Sperner
Sperner
from sperner‘s lemma to
differential equations
in banach spaces
An Introduction to Fixed Point
Theorems and their Applications
Uwe Schäfer
by
Uwe Schäfer
Priv.-Doz. Dr. Uwe Schäfer
Karlsruher Institut für Technologie (KIT)
Institut für Angewandte und Numerische Mathematik
[email protected]
Impressum
In this book we present several classical fixed point theorems and show how they
have been applied. We further indicate where they are still applied to this very
day.
Next, we present the Poincaré-Miranda theorem (by applying the Brouwer fixed
point theorem) and then apply this theorem to prove the correctness of an algo-
rithm used to solve nonlinear complementary questions, which arise, for example,
in free boundary problems.
In the next chapter we deal with so-called verification methods. The idea is to
use Brouwer’s fixed point theorem as well as the Poincaré-Miranda theorem to
verify, by use of a computer, that within a neighborhood of an approximate solu-
tion there really exists a solution of the problem f (x) = 0, where f : Rn → Rn .
Next, we present Schauder’s fixed point theorem which can be viewed as a nat-
ural generalization of Brouwer’s fixed point theorem to an infinite-dimensional
setting. The classical Peano existence theorem is presented. We also deal with
verification methods in relation to the existence of a solution of a semilinear el-
liptic partial differential equation in a neighborhood of an approximate solution.
These methods were developed in the 1990s and exploit interval arithmetic on a
computer. They still find use today in scientific research helping prove existence
via the computer.
We close with Tarski’s fixed point theorem. Though often overlooked, it is very
elegant and we show how it helps prove the existence of a solution to a parabolic
partial differential equation in a Banach space.
The book is based on a series of lectures on fixed point theorems and their appli-
cations that were given at the KIT (Karlsruhe Institute of Technology, formerly
i
Preface
known as Univerität Fridericiana Karlsruhe (TH)). The idea for this series of lec-
tures arose from the fact that fixed point theorems are used in the verification
methods mentioned above. It seems that this application is not known, even to
experts in fixed point theory.
Most authors of books about fixed point theory are functional analysts, alge-
braic topologists or mathematical economists. They do not consider numerical
methods and so the problem of roundoff errors that occur during calculations
goes unmentioned. As a consequence, no mention is made of verified results
implemented on a computer using interval arithmetic via either the Brouwer or
the Schauder fixed point theorems.
Rather than presenting the many beautiful proofs of Brouwer’s fixed point the-
orem, we opt to expose many nice applications in some detail. It is amazing
that fixed point theorems can be used in such different fields of mathematical
endeavor and we wish to share how this is so with other mathematicians.
We introduce Sperner’s lemma via a board game for children and derive Brouwer’s
fixed point theorem from it. Once that is done the following chapters can be
treated more or less independently.
I thank Prof. Dr. Joseph Diestel from the Kent State University, Ohio for his
interest, advice, and extremely valuable comments. His assiduous proofreading
turned up many infelicities of grammar, logic, and style.
My thanks are also directed to Regine Tobias, to Brigitte Maier, and to Nadine
Klumpp, of KIT Scientific Publishing, for supporting this project.
ii
Contents
1 Sperner’s Lemma 1
1.1 Simplexes and Barycentric Coordinates . . . . . . . . . . . . . . 8
1.2 The Proof of Sperner’s Lemma . . . . . . . . . . . . . . . . . . . 11
3 Nash’s Equilibrium 25
3.1 Two-Person Games . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Nash’s Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Nash’s Proof of Existence . . . . . . . . . . . . . . . . . . . . . . 29
6 Verification Methods 55
6.1 Interval Computations . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2 The Krawczyk Operator . . . . . . . . . . . . . . . . . . . . . . . 62
6.3 The Moore Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.4 The Moore-Kioustelidis Test . . . . . . . . . . . . . . . . . . . . . 65
6.5 The Frommer-Lang-Schnurr Test . . . . . . . . . . . . . . . . . . 68
6.6 A Comparison of the Tests . . . . . . . . . . . . . . . . . . . . . 69
6.7 Final Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
iii
Contents
Bibliography 125
Index 133
iv
1 Sperner’s Lemma
Suppose two children are playing a board game. The board is a big triangle that
is subdivided into smaller triangles as depicted in Figure 1.1. At the starting
point of the game a red chip lies at the left vertex, a green chip lies at the right
vertex, and a blue chip is lying at the vertex at the top.
x x
1
1 Sperner’s Lemma
x x
x x
x x
x x
x x
x x x x x x x
The rules of the game are as follows. Both children are placing alternatively
chips at the remaining vertices. (It is assumed that enough blue, red, and green
chips are available). The children can place their chips at vertices whereever
they want, but at the border edges of the big triangle there is a certain rule.
At the vertices lying on the left border edge of the big triangle it is only allowed
to place red or blue chips. See Figure 1.2. On the other hand, at the vertices
lying on the right border edge of the big triangle it is only allowed to place green
or blue chips. Finally, at the vertices lying on the border edge at the bottom of
the big triangle it is only allowed to place red or green chips.
Now, the aim of the game is to avoid a complete colored small triangle as long
as possible, where a complete colored small triangle is a small triangle with a
2
1 Sperner’s Lemma
x x
x x
x x x x
x x x x x
x x x x x x
x x x x x x x
blue, a red, and a green chip. In other words, the child who places a chip so
that a complete colored small triangle arises is the loser.
Let us consider Figure 1.3. Up to this point, both children have played very
well, because no complete colored small triangle has arisen, yet. However in this
situation the child who has to place the last chip will lose. For example, if he
(or she) places a green chip at the last vertex, then a complete colored small
triangle arises as depicted in Figure 1.4. Also a red or a blue chip will lead to a
complete colored small triangle. The reader can check this on his own.
The question arises: is there always a loser? In other words, will there arise a
complete colored small triangle during the game? This is true, indeed. In fact,
it is a consequence of Sperner’s lemma in the two-dimensional case.
3
1 Sperner’s Lemma
x x
x x x
x x x x
x x x x x
x x x x x x
x x x x x x x
How can we prove it? Let us consider Figure 1.4. Assume, we start a path from
the bottom of the big triangle, where it is only allowed to cross an edge of a
small triangle for the case that the two vertices of this edge are red and green.
We will show, that such a path either will end again at the bottom or it will
reach a complete colored small triangle. See Figure 1.5.
Since the vertices lying on the left border edge of the big triangle are blue and
red, and since the vertices lying on the right border edge of the big triangle are
blue and green, such a path can leave the big triangle neither on the left border
edge nor on the right border edge.
4
1 Sperner’s Lemma
x x
x x x
x x x x
x x x x x
x x x x x x
x x x x x x x
Figure 1.5: The path that reaches a complete colored small triangle
Furthermore, such a path cannot cross three edges of the same small triangle,
because two colors of the three vertices must be the same, if the path crosses
twice the same small triangle. As a consequence, such a path cannot have a
loop. So, such a path either will end again at the bottom or it will reach a
complete colored small triangle. See again Figure 1.5.
The remaining thing that we have to show is that there exists at least one path
that will not end again at the bottom of the big triangle.
To verify this, consider the border edge at the bottom of the big triangle.
x x x x x x x
5
1 Sperner’s Lemma
Going from left to right, since at the beginning of the game at the left vertex
there was a red chip and since at the right vertex there was a green chip, there
must be a small edge with a red chip at the left and a green chip at the right. If
afterwards there is also a small edge with a green chip at the left and a red chip
at the right, then there must be a corresponding small edge with a red chip at
the left and a green chip at the right.
Otherwise the border edge at the bottom will not end at a green vertex.
As a consequence, at the bottom of the big triangle the number of small edges
consisting of a red chip and a green chip must be odd. (Concerning this number
it does not matter if the red chip is on the left and the green chip is on the right
or vice versa.)
See Figure 1.5. At the bottom of the big triangle the number of small edges
consisting of a red and a green chip is 5.
So, at least one path that starts at an edge with a red chip and a green chip
at the bottom of the big triangle will not come back to the bottom. This path
must end in a complete colored small triangle.
This means that the board game indeed ends with a loser. Actually, Sperner’s
lemma says more: Assume the children’s game is not finished when the first
complete colored small triangle occurs but when the last chip is set.
A variant of the children’s game could be that this child who created more
complete colored small triangles than the other one has lost the game.
Also in this variant of the game there is always a loser because - and this is
Sperner’s lemma in the two-dimensional case - the number of complete colored
small triangles is odd. See Figure 1.6.
To prove this we just continue the proof for the fact that there is at least one
complete colored small triangle. Up to this point we have proved the following
two statements:
(S1) The number of edges with a red chip and a green chip at the bottom of
the big triangle is odd.
(S2) Each path from the bottom of the big triangle, where it is only allowed to
cross an edge of a small triangle for the case that the two chips are green
and red, either will end again at the bottom of the big triangle or it will
reach a complete colored small triangle.
With these two statements we can conclude that the number of complete colored
small triangles that were reached by a path started from the bottom of the big
triangle is odd.
6
1 Sperner’s Lemma
x x
x x x
x x x x
x x x x x
x x x x x x
x x x x x x x
Now, if there is another complete colored small triangle that was not reached by
a path that started from the bottom of the big triangle, then a path that starts
from this complete colored small triangle (where it is only allowed to cross an
edge of a small triangle for the case that the two chips are green and red) will
end in a further complete colored small triangle that was also not reached by a
path that started from the bottom of the big triangle. See Figure 1.7.
This means that complete colored small triangles that are not reached by a path
started from the bottom of the big triangle arise pairwise, and so we have proved
a third statement:
(S3) The number of complete colored small triangles that are not reached by a
path started from the bottom of the big triangle is even.
7
1 Sperner’s Lemma
x x
x x x
x x x x
x x x x x
x x x x x x
x x x x x x x
Together with the two statements (S1) and (S2) it follows that the number of
complete colored small triangles is odd.
8
1.1 Simplexes and Barycentric Coordinates
−−→
V0 V2
−−→
V0 V1
with
−−→ −−→ −−→ −−−→
OP = x0 · OV0 + x1 · OV1 + ... + xN · OVN
so that
9
1 Sperner’s Lemma
x
V2
x x
P
x x
x x x x
x x x x x
x x x x x x
V0 V1
x x x x x x x
Figure 1.9: Figure 1.3 combined with the vector addition (1.1)
−−→
Each point within the simplex can be reached via a linear combination of V0 V1
−−→
and V0 V2 . For example, let in Figure 1.3 the point P be the vertex without a
color. Then, (see also Figure 1.9)
10
1.2 The Proof of Sperner’s Lemma
Via
−−→ −−→ −−→ −−→ −−→ −−→ −−→ −−→ −−→
V0 V1 = OV1 − OV0 , V0 V2 = OV2 − OV0 , and V0 P = OP − OV0
In general, for n = 2, 3, ... we form the nth barycentric subdivision of the simplex
as follows. The new vertices are the points with barycentric coordinates
k0 k1 kN
, , ..., ,
n n n
The original simplex is the whole body; the subsimplexes are its cells. In general,
in the nth subdivision of a simplex in N dimensions, the number of cells tends
to infinity as n → ∞; but the diameter of each cell tends to zero. If ∆ is the
diameter of the body, then the diameter of each cell is ∆
n.
11
1 Sperner’s Lemma
Proof: The proof is by induction, where the case N = 1 is trivial and the case
N = 2 was done via the board game for children.
For the general proof we use some notation. If the vertices of a cell have the labels
m0 , m1 , ..., mk , we say the cell has type (m0 , m1 , ..., mk ). Here permutations do
not matter, but multiplicities do matter.
A cell in N dimensions has boundary elements of dimensions 0, 1, ..., N − 1.
Those of dimension 0 are vertices; those of dimensions N − 1 we will call faces.
We will call the whole cell an element of dimension N .
By F (a, b, ..., q) we mean the number of elements in the body of type (a, b, ..., q).
Let us look at the faces labeled (0, 1, ..., N − 1). Two such faces occur in each
cell of type
(0, 1, ..., N − 1, m) if m ∈ {0, 1, ..., N − 1}.
On the other hand, one face labeled (0, 1, ..., N − 1) occurs in each cell of type
(0, 1, ..., N − 1, N ).
This sum, abbreviated by S, counts every interior face (0, 1, ..., N − 1) twice,
since every interior face is shared by two cells; but the sum counts every face
(0, 1, ..., N − 1) on the boundary of the body only once, because each face on the
boundary belongs to only one cell. Therefore, the sum S satisfies
Sperner’s lemma applies with dimension N −1, and so Fb (0, 1, ..., N −1) is odd by
induction. Therefore, S is odd, which implies that F (0, 1, ..., N − 1, N ) is odd.
12
2 Brouwer’s Fixed Point Theorem
The Brouwer fixed point theorem was published by him in 1912. Three years
before, Brouwer could only prove his theorem for n = 3, and it was Hadamard
who in 1910 gave the first proof for arbitrary n.1
Brouwer’s proof differs from that of Hadamard and meanwhile, there are many
other proofs. See, for example, any of the books J. Franklin (32), J. Appell and
M. Väth (9) or H. Heuser (46). Brouwer’s fixed point theorem is a topological
result, and most proofs are topological. Here, we decided to present the proof of
Knaster-Kuratowski-Mazurkiewicz given in 1929 using the lemma of Sperner.
This chapter is organized as follows. First, we state the fixed point theorem and
we present some examples to illustrate it. Then, we give some examples to show
that no assumption can be dropped. In following sections we give the proof of
the theorem in two steps. First, we give the proof for simplexes and then, based
on this, we give the proof for convex sets.
Formally, the theorem reads as follows.
Example 2.1. Consider Figure 2.1. There you can see the graph of a function
f : [a, b] → [a, b]. Being a self-mapping means, that the graph must stay within
the square [a, b] × [a, b]. Being continuous means, that the graph has to cross
the function y = x. So, Brouwer’s fixed point theorem is more or less obvious
for the case n = 1.
Of course, there can be more than one fixed point. See Figure 2.2.
It is not that easy, to verify Brouwer’s fixed point theorem in higher dimensions.
For example, consider a cup of coffee. Imagine, you pour carefully the coffee
from one cup to anoter cup in one go, where both cups have equal size and
equal form. Then, Brouwer’s fixed point theorem says that at least one particle
of the coffee will be at the same position. It is not an easy task to verify this.
1 See (55).
13
2 Brouwer’s Fixed Point Theorem
a b x
Figure 2.1: One fixed point
At least it is not as easy as using Figure 2.1 in order to verify Brouwer’s fixed
point theorem for n = 1.
Before we give a proof of Brouwer’s fixed point theorem, we give some examples
to show that Brouwer’s fixed point theorem is no longer true if some assumptions
are dropped.
Before, let us recall the definition of a norm in Rn . A function k · k : Rn → R is
called a norm in Rn if it fulfills the following three conditions:
kxk ≥ 0 and kxk = 0 ⇔ x = 0,
kλ · xk = |λ| · kxk,
kx + yk ≤ kxk + kyk.
Here, x, y ∈ Rn and λ ∈ R. Note, that we use 0 for the number zero as well as
for the zero vector.
14
2 Brouwer’s Fixed Point Theorem
a b x
Figure 2.2: More fixed points
Perhaps the most popular norm in Rn is the Euclidean norm which is defined
as follows:
v x1
uX n
x2j with x = ... ∈ Rn .
u
kxke := t
j=1 xn
Theorem 2.2. All norms in Rn are equivalent; i.e., for any two different norms
k · k1 and k · k2 in Rn there exist two positive numbers α, β so that
15
2 Brouwer’s Fixed Point Theorem
we get
n
X n
X
kxk = k xj · ej k ≤ |xj | · kej k ≤ β · kxke
j=1 j=1
with
n
X
β= kej k.
j=1
α · kxke ≤ kxk
is valid for all x ∈ Rn . Then, for each k > 0 there is some x(k) ∈ Rn with
1
· kx(k) ke > kx(k) k.
k
1
y (k) := · x(k)
kx(k) k e
1
· ky (k) ke > ky (k) k and ky (k) ke = 1 for all k ∈ N. (2.1)
k
Due to the theorem of Bolzano-Weierstraß we can conclude that there is a con-
vergent subsequence {y (kl ) }∞ ∗
l=1 concerning k · ke ; i.e., there exists some y ∈ R
n
with
lim ky (kl ) − y ∗ ke = 0.
l→∞
Since
ky (kl ) − y ∗ k ≤ β · ky (kl ) − y ∗ ke
this subsequence is also convergent to y ∗ concerning the norm k · k. On the
one hand, (2.1) implies ky ∗ ke = 1, whence y ∗ 6= 0. On the other hand, this
contradicts to (2.1) because of k1l > ky (kl ) k.
16
2 Brouwer’s Fixed Point Theorem
f (x) = x ⇔ x · (x − 1) = 0
x, y ∈ K ⇒ λ · x + (1 − λ) · y ∈ K
The next example shows that the phrase ’f is continuous’ actually has to read
’f is continuous subject to some norm’, since the theorem was wrong if the
assumption would be changed to ’f is continuous subject to some metric’.
Example 2.5. (concerning continuity subject to some norm) A function
f : Rn → Rn is continuous subject to some norm k · k if for any x∗ ∈ Rn
17
2 Brouwer’s Fixed Point Theorem
Let Rn be equipped with the so-called discrete metric which is defined as follows.
(
∗
1 if x∗ 6= x̂,
d(x , x̂) :=
0 if x∗ = x̂.
Continuity subject to this metric concerning any function f means, that from
lim d(x(n) , x∗ ) = 0
n→∞
it follows
lim d(f (x(n) ), f (x∗ )) = 0.
n→∞
As a consequence,
f (x(n) ) = f (x∗ ) for all n ≥ n0 .
So,
lim d(f (x(n) ), f (x∗ )) = 0.
n→∞
18
2.1 The Proof for Simplexes
xj > fj (x).
19
2 Brouwer’s Fixed Point Theorem
then
−−→ −−→ −−→
x = xp · OVp + xq · OVq + ... + xs · OVs .
Because xj > fj (x) ≥ 0 it follows that
m(x) ∈ {p, q, ..., s} for all x ∈ SF .
Now, Sperner’s lemma says, that there is some cell with N + 1 vertices
X0 (n), X1 (n), ..., XN (n)
satisfying
m = 0 at the vertex X0 (n),
m = 1 at the vertex X1 (n),
..
.
m = N at the vertex XN (n).
Then, by definition of m(x) we have
−−−−−→
x0 > f0 (x) for x = OX0 (n),
−−−−−→
x1 > f1 (x) for x = OX1 (n),
.. (2.4)
.
−−−−−→
xN > fN (x) for x = OXN (n).
We can do this for any n that’s large. Since S is bounded and closed, there is a
subsequence nl so that there is X ∗ ∈ S and
−−−−−→ −−−→
lim OX0 (nl ) = OX ∗ .
l→∞
Furthermore, all the vertices of the cell are close to each other if n is large. So,
−−−−−→ −−−→
lim OXj (nl ) = OX ∗ for all j ∈ {0, 1, ..., N }.
l→∞
20
2.2 The Proof for Convex Sets
The step from a simplex to an arbitrary convex set is done using the following
definition.
Proof: Let B be an interior point of K and let u be any unit vector in RN . Then,
−−→
K contains a boundary point OB + λ · u with λ > 0. For each u there is only one
such boundary point. Why? Suppose there were two with λ1 < λ2 . Remember
that K contains some small ball Q centered at B. But then the point P1 with
−−→ −−→
OP1 = OB + λ1 · u lies interior to the convex hull of the ball Q and the point
−−→ −−→
P2 with OP2 = OB + λ2 · u. Therefore, P1 is interior to K, which is impossible
for a boundary point.
λ = λ(u) > 0
−−→ −−→
so that P with OP = OB + λ · u lies on the boundary of K. The function λ(u)
is called the radial function.
We assert that the radial function is continuous as a function of the unit vector
u. If u(0) were a point of discontinuity, then the boundary of K would contain
points Pk with
−−→ −−→
OPk = OB + λk · u(k) , k = 1, 2, 3, ...
with u(k) → u(0) but with positive λk not converging to λ(u(0) ). Since λ1 , λ2 , ...
is a bounded sequence, there is a subsequence with a limit λ∗ 6= λ(u(0) ). But
−−→
then OB + λ∗ u(0) must lie on the boundary of K, since a limit of boundary
points must be a boundary point, too. Now we would have two boundary points
P∗ and P0 with
−−→ −−→ −−→ −−→
OP∗ = OB + λ∗ · u(0) and OP0 = OB + λ(u(0) ) · u(0) .
But this is impossible, since the boundary point in the direction u(0) is unique.
21
2 Brouwer’s Fixed Point Theorem
1 −−→
Ψ(C) := −−→ · BC, if C 6= B
BC
λ −
−→
kBCke
and
Ψ(B) := 0.
On the other hand let y ∈ RN with kyke ≤ 1. Then,
−−→ y
Ψ−1 (y) := OB + λ( ) · y, if y 6= 0
kyke
and
Ψ−1 (0) := B.
By definition of the radial function λ(·) we have
−−→ !
−−→ BC
kBCke ≤ λ −−→ if C 6= B.
kBCke
Therefore, if C 6= B
−−→
kBCk
kΨ(C)ke = −−→ e ≤ 1.
BC
λ −
−→
kBCke
and
−−→
lim Ψ−1 (y (k) ) = OB, if lim y (k) = 0.
k→∞ k→∞
22
2.2 The Proof for Convex Sets
However, due to
−−−−→
BC (k)
λ −−−−→ ≥ dist(B, boundary of K) > 0
kBC (k) ke
and
y (k)
λ ≤ diameter of K
ky (k) ke
this is obvious.
Now, we are able to prove Brouwer’s fixed point theorem. For completeness, we
state it once more.
Theorem 2.5. (Brouwer’s fixed point theorem) Let K 6= ∅, K ⊆ Rn be
bounded, convex, and closed. Furthermore, let f : K → K be continuous. Then,
there exists some x∗ ∈ K so that f (x∗ ) = x∗ .
ϕ : {y ∈ Rn : kyke ≤ 1} → S
Ψ := ϕ ◦ Φ
23
2 Brouwer’s Fixed Point Theorem
Case 2. We assume that K has dimension m with m < n. Then, there exists
m orthogonal unit vectors u(1) , ..., u(m) so that for all x ∈ K there are uniquely
defined numbers y1 , ..., ym so that
With
y1
. . .
U := u(1) .. ... .. u(m) ∈ Rn×m .. ∈ R
and y := m
ym
we have x = U · y and U T · x = y. Setting
K̃ := {y ∈ Rm : y = U T x with x ∈ K}
we have that K̃ is bounded, convex, and closed with dimension m. For any
y ∈ K̃ we define
f˜(y) := U T · f (U · y).
Then, f˜ : K̃ → K̃ is continuous. Applying Case 1 with n replaced by m the
function f˜ has a fixed point y ∗ ∈ K̃; i.e.,
y ∗ = f˜(y ∗ ).
It follows
U · y ∗ = f (U · y ∗ ).
This means, that x∗ := U · y ∗ is a fixed point of f satisfying x∗ ∈ K.
24
3 Nash’s Equilibrium
Although the movie “A Beautiful Mind” from 2001 starts with the claim that
’mathematicians won World War II’, the main focus of the movie is not on
mathematicians but on the life of one particular mathematician, John Forbes
Nash, Jr., whose contributions were post-World War II.
If you talk about this movie with friends and colleagues the question will arise
about what John Nash did. Sooner or later you might have to explain the
meaning of ’Nash equilibrium’. This chapter is aimed at helping you with such
an explanation.
with the following meaning. For the case that player 1 places his bet on the
ith possibility and player 2 places his bet on the jth possibility, then player 1’s
costs are defined by
aij = eTi Aej ,
bij = eT
i Bej .
Sometimes both players are allowed to split up their bets into different possibili-
ties. We give an example. Player 1 stakes half of his money on the first possibility
25
3 Nash’s Equilibrium
and bets the other half of his money on the second possibility. Player 2 stakes a
quarter of his money on the first possibility and he stakes the remaining money
on the last possibility. Then, player 1’s costs are defined by
1
4
0
1 1
..
( 0 · · · 0)A ,
2 2
.
0
3
4
and player 2’s costs are given by
1
4
0
1 1
..
( 0 · · · 0)B .
2 2
.
0
3
4
This leads us to the definition of a mixed strategy.
Definition 3.1. For m, n ∈ N we define
m
P
Sm := {x = (xi ) ∈ Rm : xi ≥ 0, i = 1, ..., m, xi = 1},
i=1
n
P
Sn := {y = (yj ) ∈ Rn : yj ≥ 0, j = 1, ..., n, yj = 1}.
j=1
26
3.2 Nash’s Equilibrium
and
x̂T B ŷ ≤ xT By for all x ∈ S m , y ∈ S n . (3.2)
The following example will show that such a pair of strategies does not exist in
general.
Example 3.2. Let
5 0 5 10
A= and B = .
10 1 0 1
and
x̂T B ŷ ≤ xT By for all x ∈ S 2 , y ∈ S 2 . (3.4)
We will show that from (3.3) it follows that
1 0
x̂ = , ŷ = . (3.5)
0 1
27
3 Nash’s Equilibrium
and
x̂2 = 0 ⇒ x̂1 = 1 ⇒ ŷ1 = 0.
Therefore, we have ŷ1 = 0 and x̂2 = 0, which implies (3.5). Secondly, if we
substitute (3.5) in (3.4) we get a contradiction because
1
x̂T B ŷ = 10 > 0 = (0 1)B .
0
Example 3.2 shows that, in general, it is possible that no optimal pair of strate-
gies exist in a two-person game. So, it is no surprise that sometimes two compa-
nies have to abandon negotiations, since no agreement could be achieved, even
if dealing was allowed.
A third person who wants to arbitrate can take advantage of Nash’s equili-
brium.
Definition 3.2. Concerning a two-person game defined by A, B ∈ Rm×n , a
pair of strategies (x̂, ŷ), x̂ ∈ S m , ŷ ∈ S n is called a Nash equilibrium if
x̂T Aŷ ≤ xT Aŷ for all x ∈ S m ,
x̂T B ŷ ≤ x̂T By for all y ∈ S n
is valid.
28
3.3 Nash’s Proof of Existence
Proof: Let Ai· denote the ith row of A and let B·j denote the jth column of B.
Then, for (x, y) ∈ S m × S n and via
29
3 Nash’s Equilibrium
m
P
For i = 1, ..., m, ci (x, y) ≥ 0 and so ck (x, y) ≥ 0. From x ∈ S m then it
k=1
follows that x0 ∈ S m . Analogously, one can show y 0 ∈ S n . Therefore,
Now, we show
”⇐”: Let (x̂, ŷ) be a Nash equilibrium. Then, for all i = 1, ..., m it holds
x̂T Aŷ ≤ eT
i Aŷ = Ai· ŷ.
So, we have ci (x̂, ŷ) = 0 for i = 1, ..., m. By the same argument we get dj (x̂, ŷ) =
0 for j = 1, ..., n. Therefore, we get T (x̂, ŷ) = (x̂, ŷ).
”⇒”: We assume that T (x̂, ŷ) = (x̂, ŷ), although (x̂, ŷ) is not a Nash equilibrium.
Then, either there exists some x ∈ S m with xT Aŷ < x̂T Aŷ or there exists some
y ∈ S n with x̂T By < x̂T B ŷ. Since both cases lead to a contradiction in the
same way, we assume without loss of generality, that for some x ∈ S m it holds
Ai· ŷ < x̂T Aŷ for at least one i ∈ {1, ..., m}. (3.10)
Assume, that (3.10) is not true. Then, it follows that min Ai· ŷ ≥ x̂T Aŷ.
1≤i≤m
Therefore,
m
m
P P
xT Aŷ = xi Ai· ŷ ≥ min Ai· ŷ · xi = min Ai· ŷ ≥ x̂T Aŷ
i=1 1≤i≤m i=1 1≤i≤m
contradicting (3.9). So, (3.10) is shown, and it follows ci (x̂, ŷ) > 0 for at least
one i ∈ {1, ..., m}. This implies
m
X
ck (x̂, ŷ) > 0. (3.11)
k=1
30
3.3 Nash’s Proof of Existence
There exists some i ∈ {1, ..., m} with x̂i > 0 and ci (x̂, ŷ) = 0. (3.12)
Suppose, (3.12) is not true. Then, for every i ∈ {1, ..., m} it holds
x̂i
x0i = m
P < x̂i .
1+ ck (x̂, ŷ)
k=1
This means that T (x̂, ŷ) 6= (x̂, ŷ). As a consequence, (3.8) is shown.
We’ve shown the equivalence of the existence of a Nash equilibrium and the
existence of a fixed point for the map T . Now the set S m × S n is a nonempty,
closed bounded convex set and the function T is a continuous self-map on S m ×
S n ; so Brouwer’s fixed point theorem tells us T has a fixed point and, with that,
we also have a Nash equilibrium.
31
4 The Poincaré-Miranda Theorem
equals zero at some point of the interval [a, b] was proved by Bernard Bolzano
(1781-1848). In 1883-1884, Henri Poincaré announced the following generaliza-
tion without proof:
In 1886 Poincaré published his argument on the homotopy invariance of the index
which is a basis for the proof. The result obtained by Poincaré has come to be
known as the theorem of Miranda, who in 1940 showed that it was equivalent
to Brouwer’s fixed point theorem. See (55).
We will prove the Poincaré-Miranda theorem by using Brouwer’s fixed point the-
orem as it was done by Miranda in 1940. The following corollaries will generalize
the Poincaré-Miranda theorem.
33
4 The Poincaré-Miranda Theorem
and
δi− := min{|xi − Li | : x ∈ Ω, fi (x) > 0}.
Thanks to (4.3), we have δi+ > 0 and δi− > 0 for all i = 1, ..., n. Because mi < 0
and Mi > 0 for all i = 1, ..., n, there exist εi satisfying
δi+ δi−
0 < εi < min{− , }, i = 1, ..., n.
mi M i
We claim that
f˜(Ω) ⊆ Ω. (4.4)
Let i ∈ {1, .., n} be arbitrary but fixed. With x ∈ Ω we have −Li ≤ xi ≤ Li .
Firstly, if fi (x) = 0, then
it follows
0 < εi · fi (x) ≤ εi · Mi
and together with Li − xi ≥ δi− one can conclude that
δ−
−Li ≤ xi < xi + εi · fi (x) ≤ Li − δi− + εi · Mi < Li − δi− + i Mi = Li .
| {z } Mi
=f˜i (x)
34
4.1 The Theorem and its Corollaries
it follows
0 > εi · fi (x) ≥ εi · mi ,
and together with Li + xi ≥ δi+ one can conclude
δ+
Li ≥ xi > xi + εi · fi (x) ≥ −Li + δi+ + εi · mi > −Li + δi+ − i mi = −Li .
| {z } mi
=f˜i (x)
with
(k) xi
fi (x) := fi (x) − , i = 1, ..., n, k = 1, 2, 3, ... .
k
Since
(k) Li
fi (x1 , .., xi−1 , −Li , xi+1 , .., xn ) = fi (x1 , .., xi−1 , −Li , xi+1 , .., xn ) + >0
k
and
(k) Li
fi (x1 , .., xi−1 , Li , xi+1 , .., xn ) = fi (x1 , .., xi−1 , Li , xi+1 , .., xn ) − < 0,
k
Case 1.a guarantees the existence of some x(k) ∈ Ω satisfying f (k) (x(k) ) = 0.
Since Ω is compact, there exists a convergent subsequence x(kl ) ; i.e.,
35
4 The Poincaré-Miranda Theorem
Case 2: L1 ≥ 0, ..., Ln ≥ 0.
Case 2.a: L1 = 0, ..., Ln = 0. Then, Ω = {0}, and for all i = 1, ..., n we have
fi (0) ≥ 0 and fi (0) ≤ 0, which leads to f (0) = 0.
Case 2.b: Without loss of generality, we assume that
there exists
x̃1
.
. ∈ Ω̃
.
x̃k
The following corollary will show that it is not necessary that the center of the
domain is the zero vector.
If g : G → R is continuous satisfying
)
gi (x) ≤ 0 for all x ∈ G−
i
i = 1, ..., n,
gi (x) ≥ 0 for all x ∈ G+
i
36
4.1 The Theorem and its Corollaries
−gi (xc1 + x̃1 , .., xci−1 + x̃i−1 , xci − Li , xci+1 + x̃i+1 , .., xcn + x̃n ) ≥ 0
| {z }
∈ G−
i
−gi (xc1 + x̃1 , .., xci−1 + x̃i−1 , xci + Li , xci+1 + x̃i+1 , .., xcn + x̃n ) ≤ 0.
| {z }
∈ G+
i
If g : G → Rn is continous satisfying
or we have
The set G and the function g̃(x) fulfill the assumptions of Corollary 4.1. There-
fore, there exists x∗ ∈ G with g̃(x∗ ) = 0, whence g(x∗ ) = 0.
37
4 The Poincaré-Miranda Theorem
Case 2: The set {1, ..., n} can be subdivided in two sets of indices, say I
and J, so that for all i ∈ I we have either (4.6) or (4.7), whereas for all
i ∈ J there exist either x(1,i) , x(2,i) ∈ G+
i or x
(1,i)
, x(2,i) ∈ G−
i satisfying
(1,i) (2,i)
gi (x ) < 0, gi (x ) > 0.
Without loss of generality, for all i ∈ J let x(1,i) , x(2,i) ∈ G+
i . Then, we can
conclude from (4.5) that
Then, we define G̃ := {x̃ ∈ Rn0 : |xci − x̃i | ≤ Li , i = 1, ..., n0 } and the function
g̃ : G̃ → Rn0 by
g1 (x̃1 , ..., x̃n0 , xcn0 +1 − Ln0 +1 , ..., xcn − Ln )
g̃(x̃) := ..
. .
gn0 (x̃1 , ..., x̃n0 , xcn0 +1 − Ln0 +1 , ..., xcn − Ln )
Due to Case 1, we can conclude that there exists some x̃∗ ∈ G̃ with g̃(x̃∗ ) = 0.
Setting
x̃∗1
..
.
∗
x̃n
x∗ := c
0
xn0 +1 − Ln0 +1
..
.
xcn − Ln
we have x∗ ∈ G and g(x∗ ) = 0 by (4.8).
38
4.2 A Fixed Point Version
and if Li was replaced by ai . Of course, the ≤-sign is more general than the
<-sign, but it is not clear, why Miranda changed
“... for xi = ai , fi is constantly positive ...”
into
fi (x1 , x2 , ..., xi−1 , +Li , xi+1 , ..., xn ) ≤ 0.
Of course, concerning the existence of a zero, this is irrelevant thanks Corollary
4.2. However, it is worth mentioning, that one also gets the existence of a fixed
point, when using the original notation of Miranda. This will be shown in the
following theorem.
Theorem 4.2. Let Ω = {x ∈ Rn : |xi | ≤ Li , i = 1, ..., n}. Furthermore, let
g : Ω → Rn be continuous satisfying
)
gi (x1 , ..., xi−1 , −Li , xi+1 , ..., xn ) ≥ 0
for i = 1, ..., n.
gi (x1 , ..., xi−1 , +Li , xi+1 , ..., xn ) ≤ 0
fi (x1 , , .., xi−1 , −Li , xi+1 , .., xn ) = gi (x1 , .., xi−1 , −Li , xi+1 , .., xn ) + Li ≥ 0,
fi (x1 , , .., xi−1 , +Li , xi+1 , .., xn ) = gi (x1 , .., xi−1 , +Li , xi+1 , .., xn ) − Li ≤ 0.
Then, from Theorem 4.1, we can conclude that there exists x∗ ∈ Ω satisfying
f (x∗ ) = 0. Therefore, g(x∗ ) = x∗ .
Analogous corollaries for this theorem (such as Corollary 4.1 and Corollary 4.2
for Theorem 4.1) are not true as we can see in Figure 4.1 and Figure 4.2, re-
spectively.
39
4 The Poincaré-Miranda Theorem
-L L x
40
4.2 A Fixed Point Version
xc -L xc +L x
41
5 The Nonlinear Complementarity
Problem
z ≥ 0, f (z) ≥ 0, z T · f (z) = 0,
or to show that no such vector exists. Here, the inequality sign is meant com-
ponentwise; i.e.,
z ≥ 0 ⇔ z1 ≥ 0, ..., zn ≥ 0
f (z) ≥ 0 ⇔ f1 (z) ≥ 0, ..., fn (z) ≥ 0.
In 1974, Tamir published an algorithm for solving the NCP for the case that f is
a so-called Z-function. Tamir’s algorithm is a generalization of Chandrasekaran’s
algorithm which solves the linear complementarity problem for the case that the
given matrix M is a so-called Z-matrix.
Proving his results, Tamir used the iterative processes of Gauss-Seidel and of
Jacobi. In Section 5.4 we present a different proof where these iterative processes
are not used. Instead, we use the least element theory and the Poincaré-Miranda
theorem.
Before we are going to show how the Poincaré-Miranda theorem can be used to
prove the correctness of Tamir’s algorithm, we give an application of the NCP
where Tamir’s algorithm can be used.
43
5 The Nonlinear Complementarity Problem
One can show that (5.1) has a unique solution, say {ŝ, ẑ(x)}, and that
√
ŝ ≤ 2,
ẑ(x) > 0, x ∈ [0, ŝ).
See (83), (99). However, it is not possible to solve (5.1) explicitly. So, one is
satisfied to get approximations for ẑ(x) at discrete points and to get an approx-
imation for ŝ.
√
To show this, we choose n ∈ N and subdivide the interval [0, 2] equidistantly;
i.e., we define
√
2
h := , xi := i · h, i = 0, 1, ..., n + 1.
n+1
Then, we have )
ẑ(xi ) > 0, if xi ∈ [0, ŝ),
√ (5.2)
ẑ(xi ) = 0, if xi ∈ [ŝ, 2].
44
5.1 Free Boundary Problems
45
5 The Nonlinear Complementarity Problem
Tamir’s algorithm presented in the next section was developed for the NCP
defined by a so-called Z-function. So, we start with its definition.
The function given in (5.5) is a Z-function. We’ll see that Tamir’s algorithm can
be used to solve the discretized free boundary problem.
46
5.3 Tamir’s Algorithm
xi ≤ yi ⇒ gi (z) ≥ gi (x) ≥ 0,
yi < xi ⇒ gi (z) ≥ gi (y) ≥ 0.
Hence,
inf{x, y} ∈ S. (5.6)
Now, let i ∈ {1, ..., m} be fixed but arbitrary. We define
mn
Note, that not every bounded nonempty set S ⊂ Rn has a least element. See
Figure 5.1.
Definition 5.2. Let S ⊆ Rn be a nonempty set. Then, the greatest lower bound
of S is denoted by inf S. If z := inf S, then it holds
and
if v ≤ t for all t ∈ S then v ≤ z.
47
5 The Nonlinear Complementarity Problem
x2 6
@
@
@
@ S
@
r @
z -
x1
Tamir’s algorithm is given in Table 5.1, where Rk+ denotes the first orthant of
Rk ; i.e., Rk+ = {x ∈ Rk : xj ≥ 0, j = 1, ..., k}. We want to remark that the
pseudocode in Table 5.1 is not the original pseudocode presented by Tamir. We
have removed the modified Jacobi process. Instead, we use the lines 5-7.
−2 −2 −9 9 z4 −50 + z43
We will use Tamir’s algorithm to solve the NCP defined by this Z-function. At
the beginning, we have
22
−12
k = 0, z (0) = 0, f (z (0) ) =
15 ≥
6 0.
−50
48
5.3 Tamir’s Algorithm
begin
k := 0 ; z := 0 ; J := ∅ ;
if f (z) ≥ 0 then goto 10
else repeat k := k + 1;
choose ik ∈ {1, ..., n} with fik (z) < 0 ;
J := J ∪ {ik } ;
..., ik } and g (k) :
let J = {i1 , Rk+ → Rk be defined as
k
P
fi1 ( tj eij )
t1 j=1
.. ..
. 7→ .
;
k
tk P
fik ( tj eij )
j=1
5: let M (k) := {t ∈ Rk+ : g (k) (t) = 0, tj ≥ zij , j = 1, ..., k − 1};
6: if M (k) 6= ∅ then
k
P (k)
7: begin t(k) := inf M (k) ; z := tj eij end
j=1
else begin write(’NCP(f ) has no solution’); goto 20 end;
until f (z) ≥ 0 ;
10: write(’The solution is ’,z);
20: end.
Note that g (1) (1) = 0 and the function g (1) (t) is strictly increasing, since
d (1)
g (t) = 3t2 + 11 > 0.
dt
This implies M (1) = {1}. Going on with Tamir’s algorithm, we get
0 19
1 0
t(1) = 1, z (1) = (1)
0 , but f (z ) = 10 6≥ 0.
0 −52
This leads to i2 = 4 and J = {2, 4}. Now, Tamir’s algorithm considers the
function
g (2) : R2 → R2
(2) 11 −6 t1 −12 + t31
g (t1 , t2 ) := + .
−2 9 t2 −50 + t32
49
5 The Nonlinear Complementarity Problem
Now, we want to show that M (2) is a singleton. So, assume that g (2) ((a, b)) = 0
with a ≥ 1. Then,
a3 + 11a − 6b − 12 = 0,
b3 + 9b − 2a − 50 = 0.
Substituting
a3 + 11a − 12
b= (5.7)
6
in the second equation, we get
3
a3 + 11a − 12 9
h(a) := + (a3 + 11a − 12) − 2a − 50 = 0.
6 6
Since
2
a3 + 11a − 12 3a2 + 11 9 2 87
h0 (a) = 3 · · + a + > 0,
6 6 2 6
the function h(a) is strictly increasing. So, there can be only one a with h(a) = 0.
(2)
Then, via(5.7),
also b is unique. As a result, M is a singleton and we have
(2) 2
M ={ }. So, we define
3
0
2 2
t(2) = and z (2) =
0 .
3
3
Since
13
0
f (z (2) ) =
2 ≥ 0,
0
0
2
the algorithm stops with a solution z = 0 of the NCP.
3
50
5.4 The Correctness of Tamir’s Algorithm
Proof: We consider the sets S (k) := {t ∈ Rk+ : g (k) (t) ≥ 0}. Since f is a
Z-function, all g (k) are Z-functions, too. Since M (k) ⊆ S (k) , the condition
M (k) 6= ∅ implies S (k) 6= ∅, and it follows by Lemma 5.1 that every S (k) has a
least element, say t(k) . First, we will show by induction, that g (k) (t(k) ) = 0.
k = 1 : With fi1 (0) < 0 and ∅ 6= M (1) the assertion follows by the intermediate-
value theorem and the continuity of f .
k−1 k: Let t(k−1) be the least element of S (k−1) , let
!
k−1
P (k−1)
fi1 tj eij
j=1
..
o = g (k−1) (t(k−1) ) =
. , (5.8)
!
k−1
P (k−1)
fik−1 tj eij
j=1
and let
(k−1) k−1
(k) t X (k−1)
0 > gk ( ) = fik tj eij . (5.9)
0
j=1
By Lemma 5.1, S (k) has a least element t(k) . Since g (k−1) is a Z-function, for all
j ∈ {1, ..., k − 1}
Xk k−1
X (k)
(k)
0 ≤ fij tj eij ≤ fij tj eij .
j=1 j=1
(k) (k)
Therefore, (t1 , ..., tk−1 )T ∈ S (k−1) , whence
(k) (k−1)
t1 t1
. ..
. ≥ inf S (k−1) = . (5.10)
. .
(k) (k−1)
tk−1 tk−1
Since t(k) ∈ S (k) , g (k) (t(k) ) ≥ 0. In order to prove g (k) (t(k) ) = 0, we assume that
(k)
there exists some j ∈ {1, ..., k} with gj (t(k) ) > 0. By (5.10), (5.8), and (5.9)
51
5 The Nonlinear Complementarity Problem
we then have
(k−1)
k k−1
(k) t X (k)
X (k−1)
tj = ⇒ 0 < fij tj eij ≤ fij tj eij ≤ 0;
0 j j=1 j=1
(k) t(k−1)
it follows that tj > . So, we choose δ > 0 so that
0 j
(k) t(k−1) (k)
tj −δ > and gj (t(k) − δ · ej ) > 0.
0 j
t(k) ≤ z.
52
5.4 The Correctness of Tamir’s Algorithm
Hence, (zi∗1 , ..., zi∗k−1 )T ∈ S (k−1) and it follows that (zi∗1 , ..., zi∗k−1 )T ≥ t(k−1) .
Now, we consider the function g (k) on the product of intervals
(k−1) ∗
[t1 , zi1 ]
..
.
Ω= .
[t(k−1) , z ∗ ]
k−1 ik−1
[0, zi∗k ]
(k−1)
Let t ∈ Ω; i.e., tj ∈ [tj , zi∗j ] , j = 1, ..., k − 1 and tk ∈ [0, zi∗k ]. Then, for all
l ∈ {1, ..., k − 1} we have
k−1
X k−1
X
(k−1) (k−1)
fil tj eij + tl eil + tk eik ≤ fil tj eij = 0
j=1, j6=l j=1
by (5.8) and
k−1
X k
X
fil tj eij + zi∗l eil + tk eik ≥ fil zi∗j eij ≥ fil (z ∗ ) ≥ 0.
j=1, j6=l j=1
53
5 The Nonlinear Complementarity Problem
Finally,
k−1
X k−1
X (k−1)
fik tj eij + 0 · eik ≤ fik tj eij < 0
j=1 j=1
by (5.9) and
k−1
X
fik tj eij + zi∗k · eik ≥ fik (z ∗ ) ≥ 0.
j=1
Since f is continuous, g (k) is continuous, too. So, by Corollary 5.1 applied to g (k)
and Ω there must be some ξ ∈ Ω satisfying g (k) (ξ) = 0. This is a contradiction
to the assumption that M (k) = ∅ holds.
Example 5.3. We continue Example 5.1. F. Wilhelm has shown that f (z),
z ≥ 0 is injective. As a result, applying Tamir’s algorithm for solving the NCP,
all sets M (k) are either empty or a singleton. In contrast to the original paper
of Tamir, the method for calculating a zero of g (k) is not fixed in Table 5.1. So,
it is left to the programmer which method for calculating a zero is chosen.
The results presented in Table 5.2 are based on an implementation of D. Ham-
mer. The input data are n as well as the tolerance ε > 0. As the method for
calculating a zero of g (k) , Newton’s method was chosen, where
0 if k = 1
(k−1)
tstart := t
if k > 1
0
was taken as the starting point, respectively. If zi = 0 and zi+1 > 0, then
s̃ := 12 (xi + xi+1 ) was taken as an approximation for ŝ. See Table 5.2 for some
examples. Note, that the exact value of ŝ satisfies ŝ ∈ [1.393206, 1.397715]; see
(84).
54
6 Verification Methods
Many numerical methods for calculating a zero of a function f : Rn → Rn
involve calculating an approximate solution, say x̃. These methods often consist
of iterative algorithms using the stopping criterion
kf (x̃)k∞ < ε, (6.1)
where, here, kf (x̃)k∞ := max{|f1 (x̃)|, ..., |fn (x̃)|} and where ε > 0 is a given
tolerance. However, the criterion (6.1) is not sufficient in order to conclude that
there is a solution of f (x) = 0.
Example 6.1. Let f : R2 → R2 defined by
! !
f1 (x, y) min{x, − 21 x + 12 y − ε}
f (x, y) = = .
f2 (x, y) min{y, − 25 x + 25 y − 3ε}
Assume, an approximate solution is given by
(x̃, ỹ) = (0, ε).
1
Then, kf (x̃, ỹ)k∞ = 2ε < ε. However, a zero of f does not exist. Assume,
f (x∗ , y ∗ ) = (0, 0). Then, the first component gives
1 1
x∗ = 0 or − x∗ + y ∗ − ε = 0.
2 2
Case 1: Let x∗ = 0. Then, y ∗ ≥ 2ε > 0. Substituting this result in f2 (x∗ , y ∗ ) =
0, we get a contradiction; after all,
5 5 5
0 = f2 (x∗ , y ∗ ) = min{y ∗ , − · 0 + · y ∗ − 3ε} ≥ min{2ε, · 2ε − 3ε} = 2ε > 0.
2 2 2
Case 2: Let − 21 x∗ + 12 y ∗ − ε = 0 (and x∗ > 0). It follows
y ∗ = x∗ + 2ε > 0.
Substituting this result within f2 (x∗ , y ∗ ) = 0, again we get a contradiction
0 = f2 (x∗ , y ∗ ) = min{y ∗ , − 25 x∗ + 25 y ∗ − 3ε}
= min{x∗ + 2ε, − 25 x∗ + 5
2 · (x∗ + 2ε) − 3ε} = 2ε > 0.
55
6 Verification Methods
Considering (6.1), we only can guess that in the so-called interval vector
[x̃1 − ε, x̃1 + ε]
..
.
[x̃n − ε, x̃n + ε]
The basic idea is very simple: Check by use of a computer that (4.5) is valid
for
[x̃1 − ε, x̃1 + ε]
Ω=
..
.
.
[x̃n − ε, x̃n + ε]
Then, the corollaries of the Poincaré-Miranda theorem prove that there is some
x∗ ∈ Ω with f (x∗ ) = 0.
Another idea is to transform the problem f (x) = 0 into a fixed point problem
and to verify by use of a computer the assumptions of Brouwer’s fixed point
theorem.
Especially we do this in the case that the interval is described by a long term,
see Section 6.6.
56
6.1 Interval Computations
Furthermore, let IR denote the set of all real compact intervals. Then, we can
define +, −, ·, / in the set theoretic sense; i.e.,
[a] + [b] = {a + b : a ∈ [a], b ∈ [b]},
[a] − [b] = {a − b : a ∈ [a], b ∈ [b]},
[a] · [b] = {a · b : a ∈ [a], b ∈ [b]},
[a] a
= { : a ∈ [a], b ∈ [b]} if 0 6∈ [b].
[b] b
Example 6.2. Let [a] = [1, 3] and [b] = [2, 4]. Then,
[a] + [b] = {a + b : a ∈ [1, 3], b ∈ [2, 4]} = [3, 7],
[a] − [b] = {a − b : a ∈ [1, 3], b ∈ [2, 4]} = [−3, 1],
[a] · [b] = {a · b : a ∈ [1, 3], b ∈ [2, 4]} = [2, 12],
[a] a 1 3
= { : a ∈ [1, 3], b ∈ [2, 4]} = [ , ].
[b] b 4 2
Following Example 6.2, it is easy to see that the results of [a] ∗ [b], where ∗ ∈
{+, −, ·, /} can be expressed by the interval boundaries of [a] and [b]. More
precisely,
[a] + [b] = [a + b, a + b],
[a] − [b] = [a − b, a − b],
[a] · [b] = [min{a · b, a · b, a · b, a · b}, max{a · b, a · b, a · b, a · b}] (6.2)
[a] 1 1
= [a] · [ , ] if 0 6∈ [b].
[b] b b
So, for any a ∈ [a] and any b ∈ [b] we have
a ∗ b ∈ [a] ∗ [b], ∗ ∈ {+, −, ·, /}.
(For the case ∗ = / we have to assume, that 0 6∈ [b]). This means, that any term
a ∗ b is included within the interval [a] ∗ [b]. For example,
1
+ π ∈ [0.333, 0.334] + [3.141, 3.142] = [3.474, 3.476].
3
The smaller the interval will be, the better an approximation of the exact value
can be propagated. But interval arithmetic is not as easy as it seems. Some
rules, that you are used to know in R, are not valid in IR. For example, the
distributivity law, which reads
a · (b + c) = a · b + a · c,
57
6 Verification Methods
is not valid in IR. Let [a] = [1, 2], [b] = [−1, 1], and [c] = [2, 4]. Then,
whereas
[a] · [b] + [a] · [c] = [−2, 2] + [2, 8] = [0, 10].
One can show, that only the so-called sub-distributivity law
is valid in IR. We refer to the book of G. Alefeld and J. Herzberger for a detailed
introduction into interval computations.
The main thing, that we need in the following sections, is the meaning of a
so-called interval evaluation.
Example 6.3. Let f (x) = x3 + 2x2 + x + 2 and let [x] = [1, 2]. Then, an interval
evaluation is defined via the basic rules (6.2) as follows:
Clearly, if there is some unknown number, say η, that we can only include in an
interval, say η ∈ [1, 2], then we have f (η) ∈ f ([1, 2]).
1
h := , xi := i · h, i = 0, 1, ..., n + 1.
n+1
58
6.1 Interval Computations
Let ŷ(·) denote the solution, then Taylor’s formula with remainder term leads
to
1 1 1
ŷ(xi + h) = ŷ(xi ) + h · ŷ 0 (xi ) + h2 · ŷ 00 (xi ) + h3 · ŷ 000 (xi ) + h4 · ŷ 0000 (ξi )
2 6 24
and
1 1 1
ŷ(xi − h) = ŷ(xi ) − h · ŷ 0 (xi ) + h2 · ŷ 00 (xi ) − h3 · ŷ 000 (xi ) + h4 · ŷ 0000 (νi )
2 6 24
with some νi ∈ (xi − h, xi ) and some ξi ∈ (xi , xi + h). Adding both equations
we get
1 1
ν1 ∈ [x1 − h, x1 ] = [0, ], ξ1 ∈ [x1 , x1 + h] = [ , 1].
2 2
Formula (6.5) then reads
2
−ŷ(0) + 2 · ŷ( 12 ) − ŷ(1) 1 1 1
= e2 + eξ1 + eν1 .
1 2 24 2
2
59
6 Verification Methods
60
6.1 Interval Computations
The idea to use Taylor’s formula with remainder term and to use interval com-
putations in order to get inclusions of the exact solution at discrete points, was
presented by E. Hansen and R. Moore. It was applied to problem (6.3), where
the existence of a unique solution is given, but where this solution cannot be
stated explicitly as in our simple example (6.4). For the technical details we
refer to (43), for example.
By IRm×n , we denote the set of all interval matrices with m columns and n
rows. Then, with [A], [B] ∈ IRn×n , we define
and
n
X
[A] · [B] = ( [aik ] · [bkj ]),
k=1
Then, f ([x], [y]) = [−2, 1] · [1, 2]2 = [−2, 1] · [1, 4] = [−8, 4].
Finally, we want to note that not every function has an interval evaluation.
61
6 Verification Methods
f (y) − f (x(0) ) = f 0 (ξ) · (y − x(0) ) for some ξ between x(0) and y. (6.6)
If f (y) = 0, then
0 = f (x(0) ) + f 0 (ξ) · (y − x(0) ). (6.7)
Multiplying (6.7) by some r 6= 0, we get
In order to transform (6.8) into a fixed point equation, we add and subtract the
term y − x(0) , getting
where
∂f1
∂f1
∂x1 x(0) + λ1 · y − x(0) ··· ∂xn x(0) + λ1 · y − x(0)
B(x(0) , y) =
.. ..
. .
∂fn ∂fn
∂x1 x(0) + λn · y − x(0) ··· ∂xn
(0)
x + λn · y − x (0)
1
with some λ1 , ..., λn ∈ (0, 1). Note that, in general, the λi will all be distinct.
1 See Section 3.2 in (68).
62
6.2 The Krawczyk Operator
Now, let R ∈ Rn×n be a nonsingular matrix, let I denote the unit matrix, let
[x] ∈ IRn×n be an interval vector, let y ∈ [x], and let x(0) ∈ [x] be arbitrary but
∂fi
fixed. If ∂x j
([x]) can be evaluated for i, j = 1, ..., n as described in Section 6.1,
then we have
B(x(0) , y) ∈ f 0 ([x]),
where
∂f1 ∂f1
∂x1 ([x]) ··· ∂xn ([x])
f 0 ([x]) =
.. ..
. . ,
∂fn ∂fn
···
∂x1 ([x]) ∂xn ([x])
because x(0) + λi · y − x(0) ∈ [x] for i = 1, ..., n, since [x] is convex. As a result,
the Krawczyk operator can be defined as
Another technique for finding a matrix B(x(0) , y) satiyfying (6.9) goes back to
E. Hansen. For this, it is sufficient that the function f is continuous. Because
of
(0)
fi (y1 , . . . , yn ) − fi x1 , . . . , x(0) n =
(0) (0)
fi (y1 , . . . , yn ) − fi x1 , y2, . . . , yn + fi x1 , y2, . . . , yn
(0) (0) (0) (0)
−fi x1 , x2 , y3 , . . . , yn + fi x1 , x2 , y3 , . . . , yn
(0) (0) (0) (0) (0)
± · · · − fi x1 , . . . , xn−1 , yn + fi x1 , . . . , xn−1 , yn − fi x1 , . . . , x(0) n =
n fi x(0) , . . . , x(0) , yj , . . . , yn − fi x(0) , . . . , x(0) , yj+1 , . . . , yn
X 1 j−1 1 j (0)
(0)
· y j − x j
j=1 yj − xj
63
6 Verification Methods
The difference between f 0 ([x]) and H(x(0) , [x]) will be presented in the following
example.
Example 6.7. Let f (x) = x2 , x ∈ [x] ⊆ R. Then,
For example, if [x] = [3, 5] and x(0) = 4, then Hansen’s method yields
In Section 6.2 we have presented two possibilities for finding such an interval
matrix. Independently of the choice of Y (x(0) , [x]), the Krawczyk operator is
defined as
where I denotes the unit matrix, R ∈ Rn×n is a nonsingular matrix, and where
x̂ ∈ [x] is arbitrary, but fixed.
64
6.4 The Moore-Kioustelidis Test
whence
y − R · f (y) = x̂ − R · f (x̂) + (I − R · Y (x̂, y)) · (y − x̂).
Considering the continuous function
r(y) := y − R · f (y), y ∈ Rn ,
According to Brouwer’s fixed point theorem, there exists some x∗ ∈ [x] with
r(x∗ ) = x∗ . Since R is nonsingular, we have f (x∗ ) = 0.
The check for the validity of (6.11) is called Moore test. A way to find an interval
vector [x], for which it is likely to verify (6.11), has been published by G. Alefeld,
A. Gienger, and F. Potra in 1994.
65
6 Verification Methods
Applying the Poincaré-Miranda theorem, one could try to verify for all i =
1, ..., n
sup fi ([x]i,+ ) ≤ 0 ≤ inf fi ([x]i,− ) (6.13)
or
sup fi ([x]i,− ) ≤ 0 ≤ inf fi ([x]i,+ ), (6.14)
using the interval extensions explained in Section 6.1. But it is very unlikely
that this works as we will see in the following example.
A zero is given by (x∗ , y ∗ )T = (1, 0)T . However, for any interval vector [x] ∈ IR2
with (x∗ , y ∗ )T as its center, neither (6.13) nor (6.14) can be valid. To show this,
let
[1 − ε, 1 + ε]
[x] = with arbitrarily small ε > 0, δ > 0.
[−δ, δ]
Then, for any
x 1,+ 1+ε
∈ [x] =
y [−δ, δ]
we get
f1 (x, y) = f1 (1 + ε, y) = (1 + ε) · y, y ∈ [−δ, δ].
On the other hand, for any
x 1−ε
∈ [x]1,− =
y [−δ, δ]
we get
f1 (x, y) = f1 (1 − ε, y) = (1 − ε) · y, y ∈ [−δ, δ].
Thus,
δ δ
f1 (1 + ε, ) · f1 (1 − ε, ) > 0.
2 2
Therefore, the Poincaré-Miranda theorem cannot be applied.
66
6.4 The Moore-Kioustelidis Test
Proof: Let i ∈ {1, ..., n} be arbitrary, but fixed. Setting x(0) = x̂ + ti ei in (6.10),
then for every x ∈ [x]i,+ it holds
f (x) = f (x̂ + ti ei ) + Y (x̂ + ti ei , x) · (x − x̂ − ti ei ).
After multiplying by R, we get
g(x) = g(x̂ + ti ei ) + (R · Y (x̂ + ti ei , x)) · (x − x̂ − ti ei )
67
6 Verification Methods
Since
[−s1 , t1 ]
..
.
[−si−1 , ti−1 ]
i,+
[x] − x̂ − ti ei =
0 ,
[−si+1 , ti+1 ]
..
.
[−sn , tn ]
it follows that gi (x) ∈ [l]i,+ for all x ∈ [x]i,+ . Analogously, we get gi (y) ∈ [l]i,−
for all y ∈ [x]i,− . But (6.15) and (6.16) assure us that
By Corollary 4.2 it follows that there exists some x∗ ∈ [x] with g(x∗ ) = 0. Since
R is nonsingular, f (x∗ ) = 0 as well.
The check for validity of (6.15) and (6.16) is called Moore-Kioustelidis test. A
comparison to the Moore test will be given in Section 6.6. Beforehand we present
another verification test also based on the Poincaré-Miranda theorem.
If for i = 1, ..., n
sup[m(x̂)]i,− ≤ 0 ≤ inf[m(x̂)]i,+ (6.17)
or
sup[m(x̂)]i,+ ≤ 0 ≤ inf[m(x̂)]i,− , (6.18)
∗ ∗
then there exists some x ∈ [x] with f (x ) = 0.
68
6.6 A Comparison of the Tests
and
gi (y) ∈ [m(x̂)]i,− for all y ∈ [x]i,− .
Calling on Corollary 4.2, (6.17), and (6.18) we conclude that there exists some
x∗ ∈ [x] with g(x∗ ) = 0. Since R is nonsingular, it follows that f (x∗ ) = 0.
The check for the validity of (6.17) and (6.18) is called the Frommer-Lang-
Schnurr test.
Proof: Suppose that the Moore test is successful. Then, considering the ith
component one obtains
xi ≤ inf K(x̂, R, [x])i = inf x̂ − R · f (x̂) + (I − R · Y (x̂, [x])) · ([x] − x̂) .
i
and since for j 6= i we have [xj ]i,− = [xj ], we can conclude that K(x̂, R, [x])i =
n
X
x̂i − gi (x̂) + (I − R · Y (x̂, [x]))ii · ([xi ] − x̂i ) − (R · Y (x̂, [x]))ij · ([xj ]i,− − x̂j ).
j=1,j6=i
Furthermore, due to
we have
inf K(x̂, R, [x])i ≤
n
X
inf x̂i −gi (x̂)+(I −R·Y (x̂, [x]))ii ·(xi − x̂i )− (R·Y (x̂, [x]))ij ·([xj ]i,− − x̂j ) .
j=1,j6=i
69
6 Verification Methods
From the distributive law [a] · γ + [b] · γ = ([a] + [b]) · γ holding for the scalar
γ = xi − x̂i and the fact that xi = [xi ]i,− it follows that
xi ≤ inf K(x̂, R, [x])i ≤ inf xi − [m(x̂)]i,− = xi − sup[m(x̂)]i,− .
Thus,
sup[m(x̂)]i,− ≤ 0.
In the same way, one can show that
implies 0 ≤ inf[m(x̂)]i,+ .
The Moore test and the Frommer-Lang-Schnurr test require roughly the same
computational work. Therefore, the Frommer-Lang-Schnurr test is more power-
ful than the Moore test by Theorem 6.4.
Comparing the Moore-Kioustelidis test with the Frommer-Lang-Schnurr test,
it is obvious that the latter requires less computational work, since there the
function g(·) and the interval (slope) matrix Y (·, ·) are only evaluated once.
However, the following theorem will present a special situation, where the Moore-
Kioustelidis test is more powerful than the Frommer-Lang-Schnurr test.
Theorem 6.5. Let f : D → Rn , D ⊆ Rn be differentiable and let [x] ⊆ D.
Furthermore, let
Y (x̂, [x]) = f 0 ([x]).
Then, concerning Theorems 6.3 and 6.2, we have
This means, if (6.17) from Theorem 6.3 is valid, then the condition (6.15) from
Theorem 6.2 is fulfilled; analogously, if (6.18) from Theorem 6.3 is valid, then
the condition (6.16) from Theorem 6.2 is fulfilled.
whence
gi (x̂ + ti ei ) ∈ gi (x̂) + ti · R · Y (x̂, [x]) = gi (x̂) + ti · R · f 0 ([x]) . (6.20)
ii ii
70
6.6 A Comparison of the Tests
Since, in general, we have for [a], [b] ∈ IR and any γ ∈ R, that γ ∈ [a] implies
γ + [b] ⊆ [a] + [b], we get from (6.20)
n
X
gi (x̂ + ti ei ) + R · f 0 ([x]) · [−sj , tj ] ⊆ [m(x̂)]i,+ . (6.21)
ij
j=1, j6=i
Using f 0 ([x]i,+ ) ⊆ f 0 ([x]) we finally get [l]i,+ ⊆ [m(x̂)]i,+ . Analogously, one can
show that [l]i,− ⊆ [m(x̂)]i,− .
The following figure gives a summary for the three verification methods pre-
sented, where A ⇒ B means that the success of test A implies the success of
test B.
f1 (y1 , y2 ) − f1 (x̂1 , y2 )
=
y1 − x̂1
71
6 Verification Methods
Concerning [x] = ([−0.1, 0.1] [−0.1, 0.1])T and x̂ = 0, we compare the Moore
test with the Moore-Kioustelidis test. In both tests we use
0 −1 1
0
R = f (0) = 0.0525 .
0 1
Then,
f1 (y1 , y2 ) − f1 (0, y2 ) 2
= 4y22 − (y1 + 0.0125) + 0.0525 + 0.01252.
y1 − 0
Therefore, we have
0.04 , 0.0925 + 0.01252 [−0.01, 0.01]
Y (0, [x]) =
0 1
and
h i 0.01 0.01 !
0.0925+0.01252 0.04
1− 0.0525 ,1 − 0.0525 − 0.0525 , 0.0525
I − R · Y (0, [x]) =
0 0
[−0.77 , 0.24] [−0.2, 0.2]
⊆ .
0 0
Hence,
[−0.1, 0.1]
K (0, R, [x]) = 0 − R · f (0) + (I − R · Y (0, [x])) ·
[−0.1, 0.1]
[−0.077, 0.077] + [−0.02, 0.02]
⊆ ⊆ [x] .
−0.01
Therefore, the Moore test is successful, and due to Theorem 6.4 also the Frommer-
Lang-Schnurr test is successsful to verify that a zero x∗ of f is within [x]. On
the other hand, Theorem 6.2 says that
1,+ 0.1 0.1
[l] = R · f (0.1, 0) + R · Y ( , ) · [−0.1, 0.1]
1 0 [−0.1, 0.1] 12
1 1 1
= · 4 · 10−3 + · 0.5 · [−0.1, 0.1] · [−0.1, 0.1] = [−2, 18].
0.0525 0.0525 105
Thus, the Moore-Kioustelidis test fails.
Remark 6.1. Being aware of the fact that Brouwer’s fixed point theorem as
well as the Poincaré-Miranda theorem can be proved (very briefly) by using the
so-called degree of a mapping (see the book of J. Cronin and the paper of M.
N. Vrahatis, respectively), one might get the idea to verify the existence of a
zero directly by using the degree of a mapping and interval arithmetic. This was
done by A. Frommer, F. Hoxha, and B. Lang.
72
6.7 Final Remarks
This means, actually we have to consider functions f (x1 , ..., xn ; a1 , ..., am ) with
(x1 , ..., xn )T ∈ D, where D is the domain of the function and where a1 ,...,am
are parameters, that can vary in given intervals. Indeed, it is easy to extend
the corresponding theorems to this generalized case. However, the hierarchy
presented in the preceding section is no longer true as we will see in the following
example.
73
6 Verification Methods
Taking (
2x − 1, x≥1
f (x; [a]) =
[a] · x − [a] + 1, x < 1
and
Y ([x]; [a]) = convex hull of {2, [a]} = [2, 4],
we want to verify that within the interval [x] = [ 12 , 2] there is a zero of f .
Choosing R 6= 0 and x̂ = 1 we have by Theorem 6.3
and
1
[m(1)]− = R · 1 + R · [2, 4] · ( − 1) = R · [−1, 0],
2
whereas by Theorem 6.2 we have
1
[l]− = R · ([a] · − [a] + 1) = R · [−2, 1].
2
This means, that the Frommer-Lang-Schnurr test is successful, whereas the
Moore-Kioustelidis test fails.
For a hierachy in the general case we refer to (82). See also (80) for some
numerical examples.
For more examples where interval arithmetic is used to prove conjectures via
the computer we refer to (37).
74
7 Banach’s Fixed Point Theorem
Using Brouwer’s fixed point theorem one can attack the problem of showing the
existence of a fixed point. As already mentioned in Chapter 2 a fixed point
verified by Brouwer’s theorem is not necessarily unique. See Figure 2.2. The
question naturally arises how the assumptions can be strengthened so that a
verified fixed point is unique.
To get the idea, let x∗ be a fixed point of f : [a, b] → [a, b]. In order to guarantee
that x∗ is the only fixed point, the problem f (x) = x must not have a second
solution. If there was a second solution, say x̃, x̃ 6= x∗ it would follow that
x̃ − x∗ f (x̃) − f (x∗ )
1= ∗
= .
x̃ − x x̃ − x∗
If we assume f to be differentiable, the mean-value theorem says that there
exists ξ between x̃ and x∗ so that
f (x̃) − f (x∗ )
= f 0 (ξ).
x̃ − x∗
So, if f : [a, b] → [a, b] is differentiable satisfying
for some norm in Rn . (Note, however, that in (7.1) it reads f 0 (x) < 1 and not
|f 0 (x)| < 1.) So, trivially, we have the following theorem.
75
7 Banach’s Fixed Point Theorem
for all x, y ∈ K, x 6= y for some norm in Rn . Then, f has exactly one fixed
point in K.
which is a contradiction.
Of course, if the proof is so simple, there is no reason to name this result after
a mathematician. So, there must be more!
Indeed, the astonishing thing is (beyond the fact that the fixed point can be
approximated iteratively, see Theorem 7.2) the fact that the result is also true
in some vector spaces that can be infinite-dimensional. These vector spaces are
called Banach spaces. We will introduce them in the following section. In order
to emphasize this result, we mention that Brouwer’s fixed point theorem is not
valid in infinite-dimensional vector spaces.
76
7.1 Banach Spaces
4
6
2
8 6
3
1
4
2
0
0 1 2 3 4 5 6 7
Figure 7.1: Vector addition visualized in R2
x ∈ E, y ∈ E ⇒ x + y ∈ E,
(7.4)
λ ∈ R, x ∈ E ⇒ λ · x ∈ E.
Trivially, (7.4) can be verified for E = Rn for all integers n ≥ 2. More important,
however, is the fact, that we can also consider the set of all real-valued continuous
functions with domain [a, b], denoted by C([a, b]), as E. If f, g ∈ C([a, b]) and
λ ∈ R, then also
by defining
(f + g)(x) := f (x) + g(x) for all x ∈ [a, b]
and
(λ · f )(x) := λ · f (x) for all x ∈ [a, b].
Satisfying (7.4) is not the only property that the sets Rn and C([a, b]) have in
common. The idea is to find as many common properties as possible and to
prove theorems only based on these properties. Formally, a real vector space
77
7 Banach’s Fixed Point Theorem
(sometimes also called a real linear space) is a nonempty set E fulfilling (7.4)
and
x + (y + z) = (x + y) + z for all x, y, z ∈ E,
x + y = y + x for all x, y ∈ E,
there exists 0 ∈ E so that x + 0 = x for all x ∈ E,
for all x ∈ E there exists −x ∈ E so that x + (−x) = 0,
λ · (x + y) = λ · x + λ · y for all x, y ∈ E and all λ ∈ R,
(λ + ν) · x = λ · x + ν · x for all x ∈ E and for all λ, ν ∈ R,
(λ · ν) · x = λ · (ν · x) for all x ∈ E and for all λ, ν ∈ R,
1 · x = x for all x ∈ E.
A real vector space E is called a normed vector space, if there is a function
k · k : E → R satisfying
converges to x∗ and
1 qn
kx(n) − x∗ k ≤ · kx(n+1) − x(n) k ≤ · kx(1) − x(0) k.
1−q 1−q
78
7.1 Banach Spaces
Proof: First, we note that from (7.5) it follows that T is continuous. Then, we
choose arbitrarily x(0) ∈ D and consider the sequence
we get by induction
79
7 Banach’s Fixed Point Theorem
Hence,
1
kx(n) − x∗ k ≤ · kx(n+1) − x(n) k,
1−q
and (7.6) finishes the proof.
Banach’s fixed point theorem gets its full power, if one is aware of some Banach
spaces. We name just a few. (For a proof of the fact that these vector spaces
are indeed Banach spaces, we refer to any book on functional analysis).
• (Rn , k · k), where k · k can be any norm in Rn .
• (l2 , k · k), where l2 denotes the vector space of all square summable se-
quences of real numbers; i.e.,
∞
X
x= {xi }∞
i=1
2
∈ l :⇔ x2i < ∞,
i=1
and where v
u∞
uX
kxk := t x2 . i
i=1
• (l∞ , k · k), where l∞ denotes the vector space of all bounded sequences of
real numbers, and where
∞
kxk := sup |xi |.
i=1
• (c0 , k·k), where c0 denotes the vector space of all real sequences converging
to zero, and where
∞
kxk := sup |xi |.
i=1
• (C([a, b]), k·k∞ ), where C([a, b]) denotes the vector space of all real-valued,
continuous functions with domain [a, b], and where
80
7.2 Ordinary Differential Equations
A famous result by Picard-Lindelöf says that this problem has a unique solution,
if f (x, y) is continuous and if there exists L ≥ 0 so that
is valid for all x ∈ [ξ, ξ + a] and for all y, y ∈ R. Its proof can be done by
using Banach’s fixed point theorem as follows. We consider the Banach space
C([ξ, ξ + a]) with the weighted norm
Since f is continuous, y(x) is a solution of the initial value problem, if and only
if
Zx
y(x) = η + f (t, y(t))dt.
ξ
by
Zx
T (y)(x) := η + f (t, y(t))dt.
ξ
Then, the initial value problem has a unique solution, if and only if T has a
unique fixed point.
81
7 Banach’s Fixed Point Theorem
To verify the existence of a unique fixed point of T , let y, z ∈ C([ξ, ξ + a]). Then,
Rx
|T (y)(x) − T (z)(x)| = f (t, y(t)) − f (t, z(t)) dt
ξ
Rx
≤ |f (t, y(t)) − f (t, z(t))| dt
ξ
Rx
≤ L · |y(t) − z(t)| dt
ξ
Rx
= L· |y(t) − z(t)| · e−2L·t · e2L·t dt
ξ
Rx
≤ L · ky − zkw · e2L·t dt
ξ
1
≤ L · ky − zkw · 2L · e2L·x .
It follows
L
|T (y)(x) − T (z)(x)| · e−2L·x ≤ · ky − zkw for all x ∈ [ξ, ξ + a].
2L
Hence,
1
kT (y) − T (z)kw ≤
· ky − zkw .
2
Now, Banach’s fixed point theorem can be applied, and therefore T has a unique
fixed point, which is then the unique solution of the initial value problem.
The iteration y (0) (x) = η and
Zx
(n+1) (n)
y (x) := T (y )(x) = η + f (t, y (n) (t))dt, n = 0, 1, 2, 3, ...
ξ
Remark 7.1. Banach’s fixed point theorem is also called the Contraction-
Mapping Theorem, especially when E = Rn . It can also be used to prove
the Inverse Function Theorem and the Implicit Function Theorem. See (68), for
example.
Remark 7.2. In (101) splitting methods were considered for solving linear
systems of equations, say Ax = b, where A ∈ Rn×n is nonsingular and b ∈ Rn .
Expressing the matrix A in the form A = M − N with M being nonsingular is
called a splitting of the matrix A.
82
7.2 Ordinary Differential Equations
(M − N )x∗ = b,
x∗ = M −1 N x∗ + M −1 b.
Special choices of M and N lead to the well-known Jacobi, Gauss-Seidel, and the
successive overrelaxation methods. Note that no fixed point theorem is needed,
because the unique solution is guaranteed by the nonsingularity of A and we
have
83
8 Schauder’s Fixed Point Theorem
In this chapter we deal with the question if the Brouwer fixed point theorem
can be extended from Rn to other normed linear spaces. First we consider
n-dimensional normed linear spaces, second we consider infinite-dimensional
normed linear spaces.
An n-dimensional normed linear space (different from Rn ) is presented in the
following example.
Example 8.1. Let f : [0, 1] → R defined by f (x) = 1, let g : [0, 1] → R defined
by g(x) = x, and let h : [0, 1] → R defined by f (x) = x2 . Then every quadratic
function from [0, 1] to R can be expressed as a linear combination of f, g, and
h. We may write
X = span{f, g, h}.
So, using the maximum norm k · k∞ defined in Chapter 7.1 (X, k · k∞ ) is a three-
dimensional normed linear space. For instance, any solution of the ordinary
differential equation
y 000 (x) = 0, x ∈ [0, 1]
belongs to X.
As we will see in the following theorem any n-dimensional normed linear space
and Rn are more or less the same.
Theorem 8.1. Let (X, k · kX ) be an n-dimensional normed linear space. Then,
X and Rn are isomorphic; i.e., there exists a linear bijection T : X → Rn so
that both T and T −1 are continuous.
Proof: Let f (1) , f (2) , ..., f (n) be a basis of X. Then, for every f ∈ X there exist
uniquely defined x1 , x2 , ..., xn ∈ R so that
n
X
f= xj · f (j) .
j=1
n
Defining T : X → R by
x1
T (f ) = ...
xn
85
8 Schauder’s Fixed Point Theorem
xn
By Theorem 2.2 this norm is equivalent to the Euclidean norm on Rn . It follows
that there exist positive numbers α and β so that
α · kxke ≤ kxk ≤ β · kxke for all x ∈ Rn .
In other words, for x, y ∈ Rn and f, g ∈ X we have
kT −1 (x) − T −1 (y)kX = kT −1 (x − y)kX = kx − yk ≤ β · kx − yke
and
1 1
kT (f ) − T (g)ke = kT (f − g)ke ≤ · kT (f − g)k = · kf − gkX .
α α
So, T as well as T −1 are (even Lipschitz) continuous.
Proof: By Theorem 8.1 X and Rn are isomorphic; i.e., there exists a linear
bijection T : X → Rn so that both T and T −1 are continuous. Therefore,
K̃ := T (K) is nonempty, bounded, closed (due to the continuity of T ), and
convex (due to the linearity of T ) in Rn . Defining
g := T ◦ f ◦ T −1
we get g : K̃ → K̃ and we get that g is continuous. By Brouwer’s fixed point
theorem there exists y ∗ ∈ K̃ satisfying g(y ∗ ) = y ∗ . Then, x∗ := T −1 (y ∗ ) ∈ K is
a fixed point of f .
Brouwer’s fixed point theorem is not true, if one just substitutes an infinite-
dimensional normed linear space (even a very regular one) for Rn . This will be
shown in the following example.
86
8 Schauder’s Fixed Point Theorem
Example 8.2. (Kakutani, 1943) Consider the space X = l2 (Z) of all doubly-
infinite real-valued sequences {xn }∞
n=−∞ (Z denotes the set of all integers) so
that
X∞
x2n < ∞;
n=−∞
B = {x ∈ X : kxk2 ≤ 1}
is a nonempty, closed, bounded, and convex set in X. Let S(x) denote the
shifted sequence of x, that is
= kxk2 + 21 (1 − kxk2 ) = 1
2 + 1
2 · kxk2 ≤ 1.
87
8 Schauder’s Fixed Point Theorem
Since x∗ 6= 0, at least one of the numbers x∗0 and x∗1 , respectively, is not 0, in
∞
P
plain contradiction to (x∗n )2 < ∞.
n=−∞
This definition does not depend on the dimension of K. Next, we consider open
sets. Let (E, k · k) be a normed vector space and let Q ⊆ E. Q is said to be
open, if for any x ∈ Q there exists some ε > 0 so that
U (x; ε) := {y ∈ E : kx − yk < ε}
On first sight one may think that Ω is an open set. But this is not the case as
we will show now.
Let ε > 0 be arbitrary but fixed. Since L ∈ l2 , the sequence {Li }∞
i=1 has to be
convergent to 0; i.e., there exists i0 ∈ N so that
ε
Li < for all i ≥ i0 .
4
88
8 Schauder’s Fixed Point Theorem
However, it holds
3 · ε 3·ε 3·ε ε ε
|ξi0 | = + xi0 ≥ − |xi0 | ≥ − = > Li0 ,
4 4 4 4 2
U (x; ε) 6⊂ Ω;
U (x0 ; 1) := {x ∈ X : kx − x0 k ≤ 1}.
X1 := {a · x(1) : a ∈ R}.
89
8 Schauder’s Fixed Point Theorem
ϕ(a) = ky − a · x(1) k,
we have
lim ϕ(a) = ∞
|a|→∞
1
(1) (1) (1)
= ·
(y − a 1 · x ) − ky − a 1 · x k · x
ky − a1 · x(1) k
1
(1) (1)
= ·
y − (a 1 + ky − a 1 · x k) · x
≥1
ky − a1 · x(1) k
thanks to (8.3). Now, suppose x(1) , ..., x(n) ∈ X have been chosen so that
and
kx(i) − x(j) k ≥ 1 for i 6= j, i = 1, 2, ..., n, j = 1, 2, ..., n.
Then, we define
n
X
Xn := {x = tj · x(j) : t1 , ..., tn ∈ R}.
j=1
90
8 Schauder’s Fixed Point Theorem
Then,
n
X n
X n
X
kz − tj · x(j) k = k tj · x(j) − zk = kti · x(i) − (z − tj · x(j) )k.
j=1 j=1 j=1, j6=i
Hence,
n
X
(i)
ϕ(t1 , ..., tn ) ≥ |ti | · kx k − kz − tj · x(j) k
j=1, j6=i
which implies that
lim ϕ(t1 , ..., tn ) = ∞ for all i ∈ {1, ..., n}. (8.5)
|ti |→∞
91
8 Schauder’s Fixed Point Theorem
8.1 Compactness
Definition 8.1. Let X be a normed vector space. A nonempty set K ⊆ X is
called compact, if every sequence in K has a subsequence that converges to a
limit that belongs to K.
If X = Rn , then the compact sets are precisely those that are closed and
bounded; this is the celebrated theorem of Bolzano-Weierstraß. But, as we saw
in Example 8.4, this is not so for the closed unit ball of an infinite-dimensional
normed linear space, a closed, bounded set to be sure. Compactness is more spe-
cial than being just closed and bounded; indeed, it is a key ingredient exploited
by Schauder in his fixed point theorem.
We start by establishing a fundamental feature of compact sets in normed vector
spaces - the property formulated is often referred to as ’total boundedness’.
Theorem 8.2. Let X be a normed vector space and let K ⊆ X be com-
pact. Then, for every ε > 0 there exists a finite set of elements of K, say
{v (1) , ..., v (p) }, so that for any y ∈ K there exists v (i) ∈ {v (1) , ..., v (p) } with
kv (i) − yk < ε.
Proof: Assume that K fails the conclusion. That is, assume, that there is some
ε0 > 0 so that for every finite set of elements of K, say {v (1) , ..., v (n) }, there
exists some y ∈ K with
Considering the set {x(1) , x(2) } by assumption there is some x(3) ∈ K with
92
8.1 Compactness
Before we present Schauder’s fixed point theorem in the next section, we want to
give an important, classical example of a compact set in an infinite-dimensional
normed linear space, the so-called Hilbert cube.
Example 8.5. Let X = l2 and let L = {Li }∞ 2
i=1 ∈ l . Then, the set
Ω = {x ∈ l2 : |xi | ≤ Li , i = 1, 2, 3, ...}
be a sequence in Ω.
(n)
i = 1 : {x1 }∞
n=1 ⊆ [−L1 , L1 ]. Due to the theorem of Bolzano-Weierstraß there
(ϕ (n)) ∞
is a subsequence {x1 1 }n=1 and some x∗1 ∈ [−L1 , L1 ] with
(ϕ1 (n))
lim x1 = x∗1 .
n→∞
i=2: ( !)∞
(ϕ (n))
x1 1 [−L1 , L1 ]
(ϕ (n)) ⊆ .
x2 1 [−L2 , L2 ]
n=1
So, !
(ϕ (n))
x1 2 x∗1
lim = .
n→∞ (ϕ (n))
x2 2 x∗2
x∗ = {x∗i }∞
i=1 ∈ Ω (8.9)
where
ϕk (n) ⊆ ϕk−1 (n) ⊆ ... ⊆ ϕ2 (n) ⊆ ϕ1 (n) ⊆ N. (8.11)
93
8 Schauder’s Fixed Point Theorem
Concerning this k, due to (8.10) and (8.12), there exists some n0 ∈ N so that
v
u k
uX (ϕ(n)) ε
t (x
j − x∗j )2 < for all n ≥ max{n0 , k}.
j=1
2
94
8.1 Compactness
Proof: By Definition 8.2 the set T (K) is a compact set. Furthermore, since K
is closed and since T (K) ⊆ K we have
T (K) ⊆ K.
Now, let ε > 0 be arbitrary but fixed. Then, because of Theorem 8.2 there exist
v (1) , ..., v (p) ∈ T (K) so that for every k ∈ K there is some v (i) ∈ {v (1) , ..., v (p) }
satisfying kv (i) − T (k)k < ε. Therefore, considering the functions
( )
ε − kv (i) − T (k)k if kv (i) − T (k)k < ε,
mi (T (k)) := i = 1, ..., p,
0 otherwise
is well defined. Since each v (i) ∈ T (K) ⊆ K and since K is convex, it follows
that Tε (K) ⊆ K. Moreover, for every k ∈ K it holds
p
P
mi (T (k)) · kT (k) − v (i) k
i=1
kT (k) − Tε (k)k ≤ p < ε. (8.13)
P
mi (T (k))
i=1
95
8 Schauder’s Fixed Point Theorem
Since T (x(n) ) ∈ T (Kn ) ⊆ T (K) and since the set T (K) is compact, the sequence
{T (x(n) )}∞
n=1 must have a convergent subsequence, say
This results in
kx∗ − x(nl ) k = kx∗ − Tεnl (x(nl ) )k
< ε + εnl
lim x(nl ) = x∗ .
l→∞
96
8.2 Peano’s Existence Theorem
b
M := max{|f (x, y)| : (x, y) ∈ [ξ, ξ + a] × [η − b, η + b]} and α := min{a, }
M
there exists a continuously differentiable function y(x) satisfying
The proof will be done via Schauder’s fixed point theorem. Another auxiliary
result will be the theorem of Ascoli-Arzela which is based on so-called equicon-
tinuous sets of real-valued functions. So, we start with the definition of equicon-
tinuity.
Definition 8.3. Let F = {f (1) , f (2) , ...} be a set of functions where f (i) : [c, d] →
R are continuous for all f (i) ∈ F . The set F is said to be equicontinuous, if for
ε > 0 there exists some δ = δ(ε), so that for all f (i) ∈ F it holds that
Note, that for all f (i) ∈ F the same δ can be chosen. Obviously, if F = {f };
i.e., if F is a singleton, then F is equicontinuous if f is continuous. Consider the
case that F = {f (1) , f (2) } where both functions are continuous. So, for ε > 0
there exists some δ1 so that
97
8 Schauder’s Fixed Point Theorem
Example 8.6. Let F be the set of all continuous functions with domain D =
[0, 1] and let ε = 21 . Then, for every δ > 0 we can define two different values
x1 ∈ [0, 1] and x2 ∈ [0, 1] satisfying
0 < x2 − x1 < δ
and based on these two values we can define a continuous function f satisfying
f (x1 ) = 0 and f (x2 ) = 1. For example, f could be chosen as a piecewise linear
function; i.e.,
0 if x ∈ [0, x1 ],
x−x
1
f (x) = if x ∈ (x1 , x2 ],
x2 − x1
1 if x ∈ (x2 , 1].
Then, we have
1
|x2 − x1 | < δ, but |f (x2 ) − f (x1 )| = |1 − 0| = 1 6< = ε.
2
Hence, the set F is not equicontinuous.
Proof: Let A = {x1 , x2 , ...} be the set of rational numbers in [c, d]. The sequence
of real numbers {an }∞n=1 with
an := f (n) (x1 )
bn := f (pn ) (x2 )
98
8.2 Peano’s Existence Theorem
Here, {qn }∞ ∞ ∞
n=1 is a subsequence of {pn }n=1 . The third sequence {cn }n=1 with
cn := f (qn ) (x3 )
The kth row is a subsequence of the (k − 1)th row being convergent for x =
x1 , x2 , ..., xk . As a result, the diagonal sequence
)
{f (dn) (x)}∞
n=1 := f
(p1 )
(x), f (q2 ) (x), f (r3 ) (x), ...
(8.15)
is convergent for all x ∈ A.
xi ∈ Ji ∩ A.
99
8 Schauder’s Fixed Point Theorem
As a result, if x ∈ [c, d], there is some Jk with x ∈ Jk and we get |x − xk | < δ(ε),
whence via (8.16) and (8.17) it follows
|f (dm ) (x) − f (dm ) (xk )| + |f (dm ) (xk ) − f (dn ) (xk )| + |f (dn ) (xk ) − f (dn ) (x)| < ε
if m, n ≥ n0 (ε). Since x ∈ [c, d] was arbitrary, then for m, n ≥ n0 (ε) we can
conclude that
ε > max{|f (dm ) (x) − f (dn ) (x)| : x ∈ [c, d]} = kf (dm ) − f (dn) k∞ .
This means, that {f (dn) }∞n=1 is a Cauchy sequence subject to the maximum
norm. But (C([c, d]), k · k∞ ) is a Banach space, so {f (dn) }∞
n=1 is convergent in
this norm and the limit function is continuous.
T : K → K.
lim kw(n) − w∗ k∞ = 0.
n→∞
100
8.2 Peano’s Existence Theorem
Now, the theorem of Ascoli-Arzela can be applied, which means that T is com-
pact. Then Schauder’s fixed point theorem can be applied, which means that
there exists some w̃ ∈ K satisfying w̃ = T (w̃). It follows
Zx
w̃(x) = η + f (t, w̃(t))dt
ξ
101
8 Schauder’s Fixed Point Theorem
So, if n > n1 is valid, then for all f (x) ∈ f (Ω) we have some v ∈ {v (1) , ..., v (p) }
so that for all x ∈ Ω
0 0
.. ..
.
.
0 k ≤ kf (x) − vk + k 0 k ≤ 2ε.
kf (x) − h̃(n) (x)k = k
fn+1 (x) vn+1
fn+2 (x) vn+2
.. ..
. .
102
8.3 The Poincaré-Miranda Theorem in l2
and h(n) : Ωn → Rn by
f1 (x1 , x2 , ..., xn−1 , xn , x̂n+1 , x̂n+2 , ...)
h(n) (x) :=
..
. .
fn (x1 , x2 , ..., xn−1 , xn , x̂n+1 , x̂n+2 , ...)
h(n) (x(n) ) = 0,
it holds that
x̃(n) ∈ Ω and h̃(n) (x̃(n) ) = 0,
Hence, lim f (x̃(n) ) = 0, where 0 denotes the zero sequence. Since Ω is compact,
n→∞
the sequence x̃(n) has a convergent subsequence with limit x∗ ∈ Ω. W.l.o.g. we
assume that lim x̃(n) = x∗ holds. On the one hand, it follows that lim f (x̃(n) ) =
n→∞ n→∞
f (x∗ ), since f is continuous. On the other hand, it follows that f (x∗ ) = 0, since
the limit is unique.
The infinite-dimensional Poincaré-Miranda theorem can be applied to difference
equations that, in fact, are equivalent to some special infinite (denumerable)
systems of equations, and to the controllability of nonlinear repetitive processes.
See (47).
103
8 Schauder’s Fixed Point Theorem
But how can a computer, that can only represent finite many num-
bers, be used to verify solutions solving an infinite-dimensional prob-
lem?
Example 8.7. Many boundary value problems for semilinear elliptic partial
differential equations allow very stable numerical computations of approximate
solutions, but are still lacking analytical existence proofs. Plum proposes a
method which exploits the knowledge of a good numerical approximate solution,
in order to provide a rigorous proof of existence of an exact solution close to
the approximate one. This goal is achieved by Schauder’s fixed point theorem
as follows.
with α > 0 to be determined. Here, C(Ω) denotes the set of all real-valued
continuous functions with domain Ω, and kvk∞ = max{|v(x)| : x ∈ Ω}.
104
8.4 Verification Methods
The idea is to find a so-called ’weak solution’. To explain the concept of a weak
solution, we need some definitions. We define
Z
L (Ω) = {u : Ω → R : u measurable , u2 (x)dx < ∞}
2
and H := H 2 (Ω)∩ H̊ 1 (Ω), consisting of all functions u ∈ L2 (Ω) which have weak
2
∂u
derivatives ∂x i
, ∂x∂i ∂x
u
j
in L2 (Ω) (i.e., functions u for which functions
∂u ∂2u
vi =: ∈ L2 (Ω), wij =: ∈ L2 (Ω) i, j = 1, ..., n
∂xi ∂xi ∂xj
exist so that Z Z
∂ϕ
u(x) · (x)dx = − vi (x) · ϕ(x)dx (8.21)
∂xi
Ω Ω
and Z Z
∂2ϕ
u(x) · (x)dx = wij (x) · ϕ(x)dx (8.22)
∂xi ∂xj
Ω Ω
for all ϕ ∈ C0∞ (Ω))1 , and which vanish on ∂Ω in the sense that
∂u ∂ϕ(n)
ku − ϕ(n )kL2 (Ω) → 0, k − kL2 (Ω) → 0 as n → ∞ (8.23)
∂xi ∂xi
L : H → L2 (Ω)
defined by
∂f
L(u)(x) := −(4u)(x) + (x, ũ(x)) · u(x) (8.24)
∂y
has an inverse
L−1 : L2 (Ω) → H
that is bounded. Then, we define the operator
T : C(Ω) → C(Ω)
1 C ∞ (Ω)
denotes the vector space of all infinitely differentiable functions on Ω with compact
0
support in Ω.
105
8 Schauder’s Fixed Point Theorem
by
∂f
T (u)(x) := L−1 (x, ũ(x)) · u(x) − f (x, u(x)) , x ∈ Ω.
∂y
Note, that H ⊂ C(Ω). If T had a fixed point, say u∗ , then
would lead to
∂f
L(u∗ )(x) = (x, ũ(x)) · u∗ (x) − f (x, u∗ (x)), x ∈ Ω, (8.25)
∂y
which, due to (8.24), is equivalent to
∂f ∂f
− (4u∗ )(x) + (x, ũ(x)) · u∗ (x) = (x, ũ(x)) · u∗ (x) − f (x, u∗ (x)), (8.26)
∂y ∂y
x ∈ Ω; i.e.,
− (4u∗ )(x) + f (x, u∗ (x)) = 0, x ∈ Ω. (8.27)
Since T (u∗ ) ∈ H, we have that u∗ ∈ H; i.e., u∗ is a weak solution of (8.19).
(The equalities of (8.25)-(8.27), and of u∗ (x) = 0 on ∂Ω are only valid in the
sense of (8.21), (8.22), and (8.23).)
To verify a fixed point of T , Schauder’s fixed point theorem is applied based on
the observation that T is compact, since the embedding H ,→ C(Ω) is compact.
We have
T (u)(x) = L−1 ∂f ∂y (x, ũ(x)) · u(x) − f (x, u(x))
= ũ(x) − L−1 (L(ũ))(x) + L−1 ∂f ∂y (x, ũ(x)) · u(x) − f (x, u(x))
(8.24)
∂f
= ũ(x) − L−1 − (4ũ)(x) + (x, ũ(x)) · (ũ(x) − u(x)) + f (x, u(x)) .
∂y
Setting
∂f
g(x, y) := f (x, y) − f (x, ũ(x)) − (x, ũ(x)) · (y − ũ(x))
∂y
we get by Taylor’s formula
g(x, y)
→0 as y → ũ(x) (8.28)
|y − ũ(x)|
and
∂f
f (x, ũ(x)) + g(x, u(x)) = f (x, u(x)) + (x, ũ(x)) · (ũ(x) − u(x)).
∂y
106
8.4 Verification Methods
So, in order to apply Schauder’s fixed point theorem we have to find α that
satisfies
k − L−1 − 4ũ + f (·, ũ) + g(·, ũ + v) k∞ ≤ α (8.29)
T (U ) ⊆ U, (8.30)
It had been conjectured in the PDE community since the 1980’s that problem
(8.31) has at least 4 solutions for s > 0 sufficiently large. For s = 800, Breuer,
McKenna, and Plum were able to compute 4 essentially different approximate
solutions by the numerical mountain pass algorithm developed in (20), where
”essentially different” means that none of them is an elementary symmetry trans-
form of another one. With Plum’s method they could verify, by use of interval
arithmetic exploited on a computer, that in a neighborhood of each of these 4
approximate solutions, indeed there really exists a solution.
Two years after publication, Dancer and Yan gave a more general analytical
proof. They even proved that the number of solutions of problem (8.31) becomes
unbounded as s → ∞. Nevertheless, Breuer, McKenna, and Plum were first.
Other people have also considered so-called enclosure methods for the problem
(8.19) using Schauder’s fixed point theorem. We want to mention M. T. Nakao
and his co-workers. Their method avoids the computation of a bound for L−1 .
On the other hand, it requires the verified solution of large nonlinear and linear
systems, where the latter moreover have an interval right-hand side, and it
requires explicit knowledge of projection error bounds. We refer to (65).
107
9 Tarski’s Fixed Point Theorem
Continuity was the starting point to guarantee the existence of a fixed point:
) (
f : [a, b] → [a, b], there exists x∗ ∈ [a, b]
⇒
f is continuous with x∗ = f (x∗ ).
All these results exploit the fact that f is continuous. So, on first sight, it
seems implausible to get a fixed point theorem that does not assume continuity.
However, let us consider the following situation:
i.e., if x ≤ y, then f (x) ≤ f (y). Figure 9.1 shows that such a function must
have a fixed point, although the function is not continuous. It was A. Tarski,
who followed this idea. The starting point is no longer a vector space, but it is
an ordered set (in order to generalize what is meant by increasing).
x ≤ x,
x ≤ y, y ≤ x ⇒ x = y
x ≤ y, y ≤ z ⇒ x ≤ z.
109
9 Tarski’s Fixed Point Theorem
{xn }∞ ∞
n=1 ≤ {yn }n=1 :⇔ xn ≤ yn for n = 1, 2, 3, ... .
Definition 9.2. An ordered set (Ω, ≤) is said to be a complete lattice, if for any
nonempty Λ ⊆ Ω
inf Λ as well as sup Λ
exist and
inf Λ ∈ Ω as well as sup Λ ∈ Ω.
Theorem 9.1. (Tarski’s fixed point theorem, 1955) Let (Ω, ≤) be a com-
plete lattice and let Φ : Ω → Ω be a function satisfying
Λ := {ω ∈ Ω : ω ≤ Φ(ω)}.
ω ≤ ω0 for all ω ∈ Λ.
110
9 Tarski’s Fixed Point Theorem
a b x
Figure 9.1: There must be a fixed point, only due to the fact, that f is
an increasing self-mapping
As the reader might have noticed, the proof is very short and elementary. How
can it be applied? First, we need to get a hint at what we’re dealing with - what
are some pertinent examples of complete lattices?
111
9 Tarski’s Fixed Point Theorem
for all n ∈ N it holds |ωn (x, t)| ≤ Γ for all (x, t) ∈ R × [0, T ],
|ωn (x, t)| ≤ Γ for all n = 1, 2, 3, ... and for all (x, t) ∈ R × [0, T ]. (9.2)
By definition (see (9.2)) |ω̃n (x, t)| ≤ Γ for all n ∈ N and for all (x, t) ∈ R × [0, T ].
It remains to show that
So, let (x, t), (y, s) ∈ R × [0, T ], and let n ∈ N be arbitrary but fixed. Then for
all ω ∈ Λ
112
9 Tarski’s Fixed Point Theorem
a b x
Figure 9.2: No fixed point
113
9 Tarski’s Fixed Point Theorem
Finally, we want to note that Tarski’s fixed point theorem is no longer true, if
one substitutes
ω ≤ v ⇒ Φ(ω) ≤ Φ(v)
by
ω≤v ⇒ Φ(ω) ≥ Φ(v).
See Figure 9.2.
with
√0 if xm ≤ 0,
∞
fm t, {xn }n=1 = xm if 0 ≤ xm ≤ 4,
2 if 4 ≤ xm .
114
9.1 Differential Equations in Banach Spaces
t2
lim um (t) = >0 for all t ∈ (0, T ],
m→∞ 4
we would have
u(t) = {u1 (t), u2 (t), u3 (t), ...} 6∈ c0 .
Therefore, the problem (9.4) has no solution in this situation.
115
9 Tarski’s Fixed Point Theorem
with
f : R × (0, T ) × l∞ → l∞ .
I.e., the values of a solution u(x, t) have to be in l∞ . Before we attack this
problem, we need to explain what is meant by ut (x, t) and uxx (x, t).
Definition 9.3. Let E be a Banach space. Then, a function u : R × (0, T ) → E
is partially differentiable with respect to t in (x0 , t0 ) ∈ R × (0, T ), if there exists
ut (x0 , t0 ) ∈ E
so that
1
lim
· u(x0 , t) − u(x ,
0 0t ) − u t 0 0
=0
(x , t )
t→t0 t − t0
holds.
Zt Z∞ (x − ξ)2
1 −
u(x, t) := p · e 4(t − τ ) · g(ξ, τ ) dξ dτ
4π(t − τ )
0 −∞
116
9.2 Applying Tarski’s Fixed Point Theorem
For a proof we refer to (33), pp. 1-25. There, the theorem is only given for
real-valued functions, but the theorem remains true if g(x, t) has got its values
in any real Banach space E, because one has just to substitute the norm for the
absolute value. Also, if h : [a, b] → E is continuous, then
Zb Z∞
h(x)dx as well as h(x)dx
a −∞
P
have to be understood as limits of Riemann sums h(ξk ) · (xk − xk−1 ). Now,
we can attack problem (9.5).
Theorem 9.3. Let f : R × [0, T ] × l∞ → l∞ be a function with the following
properties:
i) f is continuous.
ii) There exists a constant L1 so that for all (x, t, z), (y, t, z) ∈ R × [0, T ] × l∞
it holds
kf (x, t, z) − f (y, t, z)k ≤ L1 · |x − y|.
iii) There exists a constant L2 so that for all (x, t, z), (x, t, z̃) ∈ R × [0, T ] × l∞
it holds
kf (x, t, z) − f (x, t, z̃)k ≤ L2 · kz − z̃k.
iv) There exists a constant M so that for all (x, t, z) ∈ R × [0, T ] × l∞ it holds
kf (x, t, z)k ≤ M.
with
Γ = {Γn }∞
n=1 and Γn = T · M,
117
9 Tarski’s Fixed Point Theorem
√ q
T
L3 := 2 · M · T and L4 := M + 2(L1 + L2 · L3 ) · π. Moreover, for ω ∈ Ω we
define
Zt Z∞ (x − ξ)2
1 −
Φ(ω)(x, t) := p · e 4(t − τ ) · f (ξ, τ, ω(ξ, τ )) dξ dτ.
4π(t − τ )
0 −∞
Rt R∞ 2
= M· √1 · e−α dα dτ.
π
0 −∞
118
9.2 Applying Tarski’s Fixed Point Theorem
R∞ 2
Using √1 · e−α dα = 1 we get
π
−∞
Zt
kΦ(ω)(x, t)k ≤ M · 1dτ ≤ T · M.
0
Zt Z∞ (x − ξ)2
1 − ξ−x
Φn (ω)x (x, t) = p e 4(t − τ ) fn (ξ, τ, ω(ξ, τ )) dξ dτ
4π(t − τ ) 2(t − τ )
0 −∞
M
and we get |Φn (ω)x (x, t)| ≤ √ (I
π 1
+ I2 ) with
Zt Zx (x − ξ)2
1 − |ξ − x|
I1 := p · e 4(t − τ ) · dξ dτ
4(t − τ ) 2(t − τ )
0 −∞
Zt Zx (x − ξ)2
1 − x−ξ
= p · e 4(t − τ ) · dξ dτ
4(t − τ ) 2(t − τ )
0 −∞
and
Zt Z∞ (x − ξ)2
1 − |ξ − x|
I2 := p · e 4(t − τ ) · dξ dτ
4(t − τ ) 2(t − τ )
0 x
Zt Z∞ (x − ξ)2
1 − ξ−x
= p · e 4(t − τ ) · dξ dτ .
4(t − τ ) 2(t − τ )
0 x
2
Substituting α := − (x−ξ)
4(t−τ ) , −∞ < ξ ≤ x it follows (using dα =
x−ξ
2(t−τ ) dξ) that
Zt Z0 Zt Z0
1 1
I1 = p · eα dα dτ = p · eα dα dτ
4(t − τ ) 4(t − τ )
0 −∞ 0 −∞
Zt √
1 √ √
= p dτ = [− t − τ ]|t0 = t ≤ T .
4(t − τ )
0
2
In the same way, substituting α := − (x−ξ)
4(t−τ ) , x ≤ ξ < ∞ it follows using the
119
9 Tarski’s Fixed Point Theorem
All together,
M M √ √
|Φn (ω)x (x, t)| ≤ √ (I1 + I2 ) ≤ √ · 2 · T < 2 · M · T = L3 .
π π
As in Case (B1) one can show kJ2 k ≤ M · (t − s) = M · |t − s|. So, the hard
work is kept in estimating J11 and J12 , respectively. Substituting α = √ x−ξ
4(t−τ )
120
9.2 Applying Tarski’s Fixed Point Theorem
Zs Z∞
1 2
√ √
√ · e−α ·
f (x − 2α t − τ , τ, ω(x − 2α t − τ , τ ))
kJ11 − J12 k ≤
π
0 −∞
√ √
−f (x − 2α s − τ , τ, ω(x − 2α s − τ , τ ))
dα dτ
Zs Z∞
1 2
√ √
√ · e−α · (L1 + L2 · L3 ) · x − 2α t − τ − x + 2α s − τ dα dτ
≤
π
0 −∞
Zs Z∞
1 √ √ 2
= √ · (L1 + L2 · L3 ) · ( t − τ − s − τ ) dτ · e−α 2|α| dα .
π
0 −∞
| {z } | {z }
=:F =2
Concerning F we can conclude
s
2 3 3 2 3 3 3 2 3 3
F = [−(t − τ ) 2 + (s − τ ) 2 ] = (t 2 − s 2 − (t − s) 2 ) < (t 2 − s 2 ) =: P.
3 0 3 3
3 √
Considering the function b(x) := 23 x 2 , 0 ≤ x ≤ T with b0 (x) = x by using the
mean-value theorem we get
2 3 3 √ √
P = (t 2 − s 2 ) = b(t) − b(s) = b0 (η) · (t − s) ≤ T · (t − s) = T · |t − s|
3
for some η ∈ (s, t). All together it holds
= L4 · |t − s|.
Zt Z∞ (x − ξ)2
1 −
Φn (ω)(x, t) = p · e 4(t − τ ) · fn (ξ, τ, ω(ξ, τ )) dξ dτ
4π(t − τ )
0 −∞
121
9 Tarski’s Fixed Point Theorem
Zt Z∞
1 (x−ξ)2
≤ p · e− 4(t−τ ) · fn (ξ, τ, v(ξ, τ )) dξ dτ = Φn (v)(x, t).
4π(t − τ )
0 −∞
So, all assumptions of Tarski’s fixed point theorem are fulfilled, and we can apply
it.
Finally, we want to note, that is not easy to see, how Theorem 9.3 can be proved
using Schauder’s fixed point theorem.
122
10 Further Reading
To emphasize that mathematical research never ends we add some more fixed
point theorems for further reading. To avoid any kind of rate the following list
is given in alphabetical order:
• The fixed point theorem of Browder, see (17).
• The fixed point theorem of Darbo, see (26).
• The fixed point theorem of Herzog and Kunstmann, see (45).
• The fixed point theorem of Kakutani, see (50).
• The fixed point theorem of Kneser, see (53).
• The fixed point theorem of Lagler and Volkmann, see (56).
• The fixed point theorem of Lefschetz and Hopf, see (30).
• The fixed point theorem of Lemmert, see (58).
• The fixed point theorem of Sadovskii, see (77).
• The fixed point theorem of Tychonoff, see (100).
• The fixed point theorem of Weissinger, see (108).
Some of them are (just) generalizations of others, some of them have also very
nice applications. The reader might search for more fixed point theorems or the
reader might extend the presented applications by his own or, even better, the
reader will apply some fixed point theorem in a new way. Anyway, it’s up to
you now.
123
Bibliography
[1] R. A. Adams, Sobolev spaces, Academic Press, New York, 1975.
[5] G. Alefeld, G. Mayer, The Cholesky method for interval data, Linear
Algebra Appl., 194 (1993), pp. 161-182.
[10] S. Banach, Sur les opérations dans les ensembles abstraits et leur appli-
cation aux équations intégrales, Fund. math., 3 (1922), pp. 133-181.
[12] P. Bod, On closed sets having a least element, Optimization and opera-
tions research. Lecture Notes in Econom. Math. Systems, 117 (1976), pp.
23-34.
125
Bibliography
[23] J. Crank, Free and moving boundary problems, Clarendon Press 1984.
126
Bibliography
127
Bibliography
128
Bibliography
[59] J. Milnor, John Nash and “A beautiful mind”, Notices Amer. Soc., 45
(1998), pp. 1329-1332.
[67] M. Neher, The mean value form for complex analytic functions, Com-
puting, 67 (2001), pp. 255-268.
[71] M. Plum, Existence and multiplicity proofs for semilinear elliptic bound-
ary value problems by computer assistance, Jahresber. Deutsch. Math.-
Verein., 110 (2008), pp. 19-54.
129
Bibliography
[75] S. Robinson, The problem with blondes, SIAM News, 35 (2002), p. 20.
[78] U. Schäfer, A fixed point theorem based on Miranda, Fixed Point Theory
Appl., Vol. 2007, Article ID 78706 (2007), 5 pages.
[85] U. Schäfer, Aspects for a block version of the interval Cholesky algo-
rithm, J. Comput. Appl. Math., 152 (2003), pp. 481-491.
130
Bibliography
131
Bibliography
132
Index
133
Index
open set, 88
operator, 95
Pascal-XSC, 73
Peano’s existence theorem, 97
Picard-Lindelöf, 81
Poincaré-Miranda theorem, 33
polytope, 8
prisoners’ dilemma, 29
radial function, 21
Tamir’s algorithm, 49
Tarski’s fixed point theorem, 110
Taylor’s formula, 45, 59, 106
tolerance, 54
topologically equivalent, 21
two-person game, 25
vector addition, 76
134
from sperner‘s lemma to
differential equations
in banach spaces
An Introduction to Fixed Point
Theorems and their Applications
ISBN 978-3-7315-0260-9
9 783731 502609