Nonex Kap1
Nonex Kap1
Nonex Kap1
One of the most important instrument to treat (nonlinear) problems with the aid of
functional analytic methods is the fixed point approach. This approach is an important
part of nonlinear (functional-)analysis and is deeply connected to geometric methods of
topology. We consider in this chapter the famous theorems of Banach, Brouwer and
Schauder. A more detailed description of the fixed point theory can be found for instance
in [1, 2, 6, 8, 9] dar.
x = F(x) (1.1)
F : X � x �−→ x + x0 ∈ X where x0 �= θ .
g(x) = 0
There are different possibilities to translate this equation into a fixed point problem:
1
• Consider the determination of a zero z of a function f : Rn −→ Rn . Clearly, such a
zero is a fixed point of the mapping F : Rn � x �−→ x + f(x) ∈ Rn . The formulation
of equations as a fixed point equation is possible in various ways. Consider
f : R � x �−→ x3 − 13x + 18 ∈ R .
−Δu = f(u), in Ω , u = θ in ∂Ω
f is given and the solution u has to be found. Under certain conditions one can
define the inverse (−Δ)−1 of Δ on spaces of functions u which vanish on ∂Ω . Then
we obtain the following fixed point equation:
F0 := id , Fn+1 := F ◦ Fn , n ∈ N0 .
Choosing x ∈ X, the sequence (Fn (x))n∈N is called the orbit of x under F . This orbit is
the result of the iteration
This iteration process is in the focus of the following chapters. The main question which
we will analyze is under which assumptions (concerning F and X) the orbit converges.
2
1.2 Fixed point theorem of Banach
Consider the fixed point equation (1.1) again. The essential step to prove the existence
of a solution of this equation via the iteration
xn+1 := F(xn ) , n ∈ N0 ; x0 := x , (1.4)
is to prove that the orbit (Fn (x))n∈N is a Cauchy sequence (for every x ∈ X). The
assumption which guarantees this is the assumption that F is Lipschitz continuous with
Lipschitz constant c satisfying c < 1 . This means
There exists c ∈ [0, 1) such that d(F(x), F(y)) ≤ cd(x, y) for all x, y ∈ X . (1.5)
If (1.5) is satisfyied, F is called a contraction. Notice that cn is the Lipschitz constant
of Fn .
Lemma 1.1. Let (X, d) be a metric space and let F : X −→ X be a contraction. If
c ∈ [0, 1) is the Lipschitz constant of F, then we have
1
d(x, y) ≤ (d(x, F(x)) + d(y, F(y))) , x, y ∈ X . (1.6)
1−c
Proof:
Let x, y ∈ X . Using the triangle inequality
d(x, y) ≤ d(x, F(x)) + d(F(x), F(y)) + d(F(y), y)
and using d(F(x), F(y)) ≤ cd(x, y) the result follows. �
(1.6) is a very strange inequality: it says that we can estimate how far apart any two
points x, y are just from knowing how far x is from its image F(x) and how far y is from
its image F(y) . A first application is
Corollary 1.2. Let (X, d) be a metric space and let F : X −→ X be a contraction. Then
F has at most one fixed point.
Proof:
Suppose that x, y are fixed poits of F . Then d(x, F(x)), d(y, F(y)) are zero, so by the
inequality (1.6) d(x, y) is zero too. �
Lemma 1.3. Let (X, d) be a metric space and let F : X −→ X be a contraction. Then
the orbit (Fn (x))n∈N0 is a Cauchy sequence.
Proof:
Let c ∈ [0, 1) be the Lipschitz constant of F . From the inequality (1.6) we read off
1
d(Fn (x), Fm (x)) ≤ (d(Fn (x), Fn+1 (x)) + d(Fm (x), Fm+1 (x))) , m, n ∈ N .
1−c
Recalling that ck is the Lipschitz constant of Fk we get
cn + c m
d(Fn (x), Fm (x)) ≤ d(x, F(x)) , m, n ∈ N .
1−c
Since limk ck = 0 we obtain limm,n→∞ d(Fn (x), Fm (x)) = 0 . �
3
Theorem 1.4 (Contraction Theorem, Banach 1922). Let (X, d) be a complete metric
space and let F : X −→ X be a contraction. If c ∈ [0, 1) is the Lipschitz constant of F,
then the following assertions hold:
Proof:
Ad (a), (b)
The uniqueness of a fixed point follows from Corollary 1.2. Let x ∈ X . From Lemma (1.3)
we obtain that the orbit (Fn (x))n∈N0 converges to a point z since (X, d) is complete. Since
F is continuous due the fact that F is a contraction – see (1.5) – z is a fixed point.
Ad (c)
Let x ∈ X . We may prove with the help of (1.5) inductively
1 − cm−1
m
d(F (x), F(x)) ≤ c d(Fn (x), Fn−1 (x)) , m ∈ N . (1.11)
1−c
(d), (e) are simple consequences of these inequalities. �
Theorem 1.5 (Banach’s fixed point theorem). Let (X, � · �) be a Banach space, let V be
a closed subset of X and let F : V −→ V be a contraction, that is
4
b) The successive iteration
and we conclude
1
d(xp , xp0 ) ≤d(Fp (xp0 ), Fp0 (xp0 )) .
1−c
By taking the limit p → p0 we obtain limp→p0 xp = xp0 . �
5
1
Example 1.8. X = R , � · � := | · | , M := [0, ∞), F : M � t �−→ t + ∈ M . We
1+t
have
|F(t) − F(s)| < |t − s| .
F posesses no fixed point! Notice that
Proof:
Define
1
G : Br � x �−→ (x + F(x)) ∈ X .
2
Let x ∈ Br , x �= θ , and consider u := rx�x�−1 ∈ ∂Br ⊂ Br . Then we obtain
and hence
and
1 1
�G(x)� = �x + F(x)� ≤ (�x� + �F(x)�) ≤ r .
2 2
Due to the continuity of G we obtain �G(θ)� ≤ r .
This shows that G : Br −→ Br and G is a contraction too as we conclude from
1 1
�G(x) − G(y)� ≤ (�x − y� + L�x − y�) = (1 + L)�x − y� , x, y ∈ Br .
2 2
Theorem 1.5 implies that G has a fixed point x ∈ Br . This implies x = F(x) . Due to the
fact that F is a contraction this fixed point is uniquely determined. �
Remark 1.10. What is a kind of the inverse of the contraction fixed point theorem? The
most elegant result in this direction is due to Bessaga [5].
Suppose that X is an arbitrary nonempty set and suppose that F : X −→ X has
the property that F and each of its iterates T n has a unique fixed point. Then for each
c ∈ (0, 1) there exists a metric dc on each X such that (X, dc ) is complete and for which
6
Theorem 1.11 (Edelstein, 1961). Let X be a Banach space and let V ⊂ X . Suppose that
F : V −→ V is contractive, i. e.
Then we have
Proof:
Ad a) Obvious.
Ad b) Consider the mapping V � v �−→ �v − F(v)� ∈ R . Since V is compact and since
the norm is continuous there exists x ∈ V with �x − F(x)� = inf v∈V �v − F(v)� . Now we
must have x = F(x) because otherwise we would have
Remark 1.12. The completeness that we assume in Theorem 1.5 and in the following
results is more or less necessary for the existence of fixed points; see for instance [7, 11,
10, 14, 15]. �
Remark 1.13. An application of the contraction theorem of Banach is the theorem of the
inverse mapping; see for instance [4]. �
d(xN , z) ≤ ε
where (Fn (x))n∈N ist the orbit generated by the iteration (1.17). To implement such a
stopping rule one may pursue two different concepts.
A priori concept
This results from the estimation (c) in Theorem 1.4:
cn
n
d(x , z) ≤ d(F(x), x) , n ∈ N .
1−c
A priori stopping rule
7
If δ := d(F(x), x) and N > (log(ε) + log(1 − c) − log(δ))/ log(c) then d(xN , z) ≤ ε .
Notice that such a number N exists due to the fact that 0 ≤ c < 1 . This number can be
computed before we compute the orbit (Fn (x))n∈N0 .
Suppose we take ε := 10−m in our stopping rule inequality. In order to obtain one more
decimal digit of precision we have to do (roughly) 1/| log(c)| more iterations steps. Stated
a little differently, if we need N iterative steps to get m decimal digits of precision, then
we need another N to double the precision to 2m digits. This kind of error behavior is
called linear convergence.
(1) k := 0
(2) u := xk
(7) STOP with an approximation xN for the fixed point z which satisfies d(xN , z) ≤ ε
For the realization of the stopping rules above we need the Lipschitz constant c . It is
obvious that an approximation c̃ with c̃ ∈ [c, 1) is sufficient.
Now, we consider the successive approximation method under errors. We follow mainly
[12]. We consider in the complete metric space (X, d) the fixed point equation (1.1)
x = F(x) (1.18)
8
where F : X −→ X is a given mapping and x0 is a given starting value of the iteration.
This defines an approximating family (xk )k∈N . We assume that F is a contraction with
contraction constant c ∈ [0, 1) . Then – by the contraction theorem – (xk )k∈N converges
to the unique fixed point z of the equation (1.18). Now, instead of (xk )k∈N0 we have at
hand an approximating sequence (x̃k )k∈N0 only. This sequence may be constructed by an
iteration of a Lipschitz continuous mapping F̃ : X −→ X as follows:
x̃k+1 := F̃(x̃k ) , k ∈ N0 , x̃0 := x . (1.20)
In general, we do not have limn∈N x̃k = z when F̃ is different from F .
Theorem 1.14. Let (X, d) be a complete metric space and let F : X −→ X be a
contraction with Lipschitz constant c ∈ [0, 1) and a fixed point z . Let δ > 0 and suppose
that (xk )k∈N0 , (x̃k )k∈N0 = (x̃(δ)k )k∈N0 are sequences such that the following conditions are
satisfied:
xn+1 := F(xn ) , n ∈ N0 , x0 = x . (1.21)
d(x̃n+1 , F(x̃n )) ≤ δ , n ∈ N0 . (1.22)
d(x̃0 , x0 ) ≤ δ . (1.23)
Stopping rule: Suppose there exists N(δ) ∈ N with
d(x̃n+1 , x̃n ) ≤ d(x̃n+1 , x̃n ), n = 0, . . . , N(δ) , d(x̃N(δ)+1 , x̃N(δ) ) > d(x̃N(δ) , x̃N(δ)−1 ) (1.24)
Then
lim x̃N(δ) = z (1.25)
δ↓0
Proof:
Let m ≤ N(δ) . Then
d(x̃m , xm ) ≤ d(x̃m , F(xm−1 ) ≤ δ + cd(x̃m−1 , xm−1
..
.
≤ δ + cδ + · · · + cm−1 δ + cm d(x̃0 , x0 )
δ
≤ + d(x̃0 , x0 ) .
1−c
This implies
lim d(x̃m , xm ) = 0 , m = 1, . . . , N(δ) , lim d(x̃N(δ) , xN(δ) ) = 0 . (1.26)
δ↓0 δ↓0
Since
d(x̃m , z) ≤ d(x̃m , xm ) + d(xm , z) , m = 0, . . . , N(δ), (1.27)
we have to show limδ↓0 d(xN(δ) , z) = 0 .
Case 1: limδ↓0 N(δ) = ∞ .
Since limn d(xn , z) = 0, nothing has to be proved.
Case 2: If limδ↓0 N(δ) = ∞ does not hold then there exist a sequence (δl )l∈N and q ∈ N
with Nl := N(δl ) ≤ q, l ∈ N .
Let m ≤ q . Then
d(xq , xq−1 ) ≤ cd(xq−1 , xq−2 ) ≤ · · · ≤ cq−m d(xm , xm−1 ) . (1.28)
9
Moreover, due to cq−m ≤ 1 we have
For l → ∞ the first and the second term in (1.29) converge to zero due to 1.26. From the
stopping rule we obtain
d(xNl , xNl −1 ) ≤ d(xNl +1 , xNl )
and hence
This implies
2δ
d(x̃Nl , x̃Nl −1 ) ≤
. (1.30)
1−c
Therefore, the seond term in (1.29) converges to zero too if l goes to infinity. From this
we conclude that we must have
Each term in the sum on the righthand side can be extimated analogous to (1.29) and
converges to zero as l goes to infinity. This shows
10
Theorem 1.16 (Fixed point theorem of Brouwer, 1912; Poincaré, 1883). Let B1 be the
closed unit ball in the euclidean space Rn and let F : B1 −→ B1 be a continuous mapping.
Then F has a fixed point.
Let C be a closed convex subset of Rn with nonempty interior. We may assume that
θ ∈ C . Then C is homeomorph to the closed unit ball B1 in Rn . This can easily seen
using the Minkovski functional with respect to the ball which is contained in the interior
of C . If C is an arbitrary nonempty closed convex subset in Rn then C has a nonempty
interior as a subset of the affine hull aff(C) of C . Therefore, C is homeomorph to the unit
ball B1 in Rk for some k ≤ n . Let T : B1 −→ C be such a homoemorphism. We set
G := T −1 ◦ F ◦ T : B1 −→ B1 . Then G is continuous and has a fixed point z . Then T (z)
is a fixed point of F . This consideration leads to a more general formulation of Brouwer’s
fixed point theorem.
Theorem 1.17. Let C be a nonempty convex compact subset of a finite-dimensional
normed space and let F : C −→ C be a continuous mapping. Then F has a fixed point.
Proof:
We present a proof different of the considerations above. It is more in the spirit of the
main topic in this monograph.
Without loss of generality we may assume that C is a subset of the euclidean space Rk
for some k . Since C is compact it is bounded. Suppose C ⊂ Br . Consider the mapping
PC : Rk −→ C defined by the property �x − PC (x)� = inf �x − u� .
u∈C
we conclude PC (x) = y since PC (x) is uniquely determined. For the fact liml dist(xnl , C) =
dist(x, C) we use that �dist(·, C) is Lipschitz-continuous; see Lemma 2.1.
Now consider R := PC �Br . Then G := F ◦ R is a continuous mapping from Br into Br and
has, due to Brouwer’s fixed point theorem, a fixed point x . Since G(x) ∈ C x is in C too.
Hence R(x) = x and F(x) = x . �
Remark 1.18. Deeply connected to the Brouwer fixed point theorem are the following
theorems:
• Hairy ball therorem
This theorem states that there is no nonvanishing continuous tangent vector field on
evendimensional n-spheres.
• Brötchensatz“
”
This ham sandwich theorem states that given n (measurable) objects“ in n-dimensional
”
space can be divided in half (with respect to their measure) with a single (n − 1)-
dimensional hyperplane.
11
• Theorem of Borsuk-Ulam
This theorem states that every continuous function from an n-sphere into an eu-
clidean n-space maps some pair of antipodal points to the same point.
The connection between these results is the following:
Theorem of Borsuk-Ulam =⇒ Brötchensatz“
”
Theorem of Borsuk-Ulam =⇒ Brouwers fixed point Theorem
The hairy ball Theorem =⇒ Brouwers fixed point Theorem
�
Example 1.19. The following example shows that the result does not hold in infinite-
dimensional Banach spaces.
Let H be the space of the square-summable sequences,i.e.: H = l2 . This space is a Hilbert
space endowed with the inner product
�
�(xn )n∈N |(yn )n∈N � := xn y n
n∈N
and norm � 1
�(xn )n∈N � := ( x2n ) 2 .
n∈N
We define a mapping F on l2 by
�
F((xn )n∈N ) := ( 1 − �(xn )n∈N �2 , x1 , x2 , . . . ) .
Clearly, �F((xn )n∈N )� = 1 for all (xn )n∈N ∈ l2 and hence F : B1 −→ B1 . F is continuous
since limk xk = x in l2 implies
� �
lim �F(xk ) − F(x)�2 = lim(( 1 − �xk �2 − 1 − �x�2 )2 + �xk − x�2 ) = 0 .
k k
Assume F has a fixed point x = (xn )n∈N ∈ B1 then �x� = 1 (see above) and hence
x1 = 0, x2 = x1 = 0, x3 = 0, . . . , . Thus, �x� = 0, contradicting the fact that �x� = 1 . �
In Example 1.19 we have seen that Brouwer’s fixed point theorem in a infinite-dimen-
sional Hilbert space does not hold in general. The following fixed point theorem of Kaku-
tani shows that the generalization of Brouwer’s fixed point theorem without additional
assumptions is not possible; we omit the proof.
Theorem 1.20. Let H be an infinite-dimensional Hilbert space. Then there is a contin-
uous mapping F : H −→ H which maps B1 into B1 and do not have a fixed point.
12
space and let C := B1 be the closed unit ball in X . If F : C −→ C is a nonexpansive
mapping then F has a fixed point.
Theorem 1.22. Let H be a Hilbert space, let C be a bounded closed convex subset of H
and let F : C −→ C be a nonexpansive mapping. Then F has a fixed point.
Proof:
Consider for t ∈ (0, 1] the map Ft : C � x �−→ tu + (1 − t)F(x) ∈ H . Since C is convex
the range of Ft is contained in C . We have
and we conclude from Banach’s contraction theorem that Ft has a uniquely determined
fixed point xt ∈ C: xt = tu+(1−t)F(xt ), t ∈ (0, 1] . Since C is a bounded set the diameter
diam(C) is finite. We obtain
This shows limt↓0 �xt − F(xt )� = 0 and we say that F has approximate fixed points.
Consider Fixε (F) := {z ∈ C : �F(z) − z� ≤ ε} . We conclude from the fact that limt↓0 �xt −
F(xt )� = 0 that Fixε (F) �= ∅ for all ε > 0 . Clearly, Fixε (F) is closed for all ε > 0 since F
is continuous. Moreover, Fixε (F) is convex for all ε > 0 since C is a subset of a Hilbert
space. We shall prove this in Chapter 7 in a more general context; see 7.16. Thus we know
that Fixε (F) is weakly closed for all ε > 0 and therefore weakly compact as a subset of the
weakly compact subset C . This implies ∩ε>0 Fixε (F) �= ∅ . Clearly, every x ∈ ∩ε>0 Fixε (F)
is a fixed point of F . �
The approach which is used in the proof of Theorem 1.22 is a part of a huge theory
on the iterative computation of fixed points of nonexpansive mappings. For example,
Halpern’s iteration
xn+1 := αn xn + (1 − αn )F(xn ) , n ∈ N0 , (1.33)
(see Chapter 7) is used to compute a fixed point of F when we choose the sequence (αn )n∈N0
in an appropriate way. Surely, limn αn = 0 should be satisfied. We caome back to this
iteration method in Chapter 7.
13
(c) F is called compact if F(M ∩ B) is compact for every bounded subset B of X .
�
We define φj : A −→ [0, 1] by
dist(F(x), Y\Bε (yj ))
φj (x) := �m ,x∈X.
dist(f(x), Y\Bε (yi ))
i=1
14
Now we have
if one chooses k = k(l, m) appropriate. As a consequence the sequence (yn )n∈N ; congerges.
y := limn yn . From
15
Then we may consider the iteration
v0 � v1 � w1 � w0 (1.37)
then the iteration (1.36) can be continued and we obtain sequences (vn )n∈N0 , (wn )n∈N0
with
v0 � v1 � v2 � · · · � vn � wn � · · · � w2 � w1 � w0 , n ∈ N . (1.38)
Under assumptions which imply that the interval“ {u ∈ X : v0 � u � w0 } is mapped
”
into a relative compact subset of X there exists a fixed point x ∈ X with
This approach may be used in various situations. Notice that a matrix A ∈ Rn,n may
be decomposed into A = A1 + A2 where both A1 , −A2 have nonnegative entries. Then
the mapping u �−→ A1 u is syntone and u �−→ A2 u is antitone with respect to the
usual partial order in Rn .
Another area of application is the theory of integral equations x = F(x) = Tx + y when T
is an integral operator of (Hammerstein-)type:
�
Tu(t) := κ(t, s)φ(u(s))ds , t ∈ B ,
B
We may decompose the kernel κ and/or φ such that the method above may be applied.
Example 1.30. Consider the equation
�1
1
x(t) − 1 = T (x)(t) := |t − s|(x(s) − x(s)2 )ds , t ∈ [0, 1],
0 2
in the space C[0, 1] endowed with the supremum-norm. The partial order in C[0, 1] is
given by f � g ⇐⇒ f(t) ≤ g(t) for all t ∈ [0, 1] . T may be decomposed as follows:
�1 �
1 1
T1 (u)(t) := |t − s|u(s)ds , T2 (u)(t) := − |t − s|u(s)2 ds , t ∈ [0, 1] .
0 2 0
16
1.8 Appendix: Variational principle of Ekeland and
applications
Wir beweisen das Variationslemma von Ekeland in einer endlichdimensionalen Version.
Der Beweis ist hier elementar im Gegensatz zur Situation in metrischen oder topologischen
Räumen. Sei nun in diesem und dem nächsten Abschnitt | · | stets die euklidische Norm
in Rn , assoziert zum euklidischen Skalarprodukt �·|·� .
Theorem 1.31 (Variationslemma von Ekeland, 1972). Sei f : Rn −→ R unterhalbstetig,
nach unten beschränkt. Sei ε > 0, x∗ ∈ Rn mit
Da F nach unten beschränkt und unterhalbstetig ist, und da die Niveaumengen Nr (F) :=
�x ∈ Rn |F(x) ≤ r} beschränkt und abgeschlossen sind, also kompakt sind, gibt es ein x,
das ein Minimum von F darstellt (siehe etwa [4], Satz 3.12) Nun haben wir
17
Daraus folgt √
�∇f(x), w� ≥ − ε|w| für alle w ∈ Rn ,
was nun (c) impliziert. �
i) x ≤ x (Reflexivität)
ii) x ≤ y, y ≤ x =⇒ y = x (Symmetrie)
S(x) := {y ∈ X|y ≥ x} ,
Theorem 1.33 (Brezis–Browder, 1976). Sei (X, ≤) eine halbgeordnete Menge und ϕ :
X −→ R monoton wachsend, d.h. ϕ(x) ≤ ϕ(y), falls x ≤ y. Es gelte:
i) Für jede monoton wachsende Folge (xn )n∈N mit sup{ϕ(xn )|n ∈ N} < ∞ existiert
y ∈ X mit xn ≤ y für alle n ∈ N .
Proof:
Für u ∈ X setze µ(u) := sup{ϕ(y)|y ∈ S(u)} .
Annahme: Es gibt x ∈ X mit µ(x) < ∞.
Definiere eine Folge (xn )n∈N induktiv in folgender Weise: x1 := x; sind x1 , . . . , xn bestimmt,
wähle xn+1 ∈ S(xn ) mit µ(xn ) − 1/n ≤ ϕ(xn+1 ) .
Wegen µ(xn ) ≤ µ(x) < ∞ existiert xn+1 und wegen xn+1 ≥ xn gilt µ(xn+1 ) ≤ µ(xn ) . Nun
gilt also
Wegen i) gibt es y ∈ X mit xn ≤ y für alle n ∈ N . Wähle nach ii) u ∈ X mit u ∈ S(y)
und ϕ(y) < ϕ(u) . Nun gilt xn ≤ u, ϕ(u) ≤ µ(xn ), n ∈ N. Also
1
ϕ(u) ≤ µ(xn ) ≤ ϕ(xn+1 ) + 1/n ≤ ϕ(y) + , n ∈ N,
n
d. h. ϕ(u) ≤ ϕ(y) < ϕ(u), was ein Widerspruch ist. �
18
Corollary 1.34. Sei (X, ≤) eine halbgeordnete Menge, ψ : X −→ R nach oben
beschränkt und es gelte:
w ∈ X, w ≥ v =⇒ ϕ(w) = ϕ(v) =⇒ w = v .
�
Corollary 1.35. Sei (X, d) ein metrischer Raum und sei (X, ≤) halbgeordnet. Sei η :
X −→ R streng monoton fallend und nach unten beschränkt. Es gelte:
(a) Für alle x ∈ X ist S(x) abgeschlossen.
(b) Für jede monoton wachsende Folge (xn )n∈N in X ist {xn |n ∈ N} kompakt.
Dann gilt:
∀ u ∈ X ∃ v ∈ S(u) ∀ w ∈ X (w ≥ v =⇒ w = v) .
Proof:
Sei ψ := −η. Wir weisen (10.19) nach. Sei dazu (xn )n∈N eine monoton wachsende Folge
in X . Dann existiert wegen (b) ein y ∈ X und eine Teilfolge (xnk )k∈N mit y = limk xnk .
Da xnk ∈ S(xnj ) für alle k ≥ j gilt, folgt wegen (a) y ∈ S(xnj ) ⊂ S(xn ), nj ≥ n, n ∈ N.
Also y ≤ xn für alle n ∈ N. Nun können wir Folgerung 10.27 (b) anwenden. �
Theorem 1.36. Sei (X, d) ein vollständiger metrischer Raum und sei f : X −→ (−∞, ∞]
unterhalbstetig, nach unten beschränkt und eigentlich, d. h. es gibt z ∈ X mit f(z) < ∞ .
Sei ε > 0, x∗ ∈ X mit
f(x∗ ) ≤ inf f(v) + ε . (1.44)
v∈X
19
1. f(x) ≤ f(x∗ ) ;
2. d(x, x∗ ) ≤ 1 ;
Proof:
f(x∗ ) ist endlich wegen (10.20). X � := {z ∈ X|f(z) ≤ f(x∗ )} ist abgeschlossen, da f unter-
� d) ein vollständiger metrischer Raum mit x∗ ∈ X
halbstetig ist. Also ist (X, � . Haben wir
das Resultat in (X,� d) für f� := f � bewiesen, dann ist das Resultat in (X, d) bewiesen,
|X
denn die Bedingung 3. folgt dann mit 1. aus der Zeile
�.
f(y) + εd(y, x) > f(x∗ ) + εd(y, y) ≥ f(x∗ ) ≥ f(x) , y ∈ X\X
Daraus folgt d(u, v) = 0 und daher ist die Symmetrie gezeigt. Seien u, v, w ∈ X mit
u ≤ v, v ≤ w, d. h.
20
ab, dass (xn )n∈N eine Cauchyfolge und daher konvergent ist.
Wir können nun Folgerung 10.28 anwenden: Es gibt x ∈ S(x∗ ) mit der Eigenschaft
u ≥ x =⇒ u = x . (1.45)
Damit ist 1., 2. gezeigt. Sei y ∈ X\{x}. Wegen (10.21) gilt nicht y ≥ x, d. h. f(y) − f(x) >
−εd(y, x) . �
Corollary 1.37. Sei (X, d) ein vollständiger metrischer Raum und sei f : X −→
(−∞, ∞] unterhalbstetig, nach unten beschränkt und eigentlich. Sei γ > 0, x∗ ∈ X .
Dann gibt es x ∈ X mit
Proof:
Sei Y := {x ∈ X|f(x) + γd(x, x∗ ) ≤ f(x∗ )} , g := γ1 f|Y . Y ist abgeschlossen, da f unterhalb-
stetig ist. Also ist (Y, d) ein vollständiger metrischer Raum. g ist unterhalbstetig, nach
unten beschränkt und eigentlich. Sei α := inf v∈Y g(v). Es gibt x∗∗ ∈ Y mit g(x∗∗ ) ≤ α + 1.
Also ist Satz 10.29 anwendbar mit ε = 1 :
1 1
∃ x ∈ Y ∀x ∈ Y\{x} ( f(x) < f(x) + d(x, x)) .
γ γ
Also
f(x) < f(x) + γd(x, x)) für alle x ∈ Y\{x}) , f(x) ≤ f(x∗ ) − γd(x∗ , x) .
ii) ist damit schon bewiesen.
Ist x ∈ X\Y, so gilt f(x) + γd(x, x∗ ) > f(x∗ ) und mit ii) folgt
f(x) ≤ f(x∗ ) − γd(x∗ , x) < f(x) + γ(d(x, x∗ ) − d(x∗ , x)) ≤ f(x) + γd(x, x) .
Theorem 1.38 (Variationslemma von Ekeland, 1972). Sei (X, d) ein vollständiger metrischer
Raum und sei f : X −→ (−∞, ∞] unterhalbstetig, nach unten beschränkt und eigentlich.
Sei ε > 0, x∗ ∈ X mit
f(x∗ ) ≤ inf f(v) + ε . (1.46)
v∈X
21
Proof:
1d.
Sei γ > 0. Wende Satz 10.29 an unter Verwendung der zu d äquivalenten Metrik d̃ := γ
�
√
Beachte: Es liegt folgender Kompromiss für die Wahl von γ nahe: γ := ε; siehe
Folgerung 10.17.
Das folgende Resultat stellt eine Existenzaussage im Kontext der Voraussetzungen von
Satz 10.29 bereit. Ob es als Existenzsatz Verwendung finden kann, ist zweifelhaft, wir
werden ihn aber verwenden können.
Theorem 1.39 (Takahashi, 1989). Sei (X, d) ein vollständiger metrischer Raum und sei
f : X −→ (−∞, ∞] unterhalbstetig, eigentlich und nach unten beschränkt. Es gelte:
Für alle u ∈ X mit f(u) > inf f(x) gibt es v ∈ X mit v �= u, f(v) + d(u, v) ≤ f(u) .
x∈X
Wegen f(x) > µ, gibt es v ∈ X, v �= x, mit f(v) + d(v, x) ≤ f(x) . Dies ist ein Widerspruch.
�
Die Vollständigkeit der metrischen Räume in den obigen Resultaten war wesentlich
für die Beweise. Sie ist auch wesentlich für die Gültigkeit der Resultate, denn man kann
zeigen, dass die Vollständigkeit in einem metrischen Raum schon durch die Tatsache
charakterisiert wird, dass die Aussage von Satz 10.31 für jedes f mit den dortigen Eigen-
schaften gilt.
Theorem 1.40. Sei (X, d) ein metrischer Raum. Dann sind äquivalent:
(a) (X, d) ist vollständig.
(b) Für jede unterhalbstetige Abbildung f : X −→ R mit inf v∈X f(v) ≥ 0 gilt: Ist
ε > 0, x∗ ∈ X mit f(x∗ ) ≤ inf v∈X f(v) + ε, so gibt es x ∈ X mit
1. f(x) ≤ f(x∗ ) ;
2. d(x, x∗ ) ≤ 1 ;
3. f(y) + εd(y, x) ≥ f(x) für alle y ∈ X .
Proof:
(a) =⇒ (b). Satz 10.29.
(b) =⇒ (a).
Sei (xn )n∈N eine Cauchyfolge in X . Dann existiert offenbar limn d(xn , y) für jedes y ∈ X .
Definiere f : X −→ R durch
22
f ist unterhalbstetig, da d(y, ·) stetig ist für alle y ∈ X, und es gilt inf v∈X f(v) = 0 . Sei
ε ∈ (0, 1), x∗ ∈ X mit f(x∗ ) ≤ ε . Dann gibt es nach Satz 10.29 x ∈ X mit
ein x ∈ X mit
1. f(x) ≤ f(x∗ ) .
2. �x − x∗ � ≤ γ.
ε.
3. �DG f(x)� ≤ γ
Proof:
Wähle x ∈ X gemä”s Satz 10.31. Dann gilt
ε
f(y) ≥ f(x) − �y − x�, y ∈ X . (1.47)
γ
Also
ε ε
f(x + tw) ≥ f(x) − �w�t für alle t > 0, d. h. DG f(x)(w) ≥ − �w� .
γ γ
Dies zeigt
ε
�DG f(x)� ≤ .
γ
�
Remark 1.42. Satz 10.34 zeigt, dass für ein Optimierungsproblem
Minimiere f(x), x ∈ X,
unter den dortigen Voraussetzungen Punkte x ∈ X existieren, die f fast minimieren“ und
”
die die notwendige Bedingung DG f(x) = θ fast erfüllen“:
”
ε
f(x) ≤ inf f(v) + ε , �DG f(x)� ≤ .
v∈X γ
�
23
Theorem 1.43 (Kontraktionssatz/Fixpunktsatz von Banach). Sei (X, d) ein vollständi-
ger metrischer Raum und sei F : X −→ X eine Kontraktion, d. h.
Proof:
Betrachte
f : X � x �−→ d(x, F(x)) ∈ R .
Offenbar ist f stetig, nach unten beschränkt und eigentlich. Sei x∗ ∈ X, ε := d(x∗ , F(x∗ )), γ :=
1 − L . Wende damit Satz 10.29 an: Es gibt x ∈ X mit
Dann muss x ein Fixpunkt sein, denn anderenfalls würde x := F(x) einen Widerspruch
ergeben gemä”s
d(x, F(x)) < d(F(x), F(F(x))) + (1 − L)d(x, F(x)) ≤ Ld(x, F(x)) − Ld(x, F(x)) + d(x, F(x)) .
sofort ab. �
Theorem 1.44 (Fixpunktsatz von Caristi–Kirk, 1976). Sei (X, d) vollständiger metrischer
Raum, sei f : X −→ R nach unten beschränkt und unterhalbstetig. Sei F : X −→ X; es
gelte:
d(x, F(x)) ≤ f(x) − f(F(x)) für alle x ∈ X.
Dann besitzt F einen Fixpunkt, d. h. es gibt z ∈ X mit F(z) = z.
Proof:
Annahme: Es gilt F(z) �= z für alle z ∈ X.
Nach Satz 10.29 gibt es x ∈ X mit f(z) − f(x) > −d(z, x) für alle z ∈ X\{x}) . Also
Remark 1.45. Bemerkenswert an Satz 10.37 ist, dass von der Abbildung F nicht einmal
Stetigkeitseigenschaften verlangt werden.
Unter den Voraussetzungen von Satz 10.37 liegt i. a. keine Eindeutigkeit vor, wie folgendes
Beispiel zeigt:
X := [0, 1], d(x, y) := |x − y|, f(x) := |x|, F(x) := x
Beachte: Man kann den Satz 10.29 auch aus dem Fixpunktsatz von Caristi–Kirk (Satz
10.37) herleiten. �
24
Theorem 1.46. Sei (X, d) ein vollständiger metrischer Raum und sei f : X −→ (−∞, ∞]
unterhalbstetig, eigentlich und nach unten beschränkt. Sei T : X −→ POT(X)\{∅}. Es
gelte:
∀ x ∈ X, x ∈
/ T (x) ∃ y �= x (f(y) + d(y, x) ≤ f(x)) .
Dann gibt es z ∈ X mit z ∈ T (z).
Theorem 1.47. Sei (X, d) ein vollständiger metrischer Raum und sei f : X −→ (−∞, ∞]
unterhalbstetig, eigentlich und nach unten beschränkt. Sei T : X −→ POT(X)\{∅}. Es
gelte für alle x ∈ X :
Theorem 1.48. Sei (X, d) ein vollständiger metrischer Raum und sei f : X −→ (−∞, ∞]
unterhalbstetig, eigentlich und nach unten beschränkt. Sei T : X −→ POT(X)\{∅} . Es
gelte für alle x ∈ X
f(y) + d(y, x) ≤ f(x) für alle y ∈ T (x) .
Dann gibt es z ∈ X mit z ∈ T (z).
Theorem 1.49. Sei (X, d) ein vollständiger metrischer Raumund sei f : X −→ (−∞, ∞]
unterhalbstetig, eigentlich und nach unten beschränkt. Sei T : X −→ POT(X)\{∅}
abgeschlossen. Es gelte:
Lemma 1.50. Die Sätze 10.32, 10.39, 10.40, 10.41, 10.42 sind äquivalent.
Proof:
Aus Satz 10.32 folgt Satz 10.39, denn:
Annahme: T hat keinen Fixpunkt, d. h. x ∈ / T (x) für alle x ∈ X ..
Sei u ∈ X mit f(u) > inf x∈X f(x). Da T (u) �= ∅ und da u ∈ / T (u) gibt es v ∈ T (u), v �= u,
mit f(v) + d(v, u) ≤ f(u) . Damit liefert nun Satz 10.32 die Existenz von x mit f(x) =
inf x∈X f(x) . Sei y ∈ T (x). Aus
Aus Satz 10.39 folgt die Existenz von z ∈ X mit z ∈ T (z) . Dies ist auch ein Fixpunkt von
g nach Satz 10.37.
Aus Satz 10.40 folgt Satz 10.32, denn:
25
Definiere T : X � x �−→ {y ∈ X|f(y) + d(x, y) ≤ f(x)} ∈ POT(X) . Auf Grund der
Voraussetzung in Satz 10.32 gilt T (x) �= ∅ für alle x ∈ X .
Annahme: Es gibt kein x ∈ X mit f(x) = inf v∈X f(v) . Nun gilt
Aus Satz 10.40 folgt die Existenz von z ∈ X mit {z} = T (z) . Dies zeigt, dass kein v ∈ X
esistiert mit f(y) + d(y, z) ≤ f(z), was ein Widerspruch ist.
Aus Satz 10.40 folgt Satz 10.41, denn:
Dies ist offensichtlich. Aus Satz 10.41 folgt Satz 10.42, denn:
Setze g(x) := 21 f(x), x ∈ X . Ist d(x, T (x)) = 0, dann ist x ∈ T (x), da T (x) abgeschlossen
ist. Hat also T keinen Fixpunkt, dann ist d(x, T (x)) > 0 für alle x ∈ X . Sei y ∈ T (x) so,
dass d(x, y) < 12 d(x, T (x)) . Dann haben wir
1 1
d(x, y) ≤ d(x, T (x)) ≤ (f(x) − sup f(y)) ≤ g(x) − g(y) .
2 2 y∈T (x)
Da g eigentlich, unterhalbstetig und nach unten beschränkt ist, hat T einen Fixpunkt
nach Satz 10.41.
Aus Satz 10.42 folgt Satz 10.40, denn Satz 10.40 ist ein Spezialfall von Satz 10.42. �
26
1.10 Exercises
1.) Let a < b let F : [a, b] −→ [a, b] be continuous and let F be differentiable in
(a, b) . We assume: There exists q ∈ [0, 1) with
Show:
(a) F has a uniquely determined fixed point x .
(b) The iteration xn+1 := F(xn ), x0 = x, converges for each x ∈ X to x .
Hint for (b): One can show that the function x �−→ �x − F(x)� along the iteration
is monotone decreasing.
8.) Let X be a compact metric space. Show: X is separable.
27
9.) Let C[a, b] be the space of continuous functions on [a, b] endowed with the maximum-
norm. Show that C[a, b] is separable but not compact.
10.) Consider the following system of nonlinear equations:
Find a reformulation as a fixed point equation in R2 and solve this equation (by
Brouwer’s fixed point theorem).
11.) Consider
1 2
F : R2 � (x1 , x2 ) �−→ (x1 e−x2 + x1 x2 + 3, ln(1 + x21 + x22 ) − 1) ∈ R2 .
6
Let R2 be endowed with the maximum norm � · �∞ .
(a) Show that F is Lipschitz-continuous in [0, 1] × [0, 1] with Lipschitz-constant
c = 56 .
(b) Show that there exists a uniquely determined fixed point of F .
(c) How many steps k do we need to obtain an accuracy �xk − z�∞ ≤ 10−3 using
the fixed point iteration xk+1 := F(xk ) starting with x0 := (0, 0) .
12.) Show that
x + 1/2
f : [0, ∞) � x �−→ ∈ [0, ∞)
x+1
is a contraction. Compute the fixed point of f and an approximation x1 using the
starting point x0 = 1 for the fixed point iteration.
13.) Show that the following system of nonlinear equations
sin(x1 + x2 ) − x2 = 0 , cos(x1 + x2 ) − x1 = 0
is solvable in R2 .
14.) Let A = (aij ) ∈ Rn,n be a matrix with entries aij ≥ 0 . Show that A has a
nonnegative eigenvalue λ with associated eigenvector x = (xi ) with entries xi ≥ 0 .
15.) Let F : Rn −→ Rn be a continuous mapping with
with c > 1 . Show: If F is surjective then F has a uniquely determined fixed point.
16.) Let C[0, 1] endowed with the maximum-norm � · �∞ . Let C := {x ∈ C[0, 1] : 0 =
x(0) ≤ x(s) ≤ x(1) = 1} and consider
Show:
(1) T is a linear nonexpansive mapping with �T � = 1 .
(2) T (x) ∈ C if x ∈ C .
(3) C is a bounded closed convex set.
28
(4) T posesses a uniquely determined fixed point belonging to C[0, 1]\C .
(5) There exists xt ∈ C with xt = tu + (1 − t)T (xt ) where u ∈ C, u(s) := s .
(6) inf x∈C �x − T (x)� = 0 .
(7) limn �x0 − T n (x0 )�∞ = diam(C) .
17.) Let X be a Banach space, let C ⊂ X be compact and let F : X −→ X be
continuous. Then the following statements are equivalent:
(a) inf x∈X �x − F(x)� = 0 .
(b) F posesses a fixed point.
18.) Let Xs be a Banach space, let C ⊂ X be compact and convex and let F : X −→ X
be nonexpansive. Then F posesses a fixed point.
19.) Let H be the euclidean space R2 and let C0 := {(x, y) ∈ R2 : x2 + y2 ≤ 1},
C1 := {(x, y) ∈ R2 : x ≥ 0, |y| ≤ x}, C2 := {(x, y) ∈ R2 : x ≤ 0, |y| ≤ |x|} und
C := (C0 ∩ C1 ) ∪ (C ∩ C2 ) . Show: C ⊂ co({PC (x, y) : (x, y) ∈ R2 \C}) .
20.) Let X be a Banach space and let U be a closed subspace of X . Consider the
quotient space
29
Bibliography
[1] R.P. Agarwal, M. Meehan, and D. O’Regan. Fixed point theory and applications.
Cambridge University Press, 2004.
[2] J. Appell and M. Väth. Elemente der Funktionalanalysis. Vieweg, Braunschweig,
2005.
[3] J. Baumeister. Differentialgleichungen. Skriptum 2006, Goethe–Universität Frank-
furt/Main.
[5] C. Bessaga. On the converse of the banach fixed-point principle“. Colloquio Math., 7:41–
”
43, 1959.
[6] J. Cronin. Fixed points and topological degree in nonlinear analysis. Amer. Math. Soc.,
Providence, 1964.
[7] M. Edelstein. An extension of Banach’s contraction principle. Proc. Amer. Math Soc.,
12:7–10, 1961.
[8] A. Granas and J. Dugundij. Fixed point theory. Springer, Berlin, 2003.
[9] V.I. Istratescu. Fixed point theorems: an introduction. Reidel, Dordrecht, 1988.
[10] O. Kada, T. Suzuki, and W. Takahashi. Non convex minimization theorems and fixed point
theorems in complete metric spaces. Math. Japon., 44:381–391, 1996.
[11] E. Karapinar. Edelstein type fixed point theorems. Ann. Funct. Anal., 2:51–58, 2011.
[13] J. Schröder. Anwendung von Fixpunktsätzen bei der numerischen Behandlung nichtlinearer
Gleichungen in halbgeordneten Räumen. Arch. Rational Mech. Anal., 4:177–192, 1959.
[14] S.L. Singh and S.N. Mishra. Remarks on recent fixed point theorems. Fixed point theory
and applications, pages 1–18, 2010.
[15] T. Suzuki and W. Takahashi. Fixed point theorems and characterizations of metric com-
pleteness. Topological methods in nonlinear Analysis, 8:371–382, 1996.
30