Home Solvability of an infinite system of nonlinear integral equations of Volterra-Hammerstein type
Article Open Access

Solvability of an infinite system of nonlinear integral equations of Volterra-Hammerstein type

  • Agnieszka Chlebowicz EMAIL logo
Published/Copyright: December 6, 2019

Abstract

The purpose of the paper is to study the solvability of an infinite system of integral equations of Volterra-Hammerstein type on an unbounded interval. We show that such a system of integral equations has at least one solution in the space of functions defined, continuous and bounded on the real half-axis with values in the space l1 consisting of all real sequences whose series is absolutely convergent. To prove this result we construct a suitable measure of noncompactness in the mentioned function space and we use that measure together with a fixed point theorem of Darbo type.

1 Introduction

Integral equations create a very significant part of nonlinear analysis and applied mathematics ([1, 2, 3, 4]). Many researchers, not only mathematicians, are interested in the study of the solvability of integral equations and the applicability of such equations to different problems arising in the description of real world events ([2, 3, 5, 6, 7, 8, 9]). The results obtained in the theory of integral equations are useful and widely utilized in many branches of technical sciences, as mechanics or engineering and exact sciences as physics, for example.

In the theory of integral equations the special and exceptional branch is created by infinite systems of integral equations. On the one hand such systems are very interesting subject of the study for researchers specialized in the theory of integral equations but on the other hand systems of integral equations play very crucial role in applications.

In this paper we deal with an infinite system of nonlinear integral equations of Volterra-Hammerstein type. In [10] we showed that such a system has a solution belonging to the space BC(ℝ+, c0) of functions defined, continuous and bounded on the real half-axis and with values in the sequence space c0. In the present paper we prove a stronger result. Namely, we show that infinite system of integral equations of Volterra-Hammerstein type has at least one solution in the space BC(ℝ+, l1) consisting of all functions defined, continuous and bounded on the interval ℝ+ with values in the sequence space l1. Of course each such solution belongs to the space BC(ℝ+, c0) considered in [10].

Let us mention that paper [10] was the first one in which the study of solvability of infinite systems of integral equations defined on an unbounded interval was carried out. All known up to now results have been obtained in the space of functions defined on a bounded interval (see [11, 12, 13, 14], for example).

2 Notation, definitions and auxiliary facts

In this section we recall some facts which will be utilized in the paper.

Let us start with establishing some notation. The symbols ℝ and ℕ stand for sets of real and natural numbers, respectively. Moreover, we put ℝ+ = [0, ∞).

The letter E means a Banach space normed by norm ∥⋅∥E and with zero vector θ. The symbol B(x, r) denotes the closed ball in E centered at x with radius r. In the special case when x = θ we write Br instead of B(θ, r). Moreover, if X is a subset of E then we denote by X the closure of X and by Conv X the closed convex hull of the set X. The symbols X + Y, λX(λ ∈ ℝ) will stand for the usual algebraic operations on subsets X and Y of E. For a nonempty and bounded set XE we denote by diam X the diameter of the set X. The symbol ∥XE will stand for the norm of the set XE i.e., we have

||X||E=sup{||x||E:xX}.

The fundamental notion that we use in this paper is the concept of a measure of noncompactness. We recall now the definition of a measure of noncompactness which was introduced in monograph [15]. To this end let us denote by 𝔐E the family of all nonempty and bounded subsets of E and by 𝔑E its subfamily consisting of all relatively compact sets.

Definition 2.1

A function μ : 𝔐E → ℝ+ will be called a measure of noncompactness in 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(λX + (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=1 Xn is nonempty.

The set ker μ defined in axiom (i) is called the kernel of the measure of noncompactness μ.

Let us observe that the intersection set X appearing in axiom (vi) is a member of the family ker μ[15]. This simple observation plays an essential role in our further considerations.

Now we present some properties of measures of noncompactness [15]. We say that μ is sublinear if it satisfies the following additional conditions:

(vii) μ(X + Y) ≤ μ(X) + μ(Y).

(viii) μ (λX) = |λ| μ(X) for λ ∈ ℝ.

Moreover, we say that a measure of noncompactness μ has maximum property if

(ix) μ(XY) = max{μ(X), μ(Y)}.

If the measure of noncompactness μ is such that

(x) ker μ = 𝔑E

then it is called full.

A sublinear measure of noncompactness which is full and has maximum property is said to be regular measure of noncompactness [15].

Let us mention that the first measure of noncompactness was defined in 1930 by Kuratowski [16] in the following way

α(X)=inf{ε>0:X can be covered by a finite family of sets X1,X2,,Xm such that diam Xiε for i=1,2,,m}.

The measure α(X) is called the Kuratowski measure of noncompactness. It is known (see [15]) that the Kuratowski measure of noncompactness is a regular measure.

In the similar way was defined the Hausdorff measure of noncompactness ([17, 18]):

χ(X)=inf{ε>0:X has a finite εnet in E}.

It can be shown that the measure χ(X) is also regular and that both defined above measures α(X) and χ(X) are equivalent. But despite of these similarities it turns out that the Hausdorff measure of noncompactness χ is more convenient in applications than the Kuratowski measure. The main reason is that in some classical Banach spaces we can find explicit formulas describing the Hausdorff measure of noncompactness but we do not know such formulas for the Kuratowski measure of noncompactness in any Banach space [15].

And now, taking into account our further investigations, we recall the formula expressing the Hausdorff measure of noncompactness in the space l1 (cf. [15]). So, let us call to mind that the space l1 consists of all sequences whose series is absolutely convergent, i.e.

l1={x=(xn)R:n=1|xn|<}.

It is normed by the following norm

||x||l1=||(xn)||l1=n=1|xn|.

Then we have the following formula for the Hausdorff measure of noncompactness of any bounded set X ∈ 𝔐l1:

χ(X)=limn{sup{k=n|xk|:x=(xi)X}}. (2.1)

To prove our main result we will also need a fixed point theorem involving a measure of noncompactness. The basic theorem in this subject is known fixed point theorem proved by Darbo [19]. We will use a modified version of Darbo theorem formulated below (cf. [15]).

Theorem 2.2

Let μ be an arbitrary measure of noncompactness in the Banach space E. Assume that Ω is a nonempty, bounded, closed and convex subset of 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 the operator Q has at least one fixed point in the set Ω.

3 Measures of noncompactness in the space BC(ℝ+, l1)

In [10] we investigated measures of noncompactness in the space BC(ℝ+, E) consisting of all functions defined, continuous and bounded on ℝ+ with values in a fixed Banach space E. Let us pay attention to the fact that the space BC(ℝ+, E) is a generalization of the well-known and often used classical Banach space BC(ℝ+, ℝ), therefore measures of noncompactness in the space BC(ℝ+, E) must be more complicated then known measures in BC(ℝ+, ℝ). And now we recall some basic facts about the space BC(ℝ+, E) and measures of noncompactness in this space.

Let us start with assuming that E is a given Banach space with the norm ∥⋅∥E whereby we will assume that E is an infinite dimensional space. Consider the Banach space BC(ℝ+, E) equipped with the supremum norm ∥x defined in the standard way

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

For further purposes, for a fixed T > 0 consider the Banach space CT = C([0, T], E) consisting of functions defined and continuous on the interval [0, T] with values in E and endowed with the norm

||x||T=sup{||x(t)||E:t[0,T]}.

In [10] we defined three formulas for measures of noncompactness in the space BC(ℝ+, E) and each such formula is a sum of three components. The first and the second component are the same in each formula and we start to present them. So, let us fix a set X ∈ 𝔐BC(ℝ+,E) and numbers T > 0 and ε > 0. For any function xX we define the modulus of continuity ωT(x, ε) of the function x on the interval [0, T] by the classical formula

ωT(x,ε)=sup{||x(t)x(s)||E:t,s[0,T],|ts|ε}.

Next, let us define

ωT(X,ε)=sup{ωT(x,ε):xX},ω0T(X)=limε0ωT(X,ε)

and finally, we put

ω0(X)=limTω0T(X).

Notice that both above limits exist (for details see [10]). The quantity ω0(X) is the first component of each of mentioned earlier measures of noncompactness in BC(ℝ+, E). Next, to obtain the second component, assume that y = y(X) is a measure of noncompactness in the space E. Fix number t ∈ ℝ+ and denote by X(t) the so-called cross-section of the set X at the point t:

X(t)={x(t):xX}.

Further, for a fixed T > 0 let us put

y¯T(X)=sup{y(X(t)):t[0,T]} (3.1)

and

y¯(X)=limTy¯T(X). (3.2)

Notice that the above limit exists (see [10]). The obtained quantity y(X) is the second component of all three formulas for measure of noncompactness in the space BC(ℝ+, E). Now we introduce the third component of the measure of noncompactness in BC(ℝ+, E) which describes the behaviour of the set of functions at infinity. We can do it in three ways and we will describe each of them. So, for a fixed T > 0 let us define

aT(X)=supxX{sup{||x(t)||E:tT}}. (3.3)

Next, notice that there exists the limit

a(X)=limTaT(X) (3.4)

and the quantity a(X) is considered as the third component of the measure of noncompactness in the space BC(ℝ+, E).

Instead of a(X) we can take

b(X)=limTbT(X), (3.5)

where

bT(X)=supxX{sup{||x(t)x(s)||E:t,sT}}

or we can also define

c(X)=lim suptdiam X(t), (3.6)

where diam X(t) is understood as

diam X(t)=sup{||x(t)y(t)||E:x,yX}.

And now let us consider the functions ya, yb, yc defined on the family 𝔐BC(ℝ+,E) as follows

ya(X)=ω0(X)+y¯(X)+a(X), (3.7)
yb(X)=ω0(X)+y¯(X)+b(X), (3.8)
yc(X)=ω0(X)+y¯(X)+c(X). (3.9)

In [10] we proved that under some assumptions on y the functions ya, yb, yc are measures of noncompactness in the space BC(ℝ+, E). More precisely, we have the following result.

Theorem 3.1

Assume that y is the Hausdorff measure of noncompactness in the Banach space E i.e., y = χE. Then, the functions χa, χb and χc defined by (3.7) - (3.9) are measures of noncompactness in the space BC(ℝ+, E) such that

χ(X)2χb(X),χ(X)4χc(X),χb(X)2χa(X),χc(X)2χa(X)

for an arbitrary set X ∈ 𝔐BC(ℝ+,E), where χ denotes the Hausdorff measure of noncompactness in the space BC(ℝ+, E).

For other properties of the above introduced measures of noncompactness we refer to [10]. We recall only that Theorem 3.1 remains valid if in the construction of the component y we replace the Hausdorff measure of noncompactness χ by an arbitrary regular measure of noncompactness μ equivalent to the Hausdorff measure χ[10].

Now, we are going to present formulas (3.7), (3.8) and (3.9) in the special case, for E = l1. The space l1 is one of the Banach sequence spaces and we deal with this space (and, in general, with sequence spaces) because it is strictly connected with the form of solutions of infinite systems of integral equations (see Theorem 4.2).

Therefore, we will work in the Banach space

BC(R+,l1)={x:R+l1:x is continuous and bounded on R+}.

The fundamental fact in our investigations is that every function xBC(ℝ+, l1) can be regarded as a function sequence

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

for t ∈ ℝ+, where for any fixed t the sequence (xn(t)) is a real sequence being an element of the space l1. Obviously, it means that

||x(t)||l1=n=1|xn(t)|<.

According to the formula for the norm in the space BC(ℝ+, E) given earlier we have

||x||=sup{||x(t)||l1:tR+}=sup{n=1|xn(t)|:tR+}.

Now, we are going to present explicitly the consecutive components ω0(X), χ(X) and a(X) of the measure of noncompactness χa(X) for any set X ∈ 𝔐BC(ℝ+,l1). So, let us start with ω0(X). Fix arbitrarily numbers T > 0 and ε > 0. For any x = x(t) = (xn(t)) ∈ X we have

ωT(x,ε)=sup{||x(t)x(s)||l1:t,s[0,T],|ts|ε}=sup{n=1|xn(t)xn(s)|:t,s[0,T],|ts|ε}.

Hence, we get

ωT(X,ε)=supx=(xn)X{sup{n=1|xn(t)xn(s)|:t,s[0,T],|ts|ε}}

and next

ω0T(X)=limε0{supx=(xn)X{sup{n=1|xn(t)xn(s)|:t,s[0,T],|ts|ε}}}.

Finally, we obtain

ω0(X)=limT{limε0{supx=(xn)X{sup{n=1|xn(t)xn(s)|:t,s[0,T],|ts|ε}}}}. (3.10)

Next, we are going to define the second component occuring in the formula for the measure χa(X). To this end let us assume that X ∈ 𝔐BC(ℝ+,l1) and t ∈ ℝ+ is arbitrarily fixed. Using (2.1) we have

χ(X(t))=limn{supx=(xk)X{k=n|xk(t)|}}.

Next, for a fixed T > 0 utilizing (3.1) we get

χ¯T(X)=sup{χ(X(t)):t[0,T]}=supt[0,T]limn{sup(xk(t))X(t){k=n|xk(t)|}}.

Finally, in view of (3.2) we obtain the following expression:

χ¯(X)=limTχ¯T(X)=limTsupt[0,T]{limn{supx=(xk)X{k=n|xk(t)|}}}. (3.11)

And now let us write the third component of the measure χa(X). Thus, fix an arbitrary number T > 0. Then, on the basis of (3.3), we have

aT(X)=supx=(xn)X{sup{||x(t)||l1:tT}}=supx=(xn)X{sup{n=1|xn(t)|:tT}}.

Next, by (3.4) we obtain

a(X)=limTsupx=(xn)X{suptT{n=1|xn(t)|}}. (3.12)

Finally, based on Theorem 3.1 we get that the function

χa(X)=ω0(X)+χ¯(X)+a(X)

is a measure of noncompactness in the space BC(ℝ+, l1), where ω0(X), χ(X) and a(X) are given by formulas (3.10), (3.11) and (3.12), respectively.

Observe, that keeping in mind formulas (3.5) and (3.6) together with Theorem 3.1 we obtain that the functions

χb(X)=ω0(X)+χ¯(X)+b(X)

and

χc(X)=ω0(X)+χ¯(X)+c(X)

are also measures of noncompactness in the space BC(ℝ+, l1), where

b(X)=limT{supx=(xn)X{sup{n=1|xn(t)xn(s)|:t,sT}}}

and

c(X)=lim supt{sup{||x(t)y(t)||l1:x,yX}}=lim supt{sup{n=1|xn(t)xn(s)|:x=(xn),y=(yn)X}}.

4 Theorem on the existence of solutions of infinite systems of integral equations on the real half-axis

Let us consider the following infinite system of nonlinear quadratic integral equations of the Volterra-Hammerstein type

xn(t)=an(t)+fn(t,xn(t),xn+1(t),)0tkn(t,s)gn(s,x1(s),x2(s),)ds (4.1)

for t ∈ ℝ+ and for n = 1, 2, ….

In [10] we proved that system (4.1) has at least one solution in the space BC(ℝ+, c0) := {x : ℝ+c0 : x is continuous and bounded on ℝ+}. In this paper we prove the other result, namely we show that the system (4.1) has at least one solution in the space BC(ℝ+, l1). For the convenience, from now on the space BC(ℝ+, l1) will be denoted by BC1.

At the beginning we provide a lemma which we will utilize in the proof of our main result.

Lemma 4.1

If the sequence (an) belongs to the space l1 then limni=n |ai| = 0.

Proof

The proof is an immediate consequence of the Cauchy condition for real sequences.□

In what follows we will investigate the solvability of system (4.1) in the space BC1 under the below listed assumptions.

  1. The function sequence (an(t)) is an element of the space BC1 such that limtn=1 an(t) = 0.

  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 locally equicontinuous on the set ℝ+ uniformly with respect to s ∈ ℝ+ i.e., the following condition is satisfied

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

    n=10t|kn(t,s)|dsK1

    for any t ∈ ℝ+.

  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 n=1 fn is defined on the set ℝ+ × ℝ and takes real values. Moreover, the function t n=1 fn(t, xn, xn+1, …) is continuous on ℝ+ uniformly with respect to x = (xn) ∈ l1 i.e., the following condition is satisfied

    ε>0t0R+δ>0(xi)l1tR+[|tt0|δn=1|fn(t,xn,xn+1,)fn(t0,xn,xn+1,)|ε].
  6. There exists a function l : ℝ+ → ℝ+ such that l is nondecreasing on ℝ+, continuous at 0 and the following condition is satisfied

    |fn(t,xn,xn+1,)fn(t,yn,yn+1,)|l(r)|xnyn|

    for all x = (xi), y = (yi) ∈ l1 such that ∥xl1r, ∥yl1r, for any t ∈ ℝ+ and for n = 1, 2, ….

  7. The sequence of functions (fn), where fn(t) = |fn(t, 0, 0, …)| belongs to the space BC1 and limtn=1 fn(t) = 0.

    Let us notice that on the basis of assumption (vii) we infer that there exists the finite constant F = sup { n=1 fn(t) : t ∈ ℝ+}.

    Now, we can formulate our further assumptions concerning infinite system (4.1).

  8. 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 space ℝ+ × l1 by the equality

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

    transforms the space ℝ+ × l1 into l1 and is such that the family of functions {(gx)(t)}t∈ℝ+ is equicontinuous at every point of the space l1 i.e., for each arbitrarily fixed xl1 and for a given ε > 0 there exists δ > 0 such that

    ||(gy)(t)(gx)(t)||l1ε

    for every t ∈ ℝ+ and for any yl1 such that ∥yxl1δ.

  9. The operator g defined in assumption (viii) is bounded on the space ℝ+ × l1. More precisely, there exists a positive constant G such that ∥(gx)(t)∥l1G for any xl1 and for each t ∈ ℝ+.

  10. There exists a positive solution r0 of the inequality

    A+F¯G¯K1+G¯K1rl(r)r

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

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

Now we can present our main result concerning the solvability of infinite system of integral equations (4.1).

Theorem 4.2

Under assumptions (i) - (x) infinite system (4.1) has at least one solution x(t) = (xn(t)) in the space BC1 = BC(ℝ+, l1).

Proof

Denote by F, V, Q operators defined on the space BC1 in the following way:

(Fx)(t)=((Fnx)(t))=(fn(t,x(t)))=(fn(t,xn(t),xn+1(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))

for an arbitrary element x = (xn) ∈ BC1 and for t ∈ ℝ+.

Our proof will be conducted in several steps. At the begining we show that operator Q transforms the space BC1 into itself. Thus, let us take x = x(t) = (xn(t)) ∈ BC1. Obviously this means that n=1 |xn(t)| < ∞.

Then, for arbitrary t ∈ ℝ+, using assumption (vi), we get

n=1|(Fnx)(t)|=n=1|fn(t,xn(t),xn+1(t),)|n=1|fn(t,xn(t),xn+1(t),)fn(t,0,0,)|+n=1|fn(t,0,0,)|n=1l(||x(t)||l1)|xn(t)0|+n=1f¯n(t)l(||x(t)||l1)n=1|xn(t)|+n=1f¯n(t). (4.2)

Keeping in mind the fact that xBC1, in view of assumption (vii) we obtain

n=1|(Fnx)(t)|<. (4.3)

Thus we have that (Fx)(t) ∈ l1. Moreover, from (4.2) we infer that the function $Fx$ is bounded on the set ℝ+.

Further, we are going to show that the function Fx is continuous on the interval ℝ+. To this end, let us take arbitrary t0 ∈ ℝ+ and ε > 0. It follows from the continuity of the function x that the below given condition is satisfied

t0R+ε>0δ>0tR+[|tt0|δ||x(t)x(t0)||l1ε]. (4.4)

Thus, let us choose δ1 > 0 according to (4.4). Then, for t ∈ ℝ+ such that |tt0| ≤ δ1 we obtain

n=1|(Fnx)(t)(Fnx)(t0)|=n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t0),xn+1(t0),)|n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t),xn+1(t),)|+n=1|fn(t0,xn(t),xn+1(t),)fn(t0,xn(t0),xn+1(t0),)|n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t),xn+1(t),)|+n=1l(sup{||x(t)||l1:tR+})|xn(t)xn(t0)|n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t),xn+1(t),)|+l(sup{||x(t)||l1:tR+})n=1|xn(t)xn(t0)|n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t),xn+1(t),)|+l(||x||BC1)ε. (4.5)

Next, let us choose a number δ2 > 0 according to assumption (v). Then for |tt0| ≤ δ2 we have

n=1|fn(t,xn(t),xn+1(t),)fn(t0,xn(t),xn+1(t),)|ε.

Thus, taking δ = min{δ1, δ2} we can deduce that the following estimate is satisfied

n=1|(Fnx)(t)(Fnx)(t0)|ε+l(||x||BC1)ε=ε(1+l(||x||BC1))

for any t ∈ ℝ+. Therefore, we can write

||(Fx)(t)(Fx)(t0)||ε

for any t ∈ ℝ+. It means that Fx is continuous on ℝ+.

Linking above established facts we obtain that the operator F transforms the space BC1 into itself.

Next, we are going to show that the operator V maps the space BC1 into BC1. So, let us take an arbitrary function x = x(t) = (xn(t)) ∈ BC1. We are going to prove that a function VxBC1. We start with showing the boundness of the function Vx on ℝ+. To this end observe that for an arbitrary number t ∈ ℝ+, using assumptions (iii) and (ix), we get

n=1|(Vnx)(t)|=n=1|0tkn(t,s)gn(s,x1(s),x2(s),)ds|n=10t|kn(t,s)||gn(s,x1(s),x2(s),)|dsn=10t|kn(t,s)|G¯dsG¯n=10t|kn(t,s)|dsG¯K1. (4.6)

Therefore we have that for any t ∈ ℝ+ the following inequality holds

||(Vx)(t)||l1G¯K1.

This means the function Vx is bounded on ℝ+.

To prove the continuity of the function Vx on the interval ℝ+ let us fix ε > 0, T > 0 and t0 ∈ [0, T). In virtue of (4.4) we can choose a number δ > 0 according to the continuity of the function x = x(t) at the point t0. Next, take t ∈ [0, T) satisfying |tt0| ≤ δ (without loss of generality we can assume that t > t0). Then, keeping in mind assumptions (iv), (viii) and (ix) and using the Lebesgue monotone convergence theorem [20], we arrive at the following estimates:

n=1|(Vnx)(t)(Vnx)(t0)|=n=1|0tkn(t,s)gn(s,x1(s),x2(s),)ds0t0kn(t0,s)gn(s,x1(s),x2(s),)ds|n=1|0tkn(t,s)gn(s,x1(s),x2(s),)ds0tkn(t0,s)gn(s,x1(s),x2(s),)ds|+n=1|0tkn(t0,s)gn(s,x1(s),x2(s),)ds0t0kn(t0,s)gn(s,x1(s),x2(s),)ds|n=10t|kn(t,s)kn(t0,s)||gn(s,x1(s),x2(s),)|ds+n=1t0t|kn(t0,s)||gn(s,x1(s),x2(s),)|dsn=10tωkT(ε)|gn(s,x1(s),x2(s),)|ds+n=1t0tK2|gn(s,x1(s),x2(s),)|dsωkT(ε)n=10t|gn(s,x1(s),x2(s),)|ds+K2n=1t0t|gn(s,x1(s),x2(s),)|ds=ωkT(ε)0tn=1|gn(s,x1(s),x2(s),)|ds+K2t0tn=1|gn(s,x1(s),x2(s),)|ds=ωkT(ε)0t||(gn(s,x))||l1ds+K2t0t||gn(s,x)||l1dsωkT(ε)G¯T+K2G¯ε, (4.7)

where ωkT (ε) denotes a common modulus of continuity of the function sequence tkn(t, s) on the set [0, T] (according to assumption (ii)). Obviously ωkT (ε) → 0 as ε → 0.

Hence, on the basis of (4.7) we derive the following inequality

||(Vx)(t)(Vx)(t0)||l1ωkT(ε)G¯T+K2G¯ε,

which means that the function Vx is continuous on [0, T). The number T was choosen arbitrarily therefore we deduce that the function Vx is continuous on the real half-axis ℝ+.

Futher on, we are going to prove that the function Q maps the space BC1 into itself. In order to prove this fact, notice that we can treat the space BC1 as a Banach algebra with respect to the coordinatewise multiplication of sequences. Therefore, take any function xBC1 and consider a function Qx. Keeping in mind the definition of the operator Q and established facts that the function Fx and the function Vx are continuous on ℝ+ we obtain that the function Qx is also continuous on ℝ+. Similarly, taking into account the boundness of functions Fx and Vx on the set ℝ+ we infer that Qx is also bounded on ℝ+. In order to show that Qx : ℝ+l1 let us notice that using asumption (i) and (4.3) we have

n=1|(Qnx)(t)|n=1|an(t)|+G¯K1n=1|(Fnx)(t)|<,

so for any fixed t ∈ ℝ+ we obtain that (Qx)(t) ∈ l1 and hence Qx : ℝ+l1.

Finally, linking all the above established properties of the function Qx we derive that the operator Q transforms the space BC1 into itself.

In what follows we show the existence of a number r0 > 0 such that Q transforms the ball Br0 in the space BC1 into itself. For arbitrary t ∈ ℝ+, utilizing estimates (4.2) and (4.6) as well as assumptions (x) and (vi) we obtain

n=1|(Qnx)(t)|n=1|an(t)|+n=1[|(Fnx)(t)||(Vnx)(t)|]n=1|an(t)|+n=1|(Fnx)(t)|n=1|(Vnx)(t)|A+[l(||x(t)||l1)n=1|xn(t)|+n=1f¯n(t)]G¯K1A+[l(||x(t)||l1)||x(t)||l1+n=1f¯n(t)]G¯K1A+[l(||x||BC1)||x||BC1+F¯]G¯K1=A+F¯G¯K1+G¯K1l(||x||BC1)||x||BC1.

This yields to estimate

||Qx||BC1A+F¯G¯K1+G¯K1l(||x||BC1)||x||BC1.

Taking into account the last inequality and assumption (x) we deduce that there exists a number r0 > 0 such that the operator Q transforms the ball Br0 into itself.

In what follows we show that the operator Q is continuous on the ball Br0. In order to prove this fact fix arbitrarily xBr0, ε > 0 and take a function yBr0 such that ∥xyBC1ε. Fix t ∈ ℝ+. Then, in view of assumption (vi) we have

n=1|(Fnx)(t)(Fny)(t)|=n=1|fn(t,xn(t),xn+1(t),)fn(t,yn(t),yn+1(t),)|n=1l(r0)|xn(t)yn(t)|=l(r0)n=1|xn(t)yn(t)|=l(r0)||x(t)y(t)||l1l(r0)||xy||BC1l(r0)ε,

so we get

||FxFy||BC1l(r0)ε.

This means that F is continuous on the ball Br0.

Further, let us consider the function δ = δ(ε) defined for ε > 0 in the following way

δ(ε)=sup{|gn(t,x)gn(t,y)|:x,yl1,||xy||l1ε,tR+,nN}.

Taking into account assumption (viii) we deduce that δ(ε) → 0 as ε → 0.

Further, let us assume, similarly as above, that xBr0, ε > 0 and yBr0 are such that ∥xyBC1ε. Then, for a fixed t ∈ ℝ+ we get

n=1|(Vnx)(t)(Vny)(t)|n=10t|kn(t,s)||gn(s,x1(s),x2(s),)gn(s,y1(s),y2(s),)|dsn=10t|kn(t,s)|δ(ε)dsK1δ(ε).

Consequently, we obtain

||VxVy||BC1K1δ(ε).

This proves the continuity of the operator V on the ball Br0.

Now, linking the continuity of the operators F and V on the ball Br0 and keeping in mind the representation of the operator Q written at the beginning of the proof we infer that the operator Q is continuous on Br0.

And now we have the last step of our proof in which we show that the inequality from Theorem 2.2 is satisfied for any set XBr0 and for measure of noncompactness χa defined by formula (3.7) for y = χl1.

To this end take a nonempty subset X of the ball Br0 and fix numbers ε > 0 and T > 0. Choose t, s ∈ [0, T] such that |ts| ≤ ε and consider a function x = x(t) = (xn(t)) ∈ X. Then, proceeding similarly as in (4.5) we obtain the following estimate:

n=1|(Fnx)(t)(Fnx)(s)|l(||x||BC1)n=1|xn(t)xn(s)|+n=1|fn(t,xn(t),xn+1(t),)fn(s,xn(t),xn+1(t),)|l(r0)n=1|xn(t)xn(s)|+sup{n=1|fn(t,xn,xn+1,)fn(s,xn,xn+1,)|:t,s[0,T],|ts|ε,||x||l1=||(xn)||l1r0}l(r0)ωT(x,ε)+ω1(f,ε), (4.8)

where

ω1(f,ε)=sup{n=1|fn(t,xn,xn+1,)fn(s,xn,xn+1,)|:t,s[0,T],|ts|ε,||x||l1=||(xn)||l1r0}.

Obviously, in view of assumption (v) we have that ω1(f, ε) → 0 as ε → 0.

Taking supremum with respect to t, s ∈ [0, T], |ts| ≤ ε on the left side of (4.8), we get the following estimate

ωT(Fx,ε)l(r0)ωT(x,ε)+ω1(f,ε). (4.9)

Next, let us take t, s as above. Assuming additionaly that t > s and following the estimate (4.7) we get

n=1|(Vnx)(t)(Vnx)(s)|TG¯ωkT(ε)+G¯K2ε,

where the function ωkT (ε) was introduced earlier.

Consequently this implies the following inequality:

ωT(Vx,ε)TG¯ωkT(ε)+G¯K2ε. (4.10)

Now, take any function xX and t, s ∈ ℝ+. Keeping in mind the representation of the operator Q we derive the following estimate:

||(Qx)(t)(Qx)(s)||l1||a(t)a(s)||l1+||(Fx)(t)(Vx)(t)(Fx)(s)(Vx)(s)||l1||a(t)a(s)||l1+||(Fx)(t)(Vx)(t)(Vx)(t)(Fx)(s)+(Vx)(t)(Fx)(s)(Fx)(s)(Vx)(s)||l1||a(t)a(s)||l1+||(Fx)(t)(Vx)(t)(Vx)(t)(Fx)(s)||l1+||(Vx)(t)(Fx)(s)(Fx)(s)(Vx)(s)||l1||a(t)a(s)||l1+||(Vx)(t)||l1||(Fx)(t)(Fx)(s)||l1+||(Fx)(s)||l1||(Vx)(t)(Vx)(s)||l1,

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

Further on we derive the last estimate utilizing inequalities (4.9), (4.10), (4.2) and (4.6) and assuming that t, s ∈ [0, T], |ts| ≤ ε:

ωT(Qx,ε)ωT(a,ε)+G¯K1ωT(Fx,ε)+(l(r0)r0+F¯)ωT(Vx,ε)ωT(a,ε)+G¯K1ωT(Fx,ε)+(l(r0)r0+F¯)(TG¯ωkT(ε)+G¯K2ε)ωT(a,ε)+G¯K1{l(r0)ωT(x,ε)+ω1(f,ε)}+(l(r0)r0+F¯)(TG¯ωkT(ε)+G¯K2ε).

Hence we obtain the following inequality:

ωT(QX,ε)ωT(a,ε)+G¯K1{l(r0)ωT(X,ε)+ω1(f,ε)}+(l(r0)r0+F¯)(TG¯ωkT(ε)+G¯K2ε).

Letting with ε → 0 in the above inequality and keeping in mind the properties of the functions εω1(f, ε) and ε ωkT (ε) we obtain

ω0T(QX)G¯K1l(r0)ω0T(X).

Finally, taking limit as T → ∞, we get

ω0(QX)G¯K1l(r0)ω0(X). (4.11)

Notice that we obtained the estimate for the first component ω0(X) of the measure of noncompactness χa(X) expressed by formula (3.7).

In what follows we obtain two consecutive estimations for the second and the third component of the measure of noncompactness χa(X). To this end, similarly as before, fix a set XBr0 and a function xX. Take an arbitrary number T > 0 and fix t ∈ [0, T]. Then, for any n ∈ ℕ, utilizing estimates (4.2) and (4.6) (for series from i = n to ∞), we get

i=n|(Qix)(t)|=i=n|ai(t)+(Fix)(t)(Vix)(t)|i=n|ai(t)|+i=n|(Fix)(t)(Vix)(t)|i=n|ai(t)|+i=n|(Fix)(t)|i=n|(Vix)(t)|i=n|ai(t)|+[l(||x(t)||l1)i=n|xi(t)|+i=nfi¯(t)]G¯K1i=n|ai(t)|+[l(r0)i=n|xi(t)|+i=nf¯i(t)]G¯K1.

Further, taking supremum over all x = (xi) ∈ X, we derive the following evaluation

supx=(xi)X{i=n|(Qix)(t)}i=n|ai(t)|+G¯K1l(r0)supx=(xi)X{i=n|xi(t)|}+G¯K1i=nf¯i(t).

Passing with n → ∞ and utilizing assumptions (i), (vii) and Lemma 4.1, we get

limn{supx=(xi)X{i=n|(Qix)(t)}}G¯K1l(r0)limn{supx=(xi)X{i=n|xi(t)|}}.

Finally, taking supremum over t ∈ [0, T] on both sides of the above inequality and letting with T → ∞, in view of formula (3.11) we deduce the following estimate

χ¯(QX)G¯K1l(r0)χ¯(X). (4.12)

Now, we are going to estimate the third component a(X) of the measure of noncompactness χa(X). Assume, as earlier, that XBr0, xX and T > 0. Moreover, take tT. Then, keeping in mind inequalities (4.2) and (4.6), we have

n=1|(Qnx)(t)|n=1|an(t)|+l(r0)G¯K1n=1|xn(t)|+G¯K1n=1f¯n(t).

Taking supremum over tT and x = (xn) ∈ X, we get

supx=(xn)X{suptT{n=1|(Qnx)(t)|}}suptTn=1|an(t)|+l(r0)G¯K1supx=(xn)X{suptT{n=1|xn(t)|}}+G¯K1suptT{n=1|f¯n(t)|}.

Letting with T → ∞ and utilizing assumptions (i) and (vii), we derive the following inequality

a(QX)l(r0)G¯K1a(X). (4.13)

Finally, combining (4.11), (4.12), (4.13) and formula (3.7) for y = χ, we get the following inequality

χa(QX)l(r0)G¯K1χa(X).

Hence, in view of the previously established fact that the operator Q is a continuous self-mapping of the ball Br0, utilizing Theorem 2.2 we conclude that the infinite system of Volterra-Hammerstein integral equations (4.1) has at least one solution x(t) = (xn(t)) in the space BC1 = BC(ℝ+, l1). Obviously xBr0.

The proof is complete.□

5 An example

This section is dedicated to present an example illustrating the applicability of the result contained in Theorem 4.2.

Example 5.1

Let us consider the following infinite system of integral equations having the form

xn(t)=β(1)n1tn1(n1)!+yn2sin(xn(t)+xn2(t))0t1(n2+s2)[1+(t+s)2]arctansn2[xn(s)1+n2xn2(s)+xn+1(s)1+(n+1)2xn+12(s)]ds (5.1)

for t ∈ ℝ+ and for n = 1, 2, …. Moreover, β and y are positive constants which will be specified later.

Let us observe that infinite system (5.1) is a particular case of the infinite system of quadratic integral equations of Volterra-Hammerstein type (4.1), where we put:

an(t)=β(1)n1tn1(n1)!,fn(t,xn,xn+1,)=yn2sin(xn+xn2),kn(t,s)=1(n2+s2)[1+(t+s)2],gn(t,x1,x2,)=arctantn2[xn1+n2xn2+xn+11+(n+1)2xn+12] (5.2)

for n = 1, 2, … and t, s ∈ ℝ+.

In what follows we show that the components of infinite system (5.1) defined by (5.2) satisfy assumptions of Theorem 4.2.

Indeed, for arbitrarily fixed n ∈ ℕ and t ∈ ℝ+ we have that the series

n=1an(t)=βn=1(1)n1tn1(n1)!

converges on the interval ℝ+ to the function a(t) = βet. Obviously

limtn=1an(t)=limtβet=0.

This shows that there is satisfied assumption (i). Apart from this we have that A = sup { n=1 |an(t)| : t ∈ ℝ+} = β.

Further, let us observe that the functions kn(t, s) are continuous on the set R+2 for n = 1, 2, ….

Next, fix n ∈ ℕ and s ∈ ℝ+. Then, for arbitrary t1, t2 ∈ ℝ+ we obtain

|kn(t2,s)kn(t1,s)|=1(n2+s2)|11+(t2+s)211+(t1+s)2|1n2|(t2+s)2(t1+s)2[1+(t2+s)2][1+(t1+s)2]|1n2|(t2t1)(t1+t2+2s)[1+(t2+s)2][1+(t1+s)2]|1n2|t2t1|[t1+s1+(t1+s)2+t2+s1+(t2+s)2]1n2|t2t1|(12+12)|t2t1|.

Hence we infer that the functions kn(t, s) are equicontinuous on the set ℝ+ uniformly with respect to the variable s ∈ ℝ+. Thus, these functions satisfy assumption (ii).

In what follows, for a fixed t ∈ ℝ+ we get

n=10t|kn(t,s)|ds=n=10t1n2+s211+(t+s)2dsn=11n20t11+(t+s)2ds=n=11n2t2tdu1+u2n=11n2arctan2tπ2π26=π312.

From the above estimate we deduce that the functions kn(t, s) satisfy assumption (iii) with the constant K1 = π3/12.

Now, taking arbitrary t, s ∈ ℝ+ and n ∈ ℕ, we have

|kn(t,s)|=1n2+s211+(t+s)21n21.

This means that assumption (iv) is satisfied with the constant K2 = 1.

Next, let us pay our attention to the functions fn = fn(t, xn, xn+1, …) defined by (5.2). Taking arbitrary n ∈ ℕ and an element x = (x1, x2, …) ∈ l1, we can easily infer that the first part of assumption (v) is satisfied which is a simple consequence of the fact that the series

n=1fn(t,xn,xn+1,)

is uniformly convergent on the set ℝ+ × ℝ. In fact, this conclusion follows immediately from the standard Weierstrass test. Obviously, the second part of assumption (v) is also trivially satisfied.

Now, fix arbitrarily a number r > 0 and take two elements x = (xn), y = (yn) belonging to the space l1 such that ∥xl1 = n=1 |xn| ≤ r, ∥yl1 = n=1 |yn| ≤ r. Then, for a fixed natural number n, we derive the following estimate:

|fn(t,xn,xn+1,)fn(t,yn,yn+1,)|=yn2|sin(xn+xn2)sin(yn+yn2)|yn2|(xnyn)+(xn2yn2)|yn2|xnyn|+yn2|xnyn|(|xn|+|yn|)yn2|xnyn|+yn2|xnyn|(n=1|xn|+n=1|yn|)2ymax{1,2r}|xnyn|.

Hence we infer that the functions fn(n = 1, 2, …) satisfy assumption (vi) with the function l(r) defined by the equality

l(r)=2ymax{1,2r}.

Moreover, we have that fn(t) = |fn(t, 0, 0, …)| = 0. Thus, there is satisfied assumption (vii) and F = 0.

Further on, we are going to verify assumption (viii). To this end, let us first notice that for a fixed natural number n the function gn = gn(t, x1, x2, …) defined by (5.2) on the set ℝ+ × ℝ takes real values (n = 1, 2, …).

Next, fix arbitrarily x = (xn) ∈ l1. Then, for t ∈ ℝ+ we obtain

n=1|gn(t,x1,x2,)|n=1arctantn2[|xn|1+n2xn2+|xn+1|1+(n+1)2xn+12]π2n=11n2(12+14)=3π8n=11n2=π316. (5.3)

From the above estimate it follows that the operator g defined in assumption (viii) transforms the space ℝ+ × l1 into l1. Moreover, from (5.3) we conclude that the operator g is bounded on ℝ+ × l1 and ∥(gx)(t)∥l1π3/16. Thus, the operator g satisfies assumption (ix) and we can accept that G = π3/16, where G is defined in assumption (ix).

Further, taking a number ε > 0 and choosing arbitrarily x = (xn), y = (yn) ∈ l1 such that ∥xyl1ε, for t ∈ ℝ+ we obtain

||(gy)(t)(gx)(t)||l1=n=1|gn(t,y1,y2,)gn(t,x1,x2,)|=n=1arctantn2|yn1+n2yn2+yn+11+(n+1)2yn+12xn1+n2xn2xn+11+(n+1)2xn+12|π2n=11n2[|yn1+n2yn2xn1+n2xn2|+|yn+11+(n+1)2yn+12xn+11+(n+1)2xn+12|]π2n=11n2[|yn+n2ynxn2xnn2xnyn2|(1+n2yn2)(1+n2xn2)+|yn+1+(n+1)2yn+1xn+12xn+1(n+1)2xn+1yn+12|1+(n+1)2yn+121+(n+1)2xn+12]π2n=11n2[|ynxn|+n2|xn||yn||ynxn|(1+n2yn2)(1+n2xn2)+|yn+1xn+1|+(n+1)2|xn+1||yn+1||yn+1xn+1|1+(n+1)2yn+121+(n+1)2xn+12]π2n=11n2{|ynxn|[1(1+n2yn2)(1+n2xn2)+n|yn|1+n2yn2n|xn|1+n2xn2]+|yn+1xn+1|[11+(n+1)2yn+121+(n+1)2xn+12+(n+1)|yn+1|1+(n+1)2yn+12(n+1)|xn+1|1+(n+1)2xn+12]}π2n=11n2[|ynxn|(1+1212)+|yn+1xn+1|(1+1212)]5π8n=11n2[|ynxn|+|yn+1xn+1|]5π8n=11n2k=1|ykxk|=5π348||yx||l15π348ε.

Thus we have proved that there is satisfied assumption (viii).

Finally, gathering all the above obtained constants A, F, G, K1 and taking into account the function l(r) = 2y max{1, 2r} indicated in the above calculations, we conclude that the inequality from assumption (x) has the form

β+yπ696rmax{1,2r}r. (5.4)

Further, let us assume that we are looking for a solution r of inequality (5.4) such that r 12 . In such a case inequality (5.4) has the form

β+yπ696rr. (5.5)

Assuming that y < 96/π6 we infer that, for example, the number r0 = 96β/(96 – 6) is a solution of inequality (5.4) provided β < (96 – yπ6)/192. It easily seen that in this case we have that GK1l(r0) < 1 which proves that assumption (x) is thoroughly satisfied.

Now, applying Theorem 4.2 we deduce that infinite system of integral equations (5.1) has at least one solution x = x(t) = (xn(t)) in the space BC1 = BC(ℝ+, l1).

References

[1] S. Chandrasekhar, Radiative Transfer, Dover, New York, 1960.Search in Google Scholar

[2] C. Corduneanu, Integral Equations and Applications, Cambridge University Press, Cambridge, 1991.10.1017/CBO9780511569395Search in Google Scholar

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

[4] N. Dunford and J. T. Schwartz, Linear Operators, I: General Theory, Wiley, New York, 1963.Search in Google Scholar

[5] W. Mydlarczyk, A system of Volterra integral equations with blowing up solutions, Colloq. Math. 146 (2017), 99–110.10.4064/cm6842-1-2016Search in Google Scholar

[6] W. Mydlarczyk, Coupled Volterra integral equations with blowing up solutions, J. Integral Equations Appl. 30 (2018), no. 1, 147–166.10.1216/JIE-2018-30-1-147Search in Google Scholar

[7] W. Okrasiński and Ł. Płociniczak, Solution estimates for a system of nonlinear integral equations arising in optometry, J. Integral Equations Appl. 30 (2018), no. 1, 167–179.10.1216/JIE-2018-30-1-167Search in Google Scholar

[8] W. Pogorzelski, Integral Equations and Their Applications, I, Int. Ser. Monogr. Pure Appl. Math. 88, Pergamon, Oxford, 1966.Search in Google Scholar

[9] P. P. Zabrejko, A. I. Koshelev, M. A. Krasnosel’skii, S. G. Mikhlin, L. S. Rakovschik and J. Stetsenko, Integral Equations, Nordhoff, Leyden, 1975.10.1007/978-94-010-1909-5_10Search in Google Scholar

[10] J. Banaś and A. Chlebowicz, On solutions of an infinite system of nonlinear integral equations on the real half-axis, Banach J. Math. Anal. (to appear).10.1215/17358787-2019-0008Search in Google Scholar

[11] 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

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

[13] K. Deimling, Ordinary Differential Equations in Banach Spaces, Lect. Notes Math. 596, Springer, Berlin, 1977.10.1007/BFb0091636Search in Google Scholar

[14] M. Mursaleen and S. M. H. Rizvi, Solvability of infinite systems of second order differential equations in c0 and l1 by Meir-Keeler condensing operator, Proc. Amer. Math. Soc. 144 (2016), no. 10, 4279–4289.10.1090/proc/13048Search in Google Scholar

[15] J. Banaś and K. Goebel, Measures of Noncompactness in Banach Spaces, Lect. Notes Pure Appl. Math. 60, Marcel Dekker, New York, 1980.Search in Google Scholar

[16] K. Kuratowski, Sur les espaces complets, Fund. Math. 15 (1930), 301–309.10.4064/fm-15-1-301-309Search in Google Scholar

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

[18] L.S. Goldenštein and A.S. Markus, “On the measure of non-compactness of bounded sets and linear operators” in Studies in Algebra and Mathematical Analysis, Izdat. “Karta Moldovenjaske”, Kishinev, 1965, 45–54.Search in Google Scholar

[19] G. Darbo, Punti uniti in trasformazioni a condominio non compatto, Rend. Semin. Mat. Univ. Padova 24 (1955), 84–92.Search in Google Scholar

[20] W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1987.Search in Google Scholar

Received: 2019-07-18
Accepted: 2019-09-18
Published Online: 2019-12-06

© 2019 Agnieszka Chlebowicz, published by De Gruyter

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

Articles in the same Issue

  1. Frontmatter
  2. On the moving plane method for boundary blow-up solutions to semilinear elliptic equations
  3. Regularity of solutions of the parabolic normalized p-Laplace equation
  4. Cahn–Hilliard equation on the boundary with bulk condition of Allen–Cahn type
  5. Blow-up solutions for fully nonlinear equations: Existence, asymptotic estimates and uniqueness
  6. Radon measure-valued solutions of first order scalar conservation laws
  7. Ground state solutions for a semilinear elliptic problem with critical-subcritical growth
  8. Generalized solutions of variational problems and applications
  9. Existence and non-existence results for Kirchhoff-type problems with convolution nonlinearity
  10. Nonlinear Sherman-type inequalities
  11. Global regularity for systems with p-structure depending on the symmetric gradient
  12. Homogenization of a net of periodic critically scaled boundary obstacles related to reverse osmosis “nano-composite” membranes
  13. Noncoercive resonant (p,2)-equations with concave terms
  14. Evolutionary quasi-variational and variational inequalities with constraints on the derivatives
  15. Sharp estimates on the first Dirichlet eigenvalue of nonlinear elliptic operators via maximum principle
  16. Localization and multiplicity in the homogenization of nonlinear problems
  17. Remarks on a nonlinear nonlocal operator in Orlicz spaces
  18. A Picone identity for variable exponent operators and applications
  19. On the weakly degenerate Allen-Cahn equation
  20. Continuity results for parametric nonlinear singular Dirichlet problems
  21. Construction of type I blowup solutions for a higher order semilinear parabolic equation
  22. Singularly perturbed Choquard equations with nonlinearity satisfying Berestycki-Lions assumptions
  23. Comparison results for nonlinear divergence structure elliptic PDE’s
  24. Constant sign and nodal solutions for parametric (p, 2)-equations
  25. Monotonicity formulas for coupled elliptic gradient systems with applications
  26. Berestycki-Lions conditions on ground state solutions for a Nonlinear Schrödinger equation with variable potentials
  27. A class of semipositone p-Laplacian problems with a critical growth reaction term
  28. The role of superlinear damping in the construction of solutions to drift-diffusion problems with initial data in L1
  29. Reconstruction of Tesla micro-valve using topological sensitivity analysis
  30. Lewy-Stampacchia’s inequality for a pseudomonotone parabolic problem
  31. Global well-posedness of nonlinear wave equation with weak and strong damping terms and logarithmic source term
  32. Regularity Criteria for Navier-Stokes Equations with Slip Boundary Conditions on Non-flat Boundaries via Two Velocity Components
  33. Homoclinics for singular strong force Lagrangian systems
  34. A constructive method for convex solutions of a class of nonlinear Black-Scholes equations
  35. On a class of nonlocal nonlinear Schrödinger equations with potential well
  36. Superlinear Schrödinger–Kirchhoff type problems involving the fractional p–Laplacian and critical exponent
  37. Regularity for minimizers for functionals of double phase with variable exponents
  38. Boundary blow-up solutions to the Monge-Ampère equation: Sharp conditions and asymptotic behavior
  39. Homogenisation with error estimates of attractors for damped semi-linear anisotropic wave equations
  40. A-priori bounds for quasilinear problems in critical dimension
  41. Critical growth elliptic problems involving Hardy-Littlewood-Sobolev critical exponent in non-contractible domains
  42. On the Sobolev space of functions with derivative of logarithmic order
  43. On a logarithmic Hartree equation
  44. Critical elliptic systems involving multiple strongly–coupled Hardy–type terms
  45. Sharp conditions of global existence for nonlinear Schrödinger equation with a harmonic potential
  46. Existence for (p, q) critical systems in the Heisenberg group
  47. Periodic traveling fronts for partially degenerate reaction-diffusion systems with bistable and time-periodic nonlinearity
  48. Some hemivariational inequalities in the Euclidean space
  49. Existence of standing waves for quasi-linear Schrödinger equations on Tn
  50. Periodic solutions for second order differential equations with indefinite singularities
  51. On the Hölder continuity for a class of vectorial problems
  52. Bifurcations of nontrivial solutions of a cubic Helmholtz system
  53. On the exact multiplicity of stable ground states of non-Lipschitz semilinear elliptic equations for some classes of starshaped sets
  54. Sign-changing multi-bump solutions for the Chern-Simons-Schrödinger equations in ℝ2
  55. Positive solutions for diffusive Logistic equation with refuge
  56. Null controllability for a degenerate population model in divergence form via Carleman estimates
  57. Eigenvalues for a class of singular problems involving p(x)-Biharmonic operator and q(x)-Hardy potential
  58. On the convergence analysis of a time dependent elliptic equation with discontinuous coefficients
  59. Multiplicity and concentration results for magnetic relativistic Schrödinger equations
  60. Solvability of an infinite system of nonlinear integral equations of Volterra-Hammerstein type
  61. The superposition operator in the space of functions continuous and converging at infinity on the real half-axis
  62. Estimates by gap potentials of free homotopy decompositions of critical Sobolev maps
  63. Pseudo almost periodic solutions for a class of differential equation with delays depending on state
  64. Normalized multi-bump solutions for saturable Schrödinger equations
  65. Some inequalities and superposition operator in the space of regulated functions
  66. Area Integral Characterization of Hardy space H1L related to Degenerate Schrödinger Operators
  67. Bifurcation of time-periodic solutions for the incompressible flow of nematic liquid crystals in three dimension
  68. Morrey estimates for a class of elliptic equations with drift term
  69. A singularity as a break point for the multiplicity of solutions to quasilinear elliptic problems
  70. Global and non global solutions for a class of coupled parabolic systems
  71. On the analysis of a geometrically selective turbulence model
  72. Multiplicity of positive solutions for quasilinear elliptic equations involving critical nonlinearity
  73. Lack of smoothing for bounded solutions of a semilinear parabolic equation
  74. Gradient estimates for the fundamental solution of Lévy type operator
  75. π/4-tangentiality of solutions for one-dimensional Minkowski-curvature problems
  76. On the existence and multiplicity of solutions to fractional Lane-Emden elliptic systems involving measures
  77. Anisotropic problems with unbalanced growth
  78. On a fractional thin film equation
  79. Minimum action solutions of nonhomogeneous Schrödinger equations
  80. Global existence and blow-up of weak solutions for a class of fractional p-Laplacian evolution equations
  81. Optimal rearrangement problem and normalized obstacle problem in the fractional setting
  82. A few problems connected with invariant measures of Markov maps - verification of some claims and opinions that circulate in the literature
Downloaded on 21.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2020-0045/html
Scroll to top button