Dynamical_Systems_course_notes (16)
Dynamical_Systems_course_notes (16)
Dynamical Systems
Math 637
A gentle introduction to dynamical systems and flows
R. Vandervorst ∗
May 5, 2021
VU Lecture Notes
Disclaimer
You can edit this page to suit your needs. For instance, here we have a no copyright state-
ment, a colophon and some other information. This page is based on the corresponding
page of Ken Arroyo Ohori’s thesis, with minimal changes.
No copyright
c z This book is released into the public domain using the CC0 code. To the extent
possible under law, I waive all copyright and related or neighbouring rights to this work.
To view a copy of the CC0 code, visit:
http://creativecommons.org/publicdomain/zero/1.0/
Colophon
This document was typeset with the help of KOMA-Script and LATEX using the kaobook
class.
The source code of this book is available at:
https://github.com/fmarotta/kaobook
(You are welcome to contribute!)
Publisher
First printed in May 2019 by VU Lecture Notes
I believe in intuitions and inspirations. I sometimes feel that
I am right. I do not know that I am.
– Albert Einstein
Contents
Contents v
Class material 1
1 Linear systems
Lectures 1 and 2 2
1.1 Diagionalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Exponentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Classifying eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Inhomogeneous equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5 Limit sets
Lecture 7 and 8 27
5.1 Global flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2 𝜔 -limit sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3 Attractors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6 Poincaré sections
Lecture 8 33
6.1 Poincaré maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.2 The flow box theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
8 Index theory
Lecture 11 40
8.1 Winding numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
8.2 The index of a regular Jordan curve . . . . . . . . . . . . . . . . . . . . . . 42
8.3 The fixed point index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
8.4 * Homotopy properties of the index . . . . . . . . . . . . . . . . . . . . . . 46
9 Compactification
Lecture 12 and 13 47
9.1 Stereographic and central projections . . . . . . . . . . . . . . . . . . . . . 47
9.2 Extending the vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
9.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
A 1-dimensional system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
A planar system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
A 3-dimensional vector field . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
9.4 Blowing up fixed points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
10 Bifurcations
Lecture 14 and 15 57
Appendix 58
A Elememtary calculus 59
A.1 Implicit Function Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
B Elememtary topology 60
Bibliography 61
Notation 62
Alphabetical Index 63
Class material
Linear systems
Lectures 1 and 2 1
In this chapter we consider linear systems of differential equarions with 1.1 Diagionalization . . . . . . . . 2
constant coefficients. We describe a systematic approach to determine the 1.2 Exponentials . . . . . . . . . . . 3
general solution. Most of this chapter is dedicated to the linear algebra of 1.3 Classifying eigenvalues . . . . 5
1.4 Inhomogeneous equations . . 6
exponential of square matrices.
1.1 Diagionalization
𝑥¤ = 𝑎𝑥 + 𝑏 𝑦
𝑦¤ = 𝑐𝑥 + 𝑑𝑦
𝑎 𝑏
𝐴= .
𝑐 𝑑
In the case that 𝑏 = 𝑐 = 0 the system decouples and the general solution
is given by x(𝑡) = (𝑥 0 𝑒 𝑎𝑡 , 𝑦0 𝑒 𝑑𝑡 ). In general a system is not decoupled and
we describe a method to determine then general solution.
A linear system of differential equations with 𝑛 variables x = (𝑥 1 , · · · , 𝑥 𝑛 ) ∈
ℝ 𝑛 is given as follows:
𝑎11 ··· 𝑎 1𝑛
© . .. ..
𝐴 = ..
ª
x¤ = 𝐴x , . .
®.
® (1.1)
« 𝑎𝑛1 ··· 𝑎 𝑛𝑛 ¬
𝜆1 ··· 0
© .. .. .. ª
𝐴𝐸 = 𝐸𝐷, with 𝐸 = (v1 · · · v𝑛 ) and 𝐷 = . . .
®.
®
« 0 ··· 𝜆𝑛 ¬
The eigenvectors form a basis for ℝ 𝑛 and therefore form linearly inde-
pendent columns of the square matrix 𝐸 which is therefore invertible.
This yields the following matrix decompositions:
y¤ = 𝐸 −1 x¤ = 𝐸 −1 𝐴x = 𝐸 −1 𝐴𝐸 y = 𝐷 y ,
𝑒 𝜆1 𝑡 ··· 0
© .. .. .. ª
y(𝑡) = 𝐷(𝑡)y(0), with 𝐷(𝑡) = . . .
®.
®
« 0 ··· 𝑒 𝜆𝑛 𝑡 ¬
We can now transform back to the vector x:
Remark 1.1.1 The above expression gives the general solution with
the initial vector x(0) as input. If we regard y(0) as an arbitrary vector
of points y(0) = (𝑐 1 , · · · , 𝑐 𝑛 ) then the general solution expresses as:
x(𝑡) = 𝑐 1 v1 𝑒 𝜆1 𝑡 + · · · + 𝑐 𝑛 v𝑛 𝑒 𝜆𝑛 𝑡 ,
Examples of computing x(𝑡), see [1], pp. 7-8. Unfortunately square ma-
trices 𝐴 are not always diagonalizable. In the next section we start a
systematic approach to also deal with non-diagonalizable matrices.
1.2 Exponentials
follows:
𝑎 11 · · · 𝑎1𝑛
© . .. .. ª®
x¤ = 𝐴x , 𝐴 = ..
. . ®,
« 𝑎 𝑛 1 · · · 𝑎 𝑛𝑛 ¬
we can postulate a similar formula. If in the one dimensional case we
regard 𝑎 as a 1 × 1 matrix 𝐴 = (𝑎), then 𝑥(𝑡) = 𝑒 𝑡𝐴 𝑥(0). However, if 𝐴 is
an 𝑛 × 𝑛 matrix we need to know if 𝑒 𝑡𝐴 is a well-defined matrix.3 3: An informal Taylor expansion gives:
P∞ 𝑡 𝑘 (𝑘)
x(𝑡) = 𝑘=0 𝑘 !
Moreover, x(𝑘) (0) =
x (0).
𝐴 𝑘 x(0). Upon substitution this yields
Definition 1.2.1 Let 𝐴 be a real 𝑛 × 𝑛 matrix. Define the exponential of P∞ 𝑡𝑘 𝑘
x(𝑡) = 𝐴 x(0).
𝐴 as: 𝑘=0 𝑘 !
∞
1 𝑘
𝑒 𝐴 x :=
X
𝐴 x. (1.3)
𝑘=0
𝑘 !
The next step is to argue that 𝑒 𝐴 is a well-defined matrix. For the vector
1 Linear systems
4
Lectures 1 and 2
We conclude that 𝑒 𝑡𝐴 x(0) is well-defined for all 𝑡 ∈ ℝ and for all x(0) ∈ ℝ 𝑛 .
We now propose x(𝑡) = 𝑒 𝑡𝐴 x(0) as solution for the system x¤ = 𝐴x. The
vector valued power series 𝑒 𝑡𝐴 x(0) is absolutely convergent and therefore
differentiable in 𝑡 .
Lemma 1.2.1
𝑑 𝑡𝐴
𝑒 x(0) = 𝐴𝑒 𝑡𝐴 x(0),
𝑑𝑡
for all ∈ ℝ and for all x(0) ∈ ℝ 𝑛 .
𝑑 𝑡𝐴 ∞
𝑡 𝑘−1 ∞
𝑡 𝑘−1
𝐴 𝑘 x(0) = 𝐴 𝑘−1 𝐴x(0) = 𝑒 𝑡𝐴 𝐴x(0).
X X
𝑒 x(0) =
𝑑𝑡 𝑘=1
(𝑘 − 1)! 𝑘=1
(𝑘 − 1)!
I let 𝐴1 = 𝐼 and
𝐴 𝑘 = 𝐴 − 𝜆 𝑘−1 𝐼 𝐴 𝑘−1 ;
Then,
𝑒 𝑡𝐴 = 𝑎 1 (𝑡)𝐴1 + · · · + 𝑎 𝑛 (𝑡)𝐴𝑛 . (1.5)
𝑛 𝑛 𝑛 𝑛
𝑑 X X X X
𝑎 𝑘 𝐴𝑘 = 𝑎¤ 𝑘 𝐴 𝑘 = 𝜆𝑘 𝑎 𝑘 𝐴𝑘 + 𝑎 𝑘−1 𝐴 𝑘 . (1.6)
𝑑𝑡 𝑘=1 𝑘=1 𝑘=1 𝑘=1
If the matrix 𝐴 is invertible then the eigenvalues have the property that
R 𝑒 𝜆 ≠ 0 — the real part of 𝜆. In this case we also have that x∗ = 0
is the only zero of the right hand side of x¤ = 𝐴x. The point x∗ = 0 is
called an equilibrium point for the system. Only when det 𝐴 = 0 there are
more equilibrium points. For equilibrium points we use the following
classification based on eigenvalues:
Classification of eigenvalues
The subspace are invariant, i.e. if an initial value x(0) is contained in such
a linear subspace, then 𝑒 𝑡𝐴 x(0) is in the same subspace for all 𝑡 ∈ ℝ.
Moreover,7 7: These properties follow from Remark
ℝ𝑛 = 𝐸 𝑠 ⊕ 𝐸 𝑐 ⊕ 𝐸𝑢 . 1.1.1.
When 𝜆 𝑘 < 0 then the center subsapce is trivial. This definition is more
involved in the case that 𝐴 is not diagonalizable. In order to find the
stable, unstable and center subspaces in general we consider the following
example. For the matrix
1 0 0
𝐴= 1 1 0 ®,
© ª
« − 1 − 1 −2 ¬
1 0 0 0 0 0
𝑒 𝑡𝐴 = 𝑒 𝑡 𝑡 1 0 ® + 𝑒 −2 𝑡 0 0 0 ® = 𝑒 𝑡 𝐴 1 + 𝑒 2𝑡 𝐴 2 .
© ª © ª
« 𝑡+2 1 0 ¬ « −2 −1 1 ¬
x¤ = 𝐴x + f ,
where f = f(𝑡) = ( 𝑓1 (𝑡), · · · , 𝑓𝑛 (𝑡)). For a given initial value x(0) we can
represent the solution with the following formula.
1 Linear systems
7
Lectures 1 and 2
𝑑
Proof. Consider the expression 𝑒 −𝑡𝐴 x(𝑡) . Upon differentiating, sub-
𝑑𝑡
stituting the equation and using Property (1.4) we obtain:
𝑑 −𝑡𝐴
𝑒 x(𝑡) = −𝐴𝑒 −𝑡𝐴 x(𝑡) + 𝑒 −𝑡𝐴 x¤ (𝑡)
𝑑𝑡
= −𝐴𝑒 −𝑡𝐴 x(𝑡) + 𝑒 −𝑡𝐴 𝐴x(𝑡) + 𝑒 −𝑡𝐴 f(𝑡) = 𝑒 −𝑡𝐴 f(𝑡).
2.1 Existence
We prove Theorem 2.0.1 via Picard iteration.2 To set up the problem 2: We follow the approach in cf. [2].
suitable for Picard iteration we first reformulate the equation. Let x : 𝐼 ⊂
ℝ → ℝ 𝑛 be a solution of (2.1). Upon integration of the equation we
obtain: ∫ 𝑡
f x(𝑠) 𝑑𝑠.
x(𝑡) = x0 + (2.2)
0
With Picard iteration the idea is to start with a function x0 (𝑡) = x0 and
compute the next function via the scheme
∫ 𝑡
f x 𝑘 (𝑠) 𝑑𝑠.
x 𝑘+1 (𝑡) = x0 + (2.3)
0
𝐵 𝜖 (x0 ) = x | k x − x0 k ≤ 𝜖 .
𝐾 = max k𝐷 f(x)k. (2.4)
x∈𝐵 𝜖 (x0 )
Proof of Theorem 2.0.1 (Existence part). Choose a ball 𝐵 𝜖 (x0 ) which by Lemma
2.1.1 yields a constant 𝐾 = 𝐾(x0 , 𝜖) > 0 such that k f(x) − f(y)k ≤ 𝐾k x − y k
for all x , y ∈ 𝐵 𝜖 (x0 ).5 Define 5: The constant 𝜖 > 0 is fixed for the re-
mainder of the proof.
𝑀 = max k f(x)k.
x∈𝐵 𝜖 (x0 )
Start with x0 (𝑡) = x0 and suppose the iteration scheme yields a continuous
function x 𝑘 (𝑡) such that
The next iterate x 𝑘+1 (𝑡) is also defined by the continuity of f and7 8 7: Uses condition on 𝜏.
8: If both 𝐾 = 𝑀 = 0, then 𝜏 can be any
∫ 𝑡 positive number. Indeed, 𝐵 𝜖 (x0 ) consists
k f x 𝑘 (𝑠) k𝑑𝑠 ≤ 𝑀𝜏 < 𝜖,
k x 𝑘+1 (𝑡) − x0 k ≤ of zeroes which are constant solutions that
0 exist for all time 𝑡 . We therefore assume
without loss of generality that 𝐾 ≠ 0 and
for all 𝑡 ∈ [−𝜏, 𝜏]. For the iteration scheme we may therefore assume that 𝑀 ≠ 0.
and thus
𝑚−
X1 𝑚−
X1 ∞
𝛾𝑘 ≤ 𝜖 𝛾𝑘
X
k x𝑚 (𝑡) − x𝑚0 (𝑡)k ≤ k x 𝑘+1 (𝑡) − x 𝑘 (𝑡)k ≤ 𝜖
𝑘=𝑚 0 𝑘=𝑚 0 𝑘=𝑚0
𝛾 𝑚0
=𝜖 → 0,
1−𝛾
The above proof only deals with the existence of a time interval for a
given initial point x0 . However, the proof also reveals how the constant
𝜏 = 𝜏(x0 ) behaves for x̃ in a neighborhood of x0 . Indeed, if we consider
balls 𝐵 𝜖/2 (x̃) with x̃ ∈ 𝐵 𝜖/2 (x0 ), then 𝐵 𝜖/2 (x̃) ⊂ 𝐵 𝜖 (x0 ) for all x̃ ∈ 𝐵 𝜖/2 (x0 )
by the triangle inequality. If we repeat the proof for x̃ as initial point we
obtain
𝐾˜ = max k𝐷 f(x)k ≤ max k𝐷 f(x)k = 𝐾,
x∈𝐵 𝜖/2 (x̃) x∈𝐵 𝜖 (x0 )
and similarly
to the initial value problem exists for all initial values x̃(0) = x̃ ∈ 𝐵 𝜖/2 (x0 ),
which establishes a uniform time interval for all x̃ ∈ 𝐵 𝜖/2 (x0 ).
2 The existence and uniqueness theorem
11
Lecture 3
The existence proof given above also works for vector fields that are
focally Lipschitz which is a larger class of functions than 𝐶 1 -vector fields.
We can even prove a theorem for continuous vector fields.
2.2 Uniqueness
√
Consider the differential equation 𝑥¤ = 𝑥 , 𝑥 ≥ 0 and 𝑥(0) = 0. Direct
verification shows that
(
0 for 𝑡 ≤ 𝑡0
𝑥(𝑡) = 1
4 (𝑡 − 𝑡0 )2 for 𝑡 ≥ 𝑡0 ,
are solutions of the initial value problem for all 𝑡0 ≥ 0. This those gives
√
infinitely many solutions. Note that the vector field 𝑓 (𝑥) = 𝑥 is a not
differentiable at 𝑥 = 0. The vector field is continuous however.
Proof of Theorem 2.0.1 (Uniqueness part). Suppose there exist two solu-
tions x , y : [−𝜏, 𝜏] → ℝ 𝑛 satisfying (2.1) with x(0) = y(0) = x0 . For
the difference we have, using the representation in (2.2):
∫ 𝑡
k x(𝑡) − y(𝑡)k ≤ f(x(𝑠)) − f(y(𝑠)) 𝑑𝑠
0
≤ 𝐾𝜏 max k x(𝑠) − y(𝑠)k < k x − y k𝐶 0 ,
𝑠∈[0,𝑡]
The uniqueness follows from the local Lipschitz property of the vector
field f. It is clear such a condition is not satisfied for the example above.
If we stay away from 𝑥 = 0 then the existence and uniqueness theorem
again provides a unique solution.
The uniqueness part of Theorem 2.0.1 has an immediate consequence
for the solution x(𝑡) when represented as curves in ℝ 𝑛 . Let x : 𝐼 → ℝ 𝑛
and y : 𝐽 → ℝ 𝑛 be solutions. Suppose that x(𝑡1 ) = y(𝑡2 ) and define
2 The existence and uniqueness theorem
12
Lecture 3
x̃(𝑡) := x(𝑡 − 𝑡1 ) and ỹ(𝑡) := y(𝑡 − 𝑡2 ). Then, from the uniqueness part of
Theorem 2.0.1 we conclude that z(𝑡) := 𝑥(𝑡) ˜ − 𝑦(𝑡)
˜ ≡ 0 in a neighborhood
of 𝑡 = 0. Since the function z is defined on the interval 𝐼˜ ∩ 𝐽˜, where 𝐼˜ and
𝐽˜ are the shifted time intervals the conclusion holds for all of 𝐼˜ ∩ 𝐽˜.
Two solutions x and y cross at x0 = x(𝑡1 ) = y(𝑡2 ) if z(𝑡) . 0 on 𝐼˜ ∩ 𝐽˜.
From the above argument it follows that two cannot cross. In particular
solutions cannot self-intersect. The latter leave the possibility for a
periodic solutions, i.e. x(𝑡 + 𝑡1 ) = x(𝑡) for all 𝑡 ∈ ℝ.
Theorem 2.3.1 Consider the open ball 𝐵 𝜖 (x0 ) and solutions y(𝑡), ỹ(𝑡) ∈
𝐵 𝜖 (x0 ) for 𝑡 ∈ [𝑡0 , 𝑡1 ]. Then,
∫𝑡
Proof. First consider the case 𝑐 0 > 0 and let w(𝑡) = 𝑐 0 + 0 𝑐 1 z(𝑠)𝑑𝑠 > 0.
By assumption z(𝑡) ≤ w(𝑡) and w ¤ (𝑡) = 𝑐 1 z(𝑡). This yields
¤ (𝑡) 𝑐 1 z(𝑡)
w
= ≤ 𝑐1 ,
w(𝑡) w(𝑡)
𝑑
and hence 𝑑𝑡 log w(𝑡) ≤ 𝑐 1 . Upon integration this gives
Proof. Let y(𝑡) = x(𝑡 ; x0 ), with y(𝑡0 ) = x0 and ỹ(𝑡) = x(𝑡 ; x1 ), with ỹ(𝑡0 ) =
x1 . By the continuity of x(𝑡 ; x0 ) as function of 𝑡 we have that for every
𝜖 > 0 there exists a 𝛿 1𝜖 > 0 such that 0 < |𝑡 − 𝑡0 | < 𝛿1𝜖 implies that
k x(𝑡 ; x0 ) − x0 k < 𝜖/2. Moreover, by Theorem 2.3.1 we have that
1
Choose 𝛿 𝜖 = min 𝜖𝑒 −𝐾𝛿 𝜖 /2 , 𝛿 1𝜖 . Then 0 < k x1 − x0 k + |𝑡 − 𝑡0 | < 𝛿 𝜖
3.1 Flows
x¤ = f(x), x(0) = x0 ∈ ℝ 𝑛 .
Lemma 3.1.1 For 𝑡 ∈ [−𝜏, 𝜏] and 𝑠 ∈ [−𝜎, 𝜎], with 𝜏 = 𝜏(x0 ) > 0 and
𝜎 = 𝜎 x(𝑡 ; x0 ) , it holds that
x(𝑠 + 𝑡 ; x0 ) = x 𝑠, x(𝑡 ; x0 ) .
(3.1)
The function y(𝑟) is a 𝐶 1 -function on (−𝜏, 𝑡 + 𝑠], 𝑠 < 𝜎 and satisfies the
differential equation y¤ = f(y)1 with y(0) = x0 . Since solutions are unique 1: Here we use the fact that f does not
by Theorem 2.0.1 we conclude that depend on 𝑡 explicitly.
which proves the statement for 𝑠 > 0. For 𝑠 = 0 the statement is trivial
and for 𝑠 < 0 we argue in the same way.
We refer (3.1) as the (local) group property for the solution function.
Remark 3.1.1 The fact that the group property extends the domain of
definition of x at x0 does not contradict the finite time blow-up of for
example the equation x¤ = x2 .
3 Flows, invariance and linearization
15
Lecture 4
We refer to 𝜙𝑡 as the local flow for the vector field f. A local flow depends in
a continuously differentiable way on 𝑡 , i.e.
The fact that flows are not defined globally in 𝑡 ∈ ℝ is not much of a
restriction. We will see later on that one can always study vector fields
via a normalization that yields a flow defined for all 𝑡 ∈ ℝ. Also when
we consider invariant sets the time variable will be global in most cases
as we will see in the next section.
If we consider a given local flow 𝜙𝑡 then there exists an associated vector
field: f(x) = 𝜙¤ 𝑡 (x)|𝑡=0 . The vector field is defined on all of ℝ 𝑛 and is the
velocity field of the flow lines. A given 𝐶 1 -vector field on ℝ 𝑛 yields a local
flow on ℝ 𝑛 as described above. Conversely, a local flow 𝜙𝑡 : ℝ 𝑛 → ℝ 𝑛
with the smoothness properties as described above yields a vector field
and thus are solutions of the differential equation in (2.1).
Proof. We start with the observation that there exists a 𝜏 > 0 such that
𝜙𝑡 |𝑆 is defined for all 𝑡 ∈ [−𝜏, 𝜏].5 Invariance of 𝑆 implies that 𝜙𝑡 (𝑆) ⊂ 𝑆 Due to the compactness of 𝑆. Indeed, let
5:
for all 𝑡 ∈ [−𝜏, 𝜏]. In particular, 𝐵 𝜖 (x) x∈𝑆 be an open covering for 𝑆 with
𝜖 > 0 fixed. By the compactness
of 𝑆 there
𝑛
exists a finite sub-covering 𝐵 𝜖 (x 𝑘 ) 𝑘=1
𝜙 𝑘𝑡 (𝑆) = 𝜙𝑡 ◦ · · · ◦ 𝜙𝑡 (𝑆) ⊂ 𝑆, 𝑘 ∈ ℤ,
of 𝑆 and a finite set of times 𝜏(x 𝑘 ) > 0.
Choose 𝜏 = min 𝑘 𝜏(x 𝑘 ).
| {z }
𝑘 times
which proves that 𝜙𝑡 (𝑆) ⊂ 𝑆 for all 𝑡 ∈ ℝ. Using the fact that 𝜙𝑡 is
invertible gives that 𝑆 ⊂ 𝜙−𝑡 (𝑆) for all 𝑡 ∈ ℝ and thus 𝜙𝑡 (𝑆) = 𝑆 for all
𝑡 ∈ ℝ. Conversely, if 𝜙𝑡 (𝑆) = 𝑆 for all 𝑡 ∈ ℝ, then for every x ∈ 𝑆 it holds
that 𝜙𝑡 (x) ∈ 𝑆 for all 𝑡 ∈ ℝ and thus for all 𝑡 ∈ [−𝜏, 𝜏] which proves
invariance.
The same result can be proved for forward and backward invariant sets,
i.e. a compact set 𝑆 ⊂ ℝ 𝑛 is forward invariant if and only if 𝜙𝑡 (𝑆) ⊂ 𝑆 for
all 𝑡 ≥ 0 and a compact set 𝑆 ⊂ ℝ 𝑛 is backward invariant if and only if
𝜙𝑡 (𝑆) ⊂ 𝑆 for all 𝑡 ≤ 0.
The simplest example of an invariant sets is an equilibrium point, i.e. a
point x∗ such that f(x∗ ) = 0. Equilibrium points are also referred to as
fixed points6 since they satisfy 𝜙𝑡 (x∗ ) = x∗ for all 𝑡 ∈ ℝ. The latter shows 6: We alternate between the terminology
that equilibrium point are invariant sets for the flow 𝜙𝑡 . An equilibrium equilibrium point and fixed point.
(i) 𝜙𝑇 (x• ) = x• ;
(ii) 𝜙𝑡 (x• ) ≠ x• , ∀𝑡 ∈ (0 , 𝑇).
The perodic orbit is also denoted by 𝛾(𝑡) = 𝜙𝑡 (x• ).
a The number 𝑇 in this definition is called the (minimal) period of 𝛾 .
exists a 𝑇 > 0 such that 𝜙𝑇 (x• ) = x• . Since x• is not an equilibrium point 8: This implies that 𝑓 (x• ) ≠ 0
it follows that 𝜙𝑡 (x• ) ≠ x• for some 𝑡 . Moreover, 𝜙𝑡+𝑇 (x• ) = 𝜙𝑡 (𝜙𝑇 (x• )) =
𝜙𝑡 (x• ) for all 𝑡 . By defining 𝑇0 = inf {𝑡 > 0 | 𝜙𝑡 (x• ) ≠ x• }, we satisfy
Defn. 3.2.2(ii) which shows that periodic points generate periodic orbits.
Conversely, every point x• ∈ 𝛾 of a periodic orbit is a periodic point.
3.3 Linearization
discuss the first step for analyzing equilibrium points. For an equilibrium
point x∗ we have the following Taylor expansion:
= 𝐷 f(x∗ )(x − x∗ ) + 𝑜 k x − x∗ k ,
Re𝜆 𝑖 ≠ 0 , ∀𝑖 = 1, · · · , 𝑛.
x¤ = f(x), x(0) = x0 ∈ ℝ 𝑛 ,
𝝃¤ = 𝐷 f(x∗ )𝝃.
Suppose the fixed point x∗ is hyperbolic, i.e. the eigenvalues 𝜆 𝑖 of 𝐷 f(x∗ )
satisfy Re𝜆 𝑖 ≠ 0 for all 𝑖 . Moreover precisely 𝐷 f(x∗ ) has 𝑘 eigenvalues
with Re𝜆 𝑖 < 0 (stable eigenvalues) and 𝑛 − 𝑘 eigenvalues with Re𝜆 𝑖 > 0
(unstable eigenvalues). From Section 1.3 this yields the decomposition of
ℝ 𝑛 in terms of stable and unstable invariant subspaces:
ℝ𝑛 = 𝐸 𝑠 ⊕ 𝐸𝑢 ,
Before we start the proof of the stable manifold theorem we consider some
properties of the linearization. The Taylor expansion about x∗ reads
where g(x) := f(x) − 𝐷 f(x∗ )(x − x∗ ), with g(x∗ ) = 0 and 𝐷 g(x∗ ) = 0. Since
g ∈ 𝐶 1 (ℝ 𝑛 ; ℝ 𝑛 ) we have from Taylor’s theorem that
x , x0 ∈ 𝐵 𝛿 𝜖 (x∗ ).
Let 𝐴 = 𝐷 f(x∗ ) then for the linear flow 𝑒 𝑡𝐴 it holds that 𝑒 𝑡𝐴 𝐸 𝑠 = 𝐸 𝑠 and
𝑒 𝑡𝐴 𝐸 𝑢 = 𝐸 𝑢 for all 𝑡 ∈ ℝ. From the expression
𝑒 ℎ𝐴 x − x
lim = 𝐴x ,
ℎ→0 ℎ
𝑃
0
𝐷 = 𝑉 −1 𝐴𝑉 = , (4.2)
0 𝑄
where 𝑃 is an 𝑘 × 𝑘 matrix and 𝑄 is an (𝑛 − 𝑘) × (𝑛 − 𝑘) matrix. We use
𝑉 to define new coordinates. Define
y := 𝑉 −1 (x − x∗ ),
y¤ = 𝐷 y + h(y), (4.3)
y¤ = 𝑉 −1 𝐴(x − x∗ ) + 𝑉 −1 g(x)
= 𝑉 −1 𝐴𝑉 y + 𝑉 −1 g 𝑉 y + x∗ .
for all y , y0 ∈ 𝐵 𝛿 𝜖 (0).3 The local flow associated with Equation (4.3) is 3: The Lipschitz estimate can also be ob-
denoted by 𝜓𝑡 . tained form the Lipschitz estimate for the
function g.
eigenvalues as follows:
𝑒 𝑡𝑃 𝑒 𝑡𝑃
𝑡𝐷 0 0 0 0
𝑒 = = + = 𝑈(𝑡) + 𝑉(𝑡).
0 𝑒 𝑡𝑄 0 0 0 𝑒 𝑡𝑄
Re𝜆1 ≤ · · · ≤ Re𝜆 𝑘 < −(𝛼 + 𝜎) < 0 < 𝜎 < Re𝜆 𝑘+1 ≤ · · · ≤ Re𝜆𝑛 .
𝑒 −𝛼𝑡 k a k
k yℓ (𝑡 ; a) − yℓ −1 (𝑡 ; a)k ≤ 𝑐 0 , ∀𝑡 ≥ 0 , (4.6)
2ℓ −1
and for all ℓ = 1 , · · · , 𝑚 . Now consider k y𝑚+1 (𝑡 ; a) − y𝑚 (𝑡 ; a)k . Using the
Lipschitz condition for h in (4.4) we have
∫ 𝑡
k y𝑚+1 (𝑡 ; a) − y𝑚 (𝑡 ; a)k ≤ 𝜖 k𝑈(𝑡 − 𝑠)k k y𝑚 (𝑠 ; a) − y𝑚−1 (𝑠 ; a)k𝑑𝑠
0
∫ ∞
+𝜖 k𝑉(𝑡 − 𝑠)k k y𝑚 (𝑡 ; a) − y𝑚−1 (𝑡 ; a)k𝑑𝑠,
𝑡
(4.7)
provided k y𝑚 (𝑡 ; a) − y𝑚−1 (𝑡 ; a)k < 𝛿 𝜖 for all 𝑡 ≥ 0. 5
If we use the 5: We need to choose the bounds on k a k
and 𝜖 to justify the estimate.
4 The (un)stable manifold theorem
21
Lectures 5 and 6
𝑒 −𝛼𝑡 k a k
k yℓ +1 (𝑡 ; a) − yℓ (𝑡 ; a)k ≤ 𝑐 0 , ∀𝑡 ≥ 0 , (4.11)
2ℓ
and for all ℓ ≥ 0. As in the proof of Theorem 2.0.1 consider integers
𝑚, 𝑚 0 ≥ 𝑚0 . Then, using the triangle inequality we have
𝑚−
X1 𝑚−
X1 1
k y𝑚 (𝑡 ; a) − y𝑚0 (𝑡 ; a)k𝑒 𝛼𝑡 ≤ k yℓ +1 (𝑡 ; a) − yℓ (𝑡 ; a)k𝑒 𝛼𝑡 ≤ 𝑐 0 k a k
ℓ =𝑚 0 ℓ =𝑚 0 2ℓ
∞
X 1 1
≤ 𝑐0 k a k = 𝑐0 k a k → 0,
ℓ =𝑚0 2ℓ 2𝑚0 −1
𝑦 𝑖 (0; a) = 𝜎𝑖 (𝑎 1 , · · · , 𝑎 𝑘 ), 𝑖 = 𝑘 + 1 , · · · , 𝑛,
∫∞
where 𝜎𝑖 (𝑎 1 , · · · , 𝑎 𝑘 ) = −𝜋 𝑖 0 𝑉(−𝑠)h y(0; a) 𝑑𝑠 , for 𝑖 = 1 , · · · , 𝑛 − 𝑘 ,
which we denote by 𝝈(𝜋 𝑠 a). As is the proof of Theorem 2.0.1 we have
that y(𝑡 ; a) continuously on 𝑡 and a and therefore the function 𝜎𝑖 depend
continuously on a.8 This parametrization defines a parametrized 𝑘 - 8: The continuity follows from the for-
dimensional surface (manifold)9 mula for 𝜎𝑖 using the continuity of and
the exponential decay of 𝑉 .
9: The parameter 𝜋 𝑠 a := (𝑎 1 , · · · , 𝑎 𝑘 ) is a
n o
𝑆 := 𝜋 𝑠 a , 𝝈(𝜋 𝑠 a) | 𝜋 𝑠 a ∈ 𝐵 𝛿𝑘 𝜖 /2𝑐0 (0) .
point in ℝ 𝑘 .
with decomposition
ℝ𝑛 = 𝐸𝑢 ⊕ 𝐸 𝑐 ⊕ 𝐸 𝑠 .
Theorem 4.2.2 (The center manifold theorem, II) There exists a locally
𝑐𝑢
defined 𝑛 − 𝑘 -dimensional surface 𝑊loc (x∗ ) ⊂ ℝ 𝑛 containing x∗ with tangent
𝑐 𝑢
space 𝐸 ⊕ 𝐸 at x∗ such that
𝑐𝑢 𝑐𝑢
(i) (backward invariance) 𝜙𝑡 𝑊loc (x∗ ) ⊂ 𝑊loc (x∗ ) for 𝑡 ≤ 0;
𝑐𝑢
(ii) (instability) lim𝑡→−∞ 𝜙𝑡 (x) = x∗ for all x ∈ 𝑊loc .
𝑐𝑢
The 𝑛 − 𝑘 -dimensional surface 𝑊loc (x∗ ) is given as graph over an open subset
𝑐 𝑢
in 𝐸 ⊕ 𝐸 and is called a local center-unstable manifold at x∗ .
𝑐 𝑐𝑠 𝑐𝑢
𝑊loc (x∗ ) := 𝑊loc (x∗ ) ∩ 𝑊loc (x∗ ). (4.13)
4 The (un)stable manifold theorem
24
Lectures 5 and 6
Remark 4.2.1 As for stable and unstable manifold we can also define
global version that are invariant sets for the flow.
The Grobman-Hartman theorem11 is a result for hyperbolic fixed points. 11: In many text books this theorem is
With the Grobman-Hartman theorem we establish local conjugations of a referred to as the Hartman-Grobman the-
orem.
flow. A local conjugation for a (local) flow 𝜙𝑡 at a hyperbolic fixed point x∗
to the linear flow 𝑒 𝑡𝐴 , 𝐴 = 𝐷 f(x∗ ), is a homeomorphism ℎ : 𝑈 ⊂ ℝ 𝑛 →
ℎ(𝑈) ⊂ ℝ 𝑛 , with ℎ(x∗ ) = 0, such that
Theorem 4.3.1 Let x∗ be a hyperbolic fixed point (2.1). Then, there exist a
local conjugacy for the associated local flow 𝜙𝑡 as described in (4.14).
We will not treat the proof of the Grobman-Hartman theorem in this text.
It was proves by Hartman that if the vector field is of class 𝐶 2 then a local
𝐶 1 -conjugacy exists. In this case we easily retrieve stable and unstable
manifolds from the local conjugacy.
p¤ = 𝜕q 𝐻(p , q);
(4.15)
q¤ = −𝜕p 𝐻(p , q),
𝐶2
where x = (p , q) ∈ ℝ2𝑛 , p , q ∈ ℝ 𝑛 . The function 𝐻 : ℝ2𝑛 −
−→ ℝ is called
the Hamiltonian. Since 𝐻 is of class 𝐶 2 the vector field
𝑑
𝐻 𝜙𝑡𝐻 = ∇𝐻 𝜙𝑡𝐻 · 𝑋𝐻 𝜙𝑡𝐻 = 0 ,
𝑑𝑡
4 The (un)stable manifold theorem
25
Lectures 5 and 6
which foliate the phase space. The dimension of a regular energy surface
is 2𝑛 − 1. A Hamiltonian system is an example of a conservative system.12
12: A conservative system is a system of
differential equations which allows a func-
Every motion is restricted to a given energy surface. The number 𝑛 in the tion 𝐸(x) which is constant along orbits.
Note that a conservative system is not nec-
dimension of the phase spave ℝ2𝑛 is counts the degrees of freedom pf
essarily Hamiltonian.
the Hamiltonian system.. In general Hamiltonian systems with 𝑛 ≥ 2 are
very complicated dynamical systems. However, for 𝑛 = 1 the systems are
what we call integrable. Let us consider 1-degree of freedom Hamiltonian
systems. In this case we obtain the system
𝑝¤ = 𝜕𝑞 𝐻(𝑝, 𝑞);
(4.16)
𝑞¤ = −𝜕𝑝 𝐻(𝑝, 𝑞),
𝐶2
with x = (𝑝, 𝑞) ∈ ℝ2 and 𝐻 : ℝ2 − −→ ℝ. By Lemma 4.4.1 the flow-lines of
the associated flow 𝜙𝑡𝐻 are curves given by the equation 𝐻(𝑝, 𝑞) = 𝐸 . If
we locally solve this equation at a regular point at an energy curve, say
𝑝(𝑞), then the dynamics is given by the equation
Since trace 𝐿 = 0 we have that the eigenvalues are given by the character-
istic equation 𝜆2 + det 𝐿 = 0.
Observe that det 𝐿 = det 𝑑 2 𝐻(x∗ ).14 A fixed point is a non-degenerate 14: The matrix 𝑑 2 𝐻(x∗ ) is the Hessian of
local minimum of 𝐻 if and only if 𝜕𝑝𝑝
2
𝐻(x∗ ) > 0 and det 𝑑2 𝐻(x∗ ) = det 𝐿 > 𝐻 and is the matrix of second derivatives
of 𝐻 .
0, and a non-degenerate local maximum if and only if 𝜕𝑝𝑝 2
𝐻(x∗ ) < 0 and
𝜕𝑝𝑝 𝐻(x∗ ) > 0 and det 𝑑 𝐻(x∗ ) = det 𝐿 > 0. A fixed point is a saddle for
2 2
𝑑
𝐻 𝜙𝑡 (x) = ∇𝐻 𝜙𝑡 (x) · −∇𝐻 𝜙𝑡 (x) = −k∇𝐻 𝜙𝑡 (x) k 2 ≤ 0 ,
𝑑𝑡
and < 0 whenever x ≠ x∗ , x∗ a critical point of 𝐻 . This implies that
𝐻 𝜙𝑡 (x) is strictly decreasing if we start at a point x that is not an
equilibrium point.
Gradient systems can be defined in any dimension 𝑛 and are not necessar-
ily complementary to Hamiltonian systems. A gradient system is defined
via a (smooth) potential function 𝑉 : ℝ 𝑛 → ℝ. Consider the system of
differential equations:
x¤ = −∇𝑉(x). (4.18)
For a gradient systems the nature of fixed points x∗ is determined by the
eigenvalues of the Hessian −𝑑 2 𝑉(x∗ ). Since −𝑑 2 𝑉(x∗ ) is always symmetric
non-degenerate fixed are either stable/unstable nodes, or saddle points.
Centers or spiral points, or combinations are excluded.
Limit sets
Lecture 7 and 8 5
In this chapter we study the asymptotic behavior of flows and in particular 5.1 Global flows . . . . . . . . . . 27
the notion of limit point. The most convenient way to address asymptotic 5.2 𝜔 -limit sets . . . . . . . . . . . 28
behavior is to describe dynamics by global flows. We start off with 5.3 Attractors . . . . . . . . . . . . 32
discussed global versus local flows.
f(x)
x¤ = , x(0) = x0 ∈ ℝ 𝑛 . (5.1)
1 + k f(x)k
Lemma 5.1.1 Let x0 ∈ ℝ 𝑛 . Then, the unique solution x(𝑡 ; x0 ) is defined for
all 𝑡 ∈ ℝ.
and thus
k x(𝑡 ; x0 )k < k x0 k + 𝑡. (5.2)
5 Limit sets
28
Lecture 7 and 8
Since this works for positive and negative times we have that k x(𝑡 ; x0 )k <
k x0 k + |𝑡 | . Given x0 and 𝜏(x0 ) inductively define
x 𝑘+1 = x 𝜏(x 𝑘 ); x 𝑘 , 𝑘 = 0, 1, · · · ,
where 𝜏(x 𝑘 ) > 0.2 Suppose k x 𝑘 k → ∞ as 𝑘 → ∞ and 𝜏 := 𝑘 𝜏(x 𝑘 ) < 2: The times 𝜏(x 𝑘 ) > 0 can be chosen by
P
∞. From (5.2) and the group property we obtain k x𝑛+1 k < k x0 k + virtue of Theorem 2.0.1.
P𝑛
𝑘=0
𝜏(x 𝑘 ) < k x0 k + 𝜏 < ∞, which is a contradiction. This implies that
k x(𝑡 ; x0 )k → ∞ only if 𝑡 → ∞, or 𝑡 → −∞, and thus x(𝑡 ; x0 ) exists for all
𝑡 ∈ ℝ.
f(x)
x̃0 = f(x̃), and x¤ = ,
1 + k f(x)k
𝑑x̃
where x̃0 = 𝑑𝑠 and define
∫ 𝑡
1
𝑠(𝑡) := 𝑑𝜎 > 0, 𝑡 ∈ ℝ. (5.3)
0 1 + k f x(𝜏; x0 ) k
The latter is well-defined for every 𝑡 ∈ ℝ since x(𝜏; x0 ) is defined for all
𝜏 ∈ ℝ by Lemma 5.1.1. Moreover, 𝑠(𝑡) is invertible and inverse is 𝑡(𝑠).
Now consider the function x† (𝑠 ; x0 ) := x 𝑡(𝑠); x0 and compute
𝑑 x† 𝑑 x 𝑑𝑡 h i
= x¤ 1 + k f x(𝑡 ; x0 ) k = f x(𝑡(𝑠); x0 ) = f x† (𝑠 ; x0 ,
=
𝑑𝑠 𝑑𝑡 𝑑𝑠
f(x)
x̃0 = f(x̃), and x¤ = ,
1 + k f(x)k
are topologically equivalent, i.e. the associated local flow 𝜙˜ 𝑠 and associated
global flow 𝜙𝑡 are topologically equivalent via time-reparametrization.
The method described here is not the only way to link the equation
x¤ = f(x) to a global flow. Since the orbits of the latter are obtained from
(5.1) via time-reparametrization the orbit structure of both equations is
exactly the same. We therefore, without loss of generality, may study the
global flow of (5.1).
The set of all 𝜔 -limit points of x is denoted 𝜔(x; 𝜙𝑡 ) and is called the
omega limit set, or 𝜔 -limit set. If there is no ambiguity with respect to
the flow 𝜙𝑡 we write 𝜔(x).
The set of all 𝜔 -limit points of 𝑈 is denoted 𝜔(𝑈) and is called the
omega limit set, or 𝜔 -limit set of 𝑈 .
As before we can also define 𝛼(𝑈), the alpha limit set of 𝑈 . As before
𝛼(𝑈 ; 𝜙𝑡 ) = 𝜔(𝑈 ; 𝜙−𝑡 ). The following lemma gives a use characterization
of omega limit sets, and therefore also alpha limit sets.
Proof. Fix 𝑡 > 0. Then, 𝜙𝑡𝑛 (x𝑛 ) ∈ 𝑠≥𝑡 𝜙 𝑠 (𝑈) for 𝑛 such that 𝑡 𝑛 ≥ 𝑡 .
S
Therefore, x0 ⊂ cl 𝑠≥𝑡 𝜙 𝑠 (𝑈). Since the latter is independent of 𝑡 we
S
obtain x0 ∈ 𝑡≥0 cl 𝑠≥𝑡 𝜙 𝑠 (𝑈).
T S
Now let x0 ∈ 𝑡≥0 cl 𝑠≥𝑡 𝜙 𝑠 (𝑈). Then, x0 ∈ cl 𝑠≥𝑡 𝜙 𝑠 (𝑈) for all 𝑡 ≥ 0.
T S S
We can choose point 𝑡 𝑛 → ∞ and 𝑥 𝑛 ∈ 𝑈 such that k𝜙𝑡𝑛 (x𝑛 ) − x0 k < 1/𝑛
as 𝑛 → ∞, which proves that x0 ∈ 𝜔(𝑈).
Next we give some properties of omega limit sets (and alpha limit sets)
that crucial in the application to flows.
Proposition 5.2.2 The omega limit set 𝜔(𝑈) is closed, invariant, and con-
tained in cl Γ+ (𝑈).a If 𝑈 ⊂ ℝ 𝑛 is forward invariant, then
\ \
𝜔(𝑈) = cl 𝜙𝑡 (𝑈) = 𝜙𝑡 (cl 𝑈), b (5.5)
𝑡≥0 𝑡≥ 0
!
\ \
𝜙(𝑠, 𝜔(𝑈)) = 𝜙 𝑠 cl Γ+𝑡 (𝑈) 𝜙 𝑠 cl Γ+𝑡 (𝑈)
=
𝑡≥ 0 𝑡≥ 0
\ \ \
cl 𝜙 𝑠 Γ+𝑡 (𝑈) cl Γ+𝑡+𝑠 (𝑈) = cl Γ+𝑡 (𝑈)
= =
𝑡≥ 0 𝑡≥ 0 𝑡≥𝑠
= 𝜔(𝑈),
which establishes the invariance of 𝜔(𝑈).4 4: Since forward images are nested it
holds that
In the case 𝑈 ⊂ 𝑋 is forward invariant, the group property implies,
cl Γ+ cl Γ+
\ \
𝑡 (𝑈) = 𝑡 (𝑈).
𝜙𝑡+𝑠 (𝑈) = 𝜙𝑡 𝜙 𝑠 (𝑈) ⊂ 𝜙𝑡 (𝑈) for all 𝑠, 𝑡 ≥ 0. Therefore, 𝑠≥𝑡 𝜙 𝑠 (𝑈) =
S
𝑡≥𝑠 𝑡≥ 0
𝜙𝑡 (𝑈), which completes the proof.
The same properties apply to alpha limit sets by reversing the time:
𝑡 ↦→ −𝑡 . In most applications we use the omega limit set of a compact, or
pre-compact5 set 𝑈 ⊂ ℝ 𝑛 . The following proposition provides a list of 5: Recall that a set is pre-compact in if cl 𝑈
properties in the case a compactness assumption is made. is a compact sets in ℝ 𝑛 .
Proof. For 𝑡 ≥ 𝜏, the sets cl 𝑠≥𝑡 𝜙 𝑠 (𝑈)) ⊂ cl Γ+𝜏 (𝑈) are compact, and
S
thus 𝜔(𝑈) ⊂ 𝑡≥𝜏 cl 𝑠≥𝑡 𝜙 𝑠 (𝑈)) ⊂ cl Γ+𝜏 (𝑈) is compact. Since the latter
T S
is an intersection of nested non-empty compact sets it is non-empty,
which establishes (i) and (ii).
Since 𝑈 is connected, then 𝜙𝑡 Γ+ (𝑈) is connected.
Using the precom-
pactness of Γ+ (𝑈) we derive that cl 𝜙𝑡 Γ+ (𝑈) is a nested
sequence of
compact and connected sets. Therefore, 𝑡≥0 cl 𝜙𝑡 Γ (𝑈) is connected,
+
T
which proves (iii).
Suppose6 𝑑 𝜙𝑡 (x), 𝜔(𝑈) 6→ 0 as 𝑡 → ∞, then 𝑑 𝜙𝑡𝑛 (x), 𝜔(𝑈) ≥ 𝛿 > 0,
6: Recall that the distance between a point
for some sequence 𝑡 𝑛 → ∞. Since Γ+𝜏 (𝑈) is pre-compact, the sequence x ∈ ℝ 𝑛 and a set 𝑈 ⊂ ℝ 𝑛 is defined as:
{𝜙𝑡𝑛 (x)} has a limit point y, with 𝑑(y , 𝜔(𝑈)) > 0, which is a contradiction 𝑑(x , 𝑈) := inf k x − x0 k.
x0 ∈𝑈
and therefore proves Property (iv).
The next proposition list a number of properties of omega limit sets that
do not require any compactness conditions.
and since the sets are nested the property follows. As for the intersection
we argue as follows. Note that 𝑈 ∩ 𝑉 ⊂ 𝑈 and 𝑈 ∩ 𝑉 ⊂ 𝑉 , and by
(i) 𝜔(𝑈 ∩ 𝑉) ⊂ 𝜔(𝑈) and 𝜔(𝑈 ∩ 𝑉) ⊂ 𝜔(𝑉). Combining this gives
𝜔(𝑈 ∩ 𝑉) ⊂ 𝜔(𝑈) ∩ 𝜔(𝑉).
By forward invariance of 𝜔(𝑈), 𝜙(𝑡, 𝑉) ⊂ 𝜙(𝑡, 𝜔(𝑈)) ⊂ 𝜔(𝑈). Since the
latter is closed it follows that 𝜔(𝑉) ⊂ 𝜔(𝑈), proving (iv)
Since 𝑈 ⊂ cl 𝑈 it follows that 𝜔(𝑈) ⊂ 𝜔(cl 𝑈). As for the reversed
inclusion we argue as follows. Since,
[ [ [
𝜙(𝑠, cl 𝑈) ⊂ cl 𝜙(𝑠, 𝑈) ⊂ cl 𝜙(𝑠, 𝑈)
𝑠≥𝑡 𝑠≥𝑡 𝑠≥𝑡
For the omega limit sets this implies that 𝜔(cl 𝑈) ⊂ 𝜔(𝑈), which proves
(v).
For the idempotency property we argue as follows. Since 𝜔(𝑈) is closed
and invariant we have
Remark 5.2.1 The infinite additive property is not true in general, i.e.
in general x∈𝑈 𝜔(x) ⊂ 𝜔(𝑈). In case the additive property is true for
S
arbitrary unions we say that the operator is completely additive. As a
matter of fact complete super-additivity is true for 𝜔 . With respect to
intersection we have complete sub-multiplicativity.
5 Limit sets
32
Lecture 7 and 8
5.3 Attractors
which show that every invariant sets is contained in 𝜔(𝑈) and thus
also Inv 𝑈 ⊂ 𝜔(𝑈). This implies Inv 𝑈 = 𝜔(𝑈) = 𝐴.
Attractors reveal information about the global structure of flows. In
Chapter 7 we will use trapping regions and attractors to study global
properties of flows in ℝ2 .
Poincaré sections
Lecture 8 6
In this chapter we introduce the notion of transverse section, or Poincaré 6.1 Poincaré maps . . . . . . . . . 33
section. Poincaré sections are a useful tool for analyzing the linear 6.2 The flow box theorem . . . . 35
behavior of periodic orbits. They also play a crucial role in the proof of
the Poincaré-Bendixson theorem.
Proof. The flow 𝜙𝑡 (x) is a 𝐶 1 -function in both 𝑡 and x, cf. proper ref.
Define the map
Using the above lemma we can define the first return-map, or Poincaré
map at Σ:3 3: The idea of a return map can also be de-
voloped in the absence of a periodic orbit.
𝑃 : Σ ∩ 𝐵 𝛿 (x0 ) → Σ, x ↦→ 𝑃(x) := 𝜙𝜏(x) (x). In the case a point x0 is mapped to Σ under
the flow allows one to go through the steps
of Lemma 6.1.1 in order to construct a first
For Poincaré map for periodic orbits we have the following proper- return-time and thus a Poincaré map.
ties:
The following is a variation on the first return map describes a flow near
a Poincaré section.
and therefore
ℎ −1 𝜙𝑡 (ℎ(𝑠, x)) = (𝑡 + 𝑠, x),
which is the linear flow generated by the vector field g(x) = 𝑒1 . This
shows that near non-equilibrium points 𝜙𝑡 is locally conjugate to a linear
flow.
0 < f(x) · n , ∀x ∈ 𝐿.
Proof. By the definition of omega limit set there exists a sequence of time
𝑠 𝑛 → ∞ such that z𝑛 := 𝜙 𝑠 𝑛 (x) → y ∈ 𝐿. By Lemma 6.1.3 there exists
a continuous function 𝜏 defined in a neighborhood of y with 𝜏(y) = 0,
such that y𝑛 := 𝜙𝜏(z𝑛 ) (z𝑛 ) ∈ 𝐿 for 𝑛 sufficiently large. Since z𝑛 → y we
have that 𝜏(z𝑛 ) → 0 and thus
Proof. Suppose there are two distinct intersection points y and y0. By
Lemma 7.2.3 there exist sequences of times {𝑡 𝑛 } and {𝑡 𝑛0 }, with 𝑡 𝑛 , 𝑡 𝑛0 →
∞, such that
Lemma 7.2.5 Suppose 𝜔(x) ≠ ∅ and bounded. Moreover, assume that 𝜔(x)
does not contain equilibrium points of 𝜙𝑡 . Then, there exists a periodic orbit
𝛾 ⊂ 𝜔(x) for the flow 𝜙𝑡 .
compact set 𝜔(x), and therefore 𝜔(y) ≠ ∅ and compact.6 By Proposition 6: cf. Proposition 5.2.3(i)-(ii).
5.2.4(i) and (vi) we have that 𝜔(y) ⊂ 𝜔 𝜔(x) = 𝜔(x).
Because 𝜔(y) is non-empty we can choose point z ∈ 𝜔(y). By assumption
𝜔(x) does not contain equilibrium points and therefore z is not an
equilibrium point and f(z) ≠ 0. Choose a line-segment 𝐿 at z which is
a Poincaré section for 𝜙𝑡 , i.e. z ∈ 𝜔(y) ∩ 𝐿. By Lemma 7.2.3 there exist
times 𝑡 𝑛 → ∞ such that
𝐿 3 z𝑛 = 𝜙𝑡𝑛 (y) → z.
Since 𝛾y ⊂ 𝜔(x) we have that z𝑛 ∈ 𝜔(x) ∩ 𝐿 for all 𝑛 . Lemma 7.2.4 then
implies that z𝑛 = z𝑛0 for all 𝑛 ≠ 𝑛 0. From the times 𝑡 𝑛 → ∞ we can
choose times 𝑡 𝑛 ≠ 𝑡 𝑛0 . This implies 𝜙𝑡𝑛 (y) = 𝜙𝑡𝑛0 (y) which proves that
𝛾y ⊂ 𝜔(x) is a periodic orbit.
Lemma 7.2.6 Assume 𝜔(x) is connected and 𝜔(x) does not contain equilib-
rium points of 𝜙𝑡 . Suppose 𝛾 ⊂ 𝜔(x) is a periodic orbit, then 𝜔(x) = 𝛾 .
which is a smooth vector field on ℝ2 \ O.2 The vector field 𝑋 is irrotational: 2: Note that 𝑋 does not extend to a smooth
vector field on ℝ2 .
curl 𝑋 = 0.3
3: Let 𝑋(x) = 𝑎(𝑥, 𝑦), 𝑏(𝑥, 𝑦) . Then,
curl 𝑋 := 𝜕𝑥 𝑏 − 𝜕𝑦 𝑎 .
Definition 8.1.1 Let x : [0 , 2𝜋] → ℝ2 \ O be a smooth parametrized
closed curve avoiding the origin O. Then, the line integral
∮
1
𝑊(𝛾, O) := 𝑋 · 𝑑x, (8.1)
2𝜋 𝛾
The winding number has special properties. The first step is to write The winding number can be defined with
the parametrization x(𝑡) in appropriate polar coordinates. For the point respect to any point x0 ∉ 𝛾 .
x(0) choose 𝜃0 ∈ ℝ such that x(0) = (𝑟(0) cos 𝜃0 , 𝑟(0) sin 𝜃0 ).4 Define the 4: The angle 𝜃0 is unique up to a multiple
of 2𝜋.
angular function
𝑡 𝑡
−𝑦(𝑠) 𝑥(𝑠)
¤ + 𝑥(𝑠) 𝑦(𝑠)
¤
∫ ∫
𝜃(𝑡) := 𝜃0 + 𝑋(x(𝑠)) · x¤ (𝑠)𝑑𝑠 = 𝜃0 + 𝑑𝑠.
0 0 𝑥 (𝑠) + 𝑦 (𝑠)
2 2
(8.2)
This yields the following representation for x(𝑡).
Lemma 8.1.1 The function 𝜃(𝑡) is the unique angular function such that
Proof. By definition the functions 𝑟(𝑡) and 𝜃(𝑡) are smooth functions of 𝑡.
The vectors u(𝑡) = cos 𝜃(𝑡), sin 𝜃(𝑡) and u⊥ (𝑡) = − sin 𝜃(𝑡), cos 𝜃(𝑡)
make an orthonormal frame for all 𝑡 ∈ ℝ and every vector function x(𝑡)
is uniquely represented by x(𝑡) = (x(𝑡) · u(𝑡))u(𝑡) + (x(𝑡) · u⊥ (𝑡))u⊥ (𝑡).
Note that (x(0) · 𝑢(0)) = 𝑟(0) > 0 and (x(0) · 𝑣(0)) = 0. We show now that
8 Index theory
41
Lecture 11
(x(𝑡) · u(𝑡)) = 𝑟(𝑡) and (x(𝑡) · u⊥ (𝑡)) = 0 for all 𝑡 ∈ ℝ by computing the
derivatives of the expressions (x(𝑡) · u(𝑡))/𝑟(𝑡) and (x(𝑡) · u⊥ (𝑡)). We have
Lemma 8.1.2
𝜃(2𝜋) − 𝜃(0)
𝑊(𝛾, O) = ∈ ℤ.
2𝜋
∫ 2𝜋
Proof. By definition 𝜃(0) = 𝜃0 and 𝜃(2𝜋) = 𝜃0 + 0 𝑋(x(𝑠)) · x¤ (𝑠)𝑑𝑠 =
𝜃0 + 2𝜋𝑊(𝛾, O), which proves the above expression. Since x(0) = x(2𝜋)
there exists a 𝑘 ∈ ℤ such that 𝜃(2𝜋) = 𝜃(0) + 2𝜋𝑘 = 𝜃0 + 2𝜋𝑘 , and
establishes the property that 𝑊(𝛾, O) ∈ ℤ.
𝜃0 (𝜆). Consider the path x𝜆 (0) := h(𝜆, 0) and choose 𝜃0 (0) ∈ ℝ such that 7: It is crucial for the argument in homo-
topy principle for 𝛾𝜆 to be a subset of
x0 (0) = (𝑟(0 , 0) cos 𝜃0 (0), 𝑟(0 , 0) sin 𝜃0 (0)). By Lemma 8.1 there exists a
ℝ2 \ O for all 𝜆 ∈ [0, 1].
unique function 𝜃0 (𝜆) such that
Now define
∫ 𝑡
𝜃(𝜆, 𝑡) := 𝜃0 (𝜆) + 𝑋(x𝜆 (𝑠)) · x¤ 𝜆 (𝑠)𝑑𝑠,
0
Proof. As the proof of Lemma 8.1 using the representation in (8.4).8 8: As a matter of fact 𝑊(𝛾𝜆 , O) is constant
in 𝜆 ∈ [0 , 1].
In the application of Green’s Theorem 𝛾 is positively oriented.9 In the 9: The region 𝑅 + (𝛾) is positively oriented.
case that O ∈ 𝑅 + (𝛾) we argue as follows. Define 𝑅
e+ (𝛾) = 𝑅+ (𝛾) \ 𝐵 𝜖 (O)10 The induced orientation for 𝛾 implies that
and 𝑋 is smooth vector field on 𝑅 (𝛾). Again apply Green’s Theorem:
e + the curve is traversed counter-clockwise.
10: 𝐵 𝜖 (O) is a small open ball centered at
O.
∮ ∬
1 1
𝑋 · 𝑑x = curl 𝑋 𝑑𝐴 = 0.
2𝜋 𝜕𝑅
e+ (𝛾) 2𝜋 𝑅
e+ (𝛾)
= 𝑊(𝛾, O) − 1,
x 𝑓
[0 , 2𝜋] → − ℝ2 ,
− ℝ2 →
8 Index theory
43
Lecture 11
𝜄 𝑓 (𝛾) := 𝑊( 𝑓 ◦ 𝛾, O) ∈ ℤ , (8.5)
If we express the vector field 𝑓 by 𝑓 (x) = 𝑝(x), 𝑞(x) , then the index is
given by the integral
−𝑞𝑑𝑝 + 𝑝𝑑𝑞
∮
1
𝜄 𝑓 (𝛾) = . (8.6)
2𝜋 𝛾 𝑝2 + 𝑞2
Remark 8.2.1 Note that the expression for the index in (8.1) makes
sense for any vector field defined on 𝛾 or on any neighborhood of 𝛾 .
Therefore, vector field do not have to be related to global flows and
therefore be normalized.
−𝑞𝑑𝑝 + 𝑝𝑑𝑞
∮
1
𝜄 𝑓 (𝛾) =
2𝜋 𝛾 𝑝2 + 𝑞2
−𝑞𝜕𝑥 𝑝 + 𝑝𝜕𝑥 𝑞 𝑑𝑥 + −𝑞𝜕𝑦 𝑝 + 𝑝𝜕𝑦 𝑞 𝑑𝑦
∮
1
=
2𝜋 𝛾 𝑝2 + 𝑞2
!
1
∬ h −𝑞𝜕𝑦 𝑝 + 𝑝𝜕𝑦 𝑞 i h −𝑞𝜕 𝑝 + 𝑝𝜕 𝑞 i
𝑥 𝑥
= 𝜕𝑥 − 𝜕𝑦 𝑑𝐴
2𝜋 𝑅+ (𝛾) 𝑝2 + 𝑞2 𝑝2 + 𝑞2
= 0,
which follows by carrying out the partial derivatives under the integral.
Proof. The expression in (8.6) for the index is invariant under the trans-
formation 𝑓 ↦→ − 𝑓 , which completes the proof.
Lemma 8.2.3 Let 𝛾 and 𝛾0 be two positively oriented regular Jordan curves
such that 𝛾0 ⊂ int 𝑅 + (𝛾) and 𝑆 = 𝑅 + (𝛾) \ int 𝑅 + (𝛾0) contains no equilib-
rium points. Then,
𝜄 𝑓 (𝛾) = 𝜄 𝑓 (𝛾0).
The last property of the index we mention is the case where 𝛾 is a periodic
orbit of the flow 𝜙𝑡 .14 A sktech of a proof can be found in Sect. 8.4. 14: See for instance [1] for a proof.
The index of a regular Jordan curve with respect to a vector field can be
used to assign an index the isolated fixed points of a planar flow 𝜙𝑡 .15 15: The index for a fixed point can also
be defined for flows on ℝ 𝑛 by using then
A fixed point x0 of 𝜙𝑡 is isolated if there exists an 𝜖 > 0 such that x0 is the Brouwer mapping degree as a generaliza-
only fixed point in 𝐵 𝜖 (x0 ). Let 𝛾 be a positively oriented regular Jordan tion of the winding number.
curve such that x0 is the only fixed point in 𝑅 + (𝛾). Such regular Jordan
8 Index theory
45
Lecture 11
The fixed point index can be linked to the index index of a regular Jordan
curve via the index sum formula.
Proof. Choose 𝜖 > 0 sufficiently small such that cl 𝐵 𝜖 (x𝑖 ) ⊂ int 𝑅 + (𝛾) for
all 𝑖 . Consider the set
[
𝑆 = 𝑅 + (𝛾) \ 𝐵 𝜖 (x𝑖 ).
𝑖
For some applications of the index homotopy principle for the Jordan
curves and the vector fields are useful tools.
𝜄 𝑓 (𝛾0 ) = 𝜄 𝑓 (𝛾1 ).
𝜄 𝑓 (𝛾0 ) = 𝑊( 𝑓 ◦ 𝛾0 , O) = 𝑊( 𝑓 ◦ 𝛾1 , O) = 𝜄 𝑓 (𝛾1 ),
𝜄 𝑓0 (𝛾) = 𝜄 𝑓1 (𝛾).
𝜄 𝑓0 (𝛾) = 𝑊( 𝑓0 ◦ 𝛾, O) = 𝑊( 𝑓1 ◦ 𝛾, O) = 𝜄 𝑓1 (𝛾),
We end this section with a sketch of a proof of Lemma 8.2.4. The index
is the same for positively and negatively oriented periodic orbits by
Lemma 8.2.2. We assume 𝛾 is positively oriented. Choose a 𝜖 -tubular
neighborhood of 𝑁𝜖 (𝛾) of the periodic orbit 𝛾 . Using Lemma 8.1.1 we
obtain a diffeomorphism between 𝑁𝜖 (𝛾) and the annulus 𝐴 𝜖 := {x | 1−𝜖 ≤
| x | ≤ 1 + 𝜖} such that 𝛾 maps to the unit circle 𝛾 in 𝐴 − 𝜖. The transformed
vector field on 𝐴 𝜖 is denoted by e 𝑓 . Use the homotopy principle in Lemma
8.4.2 to normalize the vector field such that e 𝑓 has unit length at the unit
circle e𝛾. Then e 𝑓 |e
𝛾 = (−𝑦, 𝑥). Summarizing we have
∮ ∫ 2𝜋
1 1
𝜄 𝑓 (𝛾) = 𝜄 e𝑓 (e
𝛾) = −𝑞𝑑𝑝 + 𝑝𝑑𝑞 = (𝑥 𝑦 0 − 𝑦𝑥 0)𝑑𝑡 = 1,
2𝜋 𝛾
e 2𝜋 0
by using the parametrization cos(𝑡), sin(𝑡) .
Compactification
Lecture 12 and 13 9
We describe a method to extend smooth vector fields on 𝑅 𝑛 to smooth 9.1 Stereographic and central projec-
vector fields on 𝑆 𝑛 ⊂ ℝ 𝑛+1 . This allows us to understand the asymptotic tions . . . . . . . . . . . . . . . . . 47
behavior of vector fields and also provides an interplay between the 9.2 Extending the vector field . . 47
9.3 Examples . . . . . . . . . . . . 49
dynamics of a vector field and its asymptotics.
A 1-dimensional system . . . 49
A planar system . . . . . . . . 49
A 3-dimensional vector field 50
9.1 Stereographic and central projections Linearization . . . . . . . . . . . 51
9.4 Blowing up fixed points . . . 52
The stereographic projection is a standard method to give a diffeomorphism
between the 𝑛 -sphere 𝑆 𝑛 \ {∞} ⊂ ℝ 𝑛+1 and ℝ 𝑛 where ∞ denotes the
north pole. The stereogarphic projection is well-known and convenient
to use in many situations. The draw-back is that all asynmptotic behavior
is mapped to one point that models infinity, the north pole.1 A different, 1: Compactification of ℝ 𝑛 using the stere-
but similar technique, makes the equator of 𝑆 𝑛 play the role of infinity. ographic projection and the 𝑛 -sphere is
also referred to as the Bendixson sphere.
This is called the central projection.2 Let us describe the central projection.
2: Compactification of ℝ 𝑛 using the cen-
Consider coordinates (𝝃, 𝜁) ⊂ ℝ 𝑛+1 , where 𝝃 = (𝜉1 , · · · , 𝜉𝑛 } and let
tral projection and the upper 𝑛 -sphere is
also referred to as the Poincaré sphere.
𝑆 𝑛 := {(𝝃, 𝜁) ∈ ℝ 𝑛+1 | |𝝃| 2 + 𝜁 2 = 1}.
Denote the upper hemisphere by 𝑆+𝑛 , i.e. all points on 𝑆 𝑛 with 𝜁 > 0.
Consider ℝ 𝑛 with coordinates x = (𝑥 1 , · · · , 𝑥 𝑛 ).
For points (𝝃, 𝜁) ∈ 𝑆+𝑛 define the points
𝝃
𝑆+𝑛 → ℝ 𝑛 , (𝝃, 𝜁) ↦→ x = ∈ ℝ𝑛 , (9.1)
𝜁
which is well-defined since 𝜁 > 0.3 Conversely, write 𝝃 = 𝜁 x. Then, 3: We can think of (𝝃, 𝜁) as homogeneous
p coordinates.
|𝝃| 2 + 𝜁 2 = 1 implies (1 + | x | 2 )𝜁 2 = 1 and thus 𝜁 = 1/ 1 + | x | 2 . This yields
the map
!
𝑛 x 1
ℝ → 𝑆+𝑛 , x ↦→ (𝝃, 𝜁) = p ,p (9.2)
1 + |x|2 1 + |x|2
This establishes the diffeomorphisms between 𝑆+𝑛 and ℝ 𝑛 . The same can
be carried out for 𝑆−𝑛 . The equator 𝑆eq
𝑛−1 = {(𝝃, 0) | |𝝃| = 1 } plays the role
This vector field is not tangent to 𝑆 𝑛 and thus we need to correct it with
the normal direction: 𝜁 𝑓 (𝝃/𝜁), 0 − 𝑐(𝝃, 𝜁). To be tangent to 𝑆 𝑛 we take
inner product with the normal, i.e.
h i
𝜁 𝑓 (𝝃/𝜁), 0 − 𝑐(𝝃, 𝜁) · (𝝃, 𝜁) = 0 ,
which yields 𝑐 = 𝜁𝝃 · 𝑓 (𝝃/𝜁). The vector field in ℝ 𝑛+1 tangent to 𝑆 𝑛 is4 4: The same expression can be obtained
by differentiating x = 𝜁𝝃 and 𝜁 .
𝜁 𝑓 (𝝃/𝜁) − 𝜁𝝃 𝝃 · 𝑓 (𝝃/𝜁) , −𝜁 2 𝝃 · 𝑓 (𝝃/𝜁) .
𝝃0 = 𝑓 ∗ (𝝃, 𝜁) − 𝝃 · 𝑓 ∗ (𝝃, 𝜁) 𝝃
𝜁0 = − 𝝃 · 𝑓 ∗ (𝝃, 𝜁) 𝜁.
For the latter system the time variable is 𝑠 . The relation to the time
variable 𝑡 of the system x¤ = 𝑓 (x) is
𝜁 𝑚−1 𝑑𝑠 = 𝑑𝑡.
𝑓 (x) =
Remark 9.2.1 If we consider a renormalized vector field x¤ = e
𝑓 (x)/(1 + | 𝑓 (x)|), then
𝜁 𝑚 𝑓 (𝝃/𝜁) 𝑓 ∗ (𝝃, 𝜁)
𝑓 ∗ (𝝃, 𝜁) =
e = ,
𝜁 𝑚 + |𝜁 𝑚 𝑓 (𝝃/𝜁)| 𝜁 𝑚 + | 𝑓 ∗ (𝝃, 𝜁)|
9.3 Examples
A 1-dimensional system
3𝜉 2 − 5𝜉 4
0 −2 0
, ,
−4 𝜁𝜉3 −𝜉4 0 −1
1 1
𝜁 2 𝑑𝑠 = 𝑑𝑡 =⇒ 𝑐 2 𝑒 −2𝑠 𝑑𝑠 ∼ 𝑑𝑡 =⇒ − 𝑐 2 𝑒 −2𝑠 ∼ 𝑡 − 02
2 2𝑐
1 𝑐0
𝑥(𝑡) ∼ ∼√
𝜁(𝑠) 1 − 2𝑡𝑐 02
A planar system
The third example in this section is a linear system in ℝ3 .8 Consider the 8: We choose a linear system to keep the
system: technicalities down to a minimum.
𝑥¤ = 𝑥 − 𝑦,
𝑦¤ = 𝑥 + 𝑦,
𝑧¤ = −𝑧
and the vector field is given by 𝑓 (𝑥, 𝑦, 𝑧) = (𝑥 − 𝑦, 𝑥 + 𝑦, −𝑧). In order to
transform to the compactified system we take, as before, 𝑚 = 1 which
gives the following system
where we used the relation 𝜉12 + 𝜉22 + 𝜉32 = 1, which is the ‘equator’ on
𝑆3 and we call it the 2-sphere at infinity. We use spherical-cylindrical
coordinates in ℝ4 : 𝜉1 = cos 𝜃 sin 𝜙 , 𝜉2 = sin 𝜃 sin 𝜙 , 𝜉3 = cos 𝜙 and
𝜁 ∈ ℝ+ , 𝜃 ∈ [0, 2𝜋] and 𝜙 ∈ [0 , 𝜋]. In terms of derivatives we have:
−𝜉2 𝜉10 + 𝜉1 𝜉20 = (sin2 𝜙)𝜃0 and 𝜉30 = −(sin 𝜙)𝜙0. In polar coordinates the
𝝃 -component of the vector field is given by:
« −2 cos 𝜙 sin2 𝜙 ¬
9 Compactification
51
Lecture 12 and 13
For 𝜃0 we obtain
(sin2 𝜙)𝜃0 = −(2 cos 𝜃 − sin 𝜃) sin 𝜃 sin2 𝜙 + (cos 𝜃 + 2 sin 𝜃) cos 𝜃 sin2 𝜙,
𝜁0 = (cos 2𝜙)𝜁.
For 𝜁 ≈ 0 we conclude that north and south pole are unstable for the
whole system and the periodic orbit is stable for the 3-dimensional
flow.
Linearization
𝜁0 = − 𝝃 · f(𝝃, 𝜁) 𝜁.
𝜉2 𝜁 𝑦 𝜉2
𝑦= , 𝑧= , and = .
𝜉1 𝜉1 𝑧 𝜁
9 Compactification
52
Lecture 12 and 13
𝜉1 𝜉2 𝑦
If we write 𝜁 = 𝑧𝜉1 , 𝜁 = 1
𝑧 and 𝜁 = 𝑧 we obtain
h 1 𝑦 1 𝑦 i 1 𝑦
𝑦 0 = 𝜉1𝑚−1 𝑧 𝑚 𝑔 , −𝑧 𝑚 𝑓 , 𝑦 and 𝑧 0 = −𝜉1𝑚−1 𝑧 𝑚+1 𝑓 , .
𝑧 𝑧 𝑧 𝑧 𝑧 𝑧
𝑦¤ = 2 𝑦,
𝑧¤ = 𝑧.
We alter the compactification via the central projection in two ways: (i)
we map to a cylinder as generalized polar coordinates and (ii) we allow a
generalization of homogeneous coordinates. In this construction we have
two parameters. The first one is a smooth convex function ℎ : ℝ 𝑛 → ℝ+
such that
(i) ℎ −1 (1) diffeomorphic to 𝑆 𝑛−1 ;
(ii) ∇ℎ(𝝃) · Ω𝝃 = 𝜆ℎ(𝝃), for some 𝜆 > 0,9 9: The function ℎ is quasi-homogeneous
of degree 𝜆.
9 Compactification
53
Lecture 12 and 13
1
𝑒 (𝑚−1)𝜏 𝝃¤ = 𝑓 † (𝝃, 𝜏) − ∇ℎ(𝝃) · 𝑓 † (𝝃, 𝜏) Ω𝝃
𝜆
1
𝝃0 = 𝑓 † (𝝃, 𝜏) − ∇ℎ(𝝃) · 𝑓 † (𝝃, 𝜏) Ω𝝃
𝜆
1
𝜏0 = − ∇ℎ(𝝃) · 𝑓 † (𝝃, 𝜏) .
𝜆
For the latter system the time variable is 𝑠 . The relation to the time
variable 𝑡 of the system x¤ = 𝑓 (x) is
𝑒 (𝑚−1)𝜏 𝑑𝑠 = 𝑑𝑡.
is smooth on S𝑛−1 × ℝ+ .11 If we consider the ‘end’ 𝜏 = +∞ we set 11: We use the convention: ℝ+ = [0 , ∞).
𝜏 = − ln 𝜁 and 𝜁 = 0 corresponds to 𝜏 = +∞, which yields the condition:
𝑥¤ = 𝑥 3 − 3𝑥 𝑦 2 + 𝑥 4 𝑦,
𝑦¤ = 3 𝑥 2 𝑦 − 𝑦 3 − 𝑥 4 + 𝑦 5 .
(9.10)
In order to analyze the origin we consider the ‘end’ 𝜏 = +∞ and 𝜏 = − ln 𝜁 .
This gives:14 14: Smooth vector field on all of S𝑛−1 ×
ℝ+ .
9 Compactification
55
Lecture 12 and 13
The correction factor on the limiting circle S1 at 𝜁 = 0 for the vector field
in the quasi-homogeneous coordinates is:
since 𝜉12 + 𝜉22 = 1. The flow on the circle is described by the equations:
From the 𝜃 -equation we have four fixed points (±1 , 0) and (0 , ±1), where
first two unstable and the latter two stable. From the 𝜁 -equation we derive
that the first fixed points are unstable for the whole system and the latter
two fixed points are stable for the whole system.16 The above equations 16: For an illustration of the blown-up
for 𝜃 and 𝜁 describe the local behavior near the degenerate fixed point fixed point see [1, Sect. 3.10].
Observe that the extension to 𝜁 = 0 does not exist with the present
choices. In order to describe ∞ we choose 𝑚 = 5. This gives:
𝜉14 𝜉2
∗
𝑓 (𝝃, 0) = . (9.15)
𝜉25
The correction factor on the limiting circle S1 at 𝜁 = 0 for the vector field
in the quasi-homogeneous coordinates is:
The right hand side is 𝜋-periodic and we find equilibria for 𝜃 equal to
0, 𝜋/4, 𝜋/2, 𝜋, 5𝜋/4, and 3𝜋/2. Without carrying out a detailed analysis
the right hand side in the 𝜃 equation reveals that the angles 𝜋/2 and
3𝜋/2 are stable, the angles 𝜋/4 and 5𝜋/4 are unstable and the angles 0
and 𝜋 are mixed stable/unstable for the flow on the circle. In order to
examine the stability in the overall system we invoke the 𝜁 equation:
[1] Lawrence Perko. ‘Differential Equations and Dynamical Systems, Third Edition’. In: (Feb. 2010), pp. 1–571
(cited on pages i, 3, 44, 50, 54, 55).
[2] M Hirsch and S Smale. ‘Differential Equations, Dynamical Systems, and Linear Algebra (Pure and
Applied Mathematics, Vol. 60)’. In: (1974) (cited on page 8).
[3] James Munkres. ‘Topology ’. In: (Jan. 2015), pp. 1–507 (cited on page 15).
Notation
The next list describes several symbols that will be later used within the body of the document.
𝑐 Speed of light in a vacuum inertial frame
ℎ Planck constant
Capitals shown are the ones that differ from Roman capitals.
Alphabetical Index