Topological Degree Theory and Applications
Topological Degree Theory and Applications
VOLUME 10
TOPOLOGICAL
DEGREE THEORY
AND APPLICATIONS
Volume 1
Method of Variation of Parameters for Dynamic Systems
V. Lakshmikantham and S.G. Deo
Volume 2
Integral and Integrodifferential Equations: Theory, Methods and Applications
Edited by Ravi P. Agarwal and Donal O’Regan
Volume 3
Theorems of Leray-Schauder Type and Applications
Donal O’Regan and Radu Precup
Volume 4
Set Valued Mappings with Applications in Nonlinear Analysis
Edited by Ravi P. Agarwal and Donal O’Regan
Volume 5
Oscillation Theory for Second Order Dynamic Equations
Ravi P. Agarwal, Said R. Grace, and Donal O’Regan
Volume 6
Theory of Fuzzy Differential Equations and Inclusions
V. Lakshmikantham and Ram N. Mohapatra
Volume 7
Monotone Flows and Rapid Convergence for Nonlinear Partial Differential Equations
V. Lakshmikantham, S. Koksal, and Raymond Bonnett
Volume 8
Nonsmooth Critical Point Theory and Nonlinear Boundary Value Problems
Leszek Gasiński and Nikolaos S. Papageorgiou
Volume 9
Nonlinear Analysis
Leszek Gasiński and Nikolaos S. Papageorgiou
Volume 10
Topological Degree Theory and Applications
Donal O’Regan, Yeol Je Cho, and Yu-Qing Chen
VOLUME 10
TOPOLOGICAL
DEGREE THEORY
AND APPLICATIONS
Donal O’Regan
Yeol Je Cho
Yu-Qing Chen
Published in 2006 by
Chapman & Hall/CRC
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
8 REFERENCES 195
where zi are the zeros of f in the region enclosed by Γ and αi are the corre-
sponding multiplicites. If we assume in addition that Γ has positive orientation
and no intersection points, then we know from Jordan’s Theorem, which will
be proved later in this chapter, that w(Γ, zi ) = 1 for all zi . Thus (2) becomes
X
w(f (Γ), 0) = αi . (3)
i
So we may say that f has at least |w(f (Γ), 0)| zeros in G. The winding number
is a very old concept which goes back to Cauchy and Gauss. Kronecker,
Hadamard, Poincare, and others extended formula (1). In 1912, Brouwer [32]
introduced the so-called Brouwer degree in Rn (see Browder [35], Sieberg [277]
for historical developments). In this chapter, we introduce the Brouwer degree
theory and its generalization to functions in V M O. This chapter is organized
as follows:
In Section 1.1 we introduce the notion of a critical point for a differentiable
function f . We then prove Sard’s Lemma, which states that the set of critical
points of a C 1 function is “small”. Our final result in this section shows how
a continuous function can be approximated by a C ∞ function.
In Section 1.2 we begin by defining the degree of a C 1 function using the
Jacobian. Also we present an integral representation which we use to define
the degree of a continuous function. Also in this section we present some
properties of our degree (see theorems 1.2.6, 1.2.12, and 1.2.13) and some
useful consequences. For example, we prove Brouwer’s and Borsuk’s fixed
point theorem, Jordan’s separation theorem and an open mapping theorem. In
addition we discuss the relation between the winding number and the degree.
In Section 1.3 we discuss some properties of the average value function and
then we introduce the degree for functions in V M O.
In Section 1.4 we use the degree theory in Section 1.2 to present some exis-
tence results for the periodic and anti-periodic first order ordinary differential
equations.
|f 0 (x) − f 0 (y)| ≤
√
nl
for all x, y ∈ Q with |x − y| ≤ m . Therefore, we have
Z 1
0
|f (x) − f (y) − f (y)(x − y)| ≤ |f 0 (y + t(x − y)) − f 0 (y)||x − y|dt
0
≤ |x − y|
√
for all x, y ∈ Q with |x − y| ≤ mnl . We decompose Q into r cubes, Qi , of
√
diameter mnl , i = 1, 2, · · · , r. Since ml is the lateral length of Qi , we have
r = mn . Now, suppose that Qi ∩Sf (Ω) 6= ∅. Choosing y ∈ Qi ∩Sf (Ω), we have √
f (y+x)−f (y) = f 0 (y)x+R(y, x) for all x ∈ Qi −y, where |R(y, x+y)| ≤ mnl .
Therefore, we have
such that f˜(x) = f (x) for all x ∈ K, where convf (K) is the convex hull of
f (K).
Proof. Since K is bounded closed subset, there exists at most countable
{ki : i = 1, 2, · · · } ⊂ K such that {ki : i = 1, 2, · · · } = K. Put
|x − ki |
d(x, K) = inf |x − y|, αi (x) = max{2 − , 0}
y∈K d(x, A)
for any x ∈
/ K and
P
(
f (x), x ∈ K,
f˜(x) = P2
i≥1
−i
i≥1 2
αi (x)f (ki )
−i α (x)
i
, x∈
/ K.
Z
fλ (x) = f˜(y)φλ (y − x)dx for all x ∈ Rn , λ > 0.
Rn
f −1 (p) = {x1 , x2 , · · · , xn }.
But we have
Jf (x) = Jf −p (x),
Z Z
φ (f (x) − p)|Jf (x)|dx = φ (x)dx = 1,
Ui Br0
deg(f, Ω, p) = deg(f, Ω, p0 ),
deg(f, Ω, p1 ) = deg(f, Ω, p2 ).
Notice that
φ (x − p2 ) − φ (x − p1 ) = divw(x),
where Z 1
w(x) = (p1 − p2 ) φ (x − p1 + t(p1 − p2 ))dt.
0
We show that there exists a function v ∈ C 1 (RN ) such that supp(v) ⊂ Ω and
Lemma 1.2.4. Let Ω ⊂ RN be open, f ∈ C 2 (Ω) and let dij be the cofactor
∂f
of ∂xji in Jf (x) and
PN
j=1 wj (f (x))dij (x) x ∈ Ω,
vi (x) =
0, otherwise.
Then (v1 (x), v1 (x), · · · , vN (x)) satisfies divv(x) = divw(f (x))Jf (x).
∂
where ∂k = ∂xk . Now, we claim that
X
N
∂i dij (x) = 0 for j = 1, 2, · · · , N.
i=1
Then we have
N
X
∂i dij (x) = (−1)i+j det(fx1 , · · · , fxi−1 , fxi+1 , · · · , ∂i fxk , · · · , fxN ).
k=1
Set
where δki = 1 for k < i, δii = 0 and δki = −δik for i, k = 1, 2, · · · , N . Hence
we have
N
X N
X N
X
(−1)j ∂i dij (x) = (−1)k−1+i γki aki = (−1)i−1+k γik aik
i=1 i,k=1 k,i=1
N
X
=− (−1)k−1+i γki aki = 0.
i,k=1
Now, we have
N
X N
X
∂i vi (x) = di,j ∂k wj (f (x))∂i fk (x) + wj (f (x))∂i dij (x).
j,k=1 j=1
PN
On the other hand, i=1 dij ∂i fk (x) = δjk Jf (x) with Kronecker’s δjk .
Therefore, it follows that
N
X
divv(x) = ∂k wj (f (x))δjk Jf (x) = divw(f (x))Jf (x).
k,j=1
Hence f has a fixed point in B(0, R). This completes the proof.
From Theorem 1.2.7, we have the well-known Brouwer fixed point theorem:
Theorem 1.2.8. Let C ⊂ Rn be a nonempty bounded closed convex subset
and f : C → C be a continuous mapping. Then f has a fixed point in C.
Proof. Take B(0, R) such that C ⊂ B(0, R) and let r : B(0, R) → C be a
retraction. By Theorem 1.2.7, there exists x0 ∈ B(0, R) such that f rx0 = x0 .
Therefore, x0 ∈ C, and so we have rx0 = x0 . This completes the proof.
Theorem 1.2.9. Let f : Rn → Rn be a continuous mapping and 0 ∈ Ω ⊂
n
R with Ω an open bounded subset. If (f (x), x) > 0 for all x ∈ ∂Ω, then
deg(f, Ω, 0) = 1.
Proof. Put H(t, x) = tx + (1 − t)f (x) for all (t, x) ∈ [0, 1] × Ω. Then
0∈
/ H([0, 1] × ∂Ω), and so we have
deg(f, Ω, 0) = deg(I, Ω, 0) = 1.
(f (x), x)
lim = +∞,
|x|→∞ |x|
then f (Rn ) = Rn .
Proof. For any p ∈ Rn , it is easy to see that there exists R > 0 such that
(f (x) − p, x) > 0 for all x ∈ ∂B(0, R), where B(0, R) is the open ball centered
at zero with radius R. By Theorem 1.2.9, we have
and so f (x) − p = 0 has a solution in B(0, R). This completes the proof.
and hence Jgk+1 (x) 6= 0. By induction, we also have gn0 (0) = g10 (0) = f 0 (0)
and so 0 is a regular value for gn . By Definition 1.2.5 and Definition 1.2.1, we
know that
X
deg(f, Ω, 0) = deg(gn , Ω, 0) = sgnJgn (0) + sgnJgn (x)
x∈g −1 (0),x6=0
If p ∈
/ (gf )(∂Ω), then
X
deg(gf, Ω, p) = deg(f, Ω, Ui )deg(g, Ui , p), (1.2.1)
i
and note
X
sgnJg (z)sgnJf (x)
x∈f −1 (z),z∈g −1 (p)
X X
= sgnJg (z)[ sgnJf (x)]
z∈g −1 (p),z∈f (Ω) x∈f −1 (z)
X
= sgnJg (z)deg(f, Ω, z)
z∈f (Ω),g(z)=p
t X
X
= sgnJg (z)deg(f, Ω, z)
i=1 z∈Ui
X
= deg(f, Ω, Ui )deg(g, Ui , p).
i
Nm = {i ∈ N : deg(f, Ω, Ui ) = m}.
Obviously, Vm = ∪i∈Nm Ui and thus we have
X X X X
deg(f, Ω, Ui )deg(g, Ui , p) = [ deg(g, Ui , p)] = deg(g, Vm , p).
i m i∈Nm m
and put
Vm0 = {z ∈ Br+1 (0) \ f0 (∂Ω) : deg(f0 , Ω, z) = m}.
Then we have Vm ∩ M0 = Vm0 ∩ M0 and
deg(g0 , Vm , p) = deg(g0 , Vm ∩ Vm0 , p) = deg(g0 , Vm0 , p).
Therefore, we have
X X
deg(g0 f0 , Ω, p) = mdeg(g0 , Vm0 , p) = mdeg(g, Vm , p).
m m
deg(h0 , Ui , Wk ) = deg(h0 , Ui , Vj )
for all j ∈ Nk . Therefore, we have
X X
1= deg(h0 , Ui , Vj )deg(h−1 , Vj , p)
k j∈Nk
X (1.2.3)
= deg(h0 , Ui , Vj )deg(h−1 , Vj , Ui ),
j
since p ∈ Ui ⊂ Rn \ (h−1 (Ω)) ⊂ Rn \ h−1 (∂Vj ). For any fixed j, the same
argument implies that
X
1= deg(h0 , Ui , Vj )deg(h0 , Ui , Vj )deg(h−1 , Vj , Ui ). (1.2.4)
i
From (1.2.3) and (1.2.4), it follows that Rn \ Ω1 and Rn \ Ω2 have the same
number of connected components. This completes the proof.
Theorem 1.2.15. Let φ : Rn → R1 be continuously differentiable, gradφ(x) 6=
0 for |x| sufficiently large and lim|x|→∞ φ(x) = +∞. Then
Z
φλ (x) = φ(y)mλ (y − x)dx for all x ∈ Rn , λ > 0.
Rn
can be extended to a unique solution on [0, ∞). We may also assume that
φ(x) ≥ 0 by adding a constant. Take r0 > 0 such that gradφ(x) 6= 0 for
|x| ≥ r0 , put M = maxx∈Br (0) φ(x). Choose r1 > r0 such that φ(x) ≥ M + 1
0
for |x| ≥ r1 and set
M1 = max φ(x).
x∈∂Br1 (0)
Rt
Again, by φ(u(t, x)) = φ(x) − 0 |gradφ(u(s, x))|2 ds, we know that if x ∈
∂Br1 (0), then φ(u(t, x)) ≤ M1 . Let β = min{|gradφ(x)| : |x| ≥ r0 and
φ(x) ≤ M1 }. Then, if |x| = r1 , we have
It is easy to check that h is continuous and h(t, x) 6= 0 for all x ∈ ∂Br1 (0).
Thus we have deg(gradφ, Br1 (0), 0) = deg(I − P, Br1 (0), 0) = 1. This com-
pletes the proof.
Theorem 1.2.16. Let Ω ⊂ Rn be an open subset and f : Ω → Rn be
continuous and locally one to one. Then f is an open mapping.
Proof. For each x0 ∈ Ω, we prove that there exists r > 0 such that
f (Br (x0 )) contains a ball with center at f (x0 ). Without loss of generality,
we may assume that x0 = 0. Otherwise, put Ω1 = Ω − {x0 } and f1 (x) =
f (x + x0 ) − f (x0 ).
1
Choose r > 0 such that f is one to one on Br (0). Set h(t, x) = f ( 1+t x) −
t
f (− 1+t x) for all (t, x) ∈ [0, 1] × Br (0). Then h is continuous and h(t, x) 6= 0
for all (t, x) ∈ [0, 1] × ∂Br (0). Otherwise, h(t, x) = 0 for some (t, x) ∈ [0, 1] ×
1 t
∂Br (0) and so then 1+t x = − 1+t x since f is one to one and x = 0, which is
a contradiction. Therefore, we have
since h(1, ·) is odd. Choose t > 0 such that t < inf{|f (x)| : x ∈ ∂Br (0)}.
Then
deg(f, Br (0), y) = deg(f, Br (0), 0).
Now, we have Bt ⊂ f (Br (0)) and thus f is open. This completes the proof.
Take > 0 small enough such that the Vi ’s are disjoint, where Vi = B(zi , ),
sgnJf (z) = sgnJf (zi ) for z ∈ Vi , Vi ⊂ B(0, 1) and the restriction of f to Vi is
a homeomorphism for i = 1, 2, · · · , k. Put Si = ∂Vi . Then f (Si ) is a Jordan
curve such that a lies in its interior region, f (Si ) has the same orientation as
Si if Jf (zi ) > 0 and the opposite orientation if Jf (zi ) < 0.
Now, set U = B(0, 1) \ ∪ki=1 Vi . Then |f (z) − a| > α in U for some α > 0.
We can divide U into small rectangles Rj such that |f (z) − f (w)| < α on
each Rj . Since the image f (∂(Rj ∩ V )) does not wind around a, we have
w(f (∂(Rj ∩ V )), a) = 0, and summing over all Rj yields
Z k Z
1 X 1
dz = dz.
f (Γ) z−a i=1 f (Si ) z−a
Proof. Since c < m(U ), there exists a compact subset K ⊂ U with m(K) >
c, and finitely many of the balls in φ, say, A1 , A2 , · · · , Am , cover K. Let B1 be
the largest of the Ai ’s (that is, choose B1 to have maximal radius), let B2 be
the largest of the Ai ’s which are disjoint from B1 and so on until the sequence
of Ai ’s is exhausted. If Aj is not one of the Bi ’s, there exists i such that
Aj ∩ Bi 6= ∅ and the radius of Aj is at most that of Bi . Therefore, Aj ⊂ Bi∗ ,
where Bi∗ is the ball concentric with Bi whose radius is three times that of
Bi . But then K ⊂ ∪i Bi∗ , so
X X
c < m(K) ≤ m(Bi∗ ) = 3n m(Bi ).
i i
Theorem 1.3.3. There is a constant β > 0 such that, for all f ∈ L1 (Ω)
and α > 0, Z
β
m({x : Hf (x) > α}) ≤ |f (x)|dx.
α Ω
k k Z
3n X 3n
X Z
c < 3n m(Bi ) ≤ |f (y)|dy ≤ |f (y)|dy.
i=1
α i=1 Bi α Ω
By letting c → m(Eα ), we obtain the desired result. This completes the proof.
Theorem 1.3.4. If f ∈ L1 (Ω), then limr→0 Ar f (x) = f (x) for almost all
x ∈ Ω.
Proof. For any > 0, there exists a continuous function g such that
Z
|f (x) − g(x)|dx < .
Ω
Continuity of g implies that for every x ∈ Ω and δ > 0, there exists r > 0
such that |g(y) − g(x)| < δ whenever |y − x| < r, and hence
Hence, if
Eα = {x : lim sup |Ar (f − g)(x) > α}, Fα = {x : |f (x) − g(x)| > α},
r→0
then we have
α
Eα ⊂ F α2 ∪ {x : H(f − g)(x) > }.
2
R
However, αm(Fα ) ≤ Fα
|f (x) − g(x)|dx < and so, by Theorem 1.3.3, we
have
2 2β
m(Eα ) ≤ + .
α α
By letting → 0, we get m(Eα ) = 0 for all α > 0. This completes the proof.
In the following, let Ω be a smooth open bounded domain in Rn or a smooth
compact Riemannian manifold. (For the definition of Riemannian manifold,
we refer the reader to [171].
Next, we introduce the following concept:
1
Definition 1.3.5. Let
1
R f ∈ 1L (Ω)R and Br ⊂ Ω a (geodesic) ball with radius
r > 0. If supr>0 m(Br ) Br m(Br ) Br |f (x) − f (y)|dxdy < ∞, then f is called
a bounded mean oscillation function. The set of all bounded mean oscillation
functions is denotedR by BM O.
1 1
R
If limr→0 m(B r) Br m(Br ) Br
|f (x) − f (y)|dxdy = 0, then f is called a
vanishing mean oscillation function. The set of all vanishing mean oscillation
functions is denoted by V M O.
Example 1.3.6. If f ∈ L1 (Ω), then f ∈ BM O.
Example 1.3.7. f (x) = | log |x|| ∈ BM O.
In the following, 1 ≤ p < +∞, let W 1,p (Ω) = {u(·) : Ω → R such that u(·) ∈
L (Ω), u0 (·) ∈ Lp (Ω)}.
p
Ãr f
lim deg(A˜r f , B(0, 1), 0) = lim deg( , B(0, 1), 0).
r→0 r→0 |Ãr f |
A˜r f (x)
Proof. Consider the homotopy Hr (t, x) = (1−t)+t|A˜r f (x)|
for all (t, x) ∈
[0, 1] × B(0, 1). By Lemma 1.3.10, we know that Hr (t, x) 6= 0 for all (t, x) ∈
[0, 1] × S n for r sufficiently small and so the conclusion follows from Theorem
1.2.6.
Remark. For more results regarding the degree defined by (1.3.1), we refer
the reader to Brezis and Nirenberg [29] (see also [25], [27], [28], [175] for more
results on the computation of the degree for Sobolev maps).
(1) There exists r > 0 such that (f (t, x), x) < 0 for all t ∈ [0, T ] and |x| = r.
has a solution.
Proof. For each x0 ∈ B(0, r), by Peano’s Theorem, the initial value prob-
lem (
x0 (t) = f (t, x(t)), t ∈ (0, t0 ),
(E 1.4.2)
x(0) = x0
has a solution for some t0 > 0. If (E 1.4.2) has two solutions x(·), y(·), then
d
|x(t) − y(t)|2 = 2(x0 (t) − y 0 (t), x(t) − y(t)) ≤ Lx0 |x(t) − y(t)| (1.4.1)
dt
for some t1 ∈ (0, t0 ) and t ∈ (0, t1 ). From (1.4.1), we get
so x(t) = y(t) for all t ∈ (0, t1 ). Therefore, x(t) = y(t) for t ∈ [0, t0 ], so the
solution of (E 1.4.2) is unique.
d
If x(t) = r, then dt |x(t)|2 = 2(x0 (t), x(t)) = (f (t, x(t)), x(t)) < 0. Thus x(t)
must stay in B(0, r) for all t ∈ [0, t0 ], so x(t) can be extended to [0, +∞) and
also x(t) ∈ B(0, r) for t ∈ [0, +∞).
Now, we define a mapping S : B(0, r) → B(0, r) as follows:
has a solution.
Proof. First, if x(t) is a solution of (E 1.4.3), then
d
(|x(t)|2 ) + |x(t)|2 ≤ 2L|x(t)| + 2|f (t)||x(t)|
dt
for some constant L > 0. Thus, if there exists a N > 0 such that |x(0)| ≤ N ,
then |x(T )| ≤ N .
Now, we define a map S : Rn → Rn by Sy = −x(T ), where x(·) is the
unique solution of (E 1.4.4) with x(0) = y. It is obvious that S is continuous,
so, by Brouwer’s fixed point theorem, Theorem 1.2.7, there exists y ∈ Rn such
that Sy = y, i.e., x(T ) = −y. Thus |x(t)| ≤ M for all t ∈ R and, consequently,
it follows that
1
∂(φ(x(t))[G(x(t)) + |x(t)|2 ]) = ∂Gx(t) + x(t).
2
Therefore, x(·) is a solution of (E 1.4.3). This completes the proof.
Corollary 1.4.3. Let A be a n × n symmetric matrix and f : R → Rn
be a continuous function such that f (t + T ) = −f (t) for all t ∈ R. Then the
following equation:
(
x0 (t) = Ax(t) + f (t), t ∈ R,
(E 1.4.5)
x(t + T ) = −x(t), t∈R
has a solution.
Proof. Since A is symmetric, put Gu = 21 (Au, u) for u ∈ Rn and then
A = ∂G. Thus the conclusion follows from Theorem 1.4.2.
1.5 Exercises
1. Let Ω ⊂ Rn be open bounded and 0 ∈ Ω and f : Ω → Rn be continuous
and (f (x), x) ≥ 0. Show 0 ∈ f (Ω).
2. Let Ω ⊂ R2 be open bounded and u(x, y), v(x, y) : Ω → R be contin-
uously differentiable functions with ux = vy and uy = −vx . Assume
that f (x, y) = (u(x, y), v(x, y)) : Ω → R2 has m many zero points in
Ω \ Sf (Ω). Show that deg(f, Ω, 0) = m.
3. Prove the fundamental theorem of algebra by using Brouwer degree.
4. Let B(0, 1) be the unit ball in Rn and f : Rn → Rn be a continuous func-
tion such that f (∂B(0, 1)) = ∂B(0, 1). Show that deg(f m , B(0, 1), 0) =
(deg(f, B(0, 1), 0)m for all positive integer m.
5. For any integer n, show that there exists an open bounded subset Ω ⊂ R
and a continuous function f : Ω → R such that deg(f, Ω, 0) = n.
6. Let Ω ⊂ Rn be open bounded, f, g ∈ C(Ω) and |g(x)| < |f (x)| for all
x ∈ ∂Ω. Show deg(f − g, Ω, 0) = deg(f, Ω, 0).
7. Let B(0, 1) be the unit ball of R2n+1 and f : ∂B(0, 1) → ∂B(0, 1) be
continuous. Show that there exists x0 ∈ ∂B(0, 1) such that f (x0 ) = x0
or f (x0 ) = −x0 .
8. Let Ω ⊂ Rn be open bounded symmetric, 0 ∈ Ω and f : ∂Ω → Rm be a
function with m < n. Show that f (x) = f (−x) for some x ∈ ∂Ω
9. Let Ω ⊂ Rn be open bounded and f ∈ C(Ω). Suppose that there exists
x0 ∈ Ω such that f satisfies the following condition:
f (x) − x0 = t(x − x0 ) for some x ∈ ∂Ω.
Then t ≤ 1. Show that f has a fixed point in Ω.
almost all x ∈ Ω.
P∞ P∞
12. If f (θ) = n=0 ai einθ ∈ L2 (S 1 , S 1 ) and n=0 n|an |2 < +∞, show that
∞
X
deg(f, S 1 ) = n|an |2 .
n=0
has a solution.
25
{xn : n ≥ m} = M ∩ {xn : n ≥ m}
⊂ ∪ki=1 B(yi , ryi ) ∩ {xn : n ≥ m}
= ∅,
lp = {(xi ) : xi ∈ R for i = 1, 2, · · · , Σ∞ p
i=1 |xi | < +∞}.
Banach space.
kx0 k = 1 and
y0 − y1
d(x0 , M ) = inf k − yk
y∈M ky0 − y1 k
= inf ky0 − y1 k−1 ky0 − y1 − ky0 − y1 kyk
y∈M
> δ0−1 δ0
= ,
δ(x0 , h)
F (x0 + h) = F x0 + F 0 (x0 )h + δ(x0 , h), lim = 0,
h→0 khk
(4) If λ ∈
/ σ(T ), then Tλ is a homeomorphism onto X;
(5) E = N (Tλk ) ⊕ R(Tλk ) for all λ 6= 0, k > 0, where dim(N (Tλk )) < +∞,
and R(N (Tλk )) is closed in E.
If this is not true, there exist xn ∈ M with kxn k = 1 such that kSxn k ≤ n−1 .
Since S = (T − λI)i = (−λ)i I + Σij=1 Cij (−λ)i−j T j , one easily sees that
(xn )∞
n=1 has a convergence subsequence (xnk ) with xnk → x0 . Thus Sx0 = 0,
Thus we have
and the degree defined in Definition 2.2.3 is well defined. For the general case,
if p ∈
/ (I − T )(∂Ω), we define deg(I − T, Ω, p) = deg(I − T − p, Ω, 0).
We recall some properties of the Leray Schauder degree as follows:
Theorem 2.2.4. The Leray Schauder degree has the following properties:
(1) (N ormality) deg(I, Ω, 0) = 1 if and only if 0 ∈ Ω;
(2) (Solvability) If deg(I − T, Ω, 0) 6= 0, then T x = x has a solution in Ω,
(3) (Homotopy) Let Tt : [0, 1]×Ω → E be continuous compact and Tt x 6= x
for all (t, x) ∈ [0, 1] × ∂Ω. Then deg(I − Tt , Ω, 0) doesn’t depend on
t ∈ [0, 1];
(4) (Additivity) Let Ω1 , Ω2 be two disjoint open subsets of Ω and 0 ∈
/
(I − T )(Ω − Ω1 ∪ Ω2 ). Then
Proof. The proof follows from the corresponding properties of the Brouwer
degree.
The following is the well-known Schauder fixed point theorem:
Theorem 2.2.5. Let C ⊂ E be a nonempty bounded closed convex subset
and T : C → C be a continuous compact mapping. Then T has a fixed point
in C.
Proof. The proof is the same as the proof of Brouwer’s fixed point theorem.
If we only require the continuous condition on the mapping T , then the
conclusion of Theorem 2.2.5 fails as the following example shows:
Example 2.2.6. Let T : l2 → l2 be a mapping defined by
T (x1 , x2 , · · · ) = (1 − kxk, x1 , x2 , · · · )
and
Σm mdeg(I − S, Vm , p) = Σm mdeg(I − S1 , Vm , p)
= Σm mdeg(I − S1 , Vm ∩ F, p).
Thus the conclusion of Theorem 2.2.8 is true. This completes the proof.
Theorem 2.2.9. Let E be a Banach space, E0 be a closed subspace of
E and Ω ⊂ E be an open bounded subset. If T : Ω → E0 is a continuous
compact mapping and p ∈ E0 , then deg(I − T, Ω, p) = deg(I − T, Ω ∩ E0 , p).
Proof. Since T (Ω) ⊂ E0 , we may choose a finite dimensional space F ⊂ E0
with p ∈ F and T1 : Ω → F such that kT x − T1 xk < in Definition 2.2.3 for
small > 0. Then we have
x∈
/ tT x for all (t, x) ∈ [0, 1] × ∂Ω.
deg(I − T, Ω, 0) = deg(I, Ω, 0) = 1
kT x − T xk < , µx 6= T x
for all µ ∈ [0, 1], x ∈ ∂Ω and ∈ (0, ), where T is the same as in Lemma
2.2.1.
If this is not true, there exist j → 0, xj ∈ ∂Ω, µj → µ0 ∈ [0, 1] such that
µj xj = Tj xj and so we have T xj − µj xj → 0. Now, we have 0 ∈ / T ∂Ω and
so µ0 6= 0, (xj ) has a subsequence converging to x0 ∈ ∂Ω and T x0 = µ0 x0 ,
which is a contradiction.
From the definition of the Leray Schauder degree, we know that
deg(I − T, Ω, 0) = deg(I − T , Ω ∩ F, 0)
deg(I − T, Ω, 0) = 0.
deg(I − T, Ω, 0) = 0.
Assume that this is not true. Then there exists a compact mapping T1 : Ω →
ConvT ∂Ω such that T1 x = T x for x ∈ ∂Ω. For k > 1, it is easy to see that
x 6= tT x + (1 − t)kT1 x for all (t, x) ∈ [0, 1] × ∂Ω.
Thus we have
deg(I − kT1 , Ω, 0) = deg(I − T, Ω, 0) 6= 0
and so kT1 x = x has a solution in Ω for k > 1, which contradicts the fact
that Ω is unbounded. Hence we have deg(I − T, Ω, 0) = 1. This completes
the proof.
Theorem 2.2.16. Let E be a Banach space, Ω with 0 ∈ Ω be an open
bounded subset of E and L : Ω → E be a linear continuous compact mapping.
If λ 6= 0 and λ−1 is not an eigenvalue of L, then
deg(I − λL, Ω, 0) = (−1)m(λ) ,
where m(λ) is the sum of the algebraic multiplicities of the eigenvalues µ
satisfying µλ > 1, and, if L has no such eigenvalues µ, then m(λ) = 0.
Proof. Put S = I − λL, then S is a homeomorphism onto E. There are at
most finitely many eigenvalues of L such that µλ > 1, say, µi , i = 1, 2, · · · , k.
Set F = ⊕ki=1 N (µi ) and W = ∩ki=1 R(µi ), where N (µi ) = {x : Lx − µi x = 0}
and R(µi ) = (L − µi )(E) for i = 1, 2, · · · , k. We know from the spectral
theory of linear compact mappings that N (µi ) are finitely dimensional spaces,
i = 1, 2, · · · , k. It is easy to see that E = F ⊕ W . There are projections
P : E → F , and Q : E → W . Set L1 = SP + Q and L2 = P + SQ. Then we
have
I − L1 = −λLP1 , I − L2 = −λLQ, S = L1 L2 ,
(I − L1 )(F ) ⊂ F, (I − L2 )(W ) ⊂ W.
Moreover, Li is one to one for i = 1, 2. Thus, by product formula, we have
deg(S, Ω, 0) = deg(L1 L2 , Ω, 0) = deg(L1 , Ω, 0)deg(L2 , Ω, 0).
By Theorem 2.2.9, we have
deg(L1 , Ω, 0) = deg(L1 , Ω ∩ F, 0)
and
deg((L2 , Ω, 0) = deg(L2 , Ω ∩ W, 0).
But I − tλL has no solution on ∂Ω ∩ W and so we have
deg(L2 , Ω ∩ W, 0) = deg(I, Ω ∩ W, 0) = 1.
On the other hand, the eigenvalues of L1 are the eigenvalues of I − λL on
F , i.e., 1 − λµi for i = 1, 2, · · · , k. Thus we have
deg(L1 , Ω ∩ F, 0) = sgndet(L1 |F ) = sgnΠki=1 (1 − λµi )dim(N (µi )
k
= (−)Σi=1 dim(N (µi )) = (−1)m(λ) .
lim H(T x, T x0 ) = 0
x→x0
kxn k = 1, kxn − xm k ≥ ,
Proof. For any x ∈ X and > 0, there exists δx > 0 such that
We may require δx < . Let {Ui }i∈I be a locally finite open refinement of
{B(x, δ2x ) : x ∈ X} and {φi }i∈I be a partition of the unity subordinated
{Ui }i∈I .
since the right hand side is convex. Now, for n = 2−n for n = 1, 2, · · · , by
the above conclusion, we have a sequence {fn } of continuous mappings such
that
fn ∈ fn−1 (x) + 2Bn (0), fn (x) ∈ T x + Bn (0)
for all x ∈ X. Indeed, assume that we have defined mappings f1 , f2 , · · · , fn
for some n > 1, respectively, and put Gx = T x ∩ (fn + B (0)) for all x ∈ X.
Then G is lower semicontinuous with convex values.
By the first step, there exists fn+1 : X → Y such that fn+1 (x) ∈ G(x) +
Bn+1 and so we have
Evidently, since {fn } is a Cauchy sequence, let f (x) = limn→∞ fn (x). Then
f is continuous since the convergence is uniform and f (x) ∈ T x for all x ∈ X.
This completes the proof.
In the above result, we considered approximate selections and continuous
selections of multi-valued mappings. In some cases, we need to find a measur-
able selection for a given multi-valued mapping. Let (Ω, A) be a measurable
space, (X, d) be a separable metric space and f : Ω → 2X be a multi-valued
mapping. Then f is called a measurable function if f −1 (B) ∈ A for all open
subset B ⊂ X.
Theorem 2.3.9. Let (Ω, A) be a measurable space, (X, d) be a separable
complete metric space and F : Ω → 2X be a measurable multi-valued mapping
with closed values. Then there exists a single valued measurable mapping
f : Ω → X such that f (x) ∈ F x for all x ∈ Ω.
Proof. Since X is separable, there exists a countable subset {x1 , x2 , · · · }
of X such that
{x1 , x2 , · · · } = X.
We shall define a sequence {fn } of measurable functions satisfying the follow-
ing:
(1) d(fn (x), fn+1 (x)) < 2−n+1 for all x ∈ Ω;
(2) d(fn (x), F (x)) < 2−n for all x ∈ Ω
and define a mapping f1 : Ω → X by f1 (z) = xk if k is the smallest integer
such that
F z ∩ B(xk , 1) 6= ∅.
Since f −1 (xk ) = F −1 (B(xk , 1) \ ∪m<k F −1 (B(xm , 1)), it follows that f1 is
measurable.
Now, assume that we have defined f1 , f2 , · · · , fk satisfying (1) and (2)
for some k > 1. For each z ∈ Ω, we have z ∈ fk−1 (xi ) for some i. Now, if
z ∈ fk−1 (xi ), then we define
fk+1 (z) = xp ,
and x ∈ Rn , where M > 0 is a constant and f (·) ∈ Lp (Rn ) for p ∈ [1, ∞).
We define a mapping F : C([0, 1]; Rn ) → Lp (Rn ) by
for all u(·) ∈ C([0, 1]; Rn ). Since F u(t) : [0, 1] → 2R is upper semicontinuous
with closed values, it is measurable. By Theorem 2.3.9, we know that there
exists a measurable selection g(t) ∈ F u(t) and, by assumption, we have
Therefore, it follows that g(·) ∈ Lp ([0, 1]) and hence F is well defined.
Next, we show that the Leray Schauder degree can be generalized to multi-
valued upper semicontinuous compact mappings with closed convex values:
Proposition 2.3.11. Let E be a real Banach space, Ω ⊂ E be an open
bounded set and T : Ω → 2E be an upper semicontinuous mapping with
closed convex values. If T Ω is relatively compact and x ∈
/ T x for all x ∈ ∂Ω,
then there exists 0 > 0 such that x 6= f x for all x ∈ ∂Ω and ∈ (0, 0 ),
where f is defined as in Lemma 2.3.7.
Proof. Suppose that the conclusion is not true. Then there exist j → 0
and xj ∈ ∂Ω such that xj = fj xj . By Lemma 2.3.7, there exist yj ∈ Ω and
zj ∈ T yj such that
fj xj → z0 , xj → z0 ∈ ∂Ω
tj fj xj + (1 − tj )fδj xj = xj .
By Lemma 2.3.7, there exist yj1 , yj2 and zj1 ∈ T yj1 , zj2 ∈ T yj2 satisfying
kyj1 − xj k < j , kyj2 − xj k < δj , kzj1 − fj xj k < j , kzj2 − fδj xj k < δj .
Then the component C of M containing (α0 , 0) has at least one of the following
properties:
(1) C ∩ ∂Ω 6= ∅.
(2) C contains an odd number of trivial zeros (αi , 0) 6= (α0 , 0), where αi−1
is an eigenvalue of S of odd algebraic multiplicity.
Proof. First, if C ∩ ∂Ω 6= ∅, then C is compact and contains another (α, 0)
with α 6= α0 . The compactness of C follows from the compactness of S and
T . Suppose C ∩ R × {0} = {(α0 , 0)}. For any δ > 0, set
C ⊂ Ω0 ⊂ Ω0 ⊂ Ω, M ∩ ∂Ω0 = ∅.
Now, we may take δ > 0 small enough such that no other eigenvalue α−1 of
S satisfies |α − α0 | ≤ 2δ and the intersection of M and the real line is given
by I = [α0 − δ, α0 + δ]. Since M ∩ ∂Ω0 = ∅, deg(I − αS − T (α, ·), Ω(α), 0) is
constant on I, where Ω(α) = {x : (α, x) ∈ Ω0 }.
Now, we choose α0 − δ < α1 < α0 < α0 + δ. For r sufficiently small, we
have
Ω0 ∩ R × {0} = (∪N
i=1 [αi − δ, αi + δ]) × {0}.
and
Therefore, we have
−1
ΣN
j=1 [deg(I − αj2 S − T (αj2 , ·), B(0, r), 0)
−deg(I − αj1 S − T (αj1 , ·), B(0, r), 0) = 0.
Since the degree has a jump at α0 and the jumps occur only at eigenvalues
of odd algebraic multiplicity, it follows that the degree has an even number
of jumps. Consequently, C contains an odd number of trivial zeros (αi , 0) 6=
(α0 , 0).
with the norm |x(·)| = maxt∈[0,t0 ] |x(t)|. Then C([0, t0 ], Rn ) is a Banach space.
Put
K = {x(·) ∈ E : x(0) = x0 , |x(t) − x0 | ≤ r1 , t ∈ [0, t0 ]}.
Then it is easy to see that K is a bounded closed convex subset.
Now, we define a mapping T : E → E by
Z t
T x(t) = x0 + f (s, x(s))ds for all x(·) ∈ E.
0
Thus the problem (E 2.5.1) has a solution. This completes the proof.
In the following, let Ω ⊂ RN be an open bounded subset with smooth
boundary and ai , b : Ω×R×RN → R, i = 1, 2, · · · , N , be continuous functions
such that
∂ai
(1) ∂ηj ξi ξj ≥ |ξ|2 for all (x, z, ξ) ∈ Ω × R × Rn ;
Since the coefficients belong to C α (Ω), it is well known that (E 2.5.3) has a
unique solution u ∈ C 2,α (Ω) (see [132]). We define a mapping T : [0, 1] ×
C 1,α (Ω) → C 1,α (Ω) by
Therefore, we have
Z T Z T
|u0i (t)|2 dt ≤ |fi (t)|2 dt. (2.5.1)
0 0
By Lemma 2.5.4, we get
T
T2
Z
|ui |2∞ ≤ |fi (t)|2 dt. (2.5.2)
4 0
Put u(t) = Σ∞ i=1 ui (t)ei . Then it follows from (2.5.2) that u is well defined. By
(2.5.1), we know that u0 (t) = Σ∞ 0 2
i=1 ui (t)ei belongs to L (0, T ; H). Therefore,
∞ 2
Σi=1 λi ui (t)ei belongs to L (0, T ; H). Since A is closed, u(t) ∈ D(A) for
almost all t ∈ R and
Au(t) = Σ∞
i=1 λi ui (t)ei for almost all t ∈ R.
Proof. Let
Z T
Wa1,2 = {u ∈ Wa : |u0 (t)|2 dt < ∞)}.
0
Since A only has a point spectrum, Lemma 2.5.5 implies that the problem
(E 2.5.8) has a unique solution u ∈ Wa1,2 . Now, we define a mapping K :
Wa → Wa as follows: For each v ∈ Wa , Kv is the unique solution of (E 2.5.8).
Next, we prove that K is continuous. Suppose that vn → v0 in Wa . Then,
by Lemma 2.5.4, |vn − v|∞ → 0 as n → ∞. Now, ∂G and F are continuous
functions, so
as n → ∞. Also, we have
Multiply both sides of (2.5.3) by (Kvn (t) − Kv0 (t))0 and integrate over (0, T ),
and we have
Z T
|(Kvn (t) − Kv0 (t))0 |2 dt
0
Z T
+ hA(Kvn (t)) − Kv0 (t)), (Kvn (t) − Kv0 (t))0 idt
0
Z T
+ h∂GKvn (t) − ∂GKv0 (t), (Kvn (t) − Kv0 (t))0 idt
0
Z T
+ hF (t, vn (t)) − F (t, v0 (t)), (Kvn (t) − Kv0 (t))0 idt = 0.
0
Thus we have
Z T 21
|(Kvn (t) − Kv0 (t))0 |2 dt
√ 0 √
≤ T |∂Gvn (·) − ∂Gv0 (·)|∞ + T |F (·, vn (·)) − F (·, v0 (·))|∞
→ 0 (n → ∞).
Therefore, it follows that Kvn (·) → Kv0 (·) in Wa . For each v ∈ Wa , again,
by (E 2.5.8), we get
Z T Z T Z T
0 2 0
((Kv(t)) ) dt + h∂Gv(t), (Kv(t)) idt + hF (t, v(t)), (Kv(t))0 idt = 0.
0 0 0
By (2.5.4) and the boundedness of ∂G, we know that K maps bounded sets of
Wa to bounded sets in Wa The compact embedding of D(A) into H implies
that K is a compact mapping. In view of Lemma 2.5.4, we may simply take
|u0 |L2 as the norm of u in Wa .
RT 1
Now, we prove that Kv 6= λv for v ∈ Wa with |v 0 |L2 > ( 0 f 2 (t)dt) 2 and
λ > 1. In fact, if this is not true, then there exist λ0 > 1 and v0 ∈ Wa with
RT 1
|v00 |L2 > ( 0 f 2 (t)dt) 2 such that Kv0 = λ0 v0 , i.e.,
Multiply both sides of (2.5.5) by v00 (t) and integrate over [0, T ], we obtain
Z T
hλ0 v00 (t) + λ0 Av0 (t) + ∂Gv0 (t) + F (t, v0 (t)), v00 (t)idt = 0.
0
From Theorem 2.5.6, we know that the problem (E 2.5.10) has a solution, so
(E 2.5.9) has a solution.
Remark. The degree theory in this chapter can be established in locally
convex spaces (see [197]) or admissible topological vector spaces (see [170],
[235], [307]).
2.6 Exercises
1. Let f (x, y) : R × R → R be a continuous
Rs function and T : C[a, b] →
C[a, b] be defined by T x(·)(s) = a f (t, x(t))dt for all x(·) ∈ C[a, b] and
s ∈ [a, b]. Show that T is continuous and compact.
2 π
Z
T x(t) = [sin t sin s + c sin 2t sin 2s][2x(s) + x3 (s)]ds
π 0
for all x(·) ∈ C[0, π]. Compute T 0 (x) and the eigenvalues of T 0 (0).
4. Let E be a infinite dimensional Banach space. Show that the unit sphere
of E is not compact.
has a solution.
14. Let g ∈ C 1 ([0, a]), h ∈ C 1 ([0, b]) with g(0) = h(0) and f : [0, a] ×
[0, b] × R2 → R be continuous such that |f (x, y, z, 0)| ≤ M (1 + |z|)
and |f (x, y, z, u) − f (x, y, z, v)| ≤ L|u − v|, where a, b, L, M are positive
constants. Show that the following equation
∂u ∂u
∂x∂y = f (x, y, u(x, y), ∂x ),
(x, y) ∈ [0, a] × [0, b],
u(x, 0) = g(x), x ∈ [0, a],
u(0, y) = h(y), y ∈ [0, b],
The Leray Schauder degree theory is very useful in solving an operator equa-
tion of the type (I − T )x = y, where T is compact. In many applications T
is not compact, so one may ask it is possible to give an analogue of the Leray
Schauder theory in the noncompact case. In 1936, Leray [184] constructed an
example to show that it is impossible to define a degree theory for mappings
with only a continuity condition. So a very natural question which arises is
the following:
For what kind of mappings in infinite dimensional spaces can we establish
a degree theory ?
Browder, Nussbaum, Sadovski, Vath, etc., showed that it is possible to
define a complete analogue of the Leray Schauder theory for condensing type
mappings T .
In this Chapter, we will introduce the degree theory for k-set contraction
mappings and condensing mappings. This chapter consists of three sections.
In Section 3.1, we define measures of non-compactness and present some
properties (see propositions 3.1.5, 3.1.7 and theorems 3.1.14, 3.1.15). Also,
countably condensing maps, etc., are defined here and, in particular, a fixed
point theorem for countably condensing self-mappings is presented in Corol-
lary 3.1.18.
Section 3.2 presents a degree theory for countably condensing mappings and
the theory is based on the use of retractions and the Leray Schauder degree.
Again, various properties and consequences are presented.
In Section 3.3, we use the degree of Section 3.2 to discuss the initial and
anti-periodic ordinary differential equations in Banach spaces.
55
Proof. (1)-(3) are obvious. To see (4), let x, y ∈ conv(A). There exist
si ∈ (0, 1), xi ∈ X, i = 1, 2, · · · , k, ti ∈ (0, 1), yi ∈ A, i = 1, 2, · · · , m, such
that x = Σki=1 si xi and y = Σmi=1 si yi . Now, we have
kx − yk = kΣki=1 si xi − Σmi=1 si yi k
= Σki=1 Σm k m
j=1 i j i − Σi=1 Σj=1 si tj yj k
s t x
≤ Σki=1 Σm
j=1 si tj kxi − yj k
≤ Σki=1 Σm
j=1 si tj diam(A).
and so
H(A, B)
≤ max{sup d(x, C) + sup d(z, B), sup d(y, C) + supz∈C d(z, A)}
x∈A z∈C y∈B
≤ max{sup d(x, C), supz∈C d(z, A)} + max{sup d(z, B), sup d(y, C)}
x∈A z∈C y∈B
= H(A, C) + H(C, B).
β(A) = inf{δ > 0 : A is covered by finitely many balls with radius δ},
For (7): Obviously, α(A) ≤ α(conv(A)). For any > 0, there exists a finite
cover {A1 , A2 , · · · , Ak } of A with diam(Ai ) < α(A) + for i = 1, 2, · · · , k.
We may also assume that Bi is convex since diam(conv(Bi )) = diam(Bi ) for
i = 1, 2, · · · , k. Put
Λ = {(λi , λ2 , · · · , λk ) : λi ≥ 0, i = 1, 2, · · · , k, Σki=1 λi = 1}
Therefore, we have
Therefore, we have
{x1 , x2 , · · · } ⊂ {x1 , x2 , · · · , xL } ∪ B(Y, rn + ) ⊂ ∪ki=1 B(zi , rn + 2).
Thus β({xn }) ≤ rn + 2, i.e.,
β({xn }) ≤ inf{rn , n ≥ 1} = lim lim sup d(xm , Xn ).
n→∞ m→∞
On the other hand, for > 0, put r = β({xi : i ≥ 1}) and there exists
finitely many wi , 1 ≤ i ≤ s, such that {x1 , x2 , · · · } ⊂ ∪si=1 B(wi , r + ). By
the construction of Xn , we know that there exists an integer K > 0 such that
d(wi , Xn ) < for i = 1, 2, · · · , s and n > K. Therefore, we have
d(xm , Xn ) ≤ inf{kxm − wi k : 1 ≤ i ≤ s} + sup{d(wi , Xn ) : 1 ≤ i ≤ s}
≤ r + 2
for m ≥ 1 and n > K. From this, we get
lim lim sup d(xm , Xn ) ≤ r.
n→∞ m→∞
Thus we have limn→∞ lim supm→∞ d(xm , Xn ) = β({xi : i ≥ 1}). This com-
pletes the proof.
Definition 3.1.9. Let X be a real normed space, T : D → X be a mapping
and α be the measure of noncompactness.
(1) T is called a k-set contraction if α(T B) ≤ kα(B) for all bounded subsets
B ⊂ D, where k > 0 is a constant;
(2) T is said to be condensing if α(T B) < α(B) for all bounded subsets
B ⊂ D with α(B) > 0.
for t, t0 ∈ [a, b] satisfying |t − t0 | < γ. From this and (6) of Proposition 3.1.6,
we infer that
for t, t0 ∈ [a, b] satisfying |t−t0 | < γ. Thus α({x(t) : x(·) ∈ B}) is a continuous
function on [a, b].
For any division of [a, b]: a = t0 < t1 < · · · < tn = b, where ti = a + i b−an ,
i = 0, 1, · · · , n. For any > 0, from the equicontinuity of B, there exists
N > 0 such that, if n > N , then
i.e.,
b
b−a
Z
lim α({Σni=1 x(ti ) : x(·) ∈ B}) = α({ x(t)dt : x(·) ∈ B}).
n→∞ n a
b−a b−a
α({Σni=1 x(ti ) : x(·) ∈ B}) ≤ Σni=1 α({x(ti ) : x(·) ∈ B}) .
n n
Therefore, it follows that
Z b Z b
α({ x(t)dt : x(·) ∈ B}) ≤ α(B(t))dt.
a a
Next, for α(B) < δ, there exist B1 , B2 , · · · , Bm ⊂ C([a, b), E) such that
diam(Bi ) ≤ δ and B ⊂ ∪m
i=1 Bi . Hence, for each t ∈ [a, b], we have
{x(t) : x(·) ∈ B} ⊂ ∪m
i=1 {x(t) : x(·) ∈ Bi }.
Also, we have
{x(t) : x(·) ∈ B} ⊂ ∪ni=1 {x(ti ) : x(·) ∈ B} + B (0) for all t ∈ [a, b].
If δ > maxt∈[a,b] α({x(t) : x(·) ∈ B}), we can find finitely many subsets
A1 , A2 , · · · , As ⊂ E such that
Proof. Put
L ⊆ convT (L ∩ Ω.
Thus we have
α(L) ≤ α(convT (L ∩ Ω) < α(L ∩ Ω),
which is a contradiction. Therefore, C is compact. This completes the proof.
Proposition 3.1.18. [291] Let E be a Banach space, Ω ⊂ E be a bounded
subset and T : Ω → E be a countably condensing mapping. Set C1 =
conv(T Ω), Cn+1 = conv(T (Cn ∩ Ω)) for n ≥ 1 and C = ∩∞ n=1 Cn . Then
C is convex and compact.
Proof. This is a special case of Lemma 7.2.1.
Corollary 3.1.19. Let C ⊂ E be a nonempty bounded closed convex
subset and let T : C → C be a continuous countably condensing mapping.
Then T has a fixed point in C.
Proof. We first assume that T is k-set countably condensing for some
k ∈ [0, 1). Let C1 = convT C and Ci+1 = convT Ci for i = 1, 2, · · · . By
Proposition 3.1.18, K = ∩∞ i=1 Ci is convex and compact and also T : K → K
is a mapping.
Now, we prove that K is non-empty. Take x0 ∈ C, then T i x0 ∈ Ci for
i ≥ 1. We have α({T i x0 , i ≥ n}) ≤ k n α({T0i , i ≥ 0}) for n ≥ 1. Obviously,
and
/ r2−1 (Ω) ∩ Ω \ r1−1 (Ω) ∩ r2−1 (Ω) ∩ Ω.
0∈
Thus the excision property of the Leray Schauder degree implies that
and
Therefore, we have
One may also define a degree by taking the set C as in Proposition 3.1.18.
This degree will coincide with the above one by the excision property of the
Leray Schauder degree.
Theorem 3.2.1. The degree defined by 3.2.1 has the following properties:
(1) (N ormality) deg(I, Ω, 0) = 1 if and only if 0 ∈ Ω;
(2) (Solvability) If deg(I − T, Ω, 0) 6= 0, then T x = x has a solution in Ω;
(3) (Homotopy) Let H(t, x) : [0, 1] × Ω → E be a continuous and countably
condensing mappings, i.e., α([0, 1] × B) < α(B) for all countable subset
B of Ω with α(B) > 0 and H(t, x) 6= x for all (t, x) ∈ [0, 1] × ∂Ω. Then
deg(I − H(t, ·), Ω, 0) doesn’t depend on t ∈ [0, 1];
(4) (Additivity) Let Ω1 , Ω2 be two disjoint open subsets of Ω and 0 ∈
/
(I − T )(Ω − Ω1 ∪ Ω2 ). Then
Proof. (1), (2), and (4) follow directly from the definition and properties
of the Leray Schauder degree.
To prove (3), we set
mapping. Suppose x 6= λT x for all λ ∈ [0, 1), x ∈ ∂Ω. Then T has a fixed
point in Ω.
Proof. We may assume that T x 6= x for all x ∈ ∂Ω. Put H(t, x) = tT x
for all (t, x) ∈ [0, 1] × Ω. It is easy to see that {H(t, ·)}t∈[0,1] is a homotopy
of countably condensing mappings. By assumption, we have H(t, x) 6= x for
all x ∈ ∂Ω. As a result, deg(I − T, Ω, 0) = deg(I, Ω, 0) = 1. Thus T x = x has
a solution in Ω, which is the desired result. This completes the proof.
Corollary 3.2.3. Let E be a Banach space, Ω ⊂ E be an open bounded
subset with 0 ∈ Ω and T : Ω → E be a continuous and countably condensing
mapping. Suppose kT xk ≤ kxk for all x ∈ ∂Ω. Then T has a fixed point in
Ω.
Proof. We may assume that T x 6= x for all x ∈ ∂Ω. Otherwise, the
conclusion is true. Thus we have T x 6= λx for all x ∈ ∂Ω and λ ≤ 1. By
Theorem 3.2.2, T has a fixed point in Ω.
Theorem 3.2.4. Let E be a Banach space and T : E → E be a continuous
and countably condensing mapping. Then one of the following conclusions
holds:
(1) T has a fixed point in E;
(2) {x : T x = λx for some λ > 1} is unbounded.
where k ∈ (0, 1) and r0 > 0 are constants. Then there exists t0 ∈ (0, 1] such
that the initial value problem
(
x0 (t) = f (t, x(t)), t ∈ (0, t0 ),
(E 3.3.1)
x(0) = x0
has a solution.
Proof. Set
r0
M = sup{kf (t, x)k : (t, x) ∈ [0, 1] × B(x0 , r0 }, t0 = min{1, }.
M
Obviously, (E 3.3.1) is equivalent to the following integral equation:
Z t
x(t) = x0 + f (s, x(s))ds. (E 3.3.2)
0
Put X = C([0, t0 ], E) with the norm kx(·)k = max{kx(t)k : t ∈ [0, t0 ]}. Then
X is a Banach space. We also set K = {x(·) ∈ X : x(t0 ) = x0 , kx(t)−x(t0 )k ≤
r0 }. Then K is a bounded closed convex subset of X.
Now, we define a mapping T : K → K by
Z t
T x(t) = x0 + f (s, x(s))ds for all x(·) ∈ K.
0
≤ t0 kα(B).
Thus T is a condensing mapping, so T has a fixed point in K, i.e., (E 3.3.2)
has a solution. Consequently, (E 3.3.1) has a solution. This completes the
proof.
Theorem 3.3.2. Let H be a real Hilbert space, T > 0 be a constant,
f (t, x) : R × H → H be a continuous mapping satisfying
α(f ([0, T ] × B) ≤ kα(B)
for all bounded subsets B of H, where k ∈ (0, 1) is a constant, and kT < 1.
If f (t + T, −x) = −f (t, x) and kf (t, x)k ≤ M kxk + g(t) for all (t, x) ∈ R × E,
where 0 ≤ M T < 2 is a constant, and g(·) ∈ L2 (0, T ), then the following
problem (
x0 (t) = f (t, x(t)), t ∈ R,
(E 3.3.3)
x(t + T ) = −x(t)
has a solution.
Proof. Let Ca = {x(·) : R → H is continuous , x(t + T ) = −x(t), t ∈ R}.
Define kx(·)ka = maxt∈[0,T ] for x(·) ∈ Ca , and it is easy to check that Ca
is a Banach space under this norm. It is simple to check that (E 3.3.3) is
equivalent to the following equation:
1 T 1 t
Z Z
x(t) = − f (s, x(s))dt + f (s, x(s))ds. (E 3.3.4)
2 t 2 0
We define a mapping S : Ca → Ca by
1 T 1 t
Z Z
Sx(t) = − f (s, x(s))dt + f (s, x(s))ds for all x(·) ∈ Ca .
2 t 2 0
For any bounded subset U of Ca , we have, by Theorem 3.1.16, that
α(SU ) ≤ kT α(U ).
Multiply both sides of (3.3.1) by x0 (t) and integrate over [0, T ], we have
Z T Z T
0
λ 2
kx (t)k dt = f (t, x(t))x0 (t)dt
0 0
Z T Z T
0
≤M kx(t)kkx (t)kdt + g(t)kx0 (t)kdt.
0 0
i.e.,
T
M T −1 T 2
Z Z
1 1
0 2
( kx (t)k dt) ≤ (1 −
2 ) ( g (t)dt) 2 .
0 2 0
3.4 Exercises
1. Let Ai ⊂ C([0, 1]), i = 1, 2, 3, be defined by
A1 = {x(·) : x(0) = 0, x(1) = 1 and 0 ≤ x(t) ≤ 1 for t ∈ [0, 1]};
A2 = {x(·) : 0 ≤ x(t) ≤ 12 , t ∈ [0, 21 ], 1
2 ≤ x(t) ≤ 1, t ∈ [ 12 , 1]} ∩ A1 ;
A3 = x(·) : 0 ≤ x(t) ≤ 32 , t ∈ [0, 21 ], 1
3 ≤ x(t) ≤ 1, t ∈ [ 21 , 1]} ∩ A1 ;
Show β(A1 ) = 12 , i = 1, 2, 3, and α(A1 ) = 1, α(A2 ) = 12 , α(A3 ) = 32 .
α1 (B) = max{ max α({x(t) : x(·) ∈ B}), max α({x0 (t) : x(·) ∈ B})}.
t∈[a,b] t∈[a,b]
has a solution u(s, t) such that u(s, ·) ∈ C([0, 1]) for all s ∈ R.
75
for x = Σ∞ i=1 αi (x)ei . In the case of a separable Hilbert space, we may choose
an orthonormal basis {ei : i ∈ N }, then the projection Pn x = Σni=1 (x, ei )ei
satisfies Pn∗ = Pn and kPn k = 1.
Example 4.1.4. Let X be a reflexive Banach space with a projection
scheme such that Pn Pm = Pmin{m,n} . Then {Pn∗ X ∗ , Pn∗ } is a projection
scheme for X ∗ .
Proof. Notice that Pn∗ Pn∗ f (x) = f (Pn2 x) = Pn∗ f (x) on X and thus Pn∗ is a
projection. We also have
We claim that X ∗ = ∪∞ ∗ ∗
i=1 Pi X . If not, there is x0 ∈ X \ {0} such that
f (x) = 0 for all f ∈ ∪∞ P ∗ X ∗ since X ∗∗ = J(X), where J(x)(f ) = f (x) for
i=1 i
all f ∈ X ∗ and x ∈ X. Thus, we have f (Pn x) = 0 for all n and f ∈ X ∗ , so
f (x) = 0 for all f ∈ X ∗ and so x = 0, which is a contradiction. Therefore,
X ∗ = ∪∞ ∗ ∗ ∗ ∗ ∗ ∗
i=1 Pi X . We also have Pn X ⊂ Pm X for n < m. Thus, for any
f ∈ X ∗ and > 0, we may choose g ∈ Pn∗ X ∗ such that kf − gk < and we
then have
∗
which gives Pm f → f as m → ∞.
Example 4.1.5. If both X and Y have Schauder basis, then there exists
an operator projection scheme.
Proof. Let {en } be a Schauder basis of X and {e0n } be a Schauder basis
of Y . Put Xn = span{e1 , e2 , · · · , en } and Yn = span{e01 , e02 , · · · , e0n }. For
x = Σ∞ ∞ 0 n n 0
i=1 αi ei and y = Σi=1 βi ei , set Pn x = Σi=1 αi ei and Qn y = Σi=1 βi ei .
Then Π = {Xn , Pn ; Yn , Qn } is an operator projection scheme.
Definition 4.1.6. Let X, Y be real Banach spaces and Π = {Xn , Pn ; Yn ,Qn }
be an operator projection scheme. Then a mapping T : D ⊂ X → Y is
called A- proper (respectively, pseudo A-proper) with respect to Π if, for any
bounded xm ∈ D ∩ Xm and Qm T xm → y, there exists a subsequence {xmk }
such that xmk → x ∈ D and T x = y, (respectively, there exists x ∈ D(T ),
such that T x = y). We denote by AΠ (D, Y ) the class of all A-proper mappings
F :D →Y.
Recall that, in a normed space X, the semi-inner products (·, ·)− and (·, ·)+
are defined by
For some properties of accretive operators, we refer the reader to [17] and
[60]. One can show that a continuous strongly accretive operator T : X → X
is A-proper, and the proof is left to the reader as an exercise.
Definition 4.1.7. Let X be a separable Banach space with a projection
scheme Π = {Xn , Pn }. Then T : D ⊆ X → X is called a P1 compact mapping
if λI − T is A-proper with respect to Π for all λ ≥ 1.
In the following, let X, Y be separable Banach spaces, S : X → Y be a linear
Fredholm mapping of index zero with N (S) 6= {0} and C : D ⊂ X → Y be
a nonlinear mapping. Consider the semilinear problem Sx − Cx = y for all
x ∈ D(L) ∩ D and y ∈ Y . Since S is Fredholm of index zero (see Chapter V),
there exist closed subspaces X 0 of X and Y 0 of Y with dimY 0 = dimN (S)
such that
X = N (S) ⊕ X 0 , Y = Y 0 ⊕ R(S).
Let P : X → N (S) be a projection, Q : Y → Y 0 be a projection and
M : N (S) → Y 0 be a isomorphism. Put T = M P . Then T is a compact
linear operator. It is known that S + T is also a Fredholm mapping with
ind(S + T ) = ind(S) = 0 and S + T is bijective with (S + T )−1 : Y → X
bounded. Set S1 = S|X 0 ∩D(S) . Then S1 is injective and closed and so S1−1 is
continuous on R(S).
Thus we have
Since P and S1−1 (I − Q) are compact, it is easy to see that (xnj )∞ j=1 has a
subsequence (x0nj )∞
j=1 converging to x0 . From the continuity of N , it follows
that Sx0nj → g + λN x0 . The closedness of S implies that
Sx0 − λN x0 = g
Qn p ∈
/ Qn T (∂(Ω ∩ Xn )) for all n > n0 .
Proof. Suppose that the assertion of Lemma 4.2.1 is not true. Then
there exists nk → ∞ and xnk ∈ ∂Ω ∩ Xnk such that Qnk p = Qnk T xnk .
Obviously, xnk ∈ ∂Ω ∩ L. Thus we have Qnk T xnk → p as k → ∞ and the A-
properness of T guarantees the existence of a subsequence (xnkl )∞
l=1 such that
xnkl → x0 ∈ ∂Ω ∩ L, and T x0 = p, which is a contradiction. This completes
the proof.
Definition 4.2.2. Let T ∈ AΠ (Ω ∩ L, Y ). Suppose that p ∈ / T (∂Ω ∩ L)
and Qn T is continuous. We define a generalized degree D(T, Ω, p) by
Thus there exists xnk ∈ Ω ∩ L such that Qnk T xnk = Qnk p. By the A-
properness of T , there is a subsequence (xnkj ) with xnkj → x0 ∈ Ω ∩ L and
T x0 = p.
(2) Since p 6∈ (Ω \ Ω1 ∪ Ω2 ) ∩ L, there exists n0 > 0 such that
/ (Ω \ Ω1 ∪ Ω2 ) ∩ Xn
Qn p ∈ for all n > n0 .
Therefore, we have
and
lim sup |deg(Qnj T, Ω2 ∩ Xnj , Qnj p)| < +∞
j→∞
Qn p ∈
/ ∪t∈[0,1] H(t, ∂Ω ∩ Xn ) for all n > n0 .
(1) If p ∈
/ (S − N )(∂Ω ∩ D(S)) and Deg(S − N, Ω ∩ L, p) 6= {0}, then
Sx − N x = p has a solution in Ω ∩ D(S);
deg(Qn (S − A), Ω ∩ Xn , 0) = 1 or − 1.
1 t
(1) H(t, x) = Sx − [ 1+t N (x) − 1+t N (−x)] is A-proper with respect to ΓS
for all t ∈ [0, 1];
(3) S(x) − N (x) 6= λ[S(−x) − T (−x)] for all x ∈ ∂Ω ∩ D(S) and λ ∈ [0, 1].
for all (t, x) ∈ [0, 1] × Ω ∩ D(S). Since QN is compact, it follows from (1) that
H is an A-proper homotopy.
Now, we claim that H(t, x) 6= 0 for all (t, x) ∈ [0, 1] × ∂Ω ∩ D(S). If this is
not true, then there exists (t0 , x0 ) ∈ [0, 1] × ∂Ω ∩ D(S) such that
If t0 ∈ (0, 1), then Sx0 −t0 N x0 −t0 p = (1−t0 )(QN x0 +Qp), so QN x0 +Qp 6=
0. Therefore, we have
0 = QSx0 = QN x0 + Qp,
(3) [Q(N x + p), x][Q(N (−x) + p), x] < 0 for all x ∈ N (S) ∩ ∂Ω.
Then Sx − N x = p has a solution in Ω ∩ D(S).
Proof.If Sx − N x = p has a solution on ∂Ω ∩ D(S), we are done, so we
may assume that Sx − N x 6= p for x ∈ ∂Ω ∩ D(S).
Consider the homotopy H : [0, 1] × Ω ∩ D(S) → Y given by
for all (t, x) ∈ [0, 1] × Ω ∩ D(S). Since QN is compact, it follows from (1)
that H(t, ·) is A-proper for all t ∈ [0, 1] and N is bounded. Therefore, H is
an A-proper homotopy.
Now, we prove that H(t, x) 6= 0 for (t, x) ∈ [0, 1]×∂Ω∩D(S). In fact, assume
the contrary, there exists (t0 , x0 ) ∈ [0, 1] × ∂Ω ∩ D(S) such that H(t0 , x0 ) = 0,
i.e.,
Sx0 − (1 − t)[Q(N (x0 ) + p)] − t0 N (x0 ) − t0 p = 0.
By (2), we know that t0 6= 1. If t0 = 0, then Sx0 = Q(N (x0 ) + p). But
R(L) ∩ Y 0 = {0}, so we have Sx0 = 0 and Q(N (x0 ) + p) = 0. Therefore,
[Q(N (x0 ) + p), x0 ] = 0, which contradicts (3). Thus we must have t0 ∈ (0, 1).
Therefore,
i.e., Q(N (x0 )+p) = 0 which contradicts (3) again. By the homotopy property
of the generalized degree, we get
Now, we prove that {0} =6 Deg(S − Q(N + p), Ω ∩ D(S), 0). To reach this
goal, we consider the homotopy H1 (t, x) : [0, 1] × Ω ∩ D(S) → Y given by
1
H1 (t, x) = Sx − [Q(N (x) + p) − t(Q(N (−x) + p)]
1+t
1
/ Deg(S − [Q(N (·) + p) − Q(N (−·) + p)].
0∈
2
Thus we have Deg(S −N −p, Ω∩D(S)) 6= {0} and, consequently, Sx−N x = p
has a solution. This completes the proof.
dim(Xn ) = dim(Yn ), X = ∪∞
n=1 Xn .
so we have
(I − P )xnk * x0 , J −1 QT xnk * J −1 QT x0
and
KP Q T xnk * KP Q T x0
and, consequently, (I − P − J −1 QT − KP Q T )x0 = y. So T is pseudo L−A-
proper with respect to Γ0 .
Proposition 4.4.8. Let X, Y be real separable Banach spaces, (Yn , Qn ) be
a projection scheme for Y . Let L : D(L) ⊂ X → Y be a Fredholm mapping
of zero index type, G ⊂ X be a bounded closed subset and N : G → Y be a
continuous compact mapping. Then L + λJP − N is A-proper with respect
Γλ,L for each λ > 0.
Proof. For any sequence (xnk ) in G ∩ D(L) ∩ Xnk with Qnk (L + λJP −
N )xnk → y, in view of the compactness of N , we may assume that N xnk →
y0 ∈ Y by taking a subsequence. Notice that
so we have
Proof. (1)-(3) follows directly from the definition and the properties of
generalized degree.
(4) Since Ker(L) = {0}, P = 0 and Q = 0, the zero mapping is L−A-proper
with respect to Γ0 . Thus degΓ0 ,J (L, Ω, 0) = deg(I, Ω, 0) = {1}.
(5) Since N is odd, the mapping I − P − (J −1 Q + KP Q )N is odd and
thus deg(I − P − (J −1 Q + KP Q )N, Ω, 0) doesn’t contain even numbers. The
conclusion follows by definition.
Corollary 4.4.11. Let L : D(L) ⊆ X → Y be a linear mapping such that
L−1 : Y → D(L) is continuous, Ω ⊂ X be an open bounded subset with 0 ∈ Ω
and N : Ω → Y be a mapping such that {L − tN }t∈[0,1] is a homotopy of L-A-
proper mappings respect to Γ0 . If Lx ∈
/ tN x for all (t, x) ∈ [0, 1] × ∂Ω ∩ D(L),
then deg(L − N, Ω, 0) = 1.
In the following, let L : D(L) ⊂ X → Y be a densely defined Fredholm
mapping of zero index type. We assume that Γ0 = (Yn , Qn ) is a projection
scheme for Y , Γλ,L is the same as in Proposition 4.4.3 and L + λJP − N is
A-proper with respect to Γλ,L for λ ∈ (0, λ0 ), where λ0 > 0 is a constant.
Suppose that 0 ∈/ (L − N )(D(L) ∩ ∂Ω). Then there exists λ1 < λ0 such that
0∈
/ (L + λJP − N )(D(L) ∩ ∂Ω) for all λ ∈ (0, λ1 ).
0∈
/ (L + λJP − N )(D(L) ∩ Ω \ (Ω1 ∪ Ω2 ))
deg(L + λj JP − N, Ω, 0)
⊆ deg(L + λj JP − N, Ω1 , 0) + deg(L + λj JP − N, Ω2 , 0)
0∈
/ ∪t∈[0,1] (L + λJP − H(t, ·))(∂Ω ∩ D(L)) for all λ ∈ (0, λ1 ).
By Theorem 4.2.3, deg(L + λJP − H(t, ·), Ω, 0) does not depend on t ∈ [0, 1]
for λ ∈ (0, min{λ0 , λ1 }). So the conclusion of (2) follows from (4.4.2).
(3) If degΓ0 (L−N, Ω, 0) 6= {0}, then there exists 0 6= m ∈ degΓ0 (L−N, Ω, 0),
so there exists λj → 0+ such that m ∈ deg(L + λj JP − N, Ω, 0). Therefore,
(L + λj JP − N )x has a solution in Ω ∩ D(L) for j = 1, 2, · · · . By letting
j → ∞, we get 0 ∈ (L − N )(D(L) ∩ Ω).
(4) The proof is left to the reader.
(5) Now, L+λJP is A-proper with respect to Γλ,L and 0 ∈ / (L+λJP )(∂Ω∩
D(L)) for all λ > 0. Since L + λJP ) is bijective, deg(L + λJP, Ω, 0) ⊆ {±1}
for all λ > 0, so we have
deg(L − N, Ω, 0) ⊆ {±1}.
Theorem 4.4.14. The generalized degree defined by (4.4.3) has the fol-
lowing properties:
Proof. The proof is standard. We prove (2) and skip the others. Since
0∈
/ ∪t∈[0,1] (L − H(t, ·)(D(L) ∩ ∂Ω), it follows that
does not depend on t ∈ [0, 1] for each λ > 0. Thus the conclusion of (2) follows
from (4.4.3). This completes the proof.
Theorem 4.4.15. Suppose that (L + λJP )−1 : Y → X is a continuous
compact mapping for each λ > 0, Ω ⊂ X is an open bounded subset with
0 ∈ Ω and N : Ω → Y is a continuous bounded mapping such that Lx 6= N x
and QN x 6= ηJP x for all x ∈ ∂Ω∩D(L) and η > 0, where P, Q are projections
as in the beginning of this section. Then deg(L − N, Ω, 0) = {1}.
Proof. Since (L + λJP )−1 : Y → X is continuous and compact for each
λ > 0, it follows that {I − (L + λJP )−1 t(N + λJP )}t∈[0,1] is a homotopy of
A-proper mappings.
Now, we claim that
for all (t, x) ∈ [0, 1]×(∂Ω∩D(L)) and λ > 0. If this is not true, then there exist
λ0 > 0 and (t0 , x0 ) ∈ [0, 1)×∂Ω such that x0 = (L+λ0 JP )−1 t0 (N x0 +λJP x0 ).
Thus we have x0 ∈ D(L) and
Lx0 + λ0 JP x0 = t0 (N x0 + λ0 JP x0 ).
deg(L − T, Ω, 0) = {1}.
Proof. Let x(·) be a solution of (E 4.5.1). Assume that |x(t)| achieves its
maximum at t = 0. Then x0 (0) = 0. Otherwise, |x(t)| can not achieve its
maximum at t = 0. Therefore, |x(0)| = |x(T )| ≤ M .
Lemma 4.5.3. Suppose the following conditions hold:
(1) There exists a constant M > 0 such that, for each solution x(·) of (E
4.4.1), |x(t)| ≤ M for t ∈ [0, T ];
where D = max{|gm |, |gM |}. Multiply (4.5.1) by x0 (t) and rearrange the
terms, we obtain
Integrate the last inequality over [µ, t] and use |x(t)| ≤ M and x0 (µ) = 0, we
obtain
αx0 (t)2 + β
ln( ) ≤ 2αM.
β
Therefore, we have
β 1
|x0 (t)| ≤ [ (e2αM − 1)] 2 = M1 .
α
This and the condition (2) implies that |x00 (t)| ≤ M2 for some M2 depending
on M, A, B, and C. This completes the proof.
Combine Lemma 4.5.1 and Lemma 4.5.2 with Lemma 4.5.3, we get the
following:
Proposition 4.5.4. Assume that the condition of Lemma 4.4.1 holds.
If there are continuous functions A(t, x), C(t, x) > 0 which are bounded on
compact subsets of [0, T ] × R and a constant B ∈ [0, 1] such that
for all (t, x) ∈ [0, T ] × [−M, M ] and r, q ∈ R, then there are constants M1 and
M2 such that, for any solution of (E 4.5.1),
(1) Let M > 0 and c, d be the same as in Lemma 4.5.1 and the condition of
Lemma 4.5.1 holds;
Then there are constants M1 , M2 > 0 such that, for λ ∈ [0, 1] and any solution
xλ (·) of (E 4.5.2),
N x(t) = f (t, x(t), x0 (t), x00 (t)) for all t ∈ [0, T ], x(·) ∈ X.
(3) There are continuous functions A(t, x), C(t, x) > 0 which are bounded
on compact subsets of [0, T ] × R and a constant B ∈ [0, 1] such that
If x(·) ∈ ∂Ω∩N (S), then kxk2 = r > M , so x ≡ r or −r. Thus the assumption
(2) implies that
Z T
QN (c) − Qg = [f (t, c, 0, 0) − g(t)]dt 6= 0
0
and Z T
[QN (c) − Qg, c] = [f (t, c, 0, 0) − g(t)]cdt = 0,
0
where c = r or −r. From Theorem 4.3.3, we know that the problem (E 4.5.1)
has a solution. This completes the proof.
A special case of (E 4.5.1) is the following:
(
x00 (t) = f (t, x(t), x0 (t)) − g(t), t ∈ [0, T ],
(E 4.5.3)
x(0) = x(T ), x0 (0) = x0 (T ).
In this case, the mapping given by N (x)(t) = f (t, x(t), x0 (t)) for all t ∈ [0, T ]
is compact, so S − λN is A-proper with respect to ΠS . From Theorem 4.5.6,
the following immediately holds:
Corollary 4.5.7. Let g(·) ∈ Y , f (t, x, r) : [0, T ] × R2 → R be a continuous
function and S, ΠS , N be the same as above. Suppose that the following
conditions hold:
|f (t, x, r)| ≤ A(t, x)r2 + C(t, x) for all (t, x) ∈ [0, T ] × [−M, M ].
and f (·) ∈ L2 ((0, 2π) × (0, π)), where δ > 0 and γ > 0 are constants.
We say that u ∈ L2 ((0, 2π) × (0, π)) is a weak solution of the problem (E
4.5.4) if
(u, vtt − vxx ) − (h(u(t, x)), v) = (f (t, x), v)
for all v ∈ C 2 ([0, 2π] × [0, π]) with v(t, 0) = v(t, π) = 0 for all t ∈ [0, 2π] and
v(2π, x) = v(0, x) for all x ∈ [0, π].
Let L : D(L) ⊂ L2 ((0, 2π)×(0, π)) → L2 ((0, 2π)×(0, π)) be the wave opera-
tor Lu = utt −uxx . Then it is well known that L is self-adjoint, densely defined,
and closed, and Ker(L) is infinite dimensional with Ker(L)⊥ = Im(L). Thus
L is a Fredholm mapping of zero index type. Let P : L2 ((0, 2π) × (0, π)) →
Ker(L) be the projection. Then (L + λP )−1 : L2 ((0, 2π) × (0, π)) → D(L) is
compact for all λ > 0.
For each η > 0, consider the following equation:
utt (t, x) − uxx (t, x) + ηu(t, x) − h(u(t, x)) = f (t, x),
t ∈ (0, 2π), x ∈ (0, π),
(E 4.5.5)
u(t, 0) = u(t, π) = 0, t ∈ (0, 2π),
u(0, x) = u(2π, x), x ∈ (0, π),
N u(t, x) = h(u(t, x)) + f (t, x) for all u(t, x) ∈ L2 ((0, 2π) × (0, π)).
Let Ω = {u(t, x) ∈ L2 ((0, 2π)×(0, π)) : kukL2 < r0 }. By (4.5.5), we know that
P N u 6= ηP u for all u ∈ C 2 ([0, 2π] × [0, π]) ∩ ∂Ω and η > 0. We may assume
that Lu 6= N u for all u ∈ C 2 ([0, 2π] × [0, π]) ∩ ∂Ω. By Theorem 4.4.16, we
have deg(L − N, Ω, 0) = {1}. Thus the problem (E 4.5.4) has a weak solution.
This completes the proof.
Remark. The results of Sections 4.1-4.3 in this chapter can be found in
[239].
4.6 Exercises
1. Let X be a separable Banach space with a projection scheme Π =
{Xn , Pn }, Ω ⊂ X be an open bounded subset with x0 ∈ Ω and T :
Ω → X be a P1 compact mapping satisfying
105
Proof. For the ”if” part, we know from our assumption that {x : T x =
0, kxk ≤ 1} is compact and thus Ker(T ) is finite dimensional. We have
X = Ker(T ) ⊕ M for some closed subspace M of X. Obviously, we have
T (M ) = Im(T ). Since T : M → Im(T ) is one to one, it follows that
Now, assume that L is a Fredholm mapping. Then there exist two linear
continuous projections P : X → X and Q : Y → Y such that
Also, we have
Now, the compactness of P implies that (xnk ) has a subsequence (xnkl ) such
that xnkl → x0 and x0 − KP Q g = P x0 − P KP Q g. Since L is closed, we have
L(x0 − KP Q g) = 0. Thus Lx0 = g. This completes the proof.
More precisely, we have the following result between the A-proper mapping
and the Fredholm mapping:
x − P x − J 0 QN x − KP Q N x = KP Q y + J 0 Qy.
x − P x − J 0 QN x − KP Q N x = KP Q y + J 0 Qy.
as k → ∞. Therefore, we have
zk = QN xnk + Qy → 0
and thus
hk = KP yk = (I − P )xnk − Pk KP Q (N xnk + y) → 0,
wk = J 0 zk = J 0 QN xnk + J 0 Qy → 0.
J 0 QN xnk → −J 0 Qy,
which immediately implies that
So (xnk )∞
k=1 has a convergence subsequence (xnkl ) with xnkl → x0 and x0 −
P x0 − J 0 QN x0 − KpQ N x0 = KpQ y + J 0 Qy. By Proposition 5.1.8, we have
Lx0 − N x0 = y. Thus L − N is A-proper with respect to Γm . This completes
the proof.
Assume now that L is a Fredholm mapping of index zero. Then, for any
isomorphism J : Im(Q) → Ker(L), the mapping JQ+KP Q is an isomorphism
from Y onto D(L) and
In fact, if y ∈ Y , we have
(JQ + KP Q )y = x ⇔ JQy = P x,
KP Q y = (I − P )x ⇔ Qy = J −1 P x,
LP KP Q y = L(I − P )x ⇔ Qy = J −1 P x,
(I − Q)y = Lx ⇔ y = (J −1 P + L)x.
Example 5.1.10. Let X be a real Banach space and T : X → X be a
linear continuous compact mapping. Then, by (5) of Theorem 2.1.15, we know
that dim(ker(I + T )) = dim(codim(I + T )) < +∞, so I + T is a Fredholm
mapping of index zero.
Example 5.1.11. Let f : [0, T ] → R be in L1 and consider the following
problem: (
x0 (t) = f (t), t ∈ (0, T ),
(E 5.1.1)
x(0) = x(T ).
We set X = C([0, T ], R), the space of all continuous function from [0, T ] to
R, Y = L1 ([0, T ], R) × R, y0 = (f (·), 0) and define a mapping L : X → Y by
Lx(·) = y0 .
which is called the coincidence degree of L and −T on Ω∩D(L). One can easily
prove that this definition does not depend on the choice of P, Q. It is known
that DJ (L + T, Ω, 0) is a constant for some J depending on orientations on
Ker(L) and Coker(L) (see [203]), so the coincidence degree in [203] is defined
only for those J 0 s. The definition given here depends on the J.
Therefore, if we only take those J such that detJ > 0, then we have DJ (T, Ω, 0) =
deg(T, Ω, 0), which is the Brouwer degree.
(2) If X = Y and we take L = I, then any continuous compact mapping T
on Ω is L-compact. If we take P = Q = 0, then KP Q = I, J = 0 : {0} → {0}
and HPJ Q F = I + T . Thus DJ (I + T, Ω, 0) = deg(I + T, Ω, 0), which is the
Leray Schauder degree.
Theorem 5.2.2. The coincidence degree of L and −T on Ω has the fol-
lowing properties:
(1) If Ω1 and Ω2 are disjoint open subsets of Ω such that 0 ∈
/ F (D(L) ∩
Ω \ (Ω1 ∪ Ω2 ), then
DJ (L + T, Ω, 0) = DJ (L + T, Ω1 ) + DJ (L + T, Ω2 , 0);
DJ (L + T1 , Ω, 0) = DJ (L + T2 , Ω, 0).
DJ (L + T, Ω, 0)
= deg(I − (A − J)(P A)−1 P, B(0, r))deg(I − P + HPAQ T, Ω, 0)
DJ (L + T, Ω, 0) = deg(I − P + KP Q T, Ω, 0).
Thus the conclusion follows from the assumption and Theorem 2.2.4.
Theorem 5.2.9. If Ω ⊂ X is open bounded with 0 ∈ Ω, Ω is symmetric
with respect to 0 and T is L-compact on Ω ∩ D(L) such that T (−x) = −T x
for all x ∈ ∂Ω ∩ D(L), then |DJ (L + T, Ω, 0)| is an odd number.
Proof. Since DJ (L + T, Ω, 0) = deg(I − P + KP Q T, Ω, 0), by the definition
of the Leray Schauder degree and Borsuk’s Theorem (Theorem 1.2.11), we
know that the conclusion is true.
In the following, let L : D(L) ⊂ X → Y be a Fredholm mapping of index
zero, and L = L1 + L2 , where L1 , L2 satisfying the following conditions:
I − P1 + H1 T1 = (I − P1 + H1 K1 )(I − P + HT ).
DJ1 (L1 + T1 , Ω, 0)
= deg((I − P1 + H1 K1 )(I − P + HT ), Ω, 0)
= Σi deg(I − P1 + H1 K1 , Ui , 0)deg(I − P + HT, Ω, Ui ),
which is called the coincidence index of L and T at a. One may easily see
that this definition is well defined by using the excision property of coincidence
degree.
The following result follows immediately from the definition:
Proposition 5.2.12. Let Ω ⊂ X be an open bounded subset and T : Ω →
Y be L-compact. If (L − T )−1 (0) = {a1 , a2 , · · · , ak } ⊂ Ω, then
DJ (L − T, Ω ∩ D(L), 0) = Σki=1 iJ (L − T, ai ).
kQT xk + kKP Q T xk
lim = 0, (5.2.2)
kxk→0 kxk
iJ (L − A − T, 0) = iJ (L − A, 0).
From the assumption (5.2.2), we know that there exists r > 0 such that
1+t 1−t
H(t, x) = Tx − T (−x) for all (t, x) ∈ [0, 1] × Ω.
2 2
1−t
Lx − T x = (−Lx − T (−x)),
1+t
(2) DJ (L + T2 , Ω, 0) 6= 0;
DJ (I − T, Ω, 0) = DJ (L − A, Ω, 0) 6= 0,
(2) T x ∈
/ Im(L) = 0 for all x ∈ Ker(L) ∩ ∂Ω;
(1) For almost all t ∈ [0, π], f (t, x, y) is continuous in (x, y);
(3) For all r > 0, there exists gr (·) ∈ L1 ([0, π], [0, +∞)) such that, For
almost all t ∈ [0, π],
Then Ker(L) = {0} and Im(L) = Y . Let a mapping N : C01 ([0, π], Rn ) → Y
be defined by
Lx = N x, x ∈ D(L). (E 5.4.2)
t π s
Z Z Z tZ s
−1
(L y)(t) = y(l)dl − y(l)dl,
π 0 0 0 0
i.e.,
1 t
Z Z π
−1
(L y)(t) = [ s(π − t)y(s)ds + t(π − s)y(s)ds,
π 0 x
(2) There exist c ≥ 0 and h(·) ∈ L1 ([0, π], R+ ) such that, for all x ∈ Rn
with |x| ≤ π(1 − a − b)−1 kgkL1 ,
Let x be a possible solution of (E 5.4.3) for some λ ∈ (0, 1). Then we have
where
√ kxk0 = maxt∈[0,π] kx(t)k. Notice that kxkL2 ≤ kx0 kL2 and kxk0 ≤
0
πkx kL2 , so (5.4.1) implies that
On the other hand, by the boundary condition, there exists si ∈ [0, π] such
that x0i (si ) = 0 for 1 ≤ i ≤ n and thus
Z t
|x0i (t)| ≤ | x00i (s)ds| ≤ kx00 kL1 ,
si
kx0 k0 ≤ r3 . (5.4.5)
where
Proof. Consider the homotopy T (α, x)) = αV (x(·)) − (1 − α)QV (x(·)) for
all α ∈ [0, 1] and x(·) ∈ X. We claim that Lx(·) 6= T (α, x(·)) for all α ∈ [0, 1]
and x(·) ∈ X. If not, there exist α ∈ [0, 1] and x(·) ∈ X such that
Multiply both sides of (5.4.6) by x0 (t) and integrate over [0, 1], one gets
Z 1
|x0 (s)|2 ds = 0.
0
(1) There exist v(·) ∈ C 1 ([0, 1], Rn ) such that lim|x|→∞ v(x) = +∞ and
β(·) ∈ L1 ([0, 1], [0, +∞)) such that (v 0 (x), f (t, x)) ≤ β(t) for all x ∈ Rn
and almost all t ∈ [0, 1].
(2) There exist r > 0 and w(·) ∈ C 1 (Rn , R) such that (v 0 (x), w(x)) > 0 for
R1
all x with |x| ≥ r and 0 (w0 (x(s)), f (s, x(s))ds) ≤ 0 for all x(·) ∈ D(L)
satisfying mint∈[0,1] |x(t)| ≥ r.
has a solution.
Proof. First, we claim that there exists r0 > 0 such that the solution xλ (·)
of the following equation:
(
x0 (t) = −(1 − λ)v 0 (x(t)) + λf (t, x(t)), t ∈ [0, 1],
(E 5.4.5)
x(0) = x(1),
Therefore, we have
Now, from kxλn k0 ≥ n and (5.4.7), we deduce mint∈[0,1] v(xλn (t)) → ∞, which
implies that mint∈[0,1] |xλn (t)| → ∞. Thus there exists N > 0 such that, for
n ≥ N,
min |xλn (t)| ≥ r. (5.4.8)
t∈[0,1]
Now, we have
d
w(xλn (t)) = −(1 − λn )(v 0 (xλn (t)), w0 (xλn (t)))
dt
+λn (w0 (xλn (t)), f (t, xλn (t))).
5.5 Exercises
1. Let X be a Banach space, T : X → X be a linear bounded Fredholm
operator and K : X → X be a linear continuous compact mapping.
Show that T + K is a Fredholm mapping.
|DJ (L − T − z, D(L) ∩ Ω, )| = 1.
(1) kLx − T xk2 ≥ kT xk2 − kLxk2 for all x ∈ (D(L) \ Ker(L)) ∩ ∂Ω;
(2) T x ∈
/ Im(L) = 0 for all x ∈ Ker(L) ∩ ∂Ω;
(3) deg(QTKer(L) , Ω ∩ Ker(L), 0) 6= 0, where Q : Y → Y is the projec-
tion such that Ker(Q) = Im(L).
(2) T x ∈
/ Im(L) = 0 for all x ∈ Ker(L) ∩ ∂Ω;
(3) deg(QTKer(L) , Ω ∩ Ker(L), 0) 6= 0, where Q : H → H is the pro-
jection such that Ker(Q) = Im(L).
Show that Lx = T x has a solution in D(L) ∩ Ω.
9. Let f : [0, π] × Rn × Rn → Rn be a function satisfying the Carathéodory
condition. Assume that the following condtions hold:
(1) There exist a, b ∈ R such that a + b < 1 and
(x − u, f (t, x, y) − f (t, u, v)) ≤ a|x − u|2 + b|x − u||y − u|
for all x, y, u, v ∈ Rn and a.e. t ∈ [0, π].
(2) There exist c ≥ 0 and h ∈ L1 ([0, π], [0, +∞)] such that, for all
x ∈ Rn with |x| ≤ π(1 − a − b)−1 kf (t, 0, 0)kL1 ,
|f (t, x, y)| ≤ c|y|2 + h(t) for all y ∈ Rn , and almost all t ∈ [0, π].
has a solution.
has a solution.
13. Let L : D(L) ⊂ X → Y be a Fredholm mapping of index zero, Ω ⊂
X be an open bounded subset, D(L) ∩ Ω 6= ∅ and T : Ω → 2Y be
a mapping with closed convex values. Assume that QT and KP Q T
are upper semicontinuous mapping such that QT (Ω) and KP Q T (Ω) are
relatively compact and Lx ∈/ T x for all x ∈ ∂Ω ∩ D(L). Construct the
coincidence degree for L and T on Ω ∩ D(L).
14. Let L : D(L) ⊂ X → Y be a Fredholm mapping of index zero, Ω ⊂ X
be an open bounded subset, D(L) ∩ Ω 6= ∅ and T : Ω → Y be a mapping
such that QT and KP Q T are continuous countably condensing mapping
and Lx 6= T x for all x ∈ ∂Ω ∩ D(L). Construct the coincidence degree
for L and T on Ω ∩ D(L).
The goal of this chapter is to introduce the degree theory for monotone-type
mapping. Chapter 6 has seven sections.
In Section 6.1, we introduce some basic geometric properties of Banach
spaces and various types of monotone and pseudomonotone maps and also
(S+ ), (S+ )0,L and L-(S+ )-mappings. Many examples and properties of these
maps are presented in Section 6.1.
Section 6.2 presents the degree theory for monotone mappings of class (S+ ).
In Section 6.3, using the results of Section 6.2, we present the degree theory
for perturbations of maximal monotone mappings and various properties are
also presented.
In Section 6.4, using the results of chapters 2, 3, we present the topological
degree for multivalued mappings of class (S+ )0,L . Some properties of this
degree are presented in theorems 6.4.4, 6.4.5, and 6.4.6.
A degree for multivalued mappings of class L-(S+ ) type is presented in
Section 6.5 (here L is a Fredholm mapping of index zero type). Various
properties are presented in Theorems 6.5.2, 6.5.3 and 6.5.5. The coincidence
degree of L and a pseudomonotone mapping is also presented in this section.
Section 6.6 presents various results concerning the computation of the topo-
logical degree for a variety of mappings.
Section 6.7 gives various existence results for the partial differential equa-
tions and evolution equations.
127
Proof. (1) For all x 6= 0, if f1 , f2 ∈ Jx, then we have (f1 + f2 )(x) = 2kxk2
and thus kf1 + f2 k = 2kxk. In addition, kf1 k = kf2 k = kxk, and so it follows
from the locally uniform convexity of E ∗ that f1 = f2 .
Next, assume that xn → x0 in E and we may assume also that (Jxn )) has
a weakly convergent sequence (Jxnk ) with Jxnk * f0 by reflexivity of E ∗ .
Then we have
Jxnk (xnk ) → kx0 k2 , Jxnk (x0 ) → f0 (x0 ).
In addition, we have
kJ(xnk )(xnk ) − J(xnk )(x0 )k = 0.
Therefore, we get f0 (x0 ) = kx0 k2 and, from this, we deduce that
kf0 k = kx0 k, kJ(xnk )k → kx0 k, kJ(xnk ) + f0 k → kx0 k = kf0 k.
∗
From the locally uniform convexity of E , we deduce J(xnk ) → Jx0 .
(2) For simplicity, we may assume that Jxn * f0 by taking a subsequence
since E ∗ is reflexive. By assumption, we have
Jxn (xn + x0 ) = kxn k2 + Jxn (x0 ) → 2kx0 k2
and so kxn + x0 k → kx0 k. From the locally uniformly convexity of E, we
deduce xn → x0 . This completes the proof.
The following is the well-known Mazur’s separation theorem for convex sets:
Theorem 6.1.9. Let X be a Banach space, C1 be an compact convex set
and C2 be a closed convex set such that C1 ∩ C2 = ∅. Then there exists a
f ∈ X ∗ \ {0} such that
sup f (x) < inf f (x).
x∈C1 x∈C2
(f − g, x − y) ≥ 0
xn * x0 , lim sup(Jxn , xn − x0 ) ≤ 0.
n→∞
for all v ∈ ∪∞
j=1 Fj , we have uj → u0 , and u0 ∈ D(T ), T u0 = h. If h = 0, then
we call T a mapping of class (S+ )0,L .
for all v ∈ ∪∞
j=1 Fj . Without loss of generality, we may assume that Jxj * f0
in E ∗ . Then we have
for all v ∈ ∪∞ ∞
j=1 Fj . But ∪j=1 Fj is dense in E, so we have f0 = h. Therefore
it follows that
Also, we have
where M > 0 is a constant and g(·) ∈ L2 (Ω). Then the following mapping A
defined by
Z
(Au, v) = [(∇u, ∇v) + f (x, u)v]dx for all u, v ∈ H01 (Ω)
Ω
|h(u)| ≤ δ|u| + γ
and f (·) ∈ L2 ((0, 2π) × (0, π)), where α > 0, β > 0, δ > 0, γ > 0 are constants.
Set
g− (u) = lim inf g(s), g+ (u) = lim sup g(s),
s→u s→u
N (u(t, x)) = {v(t, x) ∈ L2 ((0, 2π) × (0, π)) : v(t, x) ∈ Gu(t, x)}
then, since
lim sup(P (fj + h(uj )), uj − u0 ) ≥ 0
j→∞
Thus N + h is L-pseudomonotone.
Then we have
Br (0) = ∪∞
n=1 Mn .
By the Baire’s category theorem, there exists n0 such that Mn0 has nonempty
interior and so there exist z0 ∈ Br (0), r0 > 0 such that Br0 (z0 ) ⊂ Mn0 . Since
−z0 ∈ Br (0), there exists m > 0 such that
Therefore, we have
and so
2
kf k ≤ (n0 + m0 ).
r
This completes the proof.
∗
Proposition 6.1.26. Let P : E → 2E be a bounded pseudomonotone
mapping and {xn } ⊂ E be such that xn * x0 . If fn ∈ P xn such that fn * f0
and lim supn→∞ (fn , xn − x0 ) ≤ 0, then f0 ∈ P x0 and (fn , xn ) → (f0 , x0 ).
Proof. Since lim supn→∞ (fn , xn − x0 ) ≤ 0, we have
(f + P xn , x − xn ) ≥ 0
Kx = ΣN
i=1 αi (x)zi for all x ∈ C.
Obviously, we have
∩ni=1 WFi 6= ∅.
∗
If we denote by WF the weak closure of WF in E, then we have
∗
∩F ⊂E,dimF <∞ WF 6= ∅.
∗
Take (x0 , g0 ) ∈ ∩F ⊂E,dimF <∞ WF and, for any x ∈ D(T ) ∩ C, a finite dimen-
sional subspace F of E such that x, x0 ∈ F . Then there exists {(xj , P xj )} ⊂
WF such that
xj * x0 , P xj * g0 as j → ∞.
Therefore, we have
and so we have
lim sup(P xj , xj − x0 ) ≤ 0.
j→∞
Proof. We assume that both the spaces E and E ∗ are locally uniform
convex. If T is maximal monotone, then, for any p∗ ∈ E ∗ , J − p∗ is a
continuous bounded coercive monotone mapping and, thus, pseudomonotone.
By Lemma 6.1.27, there exists x0 ∈ E such that
(Jx0 − Jy0 , y0 − x0 ) ≥ 0.
lim Rλ x = x, lim Tλ x = f,
λ→0+ λ→0+
and
we have
lim kRλ x − xk = kRλ0 x − xk,
λ→0+
Therefore, {fj1 } and {fj2 } have subsequences {fj1k } and {fj2k } that converge
weakly to f 1 ∈ T1 x0 and f 2 ∈ T2 x0 , respectively. Therefore, we have
tjk fj1k + (1 − tjk )fj2k * t0 f 1 + (1 − t)f 2 ∈ t0 T1 x0 + (1 − t0 )T2 x0 .
This completes the proof.
u 0 ∈ F0 ∩ Ω ∩ Ω 1
and so
deg(T ∗ , Ω, 0) = deg(T ∗ , Ω ∩ Ω1 , 0)
= deg(T1 , Ω ∩ Ω1 , 0)
= deg(T, Ω1 , 0)
= deg(T0 , Ω0 , 0).
Therefore, we get
deg(T ∗ , Ω, 0) 6= deg(T, Ω, 0). (6.2.1)
Let
It follows from (6.2.1) that there exist t1 ∈ [0, 1] and u1 ∈ ∂Ω such that
0 ∈ Ht1 u1 . Hence there exists f1 ∈ T u1 such that
(f1 , v) = 0, (f1 , u1 ) = 0.
(3) If Ω1 and Ω2 ⊂ Ω are two open subsets with Ω = Ω1 ∪Ω2 and Ω1 ∩Ω2 = φ,
then
deg(T, Ω, 0) = deg(T, Ω1 , 0) + deg(T, Ω2 , 0);
0∈
/ (Mλ + T )(∂Ω) for all λ < λ0 ,
Mλj xj + fj = 0.
lim sup(Mλj xj , xj − x0 ) ≤ 0.
j→∞
Hence
lim sup(Mλj xj , xj ) ≤ (−f0 , x0 ). (6.3.1)
j→∞
and so we have
lim sup(fj , xj − x0 ) ≤ 0.
j→∞
Now, we define
Theorem 6.3.2. The topological degree defined by (6.3.3) has the follow-
ing properties:
(3) If Ω1 and Ω2 ⊂ Ω are two open subsets with Ω = Ω1 ∪Ω2 and Ω1 ∩Ω2 = φ,
then
(j fj − i fi , xj − xi ) ≤ 0 for i, j = 1, 2, · · · .
By letting i → ∞, we get
j (fj , xj − x0 ) ≤ 0.
Theorem 6.3.5. The topological degree defined by (6.3.4) has the follow-
ing properties:
(1) deg(J, Ω, 0) = 1 if and only if 0 ∈ J(Ω);
(2) If deg(M, Ω, 0) 6= 0, then 0 ∈ M x has a solution in Ω;
(3) If Ω1 and Ω2 ⊂ Ω are two open subsets with Ω = Ω1 ∪Ω2 and Ω1 ∩Ω2 = ∅,
then
deg(M, Ω, 0) = deg(M, Ω1 , 0) + deg(M, Ω2 , 0);
(4) If M1 , M2 are two maximal monotone mappings, Ω∩D(M1 )∩D(M2 ) 6= ∅
and
0∈
/ (tM1 +(1−t)M2 )(∂Ω∩D(M1 )∩D(M2 ))∪(tM1,λ +(1−t)M2,λ )(∂Ω)
for all t ∈ [0, 1] and λ ∈ (0, λ0 ), then
deg(M1 , Ω, 0) = deg(M2 , Ω, 0).
Proof. The proof follows from (6.3.4), (6.3.3), and Theorem 6.2.4.
Remark. A degree theory can also be developed for pseudomonotone map-
pings and generalized pseudomonotone mappings and their perturbations with
a maximal monotone mappings by using Proposition 6.1.31, which is a method
similar to the one employed above, and so we leave it to the reader as an ex-
ercise.
0∈
/ Tn (∂Ω ∩ D(T ) ∩ Fn ) for all n > N,
where Tn = jF∗ n T .
Proof. Suppose that the conclusion is not true. Then there exists xnk ∈
∂Ω ∩ D(T ) ∩ Fnk such that 0 ∈ Tnk xnk , i.e., there exists fnk ∈ T xnk such
that 0 = jF∗ n fnk for k = 1, 2, · · · . Without loss of generality, we may assume
k
that xnk * x0 . Now we have (fnk , x) = 0 for all x ∈ Fnk , k = 1, 2, · · · .
Thus (fnk , xnk ) = 0 and limk→∞ (fnk , v) = 0 for all v ∈ ∪∞ j=1 Fj . Since T is
a mapping of class (S+ )0,L , it follows that xnk → x0 ∈ ∂Ω ∩ D and 0 ∈ T x0 ,
which is a contradiction. This completes the proof.
Under the conditions of Lemma 6.4.1, we know from Section 2.3 that the
topological degree deg(Tn , Ω ∩ D(T ) ∩ Fn , 0) is well defined for sufficiently
large n, and we have the following:
Lemma 6.4.2. Let T be the same as in Lemma 6.4.1. Then there exists
an integer N > 0 such that the topological degree deg(Tn , Ω ∩ D(T ) ∩ Fn , 0)
does not depend on n > N , where Tn = jF∗ n T .
Proof. Suppose that the conclusion is not true. By Lemma 6.2.1, there
exist xnk ∈ ∂Ω∩D∩Fnk , fnk ∈ T xnk such that (fnk , xnk ) ≤ 0 and (fnk , x) = 0
for all x ∈ Fnk , k = 1, 2, · · · . We may assume that xnk * x0 . By the same
proof as in Lemma 6.4.2, we get xnk → x0 ∈ ∂Ω ∩ D and 0 ∈ T x0 , which is a
contradiction. This completes the proof.
Suppose that {Ej }∞
j=1 is another sequence of finite dimensional subspaces
∞
of L and ∪j=1 Ej = E. Then we have the following:
Lemma 6.4.3. Let T be the same as in Lemma 6.4.1. Then there exists
an integer N > 0 such that
and
Therefore, the conclusion of Lemma 6.4.3 is true. This completes the proof.
Now, let L be a dense subspace of E, Ω ⊂ E be a nonempty open bounded
∗
subset and T : D(T ) ⊂ E → 2E be a mapping of class (S+ )0,L . Assume
that Ω ∩ D(T ) ∩ F is open in F for each finite dimensional subspace F of L.
Suppose that 0 6∈ T (∂Ω ∩ D(T )). In view of lemmas 6.4.1 to 6.4.3, we may
define the topological degree by
(3) If Ω1 and Ω2 ⊂ Ω are two open subsets with Ω = Ω1 ∪Ω2 and Ω1 ∩Ω2 = ∅,
then
(4) If {Tt }t∈[0,1] is a homotopy of mappings of class (S+ )0,L with D(Tt ) = D
and 0 ∈ / Tt (∂Ω ∩ D) for all t ∈ [0, 1], then Deg(Tt , Ω ∩ D, 0) does not
depend on t ∈ [0, 1].
Proof. (1) to (3) follow easily from the definition and the properties of
degree theory in finite dimensional spaces.
We only need to prove (4). Assume that (Fj )∞ j=1 is a sequence of finite
dimensional subspaces of L with ∪∞ F
j=1 j = E. We claim that there exists an
integer N > 0 such that
0∈
/ Tt,n (∂Ω ∩ D ∩ Fn ) for all n > N, t ∈ [0, 1],
where Tt,n = jF∗ n Tt . If not, then there exist tnk → t0 , xnk ∈ ∂Ω ∩ D ∩ Fnk
with xnk * x0 , fnk ∈ Ttnk xnk such that 0 = jF∗ n fnk , which implies that
k
Since {Tt }t∈[0,1] is a homotopy of mappings of class (S+ )0,L , we get xnk →
x0 ∈ ∂Ω ∩ D and 0 ∈ Tt0 x0 , which is a contradiction. Thus the claim is true.
Now, for all n > N , we know from Section 2.3 that deg(Tt,n , Ω∩D∩L∩Fn , 0)
is a constant for t ∈ [0, 1], where Tt,n = jF∗ n Tt . In view of Lemma 6.4.3, we
see that the conclusion of (4) is true. This completes the proof.
∗
Theorem 6.4.5. Let T : D(T ) ⊂ E → 2E be a mapping of class (S+ )0,L
and Ω ⊂ E be an open bounded subset such that Ω ∩ D(T ) ∩ F is open in F
for each finite dimensional subspace F of L. If 0 ∈ Ω ∩ D(T ) and (f, x) > 0
for all x ∈ ∂Ω ∩ D(T ) and f ∈ T x, then
Deg(T, Ω ∩ D(T ), 0) = 1.
Deg(T, Ω ∩ D(T ), 0) = 1.
0∈
/ [(tPn (I − T ) + (1 − t)Pn ](∂Ω ∩ D(T ) ∩ Fn
for all n > N and t ∈ [0, 1], where Pn : H → Fn is the projection. Assume
this is not true. Then there exist tj → t0 , xj ∈ ∂Ω∩D ∩Fnj with xj * x0 and
fj ∈ T xj such that PFnj (xj −tj fj ) = 0 for j = 1, 2, · · · , where PFnj : H → Fnj
is the projection. Thus it follows that (xj − tj fj , xj ) = 0. This and our
assumption imply that tj = 1, so we have
x0 − P x0 − JQf0 − KP Q f0 = 0.
Since
we must have
JQf0 = 0, x0 − P x0 − KP Q f0 = 0.
Therefore, we have
which is a contradiction to 0 ∈
/ (L − T )(∂Ω ∩ D(L)). This completes the proof.
Now, let L be a Fredholm mapping of index zero type, Ω ⊂ H be an open
bounded subset and T : Ω → 2H be a mapping of class L-(S+ ). Suppose that
0 ∈
/ (L − T )(∂Ω ∩ D(L)). By Lemma 6.5.1, we have 0 ∈ / [I − P − (JQ +
KP Q )T ](∂Ω). As a result, deg(I − P − (JQ + KP Q )T, Ω, 0) is well defined.
We define
0∈
/ (I − P − (JQ + KP Q (tT1 + (1 − t)T2 )))x for all x ∈ ∂Ω.
Moreover, J is linear, so
and so
degJ (L − T1 , Ω, 0) = degJ (L − T2 , Ω, 0).
This completes the proof.
Suppose that L is a Fredholm mapping of index zero type, Ω ⊂ H be an
open bounded subset, Ω∩D(L)∩D(T ) 6= ∅ and T : D(T ) → 2H is a L-maximal
monotone mapping. Also, suppose that 0 ∈
/ (L−T )(∂Ω∩D(L)∩D(T )). Then,
by Lemma 6.5.1, we have
0∈
/ [I − P − (JQ + KP Q )T ](∂Ω ∩ D(T )).
0∈
/ [I − P − (JQ + KP Q )T + I](∂Ω ∩ D(T )) for all ∈ (0, 0 ).
Proof. Assume that the conclusion is false. Then there exist n → 0 and
xn ∈ ∂Ω ∩ D(T ) such that
0 ∈ (I − P − (JQ + KP Q )T + n )xn .
0∈
/ [I − P − (JQ + KP Q )T + I](∂Ω) for all ∈ [0, 0 ).
Ω1 ∩ Ω2 = ∅, / (L − T )(D(L) ∩ Ω \ (Ω1 ∪ Ω2 ),
0∈
then we have
0∈
/ ∪t∈[0,1] [L − tT1 + (1 − t)T2 ](∂Ω ∩ D(L)),
then we have
deg(T, Ω, 0) = d(J, Ω, 0) = 1.
By (6.6.2), we have
for all x ∈ D(M ) and m ∈ M x. Thus, from (6.6.4) and (6.6.6), it follows that
deg(T + M, Ω, 0) = 1.
deg(I − T, Ω, 0) = deg(I, Ω, 0) = 1.
and denote
with the norm kukw = kukp + ku0 k∗,q , where u0 is the generalized derivative
of u and k · kp , k · k∗,q are norms of the spaces Lp (0, T ; E) and Lq (0, T ; E ∗ ),
respectively. Let W ∗ be the dual space of W 1,p (0, T ; E).
∗
Theorem 6.7.1. Let A(t) : E → 2E be an operator of class (S+ ) for all
t ∈ R. Suppose the following conditions are satisfied:
(1) For all u(t) ∈ W 1,p (0, T ; E), A(t)u(t) is E ∗ -measurable on [0, T ];
(2) There exist a constant C > 0 and C1 (·) ∈ Lq (0, T ) such that
kf k∗ ≤ Ckxkp−1 + C1 (t)
for almost everywhere t ∈ [0, T ] and A(t) is a mapping of class (S+ ) and so
(fnk (t)) has a subsequence converging weakly to f0 (t) ∈ A(t)u0 (t) for almost
everywhere t ∈ [0, T ]. Hence we have g0 ∈ Sn u0 .
Let uj ∈ W 1,p (0, T ; E) with uj * u0 in W 1,p (0, T ; E) and gj ∈ Sn uj such
that
limj→∞ (gj , uj − u0 ) ≤ 0.
Then, by (6.7.1), there exist fj ∈ Auj such that
Z T h 1 i
limj→∞ (fj , uj − u0 ) + (Juj , uj − u0 ) dt ≤ 0. (6.7.2)
0 n
Now, using the fact that A(t) is a mapping of class (S+ ), the conditions (2)
and (3), we have
limj→∞ (fj (t), uj (t) − u0 (t)) ≥ 0 (6.7.3)
for almost everywhere t ∈ [0, T ] and thus uj → u0 in W 1,p . There is a
subsequence (ujk ) such that ujk (t) → u0 (t) for almost everywhere t ∈ [0, T ].
Thus, condition (2) implies that
for almost everywhere t ∈ [0, T ] and A(t) is a mapping of class (S+ ) and so
(fjk ) has a subsequence (fjk0 ) such that fjk0 (t) * f0 (t) ∈ A(t)u0 (t). Hence,
(gjk0 ) converges weakly to g0 ∈ Sn u0 , where g0 satisfies
1
(g0 , v) = (u00 , v) + (f0 , v) + (Ju0 , v) for all v ∈ W 1,p (0, T ; E).
n
where B(0, r0n ) is the open ball with radius r0n in W 1,p (0, T ; E). Therefore,
0 ∈ Sn u has a solution un ∈ B(0, r0n ), i.e., there exists fn (t) ∈ A(t)un (t) for
almost everywhere t ∈ [0, T ] such that
1
(u0n , v) + (fn , v)dt + (Jun , v) = 0 for all v ∈ W 1,p (0, T ; E). (6.7.4)
n
Put v = un in (6.7.4). Then, by the condition (3), we know that there exists
N > 0 such that
1
kun kp ≤ N, √ ku0n k∗,q ≤ N for all n ≥ 1.
n
By the condition (2), u0n is bounded in Lq (0, T ; E ∗ ).
Now, we may assume that un * u0 in W 1,p (0, T ; E). Again, by the condi-
tion (2), we know that (fn ) is bounded in Lq (0, T ; E ∗ ), and so we may assume
that fn * f0 in Lq (0, T ; E ∗ ). Let v = un − u0 . It follows from (6.7.4) that
Z T
lim (fn (s), un (s) − u0 (s))ds = 0.
n→∞ 0
Since A(t) is a mapping of class (S+ ), it is easy to show that (fn (s), un (s) −
u0 (s)) → 0 in measure and hence there exists a subsequence ((fnk , unk − u0 ))
such that (fnk (s), unk (s) − u0 (s)) → 0 for almost everywhere s ∈ [0, T ]. Thus
unk (s) → u0 (s) for almost everywhere s ∈ [0, T ] and fnk converges weakly to
f0 (s) ∈ A(s)u(s). By letting nk → ∞ in (6.7.4), we get
(1) |ai (x, u, ξ)| ≤ L(|u|m1 + kξkm−1 ) + M (x) for 1 ≤ i ≤ mn, where L > 0,
m1 < N N −m , 2 ≤ m < N are constants and M (x) ∈ L
m−1 (Ω);
N
where µ > 0 and 0 ≤ r < N −2 are constants;
(3) Σni=1 [ai (x, u, ξ) − ai (x, u, η)](ξi − ηi ) ≥ kkξ − ηkm ;
(4) f (x, z) : Ω × R → R satisfy the following conditions:
(4a) There exists Ω0 ⊂ Ω with mes(Ω0 ) = 0 such that
[
Df = {z ∈ R : f (x, .) is discontinuous at z}
x6∈Ω0
It is known that F (x, z) is nonempty closed convex and compact for all x ∈ Ωc1
(see [58]). Moreover, F (x, z) is measurable in x and z → F (x, z) is upper
semicontinuous for almost all x ∈ Ω.
q
Next, we define a mapping Ψ : W01,m (Ω) → 2L (Ω) by
Put v(x) ≡ 0. Then, by (1) and (4c), there exists a constant M > 0 such that
Z
∂uj 2
lim sup ρ2 (uj )| | dx ≤ M. (6.7.6)
j→∞ Ω ∂x
Then u˜j ∈ W01,2 (Ω). We may assume that (u˜j ) converges strongly in Lp (Ω),
p < N2N−2 , to some u˜0 . Otherwise, take a subsequence. From this, we obtain
that u˜j converges in measure to u˜0 and ρ(u˜ 0 ). Consequently, we have
∂(uj − v) m
Z
k | | dx
Ω ∂x
∂u0 ∂v
≤ −M1 − (f0 , u0 − v) + ΣN 2
i=1 ρ (u0 ) dx
∂xi ∂xi
N Z (6.7.9)
X ∂v ∂(u0 − v)
− ai (x, u0 , ) dx ≤ −(f0 , u0 − v)
i=1 Ω
∂x ∂xi
N Z
X ∂u0 ∂v ∂(u0 − v)
− [ρ2 (u0 ) + ai (x, u0 , )] dx,
i=1 Ω
∂x i ∂x ∂xi
where f0 ∈ Ψu0 . By taking limits in (6.7.9), we know that (6.7.9) is also true
for all v ∈ W01,q (Ω) for q = 2[1 − NN−2 r]−1 , where r is the same as in the
assumption (2).
Now, consider the following functional on W01,q (Ω) defined by
Z
∂u0 ∂φ
h(φ) = ΣN
i=1 ρ2 (u0 ) dx. (6.7.10)
Ω ∂xi ∂xi
u0 (x) u0 (x) 0
ρ2 (u0 ) = ∈ Lm (Ω). (6.7.13)
∂xi ∂xi
Therefore u0 ∈ D(A). Notice that (6.6.9) is also true for all v ∈ W01,m (Ω). Put
v(x) = u0 (x) + tw(x) for all t > 0 and w ∈ W01,m (Ω) in (6.7.9). Then, divide
the resulting inequality by t and pass to the limit. We obtain 0 ∈ Au0 + Ψu0 .
Finally, we prove that uj → u0 ∈ W01,m (Ω). We may assume that
uj (x) 0
ai (x, uj (x), ) * bi (x) in Lm (Ω).
∂x
1,m
Take u0j ∈ ∪∞ 0
j=1 Fj such that uj → u0 in W0 (Ω). By using the assumption
(3), we have
∂(uj − u0j ) m
Z
k | | dx
Ω ∂x
∂u0j ∂uj ∂u0j
Z
∂uj
≤ ΣNi=1 [ai (x, u j , ) − ai (x, u j , )]( − )dx
Ω ∂x ∂x ∂x ∂x
Z
∂uj 2
= (gj , uj ) − ρ2 (uj )| | dx − (fj , uj ) (6.7.14)
Ω ∂x
∂uj ∂u0j
Z
− ΣN i=1 ai (x, uj , ) dx
Ω ∂x ∂x
∂u0j ∂(uj − u0j )
Z
− ΣN i=1 ai (x, uj , ) dx.
Ω ∂x ∂x
∂(uj − u0j ) m
Z
k | | dx
Ω ∂x
Z Z (6.7.15)
2 ∂u0 2 N ∂u0 (x)
≤ −(f0 , u0 ) − ρ (u0 )| | dx − Σi=1 hi (x) dx.
Ω ∂x Ω ∂xi
for all v ∈ ∪∞
j=1 Fj . From (6.7.16) and (6.7.15), it follows that uj → u0 ∈
W01,m (Ω). This completes the proof.
From Theorem 6.7.4, we know that, if for some open bounded subset U ⊂
W01,m (Ω) and 0 ∈/ Au+Ψu for all u ∈ ∂U ∩D(A), then deg(A+Ψ, U ∩D(A), 0)
is well defined.
6.8 Exercises
∗
1. Let A : D(A) ⊆ E → 2E be a set-valued mapping of class (S+ ) and
∗
P : E → 2E be a bounded pseudomonotone mapping. Show that P +A
is a mapping of class (S+ ).
∗
2. Let A : D(A) ⊆ E → 2E be a multi-valued mapping of class (S+ ),
∗
T : D(A) → 2E be an upper semicontinuous operator with closed
convex values and T maps each bounded subset of D(A) into a relatively
compact subset of E ∗ . Show that T + A is a multi-valued mapping of
class (S+ ).
∗
3. Let A : D(A) ⊆ E → 2E be a multi-valued mapping of class (S+ ) and
∗
M : D(M ) ⊆ E → 2E be a maximal monotone operator. Show that
Mλ + A is a multi-valued mapping of class (S+ ).
∗
4. Let M : D(M ) ⊆ E → 2E be a maximal monotone operator and
∗
P : E → 2E be a bounded pseudomonotone mapping. Suppose that
(g, x − x0 )
lim = +∞ for some x0 ∈ D(M ) and all g ∈ P x.
kxk→∞ kxk
169
Proof. (1) Suppose that P is not quasinormal. Then, for any y ∈ P with
kyk = 1, there exist xn ∈ P such that
z2n = Σ2n
i=1 (xi + yi ), z2n+1 = z2n + x2n+1 ,
zn = Σni=1 xi ≤ Σ∞
i=1 (xi + yi ) ∈ P
Note that z−y ≤ x0 −x and we have φ(x) ≤ σ(P )kx0 −xk. Obviously, if x ≤ y,
then φ(x) ≥ φ(y). We claim that inf x∈D φ(x) = 0. If not, inf x∈D φ(x) = δ >
0, there exists x1 ∈ D such that φ(x1 ) > δ, so there are x1 ≤ y1 ≤ z1 such
that kz1 − y1 k > δ. Again, since φ(z1 ) > δ, there are z1 ≤ y2 ≤ z2 such that
kz2 − y2 k > δ. Repeat the above process, we get a sequence (yn ) in D such
that
y1 ≤ z1 ≤ y2 ≤ z2 ≤ · · · ≤ x0
with kzi − yi k > δ for i = 1, 2, · · · , so it does not converge, which contradicts
the fact that P is regular. Thus there exist (xn )∞ n=1 ⊂ D such that φ(xn ) → 0.
Since D is well ordered, put yn = max{x1 , x2 , · · · , xn }. Then y1 ≤ y2 ≤ · · · ≤
x0 and so φ(yn ) → 0. Since P is regular, there exists y0 ∈ E such that
yn → y0 . It is easy to see that y0 = sup{yi : i = 1, 2, · · · }.
Next, we prove that y0 = sup D. To see this, we only need to prove that y0
is an upper bound of D. For any x ∈ D, we have two cases: (1) x ≤ yn for
some n and (2) yn ≤ x.
In case (1), we have x ≤ y0 .
In case (2), we have y0 ≤ x and 0 ≤ x − y0 ≤ x − yn and thus
kx − y0 k ≤ N kyn − xk ≤ N φ(yn ) → 0 as n → ∞.
that sup D = z = (z1 , z2 , · · · ), where zi = supx∈D x(i) and x(i) is the i-th
coordinate of x for i = 1, 2 · · · . Thus, P is strongly minihedral. For any
integer n ≥ 1, let xn ∈ l∞ be defined by
(
n 1, if i ≤ n,
x (i) =
0, if i > n.
(1) P is reproducing;
(2) There exists r > 0 such that, for any x ∈ E, there exist y, z ∈ P with
kyk ≤ rkxk, kzk ≤ rkxk and x = y − z;
Proof. It is obvious that (2) is equivalent to (3), (2) implies (1), and (3)
implies (4), so we only need to prove (1) implies (2) and (4) implies (3).
(1) ⇒ (2) Since P is reproducing, we have E = ∪∞ n=1 En , where
β
ky − y1 − y2 − · · · − yn k < for n = 1, 2, · · · .
2n+1
Thus, we have
∞ 1 1
y = σi=1 yi , yn = xn − zn , xn , zn ∈ P, kxn k ≤ n
, kzn k ≤ n
2 2
for n = 1, 2, · · · . Put x = Σ∞ ∞
i=1 xi , z = Σi=1 zi . Then x, z ∈ P , kx|| ≤ 1,
β
kzk ≤ 1 and x ∈ C. Thus, 2 B ⊂ C. This completes the proof.
Ik = {n : kn ≤ k} ⊂ Ik+1 ∪ {1, 2, · · · , k}
Hence there exists yn ∈ conv((T (Ckn −1 ∩ B)) such that kxn − yn k < n−1 . In
particular, we can find some finite An ⊂ Ckn −1 ∩B such that yn ∈ conv(T An ).
Now, if we put A = ∪∞ n=1 An , then A is the required subset. To see this,
we have to check that A ∈ FB . Since C1 ⊃ C2 ⊃ · · · , we have An ⊂ Ci for
i ≤ kn − 1, and thus we have
A \ Cn = ∪∞
i=1 (Ai \ Cn ) = ∪i;n>ki −1 (Ai \ Cn ) ⊂ ∪i∈In Ai ,
(1) ind(x0 , Ω ∩ P ) = 1 if x0 ∈ Ω ∩ P ;
where r : E → P is a retraction. One can check easily this definition does not
depend on r and we have the following properties:
Theorem 7.2.3. The fixed point index ind(T, Ω∩P ) satisfies the following
properties:
(1) ind(x0 , Ω ∩ P ) = 1 if x0 ∈ Ω ∩ P ;
(2) If ind(T, Ω ∩ P ) 6= 0, then x = T x has a solution in Ω;
(3) If Ωi ⊂ Ω for i = 1, 2, Ω1 ∩ Ω2 = ∅ and 0 ∈
/ (I − T )[(Ω \ (Ω1 ∪ Ω2 )) ∩ P ],
then
ind(T, Ω ∩ P ) = ind(T, Ω1 ∩ P ) + ind(T, Ω2 ∩ P );
d(0, C) ≥ 0 .
0 = Σni=1 αi yi ,
0 6∈ Conv(T ∂Ω ∩ P )).
ind(T, Ω \ Ω1 ∩ P ) = 1.
For any m > 0, we claim that x − T x 6= tmy for all x ∈ ∂Ω ∩ P and t ∈ [0, 1].
If not, then there exist x0 ∈ ∂Ω ∩ P , t0 ∈ [0, 1] such that x0 = T x0 + t0 my ,
and so we have
Hence we get
kT x0 k 1
≤ ,
kx0 k σ−
which contradicts (7.3.1). By the homotopy property (4) of Theorem 7.2.3,
we get
ind(T, Ω ∩ P ) = ind(T + m0 y , Ω ∩ P )
and thus, we must have
ind(T + my , Ω ∩ P ) = 0
(since Ω is bounded and T (E) is bounded for any countably bounded subset
E of Ω ∩ P ). Consequently, ind(T, Ω ∩ P ) = 0. This completes the proof.
Corollary 7.3.8. Let X be a real Banach space, Ω be an open bounded
subset of X with 0 ∈ Ω and P be a quasinormal cone in X. If T : Ω ∩ P → P
is a continuous countably k-set contraction and the following conditions are
satisfied:
then ind(T, Ω ∩ P ) = 0.
Proof. If we put T1 x = µT x for all x ∈ Ω ∩ P, then T1 is countably
condensing and T1 satisfies all the conditions of Theorem 7.3.7. Thus, we
have
ind(T1 , Ω ∩ P ) = 0.
From the assumption (1) and the homotopy property (4) of Theorem 7.2.3,
we know that
ind(T, Ω ∩ P ) = ind(µT, Ω ∩ P ).
Therefore, we have ind(T, Ω ∩ P ) = 0. This completes the proof.
Corollary 7.3.9. Let X be a real Banach space, Ω be an open bounded
subset of X with 0 ∈ Ω and P be a quasinormal cone in X. If T : Ω ∩ P →
P is a continuous countably k-set contraction with k < σ, where σ is the
quasinormality constant of P and the following conditions are satisfied:
then ind(T, Ω ∩ P ) = 0.
1
Proof. We take 0 > 0 such that k + 0 < min{1, σ} and put T1 = k+ 0
T.
k
Then T1 is a countably k+0 -contraction, and we know, by the assumption
(2), that
n kT xk o 1 1
1
inf : x ∈ ∂Ω ∩ P ≥ > for all x ∈ ∂Ω ∩ P.
kxk k + 0 σ
ind(T, Ω1 ∩ P ) = 1.
ind(T, Ω1 ∩ P ) = 1.
On the other hand, we have, by the assumption (2) and Corollary 7.3.9,
that ind(T, Ω2 ∩ P ) = 0 and so
0∈
/ Conv(S(∂Ω ∩ P )).
Proof. Assume that (II) is not true. Then, by Theorem 7.3.13, we have
ind(T, Ω2 ∩ P ) = 0. On the other hand, by the assumption (2), we have
ind(T, Ω1 ∩ P ) = 1 and thus,
ind(T, Ω1 ∩ P ) = 1. (7.3.2)
ind(T, Ω2 ∩ P ) = 0. (7.3.3)
From the assumption (1), we must have t0 = 1, which contradicts the fact
that −A + T has no fixed point on ∂Ω1 ∩ P . Thus we have
ind(T (I + A)−1 , U2 \ U1 ) = −1
(1) ind(x0 , Ω ∩ P ) = 1 if x0 ∈ Ω ∩ P ;
(2) If ind(T, Ω ∩ K) 6= 0, then T x = x has a solution in Ω ∩ K;
(3) If Ω1 , Ω2 are two open bounded disjoint subsets, then ind(T, (Ω1 ∪ Ω2 ) ∩
P ) = ind(T, Ω1 ∩ P ) + ind(T, Ω2 ∩ P );
(4) Let H(t, x) : [0, 1] × Ω ∩ P → E be a continuous compact mapping
satisfying H(t, x) 6= x for all (t, x) ∈ [0, 1] × ∂Ω ∩ P and d(x, H(t, x)) 6=
d(H(t, x), P ) for all (t, x) ∈ Ω∩P with H(t, x) ∈ / P . Then ind(H(t, ·), Ω∩
P ) does not depend on t ∈ [0, 1].
Remark. For more details about the results in this section, we refer the
reader to [182].
(3) f (t, s) ≥ αsγ for all s ∈ [0, 0 ), where 0 > 0, α > 0 and 0 < γ < 1 are
constants.
Then (E 7.6.2) has a nontrivial non-negative C 2 solution.
Proof. It is well known that (E 7.6.2) is equivalent to the following integral
equation: Z 1
x(t) = G(t, s)f (s, x(s))ds, (E 7.6.3)
0
where G(t, s) is the Green function defined by
(
1
(t + 1)(2 − s), t ≤ s,
G(t, s) = 13
3 (s + 1)(2 − t), t > s.
One may easily see that M = max{G(t, s) : (t, s) ∈ [0, 1]×[0, 1]} < 34 and f, G
satisfy the conditions of Example 7.6.1. Therefore, (E 7.6.3) has a nontrivial
non-negative solution in C([0, 1]), i.e., (E 7.6.1) has a nontrivial non-negative
C 2 solution.
7.7 Exercises
1. Let Ω ⊂ Rn be measurable subset with m(Ω) < +∞, 1 ≤ p ≤ ∞ and
P ⊂ Lp (Ω) be given by P = {f (·) ∈ Lp (Ω) : f (x) ≥ 0, almost all x ∈
Ω}. Show that P is fully regular.
11. Let a(s, t) : [0, 1] × [0, 1] → [0, +∞) be a continuous function satisfying
the following conditions:
(1) [a(s, t) − a(s, r)](t − r) ≥ 0;
(2) a(s, t) ≤ ct for some constant c > 0 and all (s, t) ∈ [0, 1] × [0, 1].
Let k(s, t) : [0, 1] × [0, 1] → [0, +∞) be a continuous function such that
k(s, t) ≤ αt + β for all (s, t) ∈ [0, 1] × [0, 1], where α ∈ (0, 1), β > 0 are
constant. Show that the following integral inequation:
Z 1
x(t) + a(t, x(t)) − k(t, x(s))ds = 0 for all t ∈ [0, 1]
0
has two non-negative nontrivial solutions in C 2 ([0, 1]) by using the Green
function G(t, s) defined as
(
e−1 (at + b)[c(1 − s) + d], t ≤ s,
G(t, s) = −1
e (as + b)[(c(1 − t) + d], t > s.
195
19. V. Benci, A geometrical index for the group S 1 and some applications to
the study of periodic solutions of ordinary differential equations, Comm.
Pure Appl. Math. 34(1981), 393–432.
23. K. Borsuk, Drei Sätze über die n-dimensional Euklidische Sphäre, Fund.
Math. 20(1933), 177–190.
27. H. Brézis, Degree theory: old and new, in Topological Nonlinear Analy-
sis II: Degree, Singularity and Variations (M. Matzeu and A. Vignoli,
Ed.) Birkhauser, 1997, 87–108.
49. K. C. Chang, Critical Point Theory and Its Applications, Shanghai Sci-
ence and Technology Publishing House, Shanghai, 1986.
51. S. S. Chang and Y. Q. Chen , Degree theory for multivalued (S) type
mappings and fixed point theorems, Appl. Math. Mech. 11 (1990),
441–454.
66. Y, Q, Chen and J. K. Kim, Anti-periodic solutions for first order semi-
lnear evolution equations, Non. Funct. Anal. Appl. 8(2003), 49–57.
73. C. Conley and E. Zehnder, A Morse type index theory for flows and
periodic solutions to Hamiltonian systems, Comm. Pure Appl. Math.
37(1984), 207–253.
74. M. G. Crandall and T. Liggett, Generation of semigroups of nonlinear
transformations on general Banach spaces, Amer. J. Math. 93(1971),
265–298.
75. M. G. Crandall, H. Ishii and P. L. Lions, User’s guide to viscosity solu-
tions of second order partial differential equations, Bull. Amer. Math.
Soc. 27(1992), 1–67.
76. J. Cronin, Fixed Points and Topological Degree in Nonlinear Analysis,
Math. Surveys, No. 11, Amer. Math. Soc., 1964.
77. S. J. Daher, On a fixed point principle of Sadovskii, Nonlinear Anal.
2(1978), 643–645.
78. S. J. Daher, Fixed point theorems for nonlinear operators with a se-
quential condition, Nonlinear Anal. 3(1979), 59–63.
79. E. N. Dancer, On the indices of fixed points in cones and applications,
J. Math. Anal. Appl. 91(1983), 131–151.
80. E. N. Dancer, A new degree for S 1 -invariant gradient mappings and
applications, Anal. Nonlinear 2(1985), 329–370.
81. E. N. Dancer, Degree theory on convex sets and applications to bifurca-
tion, Calculus of variations and partial differential equations, Topics on
geometrical evolution problems and degree theory, Springer, 185–241,
2000.
82. E. N. Dancer, R. D. Nausbaum and C. A. Stuart, Quasinormal cones in
Banach spaces, Nonlinear Anal. 7(1983), 539–553.
83. E. N. Dancer and J. Toland, Degree theory for orbits of prescribed period
for flows with a first integral, Proc. London Math. Soc. 60(1990), 549–
580.
84. B. Decorogna, Weak Continuity and Weak Lower Semicontinuity of Non-
linear Functionals, Lect. Notes Math. 922, Springer, Berlin, 1982.
85. H. Debrunner and P. Flor, Ein Erweiterungssatz fur monotone Mengen,
Arch. Math. 15(1964), 445–447.
86. K. Deimling, Nonlinear Functional Analysis, Springer-Verlag, Berlin,
1985.
87. K. Deimling, Multivalued Differential Equations, W. De Gruyter, Berlin,
1992.
95. Z. Dzedzej, Fixed point index theory for a class of nonacyclic multivalued
maps, Dissert. Math. 253(1985), 1–53.
106. E. R. Fadell, Recent results in the fixed point theory of continuous maps,
Bull. Amer. Math. Soc. 76(1970), 10–29.
112. K. Fan, Some properties of convex sets related to fixed point theorems,
Math. Ann. 266(1984), 519–537.
132. E. Getzler, Degree theory for Wiener maps, J. Func. Anal. 68 (1986),
388–403.
133. E. Getler, The degree of the Nicolai map, J. Func. Anal. 74(1987),
121–138.
139. A. Granas, The Leray-Schauder index and the fixed point theory for
arbitrary ANR’s, Bull. Soc. Math. France 100(1972), 209–228.
141. D. J. Guo, A new fixed point theorem, Acta Math. Sinica 24(1981),
444–450.
147. F. B. Hang and F. H. Lin, Topology of Sobolev mappings II, Acta Math.
191(2003), 55–107.
150. M. Hazewinkel and M. van de Vel, On almost fixed point theory, Canad.
J. Math. 30(1978), 673–699.
155. S. Hu and Y. Sun, Fixed point index for weakly inward mappings, J.
Math. Anal. Appl. 172(1993), 266–273.
159. I. Ize, I. Massabo and A. Vignoli, Degree theory for equivariant maps,
315(1989), 433–510.
214. J. Milnor, Analytic proof of the hairy ball theorem and the Brouwer
fixed point theorem, Amer. Math. Monthly 85(1978), 521–524.
221. H. Mönch and G. F. von Harten, On the Cauchy problems for ordinary
differential equations in Bnach spaces, Arch. Math. 39(1982), 153–160.
239. R. S. Palais, Critical point theory and the minimax principle, Proc.
Symp. Pure Math. 15(1970), 185–212.
257. T. Riedrich, Der Raum S(0, 1) ist zulässig, Wiss Z. TU Dresden 13(1964),
1–6.
258. J. W. Roberts, A compact convex set with no extreme points, Studia
Math. 60(1977), 255–266.
259. R. T. Rockafellar, Characterization of the subdifferential of convex func-
tions, Pacific J. Math. 17(1966), 497–510.
260. R. T. Rockafellar, Convex Analysis, Princeton University Press, Prince-
ton, 1970.
261. R. T. Rockafellar, On the maximal monotonicity of subdifferential map-
pings, Pacific J. Math. 33(1970), 209–216.
262. R. T. Rockafellar, Local boundedness of nonlinear monotone operators,
Michigan Math. J. 16(1970), 397–407.
263. S. Rolewicz, Metric Linear Spaces, Polish Scientific Publishers, Warszawa,
1972.
264. E. H. Rothe, A relation between the type numbers of a critical point and
the index of the corresponding field of gradient vectors, Math. Nachr.
4(1950-51), 12–17.
265. E. H. Rothe, On the Cesari index and the Browder-Petryshyn degree, in
International Symposium on Dynamic Systems, pp. 295–312, Academic
Press, New York, 1977.
266. E. H. Rothe, Introduction to Various Aspects of Degree Theory in Ba-
nach Spaces, Amer. Math. Soc., Providence, RI, 1986.
267. W. Rudin, Functional Analysis, MacGraw-Hill, New York, 1973.
268. K. P. Rybakowski, On the homotopy index for the infinite-dimensional
semiflows, Trans. Amer. Math. Soc. 269(1982), 351–382.
269. K. P. Rybakowski and E. Zehnder, On a Morse equation in Conley’s
index theory for semiflows on metric spaces, Ergodic Theory Dyn. Syst.
5(1985), 123–143.
270. B. Rzepecki, Remarks on Schauder’s fixed point principle and its appli-
cations, Bull. Acad. Polon. Sci. Sér. Sci. Math. 27(1979), 473–480.
271. B. N. Sadovskii, Limit-compact and condensing operators, Uspekhi Mat.
Nauk. 27(1972), 81–146.
272. S. S. Schaefer, Topological Vector Spaces, Macmillan, New York, 1966.
273. J. Schauder, Der Fixpunktsatz in Funktionalraumen, Studia Math. 2(19
30), 171–180.
284. Y, Sun and J. X. Sun, Multiple positive fixed points of weakly inward
mappings, J. Math. Anal. Appl. 148(1990), 431–439.
291. M. Vath, Fixed point theorems and fixed point index for countably
condensing maps, Topological Methods in Nonlinear Anal. 13(1999),
341–363.
302. M. Willem, On a result of Rothe about the Cesari index and A-proper
mappings, Boll. Un. Mat. Ital. 17(1980), 178–182.