Difference and Differential Equations: BA Econometrics and Operations Research
Difference and Differential Equations: BA Econometrics and Operations Research
Difference and Differential Equations: BA Econometrics and Operations Research
2013-2014
Author:
G.K. Immink
Contents
1 Difference equations 5
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Asymptotic behaviour and stability . . . . . . . . . . . 13
1.1.2 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.2 Autonomous first order difference equations . . . . . . . . . . 19
1.2.1 Stability of periodic and equilibrium solutions of au-
tonomous first order difference equations . . . . . . . . 24
1.2.2 Bifurcations . . . . . . . . . . . . . . . . . . . . . . . . 32
1.2.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 37
1.3 Linear difference equations . . . . . . . . . . . . . . . . . . . . 38
1.3.1 Inhomogeneous linear difference equations . . . . . . . 44
1.3.2 First order linear difference equations . . . . . . . . . . 44
1.3.3 Stability of solutions of linear difference equations . . . 48
1.3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 50
1.4 Systems of first order difference equations . . . . . . . . . . . 51
1.4.1 Homogeneous systems of first order linear difference
equations . . . . . . . . . . . . . . . . . . . . . . . . . 56
1.4.2 Inhomogeneous systems of first order linear difference
equations . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.4.3 Systems of first order linear difference equations with
constant coefficients . . . . . . . . . . . . . . . . . . . . 58
1.4.4 Stability of solutions of systems of linear difference
equations with constant coefficients . . . . . . . . . . . 67
1.4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 69
1.5 Autonomous systems of first order difference equations . . . . 71
1.5.1 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3
1.6 Higher order linear difference equations with constant coeffi-
cients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
1.6.1 The homogeneous equation . . . . . . . . . . . . . . . 87
1.6.2 Inhomogeneous higher order difference equations; the
annihilator method . . . . . . . . . . . . . . . . . . . . 90
1.6.3 Stability of solutions of higher order difference equations 94
1.6.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Appendix 143
A.1 Change of basis and Jordan normal form . . . . . . . . . . . . 143
A.2 Computation of J n . . . . . . . . . . . . . . . . . . . . . . . . 150
A.3 Definition and properties of eA . . . . . . . . . . . . . . . . . . 151
A.4 Vector and matrix functions . . . . . . . . . . . . . . . . . . . 153
The figures in these lecture notes were produced with the program Dynamics
of H. E. Nusse and J. A. Yorke.
Chapter 1
Difference equations
1.1 Introduction
Difference equations arise in discrete, dynamic economic models.
Example 1.1.1. One of the simplest examples is the classic cobweb model:
qtd = D(pt )
qts = S(pt−1 )
qtd = qts
qtd , qts , and pt denote demand, supply and price at time t, respectively. From
the above equations we deduce the following relationship between the price
at time t and that at time t − 1:
D(pt ) = S(pt−1 )
5
or
D(pt+1 ) − S(pt ) = 0
If the demand function D is invertible, with inverse D−1 , then pt+1 can be
expressed explicitly in terms of pt :
c d−b
pt+1 = pt +
a a
In mathematics, it is customary to place the independent variable in brackets
after the dependent variable: p(t) instead of pt .
Yt = Q(Kt , Lt )
Kt+1 − Kt = sYt
Lt+1 − Lt = nLt
Here, Y , K and L denote production, capital and labour, respectively. The
rate of growth n of the labour force is a positive constant and the propensity
to save s is a number between 0 and 1. Q is a production function, assumed
homogeneous of degree 1. We can write the model in per capita form by
defining k := KL
, (capital per worker), and
Q(K, L) K
q(k) = = Q( , 1) = Q(k, 1)
L L
From the above equations we infer that k must satisfy the first order difference
equation
s 1
kt+1 = q(kt ) + kt (1.1)
1+n 1+n
6
Technical progress can be introduced into the model by replacing the first
equation with
Yt = Q(Kt , Et Lt )
Here, E is a parameter, representing the increase in labour efficiency due to
technological improvement. E is an increasing function of t, and E0 = 1.
K
Defining k := EL and
Q(K, EL) K
q(k) = = Q( , 1)
EL EL
we find that k satisfies the difference equation
Et s 1
kt+1 = q(kt ) + kt (1.2)
Et+1 1+n 1+n
Et+1 Et+1
If Et
is constant: Et
= 1 + g, with g > 0, we obtain
s 1
kt+1 = q(kt ) + kt (1.3)
(1 + n)(1 + g) (1 + n)(1 + g)
7
A difference equation is usually called linear if F is an affine function
of the last k + 1 variables, so a linear equation has the form
Then we get
y(x + h) − y(x)
=3 (1.5)
h
is transformed into
u(t + 1) − u(t)
=3
h
by the change of variables
x
x 7→ t := , y(x) 7→ u(t) := y(x) = y(th)
h
All real-valued solutions of this equation have the form
8
The difference quotient on the left-hand side of (1.5) can be used to approx-
dy
imate dx . More generally, solutions of the difference equation
y(x + h) − y(x)
= f (x, y(x))
h
are used to approximate solutions of the differential equation
dy
= f (x, y(x))
dx
The simplest numerical method to “compute” solutions of differential equa-
tions, Euler’s method, is based on this idea.
From now on, we take the shift equal to 1 and work with a discrete variable
n ∈ Z. A solution of the difference equation
to indicate that the solutions we seek are defined on D. The following exam-
ple demonstrates the importance of this addition.
consists of all functions y on N with the property that y(n) = 0 for all n ≥ 1.
y(0) can take on every real (or complex) value. However, the equation
9
Figure 1.2: 4-periodic solution of the equation y(n + 1) = 3.5y(n)(1 − y(n)),
n∈N
has no equilibrium solutions. Any 2-periodic solution must satisfy the system
consisting of the above equation and the equation y(n + 2) = y(n). For every
10
solution of the first equation we have
y(n + 2) = −(n + 1)y(n + 1) + n(−1)n+1 + 2(n + 2)
= −(n + 1)(−ny(n) + (n − 1)(−1)n + 2(n + 1)) − n(−1)n + 2(n + 2)
= n(n + 1)y(n) − (n2 + n − 1)((−1)n + 2)
Alternative notations for (1.6) that can be found in the literature are
and
F (n, y, τ y, ..., τ k y) = 0 (1.9)
11
A special case of (1.6) is the so-called recurrence relation. This is an equation
of the form
y(n + k) = f (n, y(n), y(n + 1), ..., y(n + k − 1)), n ∈ {n0 , n0 + 1, ...} (1.10)
y(n + 1) = y(n)2 − a, n ∈ N
12
and non-real when a < − 14 :
1 1 √
y1,2 = ± i −1 − 4a
2 2
If, in the above example, we restrict ourselves to real solutions, then the
equation has no equilibrium solutions for a < − 41 , a single equilibrium solu-
tion y ≡ 12 for a = − 14 and two equilibrium solutions for a > − 41 . At the
parameter-value a = − 41 , an abrupt change in the properties of the equation
occurs. Such qualitative changes, due to a change in a parameter are called
bifurcations. The value of the parameter(s) at which the bifurcation occurs
is called the bifurcation value of that parameter. There are many different
types of bifurcations, only a few of which will be reviewed here. The bifur-
cation in example 1.1.7 is a so-called ‘saddle-node bifurcation’. Bifurcations
can be represented in a bifurcation diagram, where certain characteristic
properties of the equation, such as the equilibrium values, are plotted versus
a parameter.
13
solution. A solution of the equation is said to be stable, if a small change
(‘perturbation’) of the initial values causes a (sufficiently) small change in
the entire solution, and is said to be asymptotically stable if, in the long
run, the effect of this change is no longer noticeable. Here are the precise
definitions:
Definition 1.1.8. A solution y of equation (1.10) is stable if, for every
positive number , there exists a positive number δ, such that every solution
ỹ, whose first k values do not differ more than δ from those of y, i.e.
|ỹ(n0 + j) − y(n0 + j)| ≤ δ for j = 0, ..., k − 1,
has the property that, for all n ≥ n0 ,
|ỹ(n) − y(n)| ≤
A stable solution y is called asymptotically stable or locally asymptoti-
cally stable, if there exists a positive number δ, such that, for every solution
ỹ with the property that
|ỹ(n0 + j) − y(n0 + j)| ≤ δ for j = 0, ..., k − 1
the following holds
lim (ỹ(n) − y(n)) = 0
n→∞
A stable solution y is globally asymptotically stable if
lim (ỹ(n) − y(n)) = 0
n→∞
for every solution ỹ. A stable solution that is not asymptotically stable, is
called neutrally stable. A solution of (1.10) that is not stable, is called
unstable.
If one is exclusively interested in real-valued solutions, then one only needs
to consider the effect of real-valued changes in the initial values, in order to
determine the stability of a solution. Cf. fig. 1.4 in §1.2. There, the null
solution (like all other solutions with initial value in the interval (−1, 2)) is
locally, but not globally asymptotically stable, whereas the solution y ≡ −1
is unstable.
Stability and instability are properties of solutions of equations and not
of individual sequences. Any particular sequence may be a stable solution of
one difference equation and an unstable solution of another equation, as is
demonstrated by the following example.
14
Example 1.1.9. The null sequence is a globally asymptotically stable solu-
tion of the equation
1
y(n + 1) − y(n) = 0, n ∈ N,
n+1
since, for every solution of this equation, the following relation holds
y(0)
y(n) =
n!
Hence limn→∞ y(n) = 0. On the other hand, the null sequence is an unstable
solution of
y(n + 1) − (n + 1)y(n) = 0, n ∈ N,
as, in this case, y(n) = n!y(0) for every solution of the equation, and thus
limn→∞ |y(n)| = ∞ when y(0) 6= 0. Finally, the null sequence is a neutrally
stable solution of the equation
(Why?).
15
number δ, such that limn→∞ y(n) = 0 for every solution y of (1.11), with the
property that |y(0)| ≤ δ. Let δ < 1. Another inductive argument shows that
|y(n)| ≤ δ n+1 for all n ∈ N, if |y(0)| ≤ δ. Suppose that |y(n)| ≤ δ n+1 for a
certain n ∈ N. Then
|y(n + 1)| = |y(n)|2 ≤ δ 2n+2 ≤ δ n+2
So the inequality is valid for n + 1 as well, and thus it is valid for all n ∈ N.
As limn→∞ δ n = 0, this implies that limn→∞ y(n) = 0.
The equation (1.11) has a second equilibrium solution, viz. the sequence
y1 (n) = 1 for all n ∈ N. This solution is unstable. This can be proved in
the following way. Let δ > 0 and consider the solution yδ with initial value
yδ (0) = 1 + δ. Thus, |yδ (0) − y1 (0)| = δ. A simple inductive argument shows
that
|yδ (n)| ≥ (1 + δ)n+1
for all n ∈ N. Hence it follows that limn→∞ yδ (n) = ∞. This implies that
|yδ (n) − y1 (n)| = |yδ (n) − 1| can take on arbitrarily large values. Since this
holds for any δ > 0, the condition for stability of the solution y1 is not
fulfilled. Note that the existence of a second equilibrium solution implies
that the null solution is not globally asymptotically stable!
Sometimes, by a change of variables, a difference equation can be simplified,
or reduced to an equation whose properties are already known. Thus, for
example, the change of variables: n = m + n0 , y(n) = y(m + n0 ) = z(m),
transforms the equation
F (n, y(n), ..., y(n + k)) = 0, n ∈ {n0 , n0 + 1, ...}
into
G(m, z(m), ..., z(m+k)) := F (m+n0 , y(m+n0 ), ..., y(m+n0 +k)) = 0, m ∈ N,
which is of the same type as the original equation, but now the independent
variable runs from 0 to ∞. This is why, from now on, we will usually take
n0 = 0. Another useful transformation is the following. Suppose that a
specific solution y∗ of equation (1.6) is known. Now substitute y = y∗ + z
into the equation. Then y is a solution of (1.6) iff the new dependent variable
z satisfies the equation
G(n, z(n), ..., z(n + k)) := F (n, z(n) + y∗ (n), ..., z(n + k) + y∗ (n + k)) = 0
A particular property of the new equation is that it possesses a null solution.
16
Lemma 1.1.11. The sequence y∗ is an asymptotically stable, neutrally sta-
ble, or unstable solution of the equation
y(n + k) = f (n, y(n), y(n + 1), ..., y(n + k − 1)), n∈N (1.12)
iff the null sequence is, respectively, an asymptotically stable, neutrally sta-
ble, or unstable solution of the equation
z(n + k) = g(n, z(n), z(n + 1), ..., z(n + k − 1)) :=
f (n, z(n) + y∗ (n), ..., z(n + k − 1) + y∗ (n + k − 1)) − y∗ (n + k), n∈N
(1.13)
|ỹ(n) − y∗ (n)| ≤
17
1.1.2 Exercises
1. Show that the solutions of equation (1.11) do not form a linear space.
2. Find all real, periodic solutions of the equation
y(n + 1) = y(n)2 , n∈N
18
1.2 Autonomous first order difference equa-
tions
The most general first order difference equation of the type (1.10), with
n0 = 0, is
y(n + 1) = f (n, y(n)), n ∈ N
We restrict ourselves in this section to autonomous equations. These are of
the form
y(n + 1) = f (y(n)), n ∈ N (1.15)
(1.11) is an example of such an equation. Another example is the logistic
difference equation:
Thus,
f 1 (x) = f (x), f 2 (x) = f (f (x)), etc.
If y is a solution of (1.15), then, for every k ∈ N,
y(n + k) = f k (y(n))
19
Example 1.2.1. f (x) = x2 (cf. (1.11)).
k
f k (x) = x2 for every k ∈ N. f has two fixed points: 0 and 1, both of which
are real. f 2 has four fixed points: the solutions of the 4th degree equation
x4 − x = 0
Besides the two (real) fixed points of f , this equation has two additional,
complex conjugated, non-real solutions:
2π 1 1 √
x1,2 = e± 3 i = − ± i 3
2 2
So, in this case, the difference equation (1.15) has two complex-valued peri-
odic solutions of period 2.
The positive orbit γ + (x0 ) of x0 ∈ Df subject to the map f is the set
γ + (x0 ) := {f n (x0 ), n ∈ N}
The positive (or forward) orbit of x0 , subject to the map f , is the set of all
elements of the sequence y with initial value x0 , that satisfies (1.15). We
will also call this the (positive) orbit of y. Every positive orbit is positive
invariant. The positive orbit of an equilibrium solution consists of one point,
and is also called an equilibrium point of the equation. The positive
orbit of a p-periodic solution is a set consisting of p elements, called a p-
periodic orbit. Different solutions of (1.15) may have the same positive orbit,
for instance, the positive orbit of both complex 2-periodic solutions of the
equation y(n + 1) = y(n)2 is the set {x1 , x2 } (cf. example 1.2.1). It is easily
seen that, in general, to each p-periodic orbit correspond exactly p p-periodic
solutions having this orbit (check this).
Example 1.2.2. In the case of the logistic difference equation, f (x) =
αx(1 − x) and
f 2 (x) = α2 x(1 − x)(1 − αx + αx2 )
In order to find the real-valued, periodic solutions of (1.16) with minimal
period 2, we begin by computing all real fixed points of f 2 , i.e. the solutions
of the equation
α2 x(1 − x)(1 − αx + αx2 ) = x
After eliminating the fixed points of f (cf. §1.1.2, exerc. 4), we find that the
remaining fixed points of f 2 , if there are any, should satisfy the equation
α2 x2 − α(α + 1)x + α + 1 = 0
20
If α > 3, this equation has two distinct, real-valued solutions x1 and x2 :
q
α+1± (α + 1)(α − 3)
x1,2 =
2α
Check that f (x1 ) = x2 and f (x2 ) = x1 . So, if α > 3, (1.16) has two real,
periodic solutions of minimal period 2, viz. the sequences x1 , x2 , x1 , x2 , ....
and x2 , x1 , x2 , x1 ..... Both have the same orbit: γ + (x1 ) = γ + (x2 ) = {x1 , x2 }
If {f n (x0 )}∞ +
n=0 is a periodic solution of (1.15), then ω(x0 ) = γ (x0 ). If
n
limn→∞ f (x0 ) = c, then ω(x0 ) = {c}, but the converse is not true. If
ω(x0 ) = {c} and γ + (x0 ) is unbounded, then limn→∞ f n (x0 ) does not exist
(cf. ex. 1.2.3; in such cases, it is sometimes said that ∞ is a positive limit
point of x0 ).
21
Figure 1.3: graphical determination of the first 8 elements of the solution of
the logistic difference equation (2.2), with α = 2.8, y(0) = 0.1.
The stable set of a fixed point c of f is the set of all x ∈ Df with the
property that limn→∞ f n (x) = c.
A closed, invariant set A is called an attracting set for f , if there exists
a neighbourhood U of A, such that the distance between f n (x) and A (i.e.
miny∈A |f n (x) − y|) tends to 0 as n → ∞, for all x ∈ U .
The basin of attraction of an attracting set A is the set of all x ∈ Df
such that the distance between f n (x) and A tends to 0 as n → ∞.
An attractor is an attracting set A, containing a “dense orbit”, i.e. there
exists an x ∈ A such that A = γ + (x).
The simplest example of an attractor is an asymptotically stable equi-
librium point c. If A = {c}, then A is closed and invariant (as f (c) = c).
The distance between f n (x) and A is |f n (x) − c|. By definition, there exists
a neighbourhood U of c, hence of A, such that limn→∞ f n (x) = c for all
x ∈ U . Moreover, γ + (c) = {c}, so A = γ + (c). The basin of attraction of {c}
coincides with its stable set.
Example 1.2.4. The basin of attraction of the fixed point 0, of the logistic
map
f (x) = αx(1 − x)
where α ∈ (0, 1), is the interval (1 − 1/α, 1/α) (cf. fig. 1.4).
Proof: Let x ∈ R and y(n) := f n (x), n ∈ N. We distinguish several cases.
22
Figure 1.4: solutions of the logistic difference equation, with α = 1/2, with
different initial values, varying from −2.2 to 2.5. Observe the asymptotic
behaviour of the different solutions. The basin of attraction of 0 is the interval
(−1, 2) (on the y-axis).
1. y(0) (= x) ≥ 1/α. This implies that αy(0) ≥ 1 and 1−y(0) ≤ 1−1/α < 0,
hence y(1) = αy(0)(1 − y(0)) ≤ 1 − 1/α. From here on, proceed as in case 2
below (starting from y(1) instead of y(0)).
2. y(0) ≤ 1 − 1/α. As an inductive argument shows, this implies that y(n) ≤
1−1/α for all n ∈ N. For, suppose that for some n ∈ N, y(n) ≤ 1−1/α (< 0),
then we have α(1−y(n)) ≥ 1, so y(n+1) = α(1−y(n))y(n) ≤ y(n) ≤ 1−1/α.
From 1. and 2. it follows that none of the points examined so far belongs to
the basin of attraction of 0, as in both cases y(n) ≤ 1 − 1/α for all n ≥ 1,
hence limn→∞ y(n) 6= 0. Thus, the basin of attraction of 0 must be contained
in the interval (1 − 1/α, 1/α).
3. 0 ≤ y(0) ≤ 1. If 0 ≤ y(n) ≤ 1 for some n ∈ N, then 0 ≤ 1−y(n) ≤ 1, and,
consequently, 0 ≤ y(n + 1) = αy(n)(1 − y(n)) ≤ αy(n). With the principle
of induction it follows that
0 ≤ y(n) ≤ αn y(0)
for all n ∈ N. As 0 < α < 1, limn→∞ αn = 0. With the aid of the “inclusion
theorem” we conclude that limn→∞ y(n) = 0 for all sequences y such that
0 ≤ y(0) ≤ 1. This implies that the interval [0, 1] is contained in the basin
23
of attraction of 0.
4. 1 < y(0) < 1/α. Then 1 − 1/α < 1 − y(0) < 0 and 0 < αy(0) < 1, hence
1 − 1/α < y(1) < 0. From here on, proceed as in case 5.
5. If 1 − 1/α < y(0) < 0, an inductive argument shows that 1 − 1/α <
y(n) < y(n + 1) < 0 for all n ∈ N. For, suppose that 1 − 1/α < y(n) < 0
for some n ∈ N. Then we have α(1 − y(n)) < 1, hence y(n) < y(n + 1) =
αy(n)(1 − y(n)) < 0. The sequence y is monotone increasing and bounded
above, hence it has a limit l ∈ (1 − 1/α, 0]. Letting n tend to ∞ on both
sides of the equation
y(n + 1) = αy(n)(1 − y(n))
we find that l = αl(1 − l). In the interval (1 − 1/α, 0] the latter equation has
the unique solution l = 0.
From 3., 4. and 5. we infer that limn→∞ y(n) = 0 for all sequences y with
1 − 1/α < y(0) < 1/α. Hence the interval (1 − 1/α, 1/α) is contained in the
basin of attraction of 0. As we have also proved that the basin of attraction of
0 is contained in the interval (1 − 1/α, 1/α), it must be equal to this interval.
24
Hence
f (x) − f (c)
| | < |f 0 (c)| + ρ = d
x−c
Hence it follows that
|f n (x) − c| ≤ dn |x − c|
for all x ∈ U − {c}. We will prove the second statement of the theorem by
contradiction. Suppose that y ≡ c is a stable solution. Then there exists a
neighbourhood V of c, of the form |x − c| < δ 0 , with δ 0 > 0, such that any
solution ỹ with initial value y0 ∈ V “stays within U ”, i.e.
|ỹ(n) − c| < δ
25
for all n ∈ N. Then we have, for all n ∈ N,
This equation has equilibrium solutions z ≡ 0 and z ≡ 1/2. From the facts
that f 0 (0) = 0 and f 0 ( 21 ) = 2 we deduce that the first is asymptotically stable,
whereas the second is unstable. The corresponding solutions of the original
equation are the sequences y(n) = n and y(n) = n + 21 . So, by lemma 1.1.11,
these are asymptotically stable and unstable, respectively.
Hence it follows that, if i = 1, for every solution y with initial value y(0) ∈
(−1, 1) − {0}, |y(n + 1)| < |y(n)|. Consequently, if ∈ (0, 1), then |y(0)| <
26
implies |y(n)| < for all n ∈ N, so y ≡ 0 is a stable solution of the equation
(here we can choose ‘δ 0 = ). Moreover, if y(0) ∈ (−1, 1) − {0}, the sequence
|y(n)| is monotone decreasing and thus it converges to a limit. Since this
limit must be a fixed point of |f1 |, it has to be 0 (check this). Thus we
find that all (real) solutions with initial value in (−1, 1) converge to 0, hence
the null solution of the equation y(n + 1) = y(n) − y(n)3 is asymptotically
stable. The null solution of the equation y(n + 1) = y(n) + y(n)3 , however, is
unstable. This can be seen as follows. For every solution y of the equation,
we have y(1) = y(0)(1 + y(0)2 ). By means of an inductive argument, one
easily proves that, for all n ∈ N,
This theorem can be deduced from Theorem 1.2.5 as follows. The condition
limx→0 g(x)
x
= 0 implies that limx→0 g(x) = 0, and thus, in view of the con-
tinuity of g, g(0) = 0. Take c = 0 and define f by f (x) = g(x) + ax. Then
f (0) = 0. g is differentiable at 0, with derivative
27
Here, g(x) = −αx2 and
|g(x)|
lim = lim |αx| = 0
x→0 |x| x→0
Application of Theorem 1.2.8 reproduces the known fact that the null solution
of the equation is asymptotically stable for α ∈ (0, 1) and unstable for α > 1.
Theorems 1.2.5 and 1.2.8 are based on a method that is frequently used in the
study of nonlinear difference equations, known as linearization. It consists
in approximating solutions of the nonlinear equation by solutions of a linear
equation. In the case of Theorem 1.2.8 that linear equation is
Example 1.2.10. The linear approximation of f (x) = 2x2 at the fixed point
x = 1/2 is
28
Hence, linearization of the equation
y(n + 1) = 2y(n)2
at 1/2 yields the inhomogeneous, linear difference equation
y(n + 1) = 1/2 + 2(y(n) − 1/2)
29
Theorem 1.2.11. Let c ∈ C be a fixed point of f p for some positive integer
p and suppose that f p is differentiable at c.
(i) If f is continuous at f k−1 (c) for k = 1, ..., p−1, and |(f p )0 (c)| < 1, then the
(periodic) solution y of (1.15) with initial value c is (locally) asymptotically
stable.
(ii) If |(f p )0 (c)| > 1, then the (periodic) solution y of (1.15) with initial value
c is unstable.
From the fact that x1 and x2 are solutions of the quadratic equation α2 x2 −
α(α + 1)x + α + 1 = 0, we deduce that
(f 2 )0 (x1 ) = −α2 + 2α + 4
√ 1 to −∞
The function on the right-hand side decreases monotonically from
as α increases from 3 to ∞, and it equals −1 when α = 1 + 6. Thus,
according to Theorem 1.2.11,
√ both 2-periodic solutions are
√ asymptotically
stable when 3 < α < 1 + 6 and unstable when α > 1 + 6.
30
Lemma 1.2.15. If f is a continuous map, then the orbit of any asymptoti-
cally stable periodic solution of (1.15) is a (periodic) attractor.
Proof. We give the proof for the case of an asymptotically stable, 2-periodic
solution. Let c be the initial value of the solution and A := γ + (c) =
{c, f (c)}. A being a finite set, it is closed, and it is also invariant, as
f (A) = {f (c), f 2 (c)} = {f (c), c} = A. Due to the asymptotic stability
of the solution, there exists a neighbourhood U1 of c, of the form |x − c| < δ1 ,
with δ1 > 0, such that limn→∞ f n (x)−f n (c) = 0 for all x ∈ U1 . As f n (c) ∈ A
for all n ∈ N, this implies that the distance of f n (x) to A tends to 0 for all
x ∈ U1 . It remains to prove the existence of a neighbourhood U2 of f (c), of
the form |x − f (c)| < δ2 , with δ2 > 0, such that limn→∞ f n (x) − f n (f (c)) = 0
for all x ∈ U2 . Due to the continuity of f at f (c), there exists a positive
number δ2 , such that |f (x) − f (f (c))| < δ1 for all x with the property that
|x−f (c)| < δ2 . Consequently, f (x) ∈ U1 and thus limn→∞ f n (f (x))−f n (c) =
limn→∞ f n+1 (x)−f n (f 2 (c)) = limn→∞ f n+1 (x)−f n+1 (f (c)) = 0 for these val-
ues of x. Hence there exists a neighbourhood U := U1 ∪ U2 of A such that,
for all x ∈ U , the distance of f n (x) to A tends to 0 as n → ∞.
We conclude this section with two examples of very simple economic models
that give rise to nonlinear, autonomous difference equations.
or, equivalently,
Xt+1 = (1 + βK)Xt − βXt2 ,
which is a variant of the logistic difference equation. Here, Xt denotes the
inventory level in the period t, K is the desired or maximal stock and β is a
positive constant.
qtd = D(pt )
qts = S(pet )
31
qtd = qts
pet = pet−1 + w(pt−1 − pet−1 ), 0 < w < 1
Here, qtd , qts , pt and pet denote the demand, supply, price and expected price
in the period t, respectively. We assume the demand to be a strictly de-
creasing function of the price, hence its inverse exists. Then pet satisfies the
autonomous, first order difference equation
1.2.2 Bifurcations
In this section we briefly discuss some examples of frequently occurring bi-
furcations. We already encountered one type of bifurcation in §1.2, in the
1-parameter-family of autonomous, first order difference equations
y(n + 1) = y(n)2 − a
32
Figure 1.5: Transcritical bifurcation in (2.7) at a = 1
33
Figure 1.7: Bifurcation diagram of the logistic difference equation.
where a is a real parameter. If a < 1, 0 is the unique real fixed point of f (x) =
ax − x3 . f 0 (0) = a, so the null solution of (1.22) is asymptotically stable
when −1 < a < 1 and unstable when a > 1 (the case a = 1 was discussed
in example 1.2.7).√ If a > 1, there are two additional
√ real-valued equilibrium
solutions: y ≡ ± a − 1. From the fact that f 0 (± a − 1) = 3−2a we deduce
that both solutions are asymptotically stable when 1 < a < 2. At a = 1 a
so-called pitchfork bifurcation occurs (cf. fig. 1.6).
To illustrate a fourth type of bifurcation, we consider once more the
logistic difference equation (cf. the examples 1.2.2 and 1.2.4). The function
f (x) = αx(1 − x) has two fixed points: 0 and 1 − 1/α, when α 6= 1. f 0 (0) = α
and f 0 (1 − 1/α) = 2 − α. According to Theorem 1.2.5, the null solution of
(1.16) is asymptotically stable when 0 < α < 1 and unstable when α > 1.
The second equilibrium solution is asymptotically√stable when 1 < α < 3
and unstable when α > 3. When 3 < α < 1 + 6, the equation has two
asymptotically stable periodic solutions. At the parameter values α = 1 and
α = 3, the properties of the equation change significantly, in both cases it
undergoes a bifurcation. The bifurcation at α = 1 is of the same type as
in (1.21), this is a transcritical bifurcation. At α = 3 a so-called period-
doubling or flip-bifurcation occurs: an asymptotically stable equilibrium
solution gives way to two asymptotically stable 2-periodic solutions (having
the same orbit).
The above examples all concern bifurcations of equilibrium solutions.
Similar bifurcations occur in periodic solutions of period greater than 1. Bi-
furcations can be represented in a bifurcation diagram. Usually, the values
of the asymptotically stable equilibrium (or periodic) solutions are plotted
versus the parameter. Sometimes, unstable periodic solutions are indicated
by means of dotted lines. In order to make bifurcation diagrams, like the
ones shown in figs. 1.5 through 1.7, on a computer, one roughly proceeds as
follows. For “each” value of the parameter, first, a certain number of val-
ues of the solution with a given initial value x0 , the so-called “pre-iterates”)
is computed. Next, a number of subsequent entries of the same sequence
34
Figure 1.8: The solutions of (2.2) with α = 2.8, y(0) = 0.1 and y(0) = 0.2.
35
Figure 1.9: The solutions of (2.2) with α = 3.5, y(0) = 0.1 and y(0) = 0.2.
(Also compare this figure to Fig. 1.2.)
Figure 1.10: The solutions of (2.2) with α = 4, y(0) = 0.1 and y(0) = 0.11.
36
1.2.3 Exercises
1. Let f : R → R and c ∈ R and suppose that f is continuous at f k−1 (c)
for k = 1, ..., p. Use an inductive argument to prove that f k is continuous
at c for k = 1, ..., p.
√ √
2. Prove that the interval (− 2, 2) is the basin of attraction of the fixed
point 0 of the map f (x) = x − x3 , x ∈ R.
37
1.3 Linear difference equations
The difference equation (1.9) is linear if and only if F is an affine function of
y, τ y, ..., τ k y, thus a linear difference equation has the form
F (n, y, τ y, ..., τ k y) = b + a0 y + a1 τ y + ... + ak τ k y = 0 (1.23)
Here, a0 , a1 , ..., ak and b are functions of n, i.e. sequences. For every n ∈ N
(or Z) we have
b(n) + a0 (n)y(n) + a1 (n)y(n + 1) + ... + ak (n)y(n + k) = 0 (1.24)
We can write the equation in the more compact form
k
aj τ j y = −b
X
(1.25)
j=0
38
Example 1.3.1. The solution space of the homogeneous, linear difference
equation
((−1)n − 1)y(n + 1) + y(n) = 0, n ∈ N
consists of the unique solution of this equation: the null solution, hence its
dimension is 0. On the other hand, the equation
From now on we assume that the sequence ak has no zeroes. Dividing each
term in (1.24) by ak (n), we obtain the equivalent equation
b(n) a0 (n) a1 (n) ak (n)
+ y(n) + y(n + 1) + ... + y(n + k) = 0
ak (n) ak (n) ak (n) ak (n)
It has the same form as (1.24), but now the coefficient of y(n + k) equals 1
for all n. Henceforth we consider linear difference equations of the form
In fact, this equation is a special case of a recurrence relation (to see this, take
all terms except y(n + k) to the right-hand side). In particular, any solution
is determined by its first k values y(0), ..., y(k − 1) (the initial values), which
can be chosen arbitrarily. We begin by studying the homogeneous equation
If we take the first k entries of the sequence to be zero, then all subsequent
entries will be zero as well and we end up with the null solution.
39
Proof. The k initial values y(0), ..., y(k − 1) form a k-dimensional vector.
If l ≤ k, we can choose l linearly independent such k-dimensional vectors
(y1 (0), ..., y1 (k−1)),...,(yl (0), ..., yl (k−1)) and these will determine l solutions
y1 ... yl of (1.28). These solutions are linearly independent. For, suppose
there exist numbers c1 , c2 , ..., cl with the property that c1 y1 (n) + c2 y2 (n) +
... + cl yl (n) = 0 for all n ∈ N, then, of course,
y1 (0) y2 (0) yl (0) 0
. . . .
c1
. + c2
. + ... + cl
.
=
.
(1.29)
.
.
.
.
y1 (k−1) y2 (k−1) yl (k−1) 0
The linear independence of the vectors on the left-hand side of (1.29) implies
that c1 = c2 = ... = cl = 0. Hence the equation (1.28) has at least k
linearly independent solutions. Conversely, for any number l of linearly inde-
pendent solutions y1 , ..., yl of (1.28), the vectors (y1 (0), ..., y1 (k − 1)) through
(yl (0), ..., yl (k − 1)) are linearly independent. For, suppose there exist num-
bers c1 , c2 , ..., cl such that (1.29) holds. Define a sequence y by
y := c1 y1 + c2 y2 + ... + cl yl
Then y is a solution of (1.28) with the property that y(n) = 0 for 0 ≤ n ≤
k − 1 and thus y ≡ 0, i.e.
c1 y1 (n) + c2 y2 (n) + ... + cl yl (n) = 0 for all n ∈ N
The linear independence of the solutions y1 , ..., yl now implies that c1 = c2 =
... = cl = 0, and, consequently, the vectors
(y1 (0), ..., y1 (k − 1)), ..., (yl (0), ..., yl (k − 1)) are linearly independent. Since
there can be at most k linearly independent k-dimensional vectors, we con-
clude that the dimension of the solution space of (1.28) is at most k and
thus is exactly k.
40
It is of great importance to have a basis of the solution space of (1.28) at
our disposal. In that case, any solution of (1.28) can be written as a linear
combination of the basis elements. In view of lemma 1.3.3, it suffices to choose
k linearly independent vectors (y1 (0), ..., y1 (k − 1)),..., (yk (0), ..., yk (k − 1)).
To prove that k given solutions y1 , ..., yk of (1.28) constitute a basis of the
solution space, it suffices to prove the linear independence of the vectors
(y1 (0), ..., y1 (k − 1)),..., (yk (0), ..., yk (k − 1)). This is the case iff the matrix
having these vectors for its column vectors is nonsingular, i.e. iff
y1 (0) y2 (0) . . yk (0)
y1 (1) y2 (1) . . yk (1)
. . . . .
6= 0
. . . . .
. . . . .
y1 (k − 1) y2 (k − 1) . . yk (k − 1)
The above determinant is called the Casorati determinant. For all non-
negative integers n, Cy1 ,...,yk (n) is defined by
y1 (n) y2 (n) . . yk (n)
y1 (n + 1) y2 (n + 1) . . yk (n + 1)
. . . . .
Cy1 ,...,yk (n) =
. . . . .
. . . . .
y1 (n + k − 1) y2 (n + k − 1)
. . yk (n + k − 1)
Proof. We begin by proving the lemma for the case that k = 2. For every
nonnegative integer n we have
y (n + 1) y2 (n + 1)
Cy1 ,y2 (n + 1) = 1
y1 (n + 2) y2 (n + 2)
Furthermore,
41
Inserting this into the last row of the determinant, we get
y1 (n + 1) y2 (n + 1)
−a0 (n)y1 (n) − a1 (n)y1 (n + 1) −a0 (n)y2 (n) − a1 (n)y2 (n + 1)
The last row being the sum of the two row vectors (−a0 (n)y1 (n) −a0 (n)y2 (n))
and (−a1 (n)y1 (n + 1) − a1 (n)y2 (n + 1)), and the determinant of a matrix
being a linear function of its row vectors, this is equal to
y1 (n + 1) y2 (n + 1) y1 (n + 1) y2 (n + 1)
−a0 (n) − a1 (n)
y1 (n) y2 (n) y1 (n + 1) y2 (n + 1)
The second term vanishes for all n. Interchanging rows in the first determi-
nant, we find
y1 (n) y2 (n)
Cy1 ,y2 (n + 1) = a0 (n)
= a0 (n)Cy1 ,y2 (n)
y1 (n + 1) y2 (n + 1)
42
The determinants on the right-hand side are seen to vanish, except for l = 0,
hence
y1 (n + 1) y2 (n + 1) . . yk (n + 1)
y1 (n + 2) y2 (n + 2) . . yk (n + 2)
. . . . .
Cy1 ,...,yk (n + 1) = −a0 (n)
. . . . .
. . . . .
y1 (n) y2 (n) . . yk (n)
Moreover,
0 1
Cy1 ,y2 (0) = = −1
1 0
Thus, y1 and y2 constitute a basis of the solution space of the equation. The
Casorati-determinant Cy1 ,y2 (n) satisfies the first order difference equation
hence Cy1 ,y2 (n) = −1 for all n ∈ N. (Check this by computing Cy1 ,y2 (n)
directly.)
43
1.3.1 Inhomogeneous linear difference equations
Next, we consider the inhomogeneous difference equation
k
X
Ly(n) := aj (n)y(n + j) = b(n), n ∈ N (1.31)
j=0
where b does not vanish identically (is not a null sequence). As the homo-
geneous equation Ly = 0 has other solutions besides the null solution, the
linear operator L is not invertible and equation (1.31) does not have a
unique solution. If y0 is a solution, i.e. if Ly0 = b and if ỹ is an arbitrary
solution of the homogeneous equation, so Lỹ = 0, then y0 + ỹ is another
solution of (1.31), as
L(y0 + ỹ) = Ly0 + Lỹ = b + 0 = b
Conversely, the difference of two solutions y1 and y2 of the inhomogeneous
equation (1.31) is a solution of the homogeneous equation, as
L(y1 − y2 ) = Ly1 − Ly2 = b − b = 0
Theorem 1.3.7. Let y0 be a particular solution of the inhomogeneous
linear difference equation (1.31). Every solution of (1.31) can be written
as the sum of y0 and a solution of the homogeneous equation. Conversely,
any sequence that can be written as the sum of y0 and a solution of the
homogeneous equation is a solution of (1.31).
44
Example 1.3.8. The terms of a geometric progression with common ra-
tio r ∈ R or C satisfy the equation
Here a(n) equals the common ratio r for all n, and thus is a constant sequence.
For any solution y we have
y(n) = crn , n ∈ N, c ∈ C
y(n) = cn!, n ∈ N, c ∈ C
45
and summing the expressions on both sides of the equality signs, we get
y(n) − y(0) = b(0) + b(1) + ... + b(n − 1)
As y(0) can be chosen arbitrarily, the general solution of (1.33) is
n−1
X
y(0) = c, y(n) = c + b(m) for n ≥ 1, c∈C
m=0
46
Example 1.3.10. The general solution of the inhomogeneous, first order,
linear difference equation
is
n−1
X 1
y(0) = c, y(n) = cn! + n! for n ≥ 1, c∈C
m=0 (m + 1)!
In some special cases, like in the following example, a particular solution can
be easily found.
y(n + 1) − 2y(n) = 3
has a constant (or equilibrium) solution y ≡ −3. The general solution of the
corresponding homogeneous equation is y(n) = c2n , with c ∈ C (cf. example
1.3.8). The general solution of the inhomogeneous equation is therefore
47
Here, r and v are constants: the common ratio and difference of the progres-
sion, respectively. The homogeneous equation has a solution yn0 := rn . Thus,
the general solution of the inhomogeneous equation is
n−1
y0 = c, yn = crn + rn
X
v for n ≥ 1, c ∈ C
m=0
yn = (y0 + nv)rn , n ∈ N
we have
|ỹ(n) − y(n)| ≤ for all n ∈ N
Now, z := ỹ − y is a solution of the homogeneous equation and every solution
of the homogeneous equation can be written as the difference of some solution
ỹ of (1.27) and y. Hence we can state the above condition as follows: for
every positive number there exists a positive number δ, such that, for every
solution z of the homogeneous equation, with the property that
we have
|z(n) − 0| ≤ for all n ∈ N
But this is precisely the condition for stability of the null solution of the ho-
mogeneous equation. This shows that a solution of a linear difference equa-
tion is stable if and only if the null solution of the homogeneous equation
is stable. It is easily verified that the same is true for asymptotic stabil-
ity. Consequently, the stability properties are the same for all solutions:
48
all solutions are neutrally stable, or all are asymptotically stable, or all are
unstable. Moreover, every asymptotically stable solution is globally asymp-
totically stable. For, suppose that y is an asymptotically stable solution of
(1.27). Then z ≡ 0 is an asymptotically stable solution of the homogeneous
equation. Hence there exists a positive number δ, such that limn→∞ z(n) = 0
for every solution z of the homogeneous equation, with the property that
|z(n)| ≤ δ for n = 0, ..., k − 1. Now consider an arbitrary solution z 6≡ 0
of the homogeneous equation and let µ := max{|z(0)|, ..., |z(k − 1)|}. Ob-
viously, µ > 0 and the sequence z̃ defined by z̃(n) = µδ z(n) also satisfies
the homogeneous equation. Moreover, |z̃(n)| ≤ δ for n = 0, ..., k − 1, hence
limn→∞ z̃(n) = 0, and this implies that limn→∞ z(n) = 0. Consequently, all
solutions of the homogeneous equation tend to 0 as n → ∞ and it imme-
diately follows that limn→∞ ỹ(n) − y(n) = 0 for any solution ỹ of (1.27).
Summarizing, we have the following theorem.
Theorem 1.3.14. All solutions of the linear difference equation (1.27) are
neutrally stable, globally asymptotically stable, or unstable, iff the
null solution of the homogeneous equation is neutrally stable, asymptoti-
cally stable, or unstable, respectively.
where b is an arbitrary sequence, are stable when |a| ≤ 1, as in that case every
solution y of the homogeneous equation has the property that |y(n) − 0| =
|an y(0) − 0| = |an y(0)| ≤ |y(0)| for all n ∈ N (cf. example 1.3.8). The
solutions of (1.36) are globally asymptotically stable when |a| < 1, as in that
case every solution y of the homogeneous equation, has the property that
which implies that the null solution of the homogeneous equation is asymp-
totically stable. The solutions of (1.36) are unstable when |a| > 1, as in that
case, for every solution y 6≡ 0, of the homogeneous equation, limn→∞ |y(n) −
0| = limn→∞ |y(n)| = ∞.
49
1.3.4 Exercises
1. Determine the solution of the following initial value problem
are unstable.
50
b. Determine the general solution of the above equation.
c. Determine the solution y with initial values y(0) = y(1) = 1 and
examine its stability.
51
The difference between (1.37) and (1.38) or (1.39) is mainly a matter of nota-
tion. In what follows we will not discriminate between systems of equations
and vectorial equations.
Example 1.4.1. The system of first order, linear difference equations
y1 (n + 1) = a11 (n)y1 (n) +...+ a1k (n)yk (n) +b1 (n)
y2 (n + 1) = a21 (n)y1 (n) +...+ a2k (n)yk (n) +b2 (n)
. . . .
(1.40)
. . . .
. . . .
yk (n + 1) = ak1 (n)y1 (n) +...+ akk (n)yk (n) +bk (n)
is equivalent to the linear vectorial difference equation.
y1 (n + 1) a11 (n) . . . a1k (n) y1 (n) b1 (n)
y
2 (n + 1)
a
21 (n) . . . a 2k (n)
2y (n) b2 (n)
. . . . . . . .
= + (1.41)
.
. . . . . . .
. . . . . . . .
yk (n + 1) ak1 (n) . . . akk (n) yk (n) bk (n)
or its more compact form
Here, y(n + 1), y(n) and b(n) are k-dimensional vectors and A(n) is the k × k
matrix in (1.41).
Systems of first order difference equations are important for more than one
reason. In the first place, systems of difference equations emerge naturally in
mathematical models of various problems, involving several time-dependent
variables. Secondly, any kth order, scalar difference equation can be easily
transformed into a system of k first order equations. Here is one way to do
that. As a starting point we take the kth order equation
y(n + k) = f (n, y(n), y(n + 1), ..., y(n + k − 1)), n∈N (1.43)
52
So
y1 (n) = y(n), y2 (n) = y(n + 1), ..., yk (n) = y(n + k − 1)
Then
y1 (n + 1) = y(n + 1) = y2 (n),
y2 (n + 1) = y(n + 2) = y3 (n),
........... . ............ . .....
........... . ............ . .....
yk−1 (n + 1) = y(n + k − 1) = yk (n)
If y is a solution of (1.43), we have
yk (n+1) = y(n+k) = f (n, y(n), y(n+1), ..., y(n+k−1)) = f (n, y1 (n), ..., yk (n))
Hence, the sequence of vectors or the vector function (y1 , ..., yk ), defined
by: (y1 , ..., yk )(n) = (y1 (n), ..., yk (n)), is a solution of the vectorial difference
equation
y1 (n + 1) y2 (n)
y2 (n + 1) .
. .
=
.
.
. yk (n)
yk (n + 1) f (n, y1 (n), ..., yk (n))
Example 1.4.2. Suppose that y is a solution of the kth order, linear differ-
ence equation
y(n + k) = −ak−1 (n)y(n + k − 1) − . . . − a0 (n)y(n) − b(n) (1.44)
Then the sequence of vectors or vector function (y1 , ..., yk ), defined by: (y1 , ..., yk )(n) =
(y1 (n), ..., yk (n)) = (y(n), y(n+1), ..., y(n+k−1)) is a solution of the vectorial
difference equation
y1 (n+1) 0 1 . . 0 y1 (n) 0
y2 (n+1) 0 0 1 . 0 y2 (n) 0
. . . . . . . .
= +
.
. . . . .
. .
. . . . . . . .
yk (n+1) −a0 (n) . . . −ak−1 (n) yk (n) −b(n)
(1.45)
Conversely: if (y1 , ..., yk ) is a solution of (1.45), then y1 is a solution of the
kth order scalar equation (1.44) and, for j = 1, ..., k, yj (n) = y1 (n + j − 1).
53
Example 1.4.3. The second order, linear difference equation
y1 (n + 1) = y(n + 1) = y2 (n)
y2 (n + 1) = y(n + 2) = (n + 3)y(n + 1) − (n + 1)y(n) + 3
= (n + 3)y2 (n) − (n + 1)y1 (n) + 3
y(n + 3) + ny(n) = 2n , n ∈ N
By the substitutions
y1 (n + 1) = y(n + 1) = y2 (n)
y2 (n + 1) = y(n + 2) = y3 (n) ,
n n
y3 (n + 1) = y(n + 3) = −ny(n) + 2 = −ny1 (n) + 2
y3 (n + 1) −n 0 0 y3 (n) 2n
54
Also in the context of systems of difference equations, the notions stability,
asymptotic stability and instability are used. They are defined analo-
gously to the case of a scalar difference equation. We give the definitions
for a system of first order difference equations. A solution (y1 , ..., yk ) of the
system (1.38) is called stable, iff, for every > 0 there exists a δ > 0, such
that, for every solution (ỹ1 , ..., ỹk ) with the property that
we have
As all norms on Ck are equivalent, the ∞-norm in the above definition can
be replaced with any other vector norm on Ck . (Cf. Simon & Blume, §29.4
and exercise 30.4. There, the ∞-norm is denoted by N0 .)
55
1.4.1 Homogeneous systems of first order linear differ-
ence equations
We consider a system of k first order, homogeneous, linear difference equa-
tions, in the form of the vectorial equation
Proof. The proof of the first statement is left to the reader. Let l ∈ N, let
y1 , ..., yl be solutions of (1.46) and c1 , ..., cl ∈ C. Then
c1 y1 (n) + ... + cl yl (n) = A(n − 1)...A(0)(c1 y1 (0) + ... + cl yl (0))
for all n ≥ 1. Hence it follows that c1 y1 (n) + ... + cl yl (n) = 0 for all n ∈
N iff c1 y1 (0) + ... + cl yl (0) = 0, and thus the solutions y1 , ..., yl of (1.46)
are linearly independent iff y1 (0), ..., yl (0) are linearly independent vectors
in Ck . Therefore, the dimension of the solution space equals that of Ck .
The solutions y1 , ..., yk constitute a basis of the solution space of (1.46) iff
y1 (0), ..., yk (0) are linearly independent vectors in Ck , or, equivalently, iff
det(y1 (0), ..., yk (0)) 6= 0. As
det(y1 (n), ..., yk (n)) = det A(n − 1)... det A(0) det(y1 (0), ..., yk (0))
56
for all n ≥ 1, det(y1 (n), ..., yk (n)) 6= 0 for at least one value of n iff
det(y1 (0), ..., yk (0)) 6= 0. This completes the proof of the theorem.
A k × k matrix function Y , having, for its column vectors, k linearly inde-
pendent solutions y 1 , ..., y k of the homogeneous vectorial difference equation
(1.46), is called a fundamental matrix of this equation. It is easily seen
that Y itself satisfies the equation, i.e.
Y (n + 1) = A(n)Y (n), n∈N
Moreover, Theorem 1.4.6 implies that det Y (n) 6= 0 for at least one value
of n. Conversely, the column vectors of a matrix function Y (n) with the
property that Y (n + 1) = A(n)Y (n) for all n ∈ N, also satisfy (1.46). If
det Y (n) 6= 0 for at least one value of n, then, according to Theorem 1.4.6,
these column vectors form a basis of the solution space of (1.46), hence Y
is a fundamental matrix of this equation. If Y is a fundamental matrix of
(1.46), with column vectors y 1 , ..., y k , then every solution y of (1.46) can be
written as a linear combination of y 1 , ..., y k :
c1
k
X .
y(n) = ci y i (n) = Y (n)
.
i=1
ck
where c1 , ..., ck ∈ C. In solving initial value problems of the form
y(n + 1) = A(n)y(n), y(0) = y0
it is sometimes convenient to use the particular fundamental matrix Y of
(1.46) with the property that Y (0) = I. Then the solution is given by
y(n) = Y (n)y0
If equation (1.46) is derived from a scalar, kth order, homogeneous linear
difference equation, by means of the transformation described in the previous
section, then, for j ∈ {1, ..., k}, y j has the form
57
1.4.2 Inhomogeneous systems of first order linear dif-
ference equations
Next, we consider the inhomogeneous, first order, vectorial difference equa-
tion
58
y(n + 1) = Ay(n) (1.49)
has the general solution
y(n) = An c, c ∈ Ck
An S = SJ n (1.51)
If, on the other hand, A has a single eigenvalue λ with geometric multiplicity 1
and generalized eigenvectors s1 , ..., sk , such that, for j = 2, ..., k, (A−λI)sj =
59
sj−1 , then J consists of a single Jordan block and the column vectors of An S
take the form
! ! !
n n n n−1 n n−1 n
λ s1 , λ s2 + λ s1 , ..., λn sk + λ sk−1 + ... + λn−k+1 s1
1 1 k−1
(1.52)
(Cf. the appendix, §A.1 and §A.2.)
Note that Y1 (0) = I. With the aid of §A.2 of the appendix, we find (cf. also
example A.1.6)
! ! ! !
1 0 4n n4n−1 1 0 n−1 4+n n
Y1 (n) = =4
−1 1 0 4n 1 1 −n 4−n
Verify that the column vectors of this matrix do indeed form a basis of the
solution space.
Another fundamental matrix is
! !n ! !
1 0 4 1 4n n4n−1 n−1 4 n
Y2 (n) = = =4
−1 1 0 4 −4n (4−n)4n−1 −4 4−n
60
of offspring per individual per year, is assumed to be independent of time
and equal to 0, 23 and 1, respectively. The composition of this population
satisfies the following system of equations
3
y1 (n + 1) = y (n)
2 2
+ y3 (n)
1
y2 (n + 1) = y (n)
2 1
,
1
y3 (n + 1) = y (n)
2 2
0 12 0
The eigenvalues of the above matrix are the solutions of the equation
3 1 1
−λ3 + λ + = (1 − λ)(λ2 + λ + ) = 0
4 4 4
x3 2
Hence it follows that
3
−1
0 12
4 1 −2 1 0 0 4 1 −2
1 1
A := 2 0 0 = 2 −1 0 0 − 2 1 2 −1 0
0 12 0 1 1 2 0 0 − 21 1 1 2
61
and, for all n ∈ N,
n
4 1 −2 1 0 0 2 4 2
n 1 1
A = 2 −1 0 0 − 2 1 4 −10 4
18
1 1 2 0 0 − 12 −3 3 6
0
1 1 2 0 0 − 12 −300
4 1 −2 1 0 0 200
1 1 1
= 2 −1 0 0 (− 2 )n n(− 2 )n−1 18 400
1 1
2
0 0 (− 21 )n −300
4 1 −2 2
= 9 2 −1 0 (6n + 4)(− 12 )n
50
1 n
1 1 2 −3(− 2
)
1 n
4 + (3n + 5)(− 2 )
= 100
9
2 − (3n + 2)(− 12 )n
1 + (3n − 1)(− 12 )n
62
Proof. Let {s1 , ..., sk } be a basis of generalized eigenvectors of A, and S the
matrix with column vectors s1 ,...,sk , such that the corresponding change of
coordinates takes A to ‘its’ Jordan normal form J = S −1 AS. Every solution
y of (1.49) can be written in the form
λn1
0 . . 0
0 Jk2 (λ2 )n . . 0
y(n) = SJ n c =
S
. . . . . c,
(1.54)
. . . . .
0 0 . . Jkr (λr )n
where c = (c1 , ..., ck ) ∈ Ck . J1 (λ1 ), Jk2 (λ2 ),..., Jkr (λr ) are the Jordan blocks
of J. A being a primitive matrix, with dominant eigenvalue λ1 , we have
λ1 > |λi | for i = 2, ..., r, and this implies that
lim λ−n n
1 Jki (λi ) = 0 (= the null matrix of order ki ) (1.55)
n→∞
lim λ−n −n
1 ky(n)k = lim kλ1 y(n)k = kc1 s1 k = |c1 |ks1 k (1.57)
n→∞ n→∞
Dividing (1.56) by (1.57) we find, for any solution y of (1.49) with y(0) 6= 0,
y(n) c1 s 1
lim = , provided c1 6= 0
n→∞ ky(n)k |c1 | ks1 k
It remains to be proved that c1 > 0 when y(0) >> 0, as that would imply
|c1 | = c1 . From (1.54) we deduce that c = S −1 y(0) and thus
k k
−1
y(0)j (S −1 )Tj1 = y(0) · (S −1 )∗ e1
X X
c1 = S1j y(0)j = (1.58)
j=1 j=1
63
(The scalar or inner product of two vectors x and y ∈ Ck is defined by:
x · y := kj=1 xj y j . For any matrix B, B ∗ is defined by: B ∗ = B T . Check
P
AT (S −1 )∗ e1 = (S −1 A)∗ e1 = (JS −1 )∗ e1 = (S −1 )∗ J ∗ e1 = λ1 (S −1 )∗ e1 ,
(S −1 )∗ e1 · s1 = e1 · S −1 s1 = e1 · e1 = 1
64
Theorem 1.4.13. Let A ∈ Rk×k and suppose A has an eigenvalue λ with
nonzero imaginary part and eigenvector v. Then the sequences y1 and y2
defined by
y1 (n) = Re λn v and y2 (n) = Im λn v
are linearly independent solutions of (1.49).
Example 1.4.15.
2 1 1
y(n + 1) = Ay(n) := 0 1 1
y(n) (1.59)
−1 −1 0
1
A has an eigenvalue 1 with eigenvector s1 := −1 , an eigenvalue 1 − i
0
1
with eigenvector s2 := −i
and an eigenvalue 1 + i with eigenvector
−1
1
i . Hence a basis of the solution space of (1.59) is formed by
s3 :=
−1
the equilibrium solution y1 ≡ s1 and the (complex conjugated)
√ −iπ/4 √ iπ/4sequences
{(1 − i)n s2 }∞
n=0 = {( 2e ) n
s } ∞
2 n=0 and {(1 + i) n
s 3 = {( 2e )n s3 }∞
n=0 .
The general term of the second solution can be written in the following form
−inπ/4
n
√ n e
{(1 − i) s2 = ( 2) −ie−inπ/4
e−inπ/4
By Theorem 1.4.13, a real-valued basis is formed by y1 and the sequences
∞ ∞
cos(nπ/4)
− sin(nπ/4)
n/2 n/2
Re y2 = 2 − sin(nπ/4) and Im y2 = 2 − cos(nπ/4)
− cos(nπ/4) n=0
sin(nπ/4)
n=0
65
A constant sequence y ≡ c, with c ∈ Ck , c 6= 0, is an equilibrium solution
of the homogeneous equation (1.49) iff Ac = c, i.e. iff c is an eigenvector of
A with eigenvalue 1. A periodic solution of (1.49), of period p, is a solution
y with the property that y(n + p) = y(n), so
Ap y(n) = y(n), n ∈ N
This implies that (1.49) has a p-periodic solution iff A has an eigenvalue λ
with the property that
This can be verified by inserting the above expression for y into the equation:
n n−1
An−m b(m) − An−m b(m) = b(n) for n ≥ 1
X X
y(n + 1) − Ay(n) =
m=0 m=0
and
y(1) − Ay(0) = y(1) = A0 b(0) = b(0)
Hence the general solution of (1.48) is
n−1
y(0) = c, y(n) = An c + An−m−1 b(m) for n ≥ 1 c ∈ Ck
X
m=0
66
Due to the periodicity of b,
n+p−1 n+p−1 n−1
n+p−m−1 n+p−m−1
An−m−1 b(m)
X X X
A b(m) = A b(m − p) =
m=p m=p m=0
m=0
hence also to
p−1
(I − Ap )c = Ap−m−1 b(m)
X
m=0
(why?) This shows that (1.48) has a unique periodic solution of period p,
iff the matrix I − Ap is nonsingular. The initial vector of this solution is
p−1
y(0) = (I − Ap )−1 Ap−m−1 b(m)
X
m=0
y(n) = An y(0)
67
Definition 1.4.16. Let A ∈ Ck×k . The spectrum σ(A) of A is the collection
of all eigenvalues of A. The spectral radius rσ (A) of A is defined by
Theorem 1.4.17. The solutions of (1.62) are stable if the following condi-
tions hold:
1) rσ (A) ≤ 1,
2) the algebraic and geometric multiplicities of every eigenvalue λ of A such
that |λ| = 1 are equal.
In all other cases the solutions are unstable. The solutions of (1.62) are
globally asymptotically stable iff rσ (A) < 1.
From this theorem two necessary (but not sufficient) conditions for stabil-
ity and asymptotic stability of the solutions of (1.62) can be deduced. Sup-
pose the Jordan normal form of the matrix A has diagonal entries λ1 , ..., λk
(the eigenvalues of A). If the solutions are asymptotically stable, then, ac-
cording to Theorem 1.4.17, |λi | < 1 for all i. As
k
Y k
X
det A = λi and tr A = λi
i=1 i=1
(this follows from lemma A.1.1 in the appendix), this implies that
68
EAs denote the linear subspace of Ck spanned by all generalized eigenvectors
of A, corresponding to eigenvalues λ with the property that |λ| < 1, i.e.
EAs =
M
Eλ
λ∈σ(A):|λ|<1
Theorem 1.4.18. (i) EAs is the stable set of the fixed point 0 of the linear
map f : x 7→ Ax.
(ii) If y and y∗ are solutions of the system (1.62), then limn→∞ y(n)−y∗ (n) = 0
iff y(0) − y∗ (0) ∈ EAs .
Ci,t = ci Yi,t−1
Xi = Mj . i 6= j
69
1.4.5 Exercises
1. Determine the general solution of the vectorial difference equation
!
2 −1
y(n + 1) = y(n), n∈N
1 4
!
1
and find the solution satisfying the initial condition: y(0) = .
1
!
1
2. Prove that is an eigenvector of the matrix
1+i
!
2 −1
A=
2 0
and determine the general, real-valued, solution of the equation
y(n + 1) = Ay(n), n ∈ N
0 0 1 1
1 0 1
70
7. Find the population vector y(n) in example 1.4.10, when the initial pop-
ulation is composed of 100 second-year individuals. Compute limn→∞ y(n).
Check whether there exists an initial vector y(0) such that, in the long
run, the population becomes extinct.
1
11. Find all periodic solutions of the system
!
−1 1
y(n + 1) = y(n)
a −1
71
Figure 1.11: Orbits of various points (each of which is marked by a cross),
subject to the map f (x) = (x2 , αx2 (1 − x1 )), with α = 2.01; cf. example
1.5.6.
72
Figure 1.12: Orbit of the point ( 12 , − 21 ), subject to the map f (x) = (x2 , 1−x21 );
cf. example 1.5.7.
x−1
!
5 2 x1
f (x) = , f 6 (x) = x
x1
!
1
f has a single fixed point, viz. . The only iterate to have more than
1
! !
3 −1 1
one fixed point is f , the additional fixed points being , and
−1 −1
!
−1
. The vectorial difference equation
1
! !
y1 (n + 1) y2 (n)
=
y2 (n + 1) y1 (n)−1 y2 (n)
73
!
1
has a unique equilibrium solution: y ≡ , no 2-periodic solutions and
1
three 3-periodic solutions having one and the same orbit. All other solutions
are 6-periodic.
In the case that c = 0, similarly to Theorem 1.2.8, the above theorem can be
stated as follows.
kg(x)k
lim =0 (1.66)
x→0 kxk
If rσ (A) < 1, then the null solution of the system of first order difference
equations
y(n + 1) = Ay(n) + g(y(n)) n ∈ N (1.67)
is asymptotically stable. If, on the other hand, rσ (A) > 1, then the null
solution of (1.67) is unstable.
y(n + 1) = Ay(n)
74
Example 1.5.4.
! ! !
y1 (n + 1) y1 (n) y2 (n)2
=A + , n∈N
y2 (n + 1) y2 (n) y1 (n)y2 (n)
Hence
kg(x)k1
lim = lim |x2 | = 0
x→0 kxk1 x→0
Here, a, b, c and r are positive constants. The number of prey that are eaten
during one unit of time (viz. by2 (n)y1 (n)) is proportional to both the number
of prey and the number of predators present, and the same applies to the
number of predators born during that period. Moreover, it is assumed in this
model that no predator lives more than 1 unit of time. The system (1.68)
has three equilibrium solutions, including the null solution and the solution
! !
1
y1 (n)
= 1
c for all n ∈ N (1.69)
y2 (n) b
(r − ac )
75
where x := (x1 , x2 ), is
!
1 + r − 2ax1 − bx2 −bx1
Df (x) =
cx2 cx1
Df (0, 0) is a diagonal matrix with diagonal entries 1+r and 0. From Theorem
1.5.3 it follows immediately that the null solution is unstable. In most cases
however, the more interesting question is, whether or not the equilibrium
solution (1.69) is (asymptotically) stable. In the given (real) context, this
solution exists, provided r > ac . The matrix A now reads
!
1 − ac − cb
A= 1
b
(cr − a) 1
76
Figure 1.13: The attractor of the equation in example 1.5.6, with α = 2.01.
The white area is the basin of attraction.
is converted into the system of first order equations y(n+1) = f (y(n)), where
!
x2
f (x) =
αx2 (1 − x1 )
If α 6= 1, f has two fixed points: the origin and (1 − 1/α, 1 − 1/α). The
Jacobian matrix of f is
!
0 1
Df (x) =
−αx2 α(1 − x1 )
Df (0, 0) has eigenvalues 0 and α, hence the null solution of the equation is
asymptotically stable when |α| < 1 and unstable when |α| > 1. The Jacobian
matrix at the second equilibrium point is
!
0 1
Df (1 − 1/α, 1 − 1/α) =
1−α 1
√
It has
√ eigenvalues λ = 1/2 ± 1/2 5 − 4α when α < 5/4 and λ = 1/2 ±
i/2 4α − 5 when α > 5/4. Both eigenvalues have absolute value < 1 iff 1 <
α < 2. For these values of α, the equilibrium solution y ≡ (1 − 1/α, 1 − 1/α)
is therefore asymptotically stable. At α = 2 a bifurcation occurs: for values
of α > 2, the system no longer has a point attractor, cf. fig. 1.13.
77
By a transformation of the form (1.70), it is converted into the system of
first order equations y(n + 1) = f (y(n)), where
!
x2
f (x) =
a − x21
When a < −1/4, f has no real fixed points, when a > −1/4, √ it has two
real fixed points: c± =√(c±
1 , c±
2 ), where c +
1 = c +
2 = −1/2 + 1/2 1 + 4a, and
− − ±
c1 = c2 = −1/2 − 1/2 1 + 4a. The Jacobian matrix of f at c is
!
± 0 1
Df (c ) = ±
−2c1 0
qq √
For the eigenvalues we have: λ2 = −2c±
1 , so |λ| = 2|c±
1 | = | − 1 ± 1 + 4a|.
q √
The eigenvalues of Df (c− ) have absolute value 1 + 1 + 4a > 1 for all
a > −1/4, hence the equilibrium solution y ≡ c− is qunstable for all a > −1/4.
+
√
The eigenvalues of Df (c ) have absolute value | 1 + 4a − 1| and this is
less than 1 when −1/4 < a < 3/4 and greater than 1 for all a > 3/4. Thus the
equilibrium solution y ≡ c+ is asymptotically stable for all a ∈ (−1/4, 3/4)
and unstable for all a > 3/4. f 2 has two more fixed points, in addition to
c+ and c− , viz. p1 := (c− + + −
1 , c2 ) and p2 := (c1 , c2 ). These correspond to two
2
2-periodic solutions with orbit {p1 , p2 }. The
√ Jacobian matrix √ of f at p1 and
p2 is diagonal, with diagonal entries 1 − 1 + 4a and 1 + 1 + 4a. Hence it
follows that both 2-periodic solutions are unstable for all a > −1/4. Finally,
we look for 4-periodic solutions. To that end we try to solve the equation
! !
4 a − (a − x21 )2 x1
f (x) = =
a − (a − x22 )2 x2
This requires computing the zeroes of the 4th degree function a−(a−x2 )2 −x.
Since all fixed points of f and f 2 are fixed points of f 4 as well, all zeroes of
a − x2 − x are zeroes of a − (a − x2 )2 − x. This implies that a − x2 − x divides
a−(a−x2 )2 −x. Division yields: a−(a−x2 )2 −x = (a−x2 −x)(x 2
√ −x+1−a).
When a > 3/4, the second factor has 2 real zeroes: 1/2 ± 1/2 4a − 3. Hence
78
Figure 1.14: 4-periodic attractor of the map f (x) = (x2 , 1 − x21 ) and its basin
of attraction (the white region). Which points in the white region do not
belong to the basin of attraction? Also compare this figure to fig. 5.2
The Jacobian matrix of f 4 is a diagonal matrix, with diagonal entries 4x1 (a−
x21 ) and 4x2 (a − x22 ). At the first point, both entries are equal to 4 − 4a.
Hence the orbit of this point is a 4-periodic attractor for all values of a in
the interval (3/4, 5/4). In Figure 1.14 this attractor is shown, along with its
basin of attraction, for the case a = 1.
If a fixed point c of f has the property that none of the eigenvalues of Df (c)
has absolute value equal to 1, it is called a hyperbolic fixed point. For
a hyperbolic fixed point, Theorem 1.5.2 provides a necessary and sufficient
condition for stability of the equilibrium solution y ≡ c. Note that, in that
case, neutral stability does not occur. In the case of a non-hyperbolic fixed
point, linearization doesn’t give us any information about the stability of the
corresponding equilibrium solution of (1.65). In some situations, the so-called
direct method of Lyapunov may help us out.
79
all x ∈ U − {x0 }.
(i) If V (f (x)) ≤ V (x) for all x ∈ U such that f (x) ∈ U , then x0 is a stable
equilibrium point.
(ii) If V (f (x)) < V (x) for all x ∈ U − {x0 } such that f (x) ∈ U , then x0 is
an asymptotically stable equilibrium point.
(iii) If V (f (x)) > V (x) for all x ∈ U − {x0 } such that f (x) ∈ U , then x0 is
an unstable equilibrium point.
Proof. (i) Let be a sufficiently small positive number, so that the closed
“ball” B(x0 ; ) := {x ∈ Ck : kx − x0 k ≤ }, where k.k denotes a vectornorm
on Ck , is contained in U . Then V (x) > 0 for all x ∈ B(x0 ; ) − {x0 }. f is
continuous at x0 , hence there exists a δ1 ∈ (0, ) with the property that
kf (x) − f (x0 )k ≤ for all x ∈ B(x0 ; δ1 ) (1.71)
Let
m := min{V (x) : δ1 ≤ kx − x0 k ≤ } (1.72)
Then we have: m > 0 (why?). Since V is continuous and V (x0 ) = 0, there
exists a δ2 ∈ (0, δ1 ), such that
V (x) < m for all x ∈ B(x0 ; δ2 )
We want to prove that f n (x) ∈ B(x0 ; δ1 ) for every n ∈ N and every x ∈
B(x0 ; δ2 ), as this implies
kf n (x)−f n (x0 )k = kf n (x)−x0 k ≤ δ1 < for all x ∈ B(x0 ; δ2 ) and all n ∈ N
hence the stability of the equilibrium point x0 . We give a proof by contra-
diction: suppose there exist x ∈ B(x0 ; δ2 ) and n ∈ N such that f n (x) 6∈
B(x0 ; δ1 ). Let n0 denote the smallest value of n with this property. Thus,
n0 ≥ 1, as f 0 (x) = x ∈ B(x0 ; δ2 ) ⊂ B(x0 ; δ1 ). Now we have
f n0 −1 (x) ∈ B(x0 ; δ1 ), but f n0 (x) 6∈ B(x0 ; δ1 ) (1.73)
From (1.73) and (1.71) we deduce that
δ1 < kf n0 (x) − x0 k = kf (f n0 −1 (x)) − f (x0 )k ≤
In view of (1.72), this implies V (f n0 (x)) ≥ m . On the other hand, from the
properties of V and the fact that f n (x) ∈ B(x0 ; δ1 ) ⊂ U for all n ≤ n0 − 1,
it follows that
V (f n0 (x)) ≤ V (f n0 −1 (x)) ≤ ... ≤ V (x) < m
80
Thus we have produced a contradiction and hence f n (x) ∈ B(x0 ; δ1 ) for every
n ∈ N and every x ∈ B(x0 ; δ2 ).
(iii) Choose > 0 such that B := B(x0 ; ) ⊂ U . We will show that no solution
with initial vector x ∈ B − {x0 } “stays within B”. Again we give an indirect
proof. Suppose that f n (x) ∈ B for all n ∈ N. B being bounded and closed,
thus compact, there exists a subsequence of {f n (x)}∞ n=0 converging to an
nm ∞
element of B. Denote this subsequence by {f (x)}m=1 , where n1 < n2 < ...
and let x0 = limm→∞ f nm (x). Due to the continuity of V , the sequence
V (f nm (x)) converges to V (x0 ). The assumption that V (f (x)) > V (x) > 0
for all x ∈ U − {x0 }, implies that V (f n (x)) is an increasing sequence of
positive numbers, bounded above by the maximum of V on B. Hence the
sequence is convergent and
lim V (f n (x)) = V (x0 ) > 0
n→∞
Consequently,
lim V (f n+1 (x)) = V (x0 )
n→∞
On the other hand,
lim V (f n+1 (x)) = V (f (x0 )) > V (x0 )
n→∞
81
has a fixed point at the origin. The matrix of the linearized system at (0, 0)
in this case is the identity matrix, so Theorem 1.5.2 doesn’t provide any
information about the stability of the null solution of (1.65). Now define
Then we have
V (f (x)) = x21 (1 − x22 )2 + x22 (1 − x21 )2 = x21 + x22 + x21 x22 (x21 + x22 − 4)
again has a fixed point at the origin. Df (0, 0) has eigenvalues 1 and −1, so
that Theorem 1.5.2 does not apply. We use the same function V as in the
previous example. In this case we have
1
V (f (x)) − V (x) = (x21 + x22 )2 (x21 + x22 − 4)
4
Hence V (f (x)) < V (x) for every x ∈ B(0; 2) − {0}. Thus, if y is a solution
of (1.65), with initial vector y(0) ∈ B(0; 2), then we have, for every n ∈ N,
82
x of B(0; 2). Due to the continuity of V (= the square of the 2-norm), we
have
lim ky(nm )k2 = n→∞
m→∞
lim ky(n)k2 = kxk2
and
lim ky(nm + 1)k2 = kxk2
m→∞
On the other hand, due to the continuity of f ,
lim y(nm + 1) = m→∞
m→∞
lim f (y(nm )) = f (x)
hence
lim ky(nm + 1)k2 = kf (x)k2
m→∞
Thus, V (f (x)) = V (x) and this implies x = (0, 0). We conclude that
limn→∞ ky(n)k2 = 0 for every solution with initial vector in B(0; 2), and
hence B(0; 2) is contained in the basin of attraction of (0, 0).
Theorem 1.5.11. Let f be a continuous vector function on Rk or Ck , with
fixed point x0 . Suppose there exists a positive invariant neighbourhood U of
x0 , such that, for every x ∈ U , the sequence {f n (x)}∞
n=0 is bounded and all
its accumulation points belong to U , and a continuous function V : U → R
with the property that V (x0 ) = 0, V (x) > 0 and V (f (x)) < V (x) for all
x ∈ U − {x0 }. Then U is contained in the basin of attraction of {x0 }.
Proof. Suppose that y is a solution of (1.65) with initial vector y(0) ∈
U − {x0 }. Then y is bounded and y(n) = f n (y(0)) ∈ U for all n ∈ N.
Now, 0 ≤ V (y(n + 1)) < V (y(n)) for all n ∈ N, so limn→∞ V (y(n)) exists.
Moreover, if {y(nm )}∞
m=1 with n1 < n2 < ..., is a convergent subsequence,
with limit x, then x ∈ U . Due to the continuity of V , limm→∞ V (y(nm )) =
V (x), hence limn→∞ V (y(n)) = V (x). From the continuity of f it follows
that limm→∞ f (y(nm )) = f (x), hence limm→∞ V (f (y(nm ))) = V (f (x)). On
the other hand, limm→∞ V (f (y(nm ))) = limm→∞ V ((y(nm + 1))) = V (x).
Thus, V (f (x)) = V (x) and this implies that x = x0 . The above argument
shows that every convergent subsequence of y has limit x0 and, since y is
bounded, the sequence as a whole converges to x0 .
Remark 1.5.12. If U is positive invariant and bounded, then, for every x ∈
U , the sequence {f n (x)}∞
n=0 is bounded. If U is positive invariant and closed,
then, for every x ∈ U , all accumulation points of the sequence {f n (x)}∞ n=0
belong to U . If U is positive invariant and compact, then, for every x ∈ U ,
both conditions on {f n (x)}∞n=0 in Theorem 1.5.11 are automatically fulfilled.
83
We conclude this section with two examples of simple economic models,
the dynamics of which is described by a system of nonlinear, autonomous
difference equations.
Example 1.5.13. Kaldor’s business-cycle model in discrete time
The model equations are
Yt+1 − Yt = α(I(Yt , Kt ) − S(Yt , Kt ))
Kt+1 − Kt = I(Yt , Kt ) − δKt
Here, Yt and Kt denote production and capital in the period t, respectively.
The investments I and the savings S are functions of Y and K. It is assumed
∂I
in this model that ∂Y is small, both for very small and very large values, and
∂S
larger for intermediate values of Y , whereas ∂Y is supposed relatively large
for very small and very large values of Y . Consequently, both I and S have
to be nonlinear functions of Y (otherwise all partial derivatives would be
constant).
84
1.5.1 Exercises
1. Deduce Theorem 1.5.3 from Theorem 1.5.2.
2. Prove that the system (1.68) has three unstable equilibrium solutions when
r > 2 ac .
85
9. Prove that the basin of attraction of the fixed point of the map in example
1.5.10 is B(0; 2).
Then the kth order equation (1.74) is equivalent to the first order vectorial
equation
86
0
0 1 . . 0
0
0 0 1 . 0
.
u(n + 1) = . . . . . u(n) + (1.75)
.
. . . . .
.
−a0 . . . −ak−1
−b(n)
det(A − λI) = 0
87
Pj is a polynomial in λ of degree j. Pk (λ) = det(A − λI). We will use an
inductive argument to prove that
For j = 1 we have: P1 (λ) = −ak−1 − λ. Now, let j > 1 and assume the above
equality holds for j − 1. Expanding the determinant along the first column,
we get:
−λ 1 . . 0 1 0 . . 0
0 −λ 1 . 0 −λ 1 . . 0
+(−1)j+1 .−ak−j
Pj (λ) = −λ . . . . . . . . . .
0 0 . . 1
0 0 . 1 0
−ak−j+1 . . −ak−2 −ak−1 −λ 0 . . −λ 1
The first determinant on the right-hand side equals Pj−1 (λ) and the second
one, being the determinant of a lower triangular matrix with diagonal entries
equal to 1, has the value 1. Thus, we find
Ly = −b
88
(cf. §1.3). Denoting by λ1 , ..., λk the (not necessarily distinct) solutions of
the characteristic equation, we can write the kth degree polynomial λk +
ak−1 λk−1 + ... + a0 as a product of linear factors:
L = (τ − λ1 )...(τ − λk )
{λnj }∞ n ∞
n=0 , {nλj }n=0 , ..., {n
kj −1 n ∞
λj }n=0 , j = 1, ..., r
λ2 − 4λ + 4 = (λ − 2)2 = 0
89
has a double root λ = 2. Thus, the solution space is spanned by the se-
quences: {2n }∞ n ∞
n=0 and {n2 }n=0 . The general solution is
(τ − 2)2 y = 0
(Verify this).
2 1
y(n + 3) − y(n + 1) + y(n) = 0, n∈N
3 3
has the following characteristic equation
2 1 1
λ3 − λ + = (λ + 1)(λ2 − λ + ) = 0
3 3 3
√ √
Its roots are λ1 = −1, λ2 = 61 (3 + i 3) and λ3 = 16 (3 − i 3). The last
two roots √are complex conjugated. The tangent of arg λ2 is the quotient of
Im λ2 = 61 3 and Re λ2 = 16 .3. Hence
1 π
arg λ2 = arctan √ =
3 6
The absolute value of λ2 and λ3 is the square root of the sum of the squares
of the real and imaginary parts, i.e.
s s
9 3 1 1√
|λ2 | = |λ3 | = + = = 3
36 36 3 3
(In fact, there is a simpler proof of the above identity : from the coefficient
of y(n) in the equation it follows that λ1 λ2 λ3 = − 31 . As λ1 = −1 and
λ2 λ3 = λ2 λ2 = |λ2 |2 = |λ3 |2 , we have |λ2 |2 = |λ3 |2 = 31 .) Every solution of
the equation is a linear combination of the sequences λn1 , λn2 and λn3 , i.e. of
n π n π
(−1)n , 3− 2 ein 6 and 3− 2 e−in 6
90
If one is solely interested in real-valued solutions, the last two sequences can
be replaced by the real and imaginary parts of λn2 , i.e. by
n π n π
3− 2 cos n and 3− 2 sin n ,
6 6
respectively. Thus, every real-valued solution of the equation can be writ-
ten as follows
n π n π
y(n) = c1 (−1)n + c2 3− 2 cos n + c3 3− 2 sin n , n ∈ N, c1 , c2 , c3 ∈ R
6 6
i.e.
n−1
(An−m−1 )1k b(m) for n ≥ 1
X
y(0) = 0, y(n) = −
m=0
It is not easy, in general, to extract very precise information about the long
term behaviour of y from the above formula and therefore its use in practical
applications is rather limited. In some special cases, a particular solution of
(1.74) can be found by the so-called annihilator method. This is a system-
atic procedure to find particular solutions to certain types of inhomogeneous
difference (or differential) equations, using the technique of “undetermined
coefficients”. It can be applied in the case that the function b is itself a
solution of a homogeneous linear difference equation with constant
coefficients, that is, whenever b can be written as a linear combination of
91
sequences of the form {nk λn }∞
n=0 with k ∈ N and λ ∈ C\{0} (cf. Theorem
1.6.2). We begin by writing (1.74) in the form
Ly = −b
h=0
such that
L0 b = 0
(L0 annihilates b.) Then Ly = −b implies
L0 Ly = −L0 b = 0
y(n + 1) − 3y(n) = 0
92
or, equivalently,
L0 y := (τ − 3)y = 0
The left-hand side can be written as follows
Ly := (τ 2 − 4τ + 4)y = (τ − 2)2 y
Hence, every solution of the inhomogeneous equation also satisfies the (third
order) homogeneous equation
L0 Ly = (τ − 3)(τ − 2)2 y = 0
(λ − 3)(λ − 2)2 = 0
y(n) = c1 2n + c2 n2n + c3 3n ,
Example 1.6.6.
y(n + 1) − y(n) = n2 n ∈ N
The right-hand side can be written as: n2 1n and thus satisfies a third order,
homogeneous linear difference equation, with characteristic equation (λ −
1)3 = 0, viz.
L0 y := (τ − 1)3 y ≡ 0
93
(cf. Theorem 1.6.2). Hence, every solution of the inhomogeneous equation
also satisfies the 4th order homogeneous equation
L0 Ly := (τ − 1)4 y ≡ 0
or, equivalently,
94
Then there are two corresponding solutions u and ũ of (1.75), such that
u(0) = (y(1), ..., y(k − 1)) and ũ(0) = (ỹ(1), ..., ỹ(k − 1))
hence
k ũ(0) − u(0) k∞ ≤ δ
and vice versa. From Theorem 1.4.17 we deduce the following result.
Ct = cYt−1
It = a(Yt−1 − Yt−2 ) + A
Yt = Ct + It
Here, A denotes autonomous investments, assumed independent of the time t,
c is a number between 0 and 1 and the accelerator a is a positive constant. In
this model, the income Yt satisfies the second order, linear difference equation
A
Yt = , t∈N
1−c
We are interested in the stability of the equilibrium solution and in deviations
from this equilibrium, so in the behaviour of non-equilibrium solutions close
to the equilibrium. The characteristic equation of (1.76) is
λ2 − (c + a)λ + a = 0 (1.77)
95
We distinguish 3 cases.
1). The discriminant D of (1.77) is negative. This is the case when
or, equivalently,
(a + c − 2)2 < 4(1 − c)
i.e.
√ √ √ √
(1 − 1 − c)2 = 2 − c − 2 1 − c < a < 2 − c + 2 1 − c = (1 + 1 − c)2
Then the characteristic equation (1.77) has two distinct, complex conjugated
solutions λ+ and λ− , given by
1 q
λ± = (c + a ± i 4a − (c + a)2 )
2
Furthermore, we have
|λ± |2 = λ+ λ− = a
The general, real-valued, solution of (1.76) has the form
A t
Yt = + a 2 (c1 cos φt + c2 sin φt), t ∈ N, c1 , c2 ∈ R
1−c
where
c+a
φ = arg λ+ = arccos √
2 a
The non-equilibrium solutions thus exhibit fluctuations, whose magnitude
increases with increasing a. All solutions are stable if and only if |a| ≤ 1
and asymptotically stable iff |a| < 1. In these cases, the non-equilibrium
solutions exhibit oscillations, or damped oscillations, respectively, about the
equilibrium.
2). D = (c + a)2 − 4a = 0, so
√ √
a = (1 + 1 − c)2 or a = (1 − 1 − c)2
96
with multiplicity 2. The general (real-valued) solution of (1.76) has the form
A t
Yt = + a 2 (c1 + c2 t), c1 , c2 ∈ R
1−c
The solutions are asymptotically stable iff |a| < 1. If, on the other hand,
|a| ≥ 1, all solutions are unstable.
3). D > 0, i.e.
√ √
a < (1 − 1 − c)2 or a > (1 + 1 − c)2
Then the characteristic equation (1.77) has two positive solutions λ1 and
λ2 , with the property that
λ1 + λ2 = c + a and λ1 λ2 = a
(1 − λ1 )(1 − λ2 ) = 1 − λ1 − λ2 + λ1 λ2 = 1 − c > 0
Consequently, either λ1 and λ2 are both less than 1, or they are both greater
than 1. In the first case, all solutions are asymptotically stable, whereas
in the second case they are all unstable. The general (real-valued) solution
has the form
A
Yt = + c1 λt1 + c2 λt2 , c1 , c2 ∈ R
1−c
We conclude this section with a classic example from probability theory (cf.
W. Feller: Introduction to Probability Theory and its Applications).
Example 1.6.9. A gambler plays a series of identical games. In each game,
with probability p he gains one euro and with probability q = 1 − p he looses
one. He stops playing as soon as his capital (in euro’s) has attained a certain
value K, or has vanished entirely. Suppose we want to know in what way
the probability of the latter event depends on the gambler’s starting capital.
We denote the starting capital by the variable n (the number of euro’s) and
the probability that the gambler looses all his money by y(n). At the end
of one game, with probability p his capital has increased to n + 1 and with
probability 1 − p it has decreased to n − 1. Clearly, the following relation
holds
y(n) = py(n + 1) + (1 − p)y(n − 1), n = 1, 2, .....
97
This is a homogeneous, linear, second order difference equation, which can
be rewritten in the form
1 1
y(n + 2) − y(n + 1) + ( − 1)y(n) = 0, n = 0, 1, ..... (1.78)
p p
In addition, we have: y(0) = 1 and y(K) = 0. The last two conditions are
called boundary conditions. A similar argument leads to the following
inhomogeneous difference equation for the expected number of games, z(n),
assuming a starting capital of n euro’s:
z(n) = pz(n + 1) + (1 − p)z(n − 1) + 1, n = 1, 2, .....
or, equivalently,
1 1 1
z(n + 2) − z(n + 1) + ( − 1)z(n) + = 0, n = 0, 1, ..... (1.79)
p p p
with boundary conditions: z(0) = z(K) = 0.
1.6.4 Exercises
1. Find a homogeneous, linear difference equation, with constant coefficients,
that is satisfied by the sequence y(n) = 2−n sin n π4 (n ∈ N).
2. Find a homogeneous, linear difference equation, such that the sequence
y(n) = 2−n (n ∈ N) is a neutrally stable solution.
3. Find a linear difference equation, such that the sequence y(n) = 2−n sin n π4
(n ∈ N) is an unstable solution.
4. Find a linear difference equation, with constant coefficients, such that the
sequence y(n) = 5n (n ∈ N) is an asymptotically stable solution.
5. Set up a second order, linear difference equation, with constant coeffi-
cients, for the sum S(n) of the first n terms of an arithmetic progression,
with common difference v. Let S(0) = 0 and determine S(n) as the solu-
tion of an initial value problem.
6. Set up a second order, homogeneous, linear difference equation, with con-
stant coefficients, for the sum S(n) of the first n terms of a geometric
progression, with common ratio r. Let S(0) = 0 and determine S(n) as
the solution of an initial value problem.
98
7. Set up a second order, linear difference equation, with constant coeffi-
cients, for the sum S(n) of the first n terms of an arithmetic-geometric
progression, with common ratio r and common difference v (cf. Example
1.3.13). Let S(0) = 0 and determine S(n) as the solution of an initial
value problem.
b. Convert the above initial value problem into an initial value problem
for a system of first order difference equations and solve it.
b. Convert the above equation into a system of first order equations and
determine the general solution of this system. Compare the result to
that of part a.
99
15. Determine the general solution of the equation
2 1
y(n + 3) − y(n + 1) + y(n) = n + 2n , n∈N
3 3
100
Chapter 2
Differential equations
101
is an autonomous ODE, as opposed to, for instance, the equation
y0 = t − y2
Both equations can be written in the form (2.1). In the first case, F (t, y, y 0 ) =
y 0 − ay + by 2 , in the second one F (t, y, y 0 ) = y 0 − t + y 2 . Both equations are
examples of non-linear, first order ODE.
The ODE is called linear when F is an affine function of y, y (1) ,...,y (k) .
So a linear ODE has the form
1. y 0 = g
K 0 (t) = sY (t)
L0 (t) = nL(t)
102
Y , K and L denote production, capital and labour, respectively, As in the
discrete case, n > 0, 0 < s < 1 and the production function Q is assumed
homogeneous of degree 1. Again defining k := K
L
, (capital per worker), and
Q(K, L) K
q(k) = = Q( , 1) = Q(k, 1)
L L
we find that k must satisfy the first order separable differential equation (type
2):
k 0 (t) = sq(k(t)) − nk(t) (2.2)
Here, we can choose, for example, f ≡ 1 and G(k) = sq(k) − nk.
103
Example 2.1.3. The general solution of the differential equation
y 0 (t) = 4e2t
is Z
y(t) = 4e2t dt = 2e2t + C.
is
y(t) = log(t − 1) − t−1 + C, C constant.
The solution with initial value y(2) = 1.5 is y(t) = log(t − 1) − t−1 + 2.
104
(real- or complex-valued) constant C is a solution of (2.4). This can be
verified by differentiating both sides of the equation P (y(t)) = F (t) + C. C
can be determined from the initial condition: C = P (y(t0 )) − F (t0 ). This
solution method is called the method of separation of variables.
Remark 2.1.6. Here we have assumed that G(y) 6= 0, but, more generally,
the solution is valid for all t for which y(t) is not a zero of G. If G(k) = 0
then the constant function y ≡ k is a solution of (2.4) and, conversely, if
y ≡ k is a solution of (2.4), then G(k) = 0 (check this!).
y(t) = −1/(4t2 + C). The condition y(0) = 1 then yields C = −1. As the
function y : [0, 1/2) −→ R defined by y(t) = −1/(4t2 − 1) is C 1 and different
from 0 for all t ∈ [0, 1/2), G(y(t)) is never zero and y is the solution of the
initial value problem.
Example 2.1.9. Consider the initial value problem for the logistic differen-
tial equation
y 0 (t) = a(1 − y(t)/k)y(t), y(0) = y0 , (2.5)
where the parameters a and k are positive and y0 ∈ R. Check that the
constant functions y ≡ 0 and y ≡ k satisfy the equation. Applying the
method of separation of variables, we get
Z
y0 Z
dt = a dt.
(1 − y(t)/k)y(t)
105
From example 24.11 in Simon and Blume (pp 644-645) we conclude that
y(t) = ky0 /(y0 + (k − y0 )e−at ).
Note that this expression also represents the solution of the initial value
problem when y0 = 0 or y0 = k, even though 0 and k are zeroes of “G” (the
right-hand side of (2.5)). So it represents the solution to the initial value
problem for all values of y0 . (Determine the domain of the solution y with
initial value y0 < 0.)
y0 = f y (2.6)
or, equivalently, y 0 (t) = f (t)y(t). As this is a special case of type 2, we
can apply the method of separation of variables, providedR y(t) 6= 0 for all t.
This yields y 0 (t)/y(t) dt = f (t)dt, hence log |(y(t)| = f (t)dt. If F is a
R R
and y(t) = ±eC ekt = Dekt . The initial condition then implies that D = a so
the solution is y(t) = aekt .
106
Example 2.1.12. Determine the general real-valued solution of
y0 = f y + g (2.7)
Proof. (i) Suppose that y1 is another solution of (2.7), then y10 (t) − y00 (t) =
f (t)y1 (t) + g(t) − (f (t)y0 (t) + g(t)) = f (t)(y1 (t) − y0 (t)). So the function
z := y1 − y0 solves the associated homogeneous differential equation z 0 = f z
and y1 = y0 + z.
(ii) Conversely, if z is a solution of (2.6) and y1 = y0 + z, then y10 (t) =
y00 (t) + z 0 (t) = f (t)y0 (t) + g(t) + f (t)z(t) = f (t)y1 (t) + g(t).
107
Example 2.1.14. Consider the initial value problem
y 0 (t) = −4ty(t) + 6t, y(0) = 3
It is not difficult to see that this equation has the constant solution y0 ≡ 32 :
y00 (t) = 0 = −4ty(t) + 6t = −4t 23 + 6t = 0. According to example 2.1.12,
2
the general solution of the homogeneous equation is yh (t) = Ce−2t , so the
2
general solution of the inhomogeneous equation is y(t) = 32 + Ce−2t . The
initial condition implies that C = 32 , hence the solution of the initial value
2
problem is y(t) = 32 (1 + e−2t ).
Example 2.1.15. Consider the initial value problem
y 0 = 2ty(t) + (3 − 2t2 )t2 , y(0) = 4
This equation has a solution y0 (t) = t3 : y00 (t) = 3t2 = 2t4 + 3t2 − 2t4 . Ac-
cording to example 2.1.12, the general solution of the homogeneous equation
2
is yh (t) = Cet , so the general solution of the inhomogeneous equation is
2
y(t) = t3 + Cet and from the initial condition we deduce that C = 4, hence
2
the solution of the initial value problem is y(t) = t3 + 4et .
Variation of constants.
A very general method for solving (2.7) is by means of “variation of con-
stants”, similarly to the case of a first order, inhomogeneous linear difference
equation (cf. §1.3.2). Note, however, that it doens’t necessarily lead to a
simple expression for the solutions.
Let F be a primitive of f . Then the general solution of the homogeneous
linear differential equation y 0 = f y, is y(t) = CeF (t) , where C is a (real- or
complex-valued) constant. We try to find a solution of the inhomogeneous
equation (2.7) by allowing C to vary, i.e. to depend on t. Substituting
y(t) = C(t)eF (t) into (2.7) we get a differential equation for C:
y 0 (t) = C 0 (t)eF (t) + C(t)eF (t) F 0 (t) = f (t)C(t)eF (t) + g(t)
As F 0 (t) = f (t), this is equivalent to
C 0 (t)eF (t) = g(t) ⇔ C 0 (t) = e−F (t) g(t)
which has the general solution C(t) = e−F (t) g(t) dt. Hence the general
R
108
Moreover, the solution of (2.7) with initial value y(t0 ) = y0 can be represented
as follows Z t
y(t) = eF (t) {e−F (t0 ) y0 + e−F (τ ) g(τ )dτ } (2.8)
t0
Remark 2.1.16. In the above expressions eF (t) can be replaced by any non-
vanishing solution of the homogeneous equation (2.6). (Verify this.)
2 2
Using the substitution method, we find 0t e−Aτ 4Bτ dτ = −2 B (e−At − 1).
R
A
2
So y(t) = −2 B A
+ (C + 2 B
A
)eAt solves the initial value problem. (In fact, it
is easily seen that the equation admits the constant solution y ≡ −2 B A
, cf.
example 2.1.14.)
109
2.1.5 Exercises
1. Determine the solutions of the following differential equations:
a. Explain that the above process can be modelled by the following initial
value problem:
110
f. What will be the concentration of the chemical waste in the lake on
the long run? (Underpin your answer)
111
Example 2.2.1. Consider the second order, nonlinear ODE
y1 := y, y2 := y 0
Defining y1 and y2 , by
y1 := y, y2 := y 0
we obtain the following system of first order, linear ODE’s
Example 2.2.3. Now consider the general case of a kth order linear ODE:
ak (t)y (k) (t) + ak−1 (t)y (k−1) (t) + ... + a0 (t)y(t) = b(t), t ∈ T, (2.13)
112
where it is assumed that ak (t) 6= 0 for all t ∈ T . Introducing new dependent
variables y1 through yk , as defined in (2.10), we obtain the following system
of first order, linear ODE’s
This system of first order differential equations can also be written in the
form of a first order, linear, vectorial ODE:
0 1 0 . . 0 0
y1 (t) y1 (t)
0 0 1 . . 0 0
y2 (t)
y2 (t)
d
.
. . . . . . . .
= +
dt
.
. . . . . .
.
.
. . . . . .
.
.
.
yk (t) − aak0 (t)
(t)
. . . . − ak−1 (t)
ak (t)
yk (t) b(t)
ak (t)
(2.14)
If the function y is a solution of the kth order, linear ODE (2.13), then the
vector function (y1 , ..., yk ), defined by (2.10), is a solution of the vectorial
ODE (2.14) and, conversely, if the vector function (y1 , ..., yk ) satisfies (2.14),
(j−1)
then y1 is a solution of the kth order ODE (2.13) and yj = y1 for j =
2, ..., k.
From now on, we consider systems of k first order ODE’s, of the form
113
Theorem 2.2.4 (Existence and uniqueness of solutions). Let D ⊂ Ck ,
y0 ∈ D and f : [t0 , ∞)×D → Ck be continuous on a neighbourhood of (t0 , y0 ).
Then there exists a positive number δ and a C 1 -function y : (t0 − δ, t0 + δ) →
Ck such that y 0 (t) = f (t, y(t)) and y(t0 ) = y0 . Moreover, if f is C 1 on a
neighbourhood of (t0 , y0 ), then this solution is unique.
114
ODE that is satisfied by y ≡ 0 is homogeneous (check this). Thus, an
inhomogeneous, linear vectorial ODE never has a null solution. We will
restrict ourselves to the very simple case that the coefficients of the system,
or, equivalently, the entries of the matrix function A, do not depend on time,
thus are constants.
y 0 (t) = Ay(t), t ∈ T
where A ∈ Ck×k .
{y(t) = ceat , c ∈ C}
where y0 is a given real (or complex) number, we have to choose the parameter
c such that ceat0 = y0 . Hence, this solution is
y(t) = ea(t−t0 ) y0
y 0 (t) = Ay(t)
115
is equivalent to a system of uncoupled, scalar ODE’s in the coordinates
y1 , ..., yk :
yi0 (t) = λi yi (t), i = 1, ..., k
The general solution is given by
(cf. §A.3, (45) for the definition of etA ), the general solution can be written
as follows
{y(t) = etA c, c ∈ Ck } (2.17)
We will prove that the general solution of the vectorial ODE y 0 (t) = Ay(t) is
represented by (2.17), even if A is not a diagonal matrix. For this purpose,
we need the following lemma.
Y 0 (t) = AY (t)
116
Hence it follows that
d tA −tA
(e e ) = AetA e−tA + etA .(−A)e−tA
dt
As, for every value of t, the matrix etA commutes with A, the right-hand side
vanishes identically, so we have
d tA −tA
(e e ) = 0
dt
This implies that the matrix function etA e−tA does not depend on t, hence
Thus we have shown that, for every value of t, the matrix etA is invertible,
with inverse e−tA . This property is a special case of Theorem A.3.1 in the
appendix.
Theorem 2.2.8. (i) The general solution of the vectorial ODE (2.18) has
the form
{y(t) = etA c, c ∈ Ck } (2.19)
(ii) The solution of the initial value problem
where y0 ∈ Ck , is
y(t) = e(t−t0 )A y0
Proof. (i) According to lemma 2.2.7, every vector function of the form (2.19)
is a solution of equation (2.18). It remains to be proved that all solutions are
of this form. So suppose that y is a k-dimensional vector function, satisfying
(2.18). We have to show the existence of a vector c ∈ Ck , such that y = etA c.
Noting that
d −tA
(e y(t)) = −Ae−tA y(t) + e−tA y 0 (t) = −Ae−tA y(t) + e−tA Ay(t) = 0
dt
we conclude that the vector function e−tA y(t) is a constant (vector), which
we denote by c. Then we have
117
(ii) In order to find the solution of the initial value problem, we need to
choose the vector c in such a manner that
y(t0 ) = et0 A c = y0
(The last equality is due to the fact that the matrices tA and t0 A commute,
cf. Theorem A.3.1.)
Proof. etA is a nonsingular matrix for every value of t. Hence, the linear
Pk tA Pk
combination of vector functions: i=1 αi yi = e i=1 αi ci , with αi ∈ C,
vanishes identically (i.e. equals the null vector for all values of t) if and only
if
k
X
αi ci = 0
i=1
i.e. precisely then, when the vectors c1 , ..., ck are linearly dependent. Fur-
thermore, according to the previous theorem, every solution y of (2.18) has
the form y = etA c, with c ∈ Ck . If the vectors c1 , ..., ck form a basis of Ck ,
then c is a linear combination of c1 , ..., ck and thus y is a linear combination
of the vector functions etA c1 , ..., etA ck . Hence the statements of the theorem
follow easily.
118
The solution yi is the ith column vector of the matrix function etA . If A
has k linearly independent eigenvectors s1 , ..., sk , with eigenvalues λ1 , ..., λk ,
respectively, then also the vector functions ỹi (t) := etA si , i = 1, ..., k, span
the solution space. Since, for every value of t, si is an eigenvector of the
matrix etA , with eigenvalue etλi , equation (2.18) in this case has a second
basis of solutions, of the form
The advantage of this basis over the previous one is, that the t-dependence
of the individual basis elements is immediately clear, without the need to
compute the entries of etA . The first basis is sometimes convenient in solving
initial value problems.
The matrix A in this case has an eigenvector (1, −2), with eigenvalue 2, and
an eigenvector (1, −1), with eigenvalue 3. The vector functions e2t (1, −2)
and e3t (1, −1) span the solution space of (2.20) and the general solution is
! ! ! ! !
2t 1 3t 1 e2t e3t c1 c1
y(t) = c1 e +c2 e = , ∈ C2
−2 −1 −2e2t −e3t c2 c2
c1 + c2 = 1
−2c1 − c2 = 0
119
which has the solution c1 = −1, c2 = 2. Thus, the solution of the initial
value problem, in vectorial form, is:
! !
2t 1 3t 1
y(t) = −e + 2e
−2 −1
We can summarize the above procedure for solving the initial value problem
as follows. First, write the initial vector (1, 0) as a linear combination of the
eigenvectors (1, −2) and (1, −1) of A (and thus of etA ):
! ! !
1 1 1
= −1 +2
0 −2 −1
Second, multiply each vector on the right-hand side with the corresponding
eigenvalue of etA .
A slightly different approach is by using the second part of Theorem 2.2.8:
!
tA 1
y(t) = e
0
The matrix function etA can be computed with the aid of the eigenvectors
and eigenvalues found above (cf. §A.3):
! ! !−1 !
tA 1 1 e2t 0 1 1 −e2t + 2e3t −e2t + e3t
e = =
−2 −1 0 e3t −2 −1 2e2t − 2e3t 2e2t − e3t
Any matrix function, whose column vectors form a basis of the solution space
of (2.18), is called a fundamental matrix of (2.18). Every fundamental
matrix is itself a (matrix) solution of the equation. For, if Y is a fundamental
matrix of (2.18), with column vectors y1 , ..., yk , then
d
Y (t) = (y10 (t)...yk0 (t)) = (Ay1 (t)...Ayk (t)) = A(y1 (t)...yk (t)) = AY (t)
dt
120
One example of a fundamental matrix of (2.18) is the matrix function with
column vectors yi (t) = etA ei , i = 1, ..., k, i.e. the matrix function etA . If the
vectors c1 , ..., ck form a basis of Ck , then the matrix function Y defined by
Y (t) = etA C
121
that Ax = λx, then Ax = Ax = λx = λx). The initial vector should be a
linear combination of these eigenvectors:
! ! ! ! !
1 1 1 1 1 α
=α +β =
1 2+i 2−i 2+i 2−i β
Hence ! !−1 ! !
α 1 1 1 1 1+i
= =
β 2+i 2−i 1 2 1−i
It now follows that
! !
1 1 1
y(t) = (1
+ i)e (1+3i)t
+ 1 (1 − i)e(1−3i)t
2 2 + i!! 2 2−i
1+i
= Re e(1+3i)t
!
1 + 3i !
t 1 t 1
= e cos 3t − e sin 3t
1 3
Alternatively, we can use the second part of Theorem 2.2.8 to compute this
solution:
! ! !−1 !
1 1 e(1+3i)t 0 1 1 1
y(t) = etA y(0) = (1−3i)t
2+i 2−i 0 e 2+i 2−i 1
(Verify that this vector function does indeed satisfy all requirements.) Note
that it is unnecessary, and somewhat akward, to compute the matrix function
etA .
If A is nondiagonalizable, there is no basis of Ck consisting of eigenvec-
tors of A, but there always exists a basis {s1 , ..., sk } of generalized eigen-
vectors. The solution space of (2.18) then again is spanned by the vector
functions
yi (t) = etA si , i = 1, ..., k
However, only for those values of i for which si is a genuine eigenvector, do
we have yi (t) = eλi t si .
122
Example 2.2.12. The vectorial ODE
!
0 0 1
y (t) = y(t)
0 0
is equivalent to the system
y10 (t) = y2 (t)
y20 (t) = 0
This has the general solution
y2 (t) = c2 , y1 (t) = c2 t + c1 , c1 , c2 ∈ C
In vectorial form: ! !
1 t
y(t) = c1 + c2
0 1
The matrix A has an eigenvector s1 := (1, 0), with eigenvalue 0, and a gen-
eralized eigenvector s2 := (1, 1), of degree 2, with respect to this eigenvalue.
The matrix function etA is easily computed by means of (45) in §A.3, since
An vanishes for all n > 1, so
! !
tA 0 1 1 t
e =I +t =
0 0 0 1
The column vectors of this fundamental matrix form a basis of the solution
space of the equation. Another basis consists of the vector functions
! !
tA 1 tA 1+t
e s1 = and e s2 =
0 1
Verify that both bases yield the general solution determined above.
In the general case, the matrix function etA can be computed as follows.
Suppose that, by a change of coordinates, A is transformed into a Jordan
normal form:
S −1 AS = J = Λ + N
where Λ = diag{λ1 , ..., λk } and N is a nilpotent matrix of a particular form,
which commutes with Λ (cf. §A.1). Then, due to Theorem A.3.1,
k−1
X nt Nn
S −1 etA S = etJ = etΛ etN = diag{eλ1 t , ..., eλk t } (2.21)
n=0 n!
123
In particular, for a single k × k Jordan block, with eigenvalue λ, we have
t2 tk−1
1 t 2!
. (k−1)!
tk−2
0 1 t . (k−2)!
etJk (λ) = eλt (2.22)
. . . . .
. . . . t
. . . . 1
Example 2.2.13. The matrix
!
5 1
A=
−1 3
is nondiagonalizable. In example A.1.6, in the appendix, it is shown that
! ! ! !
5 1 1 0 4 1 1 0
=
−1 3 −1 1 0 4 1 1
Hence the equation y 0 (t) = Ay(t) has a fundamental matrix of the form
!
! t 4 1 ! ! ! !
tA 1 0 0 4 1 0 4t 1 0 1 t 1 0
e = e =e
−1 1 1 1 −1 1 0 1 1 1
Another fundamental matrix is
! ! ! !
tA 1 0 4t 1 0 1 t 4t 1 t
e =e =e
−1 1 −1 1 0 1 −1 1 − t
The general solution of the equation has the form
! ! !
4t 1 t c1 4t c1 + c2 t
y(t) = e =e , c1 , c2 ∈ C
−1 1 − t c2 c2 − c1 − c2 t
The corresponding initial value problem, with initial vector
!
1
y(0) =
2
has the following solution
! ! ! ! ! !
tA 1 4t 1 0 1 t 1 0 1 4t 1 + 3t
y(t) = e =e =e
2 −1 1 0 1 1 1 2 2 − 3t
(Verify that this vector function has all the required properties.)
124
In cases where A ∈ Rk×k has eigenvalues with nonzero imaginary part and
where one is exclusively interested in real-valued solutions, it is inconvenient
to use the fundamental matrix SetJ . In such cases, the following theorem
can be used to construct a real-valued basis of the solution space of (2.18).
Theorem 2.2.14. Suppose A ∈ Rk×k has an eigenvalue λ with nonzero
imaginary part and eigenvector v. Then the functions y1 and y2 defined by
y1 (t) = Re etλ v and y2 (t) = Im etλ v
are linearly independent solutions of (2.18).
Example 2.2.15. Consider once more the equation in example 2.2.11. The
matrix A in this example has an eigenvector (1, 2 + i) with eigenvalue 1 + 3i.
With the aid of Theorem 2.2.14, we obtain the following real-valued basis of
the solution space of (2.18):
( !) !
(1+3i)t 1 t cos 3t
y1 (t) = Re e =e ,
2+i 2 cos 3t − sin 3t
( !) !
(1+3i)t 1 t sin 3t
y2 (t) = Im e =e
2+i 2 sin 3t + cos 3t
We can solve the initial value problem in example 2.2.11 by writing y as a
linear combination y = αy1 + βy2 of y1 and y2 and solving α and β from the
equation
! ! !
1 1 0
y(0) = = αy1 (0) + βy2 (0) = α +β
1 2 1
(Show that this yields the solution found in example 2.2.11).
Example 2.2.16. The matrix
−1 1 0
A := −1 1 −1
0 1 −1
1
has an eigenvector s1 = 0 with eigenvalue −1, an eigenvector s2 =
−1
1 1
1 + i with eigenvalue i and (thus) an eigenvector s3 = 1 − i with
1 1
125
eigenvector −i. Hence the solution space of the equation y 0 (t) = Ay(t) has
a basis formed by e−t s1 , eit s2 and e−it s3 . By Theorem 2.2.14, the solution
space has a real-valued basis {y1 , y2 , y3 }, where
1
1
cos t
−t it
y1 (t) = e 0 , y2 (t) = Re e 1 + i = cos t − sin t
−1
1
cos t
and
1
sin t
y3 (t) = Im eit
1 + i = sin t + cos t
1
sin t
The right-hand side being continuous on T , both sides are integrable and,
for every t ∈ T , we have
Z t Z t
d −τ A
(e y(τ ))dτ = e−τ A b(τ )dτ
t0 dτ t0
126
Multiplying both sides from the left with etA , we obtain
Z t
y(t) = e(t−t0 )A y(t0 ) + etA e−τ A b(τ )dτ, t ∈ T
t0
k
Thus, for every vector y0 ∈ C , the initial value problem
y 0 (t) = Ay(t) + b(t), y(t0 ) = y0
has the (unique) solution
Z t
y(t) = e (t−t0 )A
y0 + e tA
e−τ A b(τ )dτ, t ∈ T
t0
satisfies the equation (check this). Hence, the above expression represents
the general solution of the inhomogeneous equation. Note that the second
term on the right-hand side is the solution of the initial value problem with
initial vector 0 (the nullvector of Ck ). Setting b ≡ 0 we retrieve the general
solution of the homogeneous equation (cf. Theorem 2.2.8).
Example 2.2.17. The solution of the initial value problem
! ! !
0 4 1 0 1
y (t) = y(t) + , y(0) =
−2 1 −2et 0
is given by ! !
1 t Z
0
tA
y(t) = e + e(t−τ )A dτ
0 0 −2eτ
With the aid of example 2.2.10, we find
! ! !
1 1 e2t 0 −1
y(t) = +
−2 −1 0 e3t 2
!Z ! ! !
1 1 t e2(t−τ ) 0 −1 −1 0
+ dτ
−2 −1 0 0 e3(t−τ )
2 1 −2eτ
After some computations, we obtain
!
e3t + e2t − et
y(t) =
−e3t − 2e2t + 3et
(Verify this.)
127
2.2.4 Exercises
1. Determine the matrix function etA in the following cases:
! ! !
1 0 0 1 0 1
A= , A= , A=
1 2 1 0 −1 0
128
9. Prove that the vector functions y1 and y2 , defined by
! !
e−t e−2t
y1 (t) = , y2 (t) = , t∈R
3e−t −4e−2t
are linearly independent. Find out whether or not there exists a homoge-
neous, linear, vectorial ODE, such that both y1 and y2 are solutions.
10. Determine a fundamental matrix and the general solution of the vectorial
ODE !
5 −2
y 0 (t) = y(t)
4 −1
14. Determine the general solution of the vectorial ODE: y 0 (t) = Ay(t), where
0 1 0
A= 0
0 1
2 −5 4
129
and find the solution of the initial value problem
et
1
y 0 (t) = Ay(t) + 2et , y(0) = 1
4et 1
130
y 0 (t) = A(t)y(t) + b(t), t ∈ [t0 , ∞) (2.24)
Here, A and b are a k×k matrix function and a k-dimensional vector function
on [t0 , ∞), respectively. Analogously to the case of (a system of) linear
difference equations, we have the following theorem.
Theorem 2.3.2. All solutions of the linear, vectorial ODE (2.24) are neu-
trally stable, globally asymptotically stable, or unstable, if and only if the
null solution of the homogeneous equation is neutrally stable, asymp-
totically stable, or unstable, respectively.
Again, we consider the special case of a system of first order, linear ODE’s
with constant coefficients, i.e. an equation of the form
131
Theorem 2.3.3. The null solution of equation (2.26) is asymptotically
stable if and only if the real parts of all eigenvalues of A are negative. The
null solution is stable if A has no eigenvalues with positive real part, and, in
addition, the algebraic and geometric multiplicities of every purely imaginary
eigenvalue (i.e. with real part = 0) are equal. In all other cases, the null
solution is unstable.
From the above theorem the following necessary (but not sufficient!) con-
ditions for stability and asymptotic stability of the null solution of (2.26) (and
hence of all solutions of (2.25)) can be derived.
Re tr A ≤ 0
Re tr A < 0
132
Suppose that Re λ ≥ 0. Then |Aii | + Re λ ≥ |Aii | and, with (2.28) and
(2.29), it follows that
|Aii − λ| ≥ |Aii | (2.30)
Now consider the matrix B := A − λI. We have: Bii = Aii − λ and Bij = Aij
if j 6= i. From (2.30) and assumption (i) we deduce
X X
|Bij | = |Aij | < |Aii | ≤ |Aii − λ| = |Bii |
j6=i j6=i
kg(y)k
lim =0 (2.32)
y→0 kyk
133
If all eigenvalues of A have negative real parts, then the null solution of the
equation
y 0 (t) = Ay(t) + g(y(t)), t ∈ [0, ∞) (2.33)
is asymptotically stable. If, on the other hand, A has at least one eigen-
value with positive real part, then the null solution is unstable.
134
Like its discrete analogue, the system (2.34) has two more equilibrium solu-
tions, one of which is the following:
! !
1
y1 (t)
= 1
c for all t ∈ [0, ∞)
y2 (t) b
(r − ac )
135
is unstable: an arbitrarily small disturbance suffices to send the pendulum
swinging downward.
This second order ODE can be converted into a system of two first order
ODE’s, by means of the following transformation. Define new (dependent)
variables y1 and y2 by
y1 := y, y2 := y 0 (2.36)
If y is a solution of (2.35), then y1 and y2 satisfy the system of equations
! ! ! !
d y1 (t) 0 1 y1 (t) 0
= + , t ∈ [0, ∞)
dt y2 (t) −1 0 y2 (t) y1 (t) − sin y1 (t)
(2.37)
This is of the form (2.33), with g(y) = (0, y1 − sin y1 ). Now we have
kg(y)k1 |y1 − sin y1 | |y1 − sin y1 |
lim = lim ≤ lim = 0,
y→0 kyk1 (y1 ,y2 )→(0,0) |y1 | + |y2 | (y1 ,y2 )→(0,0) |y1 |
and thus condition (2.32) holds. The eigenvalues of the matrix A are the
solutions of the equation λ2 +1 = 0, so the numbers i and −i. Unfortunately,
however, both eigenvalues are purely imaginary, so that Theorem 2.3.7 does
not apply. This was to be expected, as the theorem provides a sufficient
condition for asymptotic stability, but not for neutral stability of the
null solution. Next, we consider the equilibrium solution y ≡ π of (2.35). It
corresponds to the solution
! !
y1 π
≡
y2 0
of the vectorial ODE (2.37) (check this). Introducing new (dependent) vari-
ables u1 and u2 , defined by
u1 := y1 − π, u2 := y2
we have ! !
d u1 (t) u2 (t)
= , t ∈ [0, ∞)
dt u2 (t) − sin(u1 (t) + π)
136
or, equivalently,
! ! ! !
d u1 (t) 0 1 u1 (t) 0
= + , t ∈ [0, ∞)
dt u2 (t) 1 0 u2 (t) sin u1 (t) − u1 (t)
(2.38)
This is of the form (2.33) again. The matrix A in this case has eigenvalues 1
and −1. By Theorem 2.3.7, the null solution of (2.38), and thus the solution
y ≡ (π, 0) of (2.37), is unstable.
We conclude the example with a version of (2.35) that takes into account the
friction experienced by the pendulum:
! !
d y1 (t) y2 (t)
=
dt y2 (t) −sin y1 (t)−wy2 (t)
! ! !
0 1 y1 (t) 0
= + , t ∈ [0, ∞)
−1 −w y2 (t) y1 (t)−sin y1 (t)
(2.40)
The eigenvalues of the 2 × 2 matrix on the right-hand side of (2.40) are the
solutions of the equation
λ2 + wλ + 1 = 0
For every positive value of w, both eigenvalues have negative real part (the
proof of this assertion is left to the reader, cf. exercise 10 below). With
the aid of Theorem 2.3.7, we conclude that the null solution of (2.40) is
asymptotically stable.
137
x ∈ U − {x0 }.
(i) If V̇ (x) := DV (x)f (x) = kj=1 ∂x
∂V
(x)fj (x) ≤ 0 for all x ∈ U , then y ≡ x0
P
j
is a stable equilibrium solution of (2.31).
(ii) If V̇ (x) < 0 for all x ∈ U − {x0 }, then x0 is an asymptotically stable
equilibrium solution. Moreover, any solution y of (2.31), with the property
that there exists r > 0 such that, for all t ∈ [0, ∞), y(t) ∈ B(x0 , r) ⊂ U ,
converges to x0 as t → ∞.
(iii) If V̇ (x) > 0 for all x ∈ U − {x0 }, then x0 is an unstable equilibrium
solution.
A function V satisfying the conditions of part (i) of Theorem 2.3.10, is called a
Lyapunov function on U , centered at x0 , for equation (2.31). A Lyapunov
function, satisfying the conditions of part (ii) of Theorem 2.3.10, is called a
strict Lyapunov function on U , centered at x0 , for equation (2.31).
Remark 2.3.11. If f is a continuous function from Rk to Rk , then any
solution y of equation (2.31) is a C 1 -function. Consequently, if V : U → R
is a C 1 -function, then so is V ◦ y and
d
V (y(t)) = DV (y(t))y 0 (t) = DV (y(t))f (y(t)) = V̇ (y(t))
dt
The condition that f should be differentiable ensures the existence and
uniqueness of solutions with a given initial value. It is a sufficient condi-
tion and is often replaced by a weaker, so-called Lipschitz condition.
If x0 is an asymptotically stable equilibrium for equation (2.31), the set of
all x in the domain of f , with the property that the solution y of (2.31),
with initial value y(0) = x, converges to x0 as t → ∞, is called the basin of
attraction of x0 .
Example 2.3.12. We can apply Theorem 2.3.10 to the example of the pen-
dulum, with or without friction, i.e. equation (2.40) with w ≥ 0. We define:
U = {x ∈ R2 : |x1 | < 2π} and
1
V (x) = x22 + 1 − cos x1
2
(the energy of the pendulum). It is easily seen that the right-hand side ≥ 0
for all x ∈ R2 and = 0 iff x2 = 0 and x1 = 0 mod 2π, hence V (x) > 0 for all
x ∈ U with the property that x2 6= 0. Furthermore,
∂V ∂V
= sin x1 , = x2 ,
∂x1 ∂x2
138
Hence it follows that
According to the first part of Theorem 2.3.10, the null solution of (2.40) is
stable, also in the frictionless case, i.e. when w = 0. However, this particular
Lyapunov function doesn’t allow us to conclude that the null solution is
asymptotically stable when w > 0, as V̇ (x) = 0 whenever x2 = 0, so for all
points on the x1 -axis, and hence (ii) doesn’t hold.
2.3.2 Exercises
1. a. Determine the general solution of the system of ODE’s
!
0 1 1
y (t) = y(t) t ∈ [0, ∞)
−2 −2
b.
0 2 1 0
−2 0 0 1
y 0 (t) = y(t), t ∈ [0, ∞)
0 0 0 2
0 0 −2 0
139
4. Determine all equilibrium solutions of the system
y10 (t) = (1 − y1 (t) − y2 (t))y1 (t)
, t ∈ [0, ∞)
y20 (t) = (2y1 (t) − 1)y2 (t)
and examine their stability.
5. Determine all equilibrium solutions of the system
y10 (t) = 2y1 (t) − y12 (t) − y1 (t)y2 (t)
, t ∈ [0, ∞)
y20 (t) = y1 (t)y2 (t) − y2 (t)
and examine their stability.
6. a. Determine the equilibrium solutions of the second order ODE
y 00 (t) + y 0 (t) + y(t)y 0 (t)2 + y(t)2 − 1 = 0, t ∈ [0, ∞)
140
b. Prove that none of these equilibrium solutions is asymptotically stable.
c. Use the function V (x1 , x2 ) = x21 + x22 to prove that the null solution is
stable.
12. a. Determine the equilibrium solutions of the system of real-valued differ-
ential equations
y10 (t) = −2y2 (t)2
y20 (t) = y1 (t)y2 (t)
b. Linearize the system around the solution (y1 , y2 ) ≡ (1, 0) and examine
the stability of all nonzero equilibrium solutions of the equation in a.
c. Use a Lyapunov function of the form V (y1 , y2 ) = y12 + by22 to examine
the stability of the null solution of the nonlinear system of differential
equations.
d. Does the system have any asymptotically stable equilibrium solutions?
13. Examine the stability of the null solution of the system of real-valued
differential equations
dy1
dt
= ay2 + by1 (y12 + y22 )
dy2
dt
= −ay1 + by2 (y12 + y22 )
where a, b ∈ R.
14. a. Convert the real-valued equation
y 00 (t) + µy 0 (t) + y(t)3 = 0 (µ > 0)
into a system of first order ODE’s.
b. Use the function V (x1 , x2 ) = 21 x41 + x22 to examine the stability of the
null solution of this system.
15. a. Convert the real-valued equation
y 00 (t) + y 0 (t)y(t)2 + y(t)3 − a2 y(t) = 0
into a system of first order differential equations.
b. Determine the equilibrium solutions of this system and examine their
stability in the case that a ∈ R, a 6= 0.
c. Examine the stability of the null solution of the system in the case that
a = 0, using Lyapunov’s direct method. Try a Lyapunov function of
the form V (x1 , x2 ) = x41 + bx22 .
141
142
Appendix
since
Le1 = 0.e1 + 1.e2 and Le2 = −1.e1 + 0.e2
! !
1 1
whereas, w.r.t. the basis {e01 , e02 } := { , }, it is represented by
1 −1
!
0 1
−1 0
as
Le01 = 0.e01 + −1.e02 and Le02 = 1.e01 + 0.e02
A change of basis of Rk , usually called a coordinate transformation, or
change of coordinates, is a bijective linear map, represented by a nonsingular
k × k matrix. Conversely, every nonsingular matrix with real entries cor-
responds to a change of coordinates in Rk . Likewise, there is a one-to-one
correspondence between changes of coordinates in Ck and nonsingular k × k
143
matrices with complex entries. In what follows, a ‘coordinate transforma-
tion’ refers to a coordinate transformation of Ck . If {e1 , ..., ek } denotes
the standard basis of Ck and {s1 , ..., sk } any other basis, then the change of
coordinates transforming the first basis into the second, is, w.r.t. the stan-
dard basis, represented by the matrix S with column vectors s1 , ..., sk . A
linear map L : Ck → Ck , which, w.r.t. the standard basis, is represented by
a matrix A, in the new coordinate system is represented by
S −1 AS
Proof. Suppose that A and B ∈ Ck×k are similar matrices. Then there
exists a nonsingular matrix S ∈ Ck×k , such that B = S −1 AS. The charac-
teristic polynomial pB of B is given by
144
A k × k matrix A is diagonalizable if and only if it has exactly k
linearly independent eigenvectors v 1 , ..., v k . These eigenvectors form a
basis of Ck and the action of A on these basis vectors simply is multiplication
with a scalar: Av i = λi v i , i = 1, ..., k, where λi is the eigenvalue of A,
corresponding to the eigenvector v i . Let S be the matrix with column vectors
v 1 , ..., v k , then
AS = (Av 1 , ..., Av k ) = (λ1 v 1 , ..., λk v k ) = (v 1 , ..., v k ) diag{λ1 , ..., λk }
Hence S −1 AS = diag{λ1 , ..., λk } (i.e. the diagonal matrix with diagonal
entries λ1 ,...,λk ). A k × k matrix with k distinct eigenvalues, always has
k linearly independent eigenvectors and thus is diagonizable. Furthermore,
every hermitian and, in particular, every real, symmetric matrix is diago-
nizable. (A k × k matrix A is hermitian iff A∗ := AT = A.) Not all matrices
are diagonizable. But every matrix can be brought to a (lower or upper)
triangular form (such as the Schur normal form). The simplest triangular
matrix, that is similar to a given matrix A, is the so-called Jordan normal
form, or Jordan canonical form, of A. If A is diagonizable, then, naturally,
its Jordan normal form is diagonal. A nondiagonizable matrix has at least
one multiple eigenvalue, whose eigenspace has dimension (the geometric
multiplicity) less than the algebraic multiplicity of the eigenvalue. To give
some idea of how such a matrix can be brought to a Jordan normal form,
we begin by considering the case that the k × k matrix A has a single eigen-
value λ, with algebraic multiplicity k and geometric multiplicity l < k. Note
that, if A has an eigenvalue λ, with geometric multiplicity l, then the matrix
B := A−λI has an eigenvalue 0, with geometric multiplicity l and vice versa.
A matrix with a single eigenvalue 0 is nilpotent.
Definition A.1.2. A square matrix B is nilpotent if there exists a positive
integer n such that B n = 0.
Lemma A.1.3. A k × k matrix B is nilpotent, if and only if B has a single
eigenvalue equal to 0. Every nilpotent k × k matrix B has the property that
B k = 0.
To begin with, we will try to bring the nilpotent matrix B = A − λI to the
simplest possible upper triangular form, by means of a coordinate transfor-
mation, i.e. we seek a basis of Ck , on which the action of B is as simple as
possible. If we succeed, the same transformation can be used for A, as
S −1 BS = S −1 AS − λI
145
If the geometric multiplicity of the eigenvalue 0 of B is l, there are l linearly
independent eigenvectors, with eigenvalue 0. We have to complement these
with k − l linearly independent vectors, to a basis of Ck . Note that every
eigenvector of B is an eigenvector of B 2 , B 3 , etc., but the converse is not
true, in general. Consider, for instance, the nilpotent matrix
0 1 0
0 0 1
B=
0 0 0
The matrix
0 0 1
B2 = 0 0 0
0 0 0
has an eigenvector (0, 1, 0), but this is not an eigenvector of B, as B(0, 1, 0)
= (1, 0, 0). (0, 1, 0) is a so-called generalized eigenvector of B, of degree
2, with respect to the eigenvalue 0. At the same time, it is a generalized
eigenvector of A w.r.t. the eigenvalue λ.
(A − λI)j−1 (x + y) = (A − λI)j−1 x 6= 0
whereas
146
If λ = 0, then Ax is a generalized eigenvector of degree j − 1 w.r.t. λ (check
this).
We can now try to supplement the eigenvectors of B = A − λI with
eigenvectors of B 2 and, if necessary, of higher order powers of B, i.e. with
generalized eigenvectors of B (hence of A), until we obtain a basis of Ck .
According to the following theorem, this can always be done.
We will give the proof for the very special case that A has a single eigenvalue
λ, and, moreover, B k−1 = (A − λI)k−1 6= 0. Thus, there exists a vector x,
such that B k−1 x 6= 0. Now define
xj := B k−j x, j = 1, ..., k
and Bx1 = B k x = 0. Hence it follows that B(x1 ...xk ) = (x1 ...xk )Nk , where
Nk is defined by
147
0 1 0 . . 0
0 0 1 . . 0
Nk =
. . . . . .
(41)
0 0 0 . . 1
0 0 0 . . 0
The matrix Nk is an upper triangular matrix of a very special type: it is
the Jordan normal form of B. The Jordan normal form of A is the matrix
Jk (λ) := λIk + Nk , i.e.
λ 1 0 . . 0
0 λ 1 . . 0
Jk (λ) =
. . . . . .
(42)
0 0 0 . . 1
0 0 0 . . λ
A matrix of the above form is called a Jordan block of order k. J1 (λ) is
defined as the 1 × 1 matrix with entry λ.
In general, the Jordan normal form of any particular matrix represent-
ing a linear map is a block matrix, whose diagonal blocks are Jordan
blocks, whereas all other blocks are null matrices. For the Jordan normal
form to be uniquely defined, an additional condition on the order in which
the different Jordan blocks are arranged, is needed. Here, we will consider
‘the’ Jordan normal form as being defined up to a permutation of the Jor-
dan blocks. Obviously, the Jordan normal form of a diagonizable matrix is
diagonal, and all its Jordan blocks are of order 1. In order to find a transfor-
mation that converts the matrix A into its Jordan normal form, one proceeds
as follows. First, determine the eigenvalues of A and, for every eigenvalue as
many linearly independent eigenvectors as possible. The maximal number of
linearly independent eigenvectors, corresponding to a particular eigenvalue
λ, equals the dimension of the eigenspace, i.e. the geometric multiplicity
of λ. If the geometric multiplicity m is less than the algebraic multiplicity
l, there are l − m linearly independent generalized eigenvectors of degree
> 1 w.r.t. λ to be determined. This is done by seeking successively, for
every previously determined generalized eigenvector x0 (note that there is
always at least one eigenvector!), as many linearly independent solutions of
the equation
(A − λI)x = x0
148
as possible. If x0 is a generalized eigenvector of degree g, then
(A − λI)g+1 x = (A − λI)g x0 = 0
and
(A − λI)g x = (A − λI)g−1 x0 6= 0
hence x is a generalized eigenvector of degree g + 1.
Example A.1.6. Consider the matrix
!
5 1
A=
−1 3
Its characteristic polynomial is
det(A − λI) = (5 − λ)(3 − λ) + 1 = (λ − 4)2
The eigenspace, associated with the eigenvalue 4, consists of the solutions of
the vectorial equation
! !
1 1 0
(A − 4I)x = x=
−1 −1 0
It is 1-dimensional and is spanned by, for example, x1 := (1, −1). The
generalized eigenvectors of degree 2 are solutions of the equation
(A − 4I)2 x = (A − 4I){(A − 4I)x} = 0
that is, the vectors x with the property that (A − 4I)x is an eigenvector.
This implies that (A − 4I)x = cx1 for some c ∈ C. For our purpose, we seek
a solution of (A − 4I)x = x1 , i.e.
! !
1 1 1
x=
−1 −1 −1
149
We conclude with some general remarks, concerning the Jordan normal form.
Let J be a k × k matrix in Jordan normal form. J is a block matrix, with the
property that all off-diagonal blocks are null matrices. A matrix with this
property is called a block-diagonal matrix. In the case considered here,
the diagonal entries are Jordan blocks, so matrices of the form (42). We will
denote them by Jki (λi ), i = 1, .., r, where r is a positive integer: 1 ≤ r ≤ k,
and ri=1 ki = k. Each Jordan block can be written in the form
P
A.2 Computation of J n
As a Jordan matrix J is a block-diagonal matrix, with diagonal entries
Jk1 (λ1 ), ..., Jkr (λr ), J n too is block-diagonal, with diagonal entries Jk1 (λ1 )n ,
..., Jkr (λr )n . We are going to examine in detail a single block, of the form
Jk (λ)n = (λIk + Nk )n
150
which is obviously true. Now suppose the statement holds for a certain
positive integer n. Then we have
Pn n
(A + B)n+1 = (A + B) An−m B m
m=0 m
n n n
B + nm=0 m
n−m+1 m
An−m B m+1
P P
= m=0 m A
Pn n!
B + n+1
n+1−m m n! n+1−m m
P
= m=0 m!(n−m)! A m=1 (m−1)!(n+1−m)! A B
n+1 n n! n+1−m m
+ m=1 m!(n+1−m)! (n + 1 − m + m)A B + B n+1
P
= A
Pn+1 (n+1)! n+1−m m Pn+1 n+1 n+1−m m
= m=0 m!(n+1−m)! A B = m=0 m
A B
Hence, by the principle of mathematical induction, the identity holds for all
positive integers.
151
converges as N → ∞, i.e. limN →∞ (SN )ij exists for every i and j ∈ {1, ..., k}.
The limit of (44) is noted eA , thus
∞
An
eA :=
X
(45)
n=0 n!
eA+B = eA eB
Proof. By definition,
∞
(A + B)n
eA+B =
X
n=0 n!
Applying lemma A.2.1, we find
∞ n ∞ X n
!
1 X n 1
eA+B An−m B m = An−m B m
X X
=
n=0 n! m=0 m n=0 m=0 (n − m)!m!
l=0 l! m=0 m!
152
Now, we have !
A 1 1
e =I +A=
0 1
and
∞ P∞
Bn
! !
1
B
X
n=0 n! 0 e 0
e = = =
n=0 n!
0 1 0 1
Hence ! !
A B e 1 B A e e
e e = and e e =
0 1 0 1
Furthermore,
! ! ! !
1 1 1 −1 1 0 1 1
A+B = =
0 0 0 1 0 0 0 1
Consequently,
!
! 1 0 !
A+B 1 −1 0 0 1 1
e = e
0 1 0 1
and thus
! ! ! !
A+B 1 −1 e 0 1 1 e e−1
e = =
0 1 0 1 0 1 0 1
153
A matrix function is continuous, or differentiable, iff all its entries are
continuous or differentiable, respectively. The derivative dA
dx
or A0 of a differ-
entiable matrix function A is the matrix function whose entries are obtained
by differentiating the entries of A, so
The computational rules for integration and differentiation are easily deduced
from those for ‘ordinary’ (i.e. scalar) functions. Thus, for example, the
derivative of the sum of two differentiable m × n matrix functions A and B
equals the sum of the derivatives:
d dA dB
(A + B) = +
dx dx dx
One should be well aware of the fact that matrix multiplication is noncom-
mutative. If A and B are a differentiable l × m and a differentiable m × n
matrix function, respectively, then the (i, j)th entry of (AB)0 is given by
m m
d X
(AB)0ij = (A0ih Bhj + Aih Bhj
0
) = (A0 B + AB 0 )ij
X
( Aih Bhj ) =
dx h=1 h=1
(AB)0 = A0 B + AB 0 ,
However, in general, the factors in each of the terms on the right-hand side
can not be interchanged. For instance, we have
d 2 dA dA
A = A+A ,
dx dx dx
but, in general,
d 2 dA
A 6= 2A ,
dx dx
154
as the following example shows.
!
1 0
A(x) =
x 2
In this case, ! !
d 2 d 1 0 0 0
A (x) = =
dx dx 3x 4 3 0
However,
! ! ! !
dA 1 0 0 0 0 0 dA 0 0
A = = and A=
dx x 2 1 0 2 0 dx 1 0
155
Index
156
inhomogeneous difference equation, 44, periodic attractor, 31, 79
90 periodic solution, 65
inhomogeneous linear difference equa- pitchfork bifurcation, 33, 34
tion, 38 positive invariant, 19
initial value problem, 12, 94 positive limit point, 21, 71
initial values, 39 positive limit set, 21
invariant set, 19, 71 pre-iterates, 34
iterate, 19, 71 predator-prey model
continuous, 134
Jordan block, 148 discrete, 74
Jordan normal form, 148
recurrence relation, 12
linear difference equation, 8, 38
linear difference operator, 38 saddle-node bifurcation, 32
linear differential equation, 102 saddle-node-bifurcation, 13
linearization, 28, 74, 134 Samuelson’s multiplier-accelerator model,
linearized difference equation, 28 94
logistic difference equation, 19, 30, 34 separation of variables, 105
logistic differential equation, 101 shift operator, 11
logistic map, 22 solution space, 38, 39, 41, 56, 89
Lyapunov function, 81, 138 spectral radius, 67
stable, 14, 49, 55, 67, 94, 130, 132,
multiplier-accelerator model 138
Samuelson’s, 94 asymptotically, 24, 30, 49, 55, 58,
neoclassical growth model in contin- 67, 73, 74, 94, 130–134, 138
uous time, 102 neutrally, 55, 58, 130, 131
neoclassical growth model in discrete stable set, 22, 68, 71
time, 6 stationary solution, 10
neutrally stable, 14, 55, 58, 130, 131 transcritical bifurcation, 33, 34
Newton’s method, 15
non-linear multiplier-accelerator model,unstable, 14, 24, 30, 49, 55, 58, 67,
84 73, 74, 94, 130–134, 138
null solution, 10
variation of constants (difference equa-
orbit, 20, 71 tions), 46
variation of constants (differential equa-
parameters, 12
tions), 108
period, 10
period-doubling bifurcation, 34
157