Sie sind auf Seite 1von 38

Contents

1 Preliminaries 3
1.1 Formal Degree Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 On the Mappings of Monotone Type . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 Construction of Degree 10
2.1 Bounded Mappings of Class (S)
+
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Uniqueness of Degree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Unbounded Mappings of Class (S)
+
. . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Generalizations for Quasimonotone Mappings . . . . . . . . . . . . . . . . . . . . . 16
3 Existence and Surjectivity Results 18
3.1 Noncoercive Mappings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2 Odd Mappings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.3 Injective Mappings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.4 Mappings of Type (M) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4 Further Applications 30
4.1 Multiplication Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2 On Solution Set Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3 On the Existence of Multiple Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 35
1
Introduction
Many problems in mathematics and mathematical physics can be reduced to a study of the set of
solutions for an equation f(x) = y, where f is a map between appropirate spaces X and Y , and
y is element of Y . Useful information about the solution set is obtained by means of topological
degree theory. The topological degree is mainly used in the study of ordinary and partial dierential
equations, integral equations, bifurcation theory and in proving xed point theorems. Degree theory
provides also a natural technique for proving so-called mapping theorems which include results
concerning the invariance of domain and generalizations of the Jordan curve theorem.
Let G be an open subset of some topological space X and f a given map of the closure of G
into another topological space Y . For each point y of Y outside the image of boundary of G under
f we assign an integer d(f, G, y). The integer-valued function d is said to be a topological degree
if it has certain normalizing, additivity and homotopy properties. The integer d(f, G, y) is a rough
estimate of the number of solutions for the equation f(x) = y. In particular, if d(f, G, y) diers
from zero, then there exists at least one solution of the equaiton f(x) = y.
The notion of the topological degree was rst introduced by Brouwer in 1912 [3] for continuous
mappings in nite demensional spaces. The Brouwer degree can be dened either by using algebraic
topology or analytically (see e.g. [11] and [19]). Leray and Schauder generalized in 1934 [18] the
degree theory for compact perturbations of identity in innite dimensional Banach spaces. The
construction of the Leray-Schauder degree is based on the Brouwer degree.
Since 1934, various generalizations of degree have been dened (see e.g. [19], [24]). However, it
was not untill 1972 and 1973 that F uhrer [13] and Amann and Weiss [1] proved the Brouwer degree
to be unquely determined by only a few conditions. These conditions provide a natural basis for
the formal denition of a classical topological degree (see [9], [19]).
In a series of articles [5], [6], [7], [8] and [9], F.E. Browder has recently extended the concept of a
topological degree to nonlinear mappings of monotone type from a real reexive Banach spaces into
its dual space. His method is based on Galerkin approximations for which the nite dimensional
Brouwer degree is well-dened.
The interest in the denition of nonlinear mappings of monotone type arises from the fact that
these properties can be veried under concrete hypotheses for the maps between Sobolev spaces
obtained from the elliptic operators of the generalized divergence form (see e.g. [9] p.18 and [17]).
In this paper we are mainly interested in the construction and uniqueness of the degree and its
direct consequences. In applications we restrict our discussion to an abstrat level. Our construction
of the degree is based on the Leray-Schauder degree and we shall follow the approach introduced
in [2].
In the rst chapter, we shall formally dene the concept of topological degree and examine
those classes of mappings for which the degree will be constructed later. We shall further extend
the notion of a degree, in a slightly modied form, to a broader class of quasimonotone mappings.
In the thrid chapter, we shall study how the degree theory can be used to deduce existence and
surjectivity results for equations involving nonlinear mappings of monotone type.
In our nal chapter, we rst study the degree of a composition of mappings; a generalization of
the multiplication theorem of the Leray-Schauder theory is achieved. At the end we examine briey
the connectedness of the solution set and the existence of multiple solutions.
For the general theory and applications we refer to [4], [9], [11], [19], [?] and for the denitions
and properties nonlinear mappings of monotone type to [23], [29]. Some elementary facts about the
weak topology and weak convergence are needed; we refer to [25], [28].
2
1 Preliminaries
1.1 Formal Degree Theory
Denition 1.1 Let X and Y be topological spaces and let O be a class of open subsets G of X. For
each G O, we associate a class F
G
of maps from G into Y . For each G O, we also associate
a class H
G
of maps [0, 1] G into Y (permissible homotopies); if H H
G
, then we simply write
h(t, x) = f
t
(x) and speak about the homotopy f
t
; 0 t 1, where t is called the parameter of the
homotopy f
t
. For any f F
G
; G O, and for any y Y f(G) , we associate an integer
d(f, G, y).
The integer-valued function d is said to be a classical topological degree if the following conditions
are satised:
(a) If d(f, G, y) ,= 0, there exists a solution of the equation f(x) = y in G.
(b) (Additivity) If D G O; D O and f F
G
, then the restriction f[
D
F
D
(the restricted
map is usually denoted by the same symbol). Let G
1
, G
2
be a pair of disjoint subsets of G
belonging to O and suppose that y / f(G (G
1
G
2
)). Then
d(f, G, y) = d(f, G
1
, y) +d(f, G
2
, y).
(c) (Invariance under Homotopy) If f
t
, 0 t 0, is a homotopy in H
G
, then f
t
F
G
for
each xed t [0, 1], and if y(t) : t [0, 1] is continuous curve in Y with y(t) / f
t
(G) for
any t [0, 1], then d(f
t
, G, y(t)) is constant in t [0, 1].
(d) (Normalizing) There exists a normalizing map j: X Y such that j[
G
F
G
for each
G O, and if y j(G), then
d(j, G, y) = 1.
The form of general theory is based on the classical degree theory in nite dimensional spaces
given by Brouwer in 1912 [3], where exactly the properties (a) - (d) listed above guarantee the
uniqueness of the degree function (see [1] or [9]). Indeed, if in Denition 1.1 we choose X = Y = R
n
,
O all open bounded sets, F
G
all continuous maps of G into R
n
, H
G
all continuous homotopies, i.e.,
continuous maps of [0, 1] G into R
n
, and j = I; the identity map, then we get the unique Brouwer
degree. The unique Leray-Schauder degree obtained by choosing X = Y a real innite dimensional
Banach space, O all open bounded sets, F
G
all continuous maps f : G X such that I f
is compact, i.e., it takes bounded sets into relatively compact sets, H
G
the class of continuous
homotopies of the form I T, where T : [0, 1] G X is compact, and j = I the identity map.
We shall now study more closely the conditions imposed on a degree function in Denition 1.1.
In view of our further study it is useful to make the following simplifying assumptions.
From now on we assume that X and Y are normed spaces and O consists of open and bounded
subsets of X. Suppose that H
G
contains at least the ane homotopies:
f
t
= (1 t)f
0
+tf
1
; t [0, 1],
where f
1
, f
2
F
G
. Since every f F
G
can be written in the form
f = (1 t)f +tf, t [0, 1],
3
f can be regarded as a constant homotopy of class H
G
. Moreover, assume that for any homotopy
f
t
in H
G
and for any continuous curve y(t): t [0, 1] in Y , also f
t
y(t) determines a permissible
homotopy. In particular, f y F
G
whenever f F
G
and y Y . Let us further assume that
f F
G
maps closed subsets of G onto closed subsets of Y . In particular, f(G) and f(G) are closed
in Y . Now we deduce three theorems which follow directly from the general structure of degree
theory. These results, much used below show characteristic features of every degree theory.
Theorem 1.1 (Boundary Dependence) Suppose that f, g F
G
are such that f = g on G and
y Y f(G). Then
d(f, G, y) = d(g, G, y).
Proof: Consider an ane homotopy f
t
= (1 t)f + th; t [0, 1]. Since f
t
(x) = f(x) for all
t [0, 1] and x G, we have y / f
t
(G) for any t in [0, 1], which by the property (c) of Deni-
tion 1.1 implies that d(f, G, y) = d(h, G, y).
For the next theorem, recall that by an open component of an open set D we mean an
equivalence class of points which can be connected by a continuous curve in D.
Theorem 1.2 Let G O and f F
G
be xed. Then d(f, G, ) is a function of the image point
y Y f(G) and constant on each open component of the open set Y f(G) .
Proof: Let be an open component of Y f(G) and let y and q be two xed elements in .
There is a continuous curve y(t) : 0 t 1 in which connects y and q. By the homo-
topy invariance property (c) we thus nd that d(f, G, y(t)) is constant in t on [0, 1]. In particular,
d(f, G, y) = d(f, G, q), which completes the proof.
Theorem 1.3 Let f F
G
and y Y f(G) be given. Then f y F
G
, 0 Y (f y)(G) and
d(f, G, y) = d(f y, G, 0).
Proof: Consider an ane homotopy f
t
= (1 t)f + t(f y), 0 t 1, and a continuous curve
y(t) = (1t)y, 0 t 1. Clearly y(t) / f
t
(G) for any t [0, 1]. Hence the assertion follows from
the homotopy invariance property (c).
In view of Theorem 1.2, it is relevant to dene the degree with respect to the component of
Y f(G) by setting
d(f, G, ) = d(f, G, y), where y .
Hence the values of d(f, G, y) and d(f, G, q) can dier only if y and q are in dierent components
of Y f(G). Thus the condition y / f(G) cannot be avoided in the denition of a degree, since
otherwise d(f, G, y) would be constant for all y, which contradicts the normalizing condition (d) of
Denition 1.1.
It is useful to examine to what extent the conditions from (a) to (d) are independent. We
can easily see that (a) follows from the additivity property of (b). Indeed, if we rst choose
G = G
2
= G
2
= and apply (b), then d(f, , y) = 0. If now f F
G
and y Y f(G), then by
choosing G
1
= G
2
= and applying (b) again we obtain d(f, G, y) = d(f, , y) + d(f, , y) = 0.
Thus if d(f, G, y) ,= 0, then y f(G) and hence y f(G) which is precisely the condition (a).
The conditions (b), (c) and (d) are independent to each other.
4
Remark 1.1 It should be noted that once a degree function is constructed for some class F
G
, then
it is usually possible to choose H
G
so simple that (c) is satised. On the other hand, the practical
value of a degree theory depends on the homotopies available. Thus it is important to pay attention
to the choice of permissible homotopies; at least to ane homotopies should be included.
Remark 1.2 It is useful to note that it usually suces to construct a degree function only with
respect to the origin. Moreover, suppose that the normalizing map j: X Y is bijective and its
inverse j
1
: Y X is locally bounded. Let us now suppose that a degree function d is constructed
with respect to the origin, i.e., it satises the conditions (a), (b) , (d) with y = 0 and the invariance
under homotopy (c) with respect to the curve y(t) = 0; t [0, 1]. Moreover, we dene
d(f, G, y) = d(f y, g, 0) for all y Y f(G). (1)
Then the conditions from (a) to (d) hold also in their original form. Indeed, conditions (a), (b) and
(c) are clearly satised. Too see (d), assume that y j(G) and denote j
t
= (1 t)j + t(j y) =
j ty; t [0, 1]. Then there exists R > 0 and k > 0 such that
|j
t
(x)|
Y
k for all t [0, 1] and for all |x|
X
R. (2)
The proof of this fact will be seen later. Hence 0 / j
t
(B
R
(0) for any t [0, 1]. The already known
properties of d together with (3) imply that
d(j, G, y) = d(j, B
R
(0), y) = d(j y, B
R
(0), 0) = d(j, B
R
(0), 0) = 1.
1.2 On the Mappings of Monotone Type
Let X be an innite dimensional real reexive Banach space and X

its dual space, and let , )


stand for the pairing between X and X

. By the results due to Trojanski, every reexive Banach


space has an equivalence norm such that X and X

are both locally uniformly convex. Thus we


assume from now on that X and X

are locally uniformly convex. This renorming is needed for the


denition of the duality mapping at the end of this section.
Denition 1.2 Let G be an open bounded subset of X. Let f: G X

.
1. The map f is said to be monotone if for all x, u G, we have
f(x) f(u), x u) 0, (3)
and strictly monotone if it is monotone and equality holds in (3) only if x = u.
2. The map f is said to be of class (S)
+
if for every sequence x
j
G with x
j
x X for
which
limsup
j
f(x
j
), x
j
x) 0,
we have x
j
x X.
5
3. The map f is said to be pseudomonotone if for every sequence x
j
G with x
j
x X
for which
limsup
j
f(x
j
), x
j
x) 0,
we have
lim
j
f(x
j
), x
j
x) = 0,
and if x G, then f(x
j
) f(x) in X

.
4. The map f is said to be quasimonotone if for every sequence x
j
G with x
j
x X,
we have
limsup
j
f(x
j
), x
j
x) 0.
Denition 1.3 Let f: X X

.
1. The map f is said to be bounded if it takes bounded subsets of X into bounded subsets of X

.
2. The map f is said to be locally bounded if for each x X there is a neighborhood U of x
such that f(U) is a bounded subset of X

.
3. The map f is said to be demicontinuous if for each x
j
in X such that x
j
x X, we
have
f(x
j
) f(x) in X

.
We note that every demicontinuous map is locally bounded. In fact, if there is x X such that,
for each j, f(B(x, 1/j)) is not bounded, then choosing x
j
B(x, 1/j) such that |f(x
j
)| j we see
that x
j
x in X and the sequence f(x
j
) is not bounded in X

. Hence f(x
j
) cannot converge
weakly in X

so that f is not demicontinuous. Also, it is easy to see that every demicontinuous


map of class (S)
+
is pseudomonotone. The class of demicontinuous quasimonotone maps is quite
large. Indeed, a demicontinuous map f: G X

is quasimonotone if for instance


(i) f is of class (S)
+
or pseudomonotone;
(ii) f is monotone and bounded or dened on the weak closure of G;
(iii) f is compact;
(iv) f is continuous from the weak topology of X to the strong topology of X

. This holds if f is
linear and compact.
In this paper the class (S)
+
is of great importance. Denote by F
G
the family of all demicontinuous
maps of class (S)
+
of G into X

and by F
G,B
all bounded maps in F
G
. If f F
G
, then the preimage
of any compact subset of X

is compact in X. To prove this we need the boundedness of closure of


G. As a special case we see that f
1
(y), the solution set of the equation f(x) = y on G, is compact
for each y X

. An immediate consequence is that f takes closed subsets of G into closed subsets


of X

. In particular, the sets f(G) and f(G) are closed.


For a demicontinuous pseudomonotone map f: G X

, f(A) is closed whenever A G is


weakly closed in X

, for instance closed and convex.


In view of application it is useful to know under what kind of perturbations F
G
remains stable.
This is the content of the next lemma.
6
Lemma 1.1 Let h: G X

be a demicontinuous map. Then the following conditions are equiva-


lent:
(i) h is quasimonotone;
(ii) For each sequence x
j
G with x
j
x, we have
liminf
j
h(x
j
), x
j
x) 0.
(iii) f +h F
G
for all f F
G
.
Proof: Assume that h is quasimonotone. If (ii) does not hold, then there exists a sequence x
j
G
with x
j
x such that liminf h(x
j
), x
j
x) < 0. Then, at least for a subsequence x
k
of x
j
,
we have
limsup
k
h(x
k
), x
k
x) = lim
k
h(x
k
), x
k
x) = liminf
j
h(x
j
), x
j
x) < 0,
which contradicts the assumed quasimonotonicity of h. Thus (ii) follows from (i).
Suppose next that (ii) holds true. Let f F
G
be given and let x
j
G such that x
j
x in
X with
limsup
j
f(x
j
) +h(x
j
), x
j
x) 0.
Now, in view of (ii), we have
limsup
j
f(x
j
), x
j
x) limsup
j
f(x
j
) +h(x
j
), x
j
x) liminf h(x
j
), x
j
x) 0,
which implies that x
j
x in X since f F
G
. Hence f +g F
G
and (iii) follows from (ii).
Assume now that (iii) holds true. If h is not quasimonotone, then there exists a sequence
x
j
G such that x
j
x in X and
limsup
j
h(x
j
), x
j
x) = q < 0.
Passing to a subsequence, if necessary, we may assume that
lim
j
h(x
j
), x
j
x) = q < 0.
Let f be a bounded map in F
G
. Such a map exists as we shall see later, for instance, the duality
mapping. It is clear that f is of class (S)
+
for any > 0 and by (iii) also f + h F
G
. Since
every map of (S)
+
is quasimonotone, we get
0 limsup
j
f(x
j
) +h(x
j
), x
j
x) = limsup
j
f(x
j
), x
j
x) q.
But the right-hand side cannot remain nonnegative for all > 0 since f is bounded. Hence we
reach a contradiction showing that h is quasimonotone and the proof is complete.
If f and g are in F
G
, then in view of the previous lemma
f F
G
for all > 0 and f +h F
G
.
Similarly, if f: G X

and h: G X

are monotone(pseudomonotone, quasimonotone), then f


for all 0 and f + h are monotone(pseudomonotone, quasimonotone). Hence these classes of
maps have a conical structure.
7
Remark 1.3 Let f F
G
be given and g: G X

compact. Then f + g F
G
for all > 0 by
Lemma 1.1. However, since also g is quasimonotone, we necessarily have that
lim
j
g(x
j
), x
j
x) = 0
for each sequence x
j
G with x
j
x in X. Thus g cannot be of the class (S)
+
and, consequently,
F
G
contains no compact maps.
Remark 1.4 If X is a Hilbert space, then certainly the identity map I is of class (S)
+
. Thus those
mappings for which the Leray-Schauder degree is dened, i.e., maps of the form I g, g compact,
are of class (S)
+
.
We shall now dene suitable homotopies in view of degree theory. A map H: [0, 1] G X

is
said to be a homotopy of class (S)
+
if for every sequence x
j
G with x
j
x in X and for
every sequence t
j
[0, 1] with t
j
t in [0, 1] for which
limsup
j
H(t
j
, x
j
), x
j
x) 0
we have x
j
x in X and H(x
j
, t
j
) H(x, t) in X

. We denote by H
G
(or H
G,B
) those homotopies
of class (S)
+
which are locally bounded(or bounded) as maps of [0, 1] G into X

. Clearly every
H H
G
is a demicontinuous map of [0, 1] G into X

and H(t, ) F
G
for any xed t in [0, 1]. It
is not hard to prove( see ......page 27) that H
G
contains all ane homotopies between any pair of
maps of F
G
. The same conclusion holds for H
G,B
. Furthermore, we say that a demicontinuous map
H: [0, 1] G X

is a quasimonotone homotopy if for every sequence x


j
G with x
j
x
in X and t
j
[0, 1] with t
j
t in [0, 1], we have
limsup
j
H(x
j
, t
j
), x
j
x) 0.
It is easy to see that quasimonotone homotopies include ane homotopies between any pair of
demicontinuous maps with common domain G. Clearly every H H
G
is also a quasimonotone
homotopy.
Remark 1.5 Pseudomonotone homotopies can also be dened. This class, although important in
application, is somewhat unsatisfactory since it does not contain ane homotopies. However, both
pseudomonotone homotopies and ane homotopies between pseudomonotone maps are contained
in the class of quasimonotone homotopies.
The following result generalizes Lemma 1.1; its proof is omitted since it is similar to that of
Lemma 1.1.
Lemma 1.2 Let h: [0, 1] G X

be a demicontinuous map. Then the following conditions are


equivalent:
(i) h is a quasimonotone homotopy;
(ii) For every sequence x
j
G and t
j
[0, 1] with x
j
x and t
j
t in [0, 1], we have
liminf
j
h(x
j
, t
j
), x
j
x) 0.
8
(iii) H +h F
G
for all H H
G
.
We close this section by recalling some facts about the duality mapping between X and X

.
Here we need the fact that X and X

are locally uniformly convex. It should be noted that Hilbert


spaces and reexive Sobolev spaces do not require any renorming. It follows form Hahn-Banach
theorem that the conditions
|J(x)| = |x| and J(x), x) = |x|
2
for all x X
determine a unique map J of X onto X

, called the duality mapping corresponding to the given


norm on X. It is bicontinuous, strictly monotone, bounded and of class (S)
+
. Moreover, J is
homogeneous, i.e., J(x) = J(x) for all x X and R.
Remark 1.6 If X is a Hilbert space, then the inverse of J is the linear isometry given by the
Frechet-Riesz theorem. On the other hand, if we assume that J is linear, we get
|x u|
2
= J(x u), x u) = J(x) Ju, x u) = |x|
2
+|u|
2
J(x), u) Ju, x)
which implies that
|x +u|
2
+|x u|
2
= 2
_
|x|
2
+|u|
2
_
for all x, u X. This means that X is a Hilbert space. Hence the duality mapping is linear if and
only if X is a Hilbert space.
Remark 1.7 The reexivity of X is closely related to the concept of the duality pairing. By
reexivity the map L dened by setting
Lx, w) = w, x) for all x X and w X

is a linear isometric homeomorphism X onto X

. Denote by J
1
the duality mapping of X

onto
X

. For any x X, we then have


|J
1
1
(Lx) = |Lx| = |x|,
and

J
1
1
(Lx), x
_
=

Lx, J
1
1
(Lx)
_
=

J
1
(J
1
1
(Lx)), J
1
1
(Lx)
_
= |J
1
1
(Lx)|
2
= |x|
2
.
Hence by denition and uniqueness of J we necessarily have
J
1
1
(Lx) = J(x) for all x X.
Thus J
1
J = L, which will be used later in section . Note that J
1
J is linear in spite of the
nonlinearity of both J
1
and J.
9
2 Construction of Degree
2.1 Bounded Mappings of Class (S)
+
Let X be an innite dimensional real reexive separable Banach space, renormed if necessary,
such that X and X

are locally uniformly convex. Let G be an open bounded subset of X, nonempty


unless otherwise indicated. Denote by F
G,B
all demicontinuous maps of class (S)
+
of G into X

and by H
G,B
all bounded homotopies of class (S)
+
of [0, 1] G into X

. As a normalizing map we
take the duality mapping J.
We rst recall the following embedding theorem due to Browder and Ton:
Proposition 2.1 Let X be a reexive separable Banach space. Then there exists a separable Hilbert
space W and a linear compact injection : W X such that (W) is dense in X.
Using Proposition 2.1 and Frechet-Riesz theorem we dene further a map

: X

W by setting
(

(w)[v) = w, (v)) for all v W, w X

(4)
where ([) stands for the inner product of W. It is easily seen that

is linear and compact, and
since (W) is dense in X, it follows that

is also injective.
To each f F
G,B
we now associate family of maps f

: > 0 dened by
f

= I +
1



f, > 0, (5)
where I denotes the identity map of X. For any xed > 0, f

maps G into X. Moreover, f

is of
type I g, g compact, a mapping for which the Leray-Schauder degree is dened.
We shall now examine how the family f

: > 0 is related to f. Since every map f F


G,B
is
also a constant homotopy, we focus our attention on the class H
G,B
.
Lemma 2.1 Assume that H H
G,B
and denote f
t
(x) = H(t, x) for all t [0, 1] and x G. If
A G is a closed set with 0 / f
t
(A) for any t [0, 1], then there exists

such that
0 / (f
t
)

(A) for all t [0, 1] and 0 < <

. (6)
Proof: If (6) does not hold, there exists a sequence
j
, t
j
[0, 1] and x
j
A with
j
0
+
and (f
t
j
)

j
(x
j
) = 0 for all j. At least for a subsequence we can write
x
j
x X, t
j
t [0, 1] and f
t
j
(x
j
) w X

.
We thus have
x
j
=
1


f
t
j
(x
j
), (7)
and since x
j
is bounded, we nd that

f
t
j
(x
j
) 0. On the other hand,

is compact and
linear and therefore

f
t
j
(x
j
)

(w). By the injectivity of

it can be deduced that w = 0,
i.e.,
f
t
j
(x
j
) 0. (8)
Using (4) and (7) we get

f
t
j
(x
j
), x
j
_
=
1

j
_
f
t
j
(x
j
),

f
t
j
(x
j
)
_
=
1

j
|

f
t
j
(x
j
)|
2
W
0.
10
Thus by (8)
limsup

f
t
j
(x
j
), x
j
x
_
= limsup
_

j
|

f
t
j
(x
j
)|
2
W
_
0,
which implies that x
j
x A and f
t
j
(x
j
) f
t
(x) in X

. Since the weak limit is unique, we nd


by (8) that f
t
(x) = 0. But this contradicts our hypothesis 0 / f
t
(A) for any t [0, 1]. Hence the
assertion (6) is true and the proof is complete.
If we restrict ourselves with a constant homotopy, i.e., a map f F
G,B
, and choose A = G,
then there exists

> 0 such that f

(x) ,= 0 for all x G and 0 < <

, provided f(x) ,= 0 for


all x G. This in turn means that the Leray-Schauder degree d
LS
(f

, G, 0) is well-dened for all


0 < <

.
Our next lemma shows that the value of d
LS
(f

, G, 0) remains stable as approaches zero.


Lemma 2.2 Let f F
G,B
and assume that f(x) ,= 0 for all x G. Then there exists

> 0 such
that f

(x) ,= 0 for all x G and 0 < <

and d
LS
(f

, G, 0) is constant for all 0 < <

.
Proof: The assertion follows from Lemma 2.1 and the homotopy invariance for the Leray-Schauder
degree. Indeed, if we take 0 <
1
<
2
<

and consider and ane Leray-Schauder homotopy


between f

1
and f

2
, it is easily seen that
d
LS
(f

1
, G, 0) = d
LS
(f

2
, G, 0).
By the preceding lemma it is now relevant to give
Denition 2.1 For any f F
G,B
with 0 / f(G), dene
d(f, G, 0) = d
LS
(f

, G, 0), 0 < <

, (9)
where

is given by Lemma 2.2. Consequently, for any y X

f(G),
d(f, G, y) = d(f y, G, 0). (10)
We are now in a position to prove that Denition 2.1 really gives us a classical topological degree
in the sense of Denition 1.1.
Theorem 2.1 The integer-valued function d dened by (9) and (10) for all f F
G,B
satises the
conditions (a) - (d) of Denition 1.1 with respect to the class H
G,B
of homotopies and with the
duality mapping J as a normalizing map.
Proof: In view of Remark 1.2, it is sucient to deal with the case y = 0 in (a), (b) and (d) and
y(t) = 0 , 0 t 1, in (c). We shall now verify the conditions (a)-(d) in this reduced form.
(a) Assume that d(f, G, 0) ,= 0. If 0 / f(G), we can apply Lemma 2.1 for A = G and, consequently,
we nd that there exists

> 0 such that d


LS
(f

, G, 0) = 0 for all (0,

). By (9) we get
d(f, G, 0) = 0, which is a contradiction. Thus 0 f(G).
(b) Let G
1
and G
2
be disjoint open subsets of G and assume that 0 / f(G (G
1
G
2
)). Applying
Lemma 2.1 for A = G (G
1
G
2
) we get

> 0 such that 0 / f

(A) for all (0,

). Since the
additivity property (b) holds for the Leray-Schauder degree, we have
d
LS
(f

, G, 0) = d
LS
(f

, G
1
, 0) +d
LS
(f

, G
2
, 0)
for all (0,

), and thus (b) follows by (9).


11
(c) Let H H
G,B
and, as usual, denote f
t
(x) = H(t, x). Assume that 0 / f
t
(G) for any t [0, 1].
For each xed t [0, 1] and > 0 we have
(f
t
)

= I +
1


f
t
: G X,
and by Lemma 4 ( for A = G) there exists

> 0 such that 0 / (f


t
)

(G) for all t [0, 1] and


(0,

). Since the homotopy H: [0, 1] D X

is demicontinuous and

is linear and compact,
the map
1


H: [0, 1] G X
is continuous for any xed > 0. Moreover, the boundedness of H together with the properties of


implies that
1


H
is also compact. Thus, for a xed , (f
t
)

determine a permissible Leray-Schauder homotopy. By


using Lemma 5 for each xed t [0, 1] and the Leray-Schauder homotopy invariance for each xed
(0,

), we have by (9)
d(f
t
, G, 0) = d
LS
((f
t
)

, G, 0) = constant
for all t [0, 1] and (0,

).
(d)Assume that 0 J(G), which implies that 0 G. We consider an ane homotopy between I
and J

. Since
J(x), (1 t)I(x) +tJ

(x)) = |x|
2
+
t

J(x)|
2
W
> 0
for all x ,= 0 and 0 t 1, and since J(x) = 0 only for x = 0, we have
0 / [(1 t)I +tJ

](G) for all t [0, 1] and > 0.


By the homotopy invariance for the Leray-Schauder degree we thus obtain
d
LS
(J

, G, 0) = d
LS
(I, G, 0) = 1 for all > 0.
Hence (d) follows from (9). The proof is now complete.
2.2 Uniqueness of Degree
Our next aim is to prove that the degree function given by Denition 4 is unique. Suppose there
exists a degree, say

d, which is dened for every map f F
G,B
and which satises the conditions
(a) - (d) of Denition 1.1 with respect to H
G,B
and J. We shall prove that d =

d. The proof is
based on the fact that the Leray-Schauder degree is unique. We rst give two lemmas.
Lemma 2.3 Let H = I T: [0, 1] G X be a Leray-Schauder homotopy, i.e., T is compact. Let
M be an open bounded subset of X such that H(t, x) M for all t [0, 1] and x G. For each
f F
M,B
, we have f H H
G,B
.
12
Proof: Clearly f H: [0, 1] G X

is bounded and demiconituous. Let sequences t


j
[0, 1]
and x
j
G be such that
_
x
j
x, t
j
t and
limsup
j
(f H)(t
j
, x
j
), x
j
x) 0.
(11)
Since T: [0, 1] G X is compact, there exists a subsequence T(t
k
, x
k
) T(t
j
, x
j
) and an
element u X such that
T(t
k
, x
k
) u in X. (12)
Thus
H(t
k
, x
k
) = x
k
T(t
k
, x
k
) x u in X,
and by denoting x u = z and H(t
k
, x
k
) = z
k
we get z
k
z. Consequently we can write by (12)
limsup
k
f(z
k
), z
k
z) = limsup
k
(f H)(t
k
, x
k
), x
k
T(t
k
, x
k
) x +u)
= limsup
k
(f H)(t
k
, x
k
), x
k
x)
limsup
j
(f H)(t
j
, x
j
), x
j
x)
0,
where the last inequality follows from (11). Since f is of class (S)
+
, we have z
k
z in X and thus
x
k
= z
k
+T(t
k
, x
k
) z +u = x.
The fact that the limit x is unique by (11) implies that the whole sequence x
j
converges to x,
i.e., x
j
x. We further have
(f H)(t
j
, x
j
) H(t, x),
which completes the proof.
Lemma 2.3 implies as a special case that J h F
G,B
whenever h: G X is a mapping of the
Leray-Schauder type, i.e., I h is compact. Moreover, J h is continuous and (J h)
1
(0) = h
1
(0),
since J(0) = 0 and the identity map is injective.
In our next lemma we use the uniqueness of the Leray-Schauder degree. We rst dene a new
degree function

d
LS
by setting

d
LS
(h, G, 0) =

d(J h, G, 0) (13)
for all Leray-Schauder mappings h: G X with 0 / h(G). Furthermore, if y Y h(G)), we
dene

d
LS
(h, G, y) =

d
LS
(h y, G, 0). (14)
Now we have
Lemma 2.4

d
LS
= d
LS
.
Proof: The assertion follows from the uniqueness of d
LS
if

d
LS
satises the conditions (a) - (d) of
Denition 1.1 in the form they have in the Leray-Schauder theory. It is sucient to deal with the
case y = 0 in (a), (b) and (d) and y(t) = 0, 0 t 1, in (c). The verications of the conditions
(a) - (d) is now eected by using the fact that

d is a degree function. In the proof of (c), invariance
under homotopy, we also use Lemma 2.3. Thus

d
LS
satises the conditions of the Leray-Schauder
degree and, by the uniqueness argument, we necessarily have

d
LS
= d
LS
.
We are now in a position to prove
13
Theorem 2.2 Suppose that

d is a degree function dened on F
G,B
satisfying conditions (a) - (d)
with respect to H
G,B
and J. Then

d coincides with the degree d given in Denition 2.1.
Proof:
2.3 Unbounded Mappings of Class (S)
+
In this section we extend the degree theory to all demicontinuous mappings of the class (S)
+
,
not only to bounded ones as in Section 2.1. By reducing suitably the domain of a demicontinuous
map of class (S)
+
we are able to use the results of Section 2.1 and 2.2 to dene the degree uniquely.
The denition is based on the following four lemmas.
Lemma 2.5 For each H H
G
and any compact set A X

, the set
K = x G: H(t, x) A for some t [0, 1]
is a compact subset of X.
Proof: We may suppose that K ,= . Let x
j
K be a given sequence. By the denition of
K there exists a sequence t
j
[0, 1] such that H(t
j
, x
j
) A for all j. For some subsequences
x
k
, t
k
and H(t
k
, x
k
) we can write t
k
t [0, 1], x
k
x X and H(t
k
, x
k
) w A.
Hence
limsup
k
H(t
k
, x
k
), x
k
x) = 0,
which implies that x
k
x and H(t
k
, x
k
) H(t, x) = w. Consequently, we have H(t, x) A which
means that x K. Thus every sequence in K has a strongly converging subsequence with the limit
in K, which proves the compactness of K.
Lemma 2.6 Let S G be any xed compact set and assume that H H
G
. Then there exists an
open set

G and R > 0 such that
(i) S

G G,
(ii) |H(t, x)| R for all t [0, 1] and x

G.
Proof: Let us assume that S ,= and introduce a family G
n
of sets, one of which satises (i)
and (ii). For each n = 1, 2, . . ., denote
D
n
=
_
x X: dist(x, S) <
1
n
_
,
where dist(x, S) denotes the distance of x from S. The compactness of S allows us to write D
n
as
D
n
=
_
x X: |x z| <
1
n
for some z S
_
.
Dening G
n
= D
n
G we get a family G
n
of open sets satisfying (i). If none of the sets G
n
satises (ii), there exist sequences t
n
[0, 1] and x
n
, x
n
G
n
, with
|H(t
n
, x
n
)| + as n . (15)
14
Since x
n
G
n
, there exists z
n
S such that |x
n
z
n
| 1/n and, moreover, the sequence z
k
S
has convergent subsequence, say z
k
z S. Thus
|x
k
z| |x
k
z
k
| +|z
k
z| 0,
which implies that x
k
z. We may assume that t
k
[0, 1] converges, say t
k
t [0, 1], and
by the demicontinuity of H we obtain H(t
k
, x
k
) H(t, z), which contradicts (15), since a weakly
convergent sequence is always bounded. Hence one of the sets G
n
, say G
n
, satises the conditions
(i) and (ii) ( in fact, the same is true for any G
n
, n n

), we denote

G = G
n
.
Remark 2.1 Lemma 2.6 actually means that by restricting the domain of H H
G
to [0, 1]

G
we have H H

G,B
.
By applying Lemma 2.6 to a constant homotopy f F
G
and to the compact set S = f
1
(0) we
get
Lemma 2.7 Assume that f F
G
and 0 / f(G). Then there exists an open set

G X such that
(i) f
1
(0)

G G,
(ii) f(

G) is bounded.
In view of Lemma 2.7 the restriction of f F
G
to G satises f F

G,B
and 0 / f(

G). Hence
the degree d(f,

G, 0) is well-dened.
Lemma 2.8 Let f F
G
be given with 0 / f(G). Then the value of d(f,

G, 0) does not depend on
the choice of

G, provided it satises the conditions (i) and (ii) of Lemma 2.7.
Denition 2.2 For any f F
G
with 0 / f(G), dene

d(f, G, 0) = d(f,

G, 0), (16)
where

G is any open set satisfying (i) and (ii) of Lemma . Consequently, if y X

f(G), we
dene

d(f, G, y) =

d(f y, G, 0). (17)
Theorem 2.3 The integer-valued function

d given by Denition 2.2 for any f F
G
satises the
conditions (a) - (d) of Denition 1.1 with respect of H
G
and J.
Proof:
Theorem 2.4 Let d
1
be any degree function which is dened on the class F
G
and which satises
the conditions (a) - (d) of Denition 1.1 with respect of H
G
and J. Then d
1
=

d.
Proof:
15
2.4 Generalizations for Quasimonotone Mappings
In this section we extend the concept of the topological degree to demicontinuous quasimonotone
maps f: G X

, where G is an open and bounded set X. However, we encounter here the diculty
that the image f(A) of a closed set A G is not necessarily closed. Therefore we shall modify
slightly the denition of a topological degree function.
Let f: G X

be a demicontinuous quasimonotone map, and let y be a point in X

such that
y / f(G). We associate to each such triple (f, G, y) an integer d
1
(f, G, y), which is said to be a
degree function in the extended sense or weak sense if the following conditions are satised:
(a

) If d
1
(f, G, y) ,= 0, then y f(G).
(b

) Let G
1
and G
2
be two open disjoint subsets of G and assume that y is a point of X

such
that y / f(G (G
1
G
2
). Then d
1
(f, G, y) = d
1
(f, G
1
, y) +d
1
(f, G
2
, y).
(c

) Let H: [0, 1] G X

be a quasimonotone homotopy, and y(t): 0 t 1 be a continuous


curve in X

. Denote H(t, x) = f
t
(x) and assume that there exists r > 0 such that for the ball
B
r
(y(t)) = w X

: |w y(t)| < r, we have f


t
(G) B
r
(y(t)) = for all 0 t 1. Then
d
1
(f
t
, G, y(t)) constant in t on [0, 1].
(d

) d
1
(J, G, y) = 1 if y J(G).
It easy to see that if d
1
exists, it determines a topological degree function also for the class F
G
in the sense of Section 2.3. By the uniqueness of

d,
d
1
(f, G, y) =

d(f, G, y) (18)
for all h F
G
with y X

such that y / h(G).


By using the same procedure as in Browder in [9], where he considers pseudomonotone maps,
we shall now construct a weak degree function for a boarder class of demicontinuous maps.
By Lemma 1.1, f +J F
G
for all > 0 if f: G X

is a demicontinuous quasimonotone map.


Similarly, by Lemma 1.2, H+J H
G
for all > 0 whenever H: G[0, 1] X

is a quasimonotone
homotopy.
The following two lemmas show how f +J is related to f.
Lemma 2.9 Let f: G X

be a demicontinuous quasimonotone map, A a subset of G, and y is a


point of X

such that y / f(A). Then there exists

> 0 such that y / (f +J)(A) for all 0 < <

.
Proof: If the assertion is not true, then there exist sequences
j
, 0
+
, and x
j
A with
f(x
j
) +
j
J(x
j
) = y for all j. Thus f(x
j
) = y
j
J(x
j
) y and, consequently, y f(A), which
leads to a contradictions.
Lemma 2.10 Let f: G X

be a demicontinuous quasimonotone map, A a subset of G, and y is


a point of X

such that y / f(G). Then



d(f +J, G, y) is constant for all 0 < <

, where

> 0
is given by Lemma 2.9 for A = A.
Proof: Let 0 <
0
< epsilon
1
<

be xed and consider an ane homotopy


(1 t)(f +
0
J) +t(f +
1
J) = f +
t
J,
where
t
= (1 t)
0
+ t
1
, 0 t 1, is between
0
and
1
. Hence y / (f +
t
J)(G) for any
t [0, 1], and the assertion follows from the homotopy invariance property of

d.
16
Denition 2.3 For any demicontinuous quasimonotone map f: G X

and for any y X

f(G)
we dene
d
QM
(f, G, y) =

d(f +J, G, y), 0 < <

, (19)
where

> 0 is given by Lemma 2.9 for A = A.


Theorem 2.5 The integer-valued function d
QM
given by Denition 2.3 is a degree function in
the weak sense on the class of all demicontinuous quasimonotone maps, which is invariant under
quasimonotone homotopies and normalized by J.
Proof:
Theorem 2.6 The degree function d
QM
in the weak sense given by Denition 2.3 is unique.
Proof:
Remark 2.2 In the view of the previous sections, it is appropriate to denote by the same symbol
d all the degree functions d,

d and d
QM
dened in this chapter.
17
3 Existence and Surjectivity Results
3.1 Noncoercive Mappings
In this section we deduce some existence and surjectivity results, which follow directly from the
basic properties of the degree function (cf. [2]). Indeed, this shows the power of the degree theory
once constructed. Results in this direction are given, for instance, in [23] without using degree
theory.
Lemma 3.1 Let f F
G
be given and assume that there exists x G such that
f(x), x x) > |f(x)||x x| for all x G. (20)
Then d(f, G, 0) = 1, and there exists at least one x
0
G such that f(x
0
) = 0.
Proof: (Cf. [2].) It follows from (20) that 0 / f(G) and therefore d(f, G, 0) is dened. Let

J
stand for the mapping

J(x) = J(x x) for all x X. By using (20) and ane homotopies we get
d(f, G, 0) = d(

J, G, 0) = d(J, G, J( x)),
where d(J, G, J( x)) = 1 since J( x) J(G). Thus d(f, G, 0) = 1 which implies that 0 f(G).
Indeed, we consider the ane (S)
+
homotopy
f
t
(x) = t

J(x) + (1 t)(J(x) J( x)), x G.
Since
f
t
(x), x x) = t J(x x), x x) + (1 t) J(x) J( x), x x)
we conclude, in view of the properties of J, that
f
t
(x) / 0 for all x G, t [0, 1].
Otherwise, for some x G and t [0, 1], we have
0 = t|x x|
2
+ (1 t) J(x) J( x), x x) .
It is clear that t ,= 1. Then we have
J(x) J( x), x x) =
t
1 t
|x x|
2
0
which implies that x = x (since J is strictly monotone); a contradiction. By homotopy invariance,
we have
d(

J, G, 0) = d(J J( x), G, 0) = d(J, G, J( x)).
We next consider the (S)
+
homotopy
g
t
(x) = t

J(x) + (1 t)f(x), x G, t [0, 1].
Obviously, g
t
(x) ,= 0 for all x G with t = 0, 1. In fact, g
t
(x) ,= 0 for all x G and t [0, 1].
Indeed, if g
t
(x) = 0 for some x G and t (0, 1), then
f(x) =
t
1 t
J(x x)
18
so that
|f(x)| =
t
1 t
|x x|.
Consequently,
f(x), x x) =
t
1 t
J(x x), x x) =
t
1 t
|x x|
2
= |f(x)||x x|;
contradicting (20). Hence, the homotopy invariance implies that
d(f, G, 0) = d(

J, G, 0) = d(J, G, J( x)) = 1.
Remark 3.1 The condition (20) is in fact a geometric one. It is easy to show that (20) is equivalent
to the condition:
f(x) ,= J(x x) for all x G, alpha 0,
which further means that there exists as ane homotopy between f and

J never vanishing on the
boundary G.
Remark 3.2 It should be noted that by Holders inequality we always have
f(x), x x) |f(x)||x x|.
Theorem 3.1 Let f: G X

be a demicontinuous quasimonotone map such that f(G) is closed


in X

and suppose that x G with


f(x), x x) > |f(x)||x x| for all x G. (21)
Then there exists at least one x
0
G such that f(x
0
) = 0 and d(f, G, 0) = 1 whenever dened.
Proof: Assume that 0 f(G). Since f(G) is closed, we nd that 0 f(G). Clearly, 0 / f(G)
by (21) and thus necessarily 0 f(G).
Assume next that 0 / f(G) which implies that d(f, G, 0) is well-dened. For any > 0 we
consider a map of class (S)
+
,
h

(x) = f(x) +J(x x), x G.


Using (21), properties of J and the triangle inequality, we get
h

(x), x x) = f(x), x x) + J(x x), x x)


> |f(x)||x x| +|J(x x)||x x|
|f(x) +J(x x)||x x|
= |h

(x)||x x| for all x G,


so that h

satises the assumptions of Lemma 3.1 for all > 0 and, consequently, we have
d(h

, G, 0) = 1 for all > 0.


Thus for each > 0 there exists x

G such that h

(x

) = 0. Choosing
j
0
+
, we have
f(x

j
) =
j
J(x

j
x) 0 as j .
19
Since f(x

j
) f(G) f(G), we conclude that 0 f(G) = f(G). Since 0 / f(G) we have
0 f(G).
Also, it is not hard to show that there exists
1
> 0 such that
(1 t)(f(x) +J(x)) +th

(x) ,= 0
for all t [0, 1], (0,
1
) and x G. For if there exist sequences t
j
[0, 1],
j
with
j
0
+
and x
j
G, then
f(x
j
) = (1 t
j
)
j
J(x
j
)
j
t
j
J(x
j
x) 0 as j
so that 0 f(G) which contradicts the assumption that 0 / f(G). Hence, by Denition 2.3 and
the homotopy invariance for d,
d(f, G, 0) = d(f +J, G, 0) = d(h

, G, 0) = 1
for any > 0 with 0 < <
1
. Thus 0 f(G), and since 0 / f(G) and f(G) is closed, we see
there to exist x
0
G such that f(x
0
) = 0.
Remark 3.3 If y is any element of X

, we can replace f by f y in Theorem 3.1. Hence the


solvability of the equation f(x) = y follows, provided there exists some x G (depending on y)
such that
f(x) y, x x) > |f(x) y||x x| for all x G.
Remark 3.4 A natural application of Theorem 3.1 occurs when G X is convex and f: G X

is pseudomonotone. Indeed, a closed convex set is weakly closed and the image of a weakly closed
set of X under a demicontinuous pseudomonotone map is always weakly closed.(see [2]).
Remark 3.5 For d we always have y f(G) if d(f, G, y) ,= 0 and f(G) is supposed to be closed.
However, the duality mapping is the only map for which we know the converse to be true, i.e.,
y J(G) implies that d(J, G, 0) ,= 0. In view of application it is desirable to nd other such maps
for which
d(f, G, y) ,= 0 if and only if y f(G). (22)
Assume that f: X X

is a demicontinuous monotone map, G is an open bounded set such


that f(G) is closed, and y is a point of X

with y / f(G). If f(x


0
) = y for some x
0
G, then
f(x) y, x x
0
) 0 > |f(x) y||x x
0
| for all x G,
which implies by Theorem 3.1 that d(f, G, y) = 1. Thus (22) holds.
Our next result gives sucient conditions for a quasimonotone map to be surjective. Let
f: X X

be a given map. We say that f has the property (B) if


(B) for any y X

there exists a neighborhood U of y such that f


1
(U) is bounded.
It is easy to see that the property (B) can be stated equivalently as
(B

) If x
j
X is such that f(x
j
) y X

, then x
j
is bounded.
It is useful to note that f has the property (B) if for instance
|f(x)| as |x| ,
which equivalently means that f
1
(C) is bounded for any bounded set C of X

.
20
Theorem 3.2 Let f: X X

be a demicontinuous quasimonotone map. Assume that f(B


R
) is
closed for each ball x X: |x| < R, R > 0, and that f has the property (B). If there exists
R > 0 such that
f(x), x)
|x|
+|f(x)| > 0 for all |x| R, (23)
then f(X) = X

, i.e., the equation f(x) = y admits a solution for any y X

.
Proof: Let y X

be xed. We can choose R

R and k > 0 such that


|f(x) ty| k for all t [0, 1] and |x| R

.
Indeed, suppose on the contrary that there exists sequences x
j
X with |x
j
| and t
j

[0, 1] such that |f(x
j
) t
j
y| 0 as j . We can assume that t
j
t, which implies f(x
j
) ty.
By the property (B), x
J
is bounded, which contradicts our assumption |x
j
| . By choosing
some
0
> 0 with
0
< (c/R

), we have for all |x| = R

and 0 < <


0
,
|f(x) +J(x) ty| |f(x) ty| |J(x)| c
0
R

> 0.
Hence (f + J)(x) ,= ty for all t [0, 1], 0 < <

and x B
R
. By Denition 2.3 and the
homotopy invariance for (S)
+
-mappings we get
d(f, B
R
, y) = d(f +J, B
R
, y) = d(f +J, B
R
, 0) = d(f, B
R
, 0).
By (23) we have
f(x), x) > |f(x)||x| for all |x| = R

and nd that the assumptions of Theorem 3.1 hold for x = 0. Thus


d(f, B
R
, y) = d(f, B
R
, 0) = 1,
which implies that y f(B
R
). Since f(B
R
) is closed, we have y f(B
R
), which completes the
proof.
Remark 3.6 The assumptions of Theorem 3.2 are satised if, for example, f: X X

is a demi-
continuous coercive pseudomonotone map, i.e.,
f(x), x)
|x|
+ as |x| +.
Some concrete examples of mappings satisfying the condition (23) but not coercive can be found in
[27].
Example 3.1 Let f: X X

be a demicontinuous monotone map which is expanding, i.e., there


exists k > 0 such that
|f(x) f(u)| k|x u| for all x, u X. (24)
By (24), |f(x)| as |x| ; thus f has the property (B). Moreover, since X is a Banach
space, f(A) is closed for each closed subset A of X. Indeed, if y
j
f(A) with y
j
y, we can
write y
j
= f(x
j
), x
j
A, and x
j
is a Cauchy sequence by (24). Thus x
j
converges, say
x
j
x A, and, consequently, f(x
j
) f(x) = y f(A). We can assume that f(0) = 0(if not,
we shall study the map f f(0)). The monotonicity of f and property (B) imply that there exists
R > 0 such that (23) holds. Hence, by Theorem 3.2, f(X) = X

(see also Section 3.3 where we


study the surjectivity of the expanding maps further).
21
3.2 Odd Mappings
In this section we shall consider the generalization of Borsuks theorem of the Leray-Schauder
theory concerning the degree of odd mappings. It can be stated as follows (see [19], [24]).
Proposition 3.1 (Borsuks theorem) Let G be an open bounded subset of X containing the
origin and symmetric, i.e., x G whenever x G and let f: G X be a map of the Leray-
Schauder type, that is I f is compact. If 0 / f(G) and f(x) = f(x) for all x G, then
d
LS
(f, G, 0) is an odd number.
We rst consider odd mappings of class (S)
+
(cf, [23] p 223). Since we do not presuppose any
boundedness of mappings, we need the following lemma.
Lemma 3.2 Let G be an open bounded symmetric subset of X with 0 G, and f F
G
be given
map with 0 / f(G). Assume that f(x) = f(x) for all x G. Then there exists a symmetric
open set

G such that
(i) f
1
(0)

G G, and
(ii) f(

G) is bounded.
Proof: The solution set f
1
(0) = x G: f(x) = 0 is obviously symmetric and 0 f
1
(0). We
have already shown in the proof of Lemma 2.6 that one of the sets
G
n
=
_
x G: dist
_
x, f
1
(0)
_
<
1
n
_
,
n = 1, 2, . . ., satises conditions (i) and (ii). The assertion now follows provided G
n
is symmetric.
Let x G
n
, then there exists z f
1
(0) such that |x z| < 1/n, and thus |(x) (z)| =
|x z| < 1/n with z f
1
(0), which implies that x G
n
.
Let G X be as in the previous lemma, and let f
1
: G X

stand for the map


f
1
(x) = f(x) for all x G.
Then f
1
is demicontinuous and of class (S)
+
whenever f has these properties.
Theorem 3.3 Let G be an open bounded symmetric set of X with 0 G, and let f F
G
be odd
in G. Then there exists in G a solution of the equation f(x) = 0 and , moreover, d(f, G, 0) is an
odd number whenever dened.
Proof: If 0 f(G), the assertion follows. Assume therefore that f(x) ,= 0 for all x G. Dene
a new map of class (S)
+
by setting
h(x) =
1
2
_
f(x) f(x)
_
, x G.
Clearly h is odd on G and coincides with f on G. Hence
d(f, G, 0) = d(h, G, 0). (25)
22
According to Lemma 3.2, there exists an open symmetric subset

G of G such that h
1
(0)

G and
the restriction h:

G X

is bounded. By the additivity property and Denition 2.1 we get


d(h, G, 0) = d(h,

G, 0) = d
LS
(h

,

G, 0) for all 0 < <

, (26)
where
h

= I +
1


h
and

> 0 is given by Lemma 2.2. Since h is odd, h

is odd on G. Consequently, by Proposition ??,


d
LS
(h

,

G, 0) = odd for all 0 < <

. (27)
The relations (25), (26) and (27) allow us to deduce that d(f, G, 0) is an odd number, which certainly
implies that 0 f(G).
Corollary 3.1 Let G be an open bounded and symmetric subset of X with 0 G and let f F
G
be such that 0 / f(G) and
f(x)
|f(x)|
,=
f(x)
|f(x)|
for all x G. (28)
Then d(f, G, 0) is odd and there exists in G a solution of the equation f(x) = 0.
Proof: Let us consider an ane homotopy of class (S)
+
given by
f
t
(x) = (1 t)f(x) +th(x), t [0, 1], x G,
where
h(x) =
1
2
_
f(x) f(x)
_
, x G.
If for some t [0, 1] and x G we have f
t
(x) = 0 then
_
1
t
2
_
f(x) =
t
2
f(x),
which contradicts (28). Thus d(f, G, 0) = d(h, G, 0) = odd, where the last equality follows from the
fact that h is odd on G. In particular, d(f, G, 0) ,= 0, which implies that 0 f(G). Hence the proof
is complete.
Our next theorem is applicable if for instance f is psuedomonotone and G is convex.
Theorem 3.4 Let G be an open bounded symmetric subset of X with 0 G and let f: G X

be a demicontinuous quasimonotone map such that f(G) is closed. If F is odd on G, then there
exists in G a solution of the equation f(x) = 0 and d(f, G, 0 = odd whenever dened.
Proof: If 0 f(G) f(G), the assertion follows. Thus we can assume that 0 / f(G).
Since (f + J)(x) = (f + J)(x) whenever f(x) = f(x), we conclude by Denition 2.3 and
Theorem 3.3 that
d(f, G, 0) = d(f +J, G, 0) = odd for all 0 < <

,
where

> 0 is given by Lemma 2.9 for A = G. Hence 0 f(G) f(G), which completes the
proof.
23
3.3 Injective Mappings
It is known from the Leray-Schauder theory that for each injective map I g: G X, g compact
with y (I g)(G), we have
d
LS
(f, G, y) = 1.
The proof of this result in based on the symmetry between I g and its inverse map [19], and there
seems to be no analogous proof for maps of class (S)
+
. However, by means of odd mappings we
shall derive a weaker result
d(f, G, y) = odd,
whenever f F
G
is a continuous injection and y f(G). An essential point is of course the fact
that the value of degree is nonzero.
We rst give an auxiliary result which entitles us to assume that 0 G, thus simplifying the
proofs. Let G be an open bounded subset of X and x
0
G any xed point. Denote
s(x) = x x
0
and s
1
(x) = x +x
0
, x X.
Now if f F
G
, then clearly f s
1
F
s(G)
. Because f(x) = (f s
1
)(s(x)) for all x G, the
degrees d(f, G, 0) and d(f s
1
, s(G), 0) are both dened if and only if 0 / f(G). Hence 0 s(G),
and it can be proved that
d(f, G, 0) = d(f s
1
, s(G), 0). (29)
Indeed, the right-hand side of (29) determines a degree function for the class F
G
in the sense of
Section 2.3. Thus (29) follows from the uniqueness of d.
The following lemma shows how to obtain a permissible homotopy of class (S)
+
between the
injection f and an odd mapping (cf. [19] p. 66). It should be noted that the continuity of f appears
necessary.
Lemma 3.3 Let B X be an open ball with center at the origin and f: B X

be a continuous
bounded injection of class (S)
+
with f(0) = 0. Then
H(t, x) = f
_
x
1 +t
_
f
_
tx
1 +t
_
t [0, 1] x B
denes a bounded homotopy of class (S)
+
.
Proof: The map H: [0, 1] B X

is clearly bounded and continuous. Let t


j
[0, 1] and
x
j
B be such that
t
j
t, x
j
x, t [0, 1], x X, and
limsup
j
H(t
j
, x
j
), x
j
x) 0.
Denote f
1
(x) = f(x), u
j
= (1 +t
j
)
1
x
j
, v
j
= t
j
u
j
, u = (1 +t)
1
x and v = tu. Now u
j
u and
v
j
v.
(a) Suppose rst that t ,= 0. Since f is bounded and t
j
,= 0 for suciently large j, we can write
0 limsup
j
H(t
j
, x
j
), x
j
x)
= limsup
j
[ f(u
j
), x
j
x) +f
1
(v
j
), x
j
x) ]
24
= limsup
j
_
(1 +t
j
)
_
f(u
j
), u
j

x
1 +t
j
_
+
1 +t
j
t
j
_
f
1
(v
j
), v
j

t
j
x
1 +t
j
_
_
= limsup
j
_
(1 +t
j
) f(u
j
), u
j
u) + (1 +t
j
)
_
f(u
j
),
x
1 +t

x
1 +t
j
_
+
1 +t
j
t
j
f
1
(v
j
), v
j
v) +
1 +t
j
t
j
_
f
1
(v
j
),
tx
1 +t
j

t
j
x
1 +t
j
_
_
= limsup
j
_
(1 +t
j
) f(u
j
), u
j
u) +
1 +t
j
t
j
f
1
(v
j
), v
j
v)
_
limsup
j
(1 +t
j
) f(u
j
), u
j
u) + liminf
j
1 +t
j
t
j
f
1
(v
j
), v
j
v)
= limsup
j
(1 +t) f(u
j
), u
j
u) + liminf
j
1 +t
t
f
1
(v
j
), v
j
v)
limsup
j
(1 +t) f(u
j
), u
j
u) ,
where the last inequality follows from the fact that (see Lemma 1.1)
liminf
j
f
1
(v
j
), v
j
v) 0.
Hence
limsup
j
f(u
j
), u
j
u) 0,
which implies that u
j
u. Thus x
j
= (1 +t
j
)u
j
(1 +t)u = x and H(t
j
, x
j
) H(t, x).
(b) Suppose now that t = 0. Then u = x, u
j
= (1 +t
j
)
1
x
j
x and v
j
= t
j
u
j
0. It follows from
the continuity of f that f
1
(v
j
) f
1
(0) = 0. Since u
j
x
j
0, we can write
0 limsup
j
H(t
j
, x
j
), x
j
x)
= limsup
j
[ f(u
j
), x
j
x) +f
1
(v
j
), x
j
x) ]
= limsup
j
f(u
j
), x
j
u
j
+u
j
x)
= limsup
j
f(u
j
), u
j
x) ,
and thus u
j
x implying x
j
x. Hence the proof is complete.
Theorem 3.5 Let G be an open bounded subset of X and f F
G
a continuous injection. If
y f(G), then d(f, G, y) = odd.
Proof: Since f y F
G
is also a continuous injection and d(f, G, y) = d(f y, G, 0), we may
assume without loss of generality that y = 0. Denote x
0
= f
1
(0) G. Since f is locally bounded,
there exists an open ball B G with center at x
0
such that the restriction f: B X

is bounded.
Moreover, we have by additivity
d(f, G, 0) = d(f, B, 0). (30)
Suppose that x
0
= 0. In view of Lemma 3.3 we consider the homotopy
f
t
(x) = f
_
x
1 +t
_
f
_
tx
1 +t
_
, t [0, 1], x B.
25
Now f
0
(x) = f(x) and f
1
(x) = f(x/2) f(x/2), which is odd everywhere on B. Consequently,
f
t
(x) ,= 0 for all t [0, 1] and x B. Hence by the homotopy invariance and Theorem 3.3,
d(f, B, 0) = d(f
1
, B, 0) = odd,
which together with (30) completes the proof in the case x
0
= 0.
If x
0
= f
1
(0) ,= 0, we use (29) and by the earlier part of the proof we have
d(f, G, 0) = d(f s
1
, s(G), 0) = odd,
which completes the proof.
As an application of Theorem 3.5 we shall prove the invariance of domain of a continuous
injection of class (S)
+
(cf, [19] p.66).
Theorem 3.6 Let D be an open subset of X ( not necessarily bounded) and f: D X

a continuous
injection of class (S)
+
. Then the set f(D) is open in X

.
Proof: Let y f(D) be xed and denote x
0
= f
1
(y) D. Let B be a ball with center at x
0
such
that B D. Now
y = f(x
0
) F(B) f(D)
by injectivity of f, y / f(B). This implies that d(f, B, y) is well-dened, and it is known in
addition that the value of d(f, B, ) is constant on every open component of the open set X

f(B).
Consequently, there exists r > 0 such that
d(f, G, q) = d(f, B, y) for all |q y| < r.
By Theorem 3.3 we have d(f, B, y) = odd, which gives
d(f, B, q) ,= 0 for all |q y| < r,
and therefore we obtain x
q
B(for all q X

, |q y| < r) such that f(x


q
) = q. Thus y f(D)
has an open neighborhood in f(D), namely,
B
r
(y) = q X

: |q y| < r f(B) f(D),


and since y f(D) was arbitrary, the assertion follows.
By using the open mapping theorem we get our next surjectivity result.
Theorem 3.7 Let f: X X

be continuous injection of class (S)


+
and assume that f has the
property (B). Then f(X) = X

.
Proof: In view of Theorem 3.6 (for D = X) it suces to prove that f(X) is closed in X

. Indeed,
if y
j
f(X), y
j
y, we can write y
j
= f(x
j
) and the sequence x
j
X is bounded. Without
loss of generality we may assume that x
j
x for some x X. Hence
limsup
j
f(x
j
), x
j
x) = lim
j
y
j
, x
j
x) = 0,
which implies x
j
x in X and, consequently, y
j
= f(x
j
) f(x) = y f(X). Thus f(X) is both
open and closed, implying f(X) = X

.
26
Corollary 3.2 Let f: X X

be a continuous map of class (S)


+
which is expanding, i.e., there
exists k > 0 such that
|f(x) f(u)| k|x u| for all x, u X. (31)
Then f(X) = X

.
Proof: By (31) the map f is injective and the inverse map f
1
is bounded because
|w y| k|f
1
(w) f
1
(y)| for all w, y f(X).
The assertion now follows from Theorem 3.7.
Corollary 3.2 relates to the following Problem (P), which was stated by Nirenberg ([22] p. 175)
in the special case if Hilbert space X.
Problem (P): Assume that X is a Banach space and let T: X X be continuous, expanding map
whose range contains an open set. Does T map X onto X?
Morel and Steinlein gave in 1984 [21] a counterexample which shows that the answer is generally
negative. However, there are several partial positive answers to Problem (P), for instance in the
following cases ( see [21]):
(i) X is nite dimensional;
(ii) T = I g, where g is compact or contraction or, more generally, a k-net contraction;
(iii) X is a Hilbert space and T a strongly monotone map, i.e.,
(T(x) T(u)[x u) c|x u|
2
for all x, u X(c > 0).
Let X be a real separable Hilbert space, and let T: X X be continuous, expanding map
of class (S)
+
, where we have identied X

with X. In view of Corollary 3.2, T(X) = X. This


is a generalization of (iii), since every strongly monotone map is of class (S)
+
. Consequently, if
T: X X is a monotone and expanding, then T(X) = X by Example 3.1, even if T is only
demicontinuous. As we showed in Example 3.1, the image of a closed set under any demicontinuous
expanding map is closed. Using this face we get
Theorem 3.8 Let X be a real separable Hilbert space and let T: X X be a continuous quasi-
monotone map which is expanding. Then T(X) = X.
Proof: Certainly, T +I is of class (S)
+
for all > 0. Since |T(x) T(u)| k|x u|,
|(T +I)(x) (T +I)(u)| (k )|x u| for all x, u X.
Thus T +I is expanding for any 0 < < k and, by Corollary 3.2, (T +I)(X) = X. For any y X
there exists x

, 0 < < k, such that T(x

) + x

= y, which implies, as 0
+
, that T(x

) y.
Since T(X) is closed, y T(X) and thus T is surjective.
We close this section by some observations concerning the maps of X

into X.
27
Remark 3.7 The reexivity of X implies that the map L dened by
Lw, x) = w, x) for all x in X and w in X

is a linear isometric homeomorphism of X onto X

. Let M be an open bounded set of X

and
h: M X a given map. We say that h is of class (S)

+
if Lh is a map of class (S)
+
of X

into X

.
We can similarly dene homotopies of class (S)

+
. For any such h we can dene a degree function
uniquely by setting

d(h, M, x) = d

(L h, M, Lx),
where x X h(M) and d

is the degree function for demicontinuous maps of class (S)


+
of X

into X

. By Remark 1.7, J

= LJ
1
is the duality mapping of X

into X

. Thus the normalizing


map for

d is J
1
, the inverse of J.
Let now f F
B,G
be a continuous injection and denote M = f(G). Then, by Theorem 3.6, M
is an open and bounded set of X

and we clearly have M = f(G) and M = f(G). It is easy


to see that f
1
: M X is a continuous bounded injection of class (S)

+
. Thus the degree of the
inverse function,

d(f
1
, M, x), is dened whenever x / f
1
(M), i.e., x / G.
The fact that the open mapping theorem holds for continuous injections of class (S)
+
, together
with the observation that the situation is symmetric with respect to f and f
1
, leads to the following
open question (generalization of Jordans theorem):
If D and D

are open bounded subsets of X and X

, respectively, and f: D D

is a homeomorphism of class (S)


+
, do X \ D and X

\ D

have the same number of


components ?
This question arises naturally since the generalization of the Jordan curve theorem holds for
the homeomorphism of the type I g, g compact, form X into X (see [24] p. 94), and there the
proof is based on the symmetry between I g and its inverse map. However, there seems to be no
analogous proof in the case of maps of class (S)
+
.
3.4 Mappings of Type (M)
Here we outline the direct application of the procedure introduced in Section 2.1, without
constructing the degree function. A function f: X X

is said to be of type (M) if we have


f(x) = w for every sequences x
j
, x
j
x in X, and f(x
j
), f(x
j
) w in X

, for which
limsup
j
f(x
j
), x
j
x) 0.
Maps of type (M) are widely used in applications; see for instance [23]. Clearly every locally
bounded map of type (M) is demicontinuous. It is not hard to see that every demicontinuous map
f: X X

which is of class (S)


+
or psuedomonotne is also of type (M). However maps of type
(M) are generally not quasimonotone and no relevant degree theory seems to be available for them.
Let f: X X

be abounded map of type (M). As in Section 2.1 the map


f

= I +
1



f: G X
is of type I g, g compact, for each > 0. In order to obtain existence results we need the following
Lemma 3.4 Let f: X X

be a bounded map of type (M). If there exist sequences


j
,
j
0
+
,
and x
j
in X, x
j
x, such that f

j
(x
j
) = 0, then f(x
j
) f(x) = 0.
28
Proof: Since f(x
j
) is bounded, there exists a weakly converging subsequence, say f(x
k
) w.
Thus

f(x
k
)

(w), and since
x
j
=
1


f(x
j
), (32)
we also have

f(x
k
) 0 and, consequently, w = 0, i.e., f(x
k
) 0. This fact does not depend on
the particular sequence f(x
k
), and hence we conclude that
f(x
j
) 0. (33)
By making use of (32) and (33) we get
limsup

f(x
j
), x
j
x)
= limsup
j
_
f(x
j
),
1


f(x
j
)
_
= limsup
j
_

j
|

f(x
j
)|
2
W
_
0,
and since f is of type (M), we deduce that f(x
j
) f(x) = 0, which completes the proof.
As an application of Lemma we shall consider two results. The rst is a generalization of
Borsuks theorem.
Theorem 3.9 Let f: X X

be a bounded map of type (M). If there exists a ball B


R
, R > 0 such
that f is odd on the boundary B
R
, i.e.,
f(x) = f(x) for all |x| = R,
then the equation f(x) = 0 has at least one solution x
0
X with |x
0
| R.
Proof: Clearly also f

is also odd on the boundary B


R
. Assume rst that there exists

> 0
0 / f

(B
R
) for all 0 < <

. Then d
LS
(f

, B
R
, 0) is well-dened and odd by Proposition 3.1.
Since d
LS
(f

, B
R
, 0) ,= 0, we can, for each with 0 < <

, nd a point x

B
R
such that
f

(x

) = 0. In view of Lemma 3.4 there exists a point x


0
X and such sequences that

j
0
+
, x

j
x
0
X and f(x

j
) f(x
0
) = 0,
and since any closed and convex set is weakly closed, we see that x
0
B
R
.
On the other hand, if there is no such

> 0 that 0 / f

(B
R
) for all 0 < <

, then we can
nd sequences
j
,
j
0
+
, and x
j
B
R
such that f

j
(x
j
) = 0. Hence the assertion again
follows from Lemma 3.4.
In a similar fashion it is easy to prove the following result (cf. Section 3.1).
Theorem 3.10 Let f: X X

be a bounded map of type (M). If there exists R > 0 such that


f(x), x) 0 for all |x| = R,
then the equation f(x) = 0 has at least one solution x
0
X with |x
0
| R.
29
4 Further Applications
4.1 Multiplication Theorem
In this section we generalize the multiplication theorem of the Leray-Schauder theory concern-
ing the computation of the degree of a composite function. It can be stated as follows (see [24] p. 93).
Proposition 4.1 (Multiplication Theorem) Suppose that G is an open bounded set of a real
Banach space X, f: G X a map of the Leray-Schauder type, that is, I f is compact, and M
an open bounded set X containing f(G). Let = M f(G) and let
j
, (j = 1, 2, 3, . . .), be the
open components of . If h: M X is a Leray-Schauder type map and y is a point of X such that
y / (h f)(G) h(M), then
d
LS
(h f, G, y) =

j
d
LS
(h,
j
, y) d
LS
(f, G,
j
). (34)
Remark 4.1 The summation in (34) is nite since h
1
(y) is a compact set, so that, being covered
by the disjoint open sets
j
, it meets only a nite number of them and the other terms in (34)
vanish.
Clearly the composition of two maps of X into X

is not dened. However, we generalize


Proposition for the case h f, where h is of class (S)
+
and f is a map of Leray-Schauder type. Let
G be an open bounded set of X and f = I g: G X such that g compact. Suppose that M is
an open bounded set of X containing f(G) and h F
M,B
. According to Lemma 2.3, h f F
G,B
.
In view of Section 2.1 we get the following maps of the Leray-Schauder type ( > 0):
h

f = f +
1


(h f): G X (35)
and
(h f)

= I +
1


(h f): G X. (36)
Let > 0 be xed and consider an ane Leray-Schauder homotopy
H

(t, x) = t(h

f)(x) + (1 t)(h f)

(x), x G, 0 t 1,
which by (35) and (36) can be written in the form
H

(t, x) = (h f)

(x) tg(x), x G, 0 t 1. (37)


The next lemma plays an essential role in this section.
Lemma 4.1 Assume that (h f)(x) ,= 0 for all x G. Then there exists

> 0 such that


H

(t, x) ,= 0 for all x G, 0 t 1 and 0 < <

. (38)
Proof: If this is not true, there exist sequences
j
, t
j
[0, 1] and x
j
G with 0
+
such that H

j
(t
j
, x
j
) = 0 or, equivalently, by (36) and (37)
x
j
+
1


(h f)(x
j
) t
j
g(x
j
) = 0 for all j. (39)
30
Because g is compact and the other sequences are bounded, we can assume without loss of generality
that
x
j
x X, t
j
t [0, 1], g(x
j
) u X and (h f)(x
j
) w X

. (40)
Consequently,

(h f)(x
j
)

(w). On the other hand, by (39)


(h f)(x
j
) =
j
x
j
+
j
t
j
g(x
j
) 0.
By properties of

, w = 0 and hence
(h f)(x
j
) 0. (41)
Using (39) we get
(h f)(x
j
), x
j
) =
_
(h f)(x
j
),
1


(h f)(x
j
) +t
j
g(x
j
)
_
=
1

j
|

(h f)(x
j
)|
2
W
+t
j
(h f)(x
j
), g(x
j
))
and further by (40) and (41)
limsup
j
(h f)(x
j
), x
j
x) = limsup
j
_

j
|

(h f)(x
j
)|
2
W
_
0,
which implies, since h f is of class (S)
+
, that x
j
x G. Using (41) and demicontinuity of
h f we get
(h f)(x
j
) (h f)(x) = 0 with x G,
which contradicts our assumption 0 / (h f)(G). Thus the proof is complete.
Now we are in a position to prove the main result of this section.
Theorem 4.1 (Generalized Multiplication Theorem) Let G be an open bounded set of X and
f = I g: g X a map of Leray-Schauder type. Let M be another open, bounded set such that
f(G) M. Assume h F
M,B
and y is a point of X

such that
y / (h f)(G) h(M). (42)
If we denote = M f(G) =
j

j
, where the open sets
j
(j = 1, 2, 3, . . .) are the components
of , then
d
LS
(h f, G, y) =

j
d
LS
(h,
j
, y) d
LS
(f, G,
j
). (43)
Proof: First we take y = 0. According to Lemma 4.1 there exists

> 0 such that


0 / [t(h

f) + (1 t)((h f)

] (G)
for all 0 t t and 0 < <

. Thus by Denition 2.1 and by the homotopy invariance for the


Leray-Schauder degree,
d(h f, G, 0) = d
LS
((h f)

, G, 0) = d
LS
(h

f, G, 0). (44)
31
for all 0 < <

. Moreover, choosing

> 0 suciently small we have 0 / h

(M) for all 0 < <

,
and, consequently,
0 / (h

f)(G) h

(M) for all 0 < <

.
Hence it follows from Proposition that
d
LS
(h

f, G, 0) =

j
d
LS
(h

,
j
, 0) d
LS
(f, G,
j
) (45)
for all 0 < <

. The right-hand side of (44) contains only a nite number of nonvanishing terms
(cf. Remark 4.1), and for each component
j
we have
d(h,
j
, 0) = d
LS
(h

,
j
, 0) for all 0 < <

. (46)
Using (44), (45) and (46) we conclude that
d(h f, G, 0) =

j
d(h,
j
, 0) d
LS
(f, G,
j
), (47)
which is exactly the relation (43) if the case y = 0.
Let now y X

be any nonzero element satisfying (42). If we denote



h() = h() y, we nd
that

h F
M,B
and 0 / (

h f)(G)

h(M). As stated above,


d(

h f, G, 0) =

j
d(

h,
j
, 0) d
LS
(f, G,
j
),
which further means
d(h f, G, y) =

j
d(h,
j
, y) d
LS
(f, G,
j
),
and the proof is complete.
As an application of Theorem 4.1 we study a simple example. Let f = I g: G X be a map
of Leray-Schauder type and z
0
X f(G). By choosing h = J, y = J(z
0
) and an open bounded
set M of X such that f(G) M and z
0
M we get
d(J f, G, y) =

j
d(J,
j
, y) d
LS
(f, G,
j
).
Since z
0
M f(G), there exists a unique
k
such that z
0

k
. Thus d(J,
j
, y) = 0 for all
j ,= k and
d(J f, G, y) = d(J,
k
, y) d
LS
(f, G,
k
) = +1 d
LS
(f, G,
k
) = d
LS
(f, G, z
0
). (48)
This shows that in fact the Leray-Schauder degree theory can be considered a part of the degree
theory for the maps of class (S)
+
. Indeed, by (48) the value of d
LS
(f, G, z
0
) is always reached by
means of d (cf. Lemma 2.4).
32
4.2 On Solution Set Structure
We shall now consider the structure of the solution set of the equation f(x) = 0. We generalize
results obtained in [20], where the authors investigate zero-epi maps, which provide an interesting
alternative to degree theory (see also [14], [15], and [16]).
Theorem 4.2 Let f F
G
be such that 0 / f(G) and d(f, G, 0) ,= 0. Suppose that for any > 0
and z f
1
(0) there exists a demicontinuous quasimonotone map h

: G X

satisfying the
following conditions:
(i) h

(z) = 0;
(ii) |h

(x)| < for all x G;


(iii) the set S

= x G : f(x) + h

(x) = 0 is -chained, i.e., for any two points a and b in S

there exists some nite sequence of points x


1
, x
2
, . . . , x
n
S

such that x
1
= a, x
n
= b and
dist(x
i
, x
i+1
) < 2, i = 1, 2, . . . , n 1.
Then the set f
1
(0) is a continuum, i.e., it is compact and connected.
Proof: Since d(f, G, 0) ,= 0, we have f
1
(0) ,= and, moreover, f
1
(0) is compact (see Lemma 2.5).
Suppose that f
1
(0) is not connected. Then f
1
(0) = K
1
K
2
, where K
1
and K
2
are compact
nonempty subsets of G with K
1
K
2
= . Denote dist(K
1
, K
2
) = 6
1
> 0 and dene two open sets
by setting
G
j
= x G : dist(K
j
, x) < 2
1
, j = 1, 2.
Then K
j
G
j
G, j = 1, 2, and clearly dist(G
1
, G
2
) 2
1
. By closedness of f(G (G
1
G
2
)) we
get
inf|f(x)| : x G (G
1
G
2
) =
2
> 0.
Let 0 <

< min(
1
,
2
). Then
_
dist(G
1
, G
2
) > 2

and
inf|f(x)| : x G (G
1
G
2
) >

.
(49)
The additivity property of d implies that
d(f, G, 0) = d(f, G
1
, 0) +d(f, G
2
, 0),
where we may suppose that, for instance, d(f, G
1
, 0) ,= 0. For any xed z K
2
and 0 < <

there exists a demicontinuous quasimonotone map h

: G X

satisfying (i), (ii) and (iii). By (i),


f(z) +h

(z) = 0, and thus z S

, which implies that G


2
S

,= . Also, by (ii), we get


|f(x)| = |h

(x)| < <

for all x S

.
Hence it follows from (49) that S

G
1
G
2
, and since S

is -chained, necessarily
S

G
2
. (50)
On the other hand, f +h

F
G
, and we consider the homotopy given by
f
t
= t(f +h

) + (1 t)f = f +th

, 0 t 1.
33
If f
t
(x) = 0 for some t [0, 1] and x G, then
|f(x)| = t|h

(x)| < ,
which means that x G
1
G
2
. Hence 0 / f
t
(G) for any t in [0, 1], and by homotopy invariance
property
d(f +h

, G
1
, 0) = d(f, G
1
, 0) ,= 0.
Hence 0 (f +h

)(G
1
), i.e., S

G
1
,= . But this contradicts (50), and the proof is complete.
Remark 4.2 The condition (iii) of Theorem 4.2 is satised if, for instance, the equation f(x) +
h

(x) = 0 has at most one solution.


Remark 4.3 Theorem 4.2 implies in particular that the solution set f
1
(0) contains either one
point or is innite.
Example 4.1 Let f: G X

be a demicontinuous monotone map of class (S)


+
with 0 / f(G).
Suppose that there exists x G such that
|f(x)| > |f( x)| for all x G.
Then, for all x G, we have
f(x), x x) = f(x) f( x), x x) +f( x), x x)
f( x), x x)
|f( x)||x x|
> |f(x)||x x|.
Then by Lemma 3.1, d(f, G, 0) = 1, and therefore f
1
(0) ,= . For any > 0 and z f
1
(0), dene
h

(x) = J(x z) for all x G,


where 0 < < (2 sup|x| : x G)
1
. Clearly
(i) h

(z) = 0;
(ii) |h

(x)| = |x z| < for all x G;


(iii) since f +h

is strictly monotone, S

contains exactly one point, z, and therefore S

is -chained.
From Theorem 4.2 we can now deduce that f
1
(0) is a continuum.
34
4.3 On the Existence of Multiple Solutions
In this section we shall outline one possible way to deal with the question of multiplicity, i.e.,
the number of solutions of the equation f(x) = 0. Our approach is based on the idea of seeking
distinct solutions in some maximal closed linear subspaces.
We shall rst present some background material.
Lemma 4.2 Let f F
G,B
be such that 0 / f(G) and d(f, G, 0) ,= 0. Then
_
y X

: y =
f(x)
|f(x)|
, x G
_
= y X

: |y| = 1.
Proof: Let y X

, |y| = 1, be given. The value of d(f, G, ) remains constant on each open


component of X

f(G). Assume 0
k
, where
k
is an open component of X

f(G).
For all suciently small values of > 0, y
k
. If, however, y
k
for all > 0, then
d(f, G, y) = d(f, G, 0) ,= 0 and y f(G) for all > 0. This is impossible since f is bounded.
Hence there exists
0
> 0 such that
0
y f(G), and the assertion follows.
For any closed linear subspace V of X, let j
V
: V X be the natural embedding for which
j
V
(x) = x for all x V . Consequently, let j

V
: X

stand for the projection which is dened


by
j

V
(w), v) = w, j
V
(v)) for all w X

and v V.
Assume G is an open bounded set of X such that V G ,= , which is true if for instance 0 G.
Clearly G V GV and
V
(GV ) GV , where
V
(GV ) denotes the boundary of GV
in V . It is not hard to see that the map
f
V
= j

V
f j
V
: G V V

is demicontinuous and of class (S)


+
. Let us denote the corresponding unique degree function by d
V
.
Before starting the main result of this section, we recall some facts about maximal closed linear
subspaces (see e.g. [25]). A closed linear subspace V X, V ,= X, is said to be maximal, for any
closed linear subspace Z X with V Z X, either Z = V or Z = X. Any maximal closed
linear subspace can be presented in the form
V = ker(w) = x X : w, x) = 0 for all w X

, w ,= 0.
Let V
1
and V
2
be maximal closed linear subspace given by V
j
= ker(w
j
), w
j
,= 0, j = 1, 2. Then
V
1
= V
2
if and only if w
1
and w
2
are linearly dependent. By the reexivity of X, any maximal
closed linear subspace N of X

can be given by
N = y X

: y, x) = 0 for all x X, x ,= 0.
We are now in a position to prove
Theorem 4.3 Let G be an open bounded subset of X with 0 G and f F
G,B
such that f(0) ,= 0.
Suppose that
(i) f(x), x) > 0 for all x G and
35
(ii) for each linearly independent pair x
1
, x
2
G f
1
(0), f(x
1
) and f(x
2
) are linearly inde-
pendent.
Then the solution set f
1
(0) is innite.
Proof: By (i), 0 / f(G), and it follows from Lemma 3.1 that d(f, G, 0) ,= 0. Hence there exits at
least one x
1
such that
f(x
1
) = 0, x
1
G
1
, x
1
,= 0.
Suppose next that we have obtained a set x
1
, x
2
, . . . , x
n
G, n 1, such that
f(x
j
) = 0, j = 1, . . . , n and x
j
,= x
i
for all j ,= i.
Certainly x
j
,= 0 for all j. Let N
0
be some maximal closed linear subspace of X

containing f(0)
and denote
N
i
= y X

: y, x
i
) = 0, i = 1, 2, . . . , n.
We can obviously nd
y X

, | y| = 1, such that y / N
0
N
1
N
n
.
Hence y ,= f(0) for all R and y, x
i
) , = 0, i = 1, 2, . . . , n. Denote V = ker( y). In view
of Lemma 4.2, we can write V = ker( y) = ker(f( u)) for some u G. Clearly u / V since
f( u), u) , = 0 by (i). As stated above, the map f
V
= j

V
f j
V
: G V V

is demicontinuous
and of class (S)
+
. Moreover, by (i)
f
V
(v), v) > 0 for all v
V
(G V ).
Thus d
V
(f
V
, G V, 0) ,= 0 by Lemma 3.1, and we get a solution x
n+1
G V of the equation
f
V
(v) = 0 on V , i.e.,
f(x
n+1
), v) = 0 for all v V. (51)
Since f( u), x
i
) , = 0, i = 1, 2, . . . , n, and f( u), x
n+1
) = 0, we have x
n+1
,= x
i
, i = 1, 2, . . . , n. If
f(x
n+1
) = 0 in X

, the proof is complete. If not, then by (51) and the denition of V ,


V = ker(f( u)) = ker(f(x
n+1
),
which implies that f(x
n+1
) = f( u) for some R, ,= 0. By virtue of (ii), x
n+1
and u are
linearly dependent, and since u / V , x
n+1
V , we necessarily have x
n+1
= 0. But this contradicts
the fact that f( u) ,= f(0) for all R. Hence f(x
n+1
) = 0 in X

, and we have obtained n + 1


distinct solutions x
1
, x
2
, . . . , x
n+1
of the equation f(x) = 0. By the induction argument the proof
is complete.
Remark 4.4 Since f
1
(0) is compact and innite in the previous theorem, it contains at least one
accumulation point.
36
References
[1] Amann, H., and S. Weiss: On the uniqueness of the topological degree, Math. Z. 130 (1973),
39-54.
[2] Berkovits, J. and V. Mustonen: On the topological degree for mappings of monotone type,
Nonlinear Anal. (to appear)
[3] Brouwer, L.E.J.:

Uber Abbildung von Mannigatigkeiten, Math. Anal. 71 (1912) 97- 115.
[4] Browder, F.E.: Nonlinear operators and nonlinear equations of evolution in Banach spaces,
Nonlinear functional analysis, Proc. Sympos. Pure Math. 18:2, Chicago, III, 1968, Amer. Math.
Soc., Providence, Rhode Island, (1976), 1-308.
[5] Browder, F.E.: Degree of mapping for nonlinear mappings of monotone type, Proc. Nat. Acad.
Sci. U.S.A. 80 (1983), 1771- 1773.
[6] Browder, F.E.: Degree of mapping for nonlinear mappings of monotone type. Densely dened
mappings, Proc. Nat. Acad. Sci. U.S.A. 80 (1983), 2405-2407.
[7] Browder, F.E.: Degree of mapping for nonlinear mappings of monotone type. Strongly nonlinear
mappings Proc. Nat. Acad. Sci. U.S.A. 80 (1983), 2408- 2409.
[8] Browder, F.E.: The degree of mapping, and its generalizations, Contemporary Mathematics
21, Topological Methods in nonlinear functional analysis, edited by S.P. Singh, S. Thomeier,
and B. Watson, American Mathematical Society, providence, Rhode Island, 1983, 15-40.
[9] Browder, F.E.: Fixed point theory and nonlinear problems, Bull. Amer. Soc. (N.S.) 9 (1983),
1- 39.
[10] Browder, F.E., and B.A.Ton: Nonlinear functional equations in Banach spaces and elliptic
super-regularization, Math. Z. 105 (1968), 177-195.
[11] Cronin, J.: Fixed points and topological degree in nonlinear analysis, Mathematical Surveys 11,
American Mathematical Society, providence, Rhode Island, (1964).
[12] Day, M.M.: Normed linear spaces, Springer-Verlag, New York (1973).
[13] F uhrer, L.: Ein elementarer analytischer Beweis zur Eindeutigkeit des Abbildungsgrades im R
n
,
Math. Nachr. 54 (1972), 259-267.
[14] Furi, M., M. Martelli, and A. Vignoli: On the solvability of nonlinear operator equations in
normed spaces, Ann. Mat. Pure Appl. (4) 124 (1980), 321-343.
[15] Furi, M. and M.P. Pera: An elementary approach to boundary value problems at resonance,
Nonlinear Anal. 4 (1980), 1081-1089.
[16] Krasnoselski

i, M.A., and A.I. Perov: Existence of solutions for certain non-linear operator
equations, Dokl. Akad. Nauk. SSSR 126 (1959), 15-18.
[17] Landes, R., and V. Mustonen: On pseudo-monotone operators and nonlinear noncoercive vari-
ational problems on unbounded domains, Math. Ann. 248 (1980), 241-246.
37
[18] Leray, J., and J. Schauder: Topologie et equations fonctionelles, Ann. Sci.

Ecole. Norm. Sup. 51
(1934), 45-78.
[19] Lloyd, N.G.: Degree theory, Cambridge University Press, Cambridge, (1978).
[20] Martelli, M., and A. Vignoli: On the structure of the solution set of nonlinear equations,
Nonlinear Anal. 7 (1983), 685-693.
[21] Morel, J. -M, and H. Steinlein: On a problem of Nirenberg concerning expanding maps, J.
Funct. Anal. 59 (1984), 145-150.
[22] Nirenberg, L.: Topics in nonlinear functional analysis, Lecture Notes, Courant Institute of
Mathematical Sciences, New York University, New York, (1974).
[23] Pascali, D., and S. Sburlan: Nonlinear mappings of monotone type, Editura Academiei,
Burkarest, (1978).
[24] Schwartz, J.T.: Nonlinear functional analysis, Gordon and Breach, New York, (1969).
[25] Taylor, A. E., and D.C. Lay: Introduction to functional analysis, John Wiley & Sons, New
York, (1980).
[26] Trojanski, S. L.: On locally uniformly convex and dierentiable norms in certain non-separable
Banach spaces, Studia Math. 37 (1970/71), 173-180.
[27] Wille, F.: Monotone operatoren mit Storungen, Arch. Rational Mech. Anal. 46 (1972), 369-388.
[28] Yosida, K.: Functional analysis, Die Grundlehren der mathematischen Wissenschaften 123,
Springer-Verlag, Berlin-Gottingen - Heidelberg, (1965).
[29] Zeilder, E.: Vorlesungen uber nichtlineare Funktionalananysis II. Monotone Operatoren,
Teubner-Texte zur mathematik, B.G. Teubner Verlagsgesellaschaft, leibzig, (1977).
Juha Berkovits
University of Oulu
Department of Mathematics
SF-90570 Oulu
Finland
38

Das könnte Ihnen auch gefallen