Skip to content
BY 4.0 license Open Access Published by De Gruyter June 19, 2020

Solvability of an infinite system of integral equations on the real half-axis

  • Józef Banaś and Weronika Woś EMAIL logo

Abstract

The aim of the paper is to investigate the solvability of an infinite system of nonlinear integral equations on the real half-axis. The considerations will be located in the space of function sequences which are bounded at every point of the half-axis. The main tool used in the investigations is the technique associated with measures of noncompactness in the space of functions defined, continuous and bounded on the real half-axis with values in the space l consisting of real bounded sequences endowed with the standard supremum norm. The essential role in our considerations is played by the fact that we will use a measure of noncompactness constructed on the basis of a measure of noncompactness in the mentioned sequence space l. An example illustrating our result will be included.

1 Introduction

The paper is dedicated to the study of the existence of solutions of an infinite system of nonlinear integral equations on the real half-axis ℝ+ = [0, ∞). More precisely, we will look for solutions of the mentioned infinite system in the space BC(ℝ+, l) consisting of functions defined, continuous and bounded on the interval ℝ+ with values in the classical sequence space l which is equipped with the standard supremum norm. Thus, any solution of our infinite system of integral equations will be treated as a sequence of functions (xn(t)) defined on ℝ+ and such that for any fixed t ∈ ℝ+ the sequence (xn(t)) is an element of the space l. Further details concerning the properties of the solutions in question will be formulated in the sequel of our paper.

The present paper is a continuation of papers [1, 2], where we constructed measures of noncompactness needed in our study. Particularly, in paper [2] we constructed measures of noncompactness in the Banach space BC(ℝ+, E) containing functions defined, continuous and bounded on the interval ℝ+ and taking values in a given Banach space E. The construction of those measures depends strongly on a given measure of noncompactness in the space E. Additionally, in the mentioned paper [2] we applied one of the constructed measures of noncompactness to the study of the solvability of an infinite system of integral equations in the space BC(ℝ+, l).

In this paper we are going to study a more general infinite system of nonlinear integral equations in the same Banach space BC(ℝ+, l) but with the use of another measure of noncompactness. Such an approach allows us to obtain an existence result concerning the mentioned infinite system, but under weaker assumptions than the existence result obtained in [2]. Thus, our result creates a generalization of the existence result contained in paper [2]. It is also worthwhile mentioning that in paper [1] it was also studied the solvability of an infinite system of integral equations in the Banach space BC(ℝ+, E) but we assumed that E is a Banach space with a regular measure of noncompactness being equivalent to the so-called Hausdorff measure of noncompactness. Let us point out that in some Banach spaces such measures of noncompactness are not known [3].

Finally, let us mention that the theory of infinite systems of integral equations has recently been intensively developed and up to now there have appeared a lot of papers concerning those infinite systems [4, 5, 6, 7, 8, 9]. That theory is closely related to the theory of infinite systems of differential equations (cf. [4, 7, 10] and references therein).

However, in the papers published up to now the authors investigated mainly infinite systems of integral equations in the Banach space C([0, T], E) consisting of functions defined and continuous on the bounded interval [0, T] with values in a Banach space E. The generalization to the space BC(ℝ+, E) is rather a quite new branch of the theory of infinite systems of integral equations (cf. [1, 2]).

2 Notations, definitions and auxiliary facts

In this section we establish the notations used in the paper and we provide definitions creating the basis of our study conducted further on. Apart from this we give some facts concerning the theory of measures of noncompactness being the basic tool utilized in our considerations.

In the paper we will denote by ℝ the set of real numbers while ℕ stands for the set of natural numbers. We will put ℝ+ = [0, ∞). Further assume that E is a Banach space with the norm ∥⋅∥E and the zero element θ. We will denote by B(x, r) the closed ball centered at x and with radius r. We write Br to denote the ball B(θ, r).

If X, Y are subsets of the Banach space E and λ ∈ ℝ then the standard algebraic operations on sets will be denoted by X + Y and λ X. Moreover, the symbol X denotes the closure of the set X while ConvX stands for the closed convex hull of the set X.

Next, denote by 𝔐E the family of all nonempty and bounded subsets of E and by 𝔑E its subfamily consisting of all relatively compact sets.

In our considerations we will accept the following definition of the concept of a measure of noncompactness (cf. [3, 4]).

Definition 2.1

A function μ : 𝔐E → ℝ+ will be called a measure of noncompactness in the space E if it satisfies the following conditions:

  1. The family ker μ = {X ∈ 𝔐E : μ(X) = 0} is nonempty and ker μ ⊂ 𝔑E.

  2. XYμ(X) ⩽ μ(Y).

  3. μ(X) = μ(X).

  4. μ(Conv X) = μ(X).

  5. μ(λ X + (1 − λ)Y) ⩽ λμ(X) + (1 − λ)μ(Y) for λ ∈ [0, 1].

  6. If (Xn) is a sequence of closed sets from 𝔐E such that Xn+1Xn for n = 1, 2, … and if limn μ(Xn) = 0 then the set X=n=1Xn is nonempty.

The family ker μ from axiom (i) will be called the kernel of the measure of noncompactness μ. Let us also observe that the set X defined in axiom (vi) is an element of the family ker μ. In fact, it follows easily from the inclusion XXn for n = 1, 2, … and axiom (ii) which implies the inequality μ(X) ⩽ μ(Xn) for any n ∈ ℕ. Hence we have that μ(X) = 0 and consequently, X ∈ ker μ. This simple conclusion is very crucial in applications.

In what follows assume that μ is a measure of noncompactness in the space E. The measure μ is called sublinear [3] if it satisfies the following additional conditions

  1. μ(X + Y) ⩽ μ(X) + μ(Y).

  2. μ(λX) = ∣λμ(X) for λ ∈ ℝ.

If the measure of noncompactness μ satisfies the condition

  1. μ(XY) = max{μ(X), μ(Y)}

then we say that it has the maximum property. Moreover, if kerμ = 𝔑E we say that μ is full. If μ is a sublinear and full measure of noncompactness with the maximum property, then μ is said to be regular.

Let us recall that from the historical point of view the first measure of noncompactness was defined in 1930 by K. Kuratowski [11], but the most important and useful measure of noncompactness is the so-called Hausdorff measure of noncompactness which was defined in [12, 13] by the formula

χ(X)=inf{ε>0:Xhas a finiteεnetinE},

where X ∈ 𝔐E. The importance of the Hausdorff measure χ is caused by the fact that in some Banach spaces like C([a, b]), c0 and lp we can give formulas expressing the measure χ in connection with the structure of the mentioned Banach spaces. Let us mention that χ is a regular measure of noncompactness [3, 14, 15].

However, in a lot of Banach spaces we are not in a position to give formulas for the Hausdorff measure of noncompactness χ. Even more, we are not able to provide formulas for full measures of noncompactness [3, 4]. Therefore, in such a situation we restrict ourselves to measures of noncompactness in the sense of Definition 2.1, which are not full (cf. also [3, 4, 14]). In the present paper we will also consider a measure of noncompactness of such a type.

Namely, assume that E is a given Banach space with the norm ∥⋅∥E and μ is a measure of noncompactness in the space E. Consider the Banach space BC(ℝ+, E) consisting of all functions x : ℝ+E which are continuous and bounded on the interval ℝ+. The space BC(ℝ+, E) will be equipped with the standard supremum norm

x=sup{x(t)E:tR+}.

Further, take an arbitrary nonempty and bounded subset X of the space BC(ℝ+, E). Fix xX and ε > 0. We define the modulus of the uniform continuity of the function x (cf.[2]) by putting

ω(x,ε)=sup{x(t)x(s)E:t,sR+,|ts|ε}.

Obviously, limε0ω(x,ε)=0 if and only if the function x is uniformly continuous on the interval ℝ+.

Further, let us define the following quantities

ω(X,ε)=sup{ω(x,ε):xX},ω0(X)=limε0ω(X,ε). (2.1)

Next, let us consider the function μ defined on the family 𝔐BC(ℝ+,E) in the following way

μ¯(X)=limTμ¯T(X), (2.2)

where μT(X) is defined by the formula

μ¯T(X)=sup{μ(X(t)):t[0,T]}

for any fixed T > 0.

Finally, for a given T > 0 let us put

aT(X)=supxX{sup{x(t)E:tT}}

and

a(X)=limTaT(X). (2.3)

Now, linking quantities defined by (2.1), (2.2) and (2.3), we can consider the following function μa defined on the family 𝔐BC(ℝ+,E):

μa(X)=ω0(X)+μ¯(X)+a(X). (2.4)

It can be shown that the function μa is a measure of noncompactness in the space BC(ℝ+, E) (cf. [2]). The kernel kerμa of the measure μa consists of all bounded subsets X of the space BC(ℝ+, E) such that functions from X are uniformly continuous and equicontinuous on ℝ+ (equivalently we can say that functions from X are equiuniformly continuous on ℝ+) and tend to zero at infinity with the same rate. Apart from this, all cross-sections X(t) = {x(t) : xX} of the set X belong to the kernel kerμ of the measure of noncompactness μ in the Banach space E (cf. [2]). The measure μa is not full and has the maximum property. If the measure μ is sublinear in E then the measure μa defined by (2.4) is also sublinear [2].

Let us notice that in the similar way as above we may define other measures of noncompactness in the space BC(ℝ+, E) (see [2]). We will not provide the definitions of those measures since we will use only the measure of noncompactness μa further on.

In what follows, taking into account our further purposes, we will consider as the Banach space E, the sequence space l endowed with the standard supremum norm.

Thus, let us consider the Banach space BC(ℝ+, l) consisting of functions x : ℝ+l which are continuous and bounded on ℝ+. Observe, that if xBC(ℝ+, l), then we can write this function in the form

x(t)=(xn(t))=(x1(t),x2(t),)

for any t ∈ ℝ+, where the sequence (xn(t)) is an element of the space l for any fixed t. The norm of the function x = x(t) = (xn(t)) is defined by the equality

x=sup{x(t)l:tR+}=suptR+{sup{|xn(t)|:n=1,2,}}.

In our further considerations the space BC(ℝ+, l) will be denoted by BC.

Now, we provide the formula expressing the measure of noncompactness μa defined by (2.4) in the space BC, provided the measure of noncompactness μ in the sequence space l is defined by the formula [3]

μ1(X)=limn{supx=(xi)X{sup{|xk|:kn}}}

for X ∈ 𝔐l. In such a case the component μ defined by (2.2) will be denoted by μ¯1.

Thus, our measure of noncompactness μa defined by (2.4) will be now denoted by μa1 and is defined as a particular case of (2.4) by the formula

μa1(X)=ω0(X)+μ¯1(X)+a(X), (2.5)

where the components on the right hand side of formula (2.5) are defined in the following way:

ω0(X)=limε0{supxX{sup{supnN|xn(t)xn(s)|:t,sR+,|ts|ε}}}, (2.6)
μ¯1(X)=limT{supt[0,T]{limn{supxX{sup{|xk(t)|:kn}}}}}, (2.7)
a(X)=limT{supxX{sup{supnN|xn(t)|:tT}}}. (2.8)

In what follows we give an other formula expressing the quantity μ defined by (2.2).

To this end we prove the following lemma.

Lemma 2.2

The following equality is satisfied

μ¯(X)=sup{μ(X(t)):tR+},

where μ is defined by formula (2.2).

Proof

Obviously, for any fixed T > 0 we have

sup{μ(X(t)):t[0,T]}sup{μ(X(t)):tR+}.

Hence, we get

μ¯(X)=limT{sup{μ(X(t)):t[0,T]}}sup{μ(X(t)):tR+}. (2.9)

To prove the converse inequality, let us denote

δ=sup{μ(X(t)):tR+}.

Further, fix an arbitrary number ε > 0. Then we can find a number t0 ∈ ℝ+ such that

δεμ(X(t0)).

Hence, for Tt0 we obtain

δεsup{μ(X(t)):t[0,T]}. (2.10)

Since the function T → sup{μ(X(t)) : t ∈ [0, T]} is nondecreasing, we get

sup{μ(X(t)):t[0,T]}limT{sup{μ(X(t)):t[0,T]}}. (2.11)

Combining (2.10) and (2.11), we have

δεlimT{sup{μ(X(t)):t[0,T]}}.

Consequently, in view of the arbitrariness of the number ε, we derive the following inequality

δlimT{sup{μ(X(t)):t[0,T]}}=μ¯(X). (2.12)

Finally, linking (2.9) and (2.12) we obtain the desired equality.□

Now, let us notice that taking into account Lemma 2.2 and formula (2.7) expressing the quantity μ in the case of the space BC, we obtain the following corollary.

Corollary 2.3

The quantity (2.7) can be expressed by the following formula

μ¯1(X)=supt0{limn{supxX{sup{|xk(t)|:kn}}}}.

At the end of this section we recall a useful fixed point theorem of Darbo type [3, 16].

To this end let us assume that E is a Banach space and μ is a measure of noncompactness (in the sense of Definition 2.1) in the space E.

Theorem 2.4

Assume that Ω is a nonempty, bounded, closed and convex subset of a Banach space E and Q : ΩΩ is a continuous operator such that there exists a constant k ∈ [0, 1) for which μ(QX) ⩽ (X) for an arbitrary nonempty subset X of Ω. Then there exists at least one fixed point of the operator Q in the set Ω.

Remark 2.5

It can be shown that the set Fix Q of all fixed points of the operator Q belongs to the family ker μ [3].

This simple observation is very essential in characterization of solutions of considered operator equations which are proved with help of Theorem 2.4.

3 Solvability of an infinite system of quadratic integral equations on the real half-axis

The aim of this section is to investigate the infinite system of the quadratic integral equations of Volterra-Hammerstein type having the form

xn(t)=an(t)+fn(t,x1(t),x2(t),)0tkn(t,s)gn(s,x1(s),x2(s),)ds, (3.1)

where t ∈ ℝ+ and n = 1, 2, ….

Our goal is to prove the existence of solutions of the infinite system of integral equations (3.1).

Considerations conducted further on will be situated in the Banach space BC = BC(ℝ+, l) described in details previously. Moreover, in our study we are going to use the measure of noncompactness μa1 (X) defined by formula (2.5).

Now, we formulate the assumptions under which the infinite system (3.1) will be considered.

  1. The sequence (an(t)) is an element of the space BC such that limtan(t)=0 uniformly with respect to n ∈ ℕ i.e., the following condition is satisfied

    ε>0T>0tTnN|an(t)|ε.

    Moreover, limn an(t) = 0 for any t ∈ ℝ+.

  2. The functions kn(t, s) = kn : R+2 → ℝ are continuous on the set R+2 (n = 1, 2, …). Moreover, the functions tkn(t, s) are equicontinuous on the set ℝ+ uniformly with respect to s ∈ ℝ+ i.e., the following condition is satisfied

    ε>0δ>0nNsR+t1,t2R+[|t2t1|δ|kn(t2,s)kn(t1,s)|ε].
  3. There exists a constant K1 > 0 such that

    0t|kn(t,s)|dsK1

    for any t ∈ ℝ+ and n = 1, 2, ….

  4. The sequence (kn(t, s)) is equibounded on R+2 i.e., there exists a constant K2 > 0 such that |kn(t, s)| ⩽ K2 for t, s ∈ ℝ+ and n = 1, 2, ….

  5. The function fn is defined on the set ℝ+ × ℝ and takes real values for n = 1, 2, …. Moreover, the function tfn(t, x1, x2, …) is uniformly continuous on ℝ+ uniformly with respect to x = (xn) ∈ l and uniformly with respect to n ∈ ℕ i.e., the following condition is satisfied

    ε>0δ>0(xi)lnNt,sR+[|ts|δ|fn(t,x1,x2,)fn(s,x1,x2,)|ε].
  6. There exists a function l : ℝ+ → ℝ+ such that l is nondecreasing on ℝ+, continuous at 0 and there exists a sequence of functions (fn) being an element of the space BC, taking nonnegative values and such that limt fn(t) = 0 uniformly with respect to n ∈ ℕ (cf. assumption (i)) and limn fn(t) = 0 for any t ∈ ℝ+. Moreover, for any r > 0 the following inequality is satisfied

    |fn(t,x1,x2,)|f¯n(t)+l(r)sup{|xi|:in}

    for each x = (xi) ∈ l such that ∥xlr, for every t ∈ ℝ+ and for n = 1, 2, ….

Observe that in view of assumption (vi) we can define the finite constant F = sup{fn(t) : t ∈ ℝ+, n = 1, 2, …}.

Now, we can formulate the other assumptions concerning the infinite system (3.1).

  1. There exists a function m : ℝ+ → ℝ+ such that m is nondecreasing on ℝ+, continuous at 0 and the following inequality is satisfied

    |fn(t,x1,x2,)fn(t,y1,y2,)|m(r)xyl

    for any r > 0, for x = (xi), y = (yi) ∈ l such that ∥xlr, ∥ylr and for all t ∈ ℝ+ and n = 1, 2, ….

  2. The function gn is defined on the set ℝ+ × ℝ and takes real values for n = 1, 2, …. Moreover, the operator g defined on the set ℝ+ × l by the formula

    (gx)(t)=(gn(t,x))=(g1(t,x),g2(t,x),)

    transforms the set ℝ+ × l into l and is such that the family of functions {(gx)(t)}t∈ℝ+ is equicontinuous on the space l i.e., for any ε > 0 there exists δ > 0 such that

    (gy)(t)(gx)(t)lε

    for any t ∈ ℝ+ and for all x, yl such that ∥xylδ.

  3. The operator g defined in assumption (viii) is bounded on the set ℝ+ × l. More precisely, there exists a positive constant G such that ∥(gx)(t)∥lG for any xl and t ∈ ℝ+.

  4. There exists a positive solution r0 of the inequality

    A+F¯G¯K1+G¯K1rl(r)r

    such that GK1 max {l(r0), m(r0)} < 1, where the constants F, G, K1 were defined above and the constant A is defined in the following way

    A=sup{|an(t)|:tR+,n=1,2,}. (3.2)

Before formulating our main result we indicate some consequences of assumption (i).

Lemma 3.1

Let the function x(t) = (xn(t)) be an element of the space BC. Then the sequence (xn) is equibounded and locally equicontinuous on+.

The proof can be conducted analogously as the proof of Lemma 4.1 in paper [1] and is therefore omitted.

Lemma 3.2

Let the sequence (an(t)) be an element of the space BC such that limt an(t) = 0 uniformly with respect to n ∈ ℕ (cf. assumption (i)). Then the sequence (an) is equibounded and equicontinuous on+.

Proof

The equiboundedness of the sequence (an) on ℝ+ follows from Lemma 3.1. To prove the equicontinuity of (an) on ℝ+ let us fix ε > 0. Keeping in mind the remaining part of the assumption in our lemma we can find a number T > 0 such that |an(t)| ⩽ ε/4 for tT and n = 1, 2, …. On the other hand, in view of Lemma 3.1 we deduce that the sequence (an) is locally equicontinuous on ℝ+. Thus, we can find a number δ > 0 such that |an(t2) – an(t1)| ⩽ ε/2 for t1, t2 ∈ [0, T] such that |t2t1| ⩽ δ and for any n = 1, 2, …. Now, let us take arbitrary t1, t2 ∈ ℝ+ such that |t2t1| ⩽ δ. Without loss of generality we can assume that t1 < t2. If t1, t2 ∈ [0, T], then in view of the above established fact we have that |an(t2) – an(t1)| ⩽ ε/2 for n = 1, 2, ….

If t1, t2T, then we obtain

|an(t2)an(t1)||an(t2)|+|an(t1)|ε2

for any n = 1, 2, ….

Assume now that t1 < Tt2. Then, fixing arbitrarily n ∈ ℕ and taking into account the above derived facts we get

|an(t2)an(t1)||an(t2)an(T)|+|an(T)an(t1)|ε2+ε2=ε.

This shows that the sequence (an) is equicontinuous on ℝ+.□

In what follows let us notice that as an immediate consequence of Lemma 3.1 we obtain the following corollary.

Corollary 3.3

The constant A defined by equality (3.2) is finite.

Now, we are in a position to formulate an existence theorem concerning the infinite system of integral equations (3.1).

Theorem 3.4

Under assumptions (i) – (x) the infinite system (3.1) has at least one solution x(t) = (xn(t)) in the space BC = BC(ℝ+, l).

Proof

At the beginning we define three operators F, V, Q on the space BC in the following way:

(Fx)(t)=((Fnx)(t))=(fn(t,x(t)))=(fn(t,x1(t),x2(t),)),(Vx)(t)=((Vnx)(t))=(0tkn(t,s)gn(s,x1(s),x2(s),)ds),(Qx)(t)=((Qnx)(t))=(an(t)+(Fnx)(t)(Vnx)(t)).

At first we show that the operator F transforms the space BC into itself.

To this end fix the function x = x(t) = (xn(t)) ∈ BC. Then, in view of assumption (vi) we have the following estimate

|(Fnx)(t)|=|fn(t,x1(t),x2(t),)|f¯n(t)+l(x(t)l)sup{|xi(t)|:in} (3.3)

for t ∈ ℝ+ and for n = 1, 2, …, where the functions fn and l = l(r) were specified in assumption (vi). Hence, on the basis of (3.3) we deduce that

FxBCF¯+l(xBC)xBC (3.4)

for any xBC. This shows that the function Fx is bounded on ℝ+.

In order to prove the continuity of the function Fx on the interval ℝ+, let us fix ε > 0. Then, from assumption (v) we obtain that there exists δ > 0 such that for t, s ∈ ℝ+ and |ts| ⩽ δ the following inequality is satisfied

|fn(t,x1,x2,)fn(s,x1,x2,)|ε

for any x = (xi) ∈ l. This implies that

(Fx)(t)(Fx)(s)lε

provided t, s ∈ ℝ+ are such that |ts| ⩽ δ. But this means that the function Fx is continuous (even uniformly continuous) on ℝ+. Finally we conclude that the operator F transforms the space BC into itself.

Now, we are going to show that the operator V acts from the space BC into itself.

Thus, similarly as above, fix a function x = x(t) = (xn(t)) ∈ BC. Then, for arbitrarily fixed numbers t ∈ ℝ+ and n ∈ ℕ, in view of assumptions (iii) and (ix), we get

|(Vnx)(t)|0t|kn(t,s)||gn(s,x1(s),x2(s),)|ds0t|kn(t,s)|G¯dsG¯0t|kn(t,s)|dsG¯K1. (3.5)

Particularly, the above estimate yields that the function Vx is bounded on the interval ℝ+.

Next, fix ε > 0 and choose a number δ > 0 according to assumption (ii). Then, for arbitrarily fixed numbers t1, t2 ∈ ℝ+ such that |t2t1| ⩽ δ, on the basis of assumptions (ii) and (ix) (assuming additionally that t1 < t2), we obtain

|(Vnx)(t2)(Vnx)(t1)|0t2kn(t2,s)gn(s,x1(s),x2(s),)ds0t2kn(t1,s)gn(s,x1(s),x2(s),)ds+0t2kn(t1,s)gn(s,x1(s),x2(s),)ds0t1kn(t1,s)gn(s,x1(s),x2(s),)ds0t2|kn(t2,s)kn(t1,s)||gn(s,x1(s),x2(s),)|ds+t1t2|kn(t1,s)||gn(s,x1(s),x2(s),)|ds0t2ωk(δ)|gn(s,x1(s),x2(s),)|ds+t1t2K2|gn(s,x1(s),x2(s),)|ds, (3.6)

where K2 is a constant appearing in assuption (iv) and ωk(δ) denotes a common modulus of continuity of the sequence of functions tkn(t, s) on the interval ℝ+ (according to assumption (ii)). Obviously we have that ωk(δ) → 0 as δ → 0.

Further on, taking into account estimate (3.6) and assumption (ix), we obtain

|(Vnx)(t2)(Vnx)(t1)|G¯ωk(δ)+K2G¯δ. (3.7)

Hence we deduce that the function Vx is continuous on the interval ℝ+. Linking boundedness of the function Vx with its continuity on ℝ+ we conclude that the operator V transforms the space BC into itself.

Now, keeping in mind the fact that the space BC = BC(ℝ+, l) forms a Banach algebra with respect to the coordinatewise multiplication of function sequences and taking into account the definition of the operator Q as well as assumption (i), we infer that for an arbitrarily fixed function x = x(t) ∈ BC the function (Qx)(t) = ((Qnx)(t)) = (an(t) + (Fnx)(t)(Vnx)(t)) acts from the interval ℝ+ into the space l. Indeed, in view of the fact that ((Fnx)(t)) ∈ l for any t ∈ ℝ+ and in the light of estimate (3.5) we get

|(Qnx)(t)||an(t)|+G¯K1|(Fnx)(t)|.

Hence, applying (3.3) we deduce that (Qx)(t) = ((Qnx)(t)) ∈ l for each t ∈ ℝ+.

Next, let us notice that the continuity of the function Qx on ℝ+ is a simple consequence of the fact that both the function Fx and the function Vx are continuous on ℝ+. Similarly we can derive that the function Qx is bounded on the interval ℝ+. Indeed, it is only sufficient to make use assumption (i) and Lemma 3.1.

Finally, let us observe that combining all the above established properties of the function Qx we conclude that the operator Q transforms the space BC into itself.

Further, let us note that based on estimates (3.4) and (3.5), for arbitrarily fixed n ∈ ℕ and t ∈ ℝ+ we get

|(Qnx)(t)||an(t)|+|(Fnx)(t)||(Vnx)(t)|A+[F¯+l(xBC)xBC]G¯K1A+F¯G¯K1+G¯K1l(xBC)xBC.

From the above estimate and assumption (x) we infer that there exists a number r0 > 0 such that the operator Q transforms the ball Br0 into itself.

In what follows we are going to show that the operator Q is continuous on the ball Br0. Keeping in mind the representation of the operator Q given at the beginning of our proof we see it is sufficient to prove the continuity of the operators F and V, separately.

To this end let us fix ε > 0 and xBr0. Next, choose an arbitrary point yBr0 such that ∥xyBCε. Then, for each fixed t ∈ ℝ+, in virtue of assumption (vii) we get

(Fy)(t)(Fx)(t)lm(r0)xylεm(r0).

Particularly this shows that the operator F is continuous at every point of the ball Br0.

To prove the continuity of the operator V on the ball Br0 let us define the function δ = δ(ε) by putting

δ(ε)=sup{|gn(t,y)gn(t,x)|:x,yl,yxlε,tR+,nN}.

Obviously, in view of assumption (viii) we have that δ(ε) → 0 as ε → 0. Next, taking x, yBr0 such that ∥yxBCε and t ∈ ℝ+, for arbitrary n ∈ ℕ we obtain

|(Vny)(t)(Vnx)(t)|0t|kn(t,s)||gn(s,y1(s),y2(s),)gn(s,x1(s),x2(s),)|ds0t|kn(t,s)|δ(ε)dsK1δ(ε).

This yields to the estimate

VxVyBCK1δ(ε).

Thus we see that the operator V is continuous on the ball Br0.

In the sequel let us fix an arbitrary number ε > 0. Next, choose a number δ > 0 according to assumption (v). Further fix a nonempty subset X of the ball Br0 and take an arbitrary function xX and n ∈ ℕ. Then, for arbitrary t, s ∈ ℝ+ such that |ts| ⩽ δ, in virtue of assumption (v) we get

|(Fnx)(t)(Fnx)(s)|=|fn(t,x1(t),x2(t),)fn(s,x1(s),x2(s),)||fn(t,x1(t),x2(t),)fn(s,x1(t),x2(t),)|+|fn(s,x1(t),x2(t),)fn(s,x1(s),x2(s),)|ε+m(r0)sup{x(t)x(s)l:t,sR+,|ts|δ}ε+m(r0)ω(x,δ).

Now, on the basis of the above estimate we obtain

ω(Fx,ε)ε+m(r0)ω(x,δ). (3.8)

Further, let us fix a number ε > 0 and choose t1, t2 ∈ ℝ+ such that |t2t1| ⩽ ε. Without loss of generality we may assume that t1 < t2. Then, in view of estimate (3.7) we derive the following inequality

|(Vnx)(t2)(Vnx)(t1)|G¯ωk(ε)+G¯K2ε.

This yields the estimate

ω(Vx,ε)G¯ωk(ε)+G¯K2ε. (3.9)

Now, taking into account the representation of the operator Q, for an arbitrary function xX and for arbitrary numbers t, s ∈ ℝ+, we obtain

(Qx)(t)(Qx)(s)la(t)a(s)l+(Vx)(t)l(Fx)(t)(Fx)(s)l+(Fx)(s)l(Vx)(t)(Vx)(s)l,

where a(t) = (an(t)).

Further, fix ε > 0 and assume that |ts| ⩽ ε. Then, from the above inequality and estimates (3.8), (3.9), (3.4), and (3.5), we get

ω(Qx,ε)ω(a,ε)+G¯K1ω(Fx,ε)+(F¯+r0l(r0))(G¯ωk(ε)+G¯K2ε)ω(a,ε)+G¯K1m(r0)ω(x,ε)+G¯K1ε+(F¯+r0l(r0))(G¯ωk(ε)+G¯K2ε).

Now, in view of Lemma 3.2 we infer that ω(a, ε) → 0 as ε → 0. Next, keeping in mind that ωk(ε) → 0 as ε → 0, from the above obtained estimate we derive the following inequality

ω(QX)G¯K1m(r0)ω0(X). (3.10)

In what follows we will consider the second term of the measure of noncompactness μa1 (cf. formula (2.5)) which is denoted by μ¯1 and is defined by formula (2.7).

To this end fix a nonempty subset X of the ball Br0 and choose an arbitrary function x = x(t) ∈ X. Further, take a natural number n and T > 0. Then, for any fixed t ∈ [0, T], in view of the representation of the operator Q and estimates (3.3) and (3.5), we obtain

|(Qnx)(t)||an(t)|+|fn(t,x1(t),x2(t),)|0t|kn(t,s)||gn(s,x1(s),x2(s),)|ds|an(t)|+[f¯n(t)+l(x(t)l)sup{|xi(t)|:in}]G¯K1.

Now, taking supremum over all xX, from the above estimate we get

supxX|(Qnx)(t)||an(t)|+G¯K1[f¯n(t)+l(r0)supxX{sup{|xi(t)|:in}}].

Hence, taking into account assumption (i) and (vi), we derive the following inequality

limnsupxX|(Qnx)(t)|G¯K1l(r0)limnsupxX{sup{|xi(t)|:in}}.

Finally, taking supremum over t ∈ [0, T] on both sides of the above inequality and next, passing with T → ∞, in virtue of formula (2.7) we get

μ¯1(QX)G¯K1l(r0)μ¯1(X). (3.11)

In order to estimate the last term a of the measure of noncompactness μa1 (cf. formula (2.5)) expressed by formula (2.8), let us take a nonempty subset X (XBr0) and choose a function xX. Further, fix arbitrarily T > 0. Then, taking tT and keeping in mind the previously obtained inequalities (3.3) and (3.5), we obtain

sup{|(Qnx)(t)|:nN}sup{|an(t)|:nN}+sup|fn(t,x1(t),x2(t),)|0t|kn(t,s)||gn(s,x1(s),x2(s),)|ds:nNsup{|an(t)|:nN}+sup[f¯n(t)+l(r0)sup{|xi(t)|:in}:nN]G¯K1sup{|an(t)|:nN}+l(r0)G¯K1sup[sup{|xi(t)|:in}:nN]+G¯K1sup{f¯n(t):nN}.

Now, taking in the above estimate supremum over tT and next, over xX, we get

supxX{suptT{sup{|(Qnx)(t)|:nN}}}suptT{sup{|an(t)|:nN}}+l(r0)G¯K1{supxX{suptT{sup{|xn(t)|:nN}}}}+G¯K1{suptT{sup{f¯n(t):nN}}}.

Further, passing with T → ∞ and taking into account assumptions (i) and (vi), we derive the following estimate

a(QX)l(r0)G¯K1a(X). (3.12)

Finally, combining estimates (3.10)-(3.12) and keeping in mind formula (2.5), we obtain the following inequality for an arbitrary nonempty subset X of the ball Br0:

μa1(QX)G¯K1max{l(r0),m(r0)}μa1(X).

Hence, in view of the fact that the operator Q is a continuous self-mapping of the ball Br0, assumption (x) and Theorem 2.4 we conclude that the infinite system of Volterra-Hamerstein integral equations (3.1) has at least one solution x = x(t) in the space BC = BC(ℝ+, l) which belongs to the ball Br0 and is uniformly continuous on the interval ℝ+.

The proof is complete. □

4 An example

Now, we are going to provide an example illustrating the existence result contained in Theorem 3.4.

To this end, let us consider the following infinite system of nonlinear Volterra-Hammerstein integral equations having the form:

xn(t)=αt1+n2+t2+βn2+t2+γxn(t)1+x12(t)+γxn+1(t)n+x22(t)0ts1+n(s2+t2)arctanx1(s)+xn(s)n+s2ds, (4.1)

for n = 1, 2, … and t ∈ ℝ+. Moreover, we assume that α, β, y appearing in the above system are positive constants.

Notice that infinite system (4.1) is a particular case of system (3.1) if we put

an(t)=αt1+n2+t2, (4.2)
fn(t,x1,x2,)=βn2+t2+yxn1+x12+yxn+1n+x22, (4.3)
kn(t,s)=s1+n(s2+t2), (4.4)
gn(t,x1,x2,)=arctanx1+xnn+t2 (4.5)

for n = 1, 2, … and t, s ∈ ℝ+.

We are going to show that the infinite system of integral equations (4.1) has a solution in the Banach space BC = BC(ℝ+, l). To this end we will apply Theorem 3.4.

Thus, we show that functions defined by (4.2)-(4.5) satisfy assumptions (i)-(x) of Theorem 3.4.

At the beginning let us observe that the function an(t) defined by (4.2) is an element of the space BC for n = 1, 2, …. In view of the inequality |an(t)|=an(t)αt1+t2 we derive that limt an(t) = 0 uniformly with respect to n ∈ ℕ. Moreover, limn an(t) = 0 for any t ∈ ℝ+. This show that the sequence (an(t)) satisfies assumption (i). Apart from this we have that A = α/2 2 , where the constant A is defined by (3.2).

Further, let us notice that the function kn(t, s) defined by (4.4) (n = 1, 2, …) is continuous on R+2 . Additionally, using standard tools of differential calculus it is easily seen that

|kn(t2,s)kn(t1,s)|1n|t2t1|

for n = 1, 2, … and for t1, t2 ∈ ℝ+. This means that the sequence of functions (kn(⋅, s)) is equicontinuous on ℝ+ uniformly with respect to s ∈ ℝ+.

Summing up, we see that there is satisfied assumption (ii).

Next, let us observe that for each n ∈ ℕ and for arbitrary t, s ∈ ℝ+ we have the following estimate

|kn(t,s)|s1+ns2s1+s212.

Hence it follows that the sequence (kn(t, s)) is equibounded on ℝ+ with the constant K2 = 12 . This shows that there is satisfied assumption (iv).

On the other hand we obtain

0t|kn(t,s)|ds=0ts1+n(s2+t2)ds=12ln1+2nt21+nt212nln212ln2.

This proves that the function sequence (kn(t, s)) satisfies assumption (iii) with the constant K1 = 12 ln 2.

Now, let us take into account the function tfn(t, x1, x2, …) defined by formula (4.3) for n = 1, 2, …. Fix arbitrary t1, t2 ∈ ℝ+ and x = (xn) ∈ l. Then, we get

|fn(t2,x1,x2,)fn(t1,x1,x2,)|β1n2+t221n2+t12β|t2t1|(t1+t2)(n2+t12)(n2+t22)=β|t2t1|t1(n2+t12)(n2+t22)+t2(n2+t12)(n2+t22)β|t2t1|121n2+t22+121n2+t12β|t2t1|

for any n = 1, 2, …. This shows that the functions fn (n = 1, 2, …) satisfy assumption (v).

In order to verify assumption (vi) let us fix a number r > 0 and choose x = (xi) ∈ l such that ∥xlr. Then, for arbitrarily fixed n ∈ ℕ and t ∈ ℝ+, we obtain

|fn(t,x1,x2,)|βn2+t2+y|xn|1+x12+|xn+1|n+x22βn2+t2+y(|xn|+|xn+1|)βn2+t2+2ysup{|xi|:in}.

This shows that the inequality from assumption (vi) is satisfied with the following functions

f¯n(t)=βn2+t2,l(r)=2y

for n = 1, 2, … . Since fn(t)⩽β/(1 + t2) we infer that limt fn(t) = 0 uniformly with respect to n ∈ ℕ. Apart from this we have that limn fn(t) = 0 for any t ∈ ℝ+.

Summing up we see that assumption (vi) is satisfied. Moreover, let us notice that

F¯=sup{f¯n(t):tR+,n=1,2,}=β.

Next, let us fix a number r > 0 and take arbitrary x = (xi), y = (yi) ∈ l such that ∥x∥lr, ∥y∥lr. Then, for a fixed n ∈ ℕ and t ∈ ℝ+, we get

|fn(t,x1,x2,)fn(t,y1,y2,)|yxn1+x12yn1+y12+yxn+1n+x22yn+1n+y22y|xn+xny12ynynx12|(1+x12)(1+y12)+y|nxn+1+xn+1y22nyn+1x22yn+1|(n+x22)(n+y22)y|xnyn|+y|(xny12yny12)+(yny12ynx12)|(1+x12)(1+y12)+yn|xn+1yn+1|(n+x22)(n+y22)+y|(xn+1y22yn+1y22)+(yn+1y22yn+1x22)|(n+x22)(n+y22)y|xnyn|+yy12|xnyn|(1+x12)(1+y12)+|yn||y12x12|(1+x12)(1+y12)+y|xn+1yn+1|+yy22|xn+1yn+1|(n+x22)(n+y22)+y|yn+1|(|x2|+|y2|)|y2x2|(n+x22)(n+y22)2y|xnyn|+yr|y1|(1+x12)(1+y12)+|x1|(1+x12)(1+y12)|x1y1|+2y|xn+1yn+1|+yr|x2y2||x2|(n+x22)(n+y22)+|y2|(n+x22)(n+y22)2y|xnyn|+yr|x1y1|+2y|xn+1yn+1|+yr|x2y2|(4y+2ry)xyl=2y(2+r)xyl.

Thus we see that assumption (vii) is satisfied with the function m(r) = 2y(2 + r).

In the next step of our proof we are going to verify assumption (viii). To this end fix arbitrarily n ∈ ℕ and consider the function gn(t, x) = gn(t, x1, x2, …) defined by formula (4.5) i.e.,

gn(t,x1,x2,)=arctanx1+xnn+t2.

Then, from the estimate

|gn(t,x1,x2,)||x1|+|xn|n+t2|x1|+|xn|n

we deduce that the operator g defined in assumption (viii) by the equality

(gx)(t)=(gn(t,x))=(g1(t,x),g2(t,x),)

transforms the set ℝ+ × l into l.

Further on, fix t ∈ ℝ+ and take x = (xi), y = (yi) ∈ l. Then we have

|gn(t,x)gn(t,y)|x1+xnn+t2y1+ynn+t2|x1y1|n+t2+|xnyn|n+t2|x1y1|n+|xnyn|n.

This allows us to derive the following estimate:

(gx)(t)(gy)(t)l=sup{|gn(t,x)gn(t,y)|:nN}sup|x1y1|n+|xnyn|n:nN2sup|xnyn|n:nN2xyl.

From the above estimate we infer that the operator g satisfies assumption (viii).

Moreover, it is easily seen that for an arbitrary xl and t ∈ ℝ+ we get

(gx)(t)l=sup{|gn(t,x)|:nN}π2.

This means that the operator g satisfies assumption (ix) with the constant G = π/2.

Finally, let us consider the first inequality from assumption (x). Obviously, in our case that inequality has the form

α22+π4ln2(β+2yr)<r. (4.6)

On the other hand, taking the second inequality required in assumption (x), we get

yπ2ln2(2+r0)<1. (4.7)

It is easy to check that choosing y<1πln2 and taking r0>α2+β2y, we can easily verify that both inequalities (4.6) and (4.7) are satisfied.

Thus, in the light of Theorem 3.4 we infer that infinite system of nonlinear integral equations (4.1) has at least one solution belonging to the ball Br0 in the space BC(ℝ+, l).

References

[1] J. Banaś and A. Chlebowicz, On solutions of an infinite system of nonlinear integral equations on the real half-axis, Banach Journal of Mathematical Analysis 13 (2019), no. 4, 944–968.10.1215/17358787-2019-0008Search in Google Scholar

[2] J. Banaś, A. Chlebowicz and W. Woś, On measures of noncompactness in the space of functions defined on the half-axis with values in a Banach space, Journal of Mathematical Analysis and Applications 489(2020), 12418710.1016/j.jmaa.2020.124187Search in Google Scholar

[3] J. Banaś and K. Goebel, Measures of Noncompactness in Banach Spaces, Lecture Notes in Pure and Applied Mathematics, vol. 60, Marcel Dekker Inc., New York, 1980.Search in Google Scholar

[4] J. Banaś and M. Mursaleen, Sequence Spaces and Measures of Noncompactness with Applications to Differential and Integral Equations, Springer, New Delhi, 2014.10.1007/978-81-322-1886-9Search in Google Scholar

[5] J. Banaś, M. Mursaleen and S.M.H. Rizvi, Existence of solutions to a boundary-value problem for an infinite system of differential equations, Electronic Journal of Differential Equations 2017 (2017), no. 262, 1–12.Search in Google Scholar

[6] J. Banaś and B. Rzepka, On solutions of infinite systems of integral equations of Hammerstein type, Journal of Nonlinear and Convex Analysis 18 (2017), no. 2, 261–278.Search in Google Scholar

[7] K. Deimling, Nonlinear Functional Analysis, Springer, Berlin, 1985.10.1007/978-3-662-00547-7Search in Google Scholar

[8] M. Mursaleen and S.M.H. Rizvi, Solvability of infinite systems of second order differential equations in c0 and l1 by Meir-Keeler condensing operators, Proceedings of the American Mathematical Society 144 (2016), no. 10, 4279–4289.10.1090/proc/13048Search in Google Scholar

[9] B. Rzepka and K. Sadarangani, On solutions of an infinite system of singular integral equations, Mathematical and Computer Modelling 45 (2007), no. 9-10, 1265–1271.10.1016/j.mcm.2006.11.006Search in Google Scholar

[10] K. Deimling, Ordinary Differential Equations in Banach Spaces, Lecture Notes in Mathematics 596, Springer, Berlin, 1977.10.1007/BFb0091636Search in Google Scholar

[11] K. Kuratowski, Sur les espaces complets, Fundamenta Mathematicae 15 (1930), no. 1, 301–309.10.4064/fm-15-1-301-309Search in Google Scholar

[12] L.S. Goldenštein, I.T. Gohberg and A.S. Markus, Investigations of some properties of bounded linear operators with their q-norms, Uč en. Zap. Kishinevsk. Univ. 29 (1957), 29–36.Search in Google Scholar

[13] L.S. Goldenštein and A.S. Markus, On a measure of noncompactness of bounded sets and linear operators, Studies in Algebra and Mathematical Analysis 1 (1965), 45–54.Search in Google Scholar

[14] R.R. Akhmerov, M.I. Kamenskii, A.S. Potapov, A.E. Rodkina and B.N. Sadovskii, Measure of Noncompactness and Condensing Operators. Operator Theory: Advanced and Applications vol. 55, Birkhäuser Verlag, Basel, 1992.10.1007/978-3-0348-5727-7Search in Google Scholar

[15] J.M. Ayerbe Toledano, T. Dominguez Benavides and G. Lopez Acedo, Measures of Noncompactness in Metric Fixed Point Theory, Birkhäuser Verlag, Basel, 1997.10.1007/978-3-0348-8920-9Search in Google Scholar

[16] G. Darbo, Punti uniti in trasformazioni a codominio non compatto, Rendiconti del Seminario Matematico della Università di Padova 24 (1955), 84–92.Search in Google Scholar

Received: 2020-03-07
Accepted: 2020-04-16
Published Online: 2020-06-19

© 2021 J. Banaś and W. Woś, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0114/html
Scroll to top button