V22003 4th Semester Project Source File
V22003 4th Semester Project Source File
V22003 4th Semester Project Source File
to be submitted by
5 Applications to ODEs 17
5.1 Solvability result for a type of differential equations in Banach spaces . . . 17
5.2 Solvability result for a type of differential equation in Hilbert spaces . . . . 19
1
1 Introduction
Degree theory, a fundamental concept in mathematics, plays a crucial role in various
branches of the field. This report delves into the extensive applications of the Brouwer
degree, exploring its significance in different mathematical contexts. The Brouwer degree, a
key tool in algebraic topology, is instrumental in understanding and analyzing a wide range
of theorems and concepts.
The report begins by examining several important theorems that utilize the Brouwer
degree, including Borsuk’s theorem, the Borsuk-Ulam theorem, the Jordan separation the-
orem, the Open mapping theorem, and the connection between the winding number and
degree. These theorems highlight the versatility and applicability of the Brouwer degree in
diverse mathematical scenarios.
Moving forward, the report delves into Degree Theory for Condensing Operators, focus-
ing on essential topics such as the Leray-Schauder degree, uniqueness of the Leray-Schauder
degree, condensing and countably condensing map, and the Housdorff metric. These dis-
cussions provide a deeper insight into the properties and applications of degree theory in
operator theory and functional analysis.
Furthermore, an exploration of Degree Theory for Countably Condensing Mapping is
undertaken, elucidating the construction of the degree for countably condensing mappings,
properties associated with this theory, and fixed point theorems involving countably con-
densing mappings. This section aims to showcase the broad scope and implications of degree
theory in the context of countably condensing mappings.
In addition to theoretical discussions, this report investigates the practical applications
of degree theory in Ordinary Differential Equations (ODEs). By presenting solvability
results for specific types of differential equations in normed space which is also complete
spaces and Hilbert spaces.
Proof. SSince ψ Ȳ \ ∞
S
i=1 Yj is closed, there exists a neighborhood N of x0 disjoint from
ψ Ȳ \ ∞ i=1 Yj , and let y be a regular value in N ; then
2
deg (ψ, X, y0 ) = deg (ψ, X\C, y0 ) .
Proof. WLOG, suppose that ψ ∈ C 1 (M̄ ) with Jψ (0) ̸= 0 because we know C 1 (M̄ ) is dense
C(M̄ ). Next, we define a mapping ζ ∈ C 1 (M̄ ) sufficiently close to ψ by induction as follows:
Let γ ∈ C 1 (R) be an odd mapping with γ ′ (0) = 0 and γ(t) = 0 if and only if t = 0. Put
ψ(z)
Mk = {z ∈ M : zk ̸= 0} and φ(z) = γ(z 1)
for all z ∈ M1 . Choose |p1 | in such a way that
p1 becomes regular value for φ on M1 . Put ζ1 (z) = ψ(z) − γ (z1 ) p1 , then 0 will be regular
value for ζ1 on M1 .
Assume that we already have an odd map ζk ∈ C 1 (M̄ ) close to ψ such that it admits
0 as a regular value for ζk on Mk . Then we set ζk+1 (z) = ζk (z) − γ (zk+1 ) pk+1 with |pk+1 |
suitably small such that 0 is regular value for ζk+1 on Mk+1 .
If z ∈ Mk+1 and zk+1 = 0, then
′
z ∈ Mk , ζk+1 (z) = ζk (z), ζk+1 (z) = ζk′ (z),
and hence Jζk+1 (z) ̸= 0. By induction, we also have ζn′ (0) = ζ1′ (0) = ψ ′ (0) and so 0 is a
regular value for ζn . By Definition of degree and gn is sufficiently close to ψ, we know that
X
deg(ψ, M, 0) = deg (ζn , M, 0) = sgn Jζn (0) + sgn Jζn (z),
z∈ζ −1 (0),z̸=0
Proof. Given a function γ(x) = ζ(x) − ζ(−x), where γ is an odd function, we seek to
prove the existence of an x such that γ(x) = 0. If this is not true, we can normalize γ(x) by
dividing it by its absolute value to obtain a new function γ : S n → S n−1 , which also remains
odd. The restriction of this new γ to the equator S n−1 then has an odd degree according to
Borsuk’s Theorem 2.1.1. However, this restriction is homotopically trivial when restricted
3
to one of the hemispheres bounded by S n−1 , leading to a contradiction, thereby concluding
the theorem.
Let θ ⊂ Rn be a bounded open set and let ψ ∈ C(θ̄). Due to the homotopy invariance
property of the degree of mapping deg(ψ, θ, y), we can conclude that this degree is an integer
that remains constant as y varies within the connected component O of Rn \ ψ(∂θ). Hence,
it is appropriate to represent this integer as deg(ψ, θ, O). The connected component that is
unbounded is symbolized as O∞ . Consequently, we obtain the following formula:
Theorem 2.2.2. [2] Let θ ⊂ Rn be a bounded subsets which are open as well, ψ ∈ C(θ),
ϕ ∈ C(Rn ), and let Oi be the connected bounded subsets of Rn \ψ(∂θ). If q ∈
/ (gf )(∂θ), then
X
deg(ϕψ, θ, q) = deg (ψ, θ, Oj ) deg (ϕ, Oj , q) , (2.2.1)
j
Proof. We initially demonstrate that 2.2.1 contains only a finite number of non-zero terms..
Let r > 0 be chosen such that the image of θ̄ under ψ is contained in the ball Br (0).
Consequently, we can deduce that the set Γ = Br (0) ∩ ϕ−1 (q) is compact. Moreover, Γ is a
subset of Rn \ψ(∂θ), which can be written as as the union of sets Oi for i ≥ 1. There exist a
finite number of indices i = 1, 2, · · · , t such that the union of Oi for i = 1 to t + 1 contains
Γ, where Ot+1 = O∞ ∩ Br+1 .. We have
and note
4
X
sgnϕ (z) sgnψ (r)
r∈ψ −1 (z),z∈ϕ−1 (q)
X X
= sgnϕ (z) sgn Jψ (r)
z∈ϕ−1 (q),z∈ψ(θ) r∈ψ −1 (z)
X
= sgnϕ (z) deg(ψ, θ, z)
z∈ψ(θ),ϕ(z)=q
t X
X
= sgn Jϕ (z) deg(ψ, θ, z)
i=1 z∈Oi
X
= deg (ψ, θ, Oi ) deg (ϕ, Oi , q) .
i
" #
X X X X
deg (ψ, θ, Oi ) deg (ϕ, Oi , q) = deg (ϕ, Oi , q) = deg (ϕ, Um , q) .
i m i∈ωm m
and put
′
Um = {z ∈ Br+1 (0)\ψ0 (∂θ) : deg (ψ0 , θ, z) = m}
5
′ ∩ Γ and
Then we have Um ∩ Γ0 = Um 0
′ ′
deg (ϕ0 , Um , q) = deg ϕ0 , Um ∩ Um , q = deg ϕ0 , Um ,q .
Therefore, we have
X X
′
deg (ϕ0 ψ0 , θ, q) = m deg ϕ0 , Um ,q = m deg (ϕ, Um , q) .
m m
Using a simple homotopy argument, we can conclude that deg (ϕ0 ψ0 , θ, q) = deg (ϕ0 ψ, θ, q).
This confirms the validity of Theorem 2.2.2.
By using the formula we have just prooven, we can proove the following form of Jordan’s
separation theorem:
deg τ , Vn , Wk = deg τ ′ , Vn , Uj ,
′
6
XX
deg τ ′ , Vn , Uj deg τ −1 , Uj , q
1=
k j∈Nk
X (2.3.1)
deg τ ′ , Vn , Uj deg τ −1 , Uj , Vn ,
=
j
By combining equations (2.3.1) and (2.3.2), we can conclude that Rn \M1 and Rn \M2
possess an equal number of connected components.
7
Proof. To establish (2.5.1 ), it is adequate to consider the scenario where p ∈
/ φ (Qφ ).
Suppose φ−1 (p) = {z1 , z2 , · · · , zk }. Our task is to demonstrate:
Z k
1 1 X
dw = sgn Jφ (zi ) (2.5.2)
2πi φ(Ω) w−p
i=1
Select a small δ > 0 such that the closures of Vi are disjoint, where Vi = B (zi , δ). For
w ∈ Vi , define sgn Jφ (w) = sgnφ (zi ), with Vi ⊂ B(0, 1) and φ restricted to Vi being a
homeomorphism for i = 1, 2, . . . , k. Let Qi = ∂Vi . Thus, φ (Qi ) forms a Jordan curve with
a inside its interior. The orientation of φ (Qi ) matches that of Qi if Jφ (zi ) > 0, and is
opposite if Jφ (zi ) < 0.
Now, let U = B(0, 1)\ ∪ki=1 Vi . Then |φ(w) − p| > ρ in U for some ρ > 0. Divide U into small
rectangles Mj such that |φ(w) − φ(s)| < ρ on each Mj . Because the image φ (∂ (Mj ∩ V))
doesn’t wind around p, we have s (φ (∂ (Mj ∩ V)) , p) = 0, and summing over all Mj yields
Z k Z
1 X 1
dw = dw.
φ(Ω) w−p φ(Qi ) w−p
i=1
Given that the orientation of φ (Qi ) is dictated by Jφ (zi ), the curve φ (Qi ) completes
precisely one revolution around a. Therefore, we obtain:
Z
1
dw = sgn Jφ (zi ) .
φ(Qi ) w − p
Proof. Since Z Λ̄ is compact within E, for any positive value of ρ, there is a limited set
{w1 , w2 , . . . , wn } from the closure of Λ.
8
n
X ψi (ω)
Zρ ω = Zwi for all ω ∈ Λ̄
Γ(ω)
i=1
Here, ψi (ω) = max{0, ρ − ∥Zω − Zwi ∥} and Γ(ω) = ni=1 ψi (ω). It can be verified that
P
Zρ meets the requirements of Lemma 3.1.1, thus concluding the theorem.
deg(ω, E, γ) := deg (ω ρ , E ∩ X ρ , γ) ,
where ω ρ = id − ω ρ .
Lemma 3.1.3. Consider a real complete normed linear space X, where E ⊂ X is a bounded
and closed subset. Let L : E → X be a compact and continuous map such that L(z) ̸= z
for every z ∈ E. Then, there exists δ0 > 0 such that for all t ∈ [0, 1] and z ∈ E, we have
z ̸= tLδ1 (z) + (1 − t)Lδ2 (z), where δi ∈ (0, δ0 ) and Lδi : E → Fδi for i = 1, 2 as in Lemma
3.1.1.
Proof. Suppose the this is false. Then there exist some δ1j → 0, δ2j → 0, tj → t0 , wj ∈ E
such that tj Lδj wj + (1 − tj ) Lδj wj = wj for j = 1, 2, · · · .
1 2
As L is a compact map, the sequence (Lwj )∞ j=1 has a converging subsequence, denoted
as (Lwjk ), which converges to y ∈ X. According to Lemma 3.1.1, Lδi wjk → y for i =
1, 2. Therefore, we have wjk → y ∈ E. Consequently, we arrive at Ly = y, leading to a
contradiction.
9
w ̸= tLρ1 w + (1 − t)Lρ2 w for all t ∈ [0, 1], w ∈ ∂Γ,
Here, with ρi ∈ (0, ρ0 ) and Lρi : Γ̄ → Rρi for i = 1, 2 as in Lemma 3.1.1, Brouwer’s
degree deg (I − Lρ , Γ ∩ Rρ , 0) is well-defined. This allows us to define
deg(I − L, Γ, 0) = deg (I − Lρ , Γ ∩ Rρ , 0)
where ρ ∈ (0, ρ0 ).
Brouwer degree doesn’t change under homotopy. Thus we have
deg (I − Lρ1 , Γ ∩ span {Rρ1 ∪ Rρ2 } , 0) = deg (I − Lρ2 , Γ ∩ span {Rρ1 ∪ Rρ2 } , 0)
Theorem 3.2.1. Suppose you have a complete normed space X, where X0 is a closed
subspace of X, and Λ is a bounded open subset of X. If L : Λ̄ → X0 is a continuous and
compact map, and p ∈ X0 , then the degree of (I − L) over Λ with respect to p is equal to
the degree of (I − L) over Λ ∩ X0 with respect to p.
10
3.3 Condensing and countably condensing map
Definition 3.3.1. In a real normed space X, let S : E → X is a mapping, where K
represents the measure of noncompactness.
1. If, for all subsets B ⊂ E, which are bounded and the inequality K(SB) ≤ kK(B)
holds, then S is called a k-set contraction, for some k > 0.
2. S is considered to be condensing if, for all subsets B ⊂ E, which are bounded and
K(B) > 0, it satisfies K(SB) < K(B).
Proposition 3.4.1. Let (Y, d) be a metric space. Then (NY , H) is a metric space. Where
NY is the collection of all bounded subsets of Y .
and so
H(M, N )
( )
≤ max sup d(p, O) + sup d(r, N ), sup d(q, O) + sup d(r, M )
p∈M r∈O q∈N r∈O
( ) ( )
≤ max sup d(p, O), sup d(r, M ) + max sup d(r, N ), sup d(q, O)
p∈M r∈O r∈O q∈N
=H(M, O) + H(O, N ).
Thus (NY , H) is a metric space.
11
Proposition 3.4.2. |K(M ) − K(N )| ≤ 2H(M, N ).
Definition 3.4.1. Consider E as a real normed space which is also complete space, where
L : Γ → E is a mapping, and K represents the measure of noncompactness.
1. L is termed a countably k-set contraction if, for all countably bounded subsets φ ⊂ Γ,
it satisfies K(Lφ) ≤ kK(φ), for some k > 0.
2. L is considered a countably k-set contraction if, for all countably bounded subsets
φ ⊂ Γ, it fulfills K(Lφ) ≤ kK(φ), where k > 0 is a constant.
3. H(t, x) : [0, 1] × Γ → E is termed a homotopy where for each t ∈ [0, 1] we have a
countably condensing map if, for all bounded and countable subsets φ ⊂ Γ with K(φ) > 0,
it satisfies K(H([0, 1] × φ)) < K(φ).
Theorem 3.4.1. Consider X as a Normed space which is complete and A ⊂ C([a, b], X)
as a bounded equicontinuous subset. Then K(A(x)) is continuous on [a, b], where A(x) =
{f (x) : f (·) ∈ A}, and
Z b Z b
K f (x)dx : f (·) ∈ A ≤ K(A(x))dx.
a a
Proof. First, we will see why K({f (x) : f (·) ∈ A}) is a function with continuity property on
the this interval [a, b]. For any ϵ > 0, since A satisfies the conditions to be equicontinuous,
we can find δ > 0 such that ∥f (x) − f (x′ )∥ < ϵ for all x, x′ ∈ [a, b] satisfying |x − x′ | < δ.
Thus we have
for x, x′ ∈ [a, b] satisfying |x − x′ | < δ. From this and (6) of Proposition 3.4.2, we infer that
f x′ : f (·) ∈ A
K({f (x) : f (·) ∈ A}) − K ≤ 2ϵ
for x, x′ ∈ [a, b] satisfying |x − x′ | < δ. Thus K({f (x) : f (·) ∈ A}) is a continuous
function on [a, b].
Given any partition of [a, b] : a = x0 < x1 < · · · < xn = b, where xi = a + i b−a n for
i = 0, 1, · · · , n. For any ϵ > 0, due to the property of equicontinuity of A, there will be
N > 0 such that if n > N , then...
for i = 1, 2, · · · , n. So we have
b xi
b−a
Z Z
Σni=1 f (xi ) − f (x)dx = Σni=1 (f (xi ) − f (x)) dx < ϵ(b − a)
n a xi−1
12
( n )! Z b
X b−a
K f (xi ) : f (·) ∈ A −K f (x)dx : f (·) ∈ A ≤ 2ϵ(b − a),
n a
i=1
i.e.,
( n )! Z b
X b−a
lim K f (xi ) : f (·) ∈ A =K f (x)dx : f (·) ∈ A .
n→∞ n a
i=1
n b−a b−a
K Σi=1 f (xi ) : f (·) ∈ A ≤ Σni=1 K ({f (xi ) : f (·) ∈ A}) .
n n
Therefore, we get the following
Z b Z b
K f (x)dx : f (·) ∈ A ≤ K(A(x))dx.
a a
This prooves the theorem.
Theorem 3.4.2. Let X be a normed space which is also complete space and P ⊂ C([c, d], X)
be a bounded subset having equicontintinuity property, where c, d ∈ R. Then
1. W ⊆ K;
3. K = conv(S(K ∩ Λ̄)).
Proposition 3.4.4. Consider X as a normed space which is also complete space, where
Λ ⊂ X is a bounded subset and L : Λ̄ → X is a countably condensing map. Define
K1 = conv(LΛ̄), Kn+1 = conv L Kn ∩ Λ̄ for n ≥ 1, and K = ∩∞
n=1 Kn . As a result, K
is both convex and compact.
13
4.1 Defining degree for a special class of mappings
Let X be a normed space which is also complete space and Θ ⊂ X be an open subset having
boundedness. Suppose J : Θ̄ → X is a continuous mapping having property of countably
condensing map, with 0 ∈
/ (I − J )(∂Θ). If 0 ∈
/ (I − J )(Θ), we define deg(I − J , Θ, 0) = 0.
Otherwise, let Ω = {x ∈ Θ̄ : Lx = x} and consider K as the subset having convexity
and compactness property described in Proposition 3.4.3. The set K is nonempty as Ω is
contained in K. The mapping J : K ∩ Θ̄ → K is well-defined, following property 3 of
Proposition 3.4.3.
To see this definition is doesn’t depend upon the choosen retraction, we have to proove,
if W1 , W2 : X → K are two retractions, then
Define W(λ, x) = λW 1 (x) + (1 − λ)κ2 (x) for all (λ, x) ∈ [0, 1] × X. Then, W(λ, ·) :
X → K acts as a retraction for each λ ∈ [0, 1]. It is evident that x ̸= J W(λ, x) for
(λ, x) ∈ [0, 1]∂ r1−1 (Θ) ∩ κ−1
2 (Θ) ∩ Θ . Therefore, utilizing the homotopy property, we
obtain
and
14
deg I − J W2 , W2−1 (Θ) ∩ Θ, 0 = deg I − J W2 , W1−1 (Θ) ∩ W2−1 (Θ) ∩ Θ, 0 .
Therefore, we have
The degree for the compact set defined in 3.4.4 can be established in a similar manner,
aligning with the previously defined degree.
4.2 Properties of the degree theory defined for countably condensing map
Theorem 4.2.1. The degree for the maps that are defined by 4.1.1 has the property
of homotopy equivalence:
Consider S(λ, ω) : [0, 1] × Ē → E as continuous map which is also countably condens-
ing map, meaning K([0, 1]×A) < K(A) for all countable subsets A of Ē with K(A) > 0,
and S(λ, ω) ̸= ω for all (λ, ω) ∈ [0, 1] × ∂E. In this case, deg(I − S(λ, ·), E, 0) remains
constant regardless of the value of λ ∈ [0, 1].
Proof. Let
independent of λ. Thus, the homotopy proprty can be drawn from the d excision property
of the Leray-Schauder degree.
Proof. Let’s assume that P z ̸= z for all z ∈ ∂Λ. Define S(λ, z) = tP z for all (λ, z) ∈
[0, 1]× Λ̄. It is evident that {S(λ, ·)}λ∈[0,1] forms a homotopy wich are countably condensing
maps for each λ. Given this, we have S(λ, z) ̸= z for each z ∈ ∂Λ. Consequently, deg(I −
P, Λ, 0) = deg(I, Λ, 0) = 1. Hence, the equation P z = z has a solution in Λ, which prooves
the theorem.
15
Corollary 4.3.1. Consider X as a normed space which is also complete space, where B ⊂ X
is an open subset having boundedness property containing 0, and P : B̄ → X is a continuous
map which is also countably condensing map. If ∥P ω∥ ≤ ∥ω∥ for all ω ∈ ∂B, then P
possesses a within B̄, such that P z = z.
Proof. Let us suppose that P ω ̸= ω for each ω ∈ ∂B. Or else, the result is true.
we have λP ω ̸= ω for each ω ∈ ∂B and λ ∈ [0, 1) because
Proof. Let’s assume that P is a map which is λ-set contraction for some λ ∈ [0, 1). Define
P1 = conv P (D) and Pi+1 = conv P Pi for i = 1, 2, . . .. According to Proposition 3.4.4, the
intersection J = ∩∞
i=1 Pi is satisfies convexity and compactness properties, and the mapping
P : J → J is established.
of J . Choose i
Now, we demonstrate the non-emptiness
i
n i
z0 ∈ D, then P z0 ∈ Pi for i ≥ 1.
It follows that J {P z0 , i ≥ n} ≤ λ J {P z0 , i ≥ 0} for n ≥ 1.
P i z0 , i ≥ n P i z0 , i ≥ 0
J =J ,
∞
Therefore, J {P i z0 , i ≥ 0} = 0. Consequently, the sequence P i z0 i=1 has a subsequence
that converges to a point in J , indicating that J is not empty. Hence, P possesses a fixed
point.
Now, assuming P is countably condensing, select z0 ∈ D, and for any λ ∈ (0, 1), define
Sz = λP z + (1 − λ)z0 for all z ∈ D. Since S is λ-set contraction, it has a point in D, such
that Sz = z. Let λn → 1. Then there exist zn ∈ D such that
λn P zn + (1 − λn ) z0 = zn ,
This means that the closure of the set of images of P zn as n ≥ 1 is equal to the closure
of the set of zn as n ≥ 1. Given that P is countably condensing, the sequence (zn )∞ n=1
contains a subsequence znj converging to a point y ∈ D. Due to the continuity of P , we
have P y = y, so we got a fixed point.
Theorem 4.3.3. Consider X as a normed space which is also complete space with P :
X → X as a countably condensing map also satisfies the continuity. This ensures one of
the following situation:
16
Proof. Consider that {z : P z = κz for some κ > 1} is a bounded subset of X. Take d > 0
such that {z : P z = κz for some κ > 1} ⊂ B(0, d). If P z = z for some z ∈ ∂B(0, d),
then we have 1st situation holds. So we may suppose that P z ̸= z for all z ∈ ∂B(0, d).
Thus by the same reasoning as in 4.3.1 z ̸= tP z for all t ∈ [0, 1] and z ∈ ∂B(0, d), so
deg(I − P, B(0, d), 0) = 1 and P has a point in B(0, d) such that P z = z.
5 Applications to ODEs
5.1 Solvability result for a type of differential equations in Banach spaces
We see applications of degree for set contractive mappings to ordinary differential equations
in normed space which is also complete spaces.
Theorem 5.1.1. Consider X as a normed space which is also complete space, where
g(x, z) : [0, 1] × B(z0 , r0 ) → X is continuous map that satisfies
where κ ∈ (0, 1) and c0 > 0 are constants. Then there exists x0 ∈ (0, 1] such that the
initial value problem (
z ′ (x) = g(x, z(x)), x ∈ (0, x0 ) ,
(5.1.1)
z(0) = z0 ,
has a solution.
Proof. n o n c o
0
J = sup ∥g(x, z)∥ : (x, z) ∈ [0, 1] × B (z0 , c0 ) , x0 = min 1, .
J
Clearly, the first theorem of ordinary differential equations (ODE) in complete normed space
is equivalent to the following integral equation:
Z x
z(x) = z0 + g(s, z(s))ds, (5.1.2)
0
Put E = C ([0, x0 ] , X) having norm ∥z(·)∥ = max {∥z(x)∥ : x ∈ [0, x0 ]}. Then E is a
normed space which is also complete space. We also define K = {z(·) ∈ X : z (x0 ) = z0 , ∥z(x) − z (x0 )∥ ≤
c0 }. Then K is a subset which is bounded and closed and satisfies convexity.
Now, LLet a mapping S : K → K is given by
Z x
Sz(x) = z0 + g(s, z(s))ds for all z(·) ∈ K. (5.1.3)
0
It is simple to verify that S satisfies continuity. Now, we have to show that S is con-
densing. Which means, for any subset which satisfies, B of K with K(B) > 0, we see by
3.4.2 and 3.4.1,
17
K(SB) = sup K({Sz(x) : z(·) ∈ B})
x∈[0,x0 ]
Z x
= sup K z0 + g(s, z(s)) : z(·) ∈ B
x∈[0,x0 ] 0
)!
n o
≤ sup K z0 + x conv (g(s, z(s)) : s ∈ [0, x0 ]) : z(·) ∈ B
x∈[0,x0 ]
!
n o
≤ sup K z0 + x conv (g ([0, x0 ] × B)
x∈[0,x0 ]
≤ x0 κK(B).
Thus S is prooven to be a condensing mapping, so S has a point in K such that Sz = z,
i.e., 5.1.2 has a solution. Which implies, 5.1.1 has a solution. This completes the proof.
Rx RK
Proof. Since U(x) = U(0) + 0 U ′ (s)ds and U(x) = U(K) − x U ′ (s)ds, and x ∈ (0, K) we
have
Z x Z K
1 ′ ′
U(x) = U (s)ds − U (s)ds ,
2 0 x
Z x Z K
1 ′
|U(x)| = U (s)ds − U ′ (s)ds ,
2 0 x
Z x Z K
1 ′
|U(x)| ≤ U (s)ds + U ′ (s)ds ,
2 0 x
Z x
1
|U(x)| ≤ U ′ (s)ds .
2 0
Since it is true for every x ∈ (0, K). Thus the conclusion follows by Holder’s inequality
√ Z K 21
K ′ 2
∥U∥∞ ≤ U (s) ds .
2 0
18
5.2 Solvability result for a type of differential equation in Hilbert spaces
Theorem 5.2.1. Let Y be a normed space additionally it is complete inner product spcace,
and C > 0, Y(x, ζ) : R × Y → Y be a continuous mapping satisfying
K(Y([0, C] × B) ≤ λK(B),
for every subsets B of Y such that B is bounded, λ ∈ (0, 1) , and we have λC < 1.
If Y(x + C, −ζ) = −Y(x, ζ) and ∥Y(x, ζ)∥ ≤ N ∥ζ∥ + J (x) for all (x, ζ) ∈ R × E, where
0 ≤ N C < 2, and J (·) ∈ C 2 (0, C), then the following problem
(
ζ ′ (x) = Y(x, ζ(x)), x ∈ R
(5.2.1)
ζ(x + C) = −ζ(x),
has a solution.
K(SU ) ≤ λCK(U )
Thus S is a λC-set contraction.
Now, we show that Sζ(·) ̸= λζ(·) for all. λ > 1 and ζ(·) ∈ La with
−1 √ Z C 12
NC C 2
∥ζ(·)∥a > 1− J (x)dx .
2 2 0
Suppose this is not true, then there exists ζ(·) ∈ La with
√ Z C 12
N C −1 C
2
∥ζ(·)∥ > 1 − J (x)dx ,
2 2 0
19
such that Sζ(·) = κζ(·) for some κ > 1. Then we have
C C Z C 21 Z C 12
NC
Z Z
′ 2 ′ 2 2 ′ 2
κ ζ (x) dx ≤ ζ (x) dx + J (x)dx ζ (x) dx ,
0 2 0 0 0
i.e.,
12 12
C
N C −1
Z Z C
′ 2 2
ζ (x) dx ≤ 1− J (x)dx .
0 2 0
Again, by Lemma 5.1.2, we get
−1 √ Z C 12
NC C 2
∥ζ(·)∥a ≤ 1− J (x)dx ,
2 2 0
This leads to a contradiction. Therefore, according to Theorem 4.3.3, S possesses a fixed
point in La , implying that the problem 5.2.1 has a solution. This concludes the proof.
20
References
[1] Louis Nirenberg. Topics in nonlinear functional analysis, volume 6. American Mathe-
matical Soc., 1974.
[2] Yeol Je Cho and Yu-Qing Chen. Topological degree theory and applications. Chapman
and Hall/CRC, 2006.
21