F + µ|dPm−2 |) = F + µF + µ2 |dPm−2 | F = M · K( For m > n we have |dPm| ··· m−n−1 ≤ F µi + µm−n |dPn|. i=0 So |dP | = F F + lim µm−n |dPn | = . lim |dPm | ≤ m→∞ 1 − µ m→∞ 1−µ (3.31) Therefore, not only have we proved that P ∈ C 1 , but also found the explicit formula for the C 0 and C 1 sizes of P which can be implemented in a computer. Our next task is to find the C 0 and C 1 sizes of πu , which projects B1 to the x2 -axis along the horizontal foliation F u . Recall that F u consists of the integral curves of the vector field defined in (3.17), i.e.,      1  Yu,1  Yu =   =  . Px Yu,2 48 Let x = (x1 , x2 ), and φ(x, t) be the solution of the vector field Yu with initial value x ∈ B1 , and let τ (x) be the real number such that the first coordinate of φ(x, τ (x)) is zero. Of course, if x1 > 0, then τ (x) < 0 and if x1 < 0, then τ (x) > 0. For simplicity, let us assume that x1 < 0. Since the first coordinate of Yu is 1 by definition, the maximum time τ (x) for any x ∈ B1 is ε0 . So we can write πu as (0, πu (x)) = φ(x, τ (x)) = (φ1 (x), φ2 (x)) with  x1 + φ(x, t) =  x2 +      t Y (φ (x))ds x1 + t 0 u,1 s    φ1 (x, t) = =  t Y (φ (x))ds t P ds x2 + 0 x φ2 (x, t) 0 u,2 s giving       0  0   x1 + τ (x)     = = . −x1 τ (x) πu (x) x2 + 0 Px dt x2 + 0 Px dt. We have |πu (x) − x2 | = | −x1 0 Px dt| ≤ |x1 | · |P | ≤ ε0 , 49 (3.32) ∂φ1 ∂φ1 ∂φ2 ∂φ1 ∂πu = 1, = 0, = 0, = for j = 1, 2, ∂x1 ∂x2 ∂x2 ∂xj ∂xj τ (x) = −x1 , ∂τ (x)/∂x1 = −1, and, ∂τ (x)/∂x2 = 0. Hence, by (3.31) and the Gronwall inequality, we have | −x1 ∂φ ∂φ ∂πu (x)| = | D P (φ(x, τ (x))) · ( 1 , 2 )ds ∂x1 ∂x1 ∂x1 0 ∂τ (x)P (φ(x, τ (x)))| + ∂x1 −x1 ∂ P ∂ P ∂πu = | ( + )ds − P (φ(x, τ (x)))| ∂x1 ∂x2 ∂x1 0 |x1 | ε F F ∂πu | |ds ≤ |P | + 0 + 1−µ 1 − µ ∂x1 0 ε F F ≤ ε0 (1 + )exp( 0 ), 1−µ 1−µ (3.33) when |P | ≤ ε0 in small B1 , which is possible due to the Inclination Lemma, as we have |P | = 0 along our initial B1top . And we have 50 ∂πu (x) − 1| ∂x2 τ (x) ∂φ ∂φ ∂τ | D P (φ(x, τ (x))) · ( 1 , 2 )ds + (x)P (φ(x, τ (x)))| ∂x2 ∂x2 ∂x2 0 τ (x) ∂ P ∂πu τ (x) ∂ P ( − 1)ds + ds| (3.34) | ∂x2 ∂x2 ∂x2 0 0 |x1 | F ε0 F ∂πu + ( )| (x) − 1|ds (3.35) 1−µ 1 − µ ∂x2 0 ε F ε0 F exp( 0 ). (3.36) 1−µ 1−µ | = = ≤ ≤ This completes the proof of Theorem 3.4.1. 51 3.5 Three dimensional vector fields Let us recall our assumptions on the vector field introduced at the beginning. We write x = (x1 , x2 , x3 ) ∈ R3 , and X(x) is a forward complete C 2 vector field : ¯ ¯ ¯ X(x) = (X1 (x), X2 (x), X3 (x)) = (ax1 + X1 (x), bx2 + X2 (x), cx3 + X3 (x)) where c < 2b < 0 < a, (3.37) ¯ Xi,x (0, 0, 0) = 0 for i, j = 1, 2, 3. j (3.38) and Let B1 = {x ∈ R3 |0 ≤ x2 ≤ ε0 , |x1 | ≤ ε0 , |x3 | ≤ ε0 } and let | · | = supx∈B | · |. Then we have a finite constant M > 0 such that 1 ¯ |Xi,x ,x (x)| ≤ M for i, j, k = 1, 2, 3. By the mean value theorem, we have some j k ¯ small constants ǫ1 > 0 and ǫ2 > 0, such that, |Xi,x (x)| ≤ M · ε0 = ǫ2 , and j ¯ |Xi (x)| ≤ ǫ2 · ε0 = ǫ1 for i, j = 1, 2, 3. Let B1top = {x ∈ B1 |x2 = ε0 }, 52 + B1 = {x ∈ B1 |x1 = ε0 }, − B1 = {x ∈ B1 |x1 = −ε0 }. Letting ϕt (x) be the local flow of X, so we may write ϕ(x, t) = (ϕ1 , ϕ2 , ϕ3 ) = (eat x1 + ϕ1 , ebt x2 + ϕ2 , ect x3 + ϕ3 ) ¯ ¯ ¯ where ϕi for i = 1, 2, 3 are the higher order terms. Set ψt (x) to be the flow of the ¯ corresponding linear vector field of X, i.e, ψt (x) = (eat x1 , ebt x2 , ect x3 ) Let f be the time-T map ϕT , for some 0 < T ≤ 1. So f : R3 → R3 is C 2 with a hyperbolic fixed point at 0. Let L = Df (0), and R3 = E u ⊕ E s ⊕ E ss be the splitting given by the eigenspaces of L. For convenience, we also write L = (Lu , Ls , Lss ) = (Lc , Lss ) with respect to the splitting E c ⊕ E ss , where E c = E u ⊕ E s and E u , E s and E ss correspond to the x1 , x2 and x3 axes respectively. In addition, we assume that the local unstable and stable manifolds of f have been u s straightened, i.e., Wloc (0) is in the x1 -axis and Wloc (0) is in the x2 x3 -plane inside B1 . Let B1ε = {x ∈ B1 : x3 = 0} 0 B2ε = {x ∈ B1 : x1 = 0, x2 = 0}. 0 53 For the most part of this section, unless we specify the index i = 1, 2, 3, we will always use this representation x = (x1 , x2 ) ∈ B1ε , and y = x3 ∈ B2ε , so that, 0 0 ˜ ˜ f (x, y) = (Lc x + f1 (x, y), Lss y + f2 (x, y)) = (f1 (x, y), f2(x, y)) (3.39) ˜ ˜ where f1 (x, y) and f2 (x, y) are the nonlinear terms. In B1 , we assume f = (f1 , f2 ) satisfies these conditions for λ1 > 1 and 0 < λ2 , λ3 < 1:   ∂f1 λ1 0  (0) =   ∂x 0 λ2 ∂f 0 < 2 (0) = λ3 < (λ2 )2 < 1 ∂y (3.40) for i = 1, 2 ˜ ˜ ˜ fi (0) = 0 and the first partial derivatives fix (0) = fiy (0) = 0, (3.41) and ˜ ˜ ˜ ˜ ˜ |fix | ≤ ǫ, |fiy | ≤ ǫ, |fixx | ≤ M1 , |fixy | ≤ M1 , |fiyy | ≤ M1 . (3.42) In this section, we consider a short initial line segment in the cone of angle less than π about the x1 -axis on the plane x2 = ε0 . 4 54 Choose small positive numbers w, ζ satisfying ε 0 < w < ζ < 0, 2 (3.43) and, consider the line segment ℓ : l(s) = (l1 , l2 , l3 ) = l(0) + sv ⊂ B1 where l(0) = (0, ε0 , w), −ζ < s < ζ, (3.44)   1     v = 0 ,     θ (3.45) and |θ| < 1. Let D0 = 0≤t≤T ϕt (l) ⊂ int(B1 ) which is possible when ℓ is short (i.e., ζ is small). And let D1 be the connected component of f (D0 ) ∩ B1 . Inductively, let Dn be the connected component of f (Dn−1 ) ∩ B1 . Define Σ = ∞ D , and j=0 j let Π(x1 , x2 , x3 ) = (x1 , x2 ) be the projection onto the x1 x2 −plane. Notice that Σ invariant almost by definition: f (Σ) ∩ B1 ⊂ Σ. The next result shows that Σ is the graph of a C 2 function g ⋆ from Π(Σ) to B2ε . 0 Theorem 3.5.1. Suppose that f, ℓ, and Σ satisfy the conditions (3.39)–(3.45) and 55 that the following inequalities hold, λ2 − ǫ > 0 , (λ3 + ǫ)(λ2 − ǫ)−1 < 1 (λ3 + ǫ)(1 + ǫ )<1 λ2 − ǫ λ3 + 2ǫ λ3 + ǫ λ + 2ǫ <1 <1 , +ǫ 3 λ2 − 2ǫ λ2 − 2ǫ (λ2 − 2ǫ)2 µ= λ3 + ǫ (λ2 − 2ǫ)2 + ǫ(λ3 + 2ǫ) < 1. (λ2 − 2ǫ)3 (3.46) (3.47) (3.48) (3.49) Then, there is a C 2 function g ⋆ : Π(Σ) → B2ε , such that g ⋆ (0, 0) = 0 and 0 Σ = {(x, g ⋆ (x))|x ∈ Π(Σ) ⊂ B1ε }. 0 Moreover, we have |g ⋆ | ≤ ε0 |Dg ⋆| ≤ |D 2 g ⋆ | ≤ ǫ λ2 − λ3 − 3ǫ 4M1 (λ2 + λ3 ) . (λ2 − 2ǫ)3 (1 − µ) (3.50) (3.51) (3.52) Once we have obtained g ⋆ as in the above theorem, we consider the graph map: h(x1 , x2 ) = (x1 , x2 , g ⋆ (x1 , x2 )). This is a C 2 diffeomorphism, and we can conjugate f to f : R2 → R2 defined by 56 h−1 f h(x1 , x2 ) = f (x1 , x2 ). This means, of course, that we can replace f by the two dimensional diffeomorphism f . We can also replace the flow φt restricted to Σ to a two-dimensional flow φt = h−1 φt h. The first step in the proof of Theorem 3.5.1 is to show that the upper part D0 of Σ is a C 2 graph. That is, Lemma 3.5.2. The set D0 is the graph {(x1 , x2 , x3 ) : x3 = g0 (x1 , x2 ))} of a C 2 function g0 defined on Π(D0 ). As we will show, this holds because X is C 1 close to its linear part, and the analogous statement is true for the linear part. Indeed, let us consider the linear vector field L(x1 , x2 , x3 ) = (ax1 , bx2 , cx3 ). Let ψ(t, x) = (ψ1 (t, x), ψ2 (t, x), ψ3 (t, x)) denote the flow of L, and let T > 0 be such that if x0 ∈ ℓ and |t| < T , then bε0 . |ψ1 (t, x0 )| < c−a Let 57 (3.53) D0 = {ψ(t, x) : |t| < T, x ∈ ℓ}. Proposition 3.5.3. Under the conditions above, the set D0 is the graph of the (real-analytic) function x c−a x3 = θx1 ( 2 ) b ε0 whose first order partial derivatives are bounded by 1. Proof: First, let us write down the solution of the linear ODE with initial condition (x10 , x20 , x30 ) ∈ B1 as follows: x1 = eat x10 (3.54) x2 = ebt x20 (3.55) x3 = ect x30 . (3.56) We want to represent x3 in terms of x1 and x2 , i.e., x3 = g0 (x1 , x2 ), the graph of which passes through the initial line segment l. By (3.55), we have x 1 et = ( 2 ) b . ε0 58 (3.57) Plugging it into (3.54), we have x −a x10 = x1 ( 2 ) b . ε0 Since x30 = θx10 , together with the above equation and (3.57), solution (3.56) becomes x c x −a x c−a x3 = ( 2 ) b · θx1 ( 2 ) b = θx1 ( 2 ) b . ε0 ε0 ε0 Let β = c−a , which satisfies β > 2, because of (3.37). b It is straightforward to compute that x ∂x3 = θ( 2 )β ∂x1 ε0 and ∂x3 θβx1 x2 β−1 = ( ) . ∂x2 ε0 ε0 ε Now, we have 0 ≤ x2 ≤ ε0 , and |θ| ≤ 1. Further, (3.53) gives |x1 | < 0 , so we β ∂x3 ∂x3 get that | | and | | are bounded by 1, and the proposition is proved. ∂x1 ∂x2 Using Lemma 3.5.5, let us prove the following lemma. Lemma 3.5.4. Given such D0 that contains l in B1 , D0 can be represented as the graph of a C 2 function g0 (x1 , x2 ) for (x1 , x2 ) ∈ Π(D0 ), where Π is the projection into the x1 x2 -plane. Moreover, we have Lip(g0 ) = sup |Dg0 | ≤ 1 x∈Π(D0 ) 59 and D 2 g0 (x) is uniformly bounded in Π(D0 ). Proof: Let the solution of the associated linear vector field of X be: ψ1 (t, l(0) + sv) = eat s ψ2 (t, l(0) + sv) = ebt ε0 ψ3 (t, l(0) + sv) = ect (w + θs). We have the solution of the vector field X: x1 (t, s) = ϕ1 (t, l(0) + sv) = ψ1 + ξ1 x2 (t, s) = ϕ2 (t, l(0) + sv) = ψ2 + ξ2 x3 (t, s) = ϕ3 (t, l(0) + sv) = ψ3 + ξ3 where ψi ’s and ξi ’s are evaluated at (t, l(0) + sv). In order to find g0 (x1 , x2 ), we express (t, s) as functions of (x1 , x2 ), so that we can write g0 (x1 , x2 ) = ϕ3 (t(x1 , x2 ), l(0) + s(x1 , x2 )v). 60 This requires that the determinant of the Jacobian does not equal to zero:   ∂(x1 , x2 ) x1,t x1,s  | = det  DET = det |  ∂(t, s) x2,t x2,s   ψ1,t + ξ1,t ψ1,s + ξ1,s  = det   ψ2,t + ξ2,t ψ2,s + ξ2,s   at s + ξ at + ξ 1,t e 1,s   ae = det   bt ε + ξ be 0 2,t 0 + ξ2,s = (aeat s + ξ1,t )ξ2,s − (eat + ξ1,s )(bebt ε0 + ξ2,t ) We have DET ≈ bε0 e(a+b)t = 0 (3.58) when ξi ’s are small enough. ∂x ∂x Here we use the notation xi,t = i and xi,s = i for i = 1, 2. ∂t ∂s In addition, since     tx2  ∂(t, s) 1  x2,s −x1,s   tx = 1 =   ∂(x1 , x2 ) DET sx1 sx2 −x2,t x1,t we have     at − ξ ξ2,s −e 1  1,s   tx1 tx2   =  . DET bt ε − ξ at s + ξ sx1 sx2 −be 0 2,t ae 1,t 61 So we may compute the first derivatives of g0 : | ∂g0 (x1 , x2 ) ∂ϕ ∂ϕ3 ∂s ∂t | = | 3· + · | ∂x1 ∂t ∂x1 ∂s ∂x1 (ψ3,t + ξ3,t ) · ξ2,s (ψ3,s + ξ3,s )(−bebt ε0 − ξ2,t ) = | + | DET DET (cect (w + θs) + ξ3,t ) · ξ2,s (ect θ + ξ3,s )(−bebt ε0 − ξ2,t ) = | + | DET DET −bθε0 e(b+c)t | , by (3.58) ≈ | bε0 e(a+b)t = |θe(c−a)t | < 1 where ξi,j for i = 2, 3, j = t, s are small enough. Similarly, we have | ∂g0 (x1 , x2 ) ∂ϕ ∂ϕ3 ∂s ∂t | = | 3· + · | ∂x2 ∂t ∂x2 ∂s ∂x2 (ψ3,t + ξ3,t ) · (−eat − ξ1,s ) (ψ3,s + ξ3,s )(aeat s + ξ1,t ) = | + | DET DET (cect (w + θs) + ξ3,t ) · (−eat − ξ1,s ) = | DET ct θ + ξ )(aeat s + ξ ) (e 3,s 1,t + | DET −ce(a+c)t (w + θs) + a · sθe(a+c)t ≈ | | , by (3.58) bε0 e(a+b)t θ(a − c)s · e(a+c)t ≈ | | , where w is close to 0 bε0 e(a+b)t (a − c)s (c−b)t ≤ |θ| · | |e <1 bε0 62 (a−c)s where ξi,j for i = 1, 3, j = t, s are small and | | < 1. bε0 Since g0 is C 2 , all its second derivatives are uniformly bounded over Π(D0 ) ⊂ B1ε , the lemma is proved. 0 Lemma 3.5.5. Let φt (x) be the solution of the vector field ¯ ¯ ¯ ¯ X(x) = (ax1 + X1 (x), bx2 + X2 (x), cx3 + X3 (x)) = (L + X)(x), where a > 0 > b > c and −c > a, with ¯ ¯ ¯ |Xi | ≤ ǫ1 , |Xi,x | ≤ ǫ2 , |D 2 X| ≤ M1 , for i, j = 1, 2, 3. j Let ψt (x) be the solution of the corresponding linear vector field, then in B1 , there exists small constants δ0 , δ and a constant M2 , such that, |φt (x) − ψt (x)| ≤ δ0 , |Dφt(x) − Dψt (x)| ≤ δ, and |D 2 φt (x) − D 2 ψt (x)| = |D 2 φt (x)| ≤ M2 . 63 Proof: By definition, we have t φt (x) − ψt (x) = x + = t 0 0 ¯ (L + X)(φs (x))ds − x − ¯ X(φs (x))ds + t 0 t 0 Lψs (x)ds L(φs − ψs )(x)ds. So by the Gronwall’s inequality, we have |φt (x) − ψt (x)| ≤ t 0 ¯ |X(φs (x))|ds + t 0 |L||φs (x) − ψs (x)|ds ≤ ǫ1 te|c|t ≤ ǫ1 T e|c|T = δ0 Next, let us compare Dφt (x) with Dψt (x), which come from the first variational equations of φt (x) and ψt (x). We have ˙ ¯ D φt(x) = DXφ (x) · Dφt (x) = (DL + D Xφ (x) )Dφt (x). t t so ˙ |D φt(x)| ≤ |Id| + t s=0 ¯ |DL + D Xφ (x) ||Dφs (x)|ds. s ¯ By our assumption in this lemma, we have each entry Xi,x (φs (x)), i, j = 1, 2, 3, j ¯ of the 3 × 3 matrix D Xφ (x) is bounded by ǫ2 , thus s ¯ ¯ ¯ ¯ |D Xφ (x) | = max {|Xi,x | + |Xi,x | + |Xi,x |} ≤ 3ǫ2 . s 1 2 3 i=1,2,3 64 Then by the Gronwall’s inequality we have ¯ |Dφt (x)| ≤ exp(|DL + D X|t) ≤ exp((|c| + 3ǫ2 )t). (3.59) Now Dφt(x) − Dψt (x) t ¯ (DL + D Xφ (x) )Dφs (x)ds − Id − DL · Dψs (x)ds s 0 0 t t ¯ D Xφ (x) Dφs (x)ds + DL(Dφs (x) − Dψs (x))ds. s 0 0 = Id + = t By the Gronwall’s inequality again and the inequality (3.59), we have ¯ |Dφt(x) − Dψt (x)| ≤ 3ǫ2 t exp(|DL + D X|t) exp(|DL|t) ¯ ≤ 3ǫ2 t exp((2|DL| + |D X|)T ) ≤ 3ǫ2 T exp((2|c| + 3ǫ2 )T ) = δ. Finally, let us find the size of |D 2 φt (x) − D 2 ψt (x)| = |D 2 φt (x) − 0|. We have t ∂ ∂φ(s, x) ∂2φ (t, x) = (ei + DXφ (x) ds) s ∂xi ∂xj ∂xj ∂xi 0 t ∂φ ∂φ ∂ 2 φ(s, x) = D 2 Xφ (x) · ( )( ) + DXφ (x) · ds. s s ∂xi ∂xj ∂xi ∂xj 0 ¯ By using (3.59) and the Gronwall’s inequality again, and since D 2 Xφ (x) = D 2 Xφ (x) s s 65 we have | ∂2φ (t, x)| ≤ ∂xi ∂xj ≤ t 0 M1 exp(2s(|c| + 3ǫ2 ))ds + t 0 (|c| + 3ǫ2 ) · | ∂ 2 φ(s, x) |ds ∂xi ∂xj M1 [exp(2T (|c| + 3ǫ2 )) − 1] · exp(T (|c| + 3ǫ2 )) 2(|c| + 3ǫ2 ) = M2 . Thus the lemma is proved. Let us be briefly reminded of the definition of our surface Σ. We have D0 = 0≤t≤T ϕt (l) ⊂ int(B1 ) for our initial line segment l. By now we have shown that D0 can be represented as a graph of a C 2 function g0 . Now let D1 be the connected component of f (D0 ) ∩ B1 . And inductively, let Dn be the connected component of f (Dn−1 ) ∩ B1 which is connected to Dn . We define Σ = ∞ D . j=0 j Theorem 3.5.6. Σ can be represented as the graph of a C 2 function g ⋆ . We will use the same Fiber Contraction method as in the previous section. The proof will require several steps. It involves obtaining a sequence g1 , g2 , · · · , of C 2 functions, such that gn converges to g ⋆ uniformly on B1 , (3.60) Dgn converges uniformly on B1 , (3.61) 66 D 2 gn converges uniformly on B1 . (3.62) The proof involves definitions of certain function spaces and associated mappings. Let us proceed to define our first function space G. Here we recall that B1ε 0 is B1 ∩ x1 x2 -plane, and B2ε is B1 ∩ x3 -axis. Let 0 0 B1 = {g|g ∈ C 0 (Π(Σ), B2ε ), Lip(g) = sup 0 ξ1 =ξ2 |g(ξ1) − g(ξ2 )| ≤ 1}. |ξ1 − ξ2 | 0 B1 is a complete metric space with respect to the supremum norm (defined above). The subspace 0 G = {g| g ∈ B1 , g(x1 , 0) = 0 f or |x1 | ≤ ε0 , g|Π(D ) = g0 } 0 is also a complete metric space equipped with the metric d(g1 , g2 ) = sup |g1 (x) − g2 (x)|. x∈B1ε 0 For x ∈ Π(Σ) ∩ Π(f (Σ)), we have the following graph transform formula Γf (g) = f2 ◦ (1, g) ◦ [f1 ◦ (1, g)]−1 . We need to show that Γf : G → G is a well defined contraction map. This will give us a continuous function g as the unique attractive fixed point of Γ. Lemma 3.5.7. Given f satisfying the conditions at the beginning of this section, if 67 λ2 − ǫ > 0 and (λ3 + ǫ)(λ2 − ǫ)−1 ≤ 1 in B1 , then Γf (g) is well-defined and maps G to itself. If, in addition, η = (λ3 + ǫ)(1 + ǫ )<1 λ2 − ǫ then Γf (g) is a contraction mapping with contraction µ. The unique fixed point of Γf (g) is a C 0 function g ⋆ with |g ⋆ | ≤ η |g − g0 |. 1−η 1 Proof: To show that Γf is well-defined, it is sufficient to show that f1 ◦ (1, g) is invertible. Let E denote the x1 x2 −plane. For u = (u1 , u2 ) ∈ E, (L | E)(u) = (λ1 u1 , λ2 u2 ). Let min(L | E) = inf |(L | E)u| > λ2 > 0. |u|=1 For two points u = v ∈ B1ε , we have 0 68 ˜ ˜ |f1 ◦ (1, g)(u) − f1 ◦ (1, g)(v)| = |(L | E)(u − v) + f1 (u, gu) − f1 (v, gv)| ˜ ˜ ≥ min(L | E)|u − v| − |f1 (u, gu) − f1 (v, gv)| ˜ ≥ λ2 |u − v| − Lip(f1 )Lip(1, g)|u − v|. ˜ Since Lip(f1 ) ≤ ǫ and Lip(1, g) ≤ 1, we have |f1 ◦ (1, g)(u) − f1 ◦ (1, g)(v)| ≥ (λ2 − ǫ)|u − v| > 0. (3.63) Hence, on B1ε , f1 ◦ (1, g) is injective, and thus, invertible. That also implies 0 Lip([f1 ◦ (1, g)]−1 ) ≤ (λ2 − ǫ)−1 . In order to have Γf (g) map G to itself, we first prove that for x1 ∈ [0, ε0 ] we have Γf (g)(x1 , 0) = 0. u Let us be reminded that, as a part of our assumptions on f , we have Wloc (0) of f is in the x1 -axis inside B1 . In addition, we have had Γf (g) = f2 ◦ (1, g) ◦ [f1 ◦ (1, g)]−1 being well defined, i.e., f1 ◦ (1, g) is a one-to-one invertible map. So we know there is a unique (¯1 , 0) ∈ B1ε with |¯1 | ≤ |x1 | such that x x 0 f1 ◦ (1, g)(¯1 , 0) = (x1 , 0) x 69 where |¯1 | = |x1 | is achieved when x1 = 0. So x Γf (g)(x1 , 0) = f2 ◦ (1, g)(¯1 , 0) = f2 (¯1 , 0, g(¯1 , 0)). x x x For g ∈ G, we have g(¯1 , 0) = 0, so x Γf (g)(x1 , 0) = f2 (¯1 , 0, 0) = 0. x Next, let us prove Lip(Γf (g)) ≤ 1 as follows. Again, for u = (u1 , u2 ), v = (v1 , v2 ) ∈ B1ε , we have 0 ˜ ˜ |f2 ◦ (1, g)(u) − f2 ◦ (1, g)(v)| = |λ3 g(u) − λ3 g(v) + f2 (u, gu) − f2 (u, gv)|. (3.64) ˜ Since Lip(f2 ) ≤ ǫ, Lip(1, g) ≤ 1, we have |f2 ◦ (1, g)(u) − f2 ◦ (1, g)(v)| ≤ (λ3 + ǫ)|u − v|. (3.65) As we know the Lipschitz constant of the composite function of any f and g over domain U can be computed as Lip(f ◦ g(x)) = sup x,y∈U |f (g(x)) − f (g(y))| |x − y| ≤ sup x,y∈U Lip(f )|g(x) − g(y)| = Lip(f )Lip(g) |x − y| 70 we have Lip(Γf (g)) ≤ (λ3 + ǫ)(λ2 − ǫ)−1 . Under the condition that (λ3 + ǫ)(λ2 − ǫ)−1 ≤ 1, we have Lip(Γf (g)) ≤ 1, and thus, Γf (g) maps G to itself. Now let us show Γ is a contraction mapping in G. Let ξ1 = [f1 ◦ (1, g1 )]−1 , ξ2 = [f1 ◦ (1, g2 )]−1 , we have |Γf (g1 ) − Γf (g2 )| (3.66) = |f2 ◦ (1, g1 ) ◦ [f1 ◦ (1, g1 )]−1 − f2 ◦ (1, g2 ) ◦ [f1 ◦ (1, g2 )]−1 | = |f2 ◦ (1, g1 )ξ1 − f2 ◦ (1, g2 )ξ2 )| ≤ |f2 ◦ (1, g1 )ξ1 − f2 ◦ (1, g2 )ξ1 | + |f2 ◦ (1, g2 )ξ1 − f2 ◦ (1, g2 )ξ2 )| ≤ Lip(f2 )|(ξ1 , g1 ξ1 ) − (ξ1 , g2 ξ1 )| + Lip(f2 )Lip((1, g2 ))|ξ1 − ξ2 | ≤ Lip(f2 )|g1 ξ1 − g2 ξ1 | + Lip(f2 )|ξ1 − ξ2 | = Lip(f2 )|g1 − g2 | + Lip(f2 )|ξ1 − ξ2 |. So we need to find |ξ1 − ξ2 |. Claim: |ξ1 − ξ2 | ≤ ǫ |g − g2 |. λ2 − ǫ 1 (3.67) Let us prove this claim. In our notations, ξ1 = [f1 ◦ (1, g1 )]−1 , ξ2 = [f1 ◦ (1, g2 )]−1 , −1 −1 let τ1 = ξ1 x and τ2 = ξ2 x, then we have x = ξ1 τ1 = f1 ◦ (1, g1 )τ1 , x = ξ2 τ2 = 71 f1 ◦ (1, g2 )τ2 . So f1 ◦ (1, g1 )τ1 = f1 ◦ (1, g2 )τ2 . As before, we write ˜ f1 (τ1 , g1 τ1 ) = Lc τ1 + f1 (τ1 , g1 τ1 ), and ˜ f1 (τ2 , g2 τ2 ) = Lc τ2 + f1 (τ2 , g2 τ2 ). So we have ˜ ˜ Lc τ1 + f1 (τ1 , g1 τ1 ) = Lc τ2 + f1 (τ2 , g2 τ2 ) ˜ ˜ Lc (τ2 − τ1 ) = f1 (τ1 , g1 τ1 ) − f1 (τ2 , g2 τ2 ) then, ˜ ˜ ˜ ˜ Lc (τ2 − τ1 ) = f1 (τ1 , g1 τ1 ) − f1 (τ2 , g1 τ2 ) + f1 (τ2 , g1 τ2 ) − f1 (τ2 , g2 τ2 ) which implies ˜ ˜ ˜ ˜ Lc (τ2 − τ1 ) − f1 (τ1 , g1 τ1 ) + f1 (τ2 , g1 τ2 ) = f1 (τ2 , g1 τ2 ) − f1 (τ2 , g2 τ2 ) ˜ ˜ f1 ◦ (1, g1 )τ2 − f1 ◦ (1, g1 )τ1 = f1 (τ2 , g1 τ2 ) − f1 (τ2 , g2 τ2 ). 72 By (3.63), i.e., |f1 ◦ (1, g1 )τ2 − f1 ◦ (1, g1 )τ1 | ≥ (λ2 − ǫ)|τ1 − τ2 | we have ˜ ˜ (λ2 − ǫ)|τ1 − τ2 | ≤ |f1 (τ2 , g1 τ2 ) − f1 (τ2 , g2 τ2 )| ˜ ≤ Lip(f1 )|(1, g1)τ2 − (1, g2 )τ2 | ≤ ǫ|g1 τ2 − g2 τ2 | ≤ ǫ|g1 − g2 |. By our notations τ1 = ξ1 x and τ2 = ξ2 x, so |ξ1 − ξ2 | ≤ ǫ |g − g2 | λ2 − ǫ 1 and the claim (3.67) is proved. Plugging (3.67) to (3.66), we have |Γf (g1 ) − Γf (g2 )| ≤ Lip(f2 )|g1 − g2 | + Lip(f2 ) ǫ |g − g2 |. λ2 − ǫ 1 ˜ ˜ As we write f2 (x1 , x2 ) = Lss x2 + f2 (x1 , x2 ) with Lip(f2 ) ≤ ǫ, we have Lip(f2 ) ≤ λ3 + ǫ. Hence, |Γf (g1 ) − Γf (g2 )| ≤ (λ3 + ǫ)(1 + 73 ǫ )|g − g2 |. λ2 − ǫ 1 We may observe that (λ3 + ǫ)(1 + ǫ ) is approximately equal to λ3 , which λ2 −ǫ is much less than 1. If the neighborhood is properly chosen, we will have η = (λ3 + ǫ)(1 + ǫ ) < 1. λ2 − ǫ So the Lemma is proved. Thus, there is a unique globally attracting fixed point g ⋆ ∈ G for the operator Γ. To show that g ⋆ is, in fact, C 2 , and to estimate its C 2 size, we modify the fiber contraction procedure of Hirsch and Pugh familiar from invariant manifold theory [4]. We construct two product bundles G × H and G × H × K and maps Φ : G × H → G × H, Ψ :G ×H×K →G ×H×K satisfying the following properties: 1. Φ is a fiber contraction over Γ, 2. Ψ is a fiber contraction over Φ, 3. if g is any map in G which happens to be C 2 and (g, Dg, D 2g) is in G × H × K, 74 then (a) Φ(g, Dg) = (Γ(g), DΓ(g)), and (b) Ψ(g, Dg, D 2g) = (Γ(g), DΓ(g), D 2Γ(g)). It will follow that, if g ∈ G is C 2 , and we set gn = Γn (¯) for each n > 0, then ¯ g the convergence conditions (3.60), (3.61), and (3.62) will hold, and the proof that g ⋆ is C 2 will be complete. In the process, we will also obtain bounds for the first and second derivatives of g ⋆. We have several tasks ahead. First, we find a C 2 initial function g in G. ¯ Recall that our original function g0 on Π(D0 ) was C 2 . Now, choose a smooth bump function ρ : R2 → R such that ρ(x) = 1 on a neighborhood of Π(D0 ) and ρ(x) = 0 on a neighborhood of B1bot = {(x1 , 0, 0)|0 ≤ x1 ≤ ε0 }. Then, letting ν(x) be the identically 0 function on R2 , and define g by ¯ g = ρ(x)g0 + (1 − ρ(x))ν(x). ¯ The function g is clearly C 2 and is evidently in G. ¯ Next, we proceed to define the function space H and the map Φ. 0 Let H = C1 (Π(Σ), L(R2 , R)) be the set of continuous functions H from Π(Σ) to the linear maps from R2 to R, and |H| ≤ 1. H becomes a complete metric space 75 equipped with the metric d(H1 , H2 ) = |H1 − H2 | induced by the norm |H| = sup |Hx (y)|. 2 ,|y|=1 x∈B1ε ,y∈R 0 We can see that H is the set of candidates of Dg. To get the fiber contraction operator over the functional space G × H, we proceed as follows. Let u(x) = [f1 ◦ (1, g)]−1 (x). By the chain rule, we take the derivative of Γf (g) = f2 ◦ (1, g) ◦ [f1 ◦ (1, g)]−1 with respect to x, we have Dx Γf (g) = (f2x + f2y Du(x) g)(f1x + f1y Du(x) g)−1 where the partial derivatives of f are evaluated at the point (u(x), g(u(x))). In this form, we define R(g, H) = (f2x + f2y Hu(x) )(f1x + f1y Hu(x) )−1 where the partial derivatives of f are evaluated at the point (u(x), g(u(x))). And let Φ : G × H → G × H be Φ(g, H) = (Γf (g), R(g, H)). By construction, if g is C 1 , then Φ(g, Dg) = (Γf (g), D(Γf (g))). 76 We shall show that Φ(g, H) = (Γf (g), R(g, H)) is a fiber contraction on G × H with the maximum metric in this product space. Lemma 3.5.8. Given f satisfying the conditions at the beginning of this section, if the following conditions are satisfied in B1 λ3 + ǫ λ + 2ǫ λ3 + 2ǫ < 1 and +ǫ 3 < 1, λ2 − 2ǫ λ2 − 2ǫ (λ2 − 2ǫ)2 (3.68) then the function g ⋆ in Lemma 3.5.7 is C 1 with |Dg ⋆ | ≤ ǫ . λ2 − λ3 − 3ǫ (3.69) Proof: Let us prove this lemma by first showing that Φ(g, H) maps G × H to itself. Since Γf (g) : G → G, we only need to show R(g, H) : G × H → H. As we can write −1 f1x + f1y Hu(x) = f1x (I + f1x f1y Hu(x) ), 77 for |H| ≤ 1, we have −1 |(f1x + f1y Hu(x) )−1 | = |{f1x (I + f1x f1y Hu(x) )}−1 | −1 −1 = |(I + f1x f1y Hu(x) )−1 f1x | −1 −1 ≤ |(I + f1x f1y Hu(x) )−1 ||f1x | 1 −1 |f1x | −1 f H min(I + f1x 1y u(x) ) 1 −1 |f1x | ≤ −1 f H 1 − |f1x 1y u(x) | −1 |f1x | ≤ −1 1 − |f1x ||f1y | 1 = . −1 |f1x |−1 − |f1y | = −1 Since |f1y | ≤ ǫ and |f1x |−1 = min(f1x ) ≥ λ2 − ǫ, we have 1 |(f1x + f1y Hu(x) )−1 | ≤ . λ2 − 2ǫ Moreover, |R(g, H)| ≤ |f2x + f2y Hu(x) ||(f1x + f1y Hu(x) )−1 | ≤ (ǫ + (λ3 + ǫ)|Hu(x) |) ≤ (3.70) 1 λ2 − 2ǫ λ3 + 2ǫ . λ2 − 2ǫ λ3 +2ǫ < 1 in some neighborhood of 0, we have proved that Φ(g, H) λ2 −2ǫ maps G × H to itself. If we have 78 Now we shall show Φ(g, H) is a fiber contraction map. Let R2 (g, H)x(y) = (f2x + f2y Hu(x) )(y), R1 (g, H)x(y) = ((f1x + f1y Hu(x) )−1 )(y), both of which are linear operators in y. So we may write R(g, H)x = R2 (g, H) ◦ R1(g, H)x, where the partial derivatives of f are evaluated at the point (u(x), g(u(x))). By what we have just computed, we know 1 |R1 (g, H)| = |(f1x + f1y Hu(x) )−1 | ≤ λ2 − 2ǫ (3.71) |R2 (g, H)| = |f2x + f2y Hu(x) | ≤ λ3 + 2ǫ. (3.72) and And by the linearity of R1 and R2 , we have |R(g, H1)x − R(g, H2 )x | ≤ |R2 (g, H1) ◦ R1 (g, H1)x − R2 (g, H2) ◦ R1 (g, H1)x | +|R2 (g, H2) ◦ R1 (g, H1 )x − R2 (g, H2) ◦ R1 (g, H2 )x | ≤ |R2 (g, H1) − R2 (g, H2 )||R1 (g, H1 )x | +|R2 (g, H2) ◦ (R1 (g, H1 )x − R1 (g, H2 )x )| 1 ≤ |f2y H2 − f2y H1 | λ2 − 2ǫ +|R2 (g, H2)||R1 (g, H1)x − R1 (g, H2)x | 1 ≤ (λ3 + ǫ)|H2 − H1 | λ2 − 2ǫ +(λ3 + 2ǫ)|R1 (g, H1)x − R1 (g, H2)x |. 79 Here |R1 (g, H1 )x − R1 (g, H2 )x | = |(f1x + f1y H1 )−1 − (f1x + f1y H2 )−1 | which involves inverse functions. So let us consider this trick: −1 −1 −1 −1 F1 − F2 = F1 (F2 − F1 )F2 which implies −1 −1 −1 −1 |F1 − F2 | ≤ |F1 ||F2 ||F1 − F2 | By the above inequality, we have |R1 (g, H1 )x − R1 (g, H2 )x | = |(f1x + f1y H1 )−1 − (f1x + f1y H2 )−1 | ≤ |(f1x + f1y H1 )−1 ||(f1x + f1y H2 )−1 ||(f1x + f1y H1 ) − (f1x + f1y H2 )| 1 )2 |f1y H1 − f1y H2 | λ2 − 2ǫ 1 ≤ ( )2 ǫ|H1 − H2 |. λ2 − 2ǫ ≤ ( Therefore, λ +ǫ λ + 2ǫ |R(g, H1 ) − R(g, H2)| ≤ ( 3 +ǫ 3 )|H1 − H2 |. λ2 − 2ǫ (λ2 − 2ǫ)2 By (3.68), we have µ= λ3 + ǫ λ + 2ǫ +ǫ 3 < 1. λ2 − 2ǫ (λ2 − 2ǫ)2 Hence, Φ is indeed a fiber contraction, and the lemma is proved. 80 Moreover, we can find the size of Dg ⋆ . By a similar argument in the previous section, we start from (3.70), which says |Dgn| = |R(gn−1 , Dgn−1 )| ≤ ( λ +ǫ ǫ + 3 |Dgn−1 |) λ2 − 2ǫ λ2 − 2ǫ Let J= ǫ , λ2 − 2ǫ and N = λ3 + ǫ , λ2 − 2ǫ we have |Dgn| ≤ J + N|Dgn−1 |. Similar to the estimate in the inequality (3.30), we have |Dg ⋆ | ≤ ǫ J = . 1−N λ2 − λ3 − 3ǫ In order to have g ⋆ in C 2 , we need another fiber contraction to compute D 2 g ⋆ . We still use the same notation and definition of G and H. Recall that 0 H = C1 (Π(Σ), L(R2 , R)), now we define another functional space K to be the set of candidates of D 2 g: K = C 0 (Π(Σ) × (R2 )2 , R) which is the set of bounded continuous functions K from Π(Σ) to the bilinear maps from R2 to R. K is a complete metric space equipped with the metric d(K1 , K2 ) = 81 |K1 − K2 | induced by the norm |K| = sup |Kx (v1 , v2 )|. 2 x∈B1ε ,|v1 |=1,|v2 |=1,v1 ,v2 ∈R 0 Similar to the previous fiber contraction constructions, we first use the chain 2 rule to find Dx Γf (g) which gives the form of our bundle map. As we know Γf (g) = f2 ◦ (1, g) ◦ [f1 ◦ (1, g)]−1 . Let u(x) = [f1 ◦ (1, g)]−1 , we have Dx Γf (g) = f2x (u(x), g(u(x)))Du(x) + f2y (u(x), g(u(x)))Dg(u(x))Du(x). Now we take the derivative of the above with respect to x one more time to obtain the following formula, all the derivatives of f are evaluated at (u(x), g(u(x))): 2 Dx Γf (g) = f2xx (Du(x))2 + f2xy Dg(u(x))(Du(x))2 + f2x D 2 u(x) (3.73) +{f2yx Du(x) + f2yy Dg(u(x))Du(x)}Dg(u(x))Du(x) +f2y {D 2 g(u(x))(Du(x))2 + Dg(u(x))D 2 u(x)}. Now let us explicitly compute Du(x) and D 2 u(x). We have Du(x) = [f1x (u(x), g(u(x))) + f1y (u(x), g(u(x)))Dg(u(x))]−1 82 and with all the derivatives of f evaluated at (u(x), g(u(x))): D 2 u(x) = −Du(x){f1xx (Du(x))2 + 2f1xy Dg(u(x))(Du(x))2 +f1yy (Dg(u(x)))2(Du(x))2 + f1y (Du(x))2 D 2 g(u(x))}. In this form, we define D1 = D1 (u, H)x = [f1x + f1y Hu(x) ]−1 and D2 = D2 (u, H, K)x = −D1 {f1xx (D1 )2 + 2f1xy Hu(x) (D1 )2 +f1yy (Hu(x) )2 (D1 )2 + f1y (D1 )2 Ku(x) }. So we define the following functional derived from (3.73) ¯ R(g, H, K) = f2xx (D1 )2 + f2xy Hu(x) (D1 )2 + f2yxD1 Hu(x) D1 +f2yy Hu(x) D1 Hu(x) D1 + f2x D2 +f2y Ku(x) (D1 )2 + f2y Hu(x) D2 where the partial derivatives of f are evaluated at the point (u(x), g(u(x))), and only the last three terms contain K. To get the fiber contraction operator in the functional space G × H × K, we let Ψ : G × H × K → G × H × K satisfy this form Ψ(g, Dg, D 2g) = (Γf (g), DΓf (g), D 2 Γf (g)). 83 So we write ¯ Ψ(g, H, K) = (Γf (g), R(g, H), R(g, H, K)) and we shall show that Ψ(g, H, K) is a fiber contraction on G × H × K with the maximum metric in this product space. Lemma 3.5.9. Given f satisfying the conditions at the beginning of this section. In addition, each one of the second derivatives of f is bounded by some constant M1 > 0. If in B1 the following condition is satisfied µ= λ3 + ǫ (λ2 − 2ǫ)2 + ǫ(λ3 + 2ǫ) < 1, (λ2 − 2ǫ)3 (3.74) then g ⋆ is C 2 with |D 2 g ⋆ | ≤ 4M1 (λ2 + λ3 ) . (λ2 − 2ǫ)3 (1 − µ) (3.75) Proof: Let us first prove that Ψ maps G × H × K to itself, and it is clear that ¯ we only need to show, for given g and H, R(g, H, K) is still in K. By our definition D1 = [f1x + f1y Hu(x) ]−1 and what we have obtained in inequalities (3.71), we have |D1 | = |(f1x + f1y Hu(x) )−1 | ≤ 84 1 . λ2 − 2ǫ Also we have |D2 | ≤ |D1 |{|f1xx ||D1 |2 + 2|f1xy ||D1 |2 +|f1yy ||D1 |2 + |f1y ||D1 |2 |K|} = |D1 |3 (|f1xx | + 2|f1xy | + |f1yy | + |f1y ||K|) ≤ 4M1 + ǫ|K| . (λ2 − 2ǫ)3 In addition, since |H| ≤ 1 and each one of the second derivatives of f is bounded by some constant M1 , we have ¯ |R(g, H, K)| ≤ |f2xx ||D1 |2 + 2|f2xy ||D1 |2 + |f2yy ||D1 |2 (3.76) +(|f2x | + |f2y |)|D2 | + |f2y ||K||D1 |2 4M1 4M1 + ǫ|K| (λ3 + ǫ)|K| + (λ3 + 2ǫ) + 2 (λ2 − 2ǫ) (λ2 − 2ǫ)3 (λ2 − 2ǫ)2 4M1 4M1 (λ3 + 2ǫ) = + (λ2 − 2ǫ)2 (λ2 − 2ǫ)3 (λ3 + ǫ) ǫ(λ3 + 2ǫ) +{ + }|K|. (λ2 − 2ǫ)3 (λ2 − 2ǫ)2 ≤ ¯ Since K is also bounded, we know R(g, H, K) is bounded. Hence, we have proved Ψ maps G × H × K to itself. Now let us show the bundle map Ψ is a fiber contraction ¯ on G × H × K. By definition, what we need to show is the map K → R(g, H, K) is 85 a contraction mapping. Again, from ¯ R(g, H, K) = f2xx (D1 )2 + f2xy Hu(x) (D1 )2 + f2yxD1 Hu(x) D1 +f2yy Hu(x) D1 Hu(x) D1 + f2y Ku(x) (D1 )2 +(f2x + f2y Hu(x) )D2 where D2 = −D1 {f1xx (D1 )2 + 2f1xy Hu(x) (D1 )2 + f1yy (Hu(x) )2 (D1 )2 } −D1 f1y (D1 )2 Ku(x) ¯ we notice that only the last two terms of R(g, H, K) involve K. We may have the terms that do not contain K canceled if we perform the following substraction, i.e., ¯ ¯ R(g, H, K1 ) − R(g, H, K2) = f2y K1 (D1 )2 − f2y K2 (D1 )2 +(f2x + f2y H){−D1 f1y (D1 )2 K1 + D1 f1y (D1 )2 K2 }. 86 By using |H| ≤ 1, we have ¯ ¯ |R(g, H, K1) − R(g, H, K2 )| ≤ |f2y K1 (D1 )2 − f2y K2 (D1 )2 | +|f2x + f2y H||D1 f1y (D1 )2 K1 − D1 f1y (D1 )2 K2 | ≤ (λ3 + ǫ)|K1 (D1 )2 − K2 (D1 )2 | (λ3 + 2ǫ)|D1 f1y (D1 )2 ||K1 − K2 | ≤ (λ3 + ǫ)|K1 − K2 ||D1 |2 (λ3 + 2ǫ)ǫ|D1 |3 |K1 − K2 | ≤ ( λ3 + ǫ (λ2 − 2ǫ)2 + ǫ(λ3 + 2ǫ) ) |K1 − K2 |. (λ2 − 2ǫ)3 µ We may observe that µ= λ3 + ǫ (λ2 − 2ǫ)2 + ǫ(λ3 + 2ǫ) λ3 ≈ . 3 (λ2 − 2ǫ) (λ2 )2 If close enough to the saddle fixed point and the strong contraction is much smaller than the weak contraction, we may well have µ < 1. So under the condition that λ3 + ǫ (λ2 − 2ǫ)2 + ǫ(λ3 + 2ǫ) <1 (λ2 − 2ǫ)3 ¯ we have proved that the map K → R(g, H, K) is a contraction mapping. Moreover, we can find the size of D 2 g ⋆ by essentially the same argument used in the previous section about the iteration (3.70). So we do not repeat the full argument here, the 87 complete details of which has been shown in the previous section. Here we start with the following estimate, derived from (3.76), which says ¯ |R(g, H, K)| ≤ 4M1 (λ3 + 2ǫ) (λ3 + ǫ) 4M1 ǫ(λ3 + 2ǫ) + + } |K|. +{ 2 3 3 (λ − 2ǫ)2 (λ2 − 2ǫ) (λ2 − 2ǫ) (λ2 − 2ǫ) 2 J N Let J= 4M1 4M1 (λ3 + 2ǫ) 4M1 (λ2 + λ3 ) + = 2 3 (λ2 − 2ǫ) (λ2 − 2ǫ) (λ2 − 2ǫ)3 and from the above µ= λ3 + ǫ ǫ(λ3 + 2ǫ) + < 1. (λ2 − 2ǫ)3 (λ2 − 2ǫ)2 Since K’s are candidates of D 2 g ⋆ , the above inequality gives us a recursive relation as ¯ |D 2 gn | = |R(gn−1 , Dgn−1 , D 2 gn−1 )| ≤ J + µ|D 2 gn−1 |. Then we have the estimate similar to the one in the inequality (3.30), i.e., |D 2 g ⋆ | ≤ 4M1 (λ2 + λ3 ) J = 1−µ (λ2 − 2ǫ)3 (1 − µ) . 88 3.6 Straightening Invariant manifolds of a three dimensional map In this section, we flatten out the local stable and unstable manifolds of a C 2 map f : R3 → R3 near its saddle fixed point at 0. We will make a C 2 transformation of the neighborhood so that the local invariant manifolds of f are flattened. First of all, let us state the Hadamard-Perron Theorem (Stable manifold Theorem). Theorem 3.6.1. Let f : M → M be a C r diffeomorphism of a smooth manifold M, u s and let p be a hyperbolic fixed point of f with associated splitting Tp M = Ep ⊕ Ep . s then W s (p) is a C r injectively immersed copy of Ep and W u (p) is a C r injectively u s immersed copy of Ep . Moreover, W s (p) is tangent at p to Ep and W u (p) is tangent u at p to Ep . Now let us start by assuming f : R3 → R3 is C 2 and satisfies the following conditions. Let L = Df (0) = (Lu , Ls ), and R2 = E u ⊕ E s be the splitting given by the hyperbolicity of L, where E u and E s correspond to the x1 −axis and x2 x3 −plane respectively. So we write ˜ ˜ f (x, y) = (Lu x + f1 (x, y), Ls y + f2 (x, y)) = (f1 (x, y), f2 (x, y)) ˜ ˜ where f1 and f2 are the higher-order nonlinear terms of f . And let Bε0 be the box around the origin of radius ε0 , B1ε = Bε0 ∩x1 -axis, and B2ε = Bε0 ∩x2 x3 -plane. 0 0 89 And we assume f = (f1 , f2 , f3 ) satisfies the following conditions in Bε0 : 1. Lu = λ1 > 1, Ls = (λ2 , λ3 ), 0 < λ3 < λ2 < 1; ˜ ˜ 2. fi (0) = 0 and the first partial derivatives fix (0) = 0, for i, j = 1, 2, 3; j ˜ ˜ 3. |fix | ≤ ǫ and |fix x | ≤ M2 , for i, j, k = 1, 2, 3. j j k u We wish to find a C 2 function g : B1ε → B2ε whose graph is Wloc (0) and is 0 0 s invariant by f . In order to find another function whose graph is Wloc (0), one may set g : B2ε → B1ε . Since the work are essentially the same, we only state the 0 0 u results of g that corresponds to Wloc (0). By using the same graph transform method appeared in the previous section, it can be shown that there is such a function g and it is C 2 . As before, we set Γf (g) = f2 ◦ (1, g) ◦ [f1 ◦ (1, g)]−1 and G = {g : B1ε → B2ε | g(0) = 0, Lip(g) ≤ 1} 0 0 which is a complete metric space with the metric d(g1 , g2 ) = sup |g1 (x) − g2 (x)|. x∈B1ε 0 We start with g0 ≡ 0, then g1 = Γf (g0 ), and inductively, gn = Γf (gn−1 ), the sequence of which converges to some function g ⋆ ∈ G. Since the proof is essentially the same as is shown in the previous section, we do not repeat the work here. We 90 will adapt a few constants, and state the lemmas and theorem accordingly. As in the previous section, ˜ ˜ f (x, y) = (Lc x + f1 (x, y), Lss y + f2 (x, y)) = (f1 (x, y), f2 (x, y)) with m(Lc ) be its weak contraction eigenvalue, and |Lss | be its strong contraction eigenvalue. In this section, we have ˜ ˜ f (x, y) = (Lu x + f1 (x, y), Ls y + f2 (x, y)) = (f1 (x, y), f2 (x, y)) where m(Lu ) = Lu = λ1 , |Ls | = λ2 . Hence, we may adapt Lemma 3.5.7 to have Lemma 3.6.2. Given f satisfying the conditions at the beginning of this section, if λ1 − ǫ > 0 and (λ2 + ǫ)(λ1 − ǫ)−1 ≤ 1 in Bε0 , then Γf (g) is well-defined and maps G to itself. If, in addition, η = (λ2 + ǫ)(1 + ǫ )<1 λ1 − ǫ then Γf (g) is a contraction mapping with contraction µ. The unique fixed point of 91 Γf (g) is a C 0 function g ⋆ with |g ⋆ | ≤ η |g − g0 |. 1−η 1 We get the following Lemma based on Lemma 3.5.8. Lemma 3.6.3. Given f satisfying the conditions at the beginning of this section, if the following conditions are satisfied in Bε0 λ2 + 2ǫ λ2 + ǫ λ + 2ǫ <1 < 1 and +ǫ 2 λ1 − 2ǫ λ1 − 2ǫ (λ1 − 2ǫ)2 then the function g ⋆ in Lemma 3.5.7 is C 1 with |Dg ⋆ | ≤ ǫ . λ1 − λ2 − 3ǫ Subsequently, we adapt Lemma 3.5.9 into Lemma 3.6.4. Given f satisfying the conditions at the beginning of this section, in addition, each one of the second derivatives of f is bounded by some constant M2 > 0. If in Bε0 the following condition is satisfied µ= λ2 + ǫ (λ1 − 2ǫ)2 + ǫ(λ2 + 2ǫ) <1 (λ1 − 2ǫ)3 then we have g ⋆ is C 2 with |D 2 g ⋆ | ≤ 4M2 (λ1 + λ2 ) (λ1 − 2ǫ)3 (1 − µ) 92 . In summary, analogous to Theorem 3.5.1 we conclude the theorem as follows. Theorem 3.6.5. Given that f satisfies the above conditions, if in Bε0 we have the following conditions satisfied λ1 − ǫ > 0 , (λ2 + ǫ)(λ1 − ǫ)−1 ≤ 1 η = (λ2 + ǫ)(1 + ǫ )<1 λ1 − ǫ λ2 + 2ǫ λ2 + ǫ λ + 2ǫ <1 , +ǫ 2 <1 λ1 − 2ǫ λ1 − 2ǫ (λ1 − 2ǫ)2 µ= λ2 + ǫ (λ1 − 2ǫ)2 + ǫ(λ2 + 2ǫ) <1 (λ1 − 2ǫ)3 u then we have a unique C 2 function g ⋆ : B1ε → B2ε , whose graph is Wloc (0) of 0 0 f . Moreover, we have |g ⋆ | ≤ ε0 |Dg ⋆| |Dg ⋆| ≤ |D 2 g ⋆ | ≤ ǫ λ1 − λ2 − 3ǫ 4M2 (λ1 + λ2 ) (λ1 − 2ǫ)3 (1 − µ) . After we have obtained C 2 parameterizations of both local stable and unstable manifolds of f at 0, which are now denoted by gs , and gu respectively. Then, in s u general, Wloc (0) (respectively Wloc (0)) is not in E s (respectively E u ). However, 93 we will show there is a C 2 diffeomorphism f which conjugates to f and has both local stable and unstable manifolds straightened. Let us consider the map ρ : Bε0 → V ⊂ R3 with ρ(x, y) = (x − gs (y), y − gu (x)). (3.77) By this definition, ρ is C 2 , ρ(0) = 0 as gs (0, 0) = gu (0) = 0, and Dρ(0) is the identity     ′ −gs (0) 1 0  1 Dρ(0) =  =  ′ −gu (0) 1 0 1 ′ ′ where we used the properties of the local manifolds that gs (0) = gu (0) = 0. Thus ρ is a diffeomorphism in some neighborhood of 0 ∈ R3 by the inverse function ′ theorem. In Bε0 ∩ f −1 (Bε0 ), such that ρ : Bε0 ∩ f −1 (Bε0 ) → V , let f = ρf ρ−1 . (3.78) ′ Then f : V → V is a C 2 diffeomorphism, with f (0) = 0 and D f (0) = L. Moreover, in Bε0 the local stable (respectively, unstable) manifold of f is in E s , i.e., x2 x3 −plane (respectively E u , i.e., x1 -axis). We can verify that property as follows: ρ(gs (y), y) = (gs (y) − gs (y), y − gu (gs (y))) = (0, y − gu (gs (y))) 94 and ρ(x, gu (x)) = (x − gs (gu (x)), gu (x) − gu (x)) = (x − gs (gu (x)), 0) Thus, up to a smooth conjugacy, the local manifolds of f is straightened. 95 BIBLIOGRAPHY 96 BIBLIOGRAPHY [1] John Guckenheimer and Philip Holmes. Nonlinear Oscillations, Dynamical Systems, and Bifurcation of Vector Fields, volume 42 of Applied Mathematical Sciences. Springer-Verlap, New York, 1983. [2] P. Hartman. On local homeomorphisms of euclidean spaces. Bol. Soc. Mat. Mexicana, 5:220–241, 1960. [3] P. Hartman. Ordinary Differential Equations, 2nd ed., volume 38 of Siam Classics in Applied Math. SIAM, 2002. [4] M. Hirsch and C. Pugh. Stable manifolds and hyperbolic sets. Proc. AMS Symp. Pure Math., 14, 1970. [5] J. Palis and F. Takens. Hyperbolicity and Sensitive Chaotic Dynamics at Homoclinic Bifurcations., volume 35 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1993. [6] Clark Robinson. Dynamical Systems, Stability, Symbolic Dynamics, and Chaos, second ed. Studies in Advanced Mathematics. CRC Press, 1999. [7] Colin Sparrow. The Lorenz Equations: Bifurcations, Chaos, and Strange Attractors, volume 41 of Applied Mathematical Sciences. Springer Verlag, New York, 1982. [8] S. Sternberg. Local Contractions and a Theorem of Poincar´. American Journal e of Mathematics, 79(4):809–824, October 1957. [9] W. Tucker. A rigorous ODE solver and Smale’s 14th problem”. Foundations of Computational Mathematics, 2:53–117, 2002. 97