Theory of ODEs 2023
Theory of ODEs 2023
James C. Robinson
March 9, 2023
Contents
2 Linear systems 32
i
3.2 One-dimensional dynamics . . . . . . . . . . . . . . . . . . . . . . . 51
3.5 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
ii
Preliminaries
1
Open, closed, and compact sets; the matrix norm
(MVC review)
Open sets
B(x, ε) ⊂ A,
Closed sets
Lemma 0.1. A set A is closed if and only if for every sequence (xn ) ∈ A with
xn → x, x ∈ A.
Compact sets
[You will see in NMT that there is a definition of compactness in terms of open
sets that is more general. It is equivalent to sequential continuity in any metric
space, in particular in Rn with its usual distance.]
2
Theorem 0.3 (Sequential Heine–Borel Theorem). The interval [0, 1] is (sequen-
tially) compact.
Proof. Take any sequence (xn ) ∈ [0, 1]; then (xn ) is a bounded sequence of real
numbers, so by the Bolzano–Weierstrass Theorem has a convergent subsequence,
xnj → x. Since 0 ≤ xnj ≤ 1, 0 ≤ x ≤ 1, i.e. x ∈ [0, 1].
[The proof in NMT using the other definition of compactness is much more painful.]
3
For the other implication, if K is compact and (xn ) ∈ K with xn → x, then as
(xn ) ∈ K and K is compact there is a subsequence xnj → y ∈ K; but xnj → x (it is
a subsequence of xn , which converges to x), so by uniqueness of limits x = y ∈ K,
i.e. K is closed. Suppose now that K is unbounded: then there exist xn ∈ K such
that |xn | ≥ n; and this sequence has no convergent subsequence.
Continuity between Euclidean spaces (or, in fact, normed spaces) looks just
like continuity in R. If f : Rn → Rm then f is continuous at x ∈ X if for every
ε > 0 there exists δ > 0 such that
(Note that the first | · | is the Euclidean norm in Rn , while the second is the
Euclidean norm in Rm .)
We can now generalise the Extreme Value Theorem from Analysis II.
Proof. Take a sequence (yn ) ∈ f (A). Then yn = f (xn ) for some sequence (xn ) ∈ A.
Since A is compact, it has a convergent subsequence xnj → x∗ ∈ A. Since f is
continuous,
f (xnj ) → f (x∗ ) =: y ∗ ∈ f (A),
so f (A) is compact.
4
Continuity in terms of open sets
The notation can be a bit confusing, this is not ‘the inverse of f applied to B’,
since the preimage as defined in (2) makes sense even when f is not invertible. So
in general, for f : X → Y we have only
f −1 (f ([0, 1])) = f −1 ([0, 1]) = [−1, 1] and f (f −1 ([−1, 1])) = f ([0, 1]) = [0, 1].
The following result shows (in particular) that we could define continuity in
terms of open sets, but in this module we will mostly use the fact that if f is
continuous (understood as in (1)) then the preimage of any open set is open (and
the preimage of any closed set is closed).
But this means that for any y ∈ B(x, δ), i.e. any y with |y − x| < δ, we must have
f (y) ∈ B(f (x), ε), i.e. |f (y) − f (x)| < ε; so f is continuous.
5
The matrix norm
We write
a11 · · · a1n
.. ..
A= . .
an1 · · · ann
as (aij )ni,j=1 . To every n × n matrix we can associate a linear map Rn → Rn , given
by x 7→ Ax; recall that
n
X
(Ax)i = aij xj .
j=1
which can be an easier definition to apply. The point is that it follows immediately
from the definition that |Ax| ≤ ∥A∥|x| for every x ∈ Rn .
It is possible to relate ∥A∥ to the entries of A. Note that for any x ∈ Rn with
|x| = 1 we have
n n
!2
X X
|Ax|2 = aij xj
i=1 j=1
n
" n
! n
!# n
X X X X
≤ a2ij 2
|xj | = |aij |2 ,
i=1 j=1 j=1 i,j=1
using the Cauchy–Schwarz inequality to move from the first to the second line. It
follows that !1/2
X n
∥A∥ ≤ |aij |2 .
i,j=1
6
However, if we let {e1 , . . . , en } be the standard basis of Rn , then we have
ai1 X n
· · · = Aei ⇒ |aij |2 = |Aei |2 ,
ain j=1
so that n n n
X X X
2 2
|aij | = |Aei | ≤ ∥A∥2 |ei |2 = n∥A∥2 .
i,j=1 i=1 i=1
It follows that
n
!1/2 n
!1/2
1 X X
√ |aij |2 ≤ ∥A∥ ≤ |aij |2 . (3)
n i,j=1 i,j=1
√
In particular, |aij | ≤ n∥A∥ for every i, j.
7
The Contraction Mapping Theorem (NMT pre-
view)
Norms
‘Trivial’ examples:
1. (R, | · |)
p
2. (Rn , | · |), where |(x1 , . . . , xn )| = x21 + · · · + x2n (we use the same notation for
the usual norm on Rn as for the absolute value in R).
3. Let V = Rn×n , the space of all n × n matrices. If A ∈ Rn×n , then the ‘matrix
norm’ of A is
|Ax|
∥A∥ = sup = sup |Ax|.
x∈Rn , x̸=0 |x| x∈Rn , |x|=1
4a. Let V = C([a, b]; R), continuous function from [a, b] into R; then
∥f ∥∞ := sup |f (t)|
t∈[a,b]
8
is a norm on V .
4b. Let
V = C([a, b]; Rn )
= {f = (f1 , . . . , fn ) : [a, b] → Rn : fi ∈ C([a, b]; R), i = 1, . . . , n},
i.e. continuous functions from [a, b] into Rn . Then
∥f ∥∞ := sup |f (t)|
t∈[a,b]
is a norm on V .
By copying almost word-for-word the proofs from Analysis I you can prove that
this kind of convergence behaves in the way you would expect, e.g. if vn → v and
un → u then vn + un → v + u (you have to use the fact that X is a vector space
to make sure that ‘adding’ makes sense).
Recall (from Analysis I) that a sequence (xn ) of real numbers converges if and
only if it is a Cauchy sequence: for every ε > 0 there exists N such that
n, m ≥ N ⇒ |xn − xm | < ε.
The point of this is it gives a test for convergence without knowing what the limit
is.
You may remember that showing that any Cauchy sequence converges is not
trivial, even when this is ‘only’ a sequence of real numbers, so it should be no
surprise that ‘being complete’ is quite a special property for a normed space.
Thankfully, many of the spaces that are often used are complete (actually, this is
probably why they are often used). In particular, we can show this for two of the
examples above.
9
Proposition 0.9. The space X = (C([a, b]; R), ∥ · ∥∞) is complete.
Proof. Suppose that (fn ) is a Cauchy sequence in X: then for every ε > 0 there
exists N such that
n, m ≥ N ⇒ ∥fn − fm ∥∞ < ε.
This means that for each fixed t, (fn (t)) is a Cauchy sequence of real numbers,
so converges to some limit: we define
Now we show that in fact fn → f uniformly on [a, b]. But we can do this using
(5): if we take m → ∞ then we obtain
Proof. Given a sequence (fk ) ∈ C([a, b]; Rn ) we have fn = (fn1 , fn2 , . . . , fkn ), where
each (fkj )k is a sequence in C([a, b]; R), so we can apply Proposition 0.9 to each of
these in turn.
10
The Contraction Mapping Theorem
T (ū) = ū,
where
T n (u) := T
| ◦ T ◦{z· · · ◦ T}(u).
n times
Proof. Pick any u0 ∈ U and set un = T n (u0 ). Now, for any n ≥ 1 we have
κn
≤ ∥u0 − u1 ∥. (7)
1−κ
11
It follows that (un ) is a Cauchy sequence in (V, ∥ · ∥); since this space is complete
we must have un → ū for some ū ∈ V . Since un ∈ U and U is closed, we must
have ū ∈ U .
To show that ū is unique, suppose that we also have T (v̄) = v̄. Then
∥ū − v̄∥ = ∥T (ū) − T (v̄)∥ ≤ κ∥ū − v̄∥,
which is impossible unless ∥ū − v̄∥ = 0, i.e. ū = v̄.
To turn this into the version we need as a ‘black box’ for Picard’s Theorem in
the module, note that we have already shown that V = C([a, b]; Rn ) is complete.
Now observe that if Y is a closed subset of Rn , then U = C([a, b]; Y ) is a closed
subset of C([a, b]; Rn ): given a sequence (fn ) ∈ C([a, b]; Y ) such that fn → f (in
the sup norm), this means that in particular for every t ∈ [a, b] we have
fn (t) → f (t);
so (fn (t)) is a convergent sequence every element of which lies in Y , and since Y
is closed this implies that f (t) ∈ Y . So f (t) ∈ Y for every t ∈ [a, b], so we have
f ∈ C([a, b]; Y ), which shows that C([a, b], Y ) is closed.
Theorem 0.12 (Contraction Mapping Theorem in spaces of continuous func-
tions). Let X = C([a, b]; Y ), where Y is a closed subset of Rn . Suppose that
P : X → X is a map such that
∥P (x) − P (y)∥∞ ≤ κ∥x − y∥∞ , x, y ∈ X,
for some 0 ≤ κ < 1, where ∥x∥∞ := supt∈[a,b] |x(t)|. Then there exists a unique
x∗ ∈ X such that
P (x∗ ) = x∗ ,
i.e. x∗ is a fixed point of P .
12
Part I
13
Chapter 1
ẋ = f (x), x ∈ U ⊂ Rn .
[Here ẋ denotes dx/dt; implicit is that x is x(t), a function of t. You could (and
many do) rewrite this equation as ẋ(t) = f (x(t)), but the extra ts just make it
look more complicated.]
However, we will start with a simple existence and uniqueness result for the
non-autonomous problem
14
Lemma 1.1. If f : Rn × R → Rn is continuous and τ > 0, then the following
statements are equivalent:
(ii) ⇒ (i) Since s 7→ f (x(s), s) is continuous, we can use the Fundamental Theorem
of Calculus to deduce that
ẋ(t) = f (x(t)),
and we have x(t0 ) = x0 as required.
In order to find solutions of the integral equation (1.2), we will use the following
version of the Contraction Mapping Theorem – the full version is proved in Norms,
Metrics, and Topologies, and also in the preliminary material.
Theorem 1.2 (Contraction Mapping Theorem in spaces of continuous functions).
Let X = C([a, b]; Y ), where Y is a closed subset of Rn . Suppose that P : X → X
is a map such that
∥P (x) − P (y)∥∞ ≤ κ∥x − y∥∞ , x, y ∈ X,
for some 0 ≤ κ < 1, where
∥x∥∞ := sup |x(t)|.
t∈[a,b]
15
Note that the whole of Rn is a closed set, so X = C([a, b]; Rn ) is allowed.
We will also need the following result, which you know for functions of a scalar
variable from Analysis III, but not for vector-valued integrals.
Lemma 1.3. If f : [a, b] → Rn is Riemann integrable then
ˆ b ˆ b
f (t) dt ≤ |f (t)| dt.
a a
´b
Proof. Let γ = f (s) ds; then
a
ˆ b ˆ b
2
|γ| = γ · γ = γ · f (s) ds = γ · f (s) ds
a a
ˆ b ˆ b ˆ b
≤ |γ · f (s)| ds ≤ |γ||f (s)| ds = |γ| |f (s)| ds.
a a a
We first give a proof of the existence and uniqueness of solutions to the non-
autonomous equation ẋ = f (x, t) when f is globally Lipschitz, i.e. when (1.3) is
satisfied.
Theorem 1.4 (Picard’s Theorem for globally Lipschitz f ). Suppose that the func-
tion f : Rn × R → Rn is continuous and that there exists L > 0 such that
Proof. Lemma 1.1 shows that finding a solution of the ODE on [t0 − τ, t0 + τ ] is
equivalent to finding a continuous function x : [t0 , t0 + τ ] that satisfies
ˆ t
x(t) = x0 + f (x(s), s) ds,
t0
16
i.e. to finding a fixed point of the operator P : X 7→ X, where
X = C 0 ([t0 − τ, t0 + τ ], Rn )
and ˆ t
[P (x)](t) := x0 + f (x(s), s) ds. (1.4)
t0
1. Being ‘Lipschitz’ is just a little weaker than being differentiable. The standard
example showing that a Lipschitz function need not be differentiable is f (x) = |x|,
which is not differentiable at zero. (In fact a Lipschitz function is differentiable
‘almost everywhere’.)
17
for some c ∈ (x, y). So if |f ′ (c)| ≤ M for all M ∈ [a, b] we have
2. As stated, the theorem only guarantees that a solution exists locally, i.e. for
times close to t0 (t ∈ [t0 − τ, t0 + τ ]). However, when f is globally Lipschitz (as it
is here) we can we can use the theorem repeatedly to extend the interval on which
the solution exists using the following simple observation.
on [t1 , t2 ]. Then (
x1 (t) t0 ≤ t ≤ t1
x(t) =
x2 (t) t1 < t ≤ t2
solves (1.5) on [0, t1 + t2 ] (note that the left derivative of x1 matches the right
derivative of x2 at 0, so x is still C 1 ). So we can try to keep extending the interval
on which the solution is defined.
We start with y(0) = y0 ; then the solution exists for all t ∈ [0, τ ]. We now solve
the equation starting at x(τ ) = y(τ ); the solution here exists for t ∈ [τ, 2τ ]; again
we apply the theorem with x(2τ ) = y(2τ ) and find a solution valid for t ∈ [2τ, 3τ ].
The gluing observation above means that if we continue in this way we end up
with a solution that exists for all t ≥ 0.
18
3. We now apply this result to the linear equation
the right-hand side is a globally Lipschitz function of x, and so we can apply the
global Picard Theorem to deduce that (1.6) has a unique solution t 7→ x(t) defined
for t ∈ [−τ, τ ] provided we choose τ < ∥A∥−1 . (In fact, we will soon see that when
f is globally Lipschitz the solution exists for all t ∈ R.) We will now deduce the
form of this solution on [−τ, τ ] using the Contracting Mapping argument.
Recall from Theorem 1.4 that the successive iterations of the Picard map P
from (1.9) converge to the solution. If we start with x0 (t) := x0 (constant) then
we obtain
ˆ t
1
x (t) = x0 + Ax0 (s) ds = x0 + tAx0
ˆ0 t ˆ t
2 1
x (t) = x0 + Ax (s) ds = x0 + A[x0 + sAx0 ] ds
0 0
t2
= x0 + tAx0 + A2 x0 . . .
2
By induction we obtain
k
k
X tj Aj
x (t) = x0 ,
j=0
j!
that the sum converges for |t| ≤ τ is guaranteed by the way we used the Contraction
Mapping Theorem (Theorem 1.2) in the proof of Theorem 1.4, although we will
see later (more directly) that the sum in fact converges for all t ∈ R (and gives
the solution for all t in this range).
19
1.3 Picard’s Theorem: locally Lipschitz case
|f (u) − f (v)| ≤ LK |u − v| u, v ∈ K.
Then for some τ = τ (x0 ) > 0 there exists a unique x : [t0 − τ, t0 + τ ] → U that
solves
ẋ = f (x), x(t0 ) = x0 (1.8)
for all t ∈ [t0 − τ, t0 + τ ].
We will soon show that in fact we can take τ (x0 ) to be uniform for all x0 ∈ K,
when K is any compact subset of U .
Proof. From our initial observations, we can restrict to the (notationally more
convenient) case t0 = 0.
K := B(x0 , r) ⊂ U,
20
where B(x0 , r) = {y ∈ Rn : |y − x0 | ≤ r} (this is possible since U is open). The
set K is compact; let L = LK and let
M = sup |f (x)|,
x∈K
which is finite since x 7→ |f (x)| is a continuous map from the compact set K into R
(so we can use the Extreme Value Theorem). [We do this to fix a set that contains
x0 on which f is bounded and Lipschitz.]
maps X into itself and is a contraction: it will then follow from the Contraction
Mapping Theorem (Theorem 1.2) that P has a unique fixed point, which will be
the solution of the ODE (1.1) by Lemma 1.1.
Take x ∈ X. Then
ˆ t
P (x)(t) = x0 + f (x(s)) ds
0
so P : W → W as claimed.
21
Given x, y ∈ W , for each t ∈ [−τ, τ ] we have
ˆ t
|P (x)(t) − P (y)(t)| = f (x(s)) − f (y(s)) ds
0
ˆ t
≤ |f (x(s)) − f (y(s))| ds
0
ˆ t
≤L |x(s) − y(s)| ds
0
≤ L|t| max |x(s) − y(s)|
s∈[0,τ ]
≤ Lτ ∥x − y∥∞ ;
so
∥P (x) − P (y)∥∞ ≤ Lτ ∥x − y∥∞ ,
and by assumption Lτ < 1. Thus P is a contraction, and it has a unique fixed
point x ∈ W , which is the solution we want.
1. If the assumptions of the theorem do not hold then solutions need not be unique.
For example, consider the equation on [0, ∞) given by
ẋ = |x|1/2 , x(0) = 0.
This has an ‘obvious’ solution x(t) = 0 for all t ≥ 0, but we can also solve by
separating variables, noting that for any solution we will have x(t) ≥ 0:
ˆ x(t) ˆ t
dx x(t)
2x1/2
= dt ⇒ 0
= t,
0 x1/2 0
which gives
t2
2x(t)1/2 = t ⇒ x(t) = .
4
is a solution.
22
This does not contradict the theorem because |x|1/2 is not locally Lipschitz
near zero: for any L > 0, if we take 0 < x < L−1/2 then
||x|1/2 − 01/2 | = x1/2 > L|x − 0| = Lx,
so no local Lipschitz constant will work near x = 0.
2. However, if f is locally Lipschitz, then the uniqueness result that is part of this
theorem means that for an autonomous equation, distinct solutions cannot cross.
3. Once again, the theorem only guarantees the existence of a solution ‘locally’,
i.e. near t = 0. And if f is only locally Lipschitz, then solutions can blow up in
a finite time. When we try to keep extending the existence time by applying the
theorem repeatedly we will end up with a solution only a finite time interval.
To see how this is consistent with repeatedly using the ‘short-time existence’
we get from Theorem 1.5, we follow the proof to see what value of τ we are allowed
to take, depending on x0 . We will take x0 > 0, since this is the situation in which
the solution blows up for positive t.
In the proof we have to choose r so that B(x0 , r) ⊂ U = R. This holds for any
r > 0, so we are (at this stage) free to choose any r. Note that, since we are in R,
B(x0 , r) = [x0 − r, x0 + r].
23
The first constraint on the local existence time τ in the proof is to take τ M ≤ r.
This requires
r
τ≤ ,
(x0 + r)2
and this is maximised when we take r = x0 , and gives τ = 1/4x0 .
The second condition that restricts τ is to take τ L < 1, where L is the Lipschitz
constant for f (x) = x2 on [x0 − r, x0 + r] = [0, 2x0 ] (give our choice of r). The
easiest way to find the Lipschitz constant of f is the use the Mean Value Theorem:
we know that there exists c ∈ (0, 2x0 ) such that
f (x) − f (y)
≤ f ′ (c),
x−y
and if we take the modulus of both sides we obtain
For f (x) = x2 we have f ′ (x) = 2x, and so whatever c ∈ (0, 2x0 ) we choose, we end
up with |f ′ (c)| < 4x0 , and so
As the solution increases, the existence times we get from our result decreases;
we can keep going up to a maximum time
1 1 1 1 1 1 1
+ + 2
+ ··· = 4 = ,
5x0 5(5/4)x0 5(5/4) x0 x0 5 1 − 5
x0
4. We used the Mean Value Theorem in R in the previous point. We can do some-
thing similar in higher dimensions. Suppose that S ⊂ Rn and that the Jacobian
24
matrix Df (x) exists for x ∈ S and x 7→ Df (x) is continuous on S. Then for each
convex2 compact subset C ⊂ S there exists an LC > 0 such that
5. As promised above, we now show that the existence time is uniform for all x0
in any compact subset K of U .
To prove this we will need the following lemma. Remember that for subsets of
n
R compact ≡ closed and bounded. The notion of ‘an open set containing A’ will
occur repeatedly: we will call such a set a neighbourhood of A.
Aε := {y ∈ Rn : dist(y, A) < ε} ⊂ V.
Note that if A = {a} is a single point, this reduces to the (trivial) result that
any open set V containing a contains B(a, ε) for some ε > 0 (‘trivial’ since it
follows immediately from the fact that a ∈ V and V is open).
Proof. Suppose that this is not the case. Then for every n ∈ N there exists a point
yn with dist(yn , A) < 1/n but yn ∈
/ V . Since dist(yn , A) < 1/n, there exists an ∈ A
such that
|an − yn | = dist(yn , A) < 1/n.
Since A is compact, the sequence (an ) has a subsequence anj that converges to
some a ∈ A.
2 A set C is convex if u, v ∈ C and λ ∈ [0, 1] implies that λu + (1 − λ)v ∈ C (the line segment
joining u and v lies in C).
3 If A is only closed (and not bounded), then the result of this lemma need not be true: consider
the sets A and V in R2 given by
2
A = {(x, 0) : x ∈ R} V = {(x, y) : |y| < e−x }.
Given any ε > 0, we can take x large enough that (x, ε/2) ∈
/ V [just take x such that
−x2
e < ε/2], so we do not have Aε ⊂ V for any ε > 0.
25
Figure 1.1: Left: a neighbourhood of a compact set A contains an ε-neighbourhood
of A for some ε > 0. Right: in the simplest case when A = {a}, any neighbourhood
contains B(a, ε) for some ε > 0.
Proof. We use Lemma 1.6 to guarantee there exists r > 0 such that
Zr := {y ∈ Rn : |y − z| ≤ r, for some z ∈ Z} ⊂ U
(the lemma gives this inclusion with |y − z| < ε for some ε > 0; now take any r
with 0 < r < ε). It follows that B(x0 , r) ⊂ U for every x0 ∈ Z.
Since Zr is compact, as in the main proof we can let M := supx∈Zr |f (x)| and
set L = LZr . The proof now follows exactly the same lines, showing that τ can be
taken uniformly for all x0 ∈ Z.
26
We will use this uniform existence time later to investigate what happens if a
solution cannot be extended to exist for all t ≥ 0.
6. In fact, we often treat a (simple) situation in which Theorem 1.5 is not directly
applicable. If we have a planar system and write it in polar coordinates then the
equations (for the radial direction) will be of the form
ṙ = f (r), r(0) = r0 ;
clearly in this case r ∈ [0, ∞), so f : [0, ∞) → R, and since [0, ∞) is not open we
cannot apply Theorem 1.5 directly. We will now analyse a differential inequality
that will enable us to deal with this case (on Examples 2).
We now investigate how solutions depend on the ‘data’, i.e. the initial condition
x0 and the ‘model’ f . We will need the following result.
27
Now integrating both sides from 0 to t gives
ˆ t
−Lt ε
e−Lt̃ dt̃ = 1 − e−Lt ,
e U (t) − U (0) ≤ ε
0 L
i.e.
ε Lt
U (t) ≤ U (0)eLt +
e −1 t ∈ [0, τ ].
L
Since U (0) = C0 and u(t) ≤ U (t) equation (1.12) now follows.
If
ẋ = f (x) x(0) = x0
and
ẏ = g(y) y(0) = y0
have solutions t 7→ x(t) and t 7→ y(t) respectively, then while x(t), y(t) ∈ U ,
ε L|t|
|x(t) − y(t)| ≤ |x0 − y0 |eL|t| +
e −1 . (1.13)
L
Since
and (1.13) now follows using Gronwall’s inequality from Lemma 1.8.
28
Note that if f = g and we consider the equation ẋ = f (x) with two different
initial conditions x0 and y0 then we get ε = 0 in this result and (1.13) becomes
The solutions are continuous with respect to the initial condition, but can separate
exponentially fast. This is at the basis of the ‘butterfly effect’: small changes in
initial conditions can magnify very rapidly. [Note, however, that you can see this
sort of behaviour even in the simple linear system ẋ = λx, when λ > 0, but there
is no ‘chaos’ here.]
ẋ = f (x) x(0) = x0
If the solution can be found explicitly, then we can compute J(x0 ) exactly.
29
Figure 1.2: Solutions of ẋ = x2 : finite-time blowup.
Otherwise, set β = sup T < ∞. If β ∈ T then the solution exists on [0, β], with
x(β) ∈ U . But now the local version of Picard’s Theorem allows us to extend the
solution beyond t = β, contradicting the definition of β. So we must have β ∈ /T
and the solution exists only on [0, β).
Theorem 1.12 (Unbounded solution). Suppose that U ⊂ Rn is open, and that
f : U → Rn is locally Lipschitz. Let J = (α, β) be the maximal interval of existence
for
ẋ = f (x) x(0) = x0 .
30
If β < ∞ then for every compact subset K of U there exists a time tK < β such
/ K for all t ≥ tK . In particular, if U = Rn , then |x(t)| → ∞ as t → β.
that x(t) ∈
Proof. Suppose that β < ∞ but that there exists a compact set K and a sequence
(tn ) with tn ↑ β such that x(tn ) ∈ K. Since K is compact, the local existence time
for solutions starting at initial conditions in K is some τ > 0, uniform over K. So
if n is large enough, tn + τ > β, a contradiction.
If U = Rn then taking the compact set K = {|x| ≤ R} we see that for each
R there exists tR < β such that |x(t)| ≥ R for all t ≥ tR , i.e. |x(t)| → ∞ as
t → β.
We have already seen the example ẋ = x2 , for which the solution blows up in
finite time. For a (much more artificial) example where the solution tends to the
boundary of U , consider
1
ẋ = f (x) = − x(0) = x0 > 0.
x
Take U = (0, ∞). To solve this equation
x2 (t) x20
q
−x dx = dt ⇒ − + =t ⇒ x(t) = x20 − 2t.
2 2
The solution is contained in U on the maximal interval J = (−∞, x20 /2); and
If U = Rn and there is no blowup then Theorem 1.12 means that the solution
must exist for all t ∈ R. For globally Lipschitz equations (e.g. linear equations)
there is no blowup (see Examples 2) which gives another proof that such equations
enjoy global existence of solutions.
31
Chapter 2
Linear systems
ẋ = Ax x(0) = x0 ,
We have remarked earlier that a unique solution of this equation exists for all
t ∈ R (since globally Lipschitz equations have unique solutions for all time). We
also showed earlier (see discussion after (1.6)) that close to t = 0 the solution is of
the form ∞
X (tA)k
x(t) = etA x0 = x0 .
k=0
k!
We now show that etA is defined for all t ∈ R, and discuss how to compute it.
32
Example 2: Jordan block
λ ε k (tλ)k εktk λk−1
A= , so (tA) = ;
0 λ 0 (tλ)k
adding entry-by-entry now yields
λt
tA e εteλt
e = ,
0 eλt
since ∞ ∞ ∞ k k
X ktk λk−1 X tk λk−1 X t λ
ε =ε = εt .
k=0
k! k=1
(k − 1)! k=0
k!
We now show that etA is always well defined, i.e. the series always converges.
Lemma 2.1. For any n × n matrix A = (aij ) the series
∞
X (tA)k
k=0
k!
converges absolutely for every t ∈ R and so etA is well defined for every t.
Proof. Write aij (k) for the ij entry of Ak , and let a = maxi,j |aij |. Then
n
X
|aij (2)| = ail alj ≤ na2 ≤ (na)2 ;
l=1
so each ij entry converges absolutely by the comparison test, hence so does the
full matrix ∞
X (tA)k
,
k=0
k!
so etA is well defined.
33
We now prove some properties of etA .
(T −1 AT )k = (T −1 AT )(T −1 AT ) · · · (T −1 AT ) = T −1 Ak T
So now ∞ X
A+B
X Aj B k
e = ;
n=0 j+k=n
j! k!
34
where ′ is over j + k ≤ 2m, 0 ≤ j ≤ m, m + 1 ≤ k ≤ 2m, and ′′ is over
P P
j + k ≤ 2m, m + 1 ≤ j ≤ 2m, 0 ≤ k ≤ m.
Now,
m 2m m 2m
X ′ Aj B k
X ∥Aj ∥ X ∥B k ∥ X ∥A∥j X ∥B∥k
≤ ≤ ;
j! k!
j=0 j! k=m+1 k! j=0
j! k=m+1 k!
| {z } | {z }
→e∥A∥ →0
The matrix
a b
A=
−b a
has eigenvalues a ± ib. We want to compute eAt in this case.
then
0 ab
BC = = CB,
−ab 0
so we can use (ii) of the above theorem. We have already shown that eBt = eat I;
for C we have
2n+1 0 (−1)n bn 2n (−1)n bn 0
C = , C = ,
−(−1)n bn 0 0 (−1)n bn
so
Ct cos bt sin bt
e =
− sin bt cos bt
35
and so
at cos bt sin bt
exp(At) = e .
− sin bt cos bt
Each entry of the matrix etA depends on t, so we can differentiate it: we define
d tA
dt
e to be the matrix obtained by differentiating each entry.
ẋ = Ax, x(0) = x0 ,
36
2.1 General linear n × n systems with constant
coefficients
ẏ = T −1 ẋ = T −1 Ax = [T −1 AT ]y.
First note that if A is a multiple of the identity then we do not need to do any
coordinate transformation put A into its canonical form. So we will exclude this
case in what follows.
[The ‘trick’ here, and in other cases, is to write AT as T AJ (moving from the first
to the second line), so that T −1 T cancels.]
37
In this case
λ1 0
Aj = .
0 λ2
If we have a repeated real eigenvalue λ for which there is only a single eigen-
vector v1 , then to find the Jordan Canonical Form we find a second vector v2 such
that
(A − λI)v2 = εv1 and so Av2 = λv2 + εv1 .
It is usual to choose ε = 1, but this more general choice will be useful later.
So here
λ ε
AJ = .
0 λ
If the matrix A is real and it has complex eigenvalues, these must occur in
complex conjugate pairs. So the only way we can have complex eigenvalues for a
2 × 2 matrix is if the eigenvalues are ρ ± iω. So there are two distinct eigenvalues:
the usual (complex) Jordan canonical form of A would be
ρ + iω 0
;
0 ρ − iω
If
Av = (ρ + iω)v (2.1)
and we split the eigenvector v into its real and imaginary parts, v = v1 + iv2 , then
we have
38
[Note that the other eigenvector is v1 − iv2 with eigenvalue ρ − iω: we get this by
taking complex conjugates on both sides of (2.1), since A is real.]
If we change to the basis1 {v1 , v2 } then with respect to this basis A becomes
so we have
ρ ω
AJ = .
−ω ρ
α1 v+ + α2 v− = 0
we have
β1 (v+ − v− ) − β2 i(v− v− ) = (β1 − iβ2 )v+ + (β1 + iβ2 )v− = 0.
Since v+ and v− are linearly independent over C, it follows that we must have
β1 − iβ2 = β1 + iβ2 = 0,
39
or
λm ε 0 ··· 0
..
0 λm ε . 0
.. .. ..
0 0 . . .
.. .. ..
. . . λm ε
0 0 ··· 0 λm
or
µ m νm ε 0
−νm µm 0 ··· 0
0 ε
µ m νm ε 0 ...
0 0
−νm µm 0 ε
... ... ..
0 0 .
.. .. .. µ m νm ε 0
. . .
−νm µm 0 ε
µ m νm
0 0 ··· 0
−νm µm
In the usual Jordan form we use ε = 1, but in fact any ε can be chosen by an
appropriate rescaling of the basis vectors, and we will use the form with small ε
later in the module.
There are now two ways to think about finding the solution of ẋ = Ax.
1. We just want to calculate eAt . Using part (i) of Lemma 2.2 we know that
x(t) = T exp(tAJ )T −1 x0 .
40
Here we don’t actually use the equation for y at all; the Jordan canonical form is
just a useful trick for finding an expression for eAt more easily.
2. But it can be useful to think about the y system: this is a linear system that
satisfies the simpler equation ẏ = AJ y.
The type of picture (node, saddle, etc...) will not be changed by this trans-
formation, it just ‘moves the eigenvectors’. We use this observation in the next
section to classify all 2D linear systems and draw their phase portraits.
Example 3D system
Consider
1 0 0
ẋ = 1 2 0 x, x(0) = x0 .
1 0 −1
The eigenvalues are 1, 2, and −1, with respective eigenvectors
2 0 0
−2 , 1 , 0 .
1 0 1
Then
1 0 0 et 0 0
AJ = T −1 AT = 0 2 0 , etAJ = 0 e2t 0 ,
0 0 −1 0 0 e−t
41
and
etA = T etAJ T −1
t
2 0 0 e 0 0 1 0 0
1
= −2 1 0 0 e2t 0 2 2 0
2
1 0 1 0 0 e−t −1 0 2
t
2 0 0 e 0 0
1
= −2 1 0 2e2t 2e2t 0
2 −t
1 0 1 −e 0 2e−t
et 0 0
= −et + e2t e2t 0 .
(e − e−t ) 0 e−t
1 t
2
So
et 0 0
x(t) = −et + e2t e2t 0 x0 .
(e − e−t ) 0 e−t
1 t
2
First we look at some systems in Jordan canonical form. Then we look at general
systems in terms of their eigenvalues. From the comments in the previous section,
the classification of the ‘canonical system’ transfers immediately to the general
case, since the only difference is a change of coordinates. [One could turn this into
a definition: the classification of a linear system ẋ = Ax is that of the ‘canonical
system’ ż = AJ z, where AJ is the (real) Jordan canonical form of A.]
• If µ < λ < 0 then we have a stable node, and for every x0 , x(t) → 0 as
t → ∞. The solution is λt
e x
x(t) = µt 01 .
e x02
42
Note that we have x2 (t) = Cx1 (t)µ/λ , and µ/λ > 1, so the solutions curves are tan-
gent to the x1 axis. (Solutions approach the origin tangent to the slower direction.)
Figure 2.1: A stable node: eigenvalues µ < λ < 0. Solutions approach the origin
tangent to the slower direction.
• For µ < 0 < λ the origin is a saddle point. For most initial conditions
|x(t)| → ∞, unless x01 = 0, in which case x(t) → 0. Solutions move on curves
xλ y ∥mu| =constant.
43
Figure 2.3: A saddle: eigenvalues µ < 0 < λ.
ẋ = λx + εy
ẏ = λy,
44
ΛE = λE = EΛ and E 2 = 0, we have
tλ tλ
tA t(Λ+E) tΛ tE e 0 1 tε e εtetλ
e =e =e e = = .
0 etλ 0 1 0 etλ
It’s easier to understand the solutions by using polar coordinates (r, θ) than by
using the explicit solution. We have
and
r2 θ̇ = xẏ − y ẋ = −br2 ⇒ θ̇ = −b.
The origin is a stable focus if a < 0, an unstable focus if a > 0, a centre if a = 0.
45
Chapter 3
ẋ = f (x),
as functions of the initial value x0 and the time T . The we can try to describe all
solutions rather than study each IVP separately. If the solutions exist for all t ∈ R
(‘global solutions’) then we can use the differential equation to define a ‘flow’, i.e. a
map ϕ : X × R → X, (x0 , t) 7→ ϕt (x0 ), where ϕt (x0 ) is the solution of (3.1) at time
t (we could also write x(t; x0 ), or something similar, if we wanted the notation to
look more like that in the previous chapter).
So ϕt (x) satisfies
d
ϕt (x) = f (ϕt (x)) ϕ0 (x) = x.
dt
46
For example, for ẋ = Ax we have ϕt (x0 ) = eAt x0 .
Note that it is possible to define a flow ‘abstractly’: it need not arise from a
differential equation.
Definition 3.1. Let X be a subset of Rn (in fact we could take (X, d) to be any
metric space). A map (t, x) 7→ ϕt (x) from R × X → X is a flow on X if
47
We now define various ‘orbits’ of a flow, which correspond to solutions of the
ODE.
Definition 3.2. The orbit of a point x ∈ X (or the trajectory through x) is the
set
O(x) := {ϕt (x) : t ∈ R} ⊂ X (3.5)
or sometimes the indexed set (ϕt (x))t∈R if the time parametrisation is important
[this indexed set contains much more information than O(x), since it gives the
direction and speed at which the orbit is traversed].
Figure 3.2: The forward and backwards orbits O± (x) through a point x.
A periodic point x is one for which ϕT (x) = x for some T > 0 but ϕt (x) ̸= x
for 0 < t < T and then
ϕt (x) = ϕt+T (x) for all t ∈ ℜ
48
and
O(x) = {ϕt (x) : 0 ≤ t ≤ T }
is called a periodic orbit.
0 b
Example: if ẋ = Ax with A = in X = R2 then
−b 0
At cos bt sin bt
ϕt (x) = e x = x;
− sin bt cos bt
the origin is a fixed point (ϕt (0) = 0 for every t ∈ R) and every point x ̸= 0 is
periodic, with O(x) a (circular) periodic orbit (of period T = 2π/b).
49
then O(y) is a homoclinic orbit.
A heteroclinic orbit connects two distinct fixed points: if x0 ̸= x1 are both fixed
points, and for some y ̸= x
ϕt (y) → x0 as t → −∞ ϕt (y) → x1 as t → ∞
Figure 3.6: A heteroclinic orbit that connects two distinct fixed points x0 and x1 .
50
Definition 3.3. A subset Λ ⊂ X is invariant under a flow ϕ if
Fixed points, periodic orbits, homoclinic and heteroclinic orbits are all invariant
sets, as are any orbit O(x), the whole state space X, and – in the case of two
dimensions – the area inside any periodic orbit, homoclinic loop, or heteroclinic
loop.
Figure 3.8: Periodic orbits, homoclinic loops, and hereoclinic loops can divide R2
into two distinct regions, an ‘inside’ and an ‘outside’.
Now consider
ẋ = f (x), x ∈ R, (3.7)
51
where f is a Lipschitz continuous function, so that the local flow ϕ is defined (i.e.
ϕt (x) exists at least for t close to zero).
Proposition 3.5. The orbits of (3.7) consist of fixed points (where f (x) = 0),
heteroclinic orbits joining distinct fixed points, and orbits which tend to ±∞ or
come from ±∞.
We saw in the Differential Equations module that we can easily sketch the phase
portrait for one-dimensional systems. We now prove a result that guarantees that
gives the right behaviour.
Figure 3.9: It is easy to sketch the phase portrait for a 1D system ẋ = f (x), by
sketching the graph of the function f .
Fix(f ) := {x ∈ X : f (x) = 0}
52
Proof. Taking x ∈ (a, b), uniqueness of solutions implies that ϕt (x) cannot cross
a and b (since they are fixed points). So ϕt (x) ∈ (a, b) and exists for all t ∈ R by
Theorem 1.12.
Now, if c < b then f (c) = δ > 0, and since f is continuous there exists ε > 0 such
that f (y) ≥ δ/2 for y ∈ [c−ε, c+ε]. Now, for some τ > 0 we have ϕτ (x) ∈ [c−ε, c];
but then if we take τ ′ > τ + 2ε/δ we have
ˆ τ′
ϕτ ′ (x) = ϕτ (x) + f (ϕt (x)) dt > c,
τ
Note that if we have an equation posed on [0, ∞) (such as a radial equation for
a planar system in polar coordinates) we have a parallel result to Proposition 3.6.
ϕtk (x) → y as k → ∞
The set
53
[the set of all ω-limit points of x] is called the omega-limit set (ω-limit set) of x.
If y is on a periodic orbit then, given any x ∈ O(y), there exists t such that
x = ϕt (y) and then if T is the period of y then
and so we can let t±k = t ± kT ; (t±k ) → ±∞, and ϕt±k (y) → x and k → ∞.
Examples in R2
ṙ = r(1 − r2 ) θ̇ = 1. (3.8)
Figure 3.10: Graph of r(1 − r2 ) and the 1D phase portrait for the r dynamics.
54
Figure 3.11: α and ω limit sets for equation (3.8).
55
Properties of the ω-limit set
Proof. First we show that ω(x) is closed. Suppose that yk ∈ ω(x), k ∈ N, with
yk → y ∗ ∈ X; we need to show that y ∗ ∈ ω(x).
For each n ∈ N we can find kn such that |ykn − y ∗ | < 1/n. Since ykn ∈ ω(x),
we can find tn > tn−1 + 1 such that
and so
Now to show that ω(x) is invariant, take y ∈ ω(x); then there exist (tj ) with
tj → ∞ such that ϕtj (x) → y. Since ϕτ : X → X is continuous for every t ∈ R, we
have
ϕτ +tj (x) = ϕτ (ϕtj (x)) → ϕτ (y)
as j → ∞; so ϕτ (y) ∈ ω(x) for any τ ∈ R, i.e. ω(x) is invariant.
If A is closed then there exists a point ā ∈ A such that dist(x, A) = |x − ā|, see
Examples Sheet 3.
Proof. Since ϕt (x) ∈ K for all t ≥ 0, it follows that in particular (ϕn (x)) is a
bounded sequence in Rn , and hence by the (Rn -generalisation of the) Bolzano–
Weierstrass Theorem, it has a convergent subsequence whose limit is an element
of ω(x) by definition, which shows that ω(x) is non-empty. Since K is compact
it is closed, so the limit of any sequence contained in K (such as ϕtn (x) for some
tn → ∞) lies in K.
56
Now suppose that dist(ϕt (x), ω(x)) does not converge to zero: this means that
there is some ε > 0 and a sequence tn → ∞ such that
Since ϕtn (x) is a sequence contained in the compact set K, it must have a conver-
gent subsequence: so ϕtnj (x) → y. Then y ∈ ω(x), so
Note that in the proofs of these two results we only ever used ϕt with t ≥ 0;
this will be useful later.
The fundamental equality to move between Cartesian and polar coordinates is, of
course
(x, y) = (r cos θ, r sin θ).
Multiply the first equation by y, the second by x, and subtract the first from the
second
ry ẋ = ṙxy − ry 2 θ̇
rxẏ = ṙxy + rx2 θ̇,
− − − − − − − − −−
r[xẏ − y ẋ] = r(x2 + y 2 )θ̇ = r3 θ̇,
57
For a quicker derivation of the expression for θ̇, ignoring possible problems
when x = 0, we can write
1 xẏ − y ẋ xẏ − y ẋ xẏ − y ẋ
θ = tan−1 (y/x) ⇒ θ̇ = y2 2
= 2 2
= .
1+ x x +y r2
x2
In both of these cases (rṙ = xẋ + y ẏ and r2 θ̇ = xẏ − y ẋ) you will most likely
end up with an expression on the right-hand side that still involves x and y. To
get a system for r and θ you then have to put x = r cos θ and y = r sin θ. [Very
occasionally it will be easier to interpret the right-hand side in x and y coordinates,
though!]
3.5 Stability
58
Example in R2 :
ẋ = y
ẏ = −x.
r 2 = x2 + y 2 ⇒ rṙ = xẋ + y ẏ = xy − yx = 0
and
xẏ − y ẋ −x2 − y 2
θ̇ = = = −1
r2 x2 + y 2
so
ṙ = 0, θ̇ = −1.
Figure 3.14: A centre: the origin is Lyapunov stable, and so is each periodic orbit
γR , R > 0.
This is easy to ‘see’, but to apply the definition carefully, first take any neighbour-
hood U of γR ; then (by Lemma 1.6) there exists ε > 0 such that
V := {(r, θ) : |r − R| < ε} ⊂ U.
59
The region V is invariant (it is bounded by two closed orbits and solutions cannot
cross), so x ∈ V ⇒ ϕt (x) ∈ V ⊂ U for all t ≥ 0
ẋ = x, ẏ = −y,
However, more useful is the following definition, which combines both kinds of
‘stability’.
60
An invariant set that is asymptotically stable if we reverse time is called asymp-
totically unstable.
In the example
ṙ = r(1 − r2 ) θ̇ = 1 (3.10)
the periodic orbit {r = 1} is asymptotically stable and the origin is asymptotically
unstable. Cartesian version: ẋ = x − y − x3 − xy 2 , ẏ = x + y − yx3 − y 3 .]
Figure 3.16: Phase portrait for equation (3.10): the origin is asymptotically un-
stable, and the circle at r = 1 is asymptotically stable.
Example: the damped Duffing equation (see Problem Sheet 5) has phase por-
trait:
61
Figure 3.17: The point p is attracting, but not stable.
Figure 3.18: Basins of attraction of two fixed points for the damped Duffing equa-
tion.
Lemma 3.15. If Λ is a compact invariant set that is attracting then B(Λ) is open.
Now suppose that y ∈ B(Λ); then, since ϕt (y) → Λ, there exists t > 0 such
that dist(ϕt (y), Λ) < ε. Since ϕt (x) is continuous in x, there exists δ > 0 such that
for all z ∈ B(y, δ), dist(ϕt (z), Λ) < ε, and so then ϕt+s (z) → Λ as s → ∞. This
shows that z ∈ B(Λ) for all z ∈ B(y, δ), so B(Λ) is open.
62
Note that if Λ is not attracting then its basin of attraction may not be open,
e.g. consider a saddle point as in Figure 3.15. Its basin of attraction is the y axis,
which is not an open subset of R2 .
Definition 3.16. If B(Λ) = X (the whole space) then we say that Λ is a global
attractor.
Easy example in R2 : ẋ = −x, ẏ = −2y, then the origin is the global attractor.
Figure 3.19: For the linear system ẋ = −x, ẏ = −2y, the origin is the global
attractor.
[Without our definition there can be more than one global attractor: for example
in the 1D system ẋ = x(1 − x2 ) both {−1, 0, 1} and [−1, 1] are global attractors.
If we alter the definition so that the global attractor has to attract bounded sets
of initial conditions at a uniform rate then we obtain uniqueness (and some other
nice properties, e.g. connectedness).]
63
Chapter 4
We now introduce some general concepts, motivated and illustrated by some par-
ticular examples.
When there is a conserved quantity (e.g. the energy) this can help us understand
the dynamics of a system of ODEs.
[p2 /2 is the kinetic energy, − cos θ the gravitational potential energy - we have set
all constants equal to 1 to make the algebra simpler] is constant:
d ∂H ∂H
H(x) = ṗ + ẋ = p(− sin θ) + sin θp = 0.
dt ∂p ∂x
Definition 4.1. A system ẋ = f (x) is said to be conservative if there exists a
non-trivial C 1 function H : X → R that is constant along all orbits, i.e.
d
H(ϕt (x))|t=0 = 0 for every x ∈ X.
dt
(By non-trivial we mean that DH ̸= 0 almost everywhere.)
65
so each ΣE is an invariant set. If we study the dynamics on each level set then we
can reduce the dimension of the problem by one.
If we want to understand the structure of the level sets then the following
‘Level Set Lemma’ is extremely useful. Recall that a critical point of a function
H : R2 → R is a point z at which ∇H(z) = 0.
Theorem 4.2. Let H : R2 → R be a C 1 function and let E be one of the connected
components of
• if E < −1 then ΣE = ∅.
• if −1 < E < 1 then ΣE is a closed curve encircling the origin made of the two
parts p
{(θ, 2(cos θ + E)) : −θE ≤ θ ≤ θE }
and its mirror image in the p = 0 axis
p
{(θ, − 2(cos θ + E)) : −θE ≤ θ ≤ θE },
We could also use the Level Set Lemma to show that this is a closed curve: the
range of θ is bounded (as we just showed), and p2 ≤ 2(1 + E), so p is bounded.
66
Since the only critical points of H are (0, 0) and (0, ±π), these level sets are closed
curves of finite length. (You can check directly that they have finite length, but it
is painful.)
In fact, this curve ΣE must be a periodic orbit: we use the fact that ΣE has
finite length ℓE , and there are no fixed points on ΣE (the only two fixed points are
(0, 0), where E = −1, and (π, 0), where E = 1).
So, since f is continuous, the trajectory along ΣE must move in only one
direction (f is never zero on ΣE , so the trajectory cannot change direction) and
|f | ≥ δ > 0 on ΣE . To see this second fact, observe that since |f | : ΣE → R is
continuous and ΣE is compact, |f | attains its lower bound on ΣE : this must be
some δ > 0, since otherwise there would be point on ΣE with |f (x)| = 0.
• if E = 1 then ΣE consists of the fixed point (π, 0) and two orbits homoclinic to
(π, 0).
• if E > 1 then ΣE consists of two disjoint closed curves that wind around the
cylinder; again, these are both periodic orbits, which represent clockwise and anti-
clockwise rotations of the pendulum. [The argument would be the same as in the
case |E| < 1, but showing that the closed curves (remember that they are actually
on the cylinder) have finite length is easier since the integrand in the arc-length
integral is bounded above.]
The fixed point (0, 0) is stable: given a neighbourhood U of (0, 0), let
ε = inf{H(x) : x ∈ X \ U } > −1
67
Figure 4.2: The phase portrait for the pendulum.
and let
V = {x : H(x) < ε}.
Then V is an invariant set entirely contained in U , so if x ∈ V then ϕt (x) ∈ V ⊂ U
for all t ≥ 0.
Suppose that an ε as in (4.1) does not exist. Then there exist xn such that
|H(xn ) − E| < 1/n and xn ∈ / U . Since xn ∈ Γ1 is bounded it is contained in a
compact set, and so there is a subsequence xnj such that xnj → x∗ . Since H is
continuous, H(x∗ ) = E, so x∗ ∈ Γ.
68
Figure 4.3: The phase portrait from Figure 4.2 ‘rolled up’ again onto the cylinder.
The ‘back’ of the cylinder looks essentially the same (but reflected).
69
Figure 4.5: The origin is stable.
[We could use the same argument for the invariant set Σ1 made of the two
homoclinic orbits (blue in Figure 4.2). If we want to use a similar argument for
either one of the orbits that make up ΣE , then we would start with the definition
for the upper portion, or the same with p < 0 for the lower portion.]
The fixed point (π, 0) is unstable, meaning not Lyapunov stable: we can find a
neighbourhood U of (π, 0) such that for any neighbourhood V of x, ϕt (x) ∈ / U for
some t > 0. Take U to be a ball of radius 1, say, centred at (π, 0). Then whatever
neighbourhood V of (π, 0) we satrt in there are points in V on the homoclinic
trajectories whose forward orbit leaves U after some time.
Similarly, neither of the homoclinic orbits taken individually are stable. How-
ever, Σ1 is stable, as indicated above.
Here the Hamiltonian H is often the energy, e.g. for the pendulum if we set
70
Figure 4.6: (π, 0) is unstable, as are the homoclinic orbits joining this fixed point
to itself.
H = p2 /2 − cos θ then
∂H ∂H
θ̇ = =p ṗ = − = − sin θ.
∂p ∂θ
Conserved quantities can be extremely useful: they can be used to reduce the
dimension of the problem by one; if X = R2 then a conserved quantity turns the
problem into a collection of one-dimensional problems. If X = R3 then dynamics
can be reduced to two-dimensional dynamics, and these are much simpler than
dynamics in any higher dimension because of the Jordan Curve Lemma (‘a closed
curve in the plane has an inside and an outside’). Trajectories cannot cross from
inside to outside a periodic orbit or a homoclinic/heteroclinic loop.
71
Example: the damped pendulum
dH
= (sin θ)p + p(−kp − sin θ) = −kp2 ≤ 0 for every (θ, p).
dt
So H is non-increasing along orbits: H(ϕt (x)) ≤ H(x) for all x ∈ X, t ≥ 0. We
only have dH/dt = 0 when p = 0, but if θ ∈ / {0, π} then ṗ ̸= 0 and the orbits are
only at such points instantaneously.
More formally, since for any θ ̸= 0, π there exists ε > 0 such that |ṗ| > ε in a
neighbourhood of x = (θ, 0), we have
ˆ τ
2
∆H ≤ ε2 t2 dt = − kτ 3 ε2 < 0,
−k |{z}
−τ 3
|p(t)|≥ε|t|
so H decreases along all orbits expect for the two fixed points.
72
Figure 4.7: Apart from at the two fixed points, ∆H < 0.
i.e. Mc is all x that have forward orbits along which H is constant and equal to
c. For example, for the damped pendulum H decreases along all orbits except for
the fixed points, so
The LaSalle Invariance Principle says that the Lyapunov function is constant
on any ω-limit set.
Theorem 4.6 (LaSalle Invariance Principle). If H is a Lyapunov function for a
flow ϕ then for every x ∈ X there exists c ∈ R such that
Proof. Pick x ∈ X. If ω(x) is empty then there is nothing to prove, as the empty
set is a subset of any set.
Otherwise, let
Now take y ∈ ω(x): then there is an increasing sequence (tj ) → ∞ such that
ϕtj (x) → y as j → ∞. Since H is continuous, H(y) = limj→∞ H(ϕtj (x)) = c.
73
All that remains is to show that y ∈ Mc , i.e. that H(ϕt (y)) = c for all t ≥ 0.
But we know that ω(x) is invariant, so if y ∈ ω(x) we have ϕt (y) ∈ ω(x), and we
have just shown that H(z) = c for all z ∈ ω(x), so H(ϕt (y)) = c as required.
Note that in this proof we only ever use properties of ϕt for t ≥ 0. [The
results on ω-limit sets from Propositions 3.9 and 3.10 only use this ‘future’ part of
ϕt .] This means that the Invariance Principle is also true if we have an Lyapunov
function that is defined on some positively invariant subset V of X (and then (4.3)
holds for all x ∈ V ); we will use this observation in the proof of Theorem 5.4.
Proof. Since there are only a finite number of fixed points, there exists δ > 0 such
that for any pair of distinct fixed points y and z, |y − z| > δ.
Suppose that ω(x) contains two distinct fixed points, y and z. Then, since
y ∈ ω(x), there exists ty > t0 such that
|ϕt x − y| = δ/2.
74
If w ̸= y is any fixed point then
So we have |ϕt x − w| ≥ δ/2 for any fixed point (including x): in other words,
dist(ϕt x, ω(x)) ≥ δ/2. But t > t0 , contradicting (4.4). It follows that ω(x) consists
of a single fixed point as claimed.
{y ∈ X : H(y) ≤ E}
La Salle’s Invariance Principle implies that there must exist c ∈ {−1, 1} such
that ω(x) ⊂ Mc , i.e. ω(x) = {(0, 0)} or {(π, 0)}, and Proposition 3.10 shows that
ϕt (x) → ω(x), so the forward orbit must converge to one of the fixed points.
We would like to be able to identify which fixed points orbits of the damped
pendulum tend to. We claim that the phase portraits are as in Figure 4.8 (for the
case k = 2 see Sheet 6, Q2).
Just two orbits tend to (π, 0) and all the others go to (0, 0). To justify this, we
will have to examine the dynamics near the fixed points, which we will do in the
next section via linearisation.
75
76
Figure 4.8: Phase portrait for the damped pendulum, depending on the value of
k. For k = 2 see Examples Sheet 6.
Figure 4.9: The phase portrait for the damped pendulum presented as in Figure
4.4.
Chapter 5
By Taylor’s Theorem,
The equation
ẏ = Df (x∗ )y (5.1)
77
is called the linearised equation near x∗ .
In this section we will see what features of the linear system (5.1) persist when
we reintroduce the nonlinear terms.
Definition 5.1. A fixed point x∗ is hyperbolic if every eigenvalue of Df (x∗ ) has
non-zero real part.
We will see that the local phase portrait near any hyperbolic fixed point x∗
‘looks like’ that of the linearised system. There are three possible cases:
• if all eigenvalues of Df (x∗ ) have negative real parts then x∗ is a sink and for the
original system x∗ is attracting;
• if all eigenvalues of Df (x∗ ) have positive real parts then x∗ is a source and for
the original system x∗ is a repellor;
At x∗ = (0, 0) we have
0 1
Df (0, 0) = ,
−1 k
78
whose eigenvalues are the solutions of λ2 + kλ + 1 = 0, so are
√
−k ± k 2 − 4
λ= ;
2
if k = 0 (ideal pendulum) then λ = ±i, so (0, 0) is not hyperbolic
if 0 < k < 2 then the origin is a stable focus
if k = 2 the origin is a stable improper node
if k > 2 the origin is a stable node
At x∗ = (π, 0) we have
0 1
Df (π, 0) = ,
1 −k
To show that sinks are attracting we will use the Adapted Norm Lemma, which
we prove here in the two-dimensional case (see separate ‘handout’ for the general
case).
Lemma 5.2 (Adapted Norm Lemma). Let A be a linear operator on a real n-
dimensional vector space E. Suppose that all the eigenvalues of A satisfy
a < Re λ < b, for some a, b ∈ R.
Then there exists a basis B = {b1 , . . . , bn } such that
a∥x∥2B ≤ ⟨x, Ax⟩B ≤ b∥x∥2B for every x ∈ E, (5.2)
where ⟨· ·⟩B denotes the usual inner product in this basis and ∥·∥B is the associated
norm.
79
If x = nj=1 xj bj and y = nj=1 yj bj we define ⟨x, y⟩B = nj=1 xj yj and set
P P P
p
∥x∥B = ⟨x, x⟩B . It is easy to show that there exist constants c1 , c2 with 0 <
c1 ≤ c2 such that
c1 ∥x∥B ≤ |x| ≤ c2 ∥x∥B , (5.3)
where |x| is the usual Euclidean norm of x. [See handout: or use the fact that all
norms on Rn are equivalent, see Norms, Metrics, & Topologies.
where we choose (if need be) ε sufficiently small so that λ ± ε ∈ [a, b]. Then,
case-by-case, for x = (x1 , x2 ) in terms of the basis B, ⟨Ax, x⟩ is given by
λ1 0 x1
x 1 x2 = λ1 |x1 |2 + λ2 |x2 |2
0 λ2 x2
λ ε x1 λx1 + εx2
x1 x2 = (x1 x2 ) = λ|x1 |2 + εx1 x2 + λ|x2 |2
0 λ x2 λx2
= λ∥x∥2B + εx1 x2
ρ ω x1 ρx1 + ωx2
x1 x2 = x1 x2 = ρ|x1 |2 + ρ|x2 |2 = ρ∥x∥2B .
−ω ρ x2 −ωx1 + ρx2
Note that in the middle case we have |x1 x2 | ≤ x21 + x22 = ∥x∥2B , so in every case
we obtain (5.2).
80
Proof. We will assume that x∗ = 0 (if not we can change coordinates, putting
y = x − x∗ ). Let A = Df (0), and let ε be small enough that
Re λ < −(a + ε)
The Adapted Norm Lemma (Lemma 5.2) shows that Rn has a basis B such
that
⟨x, Ax⟩B ≤ −(a + ε)∥x∥2B for all x ∈ Rn .
By the definition of the derivative we have
∥f (x) − Ax∥B
lim = 0.
x→0 ∥x∥B
[This limit certainly holds if we use in place of ∥ · ∥B the usual norm on Rn ; that
it also holds using ∥ · ∥B is a consequence of (5.3).] since the Cauchy–Schwarz
inequality gives
|⟨x, f (x) − Ax⟩B | ≤ ∥x∥B ∥f (x) − Ax∥B
it follows that
⟨x, f (x) − Ax⟩B
lim = 0.
x→0 ∥x∥2B
so
d
∥x∥B ≤ −a∥x∥B (5.5)
dt
and ∥x(t)∥B is decreasing (whenever x ̸= 0).
Therefore x(t) cannot escape the set V and so x(t) is defined for all t ≥ 0 and
in V for all t ≥ 0.
81
Integrating (5.5) gives ∥ϕt (x)∥B ≤ e−at ∥x∥B , or, translating the fixed point
back to x∗ ,
∥ϕt (x) − x∗ ∥B ≤ e−at ∥x − x∗ ∥B .
Now part (b) follows using (5.3).
Not that by reversing time, if all the eigenvalues λ of Df (x∗ ) have Re λ > 0
then x∗ is repelling (= attracting ‘in negative time’).
We can immediately apply this theorem to guarantee that for the damped
pendulum, (0, 0) is attracting.
We can use a similar method to show that fixed points are attracting by using
other possible choices of Lyapunov functions.
Theorem 5.4 (Lyapunov’s Stability Theorems). Let ϕ be a continuous flow on X
with a fixed point x∗ . Let N be compact subset of X that contains a neighbourhood
of x∗ and let H : X → R be a continuous function with
H(x) > H(x∗ ) for all x ∈ N \ {x∗ }
82
such that H is non-increasing along orbits in N (i.e. H(ϕt (x)) ≤ H(x) for all
x ∈ N and t ≥ 0 such that ϕs (x) ∈ N for all 0 ≤ s ≤ t). Then
Without loss of generality we can assume that H(x∗ ) = 0, so that H(x) > 0
for all x ∈ N \ {x∗ } (replace H(x) by H(x) − H(x∗ ) if need be).
We want to show that for any x ∈ V we have ϕt (x) ∈ V for all t ≥ 0. Given
some x ∈ V , let
Since H is non-increasing we have H(ϕt (x)) ≤ H(x) for all t ∈ [0, τ ). Since H
is continuous and ϕt (x) is continuous in t we must also have H(ϕT (x)) ≤ H(x);
we claim that we must also have ϕT (x) ∈ V .
1 Since H : X → [0, ∞) we could take any open set (−k, δ) with k > 0.
83
Figure 5.3: Construction for proof of Lyapunov’s First Stability Theorem.
Note that
• since ϕt (x) ∈ N for all t ∈ [0, T ) and N is compact, we have ϕT (x) ∈ N ; but
H(y) ≥ δ for all y ∈ N \ W , so we must have ϕT (x) ∈ W .
Therefore
ϕT (x) ∈ W with H(ϕT (x)) < δ
so ϕT (x) ∈ V as claimed.
(ii) If we take V as above and some x ∈ V \ {x∗ }, then ϕt (x) ∈ V for all
t ≥ 0, which (by Proposition 3.10) shows that ω(x) ̸= ∅. Now we need to show
that ω(x) = {x∗ }. This follows immediately from La Salle’s Invariance Principle
applied on the positively invariant set V : since H is strictly decreasing along
trajectories, the only set in V on which H is constant is {x∗ }.
There is a useful variant of the result in (ii), which is worth stating formally.
Corollary 5.5. Suppose that V is a contained in a compact set, and is a positively
invariant set for a flow ϕ. Assume that V contains a fixed point x∗ , H(x) > H(x∗ )
for all x ∈ V \ {x∗ }, and H is strictly decreasing along orbits in V \ {x∗ }. Then
84
ϕt (y) → x∗ as t → ∞ for all y ∈ V . If V contains a neighbourhood of x∗ then x∗
is asymptotically stable.
Proof. The proof of the first part follows exactly the proof of part (ii) in Theorem
5.4, except that now we are given the existence of a positively invariant V as an
assumption.
We can use these results to show that some fixed points that are not sinks are
nevertheless Lyapunov stable or attracting.
Example in R2 :
ẋ = −y − x3 ẏ = x5 . (5.6)
The origin is a fixed point; the linearised dynamics are ẋ = −y, ẏ = 0.
ax2k by 2m
H(x, y) = +
2k 2m
then
dH ∂H ∂H
= ẋ + ẏ = ax2k−1 (−y − x3 ) + by 2m−1 x5 ;
dt ∂x ∂y
85
Figure 5.5: The linearised system ẋ = −y, ẏ = 0 has a line of fixed points.
We can do better with a little more work: the fact that Ḣ = −x8 means that
H is strictly decreasing except when x = 0. But when x = 0, y ̸= 0, we have
|ẋ| = |y|;
in some neighbourhood of (0, y) we have |ẋ| > |y|/2. Suppose that x(0) = (0, y);
then for t small enough we have |x(t)| ≥ |yt|/2, and then
ˆ t ˆ t
8
H(x(t)) − H(x(0)) = − |x(s)| ds ≤ − |ys|8 /28 ds < 0,
0 0
so H is in fact strictly decreasing along all trajectories apart from at the origin.
It now follows that the origin is also attracting, so it is asymptotically stable.
ẋ = −y − x3 ẏ = x5 − y 3 , (5.7)
and it is easy to see that the origin is asymptotically stable without the sort of
more careful argument above.
86
Figure 5.6: The origin is Lyapunov stable for equation (5.6). It is – in fact – also
attracting.
We give the proof in the two-dimensional case; the argument in higher dimen-
sions is essentially the same, but the 2D case is a little more straightforward and
so significantly clearer.
Note that in 2D if both eigenvalues have positive real part then we can use our
previous result in the form ‘sources are repellors’. So we only need to consider the
case that we have one positive eigenvalue µ and one negative eigenvalue −λ (we
take λ, µ > 0).
87
(in the new coordinates) we have
µx R(z)
f (x, y) = + ,
−λy S(z)
where R(z) and S(z) are higher order terms: o(|z|) as z → 0 [recall that g(z) =
o(|z|) as z → 0 means that ∥g(z)∥/|z| → 0 as z → 0.]
Note that we are writing f (z) = Df (0)z + V (z), where V = (R, S). By the
definition of the derivative, |f (z) − Df (0)z| = |V (z)| = o(|z|); this is where the
inequalities for |R(z)| and |S(z)|, the components of V , come from.
d
H(z) = (y, ẏ) − (x, ẋ)
dt
= (y, −λy) + (y, S(z)) − (x, µx) − (x, R(z))
≤ −λ|y|2 − µ|x|2 + |y||S(z)| + |x||R(z)|
≤ −µ|x|2 + |z|(|R(z)| + |S(z)|).
Let
C = {z : |y| ≤ |x|};
then if z ∈ C ∩ U we have
µ 2
|R(z)|, |S(z)| ≤ |x|
2
and so within C ∩ U we have
d µ
H(z) ≤ − |x|2 :
dt 2
H is non-increasing along all orbits in C ∩ U \ {0}.
We have
1d 2 µ
|x| = µ|x|2 + R(z)x ≥ |x|2 ,
2 dt 2
88
and so |x(t)| ≥ eµt |x(0)| as long as x(t) remains in C ∩ U . This will eventually
exceed δ, so that the solution leaves U .
Thus there are points z arbitrarily close to 0 such that ϕt (z) leaves U for some
t > 0. This contradicts the Lyapunov stability of 0, and therefore the origin is
unstable.
Since x∗ is hyperbolic, Rn = E s ⊕ E u , so du + ds = n.
89
the ‘stable subspace’, where ϕLt is the linearised flow arising from (5.8), and
E u = {x ∈ Rn : ϕLt (x) → 0 as t → −∞},
the ‘unstable subspace’. [Note that E s is not a ‘stable set’ in the Lyapunov sense.]
Definition 5.7. The stable manifold W s (x∗ ) and the unstable manifold W u (x∗ )
of the fixed point x∗ are the sets
W s (x∗ ) :={x ∈ Rn : ϕt (x) → x∗ as t → ∞}
W u (x∗ ) :={x ∈ Rn : ϕt (x) → x∗ as t → −∞};
they are the nonlinear versions of the stable and unstable subspaces defined above.
Note that these are invariant: if x ∈ W s (x∗ ) then for any s ∈ R we have
ϕt (ϕs x) = ϕt+s x → x∗
as t → ∞, so ϕs x ∈ W s (x∗ ); a very similar argument works to show that W u (x∗ )
is invariant.
We will state (and use) the following result about the existence of stable and
unstable manifolds.
Theorem 5.8 (Stable and unstable manifolds). Given the above assumptions there
exists a stable manifold W s (x∗ ) of dimension ds and an unstable manifold Wu (x∗ )
of dimension du , both of which contain x∗ , and which are tangent to x∗ + E s and
x∗ + E u , respectively (at x∗ ).
90
Figure 5.9: Stable and unstable manifolds in R3 near a hyperbolic fixed point x̄:
in this picture the stable manifold is two-dimensional and the unstable manifold
is one-dimensional.
θ̇ = p
ṗ = − sin θ − kp,
91
La Salle’s Invariance Principle implies that the ω-limit set of every orbit is one of
the fixed points. The fixed point (π, 0) is a saddle, so the SUM Theorem implies
that (π, 0) is the ω-limit set of precisely two orbits and is the α-limit set of two
other orbits. The origin is then the ω-limit set of all other orbits in S 1 × R.
Figure 5.11: Stable and unstable manifolds for the damped pendulum.
Note that if dim Wxu∗ ̸= 0 then Wxs∗ forms the boundary between different
behaviours of the orbits; it is sometimes therefore called a separatrix. For example,
in the phase portrait for the Duffing equation with damping
ẋ = y
ẏ = −ky + x − x3
(see Sheet 5, Q1) the origin is saddle point and W s (0, 0) is the boundary between
the basins of attraction of the two sinks at (−1, 0) and (1, 0).
We will see the stable manifold playing the role of a ‘separatrix’ again later
when we look at predator-prey systems in certain parameter ranges.
92
5.3.1 Approximating the stable and unstable manifolds as
power series
If f is smooth enough, then we can compute W u (x∗ ) and W s (x∗ ) close to any fixed
point using power series.
ẋ = f (x, y)
ẏ = g(x, y),
We will look for the unstable manifold in the form2 y = Y (x), where Y is a
polynomial:
Y (x) = a0 + a1 x + a2 x2 + a3 x3 + · · · .
Note that we must have Y (0) = 0, because the manifold passes through the origin,
and we must have Y ′ (0) = a1 = α because we know that the unstable manifold is
tangent to the linear unstable space at the origin.
So we try
Y (x) = α + a2 x2 + a3 x3 + · · · . (5.9)
Since the manifold is invariant, if we differentiate y = Y (x) we obtain
ẏ = Y ′ (x)ẋ;
we substitute in from the origin ODE, with y = Y (x), and end up with
We now substitute the power series expansion of Y (x) in (5.9) into this equation
and solve for the aj s by equating coefficients of powers of x.
2 The only situation in which we cannot write the manifold as y = Y (x) is if the linear direction
is vertical; in this case the direction will be (0, 1), and we would have to look for the manifold
in the form x = X(y); we can use (almost) exactly the same approach. Differentiating
P∞ we
must have ẋ = X ′ (y)ẏ, and we’d look for a power series in the form X(y) = k=2 ak y k , there
being no y term since the manifold has to be tangent to the y axis at the origin.
93
For a simple example, consider the two-dimensional system
ẋ = x ẏ = −y + x2 (5.10)
The y-axis (x = 0) is the stable manifold for the nonlinear system. The SUM
Theorem says that the unstable manifold W u (0) should be tangent to the x-axis,
so we try
∞
X
Y (x) = ak x k .
k=2
(−Y + x2 ) = Y ′ (x)x,
which gives
94
Example: the damped Duffing equation near the origin.
Now we look at the unstable manifold at the origin in the damped Duffing equa-
tion, choosing k = 3/2 (which simplifies the algebra). Recall that the equations
are
ẋ = y
ẏ = −3y/2 + x − x3 .
We have ẏ = Y ′ (x)ẋ, so
3
− Y (x) + x − x3 = Y ′ (x)Y (x).
2
This gives
3 hx
+ a2 x2 +a3 x3 + · · · + x − x3
−
2 2 h
1 2 x 2 3
i
= + 2a2 x + 3a3 x + · · · + a2 x + a3 x + · · · .
2 2
95
Coefficients of x2 :
3 1 1
− a2 = a2 + 2a2 ⇒ a2 = 0.
2 2 2
Coefficients of x3 :
3 1 3a3 2
− a3 − 1 = a3 + 2a22 + ⇒ a3 = − .
2 2 2 7
To O(x3 ) we have
x 2x3
Y (x) = − .
2 7
We could do the same for the stable manifold at (0, 0); the stable direction
1
is , so we could look for the stable manifold in the form y = Y (x), where
−2
Y (x) = −2x + a2 x2 + a3 x3 + · · · . This is invariant too, so we still have ẏ = Y ′ (x)ẋ,
i.e.
3
− Y (x) + x − x3 = Y ′ (x)Y (x),
2
which now gives
3
− −2x + a2 x2 + a3 x3 + · · · ] + x − x3
2
= −2 + 2a2 x + 3a3 x2 + · · · −2x + a2 x2 + a3 x3 + · · · .
3 + 1 = 4.
The O(x2 ) term again shows that a2 = 0, and then at O(x3 ) [using the fact that
a2 = 0] we obtain
3
− a3 − 1 = −2a3 − 6a3 ⇒ a3 = 2/13;
2
to third order the stable manifold is y = −2x + (2x3 /13), and the equation on the
stable manifold is
2x3
ẋ = −2x + .
13
96
If you want to do this procedure near a fixed point (x∗ , y ∗ ) that is not the origin
there are two possible approaches: (i) look for a manifold given by the equation
Y (x) = y ∗ + α(x − x∗ ) + a2 (x − x∗ )2 + a3 (x − x∗ )3 + · · ·
We have seen that the dynamics near a hyperbolic fixed point are similar to
those of the linearised system: sinks are still attracting, and saddles have sta-
ble and unstable manifolds that play the same role as the stable and unstable
subspaces. The Hartman–Grobman Theorem shows that orbits of the linearised
system correspond to orbits of the nonlinear system close to the fixed point (the
‘pictures’ are ‘the same’).
Theorem 5.9 (Hartman–Grobman). Suppose that x∗ is a hyperbolic fixed point
of ẋ = f (x) on Rn , where f is continuously differentiable. Then there exists a
neighbourhood U of x∗ and a homeomorphism h from U onto a neighbourhood V
of 0 ∈ Rn such that h ◦ ϕt = ψt ◦ h, where ϕt is the flow associated to ẋ = f (x)
and ψt is the flow associated to the linearisation ξ˙ = Df (x∗ )ξ.
In other words, local to the fixed point h maps orbits to orbits and preserves
the time parametrisation.
97
5.4 Non-hyperbolic fixed points in R2
Examples:
(i) one eigenvalue zero, the other non-zero, e.g. both systems
ẋ = x2 ẋ = x3
ẏ = −y ẏ = −y (5.11)
have the same linearisation at zero, ẋ = 0, ẏ = −y. But the phase portraits are
very different.
Figure 5.16: Phase portraits for the nonlinear systems in (5.11): ‘saddle-node’ on
the left; ‘nonlinear saddle’ on the right.
(ii) the eigenvalues of Df (x∗ ) are purely imaginary (complex conjugate ±iω)
ẋ = y
ẏ = −x.
98
The nonlinear system can be many things:
(a) a ‘nonlinear centre’, like (0, 0) for the pendulum or (±1, 0) for the conservative
Duffing equation.
Figure 5.17: ‘Nonlinear centres’: the pendulum (left) and Duffing’s equation (right)
(d) infinitely many isolated periodic orbits encircling x∗ , accumulating there, e.g.
for
(
1 r sin 1r r ̸= 0
ẋ = y + x sin p ṙ =
x2 + y 2 0 r=0
1
ẏ = −x + y sin p θ̇ = −1.
x + y2
2
99
Figure 5.19: Infinitely many periodic orbits.
ẋ = x(A − ax)
is a simple population model for one species: A is the rate of growth of the popu-
lation without any restrictions (ẋ = Ax) and the −ax2 limits growth due to finite
resources.
Lotka–Volterra models are coupled logistic equations: for two species x and y,
ẋ = x(A − a1 x + b1 y)
ẏ = y(B + b2 x − a2 y),
where A, B, a1 , a2 > 0, and b1 and b2 could be ether sign. The sign of bi determines
the type of interaction:
(ii) b1 , b2 < 0. Competition. Both species are in competition for the same re-
source, e.g. rabbits and sheep (and grass).
(iii) b1 > 0 and b2 < 0. Predator-prey. With these parameters x is the predator
and y the prey (e.g. wolves and sheep).
100
Example: competitive case
ẋ = x(3 − x − 2y)
ẏ = y(2 − x − y).
STEP 1: Find the fixed points and study the linearisation about each fixed
point.
We have
3 − 2x − 2y −2x
Df (x, y) =
−y 2 − x − 2y
and so we have
3 0
(0,0): Df (0, 0) = with eigenvalues and eigenvectors
0 2
1 0
λ1 = 3 e 1 = , λ2 = 2 e2 =
0 1
and this is an unstable node.
−1 0
(0,2): Df (0, 2) = with eigenvalues and eigenvectors
−2 −2
1 0
λ1 = −1 e1 = , λ2 = −2 e2 =
−2 1
and this is a sable node, as is
−3 −6
(3,0): Df (3, 0) = with eigenvalues and eigenvectors
0 −1
1 3
λ1 = −3 e1 = , λ2 = −1 e2 = .
0 −1
−1 −2
(1,1): Df (1, 1) = with eigenvalues and eigenvectors
−1 −1
√
1
λ± = −1 ± 2, e± = ;
∓ √12
and this is a saddle point.
101
STEP 2. Note that all the fixed points are hyperbolic. So the local phase por-
traits near the fixed points look like the phase portraits for the linearised systems.
So now we ‘join the dots’, paying attention too to the directions of the vector
field in regions separated by the ‘nullclines’ (where ẋ = 0 or ẏ = 0).
102
Figure 5.22: Full phase portrait for the Lotka–Volterra model.
s
Note that in the final phase portrait, the stable manifold W(1,1) of the saddle
at (1, 1) divides the quadrant into two regions [the term ‘separatrix’ is sometimes
used]. Everything below this tends to (3, 0), and everything above to (0, 2). So in
this model one species eventually dies out and the other reaches a steady popula-
tion, depending on the initial conditions - apart from the very special case where
s
the initial condition lies on W(1,1) , in which case we end up with equal numbers
of both species. But this point is unstable, so any small changes will lead to
extinction of one species.
103
Chapter 6
Periodic orbits in
two-dimensional systems
104
Pick some z ∈ ω(y). If z is a fixed point then z ∈ ω(y) ⊂ ω(x), so we fall into
the first possibility in the theorem.
We want to prove that all these intersections happen at the same point, in
which case O+ (y) is periodic and the common intersection point is z.
Suppose that this is not the case, and let y1 and y2 be two consecutive, distinct
intersections of O+ (y) with Σ.
105
The region bounded by this portion of the orbit and a section of Σ,
is a ‘trapping region’: either the orbits can enter (through Σ) and then they cannot
escape (left-hand picture), or they can start inside and leave through Σ but then
never return (as in the right-hand picture).
Figure 6.3: (i) once inside orbits cannot escape; (ii) once out, orbits cannot re-
enter.
Now, since y ∈ ω(x) the orbit of x must come arbitrarily close to y; and then,
since solutions depend continuously on the initial conditions, this orbit will stay
close to the orbit through y, so must at some point cross Σ through [y1 , y2 ]. It
is then either trapped inside (so cannot return to close y), or in the second case
trapped outside, and again cannot return close to y. This contradicts the fact
that y ∈ ω(x), so we must have y1 = y2 . This means that ϕt1 (y) = ϕt2 (y), so
ϕt2 −t1 (y) = y and y lies on a periodic orbit γ that is contained in ω(x).
Finally, we want to make sure that in fact ω(x) = γ; to do this we can show
that ϕt (x) → γ as t → ∞.
106
Figure 6.4: Intersections move monotonically along Σ.
For any neighbourhood V of γ, the fact that ϕ is continuous and ϕtj (x) → γ as
j → ∞ implies that there exists a j such that ϕt (x) ∈ V for all t ∈ [tj , tj+1 ]. But
then the trapping implies that in fact ϕt (x) ∈ V for all t ≥ tj , and so ϕt (x) → γ,
so ω(x) can contain no points outside γ and we have finished the proof.
The theorem relies on the existence of a Jordan curve, that divides the state
space into two disjoint parts. So the theorem works on R2 , and also on the sphere
S 2 and the cylinder S 1 × R; but not on the torus T2 . It does not work in Rn when
n > 2.
107
Figure 6.5: A Jordan curve divides the state space into two disjoint parts: this
works on R2 and S 2 but not on the torus.
108
which is strictly negative for all θ, so there are no fixed points (apart from the
origin). To find the radial component we have
1 1
rṙ = xẋ + y ẏ =x(y + x(1 − 2r2 )) + y(−x + y(1 − r2 ))
4 2
2 2
x y
= (1 − 2r2 ) + (1 − r2 )
4 2
x2 y 2 y 2 x2 r 2 y 2 r 2
= + + − −
4 4 4 2 2
2
r 1
= (1 + sin2 θ) − r4 ,
4 2
so
r 1
ṙ = (1 + sin2 θ) − r3 .
4 2
Then ṙ ≥ 0 for all θ if 21 r − r3 ≥ 0, i.e. for 0 < r ≤ √1 ,
2
and ṙ ≤ 0 for all θ if
r − r3 ≤ 0, i.e. if r ≥ 1.
So the region
1
{(r, θ) : √ ≤ r ≤ 1}
2
is forward invariant and contains no fixed points, so it must contain at least one
periodic orbit. All orbits except that of the origin enter this region, so their ω-limit
sets are periodic orbits.
109
Figure 6.7: Setup for the Divergence Theorem.
The function g is the ‘weight function’. In the simple case g = 1 this is called the
‘Divergence Test’. The theorem can be (easily) generalised to rule out homoclinic
or heteroclinic loops.
Proof. Suppose that some periodic orbit γ does lie entirely in D, and let Ω be the
region enclosed by γ. Then since f is tangent to γ (as γ is a trajectory), f · n = 0,
where n is the outward-pointing normal to Ω. So
ˆ
g(f · n) dl = 0.
γ
But our assumption implies that the scalar function ∇·(gf ) is either (strictly) pos-
itive or negative throughout Ω, so this integral must be non-zero, a contradiction.
So there is no periodic orbit lying wholly within D.
Again we consider
ẋ = x(A − a1 x + b1 y)
ẏ = y(B + b2 x − a2 y),
110
1
We use the weight function g(x, y) = xy . Then
1 1
gf = (A − a1 x + b1 y), (B + b2 x − a2 y)
y x
and
a1 a2
∇ · (f g) = −
− <0
y x
for all x, y > 0. So there are no periodic orbits in the positive quadrant, and will
not be if either a1 or a2 are non-zero.
Are there actually periodic orbits? In this case we can find the explicit form
of the solution curves:
dy ẏ y(B − b2 x)
= = .
dx ẋ x(−A + b1 y)
Separating variables we have
ˆ ˆ
−A + b1 y) B − b2 x
dy = dx
y x
111
Figure 6.8: Periodic orbits in the ‘pike and eels’ model.
and so
−A log y + b1 y − B log x + b2 x = E, constant.
So in fact this is a conservative system with
H(x, y) := −A log y + b1 y − B log x + b2 x
conserved along orbits.
By the Level Set Lemma, the level sets of H in x, y > 0 are closed curves away
from the interior fixed point, since (i) the only critical point of H is when
∂H B ∂H A
= − + b2 = 0 and = − + b1 = 0,
∂x x ∂y y
i.e. at (B/b2 , A/b1 ) and (ii) the level sets are bounded, since
exp(b1 y + b2 x)
H(x, y) = E ⇒ = eE ,
y A xB
and eb1 y y −A → ∞ as y → 0, ∞ (similarly for eb2 x x−B ).
Let H0 = H(B/b2 , A/b1 ). Then for any E with H0 < E < ∞ the curve
H(x, y) = E is a periodic orbit. The argument is the same as we used for the
pendulum (finite length, no fixed points, so the trajectory moves in the same
direction with a speed that is bounded below).
112