Theory of ODE
Theory of ODE
theoryofodes July 4, 2007 13:20 Page ii
theoryofodes July 4, 2007 13:20 Page i
theoryofodes July 4, 2007 13:20 Page ii
theoryofodes July 4, 2007 13:20 Page i
Contents
Contents i
1 Fundamental Theory 1
1.1 ODEs and Dynamical Systems . . . . . . . . . . . . . . . . . 1
1.2 Existence of Solutions . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Uniqueness of Solutions . . . . . . . . . . . . . . . . . . . . . 10
1.4 Picard-Lindelöf Theorem . . . . . . . . . . . . . . . . . . . . . 14
2 Linear Systems 27
2.1 Constant Coefficient Linear Equations . . . . . . . . . . . . 27
2.2 Understanding the Matrix Exponential . . . . . . . . . . . . 30
2.3 Generalized Eigenspace Decomposition . . . . . . . . . . . 33
2.4 Operators on Generalized Eigenspaces . . . . . . . . . . . . 37
2.5 Real Canonical Form . . . . . . . . . . . . . . . . . . . . . . . 41
2.6 Solving Linear Systems . . . . . . . . . . . . . . . . . . . . . . 43
2.7 Qualitative Behavior of Linear Systems . . . . . . . . . . . . 50
2.8 Exponential Decay . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.9 Nonautonomous Linear Systems . . . . . . . . . . . . . . . . 56
2.10 Nearly Autonomous Linear Systems . . . . . . . . . . . . . . 61
2.11 Periodic Linear Systems . . . . . . . . . . . . . . . . . . . . . 65
3 Topological Dynamics 71
3.1 Invariant Sets and Limit Sets . . . . . . . . . . . . . . . . . . 71
3.2 Regular and Singular Points . . . . . . . . . . . . . . . . . . . 75
i
theoryofodes July 4, 2007 13:20 Page ii
Contents
4 Conjugacies 101
4.1 Hartman-Grobman Theorem: Part 1 . . . . . . . . . . . . . . 101
4.2 Hartman-Grobman Theorem: Part 2 . . . . . . . . . . . . . . 103
4.3 Hartman-Grobman Theorem: Part 3 . . . . . . . . . . . . . . 105
4.4 Hartman-Grobman Theorem: Part 4 . . . . . . . . . . . . . . 109
4.5 Hartman-Grobman Theorem: Part 5 . . . . . . . . . . . . . . 111
4.6 Constructing Conjugacies . . . . . . . . . . . . . . . . . . . . 115
4.7 Smooth Conjugacies . . . . . . . . . . . . . . . . . . . . . . . 119
ii
theoryofodes July 4, 2007 13:20 Page 1
1
Fundamental Theory
n + 1 copies
z }| {
F : dom(F ) ⊆ R × E × · · · × E → Rj . (1.1)
theoryofodes July 4, 2007 13:20 Page 2
1. Fundamental Theory
First-order Equations
y1 := x
y2 := ẋ
y3 := ẍ
..
.
yn := x (n−1) ,
G1 (t, u, p) := p1 − u2
G2 (t, u, p) := p2 − u3
G3 (t, u, p) := p3 − u4
..
.
Gn−1 (t, u, p) := pn−1 − un ,
Gn (t, u, p) := F (t, u1 , . . . , un , pn ),
where
dom(Gn ) = (t, u, p) ∈ R × E n × E n (t, u1 , . . . , un , pn ) ∈ dom(F ) .
theoryofodes July 4, 2007 13:20 Page 3
we see that x satisfies (1.2) if and only if y satisfies G(t, y(t), ẏ(t)) =
0.
x(t ) = x .
0 0
Autonomous Equations
ẋ = f (t, x) (1.4)
theoryofodes July 4, 2007 13:20 Page 4
1. Fundamental Theory
Dynamical Systems
ẋ = f (x)
(1.5)
x(0) = x 0
3. ϕ is continuous.
theoryofodes July 4, 2007 13:20 Page 5
• Ω ⊆ Rn ;
ẏ = f (y)
y(0) = x . 0
3. ϕ is continuous.
theoryofodes July 4, 2007 13:20 Page 6
1. Fundamental Theory
by
n copies
z }| {
ϕ(n, x) = F (F (· · · (F (x)) · · · )),
1.2 Existence of Solutions
Approximate Solutions
How does one go about proving that (1.7) has a solution if, unlike
the case with so many IVPs studied in introductory courses, a formula
for a solution cannot be found? One idea is to construct a sequence
of “approximate” solutions, with the approximations becoming better
and better, in some sense, as we move along the sequence. If we can
6
theoryofodes July 4, 2007 13:20 Page 7
Existence of Solutions
Note that if, for some k, xk = xk+1 then we have found a solution.
Another approach is to construct a Tonelli sequence. For each k ∈ N,
let xk (t) be defined by
a, if t0 ≤ t ≤ t0 + 1/k
xk (t) = Z t−1/k (1.8)
a + f (s, xk (s)) dx, if t ≥ t0 + 1/k
t0
(1.6)) has a solution, and will use Picard iterates to show that, under an
additional hypothesis on f , the solution of (1.7) is unique.
Existence
For the first result, we will need the following definitions and theorems.
theoryofodes July 4, 2007 13:20 Page 8
1. Fundamental Theory
Proof. For simplicity, we will only consider t ∈ [t0 , t0 + b]. For each
k ∈ N, let xk : [t0 , t0 + b] → Rn be defined by (1.8). We will show that
(xk ) converges to a solution of (1.6).
Step 1: Each xk is well-defined.
Fix k ∈ N. Note that the point (t0 , a) is in the interior of a set on which
f is well-defined. Because of the formula for xk (t) and the fact that
it is, in essence, recursively defined on intervals of width 1/k moving
steadily to the right, if xk failed to be defined on [t0 , t0 + b] then there
would be t1 ∈ [t0 + 1/k, t0 + b) for which |xk (t1 ) − a| = β. Pick the first
such t1 . Using (1.8) and the bound on f , we see that
Z Z
t1 −1/k t1 −1/k
|xk (t1 ) − a| = f (s, xk (s)) ds ≤ |f (s, xk (s))| ds
t0 t0
Z t1 −1/k
≤ M ds = M(t1 − t0 − 1/k) < M(b − 1/k)
t0
theoryofodes July 4, 2007 13:20 Page 9
Existence of Solutions
≤ M|t2 − t1 |.
Step 5: The function f (·, x(·)) is the uniform limit of (f (·, xkℓ (·)))
on [t0 , t0 + b].
Let ε > 0 be given. Since f is continuous on a compact set, it is
uniformly continuous. Thus, we can pick δ > 0 such that |f (s, p) −
f (s, q)| < ε whenever |p − q| < δ. Since (xkℓ ) converges uniformly
to x, we can pick N ∈ N such that |xkℓ (s) − x(s)| < δ whenever
s ∈ [t0 , t0 + b] and ℓ ≥ N. If ℓ ≥ N, then |f (s, xkℓ (s)) − f (s, x(s))| < ε.
Step 6: The function x is a solution of (1.6).
Fix t ∈ [t0 , t0 + b]. If t = t0 , then clearly (1.7) holds. If t > t0 , then for
ℓ sufficiently large
Zt Zt
xkℓ (t) = a + f (s, xkℓ (s)) ds − f (s, xkℓ (s)) ds. (1.9)
t0 t−1/kℓ
theoryofodes July 4, 2007 13:20 Page 10
1. Fundamental Theory
on the interval (−∞, ∞) are there? Give formulas for all of them.
If more than continuity of f is assumed, it may be possible to prove
that
ẋ = f (t, x)
(1.10)
x(t ) = a,
0
10
theoryofodes July 4, 2007 13:20 Page 11
Uniqueness of Solutions
Show that
V ′ (t) ≤ LV (t) (1.12)
t0 to t = T , show that V (T ) ≤ ε exp[L(T − t0 )].
(d) By using (1.11) and letting ε ↓ 0, show that U(T ) = 0 for all
T ∈ [t0 , t0 + b], so x1 = x2 .
theoryofodes July 4, 2007 13:20 Page 12
1. Fundamental Theory
theoryofodes July 4, 2007 13:20 Page 13
Uniqueness of Solutions
Proof. Pick λ < 1 such that d(T (x), T (y)) ≤ λd(x, y) for every x, y ∈
X. Pick any point x0 ∈ X. Define a sequence (xk ) by the recursive
formula
xk+1 = T (xk ). (1.13)
If k ≥ ℓ ≥ N, then
theoryofodes July 4, 2007 13:20 Page 14
1. Fundamental Theory
is a Banach space.
for every x ∈ X.
Theorem A closed subspace of a complete metric space is a complete
metric space.
theoryofodes July 4, 2007 13:20 Page 15
Picard-Lindelöf Theorem
theoryofodes July 4, 2007 13:20 Page 16
1. Fundamental Theory
nite subcollection of that collection whose union also contains K. The
original collection is called a cover of K, and the finite subcollection is
called a finite subcover of the original cover.
theoryofodes July 4, 2007 13:20 Page 17
Intervals of Existence
Proof.
Step 1: If I1 and I2 are open intervals of existence with correspond-
ing solutions x1 and x2 , then x1 and x2 agree on I1 ∩ I2 .
Let I = I1 ∩ I2 , and let I ∗ be the largest interval containing t0 and con-
tained in I on which x1 and x2 agree. By the Picard-Lindelöf Theorem,
I ∗ is nonempty. If I ∗ ≠ I, then I ∗ has an endpoint t1 in I. By conti-
nuity, x1 (t1 ) = x2 (t1 ) =: a1 . The Picard-Lindelöf Theorem implies that
the new IVP
ẋ = f (t, x)
(1.16)
x(t ) = a
1 1
theoryofodes July 4, 2007 13:20 Page 18
1. Fundamental Theory
covers K. Let
m
[
K ′ := [ti − 2α(ti , ai ), ti + 2α(ti , ai )] × B(ai , 2β(ti , ai )) ,
i=1
let
m
α̃ := min α(ti , ai ) i=1 ,
and let
m
theoryofodes July 4, 2007 13:20 Page 19
Intervals of Existence
Global Existence
tion
f (x)
ẋ = q .
1 + |f (x)|2
f (x) d(x, Rn \ Ω)
ẋ = q ·p ,
1 + |f (x)|2 1 + d(x, Rn \ Ω)2
19
theoryofodes July 4, 2007 13:20 Page 20
1. Fundamental Theory
theoryofodes July 4, 2007 13:20 Page 21
Dependence on Parameters
Continuous Dependence
Zt
X(t) ≤ C + K X(s) ds
t0
X(t) ≤ CeK(t−t0 )
f : [t0 − α, t0 + α] × Ω1 × Ω2 ⊆ R × Rn × Rk → Rn
theoryofodes July 4, 2007 13:20 Page 22
1. Fundamental Theory
satisfy
ẋi = f (t, xi , µi )
x (t ) = a,
i 0
then
L2
|x1 (t) − x2 (t)| ≤ |µ1 − µ2 |(eL1 |t−t0 | − 1) (1.19)
L1
for t ∈ [t0 − α, t0 + α].
Z
t
|x1 (t) − x2 (t)| = [f (s, x1 (s), µ1 ) − f (s, x2 (s), µ2 ] ds
t0
Zt
≤ |f (s, x1 (s), µ1 ) − f (s, x2 (s), µ2 )| ds
t0
Zt
≤ |f (s, x1 (s), µ1 ) − f (s, x1 (s), µ2 )| ds
t0
Zt
+ |f (s, x1 (s), µ2 ) − f (s, x2 (s), µ2 )| ds
t0
Zt
≤ [L2 |µ1 − µ2 | + L1 |x1 (s) − x2 (s)|] ds
t0
theoryofodes July 4, 2007 13:20 Page 23
Dependence on Parameters
(a) If f (t, p) < g(t, p) for every (t, p) ∈ R × R and a < b, show
that x(t) < y(t) for every t ≥ t0 .
Differentiable Dependence
That xµ , if it exists, should satisfy the IVP (1.21) is not terribly sur-
prising; (1.21) can be derived (formally) by differentiating (1.20) with
respect to µ. The real difficulty is showing that xµ exists. The key to
23
theoryofodes July 4, 2007 13:20 Page 24
1. Fundamental Theory
the proof is to use the fact that (1.21) has a solution y and then to
use the Gronwall inequality to show that difference quotients for xµ
converge to y.
Proof. Given µ, it is not hard to see that the right-hand side of the ODE
in (1.21) is continuous in t and y and is locally Lipschitz continuous
with respect to y, so by the Picard-Lindelöf Theorem we know that
(1.21) has a unique solution y(·, µ). Let
dz dw
(t, µ, ∆µ) = (t, µ, ∆µ)−fx (t, x(t, µ), µ)y(t, µ)−fµ (t, x(t, µ), µ),
dt dt
and
Hence,
dz
dt (t, µ, ∆µ) ≤ |fµ (t, x(t, µ), µ + θ2 ∆µ) − fµ (t, x(t, µ), µ)|
theoryofodes July 4, 2007 13:20 Page 25
Dependence on Parameters
and
|p(t, µ, ∆µ)| < ε, (1.23)
then Zt Zt
dz
|z(t, µ, ∆µ)| ≤ (s, µ, ∆µ) ds ≤ X(s) ds
t0 ds t0
so Zt
X(t) ≤ ε + (K + ε) X(s) ds,
t0
lim z(t, µ, ∆µ) = 0.
∆µ→0
25
theoryofodes July 4, 2007 13:20 Page 26
theoryofodes July 4, 2007 13:20 Page 27
2
Linear Systems
Linear Equations
Definition Given
f : R × Rn → Rn ,
ẋ = f (t, x) (2.1)
ODEs of the form ẋ = A(t)x + g(t) are also often called linear, al-
though they don’t satisfy the definition given above. These are called
inhomogeneous; ODEs satisfying the previous definition are called ho-
mogeneous.
27
theoryofodes July 4, 2007 13:20 Page 28
2. Linear Systems
We will show that this series converges, but first we specify a norm on
L(Rn , Rn ).
|Bx|
kBk = sup = sup |Bx|.
x≠0 |x| |x|=1
theoryofodes July 4, 2007 13:20 Page 29
Since the regular exponential series (for real arguments) has an infinite
radius of convergence, we know that the last quantity in this estimate
goes to zero as ℓ, m ↑ ∞.
Thus, eB makes sense, and, in particular, etA makes sense for each
fixed t ∈ R and each A ∈ L(Rn , Rn ). But does x(t) := etA x0 solve
(2.2)? To check that, we’ll need the following important property.
∞ X i j (i−j) ∞ X i
! j (i−j)
X B1 B2 X i B1 B2
= =
i=0 j=0
j!(i − j)! i=0 j=0 j i!
X∞
(B1 + B2 )i
= = eB1 +B2 .
i=0
i!
theoryofodes July 4, 2007 13:20 Page 30
2. Linear Systems
Transformations
Unfortunately, this cannot always be done. Over the next few sections,
we will show that what can be done, in general, is to pick Q so that
P = S + N, where S is a semisimple matrix with a fairly simple form,
N is a nilpotent matrix with a fairly simple form, and S and N com-
mute. (Recall that a matrix is semisimple if it is diagonalizable over
the complex numbers and that a matrix is nilpotent if some power of
the matrix is 0.) The forms of S and N are simple enough that we
can calculate their exponentials fairly easily, and then we can multiply
them to get the exponential of S + N.
We will spend a significant amount of time carrying out the project
described in the previous paragraph, even though it is linear algebra
that some of you have probably seen before. Since understanding the
30
theoryofodes July 4, 2007 13:20 Page 31
Eigensystems
theoryofodes July 4, 2007 13:20 Page 32
2. Linear Systems
This formula should make clear how the projections of etA x0 grow or
decay as t → ±∞.
The same sort of analysis works when the eigenvalues are (non-
trivially) complex, but the resulting formula is not as enlightening. In
addition to the difficulty of a complex change of basis, the behavior of
etλk is less clear when λk is not real.
One way around this is the following. Sort the eigenvalues (and
eigenvectors) of A so that complex conjugate eigenvalues {λ1 , λ1 , . . . ,
λm , λm } come first and are grouped together and so that real eigen-
values {λm+1 , . . . , λr } come last. For k ≤ m, set ak = Re λk ∈ R,
bk = Im λk ∈ R, yk = Re xk ∈ Rn , and zk = Im xk ∈ Rn . Then
and
theoryofodes July 4, 2007 13:20 Page 33
In order to compute etA from this formula, we’ll need to know how
to compute etAk . This can be done using the power series formula. An
alternative approach is to realize that
" # " #
x(t) tAk c
:= e
y(t) d
e ak t
cos bk t −e ak t
sin bk t
etAk =
eak t sin bk t eak t cos bk t
Putting this all together and using the form of P , we see that etA =
QetP Q−1 , where etP is the (m + r ) × (m + r ) block diagonal matrix
whose first m diagonal blocks are the 2 × 2 matrices
" #
eak t cos bk t −eak t sin bk t
eak t sin bk t eak t cos bk t
theoryofodes July 4, 2007 13:20 Page 34
2. Linear Systems
V = V1 ⊕ · · · ⊕ Vm
theoryofodes July 4, 2007 13:20 Page 35
N(T ) = x ∈ V T k x = 0 for some k > 0 ,
and let
R(T ) = x ∈ V T k u = x has a solution u for every k > 0 .
Note that N(T ) is the union of the null spaces of the positive powers of
T and R(T ) is the intersection of the ranges of the positive powers of
T . This union and intersection are each nested, and that implies that
there is a number m ∈ N such that R(T ) is the range of T m and N(T )
is the nullspace of T m .
Proof. Pick m such that R(T ) is the range of T m and N(T ) is the
nullspace of T m . Note that T |R(T ) : R(T ) → R(T ) is invertible. Given
−m m
x ∈ V , let y = T |R(T ) T x and z = x − y. Clearly, x = y + z,
y ∈ R(T ), and T m z = T m x − T m y = 0, so z ∈ N(T ). If x = ỹ + z̃ for
some other ỹ ∈ R(T ) and z̃ ∈ N(T ) then T m ỹ = T m x − T m z̃ = T m x,
so ỹ = y and z̃ = z.
theoryofodes July 4, 2007 13:20 Page 36
2. Linear Systems
(T − λk I)N(T − λj I) = N(T − λj I)
Note that dim R(B − λq I) < dim E, and R(B − λq I) is (positively) invari-
ant under B. Applying our assumption to B|R(B−λq I) : R(B − λq I) →
R(B − λq I), we get a decomposition of R(B − λq I) into the generalized
eigenspaces of B|R(B−λq I) . By the second lemma, these are just
so
theoryofodes July 4, 2007 13:20 Page 37
q q
X X
nk = dim E = dim N(B − λk I),
k=1 k=1
We’ve seen that the space on which a linear operator acts can be decom-
posed into the direct sum of generalized eigenspaces of that operator.
The operator maps each of these generalized eigenspaces into itself,
and, consequently, solutions of the differential equation starting in a
generalized eigenspace stay in that generalized eigenspace for all time.
Now we will see how the solutions within such a subspace behave by
seeing how the operator behaves on this subspace.
It may seem like nothing much can be said in general since, given a
theoryofodes July 4, 2007 13:20 Page 38
2. Linear Systems
Proof. Obviously these vectors span Z(x); the question is whether they
are linearly independent. If they were not, we could write down a linear
combination α1 S p1 x + · · · + αk S pk x, with αj ≠ 0 and 0 ≤ p1 < p2 <
· · · < pk ≤ nil(x) − 1, that added up to zero. Applying S nil(x)−p1 −1 to
this linear combination would yield α1 S nil(x)−1 x = 0, contradicting the
definition of nil(x).
theoryofodes July 4, 2007 13:20 Page 39
z1 + · · · + zk = 0. (2.5)
We will show that zj = 0 for each j. This will mean that the direct sum
Z(x1 ) ⊕ · · · ⊕ Z(xk ) exists.
Step 5: For each j, Szj = 0.
This is a consequence of Step 3, Step 4, and (2.4).
then by Step 5
39
theoryofodes July 4, 2007 13:20 Page 40
2. Linear Systems
let
This completes the proof of the first sentence in the theorem. The
second sentence follows similarly by induction.
40
theoryofodes July 4, 2007 13:20 Page 41
0 ··· ··· ··· 0
.. ..
1 . .
.. .. ..
0 . . .,
.. .. .. .. ..
. . . . .
0 ··· 0 1 0
with the corresponding basis being {x, Sx, . . . , S nil(x)−1 x}. Thus, if λ
is an eigenvalue of an operator T , then the restriction of T to a cyclic
subspace of T − λI on the generalized eigenspace N(T − λI) can be
theoryofodes July 4, 2007 13:20 Page 42
2. Linear Systems
by the matrix
a −b 0 0 ··· ··· ··· ··· 0 0
b a 0 0 ··· ··· ··· ··· 0 0
.. .. .. .. .. ..
. . . .
1 0 . .
.. .. .. .. .. ..
0 1 . . . . . .
..
..
.. .. .. .. .. ..
0 0 . . . . . . .
.
. (2.8)
.. .. .. .. .. .. ..
..
0 0 . . . . . . . .
.. .. .. .. .. .. .. ..
. . . . . . . . 0 0
.. .. .. .. .. .. .. ..
. . . . . .
. . 0 0
0 0 ··· ··· 0 0 1 0 a −b
0 0 ··· ··· 0 0 0 1 b a
λ
λ
1
1 λ
" #
λ
(2.9)
1 λ h i
λ
h i
λ
..
.
if the eigenvalue λ is real, with blocks of the form (2.7) running down
the diagonal. If the eigenvalue is complex, then the matrix representa-
tion is similar to (2.9) but with blocks of the form (2.8) instead of the
form (2.7) on the diagonal.
Finally, the matrix representation of the entire operator is block
diagonal, with blocks of the form (2.9) (or its counterpart for complex
eigenvalues). This is called the real canonical form. If we specify the
order in which blocks should appear, then matrices are similar if and
only if they have the same real canonical form.
42
theoryofodes July 4, 2007 13:20 Page 43
; ; , (b ≠ 0).
1 λ 0 µ b a
Computing etA
and 0’s elsewhere, and N has M’s off-diagonal 1’s and 2×2 identity ma-
trices. If you consider the restrictions of S and N to each of the cyclic
subspaces of A−λI into which the generalized eigenspace N(A−λI) of
A is decomposed, you’ll probably be able to see that these restrictions
commute. As a consequence of this fact (and the way Rn can be rep-
resented in terms of these cyclic subspaces), S and N commute. Thus
etM = etS etN .
43
theoryofodes July 4, 2007 13:20 Page 44
2. Linear Systems
and
" #
0 0
0 0
" #
1 0
..
.
0 1
7→
.. ..
. .
" # " #
1 0 0 0
0 1 0 0
" #
1 0
0 1
" #
t 0 ..
.
0 t
.
.. .. ..
. . .
" # " # " #
t m /m! 0 t 0 1 0
···
0 t m /m! 0 t 0 1
theoryofodes July 4, 2007 13:20 Page 45
" #
λ 0
•
0 µ
" #
λ 0
•
1 λ
" #
a −b
• , (b ≠ 0)
b a
theoryofodes July 4, 2007 13:20 Page 46
2. Linear Systems
u2
1
" #
λ 0
A=
0 µ
(λ < 0 < µ) b u1
saddle
u2
2
" #
λ 0
A=
0 µ
(λ < µ < 0) b u1
stable node
u2
3
" #
λ 0
A=
0 µ
(λ = µ < 0) b u1
stable node
46
theoryofodes July 4, 2007 13:20 Page 47
u2
4
" #
λ 0
A=
0 µ
(0 < µ < λ) b u1
unstable node
u2
5
" #
λ 0
A=
0 µ
(0 < λ = µ) b u1
unstable node
u2
6
" #
λ 0 b
A=
0 µ
(λ < µ = 0) b u1
degenerate b
47
theoryofodes July 4, 2007 13:20 Page 48
2. Linear Systems
u2
7
" #
λ 0 b
A=
0 µ
(0 = µ < λ) b u1
degenerate b
u2
8
" #
λ 0 b b b
A=
0 µ
(0 = µ = λ) b b b u1
b b b
degenerate
u2
9
" #
λ 0
A=
1 λ
(λ < 0) b u1
stable node
48
theoryofodes July 4, 2007 13:20 Page 49
u2
10
" #
λ 0
A=
1 λ
(0 < λ) b u1
unstable node
u2
11
" #
λ 0 b
A=
1 λ
(λ = 0) b u1
b
degenerate
u2
12
" #
a −b
A=
b a
(a < 0 < b) b u1
stable spiral
49
theoryofodes July 4, 2007 13:20 Page 50
2. Linear Systems
u2
13
" #
a −b
A=
b a
(b < 0 < a) b u1
unstable spiral
u2
14
" #
a −b
A=
b a
(a = 0, b > 0) b u1
center
Parameter Plane
50
theoryofodes July 4, 2007 13:20 Page 51
center
/
4
2
τ
=
δ
stable node unstable node
degenerate degenerate τ
saddle
and
E c = N(A)⊕
M M
Re u u ∈ N(A − λI) ⊕
Im u u ∈ N(A − λI) .
Re λ=0
Re λ=0
Im λ≠0 Im λ≠0
51
theoryofodes July 4, 2007 13:20 Page 52
2. Linear Systems
From our previous study of the real canonical form, we know that
Rn = E u ⊕ E s ⊕ E c .
| · |.
Given an arbitrary norm N, and letting xi be the projection of x ∈
Rn onto the ith standard basis vector ei , note that
n
X n
X n
X
N(x) = N xi ei ≤ |xi |N(ei ) ≤ |x|N(ei )
i=1 i=1 i=1
n
X
≤ N(ei ) |x|.
i=1
theoryofodes July 4, 2007 13:20 Page 53
and
\
Ec = x ∈ Rn lim |ect etA x| = lim |e−ct etA x| = 0 . (2.13)
t↓−∞ t↑∞
c>0
tA
the structure of the real canonical form, we know that Pi e x is either
of the form
p(t)eλt , (2.14)
where p(t) is a polynomial in t and λ ∈ R is an eigenvalue of A, or of
the form
p(t)eat (α cos bt + β sin bt), (2.15)
where p(t) is a polynomial in t, a+bi ∈ C \ R is an eigenvalue of A, and
α and β are real constants. Furthermore, we know that the constant
λ or a is positive if Pi corresponds to a vector in E u , is negative if Pi
corresponds to a vector in E s , and is zero if Pi corresponds to a vector
in E c .
Now, let x ∈ Rn be given. Suppose first that x ∈ E s . Then each
Pi etA x is either identically zero or has as a factor a negative exponen-
tial whose constant is the real part of an eigenvalue of A that is to the
left of the imaginary axis in the complex plane. Let σ (A) be the set of
eigenvalues of A, and set
max Re λ λ ∈ σ (A) and Re λ < 0
c= .
2
53
theoryofodes July 4, 2007 13:20 Page 54
2. Linear Systems
so
lim sup kect etA xk = ∞.
t↑∞
The preceding two paragraphs showed that (2.12) is correct. By
applying this fact to the time-reversed problem ẋ = −Ax, we find that
(2.11) is correct, as well. We now consider (2.13).
If x ∈ E c , then for each i, Pi etA x is either a polynomial or the
product of a polynomial and a periodic function. If c > 0 and we
multiply such a function of t by ect and let t ↓ −∞ or we multiply it by
e−ct and let t ↑ ∞, then the result converges to zero.
If, on the other hand, x ∉ E c then for some i, Pi etA x contains a
nontrivial exponential term. If c > 0 is sufficiently small then either
ect Pi etA x diverges as t ↓ −∞ or e−ct Pi etA x diverges as t ↑ ∞. This
completes the verification of (2.13).
Theorem
(a) The origin is a source for the equation ẋ = Ax if and only if for a
given norm k · k there are constants k, b > 0 such that
theoryofodes July 4, 2007 13:20 Page 55
Exponential Decay
(b) The origin is a sink for the equation ẋ = Ax if and only if for a given
norm k · k there are constants k, b > 0 such that
Proof. The “if” parts are a consequence of the previous theorem. The
“only if” parts follow from the proof of the previous theorem.
If " #
x1 (t)
x(t) =
x2 (t)
is a solution of ẋ = Ax, then
" #" #
d h i −1/4 0 x1
Theorem
(a) If etA is an expansion then there is some norm k · k and some con-
stant b > 0 such that
(b) If etA is a contraction then there is some norm k · k and some con-
stant b > 0 such that
theoryofodes July 4, 2007 13:20 Page 56
2. Linear Systems
Proof. The idea of the proof is to pick a basis with respect to which A
is represented by a matrix like the real canonical form but with some
small constant ε > 0 in place of the off-diagonal 1’s. (This can be done
by rescaling.) If the Euclidean norm with respect to this basis is used,
the desired estimates hold. The details of the proof may be found in
Chapter 7, §1, of Hirsch and Smale.
Exercise 9
(a) Show that if etA and etB are both contractions on Rn , and BA =
AB, then et(A+B) is a contraction.
(b) Give a concrete example that shows that (a) can fail if the as-
sumption that AB = BA is dropped.
one of the following alternatives holds:
(c) there exist constants M, N > 0 such that M < |x(t)| < N for
all t ∈ R.”
Is what they ask you to prove true? If so, prove it. If not,
determine what other possible alternatives exist, and prove that
you have accounted for all possibilities.
ẋ = A(t)x. (2.16)
56
theoryofodes July 4, 2007 13:20 Page 57
Solution Formulas
ẋ = a(t)x (2.17)
for the solution of (2.17) that satisfies the initial condition x(t0 ) = x0 .
It seems like the analogous formula for the solution of (2.16) with
initial condition x(t0 ) = x0 should be
Rt
A(τ) dτ
x(t) = e t0 x0 . (2.18)
" #
0 0
A(t) =
1 t
and
0 0
" # " # # 2 2 "
t t 2 /2 2
t
1 0 t 0 02 0 0
e = + + + ···
0 1 2 2/t 1 2! 2/t 1
" # " #
1 0 2 0 0 1 0
= + et /2 − 1 = 2 t 2 /2 2 /2
.
0 1 2/t 1 t
e −1 et
ẋ1 = 0
ẋ2 = x1 + tx2
57
theoryofodes July 4, 2007 13:20 Page 58
2. Linear Systems
theoryofodes July 4, 2007 13:20 Page 59
Note that this proof also shows that if the Wronskian of n solutions
of (2.16) is zero for some t, then it is zero for all t.
59
theoryofodes July 4, 2007 13:20 Page 60
2. Linear Systems
Initial-Value Problems
To verify this, note that
d d
x= (X(t)[X(t0 )]−1 v) = A(t)X(t)[X(t0 )]−1 v = A(t)x,
dt dt
and
x(t0 ) = G(t0 , t0 )v = X(t0 )[X(t0 )]−1 v = v.
Inhomogeneous Equations
In light of the results from the previous section when f was identi-
cally zero, it’s reasonable to look for a solution x of (2.19) of the form
x(t) := G(t, t0 )y(t), where G is as before, and y is some vector-valued
function.
Note that
theoryofodes July 4, 2007 13:20 Page 61
since G(t, t0 )G(t0 , s) = G(t, s). This is called the Variation of Con-
stants formula or the Variation of Parameters formula.
2.10 Nearly Autonomous Linear Systems
Suppose A(t) is, in some sense, close to a constant matrix A. The ques-
tion we wish to address in this section is the extent to which solutions
of the nonautonomous system
ẋ = A(t)x (2.21)
ẋ = Ax. (2.22)
theoryofodes July 4, 2007 13:20 Page 62
2. Linear Systems
Proof. The proof is very similar to the proof of the standard Gronwall
inequality. The details are left to the reader.
The first main result deals with the case when A(t) converges to A
sufficiently quickly as t ↑ ∞.
Proof. Let t0 be such that (2.23) holds. Given a solution x of (2.21), let
f (t) = (A(t)−A)x(t), and note that x satisfies the constant-coefficient
62
theoryofodes July 4, 2007 13:20 Page 63
inhomogeneous problem
ẋ = Ax + f (t). (2.24)
Setting X(t) = |x(t)|, Φ(t) = MkA(t) − Ak, and C = M|x(t0 )|, and
applying the generalized Gronwall inequality, we find that
Rt
M kA(s)−Ak ds
|x(t)| ≤ M|x(t0 )|e t0
.
The next result deals with the case when the origin is a sink for
(2.22). Will the solutions of (2.21) also all converge to the origin as
t ↑ ∞? Yes, if kA(t) − Ak is sufficiently small.
Theorem Suppose all the eigenvalues of A have negative real part. Then
there is a constant ε > 0 such that if kA(t) − Ak ≤ ε for all t sufficiently
large then every solution of (2.21) converges to 0 as t ↑ ∞.
Proof. Since the origin is a sink, we know that we can choose constants
k, b > 0 such that ketA k ≤ ke−bt for all t ≥ 0. Pick a constant ε ∈
(0, b/k), and assume that there is a time t0 ∈ R such that kA(t) − Ak ≤
ε for every t ≥ t0 .
63
theoryofodes July 4, 2007 13:20 Page 64
2. Linear Systems
|x(t)| ≤ k|x(t0 )|e(kε−b)(t−t0 )
for all t ≥ t0 . Since ε < b/k, this inequality implies that x(t) → 0 as
t ↑ ∞.
64
theoryofodes July 4, 2007 13:20 Page 65
then the eigenvalues of A(t) both have negative real part for every
t ∈ R, but " #
− cos t
x(t) := et/2 ,
sin t
which becomes unbounded as t → ∞, is a solution to (2.21).
We now consider
ẋ = A(t)x (2.26)
d d
X̃(t) = X(t + T ) = X ′ (t + T ) = A(t + T )X(t + T ) = A(t)X̃(t),
dt dt
theoryofodes July 4, 2007 13:20 Page 66
2. Linear Systems
where
X∞
(−M)k
log(I + M) := − ,
k=1
k
X(t) = P (t)etB .
66
theoryofodes July 4, 2007 13:20 Page 67
does. They depend only on A(t). To see this, let X(t) and Y (t) be
fundamental matrices with corresponding monodromy matrices C and
D. Because X(t) and Y (t) are fundamental matrices, there is a non-
singular constant matrix S such that Y (t) = X(t)S for all t ∈ R. In
particular, Y (0) = X(0)S and Y (T ) = X(T )S. Thus, C =
This means that the monodromy matrices are similar and, therefore,
have the same eigenvalues.
theoryofodes July 4, 2007 13:20 Page 68
2. Linear Systems
matrix X(t). Let x(t) = X(t)x0 . Then, clearly, x solves (2.26). The
power series formula for the matrix exponential implies that x0 is an
eigenvector of etB with eigenvalue eλt . Hence,
Let x solve (2.26), and let y(t) = [P (t)]−1 x(t), where P is as defined
previously. Then
d d
[P (t)y(t)] = x(t) = A(t)x(t) = A(t)P (t)y(t)
dt dt
= A(t)X(t)e−tB y(t).
But
d
[P (t)y(t)] = P ′ (t)y(t) + P (t)y ′ (t)
dt
= [X ′ (t)e−tB − X(t)e−tB B]y(t) + X(t)e−tB y ′ (t)
= A(t)X(t)e−tB y(t) − X(t)e−tB By(t) + X(t)e−tB y ′ (t),
so
X(t)e−tB y ′ (t) = X(t)e−tB By(t),
Theorem If all the Floquet exponents of (2.26) have negative real parts
then all solutions of (2.26) converge to 0 as t ↑ ∞.
theoryofodes July 4, 2007 13:20 Page 69
and ZT
1 2π i
λ1 + · · · + λn ≡ trace A(t) dt mod (2.29)
T 0 T
X n
Y
W (t) = ǫ(σ ) Xi,σ (i) ,
σ ∈Sn i=1
theoryofodes July 4, 2007 13:20 Page 70
2. Linear Systems
Xn
dW (t)
= Aj,j (t) det X(t) = [trace A(t)]W (t).
dt j=1
Thus, Rt Rt
W (t) = e 0 trace A(s) ds W (0) = e 0 trace A(s) ds .
In particular,
RT
trace A(s) ds
e0 = W (T ) = det X(T ) = det(P (T )eT B ) = det(P (0)eT B )
= det eT B = det C = ρ1 ρ2 · · · ρn .
that becomes unbounded as t ↑ ∞.
70
theoryofodes July 4, 2007 13:20 Page 71
3
Topological Dynamics
Orbits
Invariant Sets
theoryofodes July 4, 2007 13:20 Page 72
3. Topological Dynamics
Limit Sets
Lemma If, for each A ∈ Ω, we let A represent the topological closure of
A in Ω, then
\
ω(x) = γ + (ϕ(τ, x)) (3.1)
τ∈R
and
\
α(x) = γ − (ϕ(τ, x)). (3.2)
τ∈R
y ∈ γ + (ϕ(τ, x)).
theoryofodes July 4, 2007 13:20 Page 73
y ∈ γ + (ϕ(k, x))
so we can pick zk ∈ γ + (ϕ(k, x)) such that |zk − y| < 1/k. Since
zk ∈ γ + (ϕ(k, x)), we can pick sk ≥ 0 such that zk = ϕ(sk , ϕ(k, x)). If
we set tk = k + sk , we see that tk ≥ k, so the sequence t1 , t2 , . . . goes to
we know that
\
γ + (ϕ(τ, x)) ⊆ ω(x).
τ∈R
theoryofodes July 4, 2007 13:20 Page 74
3. Topological Dynamics
theoryofodes July 4, 2007 13:20 Page 75
each fixed k ∈ N)
ϕ(t, x) t ∈ [sk , tk ]
Examples of empty ω-limit sets are easy to find. Consider, for ex-
ample, the one-dimensional dynamical system ϕ(t, x) := x + t (gener-
ated by the differential equation ẋ = 1.
Examples of dynamical systems with nonempty, noncompact, dis-
connected ω-limit sets are a little harder to find. Consider the planar
autonomous system
ẋ = −y(1 − x 2 )
ẏ = x + y(1 − x 2 ).
theoryofodes July 4, 2007 13:20 Page 76
3. Topological Dynamics
Ω −
−−−→ Ω
hy yh
ψ(t,·)
Θ −
−−−→ Θ
theoryofodes July 4, 2007 13:20 Page 77
and " #
cos 2t − sin 2t
ψ(t, y) = y,
sin 2t cos 2t
and " #
0 −2
ẏ = y.
2 0
The functions h(x) = x and α(t, x) = 2t show that these two flows
are topologically equivalent. But these two flows are not topologically
conjugate, since, by setting t = π we see that any function h : R2 → R2
satisfying (3.5) would have to satisfy h(x) = h(−x) for all x, which
would mean that h is not invertible.
Because of examples like this, topological equivalence seems to be
the preferred concept when dealing with flows. The following theorem,
however, shows that in a neighborhood of a regular point, a smooth
flow satisfies a local version of C r -conjugacy with respect to a unidi-
rectional, constant-velocity flow.
theoryofodes July 4, 2007 13:20 Page 78
3. Topological Dynamics
in W .
have elapsed.
Step 1: ϕ(·, p) is C r +1 .
We know that
d
ϕ(t, p) = f (ϕ(t, p)). (3.8)
dt
If f is continuous then, since ϕ(·, p) is continuous, (3.8) implies that
ϕ(·, p) is C 1 . If f is C 1 , then the previous observation implies that
d
ϕ(·, p) is C 1 . Then (3.8) implies that dt ϕ(t, p) is the composition of
C 1 functions and is, therefore, C 1 ; this means that ϕ(·, p) is C 2 . Con-
tinuing inductively, we see that, since f is C r , ϕ(·, p) is C r +1 .
Step 2: ϕ(t, ·) is C r .
This is a consequence of applying differentiability with respect to pa-
rameters inductively.
Step 3: G is C r .
This is a consequence of Steps 1 and 2 and the formula for G in terms
of ϕ.
78
theoryofodes July 4, 2007 13:20 Page 79
and
∂G(y)
∂
∂p
= ϕ(0, p) · ek = ek = ek ,
∂yk y=0 ∂p
p=0
∂p
p=0
for k ≠ 1, we have
DG(0) =
αe1 e2 ··· en
,
79
theoryofodes July 4, 2007 13:20 Page 80
3. Topological Dynamics
In the previous section, we saw that all the “interesting” local behavior
of flows occurs near equilibrium points. One important aspect of the
behavior of flows has to do with whether solutions that start near a
given solution stay near it for all time and/or move closer to it as time
elapses. This question, which is the subject of stability theory, is not
just of interest when the given solution corresponds to an equilibrium
solution, so we study it–initially, at least–in a fairly broad context.
Definitions
of (3.9) and |x(t0 ) − x(t0 )| < δ then |x(t) − x(t)| < ε for all t ≥ t0 .
theoryofodes July 4, 2007 13:20 Page 81
Definitions of Stability
Note that the definition implies that stable sets are positively in-
variant.
asymptotically stable.)
Examples
b x
Orbits are ellipses with major axis along the y-axis. The equilibrium
solution at the origin is Lyapunov stable even though nearby orbits
81
theoryofodes July 4, 2007 13:20 Page 82
3. Topological Dynamics
or, equivalently, b x
ẋ = −(x 2 + y 2 )y
ẏ = (x 2 + y 2 )x.
The solution moving around the unit circle is not Lyapunov stable,
since nearby solutions move with different angular velocities. It is,
however, orbitally stable. Also, the set consisting of the unit circle is
stable.
y
3
ṙ = r (1 − r )
θ̇ = sin2 (θ/2).
b b x
The constant solution (x, y) = (1, 0) is not Lyapunov stable and the
set {(1, 0)} is not stable. However, every solution beginning near (1, 0)
converges to (1, 0) as t ↑ ∞. This shows that it is not redundant to
require Lyapunov stability (or stability) in the definition of asymptotic
stability of a solution (or a set).
ẋ = f (x) (3.10)
82
theoryofodes July 4, 2007 13:20 Page 83
Definitions of Stability
Theorem Let x be a function that solves (3.10), and let A(x) be the
corresponding orbit. Then:
We will not prove this theorem, but we will note that parts 1 and 2
are immediate results of the definitions (even if we were dealing with
a nonautonomous equation) and part 4 is also an immediate result of
the definitions (even if A were an arbitrary set).
83
theoryofodes July 4, 2007 13:20 Page 84
3. Topological Dynamics
1. ẋ = x, Ω = R, x(t) := 0
2. ẋ = x, Ω = R, x(t) := et
5. ẋ = x, Ω = (0, ∞), x(t) := et
84
theoryofodes July 4, 2007 13:20 Page 85
15. ẋ = 0, Ω = R, x(t) := 0
16. ẋ = 1, Ω = R, x(t) := t
ẋ = f (x) (3.11)
u̇ = f (u + x0 ) = f (x0 ) + Df (x0 )u + R(u) = Au + R(u), (3.12)
ẋ = A(x − x0 ) (3.14)
for x near x0 . Equation (3.13) (or sometimes (3.14)) is called the lin-
earization of (3.11) at x0 .
Now, we’ve defined (several types of) stability for equilibrium solu-
tions of (3.11) (as well as for other types of solutions and sets), but we
haven’t really given any tools for determining stability. In this lecture
we present one such tool, using the linearized equation(s) discussed
above.
85
theoryofodes July 4, 2007 13:20 Page 86
3. Topological Dynamics
for every v ∈ V .
Proof. Let n = dim V , and pick ε > 0 so small that all the eigenvalues
86
theoryofodes July 4, 2007 13:20 Page 87
of L have real part greater than c + nε. Choose a basis {v1 , . . . , vn } for
V that puts L in “modified” real canonical form with the off-diagonal
1’s replaced by ε’s, and let h·, ·i be the inner product associated with
this basis (i.e. hvi , vj i = δij ) and let k · k be the induced norm on V .
Pn
Given v = i=1 αi vi ∈ V , note that (if L = (ℓij ))
n
X n X
X n
X n X
X α2i + α2j
hv, Lvi = ℓii α2i + ℓij αi αj ≥ ℓii α2i − ε
i=1 i=1 j≠i i=1 i=1 j≠i
2
n
X n
X n
X n
X
≥ ℓii α2i − nεα2i = (ℓii − nε)α2i ≥ cα2i = ckvk2 .
i=1 i=1 i=1 i=1
Note that applying this theorem to −L also tells us that, for some
inner product,
hv, Lvi ≤ ckvk2 (3.15)
d
kx(t)k2 = 2hx(t), ẋ(t)i = 2hx(t), f (x(t))i
dt
= 2hx(t), Ax(t)i + 2hx(t), R(x(t))i
≤ 2ckx(t)k2 + 2kx(t)k · kR(x(t))k
≤ 2ckx(t)k2 − ckx(t)k2 = ckx(t)k2 .
theoryofodes July 4, 2007 13:20 Page 88
3. Topological Dynamics
The proof of the second proposition will be geometric and will con-
tain ideas that will be used to prove stronger results later in this text.
for all v ∈ E + , and define an inner product h·, ·i− (and induced norm
k · k− ) on E − such that
theoryofodes July 4, 2007 13:20 Page 89
Br
Kr
b v
(See Figure 1.) Suppose x = v + w is a solution of (3.16) that starts in
Kr at time t = 0. For as long as the solution remains in Kr ,
d
kvk2 = 2hv, v̇i = 2hv, A+ vi + 2hv, R + (v, w)i
dt
≥ 2akvk2 − 2kvk · kR + (v, w)k ≥ 2akvk2 − 2εkvk · kv + wk
1/2 √
= 2akvk2 − 2εkvk kvk2 + kwk2 ≥ 2akvk2 − 2 2εkvk2
√
= 2(a − 2ε)kvk2 ,
and
d
kwk2 = 2hw, ẇi = 2hw, A− wi + 2hw, R − (v, w)i
dt
≤ 2bkwk2 + 2kwk · kR − (v, w)k
≤ 2bkwk2 + 2εkwk · kv + wk
1/2
= 2bkwk2 + 2εkwk kvk2 + kwk2
√
≤ 2bkvk2 + 2 2εkvk2
√
= 2(b + 2ε)kvk2 .
89
theoryofodes July 4, 2007 13:20 Page 90
3. Topological Dynamics
The first estimate says that as long as the solution stays in Kr , kvk
grows exponentially, which means that the solution must eventually
leave Kr . Combining the first and second estimates, we have
d √
(kvk2 − kwk2 ) ≥ 2(a − b − 2 2ε)kvk2 > 0,
dt
nonautonomous equations. Let Ω ⊆ Rn be open and contain the ori-
gin, and suppose that f : R × Ω → Rn is a continuously differentiable
function. Suppose, furthermore, that f (t, 0) = 0 for every t ∈ R, so
x(t) := 0 is a solution of the equation
theoryofodes July 4, 2007 13:20 Page 91
Definition If V : R × D is continuously differentiable then its orbital
derivative (with respect to (3.17)) is the function V̇ : R × D → R given
by the formula
∂V ∂V
V̇ (t, x) := (t, x) + (t, x) · f (t, x).
∂t ∂x
(Here “∂V (t, x)/∂x” represents the gradient of the function V (t, ·).)
d
V (t, x(t)) = V̇ (t, x(t)).
dt
A function whose orbital derivative is always nonpositive is sometimes
called a Lyapunov function.
theoryofodes July 4, 2007 13:20 Page 92
3. Topological Dynamics
theoryofodes July 4, 2007 13:20 Page 93
Y (x(t)) t ≥ t0
V̇ (t, x(t)) t ≥ t0
d
V (t, x(t)) = V̇ (t, x(t)) ≤ −δ (3.19)
dt
Let r > 0 be given, and let
wr = min W (p) p ∈ D \ B(0, r ) ,
93
theoryofodes July 4, 2007 13:20 Page 94
3. Topological Dynamics
g ′ (t)
f (t, x) := x,
g(t)
Show that, for x near 0, V (t, x) is positive definite and V̇ (t, x) is
negative definite, but the solution 0 of (3.17) is not asymptotically
stable.)
In the previous two lectures, we have talked about two different tools
that can be used to prove that an equilibrium point x0 of an autono-
mous system
ẋ = f (x) (3.20)
theoryofodes July 4, 2007 13:20 Page 95
for all x ∈ Rn . Pick r > 0 small enough that kf (x) − Axk ≤ (c/2)kxk
whenever kxk ≤ r , let
D = x ∈ Rn kxk ≤ r ,
so V̇ is negative definite.
On the other hand, there are very simple examples to illustrate that
the direct method works in some cases where linearization doesn’t. For
example, consider ẋ = −x 3 on R. The equilibrium point at the origin
is not hyperbolic, so linearization fails to determine stability, but it is
easy to check that x 2 is positive definite and has a negative definite
orbital derivative, thus ensuring the asymptotic stability of 0.
theoryofodes July 4, 2007 13:20 Page 96
3. Topological Dynamics
of the phase portrait indicates that orbits circle the origin in the coun-
terclockwise direction, but it is not obvious whether they spiral in, spi-
ral out, or move on closed curves.
The simplest potential Lyapunov function that often turns out to be
useful is the square of the standard Euclidean norm, which in this case
is V := x 2 + y 2 . The orbital derivative is
V̇ = 2x ẋ + 2y ẏ = 2x 5 y − 2xy − 2x 4 . (3.21)
For some points (x, y) near the origin (e.g., (δ, δ)) V̇ < 0, while for
other points near the origin (e.g., (δ, −δ)) V̇ > 0, so this function
doesn’t seem to be of much use.
Sometimes when the square of the standard Euclidean norm doesn’t
work, some other homogeneous quadratic function does. Suppose we
try V := x 2 + αxy + βy 2 , with α and β to be determined. Then
Setting (x, y) = (δ, −δ2 ) for δ positive and small, we see that V̇ is not
going to be negative semidefinite, no matter what we pick α and β to
be.
If these quadratic functions don’t work, maybe something custom-
ized for the particular equation might. Note that the right-hand side
of the first equation in (3.21) sort of suggests that x 3 and y should be
treated as quantities of the same order of magnitude. Let’s try V :=
x 6 + αy 2 , for some α > 0 to be determined. Clearly, V is positive
definite, and
theoryofodes July 4, 2007 13:20 Page 97
ap bq
ab ≤ + , (3.22)
p q
1 1
+ = 1. (3.23)
p q
xp bq
g(x) := + − xb.
p q
Note that g is continuous, and g ′ (x) = x p−1 − b for every x ∈ (0, ∞).
Since limx↓0 g ′ (x) = −b < 0, limx↑∞ g ′ (x) = ∞, and g ′ is increasing on
(0, ∞), we know that g has a unique minimizer at x0 = b 1/(p−1) . Thus,
for every x ∈ [0, ∞) we see, using (3.23), that
!
b p/(p−1) bq 1 1
g(x) ≥ g(b 1/(p−1)
)= + − b p/(p−1) = + − 1 b q = 0.
p q p q
|x|6 5|y|18/5 1 5
|xy 3 | = |x||y|3 ≤ + ≤ x6 + y 2
6 6 6 6
97
theoryofodes July 4, 2007 13:20 Page 98
3. Topological Dynamics
if |y| ≤ 1, so
5 6 13 2
V ≥ x + y
6 6
if |y| ≤ 1. Also,
3|x|8 5|y|24/5 3 5
| − x 3 y 3 | = |x|3 |y|3 ≤ + ≤ x8 + y 4
8 8 8 8
if |y| ≤ 1, and
" #
6 2 6 3|x|8
2 |y|8 9 8 3 8 9 3 4
|3x y | = 3|x| |y| ≤ 3 + = x + y ≤ x8 + y
4 4 4 4 4 64
if |y| ≤ 1/2, so, in a neighborhood of 0, V is positive definite and V̇ is
negative definite, which implies that 0 is asymptotically stable.
theoryofodes July 4, 2007 13:20 Page 99
99
theoryofodes July 4, 2007 13:20 Page 100
theoryofodes July 4, 2007 13:20 Page 101
4
Conjugacies
The Principle of Linearized Stability indicates one way in which the flow
near a singular point of an autonomous ODE resembles the flow of its
linearization. The Hartman-Grobman Theorem gives further insight
into the extent of the resemblance; namely, there is a local topological
conjugacy between the two. We will spend the next 5 sections talking
about the various forms of this theorem and their proofs. This amount
of attention is justified not only by the significance of the theorem but
the general applicability of the techniques used to prove it.
Let Ω ⊆ Rn be open and let f : Ω → Rn be continuously differen-
tiable. Suppose that x0 ∈ Ω is a hyperbolic equilibrium point of the
autonomous equation
ẋ = f (x). (4.1)
Let B = Df (x0 ), and let ϕ be the (local) flow generated by (4.1). The
version of the Hartman-Grobman Theorem we’re primarily interested
in is the following.
theoryofodes July 4, 2007 13:20 Page 102
4. Conjugacies
is defined on Cb1 (Rn ). We will not define a norm on Cb1 (Rn ), but will
often use Lip, which is not a norm, to describe the size of elements of
Cb1 (Rn ).
F (h(x)) = h(Ax)
theoryofodes July 4, 2007 13:20 Page 103
kLvk ≤ ckvk
for every v ∈ V .
Proof. As in the previous lemma, the norm will be the Euclidean norm
corresponding to the modification of the real canonical basis that gives
a matrix representation of L that has ε’s in place of the off-diagonal 1’s.
With respect to this basis, it can be checked that
LT L = D + R(ε),
for every v ∈ V (where h·, ·i is the inner product that induces k·k).
kLvk ≥ ckvk
for some norm k · k. This follows trivially in the case when c ≤ 0, and
when c > 0 it follows by applying the lemma to L−1 (which exists, since
0 is not an eigenvalue of L).
103
theoryofodes July 4, 2007 13:20 Page 104
4. Conjugacies
Using the notation developed when we were deriving the real canonical
form, let
M
−
E = N(A − λI) ⊕
λ∈(−a,a)
M M
Re u u ∈ N(A − λI) ⊕ Im u u ∈ N(A − λI) ,
|λ|<a
|λ|<a
Im λ≠0 Im λ≠0
and let
M
E+ = N(A − λI) ⊕
λ∈(−∞,−a−1 )∪(a−1 ,∞)
M
M
Re u u ∈ N(A − λI) ⊕
Im u u ∈ N(A − λI) .
|λ|>a−1
|λ|>a−1
Im λ≠0 Im λ≠0
theoryofodes July 4, 2007 13:20 Page 105
This is the norm on Rn that we will use throughout our proof of the
(global) Hartman-Grobman Theorem (for maps). Note that kxk = kxk−
if x ∈ E − , and kxk = kxk+ if x ∈ E + .
Recall that we equipped Cb0 (Rn ) with the norm k · k0 defined by the
formula
kwk0 := sup kw(x)k.
x∈Rn
The norm on Rn on the right-hand side of this formula is that given
in (4.2). Recall also that we will use the functional Lip defined by the
formula
kw(x1 ) − w(x2 )k
Lip(w) := sup
x1 ,x2 ∈Rn kx1 − x2 k
x1 ≠x2
x∈Rn
Since Rn = E − ⊕ E + , it follows that
w = P − ◦ w + P + ◦ w.
theoryofodes July 4, 2007 13:20 Page 106
4. Conjugacies
Choose, and fix, a function g ∈ C1b (Rn ) for which Lip(g) < ε. The
(global) Hartman-Grobman Theorem (for maps) will be proved by con-
structing a map Θ from Cb0 (Rn ) to Cb0 (Rn ) whose fixed points would be
precisely those objects v which, when added to the identity I, would
yield solutions h to the conjugacy equation
(A + g) ◦ h = h ◦ A, (4.3)
Lv = Ψ (v), (4.4)
Lv = v − A ◦ v ◦ A−1 =: (id −A)v.
Inverting L
theoryofodes July 4, 2007 13:20 Page 107
1
k(id −G)−1 kL(X,X) ≤ .
1−c
1
kHkL(X,X) ≤ .
1−c
1
k(id −G−1 )−1 kL(X,X) ≤ .
1−c
107
theoryofodes July 4, 2007 13:20 Page 108
4. Conjugacies
Since id −G = −G(id −G−1 ) = −(id −G−1 )G, it is not hard to check that
−(id −G−1 )−1 G−1 is the inverse of id −G and that
c
k − (id −G−1 )−1 G−1 kL(X,X) ≤ .
1−c
The first half of this lemma is useful for inverting small perturba-
tions of the identity, while the second half is useful for inverting large
perturbations of the identity. It should, therefore, not be too surprising
that we will apply the first half with G = A− and the second half with
G = A+ (since A compresses things in the E − direction and stretches
things in the E + direction).
First, consider A− . If w ∈ Cb0 (E − ), then
so the operator norm of A− is bounded by a. Applying the first half
of the lemma with X = Cb0 (E − ), G = A− , and c = a, we find that L− is
invertible, and its inverse has operator norm bounded by (1 − a)−1 .
Next, consider A+ . It is not hard to see that A+ is invertible, and
+ −1
(A ) w = A−1 ◦ w ◦ A. If w ∈ Cb0 (E + ), then (because the eigenvalues
of the restriction of A−1 to E + all have magnitude less than a)
theoryofodes July 4, 2007 13:20 Page 109
Recall that we are looking for fixed points v of the map Θ := L−1 ◦ Ψ ,
where Ψ (v) := g ◦ (I + v) ◦ A−1 . We have established that L−1 is a well-
defined linear map from Cb0 (Rn ) to Cb0 (Rn ) and that its operator norm
is bounded by (1−a)−1 . This means that Θ is a well-defined (nonlinear)
map from Cb0 (Rn ) to Cb0 (Rn ); furthermore, if v1 , v2 ∈ Cb0 (Rn ), then
1
kΘ(v1 )−Θ(v2 )k0 = kL−1 (Ψ (v1 )−Ψ (v2))k0 ≤ kΨ (v1 )−Ψ (v2 )k0
1−a
1
= kg ◦ (I + v1 ) ◦ A−1 − g ◦ (I + v2 ) ◦ A−1 k0
1−a
1
= sup kg(A−1 x + v1 (A−1 x)) − g(A−1 x + v2 (A−1 x))k
1 − a x∈Rn
ε
≤ sup k(A−1 x + v1 (A−1 x)) − (A−1 x + v2 (A−1 x))k
1 − a x∈Rn
ε
= sup kv1 (A−1 x) − v2 (A−1 x)k
1 − a x∈Rn
ε ε
= sup kv1 (y) − v2 (y)k = kv1 − v2 k0 .
1 − a y∈Rn 1−a
theoryofodes July 4, 2007 13:20 Page 110
4. Conjugacies
Injectivity
kAn zk = kAn x1 − An x2 k
= k(h(An x1 ) − v(An x1 )) − (h(An x2 ) − v(An x2 ))k
= kv(An x1 ) − v(An x2 )k ≤ 2kvk0 .
Because of the way the norm was chosen, we then know that for n ≥ 0
as n ↓ −∞. Hence, z = P − z + P + z = 0, so x1 = x2 .
110
theoryofodes July 4, 2007 13:20 Page 111
Surjectivity
kh(xk ) − yk ≤ 1
for every k. This implies that kh(xk )k ≤ kyk + 1, which in turn implies
that kxk k ≤ kyk + kvk0 + 1. Thus, the sequence (xk ) is bounded
and therefore has a subsequence (xkℓ ) converging to some point x0 ∈
Rn . By continuity of h, (h(xkℓ )) converges to h(x0 ), which means that
h(x0 ) = y. Hence, h(Rn ) is closed.
ẋ = f (x) (4.6)
111
theoryofodes July 4, 2007 13:20 Page 112
4. Conjugacies
ẋ = f˜(x) (4.7)
and
ẋ = Bx (4.8)
will also hold between (4.6) and (4.8) for x small.
Pick β : [0, ∞) → [0, 1] to be a C ∞ function satisfying
1 if s ≤ 1
β(s) =
0 if s ≥ 2,
and let C = sups∈[0,∞) |β′ (s)|. Given ε > 0, pick r > 0 so small that
ε
kDf (x) − Bk <
2C + 1
theoryofodes July 4, 2007 13:20 Page 113
ẋ = Bx + (f˜(x) − Bx),
we find that
Z1
g(x) = e(1−s)B [f˜(ϕ(s, x)) − Bϕ(s, x)] ds,
0
so
kg(x) − g(y)k
Z1
≤ ke(1−s)B kk(f˜(ϕ(s, x)) − Bϕ(s, x)) − (f˜(ϕ(s, y)) − Bϕ(s, y))k ds
0
Z1
≤ε ke(1−s)B kkϕ(s, x) − ϕ(s, y)k ds
0
Z1
≤ kx − ykε ke(1−s)B kke(kBk+ε)s − 1k ds,
0
Conjugacy for t = 1
theoryofodes July 4, 2007 13:20 Page 114
4. Conjugacies
Conjugacy for t ≠ 1
= ϕ(t, h(eB e−tB x)) = ϕ(t, h(e−tB eB x))) = h̃(eB x),
h̃ − I = ϕ(t, ·) ◦ h ◦ e−tB − I
= (ϕ(t, ·) − etB ) ◦ h ◦ e−tB + etB ◦ (h − I) ◦ e−tB =: v1 + v2 .
The fact that ϕ(t, x) and etB x agree for large x implies that ϕ(t, ·) −
etB is bounded, so v1 is bounded, as well. The fact that h−I is bounded
implies that v2 is bounded. Hence, h̃ − I is bounded.
The uniqueness part of the global Hartman-Grobman Theorem for
maps now implies that h and h̃ must be the same function. Using this
fact and substituting y = e−tB x in (4.10) yields
for every y ∈ Rn and every t ∈ Rn . This means that the flows gen-
erated by (4.8) and (4.7) are globally topologically conjugate, and the
flows generated by (4.8) and (4.6) are locally topologically conjugate.
114
theoryofodes July 4, 2007 13:20 Page 115
Constructing Conjugacies
ẋ = f (x) (4.11)
and
ẋ = g(x), (4.12)
Substituting (4.13) into the left-hand side of (4.14) and then replacing
ψ(t, x) by x, we get the equivalent equation
theoryofodes July 4, 2007 13:20 Page 116
4. Conjugacies
and
ẋ = bx (4.17)
for some constant C. Clearly, (4.19) does not define a topological con-
jugacy for a single constant C, because it fails to be injective on R;
however, the formula
x|x|a/b−1 if x ≠ 0
h(x) = (4.20)
0 if x = 0,
which is (4.13) for this example, shows that it does, in fact, define a
topological conjugacy when ab > 0. (Note that in no case is this a
C 1 -conjugacy, since either h′ (0) or (h−1 )′ (0) does not exist.)
116
theoryofodes July 4, 2007 13:20 Page 117
Constructing Conjugacies
• If b > 0 and a < 0, then (4.23) implies that h(1) and h(b) have
opposite signs even though 1 and b have the same sign; conse-
quently, the Intermediate Value Theorem yields a contradiction
to injectivity.
117
theoryofodes July 4, 2007 13:20 Page 118
4. Conjugacies
ż = F (z)
and
ż = Az
118
theoryofodes July 4, 2007 13:20 Page 119
Smooth Conjugacies
Hartman’s Example
where α > γ > 0 and ε ≠ 0. We will not cut off this vector field but will
where a = eα and c = e−γ . Note that a > ac > 1 > c > 0. The time-1
map B of the linearization of the differential equation is given by
x ax
B
y = acy .
y cz
theoryofodes July 4, 2007 13:20 Page 120
4. Conjugacies
whenever u, βu ∈ U. Then if |β| < 1 < |α| or |α| < 1 < |β|, j(u) = 0
for every u ∈ U.
Proof. If |β| < 1 < |α|, fix u ∈ U and apply (4.28) inductively to get
f (0, z) = 0 (4.32)
for every z near zero. Setting z = 0 in (4.27) and applying the Lemma
gives
h(x, 0) = 0 (4.33)
120
theoryofodes July 4, 2007 13:20 Page 121
Smooth Conjugacies
for every x near zero. Setting x = 0 in (4.26), using (4.32), and applying
the Lemma gives
g(0, z) = 0 (4.34)
for every z near zero. If we set z = 0 in (4.26), use (4.33), and then dif-
ferentiate both sides with respect to x, we get cgx (x, 0) = gx (ax, 0);
applying the Lemma yields
gx (x, 0) = 0 (4.35)
for every x near zero. Setting z = 0 in (4.34) and using (4.35), we get
g(x, 0) = 0 (4.36)
for every n ∈ N.
The existence of gx (0, z) and gz (x, 0) along with equations (4.34)
and (4.36) imply that an g(a−n x, z) and c −n g(x, c n z) stay bounded as
n ↑ ∞. Using this fact, and letting n ↑ ∞ in (4.39), we get
f (x, 0)h(0, z) = 0,
theoryofodes July 4, 2007 13:20 Page 122
4. Conjugacies
ẋ = f (x) (4.40)
and
n
X
λs = mk λk .
k=1
If Df (0) is not resonant, we say that it is nonresonant.
theoryofodes July 4, 2007 13:20 Page 123
Smooth Conjugacies
123
theoryofodes July 4, 2007 13:20 Page 124
theoryofodes July 4, 2007 13:20 Page 125
5
Invariant Manifolds
be matched up by smooth changes of variable. In this section, we will
discuss the Stable Manifold Theorem on an informal level and discuss
two different approaches to proving it.
Let f : Ω ⊆ Rn → Rn be C 1 , and let ϕ : R × Ω → Ω be the flow
generated by the differential equation
ẋ = f (x). (5.1)
theoryofodes July 4, 2007 13:20 Page 126
5. Invariant Manifolds
Note that:
s u
• Wloc (x0 ) ⊆ W s (x0 ), and Wloc (x0 ) ⊆ W u (x0 ).
s u
• Wloc (x0 ) and Wloc (x0 ) are both nonempty, since they each contain
x0 .
s u
• Wloc (x0 ) is positively invariant, and Wloc (x0 ) is negatively invari-
ant.
s u
• Wloc (x0 ) is not necessarily W s (x0 ) ∩ U, and Wloc (x0 ) is not nec-
essarily W u (x0 ) ∩ U.
s
Wloc (x0 ) is not necessarily invariant, since it might not be negatively
u
invariant, and Wloc (x0 ) is not necessarily invariant, since it might not
be positively invariant. They do, however, possess what is known as
relative invariance.
theoryofodes July 4, 2007 13:20 Page 127
s
Wloc (x0 ) is negatively invariant relative to U and is therefore invari-
u
ant relative to U. Wloc (x0 ) is positively invariant relative to U and is
therefore invariant relative to U.
Recall that a (k-)manifold is a set that is locally homeomorphic to
an open subset of Rk . Although the word “manifold” appeared in the
s u
names of Wloc (x0 ), Wloc (x0 ), W s (x0 ), and W u (x0 ), it is not obvious
from the defintions of these sets that they are, indeed, manifolds. One
of the consequences of the Stable Manifold Theorem is that, if U is
s u
sufficiently small, Wloc (x0 ) and Wloc (x0 ) are manifolds. (W s (x0 ) and
W u (x0 ) are what are known as immersed manifolds.)
For simplicity, let’s now assume that x0 = 0. Let E s be the stable
subspace of Df (0), and let E u be the unstable subspace of Df (0). If
f is linear, then W s (0) = E s and W u (0) = E u . The Stable Manifold
Theorem says that in the nonlinear case not only are the Stable and
Unstable Manifolds indeed manifolds, but they are tangent to E s and
E u , respectively, at the origin. This is information that the Hartman-
Grobman Theorem does not provide.
and
u
Wloc (0) = x + hu (x) x ∈ Uu .
Furthermore, not only do solutions of (5.1) in the stable manifold con-
verge to 0 as t ↑ ∞, they do so exponentially quickly. (A similar state-
ment can be made about the unstable manifold.)
Liapunov-Perron Approach
theoryofodes July 4, 2007 13:20 Page 128
5. Invariant Manifolds
A fixed point x of this functional will solve (5.2), will have a range
contained in the stable manifold, and will satisfy xs (0) = as . If we set
hs (as ) = xu (0) and define hs similarly for other inputs, the graph of
hs will be the stable manifold.
Hadamard Approach
theoryofodes July 4, 2007 13:20 Page 129
in other words,
Statements
ẋ = f (x). (5.5)
decomposition of Rn corresponding to the matrix Df (0). Then there is a
norm k · k on Rn , a number r > 0, and a C k function h : E s (r ) → E u (r )
such that h(0) = 0 and Dh(0) = 0 and such that the local stable mani-
s
fold Wloc (0) of 0 relative to B(r ) := E s (r ) ⊕ E u (r ) is the set
vs + h(vs ) vs ∈ E s (r ) .
Two immediate and obvious corollaries, which we will not state ex-
plicitly, describe the stable manifolds of other equilibrium points (via
translation) and describe unstable manifolds (by time reversal).
We will actually prove this theorem by first proving an analogous
theorem for maps (much as we did with the Hartman-Grobman Theo-
rem). Given a neighborhood U of a fixed point p of a map F , we can
define the local stable manifold of p (relative to U) as
n o
s
Wloc (p) := x ∈ U F j (x) ∈ U for every j ∈ N and lim F j (x) = p .
j↑∞
129
theoryofodes July 4, 2007 13:20 Page 130
5. Invariant Manifolds
s
Wloc (0) = vs + h(vs ) vs ∈ E s (r )
n o
= v ∈ B(r ) F j (v) ∈ B(r ) for every j ∈ N
n o
= v ∈ B(r ) F j (v) ∈ B(r ) and kF j (v)k ≤ µ̃ j kvk for all j ∈ N .
Preliminaries
The proof of the Stable Manifold Theorem for Maps will be broken up
into a series of lemmas. Before stating and proving those lemmas, we
need to lay a foundation by introducing some terminology and notation
theoryofodes July 4, 2007 13:20 Page 131
and
kAuu (0)vu k
m(Auu (0)) := inf > λ.
vu ≠0 kvu k
(The functional m(·) defined implicitly in the last formula is some-
times called the minimum norm even though it is not a norm.) For
a vector in v ∈ Rn , let kvk = max{kπs vk, kπu vk}. This will be
n
the norm on R that will be used throughout the proof. Note that
B(r ) := E (r ) ⊕ E u (r ) is the closed ball of radius r in Rn by this
s
norm.
Next, we choose r . Fix α > 0. Pick ε > 0 small enough that
and note that Wrs is the set of all points in B(r ) that produce forward
semiorbits (under the discrete dynamical system generated by F ) that
s
stay in B(r ) for all forward iterates. By definition, Wloc (0) ⊆ Wrs ; we
will show that these two sets are, in fact, equal.
Two other types of geometric sets play vital roles in the proof: cones
and disks. The cones are of two types: stable and unstable. The stable
cone (of “slope” α) is
C s (α) := v ∈ Rn kπu vk ≤ αkπs vk ,
theoryofodes July 4, 2007 13:20 Page 132
5. Invariant Manifolds
The first lemma shows that if the derivative of the map is applied to a
point in the unstable cone, the image is also in the unstable cone.
and
kπs DF (p)vk = kAss (p)vs + Asu (p)vu k ≤ kAss (p)vs k + kAsu (p)vu k
≤ kAss (p)kkvs k + kAsu (p)kkvu k ≤ µkvs k + εkvu k
≤ (µ/α + ε)kvu k.
so DF (p)v ∈ C u (α).
The main part of the second lemma is that moving unstable cones are
positively invariant. More precisely, if two points are in B(r ) and one
132
theoryofodes July 4, 2007 13:20 Page 133
theoryofodes July 4, 2007 13:20 Page 134
5. Invariant Manifolds
Similarly,
theoryofodes July 4, 2007 13:20 Page 135
for each j ∈ N.
g(x) = y, (5.7)
theoryofodes July 4, 2007 13:20 Page 136
5. Invariant Manifolds
so
kg(0)k = kπu F (ψ0 (0))k = kρ(ψ0 (0))k ≤ εkψ0 (0)k ≤ εr . (5.8)
1 1 + ε/α + 2ε
kG(x)k ≤ (εr + r + (ε + ε/α)r ) = r < r,
λ λ
so G(x) ∈ E u (r ).
That completes the verification that (5.7) has a solution for each
y ∈ E u (r ) and, therefore, that dom(ψ1 ) = E u (r ). To finish the proof,
we need to show that ψ1 is C 1 . Let g̃ be the restriction of g to g −1 (D1 ),
and observe that
ψ1 ◦ g̃ = πs ◦ F ◦ (I + ψ0 ). (5.9)
ψ1 = πs ◦ F ◦ (I + ψ0 ) ◦ g̃ −1
136
theoryofodes July 4, 2007 13:20 Page 137
Recall that Wrs was defined to be all points in the box B(r ) that pro-
duced forward orbits that remain confined within B(r ). The next
lemma shows that this set is a manifold.
D := {vs } + E u (r ).
s
Wloc (0) is a Lipschitz Manifold
s
Our next lemma shows that Wloc (0) = Wrs and that, in fact, orbits in
this set converge to 0 exponentially. (The constant µ̃ in the statement
of the theorem can be chosen to be µ + ε if α ≤ 1.)
s
In particular, Wrs = Wloc (0).
137
theoryofodes July 4, 2007 13:20 Page 138
5. Invariant Manifolds
s
Wloc (0) is C1
is C 1 , and Dh(0) = 0.
Proof. Let q ∈ Wrs be given. We will first come up with a candidate for
a plane that is tangent to Wrs at q, and then we will show that it really
works.
For each j ∈ N and each p ∈ Wrs , define
and let
C s,0 (p) := C s (α).
By definition (and by the invertibility of DF (v) for all v ∈ B(r )),
C s,j (p) is the image of the stable cone under an invertible linear trans-
formation. Note that
theoryofodes July 4, 2007 13:20 Page 139
and
C s,3 (p) = [D(F 3 )(p)]−1 C s (α) = [DF (F 2 (p))DF (F (p))DF (p)]−1 C s (α)
= [DF (p)]−1 [DF (F (p))]−1 [DF (F 2 (p))]−1 C s (α)
= [DF (p)]−1 [DF (F (p))]−1 C s,1 (F 2 (p))
⊂ [DF (p)]−1 [DF (F (p))]−1 C s (α) = C s,2 (p).
The plane that we will show is the tangent plane to Wrs at q is the
intersection
∞
\
C s,∞ (q) := C s,j (q)
j=0
theoryofodes July 4, 2007 13:20 Page 140
5. Invariant Manifolds
Since D(F j )(q)C s,j (q) = C s (α), it must, therefore, be the case that
(λ − ε/α)j kπu x − πu yk
≤ 2α.
(µ + εα)j kπs xk
(b) The estimate (5.11) shows that the size of the largest angle between
two vectors in C s,j (q) having the same projection onto E s goes
to zero as j ↑ ∞.
(c) Also, the estimates in the proof of the lemma on linear invariance
of the unstable cone show that the size of the minimal angle be-
tween a vector in C s,1 (F j (q)) and a vector outside of C s,0 (F j (q))
is bounded away from zero. Since
and
this means that the size of the minimal angle between a vector
in C s,j+1 (q) and a vector outside of C s,j (q) is also bounded away
from zero (for fixed j).
140
theoryofodes July 4, 2007 13:20 Page 141
From the continuity of F and the α-Lipschitz continuity of h, we know
that we can pick δ > 0 such that
theoryofodes July 4, 2007 13:20 Page 142
5. Invariant Manifolds
F −j−1 ({F j+1 (q)} + C s (α, η)) ⊆ {q} + C s,j (q). (5.16)
Higher Differentiability
Proof. We’ve already seen that this holds for k = 1. We show that it
is true for all k by induction. Let k ≥ 2, and assume that the lemma
works for k − 1. Define a new map H : Rn × Rn → Rn × Rn by the
formula " #! " #
p F (p)
H := .
v DF (p)v
142
theoryofodes July 4, 2007 13:20 Page 143
and, in general,
" #! " #
j p F j (p)
H = .
v D(F j )(p)v
Also,
" #! DF (p) 0
p
DH = ,
v
D 2 F (p)v DF (p)
so
" #! DF (0) 0
0
DH = ,
0
0 DF (0)
which is hyperbolic and invertible, since DF (0) is. Applying the induc-
tion hypothesis, we can conclude that the fixed point of H at the origin
(in Rn × Rn ) has a local stable manifold W that is C k−1 .
Fix q ∈ Wrs , and note that F j (q) → 0 as j ↑ ∞ and
n o
Pq = v ∈ Rn lim D(F j )(q)v = 0 .
j↑∞
Flows
Now we discuss how the Stable Manifold Theorem for maps implies the
Stable Manifold Theorem for flows. Given f : Ω → Rn satisfying f (0) =
0, let F = ϕ(1, ·), where ϕ is the flow generated by the differential
equation
ẋ = f (x). (5.17)
143
theoryofodes July 4, 2007 13:20 Page 144
5. Invariant Manifolds
d
g(t) = Df (0)g(t)
dt
and g(0) = I, so
g(t) = etDf (0) .
Thus, DF (0) is invertible, and if (5.17) has a hyperbolic equilibrium at
the origin then DF (0) is hyperbolic.
Since F satisfies the hypotheses of the Stable Manifold Theorem for
maps, we know that F has a local stable manifold Wrs on some box
B(r ). Assume that α < 1 and that r is small enough that the vector
field of (5.17) points into B(r ) on C s (α) ∩ ∂B(r ). (See the estimates
in Section 3.4.) The requirements for a point to be in Wrs are no more
restrictive then the requirements to be in the local stable manifold Wrs
of the origin with respect to the flow, so Wrs ⊆ Wrs .
We claim that, in fact, these two sets are equal. Suppose they are
not. Then there is a point q ∈ Wrs \Wrs . Let x(t) be the solution of (5.17)
satisfying x(0) = q. Since limj↑∞ F j (q) = 0 and, in a neighborhood of
the origin, there is a bound on the factor by which x(t) can grow in 1
unit of time, we know that x(t) → 0 as t ↑ ∞. Among other things, this
implies that
theoryofodes July 4, 2007 13:20 Page 145
Center Manifolds
Since Wrs is a closed set and x is continuous, (a) and (b) say that we can
pick t0 to be the earliest time such that x(t) ∈ Wrs for every t ≥ t0 .
Now, consider the location of x(t) for t in the interval [t0 − 1, t0 ).
Since x(0) ∈ Wrs , we know that x(j) ∈ Wrs for every j ∈ N. In particu-
lar, we can choose t1 ∈ [t0 − 1, t0 ) such that x(t1 ) ∈ Wrs . By definition
of t0 , we can choose t2 ∈ (t1 , t0 ) such that x(t2 ) ∉ Wrs . By the continu-
ity of x and the closedness of Wrs , we can pick t3 to the be the last time
before t2 such that x(t3 ) ∈ Wrs . By definition of Wrs , if t ∈ [t0 − 1, t0 )
and x(t) ∉ Wrs , then x(t) ∉ B(r ); hence, x(t) must leave B(r ) at time
t = t3 . But this contradicts the fact that the vector field points into
B(r ) at x(t3 ), since x(t3 ) ∈ C s (α) ∩ ∂B(r ). This contradiction implies
that no point q ∈ Wrs \ Wrs exists; i.e., Wrs = Wrs .
The exponential decay of solutions of the flow on the local stable
manifold is a consequence of the similar decay estimate for the map,
along with the observation that, near 0, there is a bound to the factor
by which a solution can grown in 1 unit of time.
Definition
ẋ = Ax (5.18)
and
\n o
Ec = x ∈ Rn lim |ect etA x| = lim |e−ct etA x| = 0 .
t↓−∞ t↑∞
c>0
The Stable Manifold Theorem tells us that for the nonlinear differential
equation
ẋ = f (x), (5.19)
145
theoryofodes July 4, 2007 13:20 Page 146
5. Invariant Manifolds
with f (0) = 0, the stable manifold W s (0) and the unstable manifold
W u (0) have characterizations similar to E s and E u , respectively:
[n o
W s (0) = x ∈ Rn lim |ect ϕ(t, x)| = 0 ,
t↑∞
c>0
and
[n o
W u (0) = x ∈ Rn lim |e−ct ϕ(t, x)| = 0 ,
t↓−∞
c>0
where ϕ is the flow generated by (5.19). (This was only verified when
the equilibrium point at the origin was hyperbolic, but a similar result
holds in general.)
Is there a useful way to modify the characterization of E c similarly
to get a characterization of a center manifold W c (0)? Not really. The
main problem is that the characterizations of E s and E u only depend
on the local behavior of solutions when they are near the origin, but
the characterization of E c depends on the behavior of solutions that
are, possibly, far from 0.
Still, the idea of a center manifold as some sort of nonlinear ana-
logue of E c (0) is useful. Here’s one widely-used definition:
Nonuniqueness
While the fact that stable and unstable manifolds are really manifolds
is a theorem (namely, the Stable Manifold Theorem), a center manifold
is a manifold by definition. Also, note that we refer to the stable mani-
fold and the unstable manifold, but we refer to a center manifold. This
is because center manifolds are not necessarily unique. An extremely
simple example of nonuniqueness (commonly credited to Kelley) is the
planar system
ẋ = x 2
ẏ = −y.
146
theoryofodes July 4, 2007 13:20 Page 147
Center Manifolds
Clearly, E c is the x-axis, and solving the system explicitly reveals that
for any constant c ∈ R the curve
(x, y) ∈ R2 x < 0 and y = ce1/x ∪ (x, 0) ∈ R2 x ≥ 0
is a center manifold.
Existence
There is a Center Manifold Theorem just like there was a Stable Man-
ifold Theorem. However, the goal of the Center Manifold Theorem is
not to characterize a center manifold; that is done by the definition.
The Center Manifold Theorem asserts the existence of a center mani-
fold.
We will not state this theorem precisely nor prove it, but we can give
some indication how the proof of existence of a center manifold might
go. Suppose that none of the eigenvalues of Df (0) have real part equal
to α, where α is a given real number. Then we can split the eigenval-
ues up into two sets: Those with real part less than α and those with
real part greater than α. Let E − be the vector space spanned by the
theoryofodes July 4, 2007 13:20 Page 148
5. Invariant Manifolds
is a center manifold.
the function h whose graph is the center manifold is a solution of the
equation Mh = 0.
Approximation
theoryofodes July 4, 2007 13:20 Page 149
(Mφ)(x) = O(|x|q )
as x → 0.
Stability
asymptotically stable, or unstable) matches the stability type of the
origin as an equilibrium solution of (5.22).
These results of Carr are sometimes useful in computing the stabil-
ity type of the origin. Consider, for example, the following system:
ẋ = xy + ax 3 + by 2 x
ẏ = −y + cx 2 + dx 2 y,
where x and y are real variables and a, b, c, and d are real parameters.
We know that there is a center manifold, tangent to the x-axis at the
origin, that is (locally) of the form y = h(x). The reduced equation on
the center manifold is
theoryofodes July 4, 2007 13:20 Page 150
5. Invariant Manifolds
ẋ = (a + c)x 3 + O(x 5 ).
ẋ = (cd + bc 2 )x 5 + O(x 7 ),
ẋ = −b 2 c 3 x 7 + O(x 9 ).
Bifurcation Theory
theoryofodes July 4, 2007 13:20 Page 151
with the eigenvalues of A lying on the imaginary axis and the eigen-
values of B lying in the open right half-plane. Since the parameter ε
does not depend on time, we can append the equation ε̇ = 0 to get the
expanded system
u̇ = Au + f (u, v, ε)
v̇ = Bv + g(u, v, ε) (5.25)
ε̇ = 0.
151
theoryofodes July 4, 2007 13:20 Page 152
theoryofodes July 4, 2007 13:20 Page 153
6
Periodic Orbits
Theorem (Poincaré-Bendixson) Every nonempty, compact ω-limit set of
a C 1 planar flow that does not contain an equilibrium point is a nonde-
generate periodic orbit.
theoryofodes July 4, 2007 13:20 Page 154
6. Periodic Orbits
and p3 is between p1 and p2 . Note that the union of the line segment
p1 p2 from p1 to p2 with the curve ϕ([t1 , t2 ], p) is a simple closed
curve in the plane, so by the Jordan Curve Theorem it has an “inside” I
and an “outside” O. Assuming, without loss of generality, that f points
into I all along the “interior” of p1 p2 , we get a picture something like:
p1 b
bp2
Note that
I ∪ p1 p2 ∪ ϕ([t1 , t2 ], p)
theoryofodes July 4, 2007 13:20 Page 155
Poincaré-Bendixson Theorem
The proof of the next lemma uses something called a flow box. A
flow box is a (topological) box such that f points into the box along
one side, points out of the box along the opposite side, and is tangent
to the other two sides, and the restriction of ϕ to the box is conjugate
to unidirectional, constant-velocity flow. The existence of a flow box
around any regular point of ϕ is a consequence of the C r -rectification
Theorem.
Proof. Suppose that for some point x and some transversal S, ω(x)
intersects S at two distinct points p1 and p2 . Since p1 and p2 are
on a transversal, they are regular points, so we can choose disjoint
subintervals S1 and S2 of S containing, respectively, p1 and p2 , and,
for some ε > 0, define flow boxes B1 and B2 by
Now, the fact that p1 , p2 ∈ ω(x) means that we can pick an in-
creasing sequence of times t1 , t2 , . . . such that ϕ(tj , x) ∈ B1 if j is
odd and ϕ(tj , x) ∈ B2 if j is even. In fact, because of the nature of
the flow in B1 and B2 , we can assume that ϕ(tj , x) ∈ S for each j.
Although the sequence ϕ(t1 , x), ϕ(t2 , x), . . . is monotone on the tra-
jectory T := γ(x), it is not monotone on S, contradicting the previous
lemma.
theoryofodes July 4, 2007 13:20 Page 156
6. Periodic Orbits
Proof. Fix q ∈ P. Pick T > 0 such that ϕ(T , q) = q. Let ε > 0 be given.
By continuous dependence, we can pick δ > 0 such that |ϕ(t, y) −
ϕ(t, q)| < ε whenever t ∈ [0, 3T /2] and |y − q| < δ. Pick a transversal
S of length less than δ with q in its “interior”, and create a flow box
B := ϕ(t, u) u ∈ S, t ∈ [−ρ, ρ]
for some ρ ∈ (0, T /2]. By continuity of ϕ(T , ·), we know that we can
pick a subinterval S′ of S that contains q and that satisfies ϕ(T , S′ ) ⊂
t ≥ 0 ϕ(t, x) ∈ S′ .
Now, let t ≥ t1 be given. Then t ∈ [tj , tj+1 ) for some j ≥ 1. For this
j,
since, by (6.1), |t−tj | < |tj+1 −tj | < 3T /2 and since, because ϕ(tj , x) ∈
S′ ⊆ S, |q − ϕ(tj , x)| < δ.
156
theoryofodes July 4, 2007 13:20 Page 157
Lienard’s Equation
iC
b b
L R
iL iR
iL = iR = −iC , (6.2)
157
theoryofodes July 4, 2007 13:20 Page 158
6. Periodic Orbits
dVC
C = iC , (6.4)
dt
and by Faraday’s Law
diL
L
= VL , (6.5)
dt
where L is the inductance of the inductor. We assume that the resistor
behaves nonlinearly and satisfies the generalized form of Ohm’s Law:
VR = F (iR ). (6.6)
1
ẋ = VL ,
L
so by (6.3), (6.4), (6.6), and (6.2)
1 dVL 1 1 1 diR
ẍ = = (V̇C − V̇R ) = iC − F ′ (iR )
L dt L L C dt
1 1
= (−x) − f (x)ẋ
L C
Hence,
1 1
ẍ + f (x)ẋ + x = 0.
L LC
By rescaling f and t (or, equivalently, by choosing units judiciously),
we get Lienard’s Equation:
ẍ + f (x)ẋ + x = 0.
(i) F (0) = 0;
(iii) F is odd;
158
theoryofodes July 4, 2007 13:20 Page 159
Lienard’s Equation
(iv) F (x) → ∞ as x ↑ ∞;
(v) for some β > 0, F (β) = 0 and F is positive and increasing on (β, ∞);
periodic orbit, and it is the ω-limit set of all points other than the origin.
d
(x 2 + y 2 ) = 2(x ẋ + y ẏ) = −2xF (x),
dt
theoryofodes July 4, 2007 13:20 Page 160
6. Periodic Orbits
D A
C B
and let
S5 := (x, y) ∈ R2 x 2 + y 2 = ε2 ,
160
theoryofodes July 4, 2007 13:20 Page 161
Lienard’s Theorem
for some small ε. Then it is not hard to see that ∪5i=1 Si is the bound-
ary of a compact, positively invariant region that does not contain an
equilibrium point.
y0
−ỹ0
ỹ0
−y0
theoryofodes July 4, 2007 13:20 Page 162
6. Periodic Orbits
Zd Zd
Ṙ −x(y)F (x(y))
R(c, d) − R(a, b) = dy = dy
b ẏ b −x(y)
Zd
= F (x(y)) dy. (6.9)
b
We will chop the orbit segment connecting (0, y0 ) to (0, ỹ0 ) up into
pieces and use (6.8) and (6.9) to estimate the change ∆R in R along
each piece and, therefore, along the whole orbit segment.
Let σ = β + 1, and let
B = sup |F (x)|.
0≤x≤σ
R := (x, y) ∈ R2 x ∈ [0, σ ], y ∈ [B + σ , ∞) .
In R,
dy x σ
dx = y − F (x) ≤ σ = 1;
theoryofodes July 4, 2007 13:20 Page 163
Lienard’s Theorem
(0, y0 )
(σ , yσ )
(σ , ỹσ )
(0, ỹ0 )
σ 2B
= →0
y0 − B − σ
as y0 → ∞. A similar estimate shows that |R(0, ỹ0 ) − R(σ , ỹσ )| → 0 as
y0 → ∞.
On the middle segment, the x-coordinate is a function x(y) of the
y-coordinate y. Hence,
Z ỹσ
R(σ , ỹσ ) − R(σ , yσ ) = F (x(y)) dy ≤ −|yσ − ỹσ |F (σ ) → −∞
yσ
as y0 → ∞.
Putting these three estimates together, we see that
theoryofodes July 4, 2007 13:20 Page 164
6. Periodic Orbits
y-axis at a point (0, ỹ0 ). Such orbit segments are “nested” and fill
up the open right half-plane. We need to show that only one of them
satisfies ỹ0 = −y0 . In other words, we claim that there is only one
segment that gives
Now, if such a segment hits the x-axis on [0, β], then x ≤ β all along
that segment, and F (x) ≤ 0 with equality only if (x, y) = (β, 0). Let
x(y) be the x-coordinate as a function of y and observe that
Z ỹ0
R(0, ỹ0 ) − R(0, y0 ) = F (x(y)) dy > 0. (6.10)
y0
theoryofodes July 4, 2007 13:20 Page 165
Lienard’s Theorem
(0, Y0 ) b
b b (β, Yβ )
(0, y0 )
b b
(λ, yβ )
(β, yβ )
(β, ỹβ )
b b
(µ, ỹβ )
b
(0, ỹ0 )
b
b (β, Ỹβ )
(0, Ỹ0 )
Note that
Let X(Y ) and x(y) give, respectively, the first coordinate of a point
on the outer and inner orbit segments as a function of the second co-
ordinate. Similarly, let Y (X) and y(x) give the second coordinates as
functions of the first coordinates (on the segments where that’s possi-
ble). Estimating, we find that
Z0 Z0
−XF (X) −xF (x)
∆1 = dX < dx = R(0, ỹ0 ) − R(β, ỹβ ),
β Y (X) − F (X) β y(x) − F (x)
(6.12)
Z Ỹβ
∆2 = F (X(Y )) dY < 0, (6.13)
ỹβ
Z ỹβ Z ỹβ
∆3 = F (X(Y )) dY < F (x(y)) dy = R(β, ỹβ ) − R(β, yβ ), (6.14)
yβ yβ
165
theoryofodes July 4, 2007 13:20 Page 166
6. Periodic Orbits
Z yβ
∆4 = F (X(Y )) dY < 0, (6.15)
Yβ
and
Zβ Zβ
−XF (X) −xF (x)
∆5 = dX < dx = R(β, yβ ) − R(0, y0 ).
0 Y (X) − F (X) 0 y(x) − F (x)
(6.16)
By plugging, (6.12), (6.13), (6.14), (6.15), and (6.16) into (6.11), we see
that
166