Next Article in Journal
Smooth Versions of the Mann–Whitney–Wilcoxon Statistics
Previous Article in Journal
Dynamic Behaviors of an Obligate Commensal Symbiosis Model with Crowley–Martin Functional Responses
Previous Article in Special Issue
The Space of Functions with Tempered Increments on a Locally Compact and Countable at Infinity Metric Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Survey of Measures of Noncompactness and Their Applications

by
Moosa Gabeleh
1,†,
Eberhard Malkowsky
2,†,
Mohammad Mursaleen
3,4,† and
Vladimir Rakočević
5,*,†
1
Department of Mathematics, Ayatollah Boroujerdi University, Boroujerd RQHV3W6, Iran
2
Department of Mathematics, State University of Novi Pazar, Vuka Karadžića bb, 36300 Novi Pazar, Serbia
3
Department of Medical Research, China Medical University Hospital, China Medical University (Taiwan), No. 91, Xueshi Rd, North District, Taichung City 404, Taiwan
4
Department of Mathematics, Aligarh Muslim University, Aligarh 202002, UP, India
5
Department of Mathematics, Faculty of Sciences and Mathematics, University of Niš, Višegradska 33, 18000 Niš, Serbia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Axioms 2022, 11(6), 299; https://doi.org/10.3390/axioms11060299
Submission received: 22 May 2022 / Revised: 15 June 2022 / Accepted: 17 June 2022 / Published: 20 June 2022
(This article belongs to the Special Issue Operator Theory and Its Applications)

Abstract

:
We present a survey of the theory of measures of noncompactness and discuss some fixed point theorems of Darbo’s type. We apply the technique of measures of noncompactness to the characterization of classes of compact operators between certain sequence spaces, in solving infinite systems of integral equations in some sequence spaces. We also present some recent results related to the existence of best proximity points (pairs) for some classes of cyclic and noncyclic condensing operators in Banach spaces equipped with a suitable measure of noncompactness. Finally, we discuss the existence of an optimal solution for systems of integro–differentials.

1. Introduction, Notations and Preliminaries

Measures of noncompactness play an important role in nonlinear functional analysis. They are important tools in metric fixed point theory, the theory of operator equations in Banach spaces, and the characterizations of classes of compact operators. They are also applied in the studies of various kinds of differential and integral equations.
The first measure of noncompactness, the function α , was defined and studied by Kuratowski [1] in 1930. In 1955, Darbo [2] was the first to use the function α to prove his famous fixed point theorem, Theorem 9.
The second measure of noncompactness was introduced by Goldenštein et al. [3,4], namely the Hausdorff or ball measure of noncompactness denoted by χ .
We refer to [5,6,7,8,9,10] for comprehensive studies.
Throughout, we use the following standard notations.
Let ( X , d ) be a metric space, x X and r > 0 . A subset M of X is relatively compact if it has compact closure M ¯ . Further,
B ( x , r ) = B X ( x , r ) = { y X : d ( y , x ) < r } , B ¯ ( x , r ) = B ¯ X ( x , r ) = { y X : d ( y , x ) r } and S ( x , r ) = S X ( x , r ) = { y X : d ( y , x ) = r }
denote the open and closed ball, and the sphere of radius r centered at x, respectively. If X is a normed space, x = 0 and r = 1 , then we write B X = B X ( 0 , 1 ) , B ¯ X = B ¯ X ( 0 , 1 ) and S X = S X ( 0 , 1 ) . Let S and S ˜ be subsets of a metric space ( X , d ) , then:
diam ( S ) = sup { d ( s 1 , s 2 ) : s 1 , s 2 S } , dist ( S , S ˜ ) = inf { d ( s , s ˜ ) : s S , s ˜ S ˜ } and dist ( x , S ) = dist ( { x } , S )
are called the diameter of S, the distance of S, and S ˜ , and the distance of the point x and the set S, respectively.
If M , S X and ε > 0 , then S is called an ε –net of M, if, for every x M , there exists s S such that d ( x , s ) < ε ; if S is finite, then S is a finite ε –net of M.
A sequence ( b n ) in a linear metric space X is called a Schauder basis for X if for every x X there exists a unique sequence ( λ n ) n = 1 of scalars such that:
x = n = 1 λ n x n .
Let X and Y be Banach spaces. Then, B ( X , Y ) denotes the Banach space of all bounded linear operators from X into Y with the operator norm:
L = sup { L ( x ) : x = 1 } for   all L B ( X , Y ) ;
we write B ( X ) = B ( X , X ) , for short. In particular, X * = B ( X , C ) denotes the set of all continuous linear functionals on X with the norm:
f = sup { | f ( x ) | : x = 1 } for   all f X * ;
X * is also referred to as the continuous dual of X.
An operator L : X Y is compact if L maps bounded subsets of X to relatively compact subsets of Y, or equivalently, for any bounded sequence ( x n ) in X, the sequence ( L ( x n ) ) has a convergent subsequence in Y. The set of all compact operators from X to Y is denoted by C ( X , Y ) ; we write C ( X ) = C ( X , Y ) , for short.

B k Spaces

The study of operators, in particular of matrix transformations, between sequence spaces is an important field of applications of measures of noncompactness. Here, we mention the standard notations and list the necessary results concerning B K spaces. We recommend the monographs [9,11,12,13,14,15] for the study of the theory of B K spaces.
We denote by ω the set of all complex sequences x = ( x k ) k = 1 , and by , c, c 0 , and ϕ the subsets of ω of all bounded, convergent, null, and finite sequences, and write:
p = x = ( x k ) k = 1 ω : k = 1 | x k | p < for 1 p < .
Moreover, c s and b s denote the sets of all convergent and bounded series of complex numbers, respectively. Finally we write:
b v = x = ( x k ) k = 1 ω : Δ x = ( x k x k + 1 ) k = 1 1
for the set of all sequences of bounded variation, and b v 0 = b v c 0 .
We write e = ( e k ) k = 1 and e ( n ) = ( e k ( n ) ) k = 1   ( n N ) for the sequences with e k = 1 for all k, and e n ( n ) = 1 and e k ( n ) = 0 for k n .
Example 1.
The following facts are well known.
(a) 
The set ω is a Fréchet space, that is, a complete linear metric space, with respect to:
d ω ( x , y ) = k = 1 1 2 k · | x k y k | 1 + | x k y k | ( x , y ω )
and convergence in ( ω , d ω ) and coordinatewise convergence are equivalent; this means:
lim n d ( x k ( n ) ) k = 1 , ( x k ) k = 1 = 0   i f   a n d   o n l y   i f lim n x k ( n ) = x k   f o r   e a c h   k .
(b) 
The sets , c, c 0 , p for 1 p < , b s , c s , b v and b v 0 are Banach spaces with respect to their natural norms defined by:
x = sup k | x k | o n   , c , c 0 , x p = k = 1 | x k | p 1 / p   o n   p , x b s = sup n k = 1 n x k   o n   b s , c s , x b v = k = 1 | x k x k + 1 | + lim k x k   o n   b v ,
and
x b v 0 = k = 1 | x k x k + 1 |   o n   b v 0 .
Now, we recall the definition of F K spaces, and their special cases B K spaces. F K and B K were first studied by Zeller [16,17,18].
Definition 1.
A Fréchet sequence space ( X , d ) is called an F K space if d is stronger than d ω , that is, if the inclusion map ι : ( X , d ) ( ω , d ω ) with ι ( x ) = x for all x X is continuous. An F K space is called a B K space if its metric is given by a norm.
We note that, by Example 1 (a), a Fréchet sequence space ( X , d ) is an F K space if convergence in d implies coordinatewise convergence.
Now, we recall the concept of the A K property.
Definition 2.
An F K space X has A K , if every sequence x = ( x k ) k = 1 X has a unique representation:
x = k = 1 x k e ( k ) ,   t h a t   i s , x = lim n x [ n ] ,
where x [ n ] = k = 1 n x k e ( k ) is then–section x.
Example 2.
The following facts are well known.
(a) 
The F K space ( X , d ω ) has A K .
(b) 
The Banach spaces of Example 1 (a) are B K space with respect to their natural norms.
(c) 
The B K spaces c 0 , p ( 1 p < ) , c s and b v 0 have A K ; every sequence x = ( x k ) k = 1 c has a unique representation:
x = ξ e + k = 1 ( x k ξ ) e ( k ) ,
where ξ = lim k x k ; and b s have no Schauder bases.
We also recall the following concepts.
Let X ϕ . Then, the set,
X β = { a = ( a k ) k = 1 : a x = ( a k x k ) k = 1 c s for all x = ( x k ) k = 1 X }
is called the β–dual of X.
Theorem 1
([13], Theorem 7.2.9).
Let X ϕ be an F K space. Then, X β X ; this means that there is a linear one–to–one map T : X β X . If X has A K then T is onto.
Let A = ( a n k ) n , k = 1 be an infinite matrix of complex entries, A n = ( a n k ) k = 1 denote the sequence in the n t h row of A, x = ( x k ) k = 1 be a sequence and X and Y be subsets of ω . Then
A n x = k = 1 a n k x k for n N
is called the n t h A transform of the sequence x, and A x = ( A n x ) n = 1 is called the A transform of the sequence x (provided all the series converge). Furthermore,
X A = { x = ( x k ) k = 1 ω : A x X }
is the matrix domain of A in X. Finally ( X , Y ) denotes the class of all infinite matrices A with X Y A .
Now, we state the probably most important result concerning matrix transformations.
Theorem 2
([13], Theorem 4.2.7). Matrix transformations between F K spaces are continuous.
Finally, we state the relation between the classes B ( X , Y ) and ( X , Y ) for B K spaces X and Y; the first part is a special case of Theorem 2, and the second part is ([9], Theorem 9.3.3).
Theorem 3.
Let X and Y be B K spaces.
(a) 
Then, ( X , Y ) B ( X , Y ) ; this means, every matrix A ( X , Y ) defines an operator L A B ( X , Y ) , where:
L A ( x ) = A x for all x = ( x k ) k = 1 X .
(b) 
If X has A K then B ( X , Y ) ( X , Y ) ; this means, every operator L B ( X , Y ) is given by a matrix A ( X , Y ) , where:
A x = L ( x ) for all x = ( x k ) k = 1 X .
Example 3
([13], 8.4.1D). We have L B ( 1 ) if and only if:
L = A ( 1 , 1 ) = sup k n = 1 | a n k | < ,
where A = ( a n k ) n , k = 1 represents L as in (4).
Proof. 
By Example 2 (c), 1 is a B K space with A K , hence L B ( 1 ) if and only if A ( 1 , 1 ) with L ( x ) = A x for all x 1 .
(i) Let L B ( 1 ) . Then we obtain for each k N ,
L ( e ( k ) ) 1 = n = 1 | A n e ( k ) | = n = 1 | a n k | L · e ( k ) 1 = L ,
whence,
sup k n = 1 | a n k | L < ,
since k N was arbitrary.
(ii) Conversely, we assume that sup k n = 1 | a n k | < . Then it follows for all x 1 that:
L ( x ) 1 = n = 1 | A n x | n = 1 k = 1 | a n k x k | = k = 1 | x k | n = 1 | a n k | sup k n = 1 | a n k | · x 1 < ,
hence L B ( 1 ) and:
L sup k n = 1 | a n k | .
Finally (6) and (7) yield (5). □

2. Measures of Noncompactness and Their Properties

We start with the axioms of a measure of noncompactness on M X , the bounded subsets of a complete metric space ( X , d ) ; they can be found, for instance, in ([7], Definition II, 1.1).
We will also consider the axioms of measures of noncompactness in Banach spaces as in [5,6].
Definition 3.
Let X be a complete metric space. A set function ϕ : M X [ 0 , ) is a measure of noncompactness on M X , if the following conditions are satisfied for all sets Q , Q 1 , Q 2 M X ,
( M N C . 1 ) ϕ ( Q ) = 0   i f   a n d   o n l y   i f   Q ¯   i s   c o m p a c t ( r e g u l a r i t y ) ( M N C . 2 ) ϕ ( Q ) = ϕ ( Q ¯ ) ( i n v a r i a n c e   u n d e r   c l o s u r e ) ( M N C . 3 ) ϕ ( Q 1 Q 2 ) = max { ϕ ( Q 1 ) , ϕ ( Q 2 ) } ( s e m i a d d i t i v i t y ) .
Example 4.
Let X be a complete metric space and ϕ for all Q M X be defined by ϕ ( Q ) = 0 if Q is relatively compact, and ϕ ( Q ) = 1 otherwise. Then ϕ is a measure of noncompactness, the so–called trivial measure of noncompactness.
The following properties are easily obtained from Definition 3.
Proposition 1.
Let ϕ be a measure of noncompactness on a complete metric space X. Then ϕ has the following properties:
Q Q ˜   i m p l i e s   ϕ ( Q ) ϕ ( Q ˜ ) ( monotonicity ) ,
ϕ ( Q 1 Q 2 ) min { ϕ ( Q 1 ) , ϕ ( Q 2 ) }   f o r   a l l   Q 1 , Q 2 M X .
I f   Q   i s   f i n i t e   t h e n   ϕ ( Q ) = 0 ( non–singularity ) .
Generalized Cantor’s intersection property I f   ( Q n )   i s   a   d e c r e a s i n g   s e q u e n c e   o f   n o n e m p t y ,   b o u n d e d   a n d   c l o s e d   s e t s   i n   X , a n d lim n ϕ ( Q n ) = 0 ,   t h e n   t h e   i n t e r s e c t i o n Q = Q n i s   c o m p a c t .
Definition 4.
Let ( X , d ) be a complete metric space.
(a) 
The function α : M X [ 0 , ) with:
α ( Q ) = inf ε > 0 : Q k = 1 n S k , S k X , d i a m ( S k ) < ε ( k = 1 , 2 , , n ; n N )
for all Q M X is called the Kuratowski measure of noncompactness; the real number α ( Q ) is called the Kuratowski measure of noncompactness of Q.
(b) 
The function χ : M X [ 0 , ) with:
χ ( Q ) = inf ε > 0 : Q k = 1 n B ( x k , r k ) , B ( x k , r k ) X , r k < ε ( k = 1 , 2 , , n ; n N )
for all Q M X is called the Hausdorff or ball measure of noncompactness; the real number χ ( Q ) is called the Hausdorff or ball measure of noncompactness of Q.
(c) 
We recall that a subset S of ( X , d ) is said to be r–separated or r–discrete, if d ( x , y ) r for all distinct elements of S; the set S is called an r–separation. The function β : M X [ 0 , ) with:
β ( Q ) = sup { r > 0 : Q   h a s   a n   i n f i n i t e   r–separation } ,
or equivalently,
β ( Q ) = inf { r > 0 : Q   d o e s   n o t   a n   i n f i n i t e   r–separation }
for all Q M X is called the separation or Istrǎţescu measure of noncompactness; the real number β ( Q ) is called the separation or Istrǎţescu measure of noncompactness of Q.
Remark 1.(a) If it is required that the centers of the balls that cover Q belong to Q then the real number χ i ( Q ) is referred to a the inner Hausdorff measure on noncompactness of Q, and the function χ i : M X [ 0 , ) is called the inner Hausdorff measure on noncompactness.
(b) 
(([9], Remark 7.7.3) The function χ i is not a measure of noncompactness in the sense of Definition 3; it satisfies the conditions in (MNC.1) and (MNC.2), but (MNC.3) and (8) do not hold, in general. It can be shown that:
χ ( Q ) χ i ( Q ) α ( Q )   f o r   a l l   Q M X .
The following results hold (([9], Theorems 7.6.3, 7.7.5 (a)) for α and χ , and ([7], Remark II.3.2) for β ).
Theorem 4.
Let X be a complete metric space, and ϕ be any of the functions α, χ or β. Then ϕ is a measure of noncompactness which also satisfies the conditions in (8)–(11).
If X is a Banach space, then some measures of noncompactness my satisfy some additional conditions. The convex hull of a subset M of a linear space is denoted and defined by:
co ( M ) = { C M : C convex } .
The following results hold for α and χ by ([7], Proposition II.2.3 and Theorem II.2.4) and for β by ([7], Remark II.3.2 and Theorems II.3.4 and II.3.6).
Theorem 5.
Let X be a Banach space, and ϕ be any of the functions α, χ or β. Then we have for all Q , Q 1 , Q 2 M X :
ϕ ( λ Q ) = | λ | ϕ ( Q )   f o r   a l l   λ C semi–homogeneity )
ϕ ( Q 1 + Q 2 ) ϕ ( Q 1 ) + ϕ ( Q 2 ) ( algebraic semi–additivity )
ϕ ( x + Q ) = ϕ ( Q )   f o r   a l l   x X ( translation invariance )
ϕ ( c o ( Q ) ) = ϕ ( Q ) ( invariance under the passage to the convex hull ) .
Remark 2.
Let X be an infinite dimensional Banach space.
(a) 
([7], Corollary II.2.6) Then,
α ( B X ) = α ( B ¯ X ) = α ( S X ) = 2   a n d   χ ( B X ) = χ ( B ¯ X ) = χ ( S X ) = 1 .
(b) 
([7], Remark II.3.2) The functions α, β and χ are equivalent, that is,
χ ( Q ) β ( Q ) α ( Q ) 2 χ ( Q )   f o r   a l l   Q M X .
(c) 
The Kuratowski and Hausdorff measures of noncompactness are closely related to the geometric properties of the space; the inequality χ ( Q ) α ( Q ) can be improved in some spaces ([19,20]).
For instance, in Hilbert spaces H ([5,21]):
2 χ ( Q ) α ( Q ) 2 χ ( Q )   f o r   a l l   Q M H ,
and in p for 1 p < ,
χ p ( Q ) α ( Q ) 2 χ ( Q )   f o r   a l l   Q M p .
(d) 
Studies on inequivalent measures of noncompactness can be found, for instance, in [22,23].
Whereas α ( B X ) and χ ( B X ) in infinite dimensional Banach spaces X are independent of the space, this is not true for β . The following result holds by ([7], Remark II.3.11 and Theorem II.3.12) for 1 p 2 and 2 < p < , respectively.
Remark 3.
Let 1 p < . Then β ( B p ) = 2 1 / p .
There is a relation between the Hausdorff distance (Definition 5) and χ .
Definition 5.
Let ( X , d ) be a metric space. The function d H : M X × M X R defined by:
d H ( S , S ˜ ) = max sup s S dist ( s , S ˜ ) , sup s ˜ S ˜ dist ( s ˜ , S ) for   all S , S ˜ M X
is called the Hausdorff distance; the value d H ( S , S ˜ ) is called the Hausdorff distance of the sets S and S ˜ .
Remark 4
([9], Theorem 7.4.2). It is well known that if ( X , d ) is a metric space, then ( M X , d H ) is a semimetric space and ( M X c , d H ) is a metric space, where M X c denotes the class of closed subsets in M X .
We also mention the following result.
Theorem 6
([9], Theorem 7.7.14). Let ( X , d ) be a complete metric space, and N X c denote the class of all nonempty compact sets in M X . Then we have:
χ ( Q 1 ) χ ( Q 2 ) d H ( Q 1 , Q 2 )   f o r   a l l   Q 1 , Q 2 M X ,
and,
χ ( Q ) = d H ( Q , N X c )   f o r   a l l   Q M X .
Now, we list the axioms for measures of noncompactness in as stated by Banaś and Goebel [5].
Definition 6
([5], Definition 3.1.1). Let X be a Banach space.
A function ψ : M X [ 0 , ) is a measure of noncompactness on X if it satisfies the conditions (MNC.2) (invariance under closure), (8) (monotonicity), (14) (invariance under the passage to the convex hull), and,
(i) 
The family ker ( ψ ) = { Q M X : ψ ( Q ) = 0 } is contained in the family of all relatively compact subsets of X (compare this with (MNC.1));
(ii) 
i=If ( Q n ) is a decreasing sequence of sets in M X c with lim n ψ ( Q n ) = 0 , then
Q = n = 1 Q n k e r ( ψ )
(compare with (11) (Cantor’s generalized intersection property));
(iii) 
ψ ( λ Q + ( 1 λ ) Q ˜ ) λ ψ ( Q ) + ( 1 λ ) ψ ( Q ˜ ) for all λ ( 0 , 1 ) and all Q , Q ˜ M X (convexity condition).
Remark 5.(a) The functions α, χ, and β are measures on noncompactness in the sense of Definition 6. (b) The family ker ( ψ ) is referred to as the kernel of the measure of noncompactness ψ.
(c) 
A measure of noncompactness is said to be sublinear if it satisfies the conditions (12) and (13) (semi–homogeneity and algebraic semi–additivity). If ker ( ψ ) = N , the family of all nonempty relatively compact sets, then ψ is said to be full.
Remark 6.
The term measure of noncompactness will always be used in the sense of Definition 3 unless explicitly stated otherwise.
As an important application of the Hausdorff measure of noncompactness χ we are now going to state the famous Goldenštein, Go’hberg, Markus theorem [3] which provides an estimate for χ ( Q ) in Banach spaces with a Schauder basis.
Theorem 7 
(Goldenštein, Go’hberg, Markus ([3] or [9], Theorem 7.9.3)).
Let X be a Banach space with a Schauder basis ( b k ) and the functions μ n : M X [ 0 , ) for n = 1 , 2 , be defined by:
μ n ( Q ) = sup x Q R n ( x ) ,
where R n : X X for each n is the function with:
R n ( x ) = k = n + 1 λ k x k   f o r   a l l   x = k = 1 λ k x k X .
Then, we have for all Q M X :
1 a lim sup n μ n ( Q ) χ ( Q ) inf n μ n ( Q ) lim sup n μ n ( Q ) ,
where a = lim sup n R n is the basis constant.
The following corollary of Theorem 7 is very useful for B K spaces with A K with a so–called monotonous norm · , that is, a norm for which x x ˜ whenever x , x ˜ X with | x k | | x ˜ k | for all k.
Corollary 1
([9], Lemma 9.8.1).
(a) 
Let ( X , · ) be a monotonous B K space with A K and R n ( x ) = x x [ n ] ( x X ) for each n. Then we have:
χ ( Q ) = lim n sup x Q R n ( x )   f o r   a l l   Q M X .
(b) 
Let R n : c c for n = 1 , 2 , be defined by R n ( x ) = k = n + 1 ( x k ξ ) e ( k ) for all x = ξ e + k = 1 ( x k ξ ) e ( k ) c , where ξ = lim k x k . Then,
lim n sup x Q R n ( x )   e x i s t s   f o r   a l l   Q M c ,
and a = lim n R n = 2 .
Example 5.(a) Since p ( 1 p < ) and c 0 are monotonous B K spaces with A K , Corollary 1 (a) yields:
χ ( Q ) = lim n sup x Q k = n + 1 | x k | p 1 / p ( Q M p ) lim n sup x Q sup k n + 1 | x k | ( Q M c 0 ) .
(b) 
We obtain from Corollary 1(b):
1 2 lim n sup x Q R n ( x ) χ ( Q ) inf n sup x Q R n ( x ) ( Q M c ) ,
where, for each x c with ξ x = lim k x k ,
R n ( x ) = sup k n | x k ξ | .

Measures of Noncompactness of Operators

Contractive and condensing maps play an important role in fixed point theory, for instance in Banach’s and Darbo’s eminent fixed point theorems. Now, we are going to introduce these concepts, and measures of noncompactness of operators.
Definition 7
([7], Definition II.5.1). Let X and Y be complete metric spaces, ϕ and ψ be measures of noncompactness on X and Y, respectively, and L : D X Y be a map. Then:
(a) 
L is a ( ϕ , ψ ) –contractive operator with constant k > 0 , or k ( ϕ , ψ ) –contractive, for short, if L is continuous and satisfies:
ψ ( L ( Q ) ) k · ϕ ( Q )   f o r   e v e r y   Q M D .
If X = Y and ψ = ϕ , L is referred to as a k ϕ –contractive operator.
(b) 
L is a ( ϕ , ψ ) –condensing operator with constant k > 0 , or k ( ϕ , ψ ) –condensing, for short, if L is continuous and satisfies
ψ ( L ( Q ) ) < k · ϕ ( Q )   f o r   e v e r y   n o n r e l a t i v e l y   c o m p a c t   s e t   Q M D .
If X = Y and ψ = ϕ , L is referred to as a k ϕ –condensing operator. Moreover, if k = 1 , then L is said to be a ϕ–condensing operator.
Remark 7
([7], Proposition II.5.3).
(a) 
If ϕ = α , the Kuratowski measure of noncompactness, then the k α –contractive (condensing) operators are calledk–set contractive (condensing).
(b) 
If ϕ = χ , the Hausdorff measure of noncompactness, then the k χ –contractive (condensing) operators are calledk–ball contractive (condensing).
(c) 
Every compact operator is k ( ϕ , ψ ) –contractive and k ( ϕ , ψ ) –condensing for all k > 0 .
(d) 
Every k ( ϕ , ψ ) –condensing operator is k ( ϕ , ψ ) –contractive, but the converse is not true, in general.
(e) 
An example of a set–condensing operator which is not k–set–contractive for any k [ 0 , 1 ) can be found in ([7], Example II.6).
We recall that a map f from a metric space ( X , d ) into itself is called a contraction if there exists a constant c ( 0 , 1 ) such that:
d ( f ( x ) , f ( y ) ) c · d ( x , y ) for all x , y X .
Using the above concepts, we can now state the famous fixed point theorems by Banach et al. Banach’s fixed point theorem is also referred to as the Banach contraction principle. We recommend the monographs [24,25,26,27,28] and the survey paper [29] for further studies on fixed point theorems.
Theorem 8
(Banach’s fixed point theorem). Every contraction from a complete metric space into itself has a unique fixed point.
Theorem 9
(Darbo’s fixed point theorem [2]). Let X be a Banach space and C M X c be nonempty and convex. If L : C C is a k–contractive set operator for some k ( 0 , 1 ) , then L has a fixed point in C.
Darbo’s fixed point theorem is a generalization of Schauder’s fixed point theorem.
Theorem 10 
(Schauder’s fixed point theorem) ([30], Theorem 1). Every continuous map from a nonempty, compact and convex subset C of a Banach space into C has a fixed point.
The next result is a generalization of Theorem 9.
Theorem 11 
(Darbo–Sadovskiĭ ([31,32] or ([7], Theorem II.5.4))).
Let X be a Banach space, ϕ be a measure of noncompactness which is invaraint under the passage to the convex hull, C M X c be nonempty and convex, and L : M M be a ϕ–condensing operator. Then L has a fixed point.
The following example shows that Theorem 11 need not hold for one–contractive operators f.
Example 6
([7], Example II.7).
We define the operator f : B ¯ 2 B ¯ 2 by:
f ( x ) = f ( x k ) k = 1 = 1 x 2 2 , x 1 , x 2 , for all x = ( x k ) k = 1 B ¯ 2 .
Then, we can write f = g + h , where g is the mapping with:
g ( x ) = g ( x k ) k = 1 = 1 x 2 2 e ( 1 ) ,
and h ( x ) = ( 0 , x 1 , x 2 , ) is an isometry.
Then f is a well–defined, continuous operator, and every bounded subset Q in B ¯ 2 satisfies:
α ( f ( Q ) ) α ( g ( Q ) ) + α ( h ( Q ) ) = α ( h ( Q ) ) = α ( Q ) .
Consequently, f is a one–set–contractive operator, but f has no fixed points.
If f had a fixed point x B ¯ 2 , then we would have x k = x k + 1 for all k. Since x 2 , this would imply x k = 0 for all k, and then f ( x ) = 1 x 2 2 e ( 0 ) = e ( 0 ) = ( 0 , 0 , ) . This is a contradiction.
Definition 8
([9], Definition 7.11.1).
Let ϕ and ψ be measures of noncompactness on the Banach spaces X and Y, respectively.
(a) 
An operator L : X Y is said to be ( ϕ , ψ ) –bounded , if:
L ( Q ) M Y for all Q M X ,
and if there exist a nonnegative real number c such that:
ψ ( L ( Q ) ) c · ϕ ( Q ) for all Q M X .
(b) 
If an operator L is ( ϕ , ψ ) –bounded, then the number,
L ( ϕ , ψ ) = inf { c 0 : ( 17 ) h o l d s }
is called the ( ϕ , ψ ) –operator norm of Lor ( ϕ , ψ ) –measure of noncompactness of L.
If ψ = ϕ , we write L ϕ = L ( ϕ , ψ ) , for short.
Remark 8.
A ( ϕ , ψ ) –bounded operator is a c–contractive ( ϕ , ψ ) – operator between Banach spaces for some c by Definitions 8 (a) and 7 (a).
Theorem 12
([9], Theorem 7.11.4). Let X and Y be infinite dimensional Banach spaces and L B ( X , Y ) . Then we have:
L χ = χ ( L ( S X ) ) = χ ( L ( B X ) ) = χ ( L ( B ¯ X ) ) .
Theorem 13
([9], Theorem 7.11.5). Let X, Y, and Z be Banach spaces, L B ( X , Y ) and L ˜ B ( Y , Z ) . Then · χ is a seminorm on B ( X , Y ) , and:
L χ =   0   i f   a n d   o n l y   i f   L C ( X , Y ) ; L χ   L   f o r   a l l   L B ( X , Y ) ; L ˜ L χ   L ˜ χ · L χ   f o r   a l l   L B ( X , Y ) a n d   L ˜ B ( Y , Z ) .
In Example (3), we characterized the class B ( 1 ) and established a formula for the norm of operators in B ( 1 ) . Now we characterize the class C ( 1 ) .
Example 7
(Goldenštein, Go’hberg, Markus). ([3] or ([9], Theorem 7.9.3)) Let L B ( 1 ) . Then:
L χ = lim m sup k n = m | a n k | ,
where A = ( a n k ) n , k = 1 represents L.
Furthermore, L C ( 1 ) if and only if:
lim m sup k n = m | a n k | = 0 .
Proof. 
Let A = ( a n k ) n , k = 1 be any infinite matrix. Then, for each m N , let A < m > be the matrix with the rows A n < m > = 0 for n m and A n < m > = A n for n m + 1 . It is clear that:
R m L ( x ) = A < m > ( x ) for all x 1 .
Now (19), Example 5 (a) and (5) in Example 3 imply:
L χ = χ L ( B ¯ 1 ) = lim m sup x 1 = 1 R m L ( x ) 1 = lim m A < m > ( 1 , 1 ) = lim m sup k n = m + 1 | a n k | ,
whence (21).
Furthermore, it follows from (21) and (20) that L C ( 1 ) if and only if (22) is satisfied. □

3. Bounded and Compact Operators on the Generalized Hahn Space

Here, we apply the results of Section 1 and Section 2 to the characterizations of classes of bounded and compact linear operators from the generalized Hahn space h d into itself and into the spaces of sequences that are strongly summable by the Cesàro method of order one, with index p 1 , and into the spaces of strongly convergent sequences. We also establish identities or estimates for the Hausdorff measure of noncompactness of those operators.
For further studies on the generalized Hahn space we recommend the research papers [33,34,35].

The Properties of Our Sequence Spaces

We recall the definition of the operators Δ , Δ : ω ω of the forward and backward differences given for all sequences x = ( x k ) k = 1 by:
Δ x k = x k x k + 1 and Δ x k = x k x k 1 for k = 1 , 2 , , respectively .
Throughout, we use the convention that every term with an index 0 is equal to 0.
The original Hahn space:
h = x = ( x k ) k = 1 ω : k = 1 k | Δ x k | < c 0
was introduced by Hahn in 1922 [36] in connection with the theory of singular integrals. K. C. Rao showed [37] that h is a B K space with A K with the norm:
x = k = 1 k | Δ x k | ( x h ) .
Goes [38] introduced and studied the generalized Hahn space:
h d = x = ( x k ) k = 1 ω : k = 1 d k | Δ x k | < c 0 ,
where d = ( d k ) k = 1 is a given sequence of positive real numbers d k ( k = 1 , 2 , ) . If d k = k for all k, then h d reduces to the original Hahn space h, and if d = e then h e = b v 0 .
Let 1 p < . The sets:
w 0 p = x = ( x k ) k = 1 ω : lim n 1 n k = 1 n | x k | p = 0 , w p = x = ( x k ) k = 1 ω : x ξ e w 0 p for some ξ C
and
w p = x = ( x k ) k = 1 ω : sup n 1 n k = 1 n | x k | p <
of sequences that are strongly summable to zero, strongly summable and strongly bounded by the Cesàro method of order 1, with index p, were first introduced and studied by I. J. Maddox [39]. We write w 0 = w 0 1 , w = w 1 and w = w , for short.
The sets:
[ c 0 ] = x = ( x k ) k = 1 ω : Δ ( ( k x k ) k = 1 ) w 0 , = x = ( x k ) k = 1 ω : x ξ [ c 0 ] for some ξ C
and:
[ c ] = x = ( x k ) k = 1 ω : Δ ( ( k x k ) k = 1 ) w
of sequences that are strongly convergent to zero, strongly convergent, and strongly bounded were introduced and studied by Kuttner and Thorpe [40] and later generalized and studied in [41,42].
Throughout, we assume that the sequence d for h d is always a monotone increasing unbounded sequence of positive real numbers.
The following result holds.
Theorem 14
([43], Proposition 2.1). The space ( h d , · h d ) is a B K space with A K , where:
x h d = k = 1 d k | Δ x k | for all x h d .
The following example shows that h d may not have A K , in general, if the sequence d is not monotone increasing.
Example 8.
Let d = ( d k ) k = 1 and x = ( x k ) k = 1 be the sequences with:
d k = k ( k = 2 ν ) 1 ( k 2 ν ) ( ν = 0 , 1 , )
and:
x K = 0 ( k = 1 , 2 , 3 )   a n d   x k = 1 k ( k 2 ν + 1 ) 1 k 1 ( k = 2 ν + 1 ) ( ν = 2 , 3 , ) .
Then, clearly, x c 0 , and also,
x h d = 1 4 + k = 4 , k 2 ν | x k x k + 1 | ,
where,
x k x k + 1 = 1 k 1 1 k + 1 = 2 k 2 1 ( k = 2 ν + 1 ) 1 k 1 k + 1 = 1 k ( k 1 ) ( k 2 ν + 1 ) ( ν = 2 , 3 , ) ,
hence, x h d < . Thus, we have x h d .
On the other hand, let ν 2 be given. Then we have for x [ 2 ν ] ,
x x [ 2 ν ] h d d 2 ν x 2 ν x 2 ν + 1 x 2 ν [ 2 ν ] x 2 ν + 1 [ 2 ν ] = 2 ν 1 2 ν 1 2 ν 1 2 ν 0 = 1 ,
hence x [ m ] 0 as m .
Let:
b s d = a ω : sup n 1 d n k = 1 n a k <
and:
a b s d = sup n 1 d n k = 1 n a k   for   all   a b s d .
Remark 9.
Since b s is a B K space with a b s = sup n | k = 1 n a k | for all a b s by ([13], Example 4.3.17), and b s d is the matrix domain in b s of the triangle T = ( t n k ) n , k = 1 with t n k = 1 / d n for 1 k n and n = 1 , 2 , , b s d is a B K space with · b s d by ([13], Theorem 4.3.12).
Theorem 15
([43], Proposition 2.3). The spaces h d * and h d β of h d are norm isomorphic.
Now, we list the fundamental topological properties of the sets w 0 p , w p , w p ( 1 p < ) , [ c 0 ] , [ c ] and [ c ] . The results are analogous to those for c 0 , c and in Example 2.
Theorem 16. 
(a) ([39]) Let 1 p < . Then the sets w 0 p , w p , and w p are B K spaces with their natural norms:
x w p = sup n 1 n k = 1 n | x | p 1 / p ;
w 0 p is a closed subspace of w p and w p is a closed subspace of w ; w 0 p has A K , every sequence x = ( x k ) k = 1 w p has a unique representation (2), where ξ is the unique complex number such that x ξ · e w 0 p ; w p has no Schauder basis.
(b) 
([42], Theorem 2)The sets [ c 0 ] , [ c ] , and [ c ] are B K spaces with their natural norms
x [ c ] = sup n 1 n k = 1 n | Δ ( k x k ) | ;
[ c 0 ] is a closed subspace of [ c ] and [ c ] is a closed subspace of [ c ] ; [ c 0 ] has A K , every sequence x = ( x k ) k = 1 [ c ] has a unique representation (2), where ξ is the unique complex number such that x ξ · e [ c 0 ] ; [ c ] has no Schauder basis.

3.1. Some Classes of Bounded Linear Operators on the Generalized Hahn Space

In this subsection, we characterize the classes B ( h d , Y ) where Y is any of the spaces h d , w 0 p , w p , w p for 1 p < , [ c 0 ] , [ c ] and [ c ] . We also establish formulas for the norm of the corresponding operators.
We recall the following concept and results needed in the proofs of our characterizations.
Definition 9. 
([13], Definition 7.4.2) Let X be a B K space. A subset E of the set ϕ called a determining set for X if D ( X ) = B ¯ X ϕ is the absolutely convex hull of E.
Proposition 2
([43], Proposition 3.2). Let,
s ( d , k ) = 1 d k · e [ k ]   f o r   e a c h   k N ,   a n d   E = { s ( d , k ) : k N } .
Then E is a determining set for h d .
Proposition 3
([13], Theorem 8.3.4).
Let X be a B K space with A K , E be a determining set for X, and Y be an F K space. Then, A ( X , Y ) if and only if:
(i) The columns of A belong to Y, that is, A k = ( a n k ) n = 1 Y for all k,
and,
(ii) L ( E ) is a bounded subset of Y, where L ( x ) = A x for all x X
Since ( h d ) is a B K space with A K by Theorem 14, and the spaces Y for Y = w 0 p , w p , w p ( 1 p < ) , [ c 0 ] , [ c ] , and [ c ] are B K spaces by Theorem 16, it follows from Theorem 3 that L B ( h d , Y ) if and only if A ( h d , Y ) , where A is the infinite matrix that represents L as in (4). We are going to use this throughout.
Theorem 17
([43], Theorem 3.9 and Corollary 3.15 (a)).
We have L B ( h d ) if and only if:
lim n a n k = 0 , for all k ,
and:
A ( h d , h d ) = sup m 1 d m n = 1 d n k = 1 m ( a n k a n + 1 , k ) < .
Moreover, if L B ( h d ) then:
L = A ( h d , h d ) .
Proof. 
Since h d is a B K spaces with A K by Theorem 14, we apply Proposition 3 and observe that:
E = y ( m ) = 1 d m e [ m ] : m N
is a determining set for h d by Proposition 2.
First, the condition in (ii) of Proposition 3 is:
sup m A y ( m ) h d < for all y ( m ) E
and:
A y ( m ) c 0 for all y ( m ) E .
First, we obtain:
A y ( m ) h d = n = 1 d n A n y ( m ) A n + 1 y ( m ) = 1 d m n = 1 d n k = 1 m ( a n k a n + 1 , k )
for all m N , and so (26) is (24).
It is easy to see that (27) and (23) are equivalent.
Now, we show that condition (i) in Proposition 3 is redundant. Since A k c 0 for each k by (23), it follows from (24) that:
A k h d = n = 1 d n | a n k a n + 1 , k | = n = 1 d n j = 1 k ( a n j a n + 1 , j ) j = 1 k 1 ( a n j a n + 1 , j ) d k n = 1 d n A n y ( k ) A n y ( k ) + d k 1 n = 1 d n A n y ( k 1 ) A n y ( k 1 ) = d k A y ( k ) h d + d k 1 A y ( k 1 ) h d = d k A y ( k ) h d + d k 1 A y ( k 1 ) h d < for all k .
Finally, we show that L B ( h d ) implies (25).
We write B for the matrix with the rows B n = A n A n + 1 for all n. Let m N be given. Then, we have by Abel’s summation by parts for each n:
L n ( x [ m ] ) L n + 1 ( x [ m ] ) = k = 1 m b n k x [ m ] = k = 1 m 1 Δ x k j = 1 k b n j + x m j = 1 m b n j = k = 1 m 1 d k Δ x k 1 d k j = 1 k b n j + d m x m 1 d m j = 1 m b n j .
Since h d has A K and x h d , it follows that:
0 | d m x m | = k = m d k | Δ x k [ m ] | k = 1 d k Δ ( x k [ m ] x k ) + k = m d k | Δ x k | = x [ m ] x h d + k = m d k | Δ x k | 0   a s   m .
Furthermore, each functional L n is continuous, since h d is a B K space, and so for each n N and all x h d :
L n ( x ) L n + 1 ( x ) = k = 1 d k Δ x k 1 d k j = 1 k b n j ,
hence for all x h d
L ( x ) h d = n = 1 d n | L n ( x ) L n + 1 ( x ) | n = 1 d n k = 1 d k | Δ x k | 1 d k j = 1 k b n j = k = 1 d k | Δ x k | 1 d k n = 1 d n j = 1 k b n j sup k 1 d k n = 1 d n j = 1 k b n j · x h d = sup k 1 d k n = 1 d n j = 1 k ( a n j a n + 1 , j · x h d ,
that is,
L A ( h d , h d ) .
To show the converse inequality, let m N be given and x ( m ) = e [ m ] / d m . Then it follows that:
L ( x ( m ) ) h d = n = 1 d n A n x ( m ) A n + 1 x ( m ) = 1 d m n = 1 d n k = 1 m ( a n k a n + 1 , k ) L .
Since m N was arbitrary, we obtain A ( h d , h d ) L and this and (28) together imply (25). □
Remark 10.
It was shown in ([37], Proposition 10) that A ( h , h ) if and only if:
( i ) : lim n a n k = 0 ,
( ii ) : n = 1 n | a n m a n + 1 , m |   c o n v e r g e s   f o r   m = 1 , 2 , ,
( iii ) : sup m 1 m n = 1 n k = 1 m ( a n k a n + 1 , k ) < .
It seems that the condition in (ii) is redundant.
Proof. 
We show more generally that (24) implies:
( iv ) : n = 1 d n | a n m a n + 1 , m |   converges   for   all   m .
Let (24) be satisfied. Then:
M m = n = 1 d n k = 1 m ( a n k a n + 1 , k ) < for all m ,
hence:
n = 1 d n | a n m a n + 1 , m | = n = 1 d n k = 1 m ( a n k a n + 1 , k ) k = 1 m 1 ( a n k a n + 1 , k ) M m + M m 1 < for all m ,
that is, (iv) is satisfied. □
Theorem 18 
(([44], Theorem 3.3) for p = 1 and ([45], Theorem 3.4) for p > 1 ).
We have:
(a) 
L B ( h d , w p ) if and only if:
A ( h d , w p ) = sup l , m 1 d m 1 l n = 1 l k = 1 m a n k p 1 / p < ;
(b) 
L B ( h d , w p ) if and only if (29) holds and:
f o r   e a c h   k N ,   t h e r e   e x i s t s   α k C   s u c h   t h a t lim l 1 l n = 1 l | a n k α k | p = 0 ;
(c) 
L B ( h d , w 0 p ) if and only if (29) holds and:
lim l 1 l n = 1 l | a n k | p = 0   f o r   e a c h   k .
(d) 
If L B ( h d , Y ) for Y { w p , w p , w 0 p } , then,
L = A ( h d , w p ) .
Theorem 19
([46], Theorem 2.4). We have:
(a) 
L B ( h d , [ c ] ) if and only if,
A ( h d , [ c ] ) = sup l , m 1 l d m n = 1 l n k = 1 m a n k ( n 1 ) k = 1 m a n 1 , k < ;
(b) 
L B ( h d , [ c ] ) if and only if (33) holds and,
f o r   e a c h   k N ,   t h e r e   e x i s t s   α k C   s u c h   t h a t lim l 1 l n = 1 l | n a n k ( n 1 ) a n 1 , k α k | = 0 ;
(c) 
L B ( h d , [ c 0 ] ) if and only if (33) holds and,
lim l 1 l n = 1 l | n a n k ( n 1 ) a n 1 , k | = 0   f o r   e a c h   k .
(d) 
If L B ( h d , Y ) for Y { [ c 0 ] , [ c ] , [ c ] } , then,
L = A ( h d , [ c ] ) .

3.2. Some Classes of Compact Operators on the Generalized Hahn Space

Now, we study the Hausdorff measure of the bounded linear operators of Section 3.1 and the related classes of compact operators.
First, we consider the case of C ( h d ) .
Lemma 1
([43], Lemma 4.5). Let ( α n ) n = 1 , ( β n ) n = 1 and ( γ n ) n = 1 be given sequences of complex numbers, and A = ( a n k ) n , k = 1 be the tridiagonal matrix with:
a n k = α n ( k = n ) β n ( k = n + 1 ) γ n 1 ( k = n 1 ) 0 ( k n , n + 1 , n 1 ) ( n = 1 , 2 , ) .
Putting,
c m = 1 d m n = 1 d n k = 1 m ( a n k a n + 1 , k ) ,
we obtain,
c m = 1 d m n = 1 m 2 d n Δ ( α n + β n + γ n 1 ) + d m 1 Δ ( α m 1 + γ m 2 ) + β m 1 + d m α m + Δ γ m 1 + d m + 1 | γ m | .
For X , Y ω , M ( X , Y ) = { z ω : z x = ( z k x k ) Y for all x = ( x k ) X } is the multiplier of X in Y.
We obtain some useful special cases of Lemma 1.
Corollary 2
([43], Remark 4.6). (a) If α n = z n , β n = γ n = 0 for all n, then (37) reduces to:
c m = 1 d m n = 1 m 1 d n | Δ z n | + | z m | for all m ,
so z M ( h d , h d ) if and only if:
sup m c m < ,   o r   e q u i v a l e n t l y ,   1 d m · z [ m 1 ] h d m = 1 .
(b) 
Let l N be given and z = e e [ l ] , then we obtain from Part (a):
c m ( l ) = 0 ( 1 m l ) 1 + d l d m ( m l + 1 )
and so, since h d has A K ,
lim sup l R l = lim sup l sup m l c m ( l ) = lim sup l sup m l 1 + d l d m = 2 .
In the next result, we use the notation introduced at the beginning of the proof of Example 7.
Theorem 20. 
(a) ([43], Theorem 4.8 (a))Let L B ( h d ) . We write:
γ m ( l ) = 1 d m d l k = 1 m a l + 1 , k + n = l + 1 d n k = 1 m ( a n k a n + 1 , k )   f o r   a l l   m   a n d   l .
Then, we have:
1 2 · lim sup l sup m γ m ( l ) L χ lim sup l sup m γ m ( l ) .
(b) 
([43], Corollary 4.10 (d))The operator L B ( h d ) is compact if and only if:
lim l sup m γ m ( l ) = 0 ;
Proof. 
(a) We apply (16) with a = 2 by (38). We have by (24) and (25) for all l:
L < l > = A < l > ( h d , h d ) = sup m 1 d m n = 1 d n k = 1 m ( a n k < l > a n + 1 , k < l + 1 > ) = sup m γ m ( l )
and (39) follows by (19) and (16).
(b) Part (b) follows from (39) by (20). □
Theorem 21. 
(a) Let L B ( h d , w p ) . Then we have:
1 2 · lim r sup m ; l r 1 d m 1 l n = r l k = 1 m ( a n k α k ) p 1 / p L χ lim r sup m ; l r 1 d m 1 l n = r l k = 1 m ( a n k α k ) p 1 / p ,
where the complex numbers α k are defined in (30).
(b) 
Let L B ( h d , w 0 p ) . Then we have:
L χ = lim r sup m ; l r 1 d m 1 l n = r l k = 1 m a n k p 1 / p .
(c) 
Let L B ( h d , w p ) . Then L C ( h d , w p ) if and only if:
lim r sup m ; l r 1 d m 1 l n = r l k = 1 m ( a n k α k ) p 1 / p = 0 ,
where the complex numbers α k are defined in (30).
(d) 
Let L B ( h d , w 0 p ) . Then L C ( h d , w p ) if and only if:
lim r sup m ; l r 1 d m 1 l n = r l k = 1 m a n k p 1 / p = 0 .
Remark 11.
Parts (a) and (b) in Theorem 21 are ([44], Theorem 3.3) for p = 1 and ([45], Theorem 3.4) for p > 1 .
Parts (c) and (d) in Theorem 21 are ([44], Corollary 3.4) for p = 1 and ([45], Corollary 3.5) for p > 1 .
Theorem 22
([46], Theorem 3.4 and Corollary 3.5).
(a) 
Let L B ( h d , [ c ] ) . Then we have:
1 2 · lim r sup m ; l r 1 l d m n = r l k = 1 m ( n a n k ( n 1 ) a n 1 , k α k ) L χ lim r sup m ; l r 1 l d m n = r l k = 1 m ( n a n k ( n 1 ) a n 1 , k α k ) ,
where the complex numbers α k are defined in (34).
(b) 
Let L B ( h d , [ c 0 ] ) . Then we have:
L χ = lim r sup m ; l r 1 l d m n = r l k = 1 m ( n a n k ( n 1 ) a n 1 , k .
(c) 
Let L B ( h d , [ c ] ) . Then L C ( h d , [ c ] ) if and only if:
lim r sup m ; l r 1 l d m n = r l k = 1 m ( n a n k ( n 1 ) a n 1 , k α k ) = 0 ,
where the complex numbers α k are defined in (34).
(d) 
Let L B ( h d , [ c 0 ] ) . Then, L C ( h d , [ c 0 ] ) if and only if:
lim r sup m ; l r 1 l d m n = r l k = 1 m n a n k ( n 1 ) a n 1 , k = 0 .

4. Some Applications

We apply Theorem 17, Corollary 2 (a) and Theorem 20 (b) and get results by Sawano and El–Shabrawy ([47], Corollary 5.1 and Lemma 5.1).
Rhaly [48] defined the generalized Cesàro operator C t on ω for t [ 0 , 1 ) by the triangle C t = ( a n k ( t ) ) n , k = 0 , where a n k = t n k / ( n + 1 ) for 0 k n and n = 0 , 1 , .
Example 9
([47], Corollary 5.1). We have C t ( h , h ) for 0 t < 1 .
Proof. 
Clearly lim n a n k ( t ) = 0 for each k, so (23) in Theorem 17 holds.
We put a n k = a n k ( t ) for all n and k. We need show that (24) also holds.
We put:
c m ( n ) = n k = 1 m ( a n k a n + 1 , k )   and   c m = 1 m n = 1 c m ( n )   for   all   m   and   n .
If t = 0 , then C 0 = diag ( 1 / ( n + 1 ) ) is the diagonal matrix with the entries 1 / ( n + 1 ) on its diagonal.
Let m N be arbitrary.
For n m 1 , we obtain:
c m ( n ) = n 1 n + 1 1 n + 2 .
For n m , we have c m ( n ) = m / ( m + 1 ) for n = m and c m ( n ) = 0 for n m + 1 . For all m, it follows that:
c m = 1 m n = 1 m 1 c m ( n ) + c m ( m ) n = 1 m 1 1 n + 1 1 n + 2 + 1 m · m m + 1 1 2 + 1 2 1 ,
and so (24) also holds.
Now, let t ( 0 , 1 ) , and m N be arbitrary.
If n m 1 , then a n , k a n + 1 , k 0 for 0 k n and a n , k = 0 for k n + 1 . We get:
c m ( n ) k = 1 n t n k + 1 = k = 0 n t k 1 1 t
If n m , then a n k a n + 1 , k 0 for all k m . We get:
c m ( n ) k = 0 m t n k t n m k = 0 m t m k t n m 1 t .
Finally (44) and (45) imply:
c m = 1 m n = 1 m 1 c n ( m ) + n = m c n ( m ) 1 m m 1 1 t + n = 0 t n 1 t 2 t ( 1 t ) 2   for   all   m ,
hence, sup m c m < . Thus, (24) also holds. □
If d k = k for all k of the following example gives ([47], Lemma 5.1).
Example 10.
Let ( λ k ) k = 1 be a decreasing sequence of positive real numbers which converges to 0 and D ( λ ) = diag ( λ 1 , λ 2 , ) denote the diagonal matrix with the sequence λ on its diagonal. Then L D ( λ ) C ( h d ) .
Proof. 
Since d k d k + 1 and λ k λ k + 1 for all k, we have for all m:
c m = 1 d m k = 1 m 1 d k | Δ λ k | + | λ m | k = 1 m 1 ( λ k λ k + 1 ) + λ m = λ 1 ,
hence, λ M ( h d , h d ) by Corollary 2 (a), that is, L D ( λ ) B ( h d ) .
If l N is arbitrary, then γ m ( l ) = 0 for all m l , and:
γ m ( l ) = 1 d m n = l m 1 d n | Δ λ n | + | λ m | λ l + 1 + n = l + 1 m 1 ( λ n λ n + 1 ) + λ m = 2 λ l + 1 λ m + λ m 2 λ l + 1
for all m l + 1 . Hence,
0 lim l sup m γ m ( l ) 2 lim l λ l + 1 = 0 ,
and so L D ( λ ) is compact by Theeorem 20 (b). □
We obtain the following results for the classical Hahn space h.
Remark 12.
We have:
(a) 
(([44], Example 3.5) for p = 1 and ([45], Example 3.6) for 1 < p < ) L C 1 C ( h , w 0 p ) for 1 p < and L C 1 = 1 ;
(b) 
([46], Example 3.6) L C 1 C ( h d , [ c 0 ] ) and L C 1 = 1 .
If X and Y are Banach spaces, L B ( X , Y ) , then we denote by N ( L ) and R ( L ) denote the null space and the range of L, respectively. Now, L is called a Fredholm operator, if R ( L ) is closed, dim N ( T ) , dim X / R ( L ) | . In this case, the index L is given by i ( L ) = dim N ( L ) dim X / R ( L ) . Furthermore, if L B ( X , X ) and L χ < 1 , then I L is a Fredholm operator with i ( I L ) = 0 ([49] or ([9], Section 7.13)).
Corollary 3
([43], Corollary 4.13). Let α = ( α n ) n = 1 , β = ( β n ) n = 1 and γ = ( γ n ) n = 1 be given complex sequences, and:
A ( γ , α , β ) = α 1 β 1 0 0 γ 1 α 2 β 2 0 0 γ 2 α 3 β 3 0 0 0 γ n 1 α n β n 0 0 .
Then, the operator L B ( h d ) represented by the matrix:
A ( γ , α , β ) = A ( 0 , α , 0 ) + A ( γ , 0 , 0 ) + A ( 0 , 0 , β )
is Fredholm with i ( A ( α , β , γ ) ) = 0 , if A ( 0 , α , 0 ) is Fredholm with i ( A ( 0 , α , 0 ) ) = 0 and A ( γ , 0 , 0 ) and A ( 0 , 0 , β ) are compact.
Example 11
([43], Example 4.14). If d k = k , α k = 1 1 / k and β k = γ k = 1 / k for all k, then L B ( h d ) represented by A ( γ , α , β ) is Fredholm.
Proof. 
We write c m ( < l > ) ( α e ) , c m < l > ( γ ) and c m < l > ( β ) for the expressions in (37) for the matrices A ( 0 , α e , 0 ) , A ( γ , 0 , 0 , ) and A ( 0 , 0 , β ) . Then we get from (37):
c m < l > ( α e ) = 1 d m d l | α l + 1 1 | + n = l + 1 m 1 d n Δ α n + d m | α m 1 | = 1 m l l + 1 + n = l + 1 m 1 n 1 n 1 n + 1 + m m 2 l + n = l + 1 1 n 1 n + 1 .
Consequently:
L A ( 0 , α , 0 ) I χ = lim sup l sup m c m < l > ( α e ) < 1 ,
hence, L A ( 0 , α , 0 ) I is Fredholm.
Furthermore, (37) yields:
c m < l > ( γ ) = 1 d m d l | γ l | + n = l + 1 m d n Δ γ n 1 + d m + 1 | γ m | 1 m + n = 1 1 n 1 1 n + m + 1 m 2 3 l + 2 n = l 1 n 2 .
Thus,
L A ( γ , 0 , 0 ) χ = lim sup l sup m c m < l > ( γ ) = 0 ,
and L A ( γ , 0 , 0 ) is compact.
Analogously, we can show that the L A ( 0 , 0 , β ) is compact.
Thus, L A ( γ , α , β ) is Fredholm by Corollary 3. □

5. Some Mathematical Background

Now, we apply measures of noncompactness to the solvability of infinite systems of integral equations.
The notation M N C will stand for measures of noncompactness in Banach spaces in the sense of Banaś and Goebel given in Definition 6.
Hyperconvex spaces were introduced by Aronszajn and Panitchpakdi [50]. They are very important in metric fixed point theory, see [51] and the references therein.
Definition 10.
A metric space ( X , d ) is hyperconvex if every class of closed balls { B ¯ ( x i , r i ) } i I with d ( x i , x j ) r i + r j satisfies:
i I B ( x i , r i ) .
The following result holds.
Theorem 23
([52]). Let X be a hyperconvex metric space, x 0 X and let f be a continuous self–map of X. If the following implication:
(V is is ometric to  ε f ( V ) or V = f ( V ) { x 0 } ) ( α ( V ) = 0
where ε f ( V ) denotes hyperconvex hull of f ( V ) , holds for every subset V X , then f has a fixed point.
Theorem 23 can be applied in certain cases of continuous self–maps in hyperconvex metric spaces, where Darbo’s fixed point theorem, Theorem 9, or Darbo–Sadovskiĭ type fixed point theorems such as Theorem 11 are not applicable. This is illustrated in the following example.
Example 12
([52]). Consider R 2 with the radial metric:
d ( v 1 , v 2 ) = ρ ( v 1 , v 2 ) i f   0 ,   v 1 ,   v 2   a r e   c o l l i n e a r , ρ ( v 1 , 0 ) + ρ ( v 2 , 0 ) o t h e r w i s e ,
where ρ denotes the usual Euclidean metric and v 1 = ( x 1 , y 1 ) , v 2 = ( x 2 , y 2 ) R 2 . Define the map, f : R 2 R 2 by f ( x , y ) = ( h x , h y ) for ( x , y ) R 2 and h > 1 . Then f does not satisfy Darbo’s condensing condition, but it satisfies the hypotheses of Theorem 23. Hence, f has a fixed point.
Samadi [53] gave the following extension of Darbo’s fixed point theorem.
Theorem 24.
Let C be a bounded, closed and convex subset of a Banach space E. Assume T : C C is a continuous operator satisfying:
θ ( μ ( X ) ) + f ( μ ( T ( X ) ) ) f ( μ ( X ) )
for all nonempty subsets X of C, where μ is an arbitrary M N C on E and ( θ , f ) Δ , where Δ is the set of all pairs ( θ , f ) that satisfy the following conditions:
( Δ 1 )
θ ( t n ) 0 for each strictly increasing sequence { t n } ;
( Δ 2 )
f is strictly increasing function;
( Δ 3 )
for each sequence { α n } of positive numbers, lim n α n = 0 if and only if lim n f ( α n ) = .
( Δ 4 )
If { t n } is a decreasing sequence such that t n 0 and θ ( t n ) < f ( t n ) f ( t n + 1 ) , then we have n = 1 t n < .

5.1. Meir–Keeler Generalization

We continue with the famous result by Meir–Keeler [54] of 1969.
Definition 11.
Let ( X , d ) be a metric space. A self– map T on X is a Meir–Keeler contraction (MKC) if for any ε > 0 , there exists δ > 0 such that:
ε d ( x , y ) < ε + δ implies d ( T x , T y ) < ε ,
for all x , y X .
Theorem 25
([54]). Let ( X , d ) be a complete metric space. If T : X X is a Meir–Keeler contraction, then T has a unique fixed point.
Definition 12
([55]). Let C be a nonempty subset of a Banach space E and μ be an M N C on E. We say that an operator T : C C is a Meir–Keeler condensing operator if for any ε > 0 , there exists δ > 0 such that:
ε μ ( X ) < ε + δ implies μ ( T ( X ) ) < ε ,
for any bounded subset X of C.
We note that any M K C is also a Meir–Keeler condensing operator, if we take the M N C as diam ( X ) .
Theorem 26
([55]). Let C be a closed, bounded, and convex subset of a Banach space E and μ be an arbitrary M N C on E. If T : C C is continuous and a Meir–Keeler condensing operator, then T has at least one fixed point and the set of all fixed points of T in C is compact.
The characterization of Meir–Keeler contractions in metric spaces was studied by Lim [56] and Suzuki [57] by introducing notion of L–functions.
Definition 13
([56]). A slef–map ϕ on R + is called an L–function if ϕ ( 0 ) = 0 , ϕ ( s ) > 0 for s ( 0 , ) , and for every s ( 0 , ) there exists δ > 0 such that ϕ ( t ) s , for any t [ s , s + δ ] .
Theorem 27
([55]). Let, C be a bounded subset of a Banach space E, μ be an arbitrary M N C on E and T : C C be a continuous operator. Then, T is a Meir–Keeler condensing operator if and only if there exists an L–function ϕ such that:
μ ( T ( X ) ) < ϕ ( μ ( X ) ) ,
for all closed and bounded subset X of C with μ ( X ) 0 .
We need the following concept.
Definition 14
([58]). Let ( X , d ) be a metric space. Then, a mapping T : X X is said to be contractive if:
d ( T ( x ) , T ( y ) ) < d ( x , y )
for all x , y X with x y .
Theorem 28
(Edelstein [58]). Let ( X , d ) be a compact metric space. If T is a contractive map on X, then there exists a unique fixed point z X .
Definition 15.
Let C be a bounded subset of a Banach space E, and μ an M N C on E. Then, a self–map T on C is an asymptotic Meir–Keeler condensing operator if there exists a sequence ( ϕ n ) of self–maps on R + satisfying the following conditions:
(A1) 
For each ε > 0 , there exists δ > 0 and ν N such that ϕ ν ( t ) ε for any t [ ε , ε + δ ] ,
(A2) 
μ ( T n ( C ) ) < ϕ n ( μ ( C ) ) , n N .
In the next theorem, the convexity condition of the set C in the previous results is replaced by assumption that the operator T is contractive.
Theorem 29.
Let C be a bounded and closed (not necessarily convex) subset of a Banach space E, and μ be an M N C on E. Let T : C C be a contractive and asymptotic Meir–Keeler condensing operator. Then, T has a unique fixed point in C.
Proof. 
We define a sequence ( C n ) by putting C 0 = C and C n = T n C ¯ for n 1 . Since T is contractive and continuous, it follows that T ( A ¯ ) T ( A ) ¯ . This inclusion yields T n + 1 C T n C , so C n + 1 C n and T ( C n ) C n . If μ ( C N ) = 0 for some integer N 0 , then C N is compact. Hence, T has a fixed point by Theorem 28. Now we suppose that μ ( C n ) 0 for n 0 . We put ε n = μ ( C n ) and r = inf n N ε n . We prove r = 0 . If r 0 , then by the definition of r, and the conditions in (A1) and (A2), there exist n 0 N , δ r > 0 , and ν N such that ϕ ν ( t ) r for any t [ r , r + δ r ] and r ε n 0 < r + δ r . Consequently,
ε n 0 + ν = μ ( C n 0 + ν ) = μ ( T n 0 + ν ( C ) ) < ϕ ν ( μ ( T n 0 ( C ) ) ) = ϕ ν ( μ ( C n 0 ) ) r .
This is a contradiction, so r = 0 . Hence, lim n μ ( C n ) = 0 . Since C n + 1 C n and T C n C n for all n 1 , the generalized Cantor intersection property of the M N C μ yields the C = n = 1 C n is nonempty and closed, invariant under T, and belongs to k e r μ . Then, by Theorem 28, T has a unique fixed point in C . Furthermore, since F T = { x X : T ( x ) = x } C n for all n 0 , it follows that F T C and T has a unique fixed point in C. □

5.2. Darbo-Type Theorem for Commuting Operators

Now we are going to discuss some fixed point theorems obtained in [59,60] for commuting maps in locally convex spaces and Banach spaces, satisfying the following inequalities:
α ( S ( A ) ) k sup i I ( α ( T i ( A ) ) )
and:
α ( S ( A ) ) < sup i I ( α ( T i ( A ) , α ( A ) ) .
We briefly describe M N C ’s on locally convex spaces. Let X be a Hausdorff complete and locally convex space whose topology is defined by family of equicontinuous seminorms P . A local base of closed 0–neighborhood of X is generated by the sets:
{ x X : max 1 i n p i ( x ) ε } , ε > 0 , p i P .
Let B denote the family of all bounded subsets of X and Φ be the space of all functions ϕ : P R + with the partial order “ ϕ 1 ϕ 2 if and only if ϕ 1 ( p ) ϕ 2 ( p ) for all p P ”.
Definition 16.
A measure of noncompactness on a locally convex space is the function γ from B into Φ such that for each B B , we have that γ ( B ) is a function from P into R + , such that:
γ ( B ) ( p ) = inf { d > 0 : B is a finite union of B i , sup { p ( x y ) : x , y B i } d p P } .
Remark 13
([60]). On a Hausdorff, complete locally convex space, γ satisfies the generalized Cantor intersection property.
Definition 17.
A mapping T of a convex set M is said to be affine if:
T ( k x + ( 1 k ) y ) = k T x + ( 1 k ) T y ,
whenever 0 < k < 1 and x , y M .
The following result holds.
Theorem 30
([60]). Let X be a Hausdorff complete and locally convex space, Ω be a convex, closed and bounded subset of X, I be an index set, and { T i } i I , S be a continuous function from Ω into Ω such that the following conditions hold:
(a) 
For any i I , T i commutes with S.
(b) 
For any A Ω and i I , we have T i ( c o ¯ ( A ) ) c o ¯ ( T i ( A ) ) .
(c) 
There exists 0 < k < 1 such that for any A Ω   α ( S ( A ) ) ( p ) k sup i I α ( T i ( A ) ) ( p ) , p P .
Then we have:
(1) 
The set { x Ω : S x = x } is nonempty and compact.
(2) 
For any i I , set { x Ω : T i x = x } is nonempty, closed and invariant by S.
(3) 
If T i is affine and { T i } i I is a commuting family then T i and S have a common fixed point and the set { x Ω : T i ( x ) = S ( x ) = x } is compact.
(4) 
If { T i } i I is a commuting family and S is affine, then there exists a common fixed point for the mapping { T i } i I .
Remark 14.
If T i is the identity function for any i I , above theorem becomes generalization of Darbo’s fixed point theorem in the structure of locally convex spaces.
The following theorem due to [59] generalizes the Sadovskiĭ fixed point theorem for commuting operators.
Theorem 31.
Let X be a Hausdorff complete and locally convex space, Ω be a convex, closed and bounded subset of X, I be an index set, and { T i } i I , S be a continuous function from Ω into Ω such that:
(a) 
For each i I , T i commutes with S.
(b) 
For each i I , T i is linear map.
(c) 
There exists j I such that for each A Ω and p P , with α ( A ) ( p ) 0 , we have:
α ( S ( A ) ) ( p ) < sup ( α ( T j ( A ) ) ( p ) , α ( A ) ( p ) ) .
Then we have:
(1) 
T j and S have a fixed point, and { x Ω : T j x = x } is compact.
(2) 
If { T i } i I is a commuting family and S is affine, then there exists a common fixed point for the mapping in { T i } i I .
Remark 15.
If T j is the identity function, then above theorem becomes a generalization of Sadovskii’s fixed point theorem.
It is well known for operators S and T that if the composition operator S T has a fixed point, then S and T do not necessarily poss a fixed point or a common fixed point. It becomes interesting to investigate the conditions which force the operator S, T to have a common fixed point. This result is also helpful in obtaining existence results for common solutions of a certain type of equations.
Theorem 32
([59]). Let X be a Banach space and Ω be a convex, closed, and bounded subset of X. Let T and S be two continuous functions from Ω into Ω such that:
(a) 
S T = T S ;
(b) 
T is affine;
(c) 
There exist k ( 0 , 1 ) such that for any A Ω we have α ( S T ( A ) ) k α ( A ) .
Then, the set { x Ω : T x = S x = x } is nonempty and compact.
Proof. 
The operator H with H ( x ) = k S ( T ( x ) ) + ( 1 k ) T ( x ) is a continuous self–map H on Ω and, commutes with T.
The semi–homogeneity and sub–additive property of the M N C   α imply:
α ( H ( A ) ) = α ( k S ( T ( A ) ) ) + ( 1 k ) T ( A ) k 2 α ( A ) + ( 1 k ) α ( T ( A ) )
for any A Ω . Since k ( 0 , 1 ) , k 2 < k , and we have k 2 + 1 k < k + 1 k . Hence, it follows from Theorem 30 that F 0 = { x Ω : H x = T x = x } is compact.
Moreover, we have for any x F 0 :
H ( x ) = k S T ( x ) + ( 1 k ) T ( x ) = T x = x implies S x = x .
So S and T have a common fixed point. We put F = { x Ω : S x = T x = x } . Then,
α ( F ) = α ( S T ( F ) ) k α ( F )
implies α ( F ) = 0 . Since S and T are continuous, F is compact. □
Remark 16.
If the operator T is equal to the identity function, then we obtain Darbo’s fixed point theorem from Theorem 32.
Theorem 33
([59]). Let X be a Banach space and Ω be a convex, closed and bounded subset of X. Let T 1 , T 2 , and S be two continuous self–maps on Ω such that:
(a) 
T 1 T 2 = T 2 T 1 ;
(b) 
T 1 , T 2 are affine;
(c) 
There exist k ( 0 , 1 ) such that for any A Ω we have α ( S ( A ) ) k α ( A ) .
Then, the set { x Ω : S x = T 1 x = T 2 x = x } is compact.

6. Applications to Integral Equations

Now we apply measures of noncompactness to solve some differential and integral equations, and systems of linear equations in sequence spaces. Furthermore, we discuss existence results obtained by various authors, for the solution of integral equations in some sequence spaces.
We use the standard notations and results for functions of bounded variation, their total variation and the Riemann–Stieltjes integral (cf. [55]).

6.1. Infinite System of Integral Equations of Volterra–Stieltjes Type In Sequence Spaces p and C 0

We study the solutions for an infinite system of integral equations of the Volterra–Stieltjes type of the form (see [61]):
u n ( t , x ) = F n ( t , s , f 1 ( t , u ( t , x ) ) 0 t 0 x g n t , s , x , y , u ( t , x ) d y g 2 ( x , y ) d s g 1 ( t , s ) , ( T u ) ( t , x ) 0 V n ( t , s , u ( t , x ) ) d s ) ; u ( t , x ) = u i ( t , x ) i = 1 , u i ( t , x ) B C ( R + × R + , R ) ,
where B C ( R + × R + , R ) is the space of all real functions u ( t , x ) = u : R + × R + R , which are defined, continuous, and bounded on the set R + × R + with the supremum norm:
u = sup | u ( t , x ) | : ( t , x ) R + × R + .

6.1.1. Solution in the Space p   ( 1 p < )

We consider the following hypotheses:
( H 1 ) :
F n : R + × R + × R × R R is continuous and there are reals τ > 0 with:
| F n ( t , s , x 1 , y 1 ) F n ( t , s , x 2 , y 2 ) | p e τ ( | x 1 x 2 | p + | y 1 y 2 | p ) ,
for all t , s R + and x 1 , x 2 , y 1 , y 2 R . Moreover, we have:
lim i Σ i = 1 | F i ( t , s , 0 , 0 ) | p = 0 , N 1 = Σ i = 1 | F i ( t , s , 0 , 0 ) | p .
( H 2 ) :
f 1 : R + × R R is continuous with f 0 = sup t R + | f ( t , 0 ) | and there are reals τ > 0 with:
| f 1 ( t , u ( t , x ) ) f 1 ( t , v ( t , x ) ) | p e τ u ( t , x ) v ( t , x ) p , | f 1 ( t , u ( t , x ) ) | p e τ u ( t , x ) p .
for all t , x R + and:
u ( t , x ) = u i ( t , x ) i = 1 , v ( t , x ) = v i ( t , x ) i = 1 p .
( H 3 ) :
T : B C ( R + × R + , p ) B C ( R + × R + , R ) is a continuous operator satisfying:
| ( T u ) ( t , x ) ( T v ) ( t , x ) | u ( t , x ) v ( t , x ) p , | ( T u ) ( t , x ) | 1 .
for all u , v B C ( R + × R + , p ) and t , x R + .
( H 4 ) :
For any fixed t > 0 the function s g i ( t , s ) is of bounded variation on the interval [ 0 , t ] and the function t s = 0 t g i ( t , s ) is bounded over R + .
( H 5 ) :
g n : R + × R + × R + × R + × R R is continuous and there exist continuous functions a n : R + × R + R + such that:
| g n ( t , s , x , y , u ( t , x ) ) | a n ( t , s ) , lim t Σ n 1 0 t | g n ( t , s , x , y , u ( t , x ) ) g n ( t , s , x , y , v ( t , x ) ) | d s q = 0 t g 1 ( t , q ) = 0 , φ k = sup { Σ n k | 0 t 0 x g n t , s , x , y , u ( t , x ) d y g 2 ( x , y ) d s g 1 ( t , s ) | ; t , s , x , y R + , u ( t , x ) R } .
We also put:
A = sup Σ n = 1 0 t a n ( t , s ) d s p = 0 s g 1 ( t , p ) , t R + , G = sup y = 0 x g 2 ( x , y ) ; x R + , lim k φ k = 0 .
( H 6 ) :
V n : R + × R + × R R is a continuous function and there exists a continuous function k : R + × R + R + such that the function s k ( t , s ) is integrable over R + satisfying:
| V n ( t , s , u ( t , x ) ) | k ( t , s ) | u n ( t , x ) | p , | V n ( t , s , u ( t , x ) ) V n ( t , s , v ( t , x ) | | u n ( t , x ) v n ( t , x ) | p k ( t , s ) .
for all t , s , x R + and u , v p . We put:
M = sup t R + 0 k ( t , s ) d s .
( H 7 ) :
There exists a solution r 0 > 0 with:
2 2 p e 2 τ r 0 p ( G A ) p + 2 2 p e τ f 0 p ( G A ) p + 2 p e τ r 0 p M p + 2 p N 1 r 0 p ,
Moreover, assume that 2 p M < 1 .
Theorem 34.
Under the assumptions ( H 1 ) ( H 7 ) , Equation (47) has at least one solution u ( t , x ) = u i ( t , x ) i = 1 in p .
Example 13.
Here, we investigate the system of integral equations:
u n ( t , x ) = ( e τ t n ) 1 p 2 sin ( ( e t τ ) 1 p sin u ( t , x ) p 2 × 0 t 0 x arctan 1 2 n × e 3 t + s 8 + | x | + | y | + | u n ( t , x ) | e x 1 + y 2 e 2 x e t 1 + t 2 d y d s + cos 1 1 + u ( t , x ) l p 0 e s 1 + t 8 sin | u n ( t , x ) | d s ) .
We observe that Equation (48) is a special case of (47) putting:
F n t , s , x , y = ( e τ t n ) 1 p 2 sin x + y , g n ( t , s , x , y , u ( t , x ) ) = arctan 1 2 n × e 3 t + s 8 + | x | + | y | + | u n ( t , x ) | , f 1 ( t , u ( t , x ) ) = ( e t τ ) 1 p sin u ( t , x ) p 2 , a n ( t , s ) = 1 2 n e 3 t + s , g 1 ( t , s ) = s e t 1 + t 2 , g 2 ( x , y ) = arctan y e x , V n ( t , s , u ( t , x ) ) = e s 1 + t 8 sin | u n ( t , x ) | , k ( t , s ) = e s 1 + t 8 , ( T u ) ( t , x ) = cos 1 1 + u ( t , x ) l p .
Obviously, F n and f 1 satisfy ( H 1 ) and ( H 2 ) with N 1 = 0 and f 0 = 0 , T satisfies ( H 3 ) . To check ( H 5 ) , we assume t , s x , y R + and u , u p . It follows that:
| g n ( t , s , x , y , u ( t , x ) ) | 1 2 n e 3 t + s = a n ( t , s ) .
We obtain from g 1 s = e t 1 + t 2 > 0 that: q = 0 s g 1 ( t , q ) = g 1 ( t , s ) g 1 ( t , 0 ) = s e t 1 + t 2 . Consequently, we have:
lim t 0 t a n ( t , s ) d s q = 0 s g 1 ( t , q ) = lim t 0 t 1 2 n e 3 t + s ( e t 1 + t 2 ) d s = lim t 1 2 n e 2 t + s 1 + t 2 | 0 t = 0 ,
hence,
lim t Σ n 1 0 t | g n ( t , s , x , y , u ( t , x ) ) g n ( t , s , x , y , v ( t , x ) ) | d s q = 0 t g 1 ( t , q ) = 0 , A = sup Σ i = 1 0 t a n ( t , s ) d s p = 0 s g 1 ( t , s ) , t R + , φ k = sup { Σ n k 0 t 0 x g n t , s , x , y , u ( t , x ) d y g 2 ( x , y ) d s g 1 ( t , s ) ; t , s , x , y R + , u ( t , x ) p } G e 2 t 1 + t 2 e t 1 + t 2 Σ n k 1 2 n .
Thus, φ k 0 . Furthermore, V n ( t , s , u ( t , x ) ) = e s 1 + t 8 sin | u n ( t , x ) | verifies ( H 6 ) with k ( t , s ) = e s 1 + t 8 and M = 1 . To establish that g 1 and g 2 satisfy assumption ( H 4 ) , we observe that g 1 and g 2 are increasing on every interval [ 0 , t ] and g 2 is bounded on the triangle 2 . Therefore, the function y g 2 ( x , y ) is of bounded variation on [ 0 , x ] and:
y = 0 x g 2 ( x , y ) = g 2 ( x , y ) g 2 ( x , 0 ) = g 2 ( x , y ) π 4 .
Thus, G π / 4 and we may choose G = π / 4 .
Therefore, by Theorem 34, the infinite system (48) has at least one solution in p .

6.1.2. Solution in the Space C 0

Now we study the system (47) and consider the following assumptions.
( D 1 )
F n : R + × R + × R × R R is continuous and there exist positive reals τ with:
| F n ( t , s , x 1 , y 1 ) F n ( t , s , x 2 , y 2 ) | e τ ( | x 1 x 2 | + | y 1 y 2 | ) ,
for all t , s R + and x 1 , x 2 , y 1 , y 2 R . Moreover, assume:
lim i | F i ( t , s , 0 , 0 ) | = 0 , M 1 = sup | F i ( t , s , 0 , 0 ) | ; t , s R + , i 1 .
( D 2 )
f 1 : R + × R R is continuous with f 0 = sup t R + | f ( t , 0 ) | and there exist positive reals τ with:
| f 1 ( t , u ( t , x ) ) f 1 ( t , v ( t , x ) ) | e τ sup n 1 | u i ( t , x ) v i ( t , x ) | ; i n , | f 1 ( t , u ( t , x ) ) | e τ sup n 1 | u i ( t , x ) | ; i n
for all t , x R + and u ( t , x ) = u i ( t , x ) , v ( t , x ) = v i ( t , x ) c 0 .
( D 3 )
T : B C ( R + × R + , c 0 ) B C ( R + × R + , R ) is a continuous operator satisfying:
| ( T u ) ( t , x ) ( T v ) ( t , x ) | sup n 1 | u i ( t , x ) v i ( t , x ) | ; i n , | ( T u ) ( t , x ) | 1 .
for all u , v B C ( R + × R + , c 0 ) and t , x R + .
( D 4 )
For any fixed t > 0 the functions s g i ( t , s ) are of bounded variation on [ 0 , t ] and the functions t s = 0 t g i ( t , s ) are bounded on R + . Furthermore, for arbitrary, fixed positive T, the function w z = 0 w g i ( w , z ) is continuous on [ 0 , T ] for i = 1 , 2 .
( D 5 )
g n : R + × R + × R + × R + × R R is continuous and there exist continuous functions a n : R + × R + R + with:
| g n ( t , s , x , y , u ( t , x ) ) | a n ( t , s ) , lim t 0 t | g n ( t , s , x , y , u ( t , x ) ) g n ( t , s , x , y , v ( t , x ) ) | d s q = 0 t g 1 ( t , q ) = 0 ,
for all t , s , x , y R + and u , v R . Furthermore, we suppose that:
lim n 0 t a n ( t , s ) d s p = 0 s g 1 ( t , p ) = 0 , A = sup 0 t a n ( t , s ) d s p = 0 s g 1 ( t , p ) ; n N , G = sup y = 0 x g 2 ( x , y ) ; x R + , G 1 = sup z = 0 w g 1 ( w , z ) ; w [ 0 , T ] .
where T is an arbitrary fixed positive real number.
( D 6 )
V n : R + × R + × R R is a continuous function and there exists continuous function k : R + × R + R + such that the function s k ( t , s ) is integrable over R + and the following conditions hold:
| V n ( t , s , u ( t , x ) ) | k ( t , s ) sup n 1 | u i ( t , x ) | ; i n , | V n ( t , s , u ( t , x ) ) V n ( t , s , v ( t , x ) | sup n 1 | u i ( t , x ) v i ( t , x ) ; i n k ( t , s ) .
for all t , s , x R + and u , v c 0 . Furthermore, we suppose:
M = sup t R + 0 k ( t , s ) d s < 1 , e 2 τ G A + f 0 G A e τ + M e τ + M e τ < 1 .
Theorem 35.
If the infinite system (47) satisfies ( D 1 ) ( D 6 ) , then it has at least one solution u ( t ) = ( u i ( t , x ) ) i = 1 in c 0 .
Example 14.
Now we investigate:
u n ( t , x ) = e t s τ n arctan e τ Σ k n | u k ( t , x ) | 1 + k 2 ( H n ) ( u ) 5 + ( D n ) ( u ) 7 3
in c 0 . Writing:
( D n ) ( u ) = e 100 Σ k n sin | u k ( t , x ) | ( 1 + k 2 ) 0 e t s n Σ k n | u k ( t , x ) | 10 n ( 1 + k 2 ) d s , ( H n ) ( u ) = 0 t 0 x arctan e s + t 2 n 8 + | u ( t , x ) | e 2 t 1 + t 2 × e x 1 + y 2 e 2 x d y d s , F n ( t , s , x , y ) = e τ t s n x 5 + y 7 3 , f 1 ( t , u ( t , x ) ) = arctan e τ Σ k n | u k ( t , x ) | 1 + k 2 , g n ( t , s , x , y , u ( t , x ) ) = arctan e s + t 2 n 8 + | u ( t , x ) | , g 1 ( t , s ) = s e 2 t 1 + t 2 , g 2 ( x , y ) = arctan y e x , V n ( t , s , u ( t , x ) ) = e t s n Σ k n | u k ( t , x ) | 10 n ( 1 + k 2 ) , k ( t , s ) = e t s , ( T u ) ( t , x ) = e 100 Σ k n sin | u k ( t , x ) | ( 1 + k 2 ) n N ,
in (47), we obtain (49). We observe that F n and f 1 satisfy ( D 1 ) and ( D 2 ) . Indeed, we have:
| F n ( t , x 1 , y 1 ) F n ( t , x 2 , y 2 ) | = e τ n t x 1 5 + y 1 7 3 x 1 5 + y 1 7 3 e τ x 1 5 + y 1 7 x 2 5 y 2 7 3 e τ x 1 x 2 5 + y 1 y 2 7 3 e τ | x 1 x 2 | + | y 1 y 2 | , M 1 = 0 , lim n F n t , s , 0 , 0 = 0 , | f 1 t , u ( t , x ) | sup n 1 | u i ( t , x ) | ; i n , | f 1 t , u ( t , x ) f 1 t , v ( t , x ) | sup n 1 | u i ( t , x ) | | v i ( t , x ) | ; i n .
Obviously, T satisfies ( D 3 ) and:
| ( T u ) ( t , x ) | e 100 π 2 6 sup n 1 | u i ( t , x ) | ; i n , | ( T u ) ( t , x ) ( T v ) ( t , x ) | e τ π 2 6 sup n 1 | u i ( t , x ) v i ( t , x ) | ; i n .
Moreover, since:
g 1 s = e 2 t 1 + t 2 > 0 ,
the function g 1 is increasing and we obtain:
q = 0 s g 1 ( t , q ) = g 1 ( t , s ) g 1 ( t , 0 ) = g 1 ( t , s ) = s e 2 t 1 + t 2 > 0 .
Consequently,
| g n ( t , s , x , y , u ( t , x ) ) | e s + t 2 n , lim t 0 t | g n ( t , s , x , y , u ( t , x ) ) g n ( t , s , x , y , v ( t , x ) ) | d s q = 0 t g 1 ( t , q ) 2 lim t 0 t e t + s e 2 t 1 + t 2 d s = 0 .
Again, we have:
q = 0 y g 2 ( x , y ) = g 2 ( x , y ) g 2 ( x , 0 ) = g 2 ( x , y ) π 4 , lim n 0 t a n ( t , s ) d s q = 0 s g 1 ( t , q ) = lim n 2 n ( 1 1 + t 2 e t 1 + t 2 ) = 0 .
Hence, G = π / 4 and A < . We also have that:
V n ( t , s , u ( t , x ) ) = e t s n Σ k n | u k ( t , x ) | 10 n ( 1 + k 2 )
satisfies assumption ( D 6 ) with k ( t , s ) = e t s and M = 1 . Since the function h z = 0 w g i ( h , z ) is continuous on [ 0 , T ] , we can put G 1 = sup z = 0 w g 1 ( w , z ) : w [ 0 , T ] , where T is an arbitrary, fixed, psoitive real number. Thus, Theorem 35 implies that the infinite system (49) has at least one solution in c 0 .

6.2. Infinite System of Integral Equations in Two Variables of Hammerstein Type in Sequence Spaces C 0 and 1

In this subsection, we study the following infinite system of Hammerstein-type integral equations in two variables:
v n ( s , t ) = r n ( s , t ) + a b a b K n ( s , t , τ 1 , τ 2 ) f n ( τ 1 , τ 2 , v ( τ 1 , τ 2 ) ) d τ 1 d τ 2 ,
where ( s , t ) [ a , b ] × [ a , b ] in c 0 and 1 . The solvability of (50) is studied in [62] using the idea of measure of noncompactness (MNC).
To find the condition under which (50) has a solution in c 0 we need the following assumptions:
(A1
The functions ( f j ) j = 1 are real valued and continuous defined on the set I 2 × R . The operator Q defined on the space I 2 × c 0 as:
( s , t , v ) Q v ( s , t ) = f 1 ( s , t , v ) , f 2 ( s , t , v ) , f 3 ( s , t , v ) ,
maps I 2 × c 0 into c 0 . The set of all such functions Q v ( s , t ) ( s , t ) I 2 is equicontinuous at every point of c 0 , that is, given ϵ , δ > 0 :
u v c 0 δ implies ( Q u ) ( s , t ) ( Q v ) ( s , t ) c 0 ϵ .
(A2
For each fixed ( s , t ) I 2 , v ( s , t ) = v j ( s , t ) C ( I 2 , c 0 ) :
f n ( s , t , v ( s , t ) ) p n ( s , t ) + q n ( s , t ) sup j n | v j | n N ,
where p j ( s , t ) and q j ( s , t ) are real–valued continuous functions on I 2 . The function sequence q j ( s , t ) j N is equibounded on I 2 and the function sequence p j ( s , t ) j N converges uniformly on I 2 to a function vanishing identically on I 2 .
(A3
The functions K n : I 4 R are continuous on I 4 , ( n = 1 , 2 , ) , and K n ( s , t , x , y ) are equicontinuous with respect to ( s , t ) that is, for every ϵ > 0 there exists δ > 0 with:
| K n ( s 2 , t 2 , x , y ) K n ( s 1 , t 1 , x , y ) | ϵ , whenever | s 2 s 1 | δ , | t 2 t 1 | δ ,
for all ( x , y ) I 2 . Furthermore, the function sequence ( K n ( s , t , x , y ) ) is equibounded on the set I 4 and:
K = sup | K n ( s , t , x , y ) | : ( s , t ) , ( x , y ) I 2 , n = 1 , 2 , < .
(A4
The functions r n : I 2 R are continuous and the function sequence ( r n ) is uniformly convergent to zero on I 2 . Moreover,
R = sup | r n ( s , t ) | : ( s , t ) I 2 : n = 1 , 2 , < .
Keeping assumption ( A 2 ) under consideration, we define the following finite constants:
Q = sup q n ( s , t ) : ( s , t ) I 2 , n N ,
P = sup p n ( s , t ) : ( s , t ) I 2 , n N .
Theorem 36.
If the infinite system (50) satisfies ( A 1 ) ( A 4 ) , then it has at least one solution v ( s , t ) = v j ( s , t ) j N in c 0 for fixed ( s , t ) I 2 , whenever ( b a ) 2 K Q < 1 .
Example 15.
We study the infinite system of Hammerstein-type integral equations in two variables:
v n ( s , t ) = 1 n arctan ( s + t ) n + 1 2 1 2 sin s + t + τ 1 + τ 2 n ln 1 + 4 n 2 + ( τ 1 + τ 2 ) 2 4 + sup k n { | v k ( τ 1 , τ 2 ) | } 4 ( τ 1 + τ 2 ) 2 + n 2 d τ 1 d τ 2
for ( s , t ) [ 1 , 2 ] × [ 1 , 2 ] and n = 1 , 2 , .
Comparing (51) with (50) we have:
r n ( s , t ) = 1 n arctan ( s + t ) n , K n ( s , t , x , y ) = sin s + t + x + y n , f n ( τ 1 , τ 2 , v ( τ 1 , τ 2 ) ) = ln 1 + 4 n 2 + ( τ 1 + τ 2 ) 2 4 + sup k n { | v k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] = ln 1 + 1 + ( τ 1 + τ 2 ) 2 sup k n { | v k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] .
Denoting, by I 2 the interval [ 1 , 2 ] , we show that the assumptions of the Theorem 36 are satisfied. It is obvious that the operator F 1 defined by:
F 1 v ( s , t ) = f n ( s , t , v ( s , t ) ) ,
transforms the space I 2 2 × c 0 into c 0 .
Now, we establish that the family of functions F 1 v ( s , t ) ( s , t ) I 2 2 is equicontinuous at an arbitrary point v c 0 . Fix ϵ > 0 , n N , v c 0 and ( s , t ) I 2 2 , let u c 0 such that u v c 0 ϵ . Then,
f n ( s , t , v ) f n ( s , t , u ) = ln 1 + 1 + ( τ 1 + τ 2 ) 2 sup k n { | v k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] ln 1 + 1 + ( τ 1 + τ 2 ) 2 sup k n { | u k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] ( τ 1 + τ 2 ) 2 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] sup k n { | v k ( τ 1 , τ 2 ) | } sup k n { | u k ( τ 1 , τ 2 ) | } 1 16 sup k n | v k u k | .
Hence,
f n ( s , t , v ) f n ( s , t , u ) 1 16 v u c 0 ϵ 16 ,
so the family F 1 v ( s , t ) ( s , t ) I 2 2 is equicontinuous.
Now, fix ( s , t ) I 2 2 , v c 0 and n N , then:
f n ( s , t , v ) = ln 1 + 1 + ( τ 1 + τ 2 ) 2 sup k n { | v k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] 1 + ( τ 1 + τ 2 ) 2 sup k n { | v k ( τ 1 , τ 2 ) | } 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] = 1 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] + ( τ 1 + τ 2 ) 2 4 [ ( τ 1 + τ 2 ) 2 + n 2 ] sup k n { | v k ( τ 1 , τ 2 ) | }
We put p n ( s , t ) = 1 4 [ ( s + t ) 2 + n 2 ] and q n ( s , t ) = ( s + t ) 2 4 [ ( s + t ) 2 + n 2 ] . Then, clearly p n ( s , t ) and q n ( s , t ) are real–valued functions and p n ( s , t ) converges uniformly to zero.
Further, | q n ( s , t ) | 1 / 4 for all n = 1 , 2 , .
Hence, P = 1 / 4 and Q = sup s , t ) I 2 { q n ( s , t ) } = 1 / 4 .
The functions K n ( s , t , x , y ) are continuous on I 2 4 = [ 1 , 2 ] × [ 1 , 2 ] × [ 1 , 2 ] × [ 1 , 2 ] and the function sequence K n ( s , t , x , y ) is equibounded on I 2 4 . Moreover,
K = sup | K n ( s , t , x , y ) | : ( s , t ) , ( x , y ) I 2 2 , n N = 1 .
Now, fix ϵ > 0 , ( x , y ) I 2 2 and n N then for arbitrary ( s 1 , t 1 ) , ( s 2 , t 2 ) I 2 with:
| s 2 s 1 | ϵ 2 , | t 2 t 1 | ϵ 2 .
We have:
K n ( s 2 , t 2 , x , y ) K n ( s 1 , t 1 , x , y ) s 2 + t 1 + x + y n s 1 + t 1 + x + y n = 1 n ( s 2 s 1 ) + ( t 2 t 1 ) 1 n ( | s 2 s 1 | + | t 2 t 1 | ) ϵ .
Therefore, K n ( s , t , x , y ) is equicontinuous.
Thus, r n ( s , t ) , is continuous for all ( s , t ) I 2 2 and for all n and r n ( s , t ) converges uniformly to zero.
The value of the factor ( b a ) 2 K Q = 1 / 4 < 1 . Thus, by Theorem 36, the infinite system in (50) has a solution in c 0 , which belongs to the ball B R 0 c 0 where:
R 0 = R + ( b a ) 2 K Q 1 ( b a ) 2 K Q = arctan 4 + 1 4 1 1 4 = 4 3 arctan ( 4 ) .

Solution in the Space 1

The existence of a solution for the system (50) is found in the space 1 keeping the following assumptions under consideration:
(C1
The functions ( f j ) j = 1 are real valued and continuous defined on the set I 2 × R . The operator Q defined on the space I 2 × 1 as:
( s , t , v ) Q v ( s , t ) = f 1 ( s , t , v ) , f 2 ( s , t , v ) , f 3 ( s , t , v ) , ,
maps I 2 × 1 into 1 . The set of all such functions Q v ( s , t ) ( s , t ) I 2 is equicontinuous at every point of the space 1 , that is, given ϵ , δ > 0 ,
u v 1 δ implies ( Q u ) ( s , t ) ( Q v ) ( s , t ) 1 ϵ .
(C1
For fixed ( s , t ) I 2 , v ( s , t ) = v j ( s , t ) C ( I 2 , 1 ) , the following inequality holds:
f n ( s , t , v ( s , t ) ) a n ( s , t ) + d n ( s , t ) | v n | , n = 1 , 2 , 3 , ,
where a j ( s , t ) and d j ( s , t ) are real–valued continuous functions on I 2 . The function series n = 1 a n ( s , t ) is uniformly convergent on I 2 and the function sequence d j ( s , t ) j N is equibounded on I 2 . The function a ( s , t ) given by a ( s , t ) = n = 1 a n ( s , t ) is continuous on I 2 and the constants D , A defined as:
D = sup d n ( s , t ) : ( s , t ) I 2 , n N , A = max a ( s , t ) : ( s , t ) I 2 ,
are finite.
(C3
The functions K n : I 4 R are continuous on I 4 ( n = 1 , 2 , ) . Furthermore, these functions K n ( s , t , x , y ) are equicontinuous with respect to ( s , t ) , that is, for all ϵ > 0 there exists a δ > 0 such that:
| K n ( s 2 , t 2 , x , y ) K n ( s 1 , t 1 , x , y ) | ϵ whenever | s 2 s 1 | δ , | t 2 t 1 | δ ,
for all ( x , y ) I 2 . Moreover, the function sequence ( K n ( s , t , x , y ) ) is equibounded on the set I 4 and:
K = sup | K n ( s , t , x , y ) | : ( s , t ) , ( x , y ) I 2 , n = 1 , 2 , < .
(C4
The functions r n : I 2 R are continuous and the function sequence ( r n ) C ( I 2 , 1 ) .
Remark 17.
Since I 2 = [ a , b ] × [ a , b ] is a compact subset of R 2 , so the assumption of continuity in ( C 4 ) implies that r n : I 2 R is uniformly continuous, which implies that the function sequence r n ( s , t ) is equicontinuous on I 2 , as for every ϵ > 0 there is a δ > 0 , such that for all ( s 1 , t 1 ) , ( s 2 , t 2 ) I 2 ,
r n ( s 1 , t 1 ) r n ( s 2 , t 2 ) 1 n = 1 | r n ( s 2 , t 2 ) r n ( s 2 , t 2 ) | ϵ ,
whenever | ( s 1 , t 1 ) ( s 2 , t 2 ) | < δ . Furthermore, by (52), the function series n = 1 r n ( s , t ) is obviously convergent on I 2 and the function:
r ( s , t ) = n = 1 r n ( s , t ) ,
is continuous on I 2 . Furthermore,
R = max { r ( s , t ) : ( s , t ) I 2 } < .
Theorem 37.
If the system (50) satisfies ( C 1 ) ( C 4 ) , then it has at least one solution v ( s , t ) = v j ( s , t ) j N in 1 for fixed ( s , t ) I 2 , whenever ( b a ) 2 K D < 1 .
Example 16.
We study the infinite system of Hammerstein-type integral equations in two variables:
v n ( s , t ) = α n 2 ln [ ( s + t ) + n ] + 1 2 1 2 arctan ( s + t + τ 1 + τ 2 + n ) ( ( τ 1 + τ 2 ) 2 e n ( τ 1 + τ 2 ) + sin n ( τ 1 + τ 2 ) ( τ 1 + τ 2 ) 2 + n 3 · v n 2 ( τ 1 , τ 2 ) 1 + v 1 2 ( τ 1 , τ 2 ) + + v n 2 ( τ 1 , τ 2 ) ) d τ 1 d τ 2
for ( s , t ) [ 1 , 2 ] × [ 1 , 2 ] , α > 0 a constant.
Comparing the system with (50) we have:
r n ( s , t ) = α n 2 ln [ ( s + t ) + n ] , K n ( s , t , x , y ) = arctan ( s + t + x + y + n ) , f n ( s , t , v 1 , v 2 , ) = ( s + t ) 2 e n ( s + t ) + sin n ( s + t ) ( s + t ) 2 + n 3 · v n 2 ( s , t ) 1 + v 1 2 ( s , t ) + + v n 2 ( s , t ) .
for ( s , t ) , ( τ 1 , τ 2 ) [ 1 , 2 ] × [ 1 , 2 ] and n = 1 , 2 , .
Clearly, r n ( s , t ) is continuous on I 1 2 = [ 1 , 2 ] × [ 1 , 2 ] .
Moreover, for fixed ( s 1 , t 1 ) , ( s 2 , t 2 ) I 1 2 , we see that:
r n ( s 1 , t 1 ) r n ( s 2 , t 2 ) = n = 1 | r n ( s 1 , t 1 ) r n ( s 2 , t 2 ) | = α n = 1 1 n 2 ln [ ( s 1 + t 1 ) + n ] ln [ ( s 2 + t 2 ) + n ] = α n = 1 1 n 2 ln 1 + s 1 + t 1 s 2 t 2 s 2 + t 2 + n α n = 1 1 n 3 | s 1 + t 1 s 2 t 2 | α [ | s 1 s 2 | + | t 1 t 2 | ] ζ ( 3 ) ,
where ζ ( s ) denotes Riemann zeta function.
Choosing δ = ϵ / ( α ζ ( 3 ) ) , so that | s 1 s 2 | < δ 2 , | t 1 t 2 | < δ / 2 , we obtain:
r n ( s 1 , t 1 ) r n ( s 2 , t 2 ) < ϵ .
Furthermore, for every ( s , t ) I 1 2 we have:
r n ( s , t ) α n 2 ln ( 4 + n ) α n 2 4 + n α 2 n 2 + 1 n 3 / 2 .
Hence,
R = max n = 1 r n ( s , t ) : ( s , t ) I 1 2 = α ( 2 ζ ( 2 ) + ζ ( 1.5 ) ) < .
Thus, assumption ( C 4 ) and Remark 17 are satisfied.
Then, the function K n ( s , t , x , y ) is continuous in I 1 4 and:
K n ( s , t , x , y ) = | arctan ( s + t + x + y + n ) | π 2 .
Thus, the function sequence K n is equibounded on I 1 4 . Moreover, for fixed ( s 1 , t 1 ) , ( s 2 , t 2 ) I 1 2 and n N , we have for ( x , y ) I 1 2 :
| K n ( s 1 , t 1 , x , y ) K n ( s 2 , t 2 , x , y ) | = | arctan ( s 1 + t 1 + x + y + n ) arctan ( s 2 + t 2 + x + y + n ) | | s 1 s 2 | + | t 1 t 2 | .
Therefore, the function sequence K n ( s , t , x , y ) is equicontinuous with respect to ( s , t ) I 1 2 uniformly with respect to ( x , y ) I 1 2 , the value of the constant K given as:
K = sup { K n ( s , t , x , y ) : ( s , t ) , ( x , y ) I 1 2 , n N } = π 2 .
Hence, all assumptions of ( C 3 ) are satisfied.
Again,
| f n ( s , t , v ) | ( s + t ) 2 e n ( s + t ) + sin n ( s + t ) ( s + t ) 2 + n 3 · v n 2 1 + v 1 2 + + v n 2 ( s + t ) 2 e n ( s + t ) + 1 ( s + t ) 2 + n 3 · v n 2 1 + v 1 2 + + v n 2 ( s + t ) 2 e n ( s + t ) + 1 ( s + t ) 2 + n 3 · | v n | 1 + v n 2 ( | v n | ) ( s + t ) 2 e n ( s + t ) + 1 2 [ ( s + t ) 2 + n 3 ] | v n | .
Taking, a n ( s , t ) = ( s + t ) 2 e n ( s + t ) and d n ( s , t ) = 1 2 [ ( s + t ) 2 + n 3 ] gives:
| f n ( s , t , v ) | a n ( s , t ) + d n ( s , t ) | v n | .
Obviously, the functions a n ( s , t ) are continuous on I 1 2 , for any ( s , t ) I 1 2 we have | a n ( s , t ) |   ( 4 / n 3 ) · e 2 , and the function series a ( s , t ) = n = 1 a n ( s , t ) = ( s + t ) 2 e s + t 1 is uniformly convergent on the interval I 1 2 .
Furthermore,
| d n ( s , t ) | = 1 2 [ ( s + t ) 2 + n 3 ] 1 2 n 3 1 2 ,
for all n N . Hence, the function sequence ( h n ( s , t ) ) is equibounded on I 1 2 . The value of the constants A , D are:
A = max a ( s , t ) : ( s , t ) I 1 2 = 16 e 2 1 ; D = 1 2 ,
and ( b a ) 2 KD = π 8 . Using (54), (55), (56), and equation (11) of [62], we obtain:
R 1 = α 2 ζ ( 2 ) + ζ ( 3 ) + ( 2 1 ) 2 × 1 2 × 16 e 2 1 1 π 8 1.84 for α = 0.10 .
Finally, we check whether the assumption ( C 1 ) is satisfied. Fix v = ( v n ) B R 1 1 and ϵ > 0 , then for any u = ( u n ) B R 1 with u v 1 ϵ , then for fixed ( s , t ) I 1 2 , we have:
Q u ( s , t ) Q v ( s , t ) 1 = n = 1 f n ( s , t , u ) f n ( s , t , v ) n = 1 sin n ( s + t ) ( s + t ) 2 + n 3 u n 2 1 + u 1 2 + + u n 2 v n 2 1 + v 1 2 + + v n 2 n = 1 1 n 3 | u n 2 ( 1 + v 1 2 + + v n 2 ) v n 2 ( 1 + u 1 2 + + u n 2 ) | n = 1 1 n 3 [ | u n 2 v n 2 | + | u n 2 ( v 1 2 + + v n 2 ) u n 2 ( u 1 2 + + u n 2 ) | + | u n 2 ( u 1 2 + + u n 2 ) v n 2 ( u 1 2 + + u n 2 ) | ] n = 1 1 n 3 | u n 2 v n 2 | + u n 2 ( | v 1 2 u 1 2 | + + | v n 2 u n 2 | ) + | u n 2 v n 2 | ( u 1 2 + + u n 2 ) .
Since, v n , u n B R 1 , n N so | v n | R 1 , | u n | < R 1 so:
Q u ( s , t ) Q v ( s , t ) 1 n = 1 1 n 3 ( | u n v n | ( | u n | + | v n | ) ( 1 + u 1 2 + + u n 2 ) + u n 2 ( | v 1 u 1 | ( | v 1 | + | u 1 | ) + + | v n u n | ( | v n | + | u n | ) ) < 2 R 1 n = 1 1 n 3 | u n v n | ( 1 + n R 1 2 ) + R 1 2 i = 1 n | v i u i | = 2 R 1 u v 1 n = 1 1 n 3 ( 1 + n R 1 2 ) + R 1 2 = 2 R 1 u v 1 [ 1 + R 1 2 ] ζ ( 3 ) + R 1 2 ζ ( 2 ) .
Thus, choose:
δ = ϵ 2 R 1 [ 1 + R 1 2 ] ζ ( 3 ) + R 1 2 ζ ( 2 ) ,
then for u v 1 < δ we have:
Q u ( s , t ) Q v ( s , t ) 1 < ϵ .
Hence, the assumption ( C 1 ) is also satisfied, therefore by Theorem 37, we conclude that the system in (53) has a solution in B R 1 1 , where R 1 is given by (57).

6.3. Solvability of an Infinite System Of Integral Equations of Volterra–Hammerstein Type on the Real Half–Axis

Here, we consider one more recent application of a measure of noncompactness and Darbo’s fixed point theorem to the solvability of an infinite system of integral equations of Volterra–Hammerstein type:
x n ( t ) = a n ( t ) + f n ( t , x 1 , x 2 ) 0 t k n ( t , s ) g n ( s , x 1 , x 2 , ) d s
where t R + and n N , on the real half–axis ([63], Theorem 3.4). The paper [63] is in continuation of the papers [64,65].
In [63], the authors construct a measure of noncompactness on the space B C ( R + , ) of all functions x : R + that are continuous and bounded on R + . If x B C ( R + , ) , then x ( t ) = ( x n ( t ) ) for each t R + ; B C ( R + , ) is a Banach space with:
x = sup t R + x = sup t R + sup n | x n ( t ) | for all x B C ( R + , ) .
The following assumptions are made for the system (58):
(i)
The sequence ( a n ( t ) ) B C ( R + , ) satisfies lim t a n ( t ) = 0 uniformly in n, that is,
for all ε > 0 there exists T > 0 such that for all t T and all n N | a n ( t ) | ε ,
and also ( a n ( t ) ) c 0 for all t R + .
(ii)
The functions k n ( t , s ) = k n : R + 2 R are continuous on R + 2 for n = 1 , 2 , . Moreover the functions t k n ( t , s ) are equicontinuous on R + uniformly with respect to s R + , that is,
for all ε > 0 there exists δ > 0 such that for all n N , all s R + and all t 1 , t 2 R + | t 1 t 2 | δ implies | k n ( t 2 , s ) k n ( t 1 , s ) | ε .
(iii)
There exists a positive constant K 1 such that:
0 t | k n ( t , s ) | d s K 1
for any t R + and n = 1 , 2 , .
(iv)
The sequence ( k n ( t , s ) ) is equibounded on R + 2 , that is, there exists a positive constant K 2 such that | k n ( t , s ) | K 2 for all t , s R + and n = 1 , 2 , .
(v)
The functions f n are defined on the R + × R and take real values for n = 1 , 2 , . Moreover, the function t f n ( t , x 1 , x 2 , ) is uniformly continuous on R + with respect to x = ( x n ) and uniformly with respect to n N , that is, the following condition is satisfied:
for all ε > 0 there exists δ > 0 such that for all ( x i ) , all n N and all t , s R + | t s | δ implies | f n ( t , x 1 , x 2 , ) f n ( s , x 1 , x 2 , ) | ε .
(vi)
There exists a function l : R + R + such that l is nondecreasing on R + , continuous at 0 and there exists a sequence of functions ( f n ) in B C ( R + , ) , taking nonnegative values and such that lim t f n ( t ) = 0 uniformly with respect to n N (cf. assumption (i)) and lim n f n ( t ) = 0 for any t R + . Moreover, for any r > 0 the following inequality is satisfied:
| f n ( t , x 1 , x 2 , ) | f ¯ n ( t ) + l ( r ) sup { | x i | : i > n }
for each x = ( x i ) such that x r , for every t R + and for n = 1 , 2 , .
Let F ¯ = sup { f ¯ n ( t ) : n N , t R + .
(vii)
There exists a nondecreasing function m : R + R + which is continuous at 0 and satisfies:
| f n ( t , x 1 , x 2 , ) f n ( t , y 1 , y 2 , ) | m ( r ) x y
for any r > 0 , for x = ( x i ) , y = ( y i ) such that x , y r and for all t R + and n = 1 , 2 , .
(viii)
The functions g n are defined on the set R + × R and take real values for n = 1 , 2 , . Moreover, the operator g defined on R + × by:
( g x ) ( t ) = ( g n ( t , x ) ) = ( g 1 ( t , x ) , g 2 ( t , x ) , )
transforms the set R + × into and is such that the family of functions { ( g x ) ( t ) } t R + is equicontinuous on , that is, for all ε > 0 there exists δ > 0 such that:
( g y ) ( t ) ( g x ) ( t ) ε
for all t R + and all x , y such that x y δ .
(ix)
The operator g defined in assumption (viii) is bounded on the set R + × , that is, there exists a positive constant G such that ( g x ) ( t ) G for all x and all t R + .
(x)
There exists a positive solution r 0 of the inequality:
A + F ¯ G ¯ K 1 + G ¯ K 1 r l ( r ) r
such that G ¯ K 1 max { l ( r 0 ) , m ( r 0 ) } < 1 , where the constants F ¯ , G ¯ , K 1 were defined above and the constant A is defined by:
A = sup { | a n ( t ) | : t R + , n = 1 , 2 , } .
Theorem 38. 
([63], Theorem 3.4) Under the assumptions (i)–(x), the infinite system (58) has at least one solution x ( t ) = ( x n ( t ) ) in B C ( R + , ) .
Remark 18.
An example of the application of Theorem 38 can be found in ([63], Section 4).
We also recommend the paper [66].
Recently, in 2021 [67], a new sequence space related to the space p ( 1 p < ) was defined. The authors showed that it is a B K space with a Schauder basis. They established a formula for the Hausdorff measure of noncompactness for the bounded sets in the new sequence space. Then, Darbo’s fixed point theorem is applied to study the existence results for some infinite system of Langevin equations.

6.4. Periodic Mild Solutions for a Class of Functional Evolution Equations

In [68], the authors showed that the Poincaré operator is condensing with respect to the Kuratowski measure of noncompactness in a determined phase space. They also obtained periodic solutions from bounded solutions by applying Sadovskiĭ’s fixed point theorem.
Consider the existence of periodic mild solutions to the class of functional differential equations with infinite delay and non-instantaneous impulses:
u ( t ) + A ( t ) u ( t ) = f ( t , u ( t ) , u t ) if t I k , k = 0 , 1 , . . . , u ( t ) = g k ( t , u ( t k ) ) if t J k , k = 1 , 2 , . . . , u ( t ) = f ( t ) if t R : = ( , 0 ] ,
where I 0 = [ 0 , t 1 ] , I k = ( s k , t k + 1 ] , J k = ( t k , s k ] , 0 = s 0 < t 1 = s 1 = t 2 < < s m 1 = t m = s m = t m + 1 = T = s m + 1 = t m + 2 = < + , ( E , · E ) is a real Banach space, f : I k × E × B E , k = 0 , , g k : J k × E E , k = 1 , 2 , , are given functions T–periodic in t , T > 0 , B is an abstract phase space to be specified later, and ϕ : R E is a given function. Here, { A ( t ) } t > 0 is a T–periodic family of unbounded operators from E into E that generate an evolution system of operators { U ( t , s ) } ( t , s ) R + × R + for ( t , s ) Λ = { ( t , s ) R + × R + : 0 s t < + } , where R + : = [ 0 , + ) .
For any continuous function u and any t R + , we denote by u t the element of B defined by u t ( θ ) = u ( t + θ ) for θ R = ( , 0 ] . Here, u t ( · ) represents the history of the state up to the present time t. We assume that the histories u t belong to B .
By a periodic mild solution of problem (59), we mean a measurable and T–periodic function u that satisfies:
u ( t ) = U ( t , 0 ) ϕ ( 0 ) + 0 t U ( t , s ) f ( s , u ( s ) , u s ) d s if t I 0 U ( t , s k ) g k ( s k , u ( s k ) ) + s k t U ( t , s ) f ( s , u ( s ) , u s ) d s , if t I k , k = 1 , . . . , m g k ( t , u ( t k ) ) if t J k , k = 1 , . . . , m ϕ ( t ) if t R .
We use the following assumptions.
( H 1 ) The functions f and g k are continuous, and map bounded sets into bounded sets.
( H 2 ) The function t f ( t , u , v ) is measurable on I k for k = 0 , , m and for each u , v E × B . Furthermore, the functions u f ( t , u , v ) and v f ( t , u , v ) are continuous on E × B for a.e. t I k for k = 0 , , m .
( H 3 ) There is a positive constant T with f ( t + T , u , v ) = f ( t , u , v ) , A ( t + T ) = A ( t ) for t I k and k = 0 , , m , u , v E × B , and g k ( t + T , z ) = g k ( t , z ) for t J k , k = 1 , and m , z E .
( H 4 ) There exist continuous functions p : I k R + and q : J k R + with:
f ( t , u , v ) p ( t ) for a . e . t I k , k = 0 , . . . , m , and each u , v E × B ,
and,
g k ( t , z ) q ( t ) for a . e . t J k , and each z E , k = 0 , . . . , m .
( H 5 ) For bounded and measurable sets B ( t ) E and B t B for t R +
B ( t ) = { u ( t ) : u C ( I ) } and B t = { u t : u t B } ,
implies,
α ( f ( t , B ( t ) , B t ) ) p ( t ) α ( B ) for a . e . t I k , ( k = 0 , , m ) ,
and,
α ( g k ( t , B ) ) q ( t ) α ( B ) for a . e . t J k , ( k = 1 , , m ) ,
where α is Kuratowski’s measure of noncompactness on the Banach space E.
Further, set:
Δ = { ( t , s ) J × J : 0 s t T } ,
M = s u p ( t , s ) Δ U ( t , s ) B ( E ) , p * = sup t I k p ( t ) and q * = sup t J k q ( t ) .
We shall state the main result of the paper [69].
Theorem 39. 
([69], Theorem 3.2) If ( H 1 ) ( H 5 ) are satisfied and 4 M T p * < 1 , then Problem (59) has at least one T–periodic mild solution on R .
The authors also present an example to illustrate Theorem 39.
We also mention that fixed point theorems in b–metric spaces were recently considered.
Remark 19.
Recently, in 2021 [69], the authors introduced and studied two generalized contractions, the generalized F t s –contraction and the generalized ( ψ , ϕ , F t s ) –contraction. Two fixed point theorems were established in ordered b–metric spaces. An example is presented to illustrate the fixed point theorem of the generalized F t s –contraction.
It would be interesting to prove related results in the framework of measures of noncompactnes.

7. Some Mathematical Background

Here, we present some recent results connected to the existence of best proximity points (pairs) for some classes of cyclic and noncyclic condensing operators in Banach spaces with respect to a suitable measure of noncompactness. We also discuss the existence of an optimal solution for systems of integro–differentials.
Recently, many studies [70,71,72,73,74] applied generalizations of Darbo–Sadovskii’s fixed point theorem, Theorem 11, concerning the existence of solutions for several classes of functional integral equations.
In the following survey, we present some recent existence results of best proximity points (pairs) as a generalization of fixed points and obtain other extensions of Schauder’s fixed point problem as well as Darbo–Sadovskii’s fixed point theorem. As applications of our conclusions, we study the existence of optimal solutions for various classes of differential equations.
We recall that a Banach space X is said to be strictly convex provided that the following implication holds for x , y , p X and R > 0 :
x p R , y p R , x y implies x + y 2 p < R .
It is well known that Hilbert spaces and p spaces ( 1 < p < ) are strictly convex Banach spaces. Furthermore, the Banach space 1 with the norm:
x = x 1 + x 2 , for all x 1 ,
where, . 1 and . 2 are the norms on 1 and 2 , respectively, is strictly convex.
Suppose A is a nonempty subset of a normed linear space X and T maps A into X. It is clear that the necessary (but not sufficient) condition for the existence of a fixed point of T is that the intersection of A and T ( A ) is nonempty. If T does not have any fixed point, then the distance between x and T x is positive for any x in A. In this case, it is our purpose to find an element x in A so that the distance of x and T x is minimum. Such a point is called a best approximant point of T in A. The first best approximation theorem due to Ky Fan ([75]) states that if A is a compact and convex subset of a normed linear space X and T : A X is a continuous map from A, then T has a best approximant point in A. An interesting extension of Ky Fan’s theorem can be considered when T : A B , where subset B X . In this case, it is interesting to study the existence of the best proximity points; that is, points in A that estimate the distance between A and B. The existence of best proximity points for various classes of non-self mappings is a subject in optimization theory, which recently attracted the attention of many authors (see [76,77,78,79], and the references therein).
Let A , B be subsets of a normed linear space X. We say that a pair ( A , B ) of subsets of a Banach space X satisfies a certain property if both A and B satisfy that property. For example, ( A , B ) is convex if and only if both A and B are convex; ( A , B ) ( C , D ) A C , B D . From now on, B ( x ; r ) will denote the closed ball in the Banach space X centered at x X with radius r > 0 . The closed and convex hull of a set A will be denoted by con ¯ ( A ) . Furthermore, diam ( A ) stands for the diameter of the set A. Moreover, for the pair ( A , B ) we define:
A 0 = { x A : y B | x y = dist ( A , B ) } ,
B 0 = { y B : x A | x y = dist ( A , B ) } .
It is known that if ( A , B ) is a nonempty, weakly compact, and convex pair in a Banach space X, then the pair ( A 0 , B 0 ) is also nonempty, weakly compact, and convex.
Definition 18.
A nonempty pair ( A , B ) in a normed linear space X is said to be proximinal if A = A 0 and B = B 0 .
A map T : A B A B is cyclic relatively nonexpansive if T is cyclic, that is, T ( A ) B , T ( B ) A and T x T y x y , whenever x A and y B . In particular, if A = B , then T is called a nonexpansive self–map. A point x 🟉 A B is a best proximity point for the map T if:
x 🟉 T x 🟉 = dist ( A , B ) : = inf { x y : x A , y B } .
In fact, best proximity point theorems have been studied to find necessary conditions such that the minimization problem:
min x A B x T x ,
has at least one solution.
A map T : A B A B is noncyclic relatively nonexpansive if T is noncyclic, that is, T ( A ) A , T ( B ) B and T x T y x y for any ( x , y ) A × B . Clearly, the class of noncyclic relatively nonexpansive maps contains the class of nonexpansive maps. Noncyclic relatively nonexpansive maps may not necessarily be continuous. A point ( p , q ) A × B is a best proximity pair if it is a solution of the following minimization problem:
min x A x T x , min y B y T y , and min ( x , y ) A × B x y .
Clearly, ( p , q ) A × B is a solution of the problem (61) if and only if:
p = T p , q = T q , and p q = dist ( A , B ) .
In 2017, M. Gabeleh, proved the following existence theorems by using a concept of proximal diametral sequences (we also refer to [80] for the same results which were based on a geometric notion of proximal normal structure).
Theorem 40
([81]). Let ( A , B ) be a nonempty, compact, and convex pair in a Banach space X. If T is cyclic relatively nonexpansive mapping, then T has a best proximity point.
Theorem 41
([81]). Let ( A , B ) be a nonempty, compact, and convex pair in a strictly convex Banach space X. If T is noncyclic relatively nonexpansive mapping, then T has a best proximity pair.
Finally, we state Mazur’s lemma.
Lemma 2
([82]). Let A be a nonempty and compact subset of a Banach space X. Then con ¯ ( A ) is compact.

8. Cyclic (Noncyclic) Condensing Operators

We start with an extension of Theorem 40.
Definition 19.
Let ( A , B ) be a bounded pair in a Banach space X and T : A B A B a cyclic (noncyclic) map. Then, T is called compact whenever both T | A and T | B are compact, that is, the pair ( T ( A ) ¯ , T ( B ) ¯ ) is compact.
The next result generalizes Schauder’s fixed point theorem, Theorem 10.
Theorem 42. 
([83], Theorem 3.2) Let ( A , B ) be a bounded, closed, and convex pair in a Banach space X such that A 0 . Also, let T : A B A B be a cyclic relatively nonexpansive map. If T is compact, then T has a best proximity point.
Proof. 
Put K 1 = con ¯ ( T ( B ) ) and K 2 = con ¯ ( T ( A ) ) . Let x A 0 . Then there exists y B with x y = dist ( A , B ) . Since T is a cyclic relatively nonexpansive map,
dist ( K 1 , K 2 ) T y T x x y = dist ( A , B ) .
Thus, dist ( K 1 , K 2 ) = dist ( A , B ) . It follows from Mazur’s lemma that the pair ( K 1 , K 2 ) is compact and clearly is convex. Since T ( A ) B , we get con ¯ ( T ( A ) ) B . Hence,
T ( K 2 ) = T ( con ¯ ( T ( A ) ) ) T ( B ) con ¯ ( T ( B ) ) = K 1 .
Analogously, T ( K 1 ) K 2 , and so T is cyclic on K 1 K 2 . It follows from Theorem 40 that there exists a point x * K 1 K 2 with x * T x * = dist ( K 1 , K 2 ) ( = dist ( A , B ) ) , and the result follows. □
Theorem 43
([83], Theorem 4.1). Let ( A , B ) be a bounded, closed, and convex pair in a strictly convex Banach space X such that A 0 . Furthermore, let T : A B A B be a noncyclic relatively nonexpansive map. If T is compact, then T has a best proximity pair.
Proof. 
We assume K 1 = con ¯ ( T ( A ) ) and K 2 = con ¯ ( T ( B ) ) . Then dist ( K 1 , K 2 ) = dist ( A , B ) . Moreover, con ¯ ( T ( A ) ) A , so:
T ( K 1 ) = T ( con ¯ ( T ( A ) ) ) T ( A ) con ¯ ( T ( A ) ) = K 1 .
Analogously, T ( K 2 ) K 2 . Therefore, T is noncyclic on K 1 K 2 . On the other hand, from Lemma 2 ( K 1 , K 2 ) is compact and convex in a strictly convex Banach space X. By Theorem 41 that there exists ( p , q ) K 1 × K 2 with:
p = T p , q = T q , and p q = dist ( K 1 , K 2 ) ( = dist ( A , B ) ) ,
that is, ( p , q ) is a best proximity pair for the map T. □
Notation. Let ( A , B ) be a pair in a normed linear space X and T : A B A B be a cyclic (noncyclic) map. The set of all nonempty, bounded, closed, convex, proximinal, and T–invariant pairs ( C , D ) ( A , B ) with dist ( C , D ) = dist ( A , B ) is denoted by M T ( A , B ) . Notice that M T ( A , B ) may be empty, but in particular if ( A , B ) is a weakly compact and convex pair in a Banach space X and T is cyclic (noncyclic) relatively nonexpansive, then ( A 0 , B 0 ) M T ( A , B ) (see [84,85] for more details).
Definition 20 
(Gabeleh-Markin, (2018) [83]). Let ( A , B ) be a convex pair in a Banach space X and μ an MNC on X. A map T : A B A B is said to be a cyclic (noncylic) condensing operator if there exists r ( 0 , 1 ) such that for any ( K 1 , K 2 ) M T ( A , B ) ,
μ ( T ( K 1 ) T ( K 2 ) ) r μ ( K 1 K 2 ) .
Definition 21 
(Gabeleh-Vetro, (2019) [86]). Let ( A , B ) be a convex pair in a Banach space X and μ be an MNC on X. A map T : A B A B is said to be a cyclic (noncyclic) generalized condensing operator provided that T is cyclic (noncyclic) map and for any ( C , D ) M T ( A , B ) there exist ψ Ψ and l N such that:
μ ( C 2 l D 2 l ) ψ ( μ ( C D ) ) .
Notation. Let Φ denote the set of all functions φ : [ 0 , ) [ 0 , 1 ) such that:
φ ( t n ) 1 t n 0 .
Definition 22 
(Gabeleh-Moshokoa-Vetro, (2019) [87]). Let ( A , B ) be a convex pair in a Banach space X and μ be an MNC on X. A map T : A B A B is said to be a noncyclic (cyclic) φ-condensing operator for some φ Φ provided that for any ( K 1 , K 2 ) M T ( A , B ) we have:
μ ( T ( K 1 ) T ( K 2 ) ) φ ( μ ( K 1 K 2 ) ) μ ( K 1 K 2 ) .
Example 17.
Let ( A , B ) be a convex pair in a Banach space X such that B is compact and α is the Kuratowski measure of noncompactness on X. Assume that T : A B A B is a cyclic maps so that T | A is contraction with the contraction constant r ] 0 , 1 [ . Then T is a cyclic condensing operator.
Proof. 
Suppose ( H 1 , H 2 ) ( A , B ) is a nonempty, bounded, closed, convex, and proximinal pair, which is T-invariant and dist ( H 1 , H 2 ) = dist ( A , B ) . Since B is compact, α ( T ( H 2 ) ) = 0 and so,
α ( T ( H 1 ) T ( H 2 ) ) = max { α ( T ( H 1 ) ) , α ( T ( H 2 ) ) }
= α ( T ( H 1 ) ) r α ( H 1 ) r α ( H 1 H 2 ) ,
and the result follows. □
We recall that ( A , B ) in a metric space ( X , d ) is be proximal compactness ([88]) provided that every net { ( x α , y α ) } of A × B satisfying the condition that d ( x α , y α ) dist ( A , B ) , has a convergent subnet in A × B .
Example 18.
Let ( A , B ) be a convex and a proximal compactness pair in a Banach space X and μ be a measure of noncompactness on X. Then, every cyclic relatively nonexpansive map T : A B A B is a condensing operator.
Proof. 
Suppose ( H 1 , H 2 ) ( A , B ) is a bounded, closed, convex, and proximinal pair, which is T–invariant and dist ( H 1 , H 2 ) = dist ( A , B ) . We prove that ( T ( H 1 ) , T ( H 2 ) ) is a relatively compact pair. Let { x n } be a sequence in H 1 . Since the ( H 1 , H 2 ) is proximinal, there exists a sequence { y n } in H 2 such that x n y n = dist ( A , B ) for all n 1 . Then,
T x n T y n x n y n = dist ( A , B ) , n 1 .
Since ( A , B ) is a proximal compactness pair, the sequence { ( T x n , T y n ) } has a convergent subsequence which implies that ( T ( H 1 ) , T ( H 2 ) ) is relatively compact. Therefore, μ ( T ( H 1 ) T ( H 2 ) ) = 0 , which concludes that T is a condensing operator for any r [ 0 , 1 [ . □

9. Existence Results

In this section, we present some existence theorems of best proximity points for the aforesaid classes of condensing operators, which are new extensions of Darbo’s fixed point problem.
Theorem 44
([83]). Let ( A , B ) be a bounded, closed, and convex pair in a Banach space X such that A 0 and μ is an MNC on X. Suppose T : A B A B is a cyclic relatively nonexpansive map, which is condensing in the sense of Definition 21. Then, T has a best proximity point.
Proof. 
Note that ( A 0 , B 0 ) is a closed, convex, and proximinal pair, which is T-invariant because of the fact that T is a cyclic relatively nonexpansive map. Let ( x 0 , y 0 ) A 0 × B 0 be such that x 0 y 0 = dist ( A , B ) and suppose C is a family of all nonempty, closed, convex, proximinal, and T-invariant pairs ( E , F ) ( A , B ) such that ( x 0 , y 0 ) E × F . Then, ( A 0 , B 0 ) C . Put:
( K 1 , K 2 ) = ( E , F ) C ( E , F ) C ,
and define N = con ¯ ( T ( K 1 ) { y 0 } ) and M = con ¯ ( T ( K 2 ) { x 0 } ) . Thus ( x 0 , y 0 ) M × N and ( M , N ) ( K 1 , K 2 ) . Moreover,
T ( M ) T ( K 1 ) N , T ( N ) T ( K 2 ) M ,
that is, T is cyclic on M N . Besides, if x M , then x = j = 1 n 1 c j T ( y j ) + c n x 0 , where y j K 2 for all j { 1 , 2 , , n 1 } for which c j 0 , j = 1 n c j = 1 . Since ( K 1 , K 2 ) is proximinal, there exists x j K 1 so that x j y j = dist ( A , B ) for all j { 1 , 2 , , n 1 } . Now, if y = j = 1 n 1 c j T ( x j ) + c n y 0 , then y N and we have:
x y = ( j = 1 n 1 c j T ( y j ) + c n x 0 ) ( j = 1 n 1 c j T ( x j ) + c n y 0 ) j = 1 n 1 c j T ( y j ) T ( x j ) + c n x 0 y 0 [ j = 1 n 1 c j dist ( A , B ) ] + c n dist ( A , B ) = dist ( A , B ) .
Therefore, M 0 = M . Similarly, N 0 = N and so ( M , N ) is proximinal. Hence, ( M , N ) C . It follows from the definition of ( K 1 , K 2 ) that M = K 1 and N = K 2 . On the other hand, since T is a condensing operator, we have:
μ ( M N ) = max { μ ( M ) , μ ( N ) } = max { μ ( con ¯ ( T ( K 2 ) { x 0 } ) ) , μ ( con ¯ ( T ( K 1 ) { y 0 } ) ) } = max { μ ( T ( K 2 ) ) , μ ( T ( K 1 ) ) } = max { μ ( T ( N ) ) , μ ( T ( M ) ) } = μ ( T ( N ) T ( M ) ) r μ ( M N ) .
This implies that max { μ ( M ) , μ ( N ) } = μ ( M N ) = 0 . Thereby, ( M , N ) is a compact and convex pair with dist ( M , N ) = dist ( A , B ) such that T : M N M N is a cyclic relatively nonexpansive map. Now from Theorem 42, we conclude that T has a best proximity point. □
In the case that T is noncyclic in the above theorem, we need the strict convexity of the Banach space X.
Theorem 45
([83]). Let ( A , B ) be a bounded, closed, and convex pair in a strictly convex Banach space X such that A 0 and μ is an MNC on X. If T : A B A B is a noncyclic relatively nonexpansive map, which is condensing in the sense of Definition 21, then T has a best proximity pair.
Proof. 
We note that ( A 0 , B 0 ) is closed, convex, and proximinal, which is T-invariant. Let ( u , v ) A 0 × B 0 be such that u v = dist ( A , B ) and suppose F is a family of all nonempty, closed, convex, proximinal, and T-invariant pairs ( E , F ) ( A , B ) such that ( u , v ) E × F . Then, ( A 0 , B 0 ) F . Put,
( K 1 , K 2 ) = ( E , F ) C ( E , F ) F ,
and set H 1 = con ¯ ( T ( K 1 ) { u } ) and H 2 = con ¯ ( T ( K 2 ) { v } ) . Thus, ( u , v ) H 1 × H 2 and ( H 1 , H 2 ) ( K 1 , K 2 ) . Further,
T ( H 1 ) T ( K 1 ) H 1 , T ( H 2 ) T ( K 2 ) H 2 .
Therefore, T is noncyclic on H 1 H 2 . Moreover, if x H 1 , then x = j = 1 n 1 c j T ( u j ) + c n u , where u j K 1 for all j { 1 , 2 , , n 1 } for which c j 0 , j = 1 n c j = 1 . In view of the fact that ( K 1 , K 2 ) is proximinal, there exists v j K 2 so that u j v j = dist ( A , B ) for all j { 1 , 2 , , n 1 } . Now, if we define y = j = 1 n 1 c j T ( v j ) + c n v , then y H 2 and:
x y = ( j = 1 n 1 c j T ( u j ) + c n u ) ( j = 1 n 1 c j T ( v j ) + c n v ) j = 1 n 1 c j T ( u j ) T ( v j ) + c n u v dist ( A , B ) .
Hence, ( H 1 ) 0 = H 1 . By a similar argument, we can see that ( H 2 ) 0 = H 2 , that is, ( H 1 , H 2 ) is a proximinal pair. This concludes that ( H 1 , H 2 ) F and by the definition of ( K 1 , K 2 ) we must have H 1 = K 1 and H 2 = K 2 . Now, since T is a condensing operator,
μ ( H 1 H 2 ) = max { μ ( H 1 ) , μ ( H 2 ) } = max { μ ( con ¯ ( T ( K 1 ) { u } ) ) , μ ( con ¯ ( T ( K 2 ) { v } ) ) } = max { μ ( T ( K 1 ) ) , μ ( T ( K 2 ) ) } = max { μ ( T ( H 1 ) ) , μ ( T ( H 2 ) ) } r μ ( H 1 H 2 ) .
Thereby, max { μ ( H 1 ) , μ ( H 2 ) } = 0 , and so, ( H 1 , H 2 ) is a compact and convex pair with dist ( H 1 , H 2 ) = dist ( A , B ) and that T : H 1 H 2 H 1 H 2 is a noncyclic relatively nonexpansive map. Now the result follows from Theorem 43. □
We now present some extensions of Theorem 44 and Theorem 45.
Theorem 46
([86]). Let ( A , B ) be a bounded, closed, and convex pair in a Banach space X such that A 0 and μ be an MNC on X. Let T : A B A B be a cyclic relatively nonexpansive map which is Meir–Keeler condensing. Then, T has a best proximity point.
Proof. 
Put G 0 : = A 0 and H 0 : = B 0 , and for all n N define:
G n = con ¯ ( T ( G n 1 ) ) , H n = con ¯ ( T ( H n 1 ) ) .
We now haveL
G 1 = con ¯ ( T ( G 0 ) ) = con ¯ ( T ( A 0 ) ) B 0 = H 0 .
Thus, T ( G 1 ) T ( H 0 ) and so G 2 = con ¯ ( T ( G 1 ) ) con ¯ ( T ( H 0 ) ) = H 1 . Continuing this process, and by induction, we conclude that G n + 1 H n . Similarly, we can see that H n G n 1 for all n N . This implies that:
G n + 2 H n + 1 G n H n 1 , n N .
Hence, { ( G 2 n , H 2 n ) } n 0 is a decreasing sequence of nonempty, closed, and convex pairs in A 0 × B 0 . Moreover,
T ( H 2 n ) T ( G 2 n 1 ) con ¯ ( T ( G 2 n 1 ) ) = G 2 n ,
T ( G 2 n ) T ( H 2 n 1 ) con ¯ ( T ( H 2 n 1 ) ) = H 2 n .
Thereby, for all n N the pair ( G 2 n , H 2 n ) is T-invariant. On the other hand, if ( p , q ) A 0 × B 0 is a proximal pair, then:
dist ( G 2 n , H 2 n ) T 2 n p T 2 n q p q = dist ( A , B ) , n N .
We shall show by induction that the pair ( G n , H n ) is proximinal for all n N { 0 } . It is obvious if n = 0 . Suppose that ( G n , H n ) is proximinal. Let x G n + 1 = con ¯ ( T ( G n ) ) be an arbitrary element. Then x = j = 1 k λ j T ( x j ) with x j G n where k N , λ j 0 and j = 1 k λ j = 1 . The proximinality of the pair ( G n , H n ) implies that for all 1 j k there exists y j H n such that x j y j = dist ( G n , H n ) ( = dist ( A , B ) ) . Put y = j = 1 k λ j T ( y j ) . Then y con ¯ ( T ( H n ) ) = H n + 1 and:
x y = j = 1 k λ j T ( x j ) j = 1 k λ j T ( y j ) j = 1 k λ j x j y j = dist ( A , B ) ,
and so the pair ( G n + 1 , H n + 1 ) is proximinal. We now consider the following possible cases.
If max { μ ( G 2 k ) , μ ( H 2 k ) } = 0 for some k N , then:
T : G 2 k H 2 k G 2 k H 2 k
is a compact and cyclic relatively nonexpansive map. Now, from Theorem 42, the result follows.
Assume that max { μ ( G 2 n ) , μ ( H 2 n ) } > 0 for all n N . Put ε n : = μ ( G 2 n H 2 n ) . Since T is a cyclic Meir–Keeler condensing operator, there exists δ n : = δ ( ε n ) such that:
μ ( T ( G 2 n ) T ( H 2 n ) ) < ε n , n N .
Further, for all n N we have:
ε n + 1 = μ ( G 2 n + 2 H 2 n + 2 ) = max { μ ( G 2 n + 2 ) , μ ( H 2 n + 2 ) } max { μ ( G 2 n ) , μ ( H 2 n ) } = ε n .
Thus, { ε n } is a decreasing sequence of positive real numbers. Assume that lim n ε n = r . We claim that r = 0 . Suppose the contrary. Then there exists l N such that r ε l < r + δ ( r ) . Again, using the fact that T is a cyclic Meir–Keeler condensing operator, we conclude that:
ε l + 1 = μ ( G 2 l + 2 H 2 l + 2 ) = max { μ ( G 2 l + 2 ) , μ ( H 2 l + 2 ) } max { μ ( H 2 l + 1 ) , μ ( G 2 l + 1 ) } = max { μ ( con ¯ ( T ( H 2 l ) ) ) , μ ( con ¯ ( T ( G 2 l ) ) ) } = max { μ ( T ( H 2 l ) ) , μ ( T ( G 2 l ) ) } = μ ( T ( G 2 l ) T ( H 2 l ) ) < r ,
which is a contradiction. Therefore,
lim n μ ( G 2 n H 2 n ) = max { lim n μ ( G 2 n ) , lim n μ ( H 2 n ) } = 0 .
Set:
G = n = 0 G 2 n and H = n = 0 H 2 n .
Then the pair ( C , D ) is compact. It is also convex and T–invariant with dist ( A , B ) = dist ( G , H ) . This ensures that T has a best proximity point. □
Theorem 47
([86]). Let ( A , B ) be a weakly compact and convex pair in a Banach space X and μ be an MNC on X. Let T : A B A B be a cyclic relatively nonexpansive map which is a generalized condensing operator in the sense of Definition 20. Then T has a best proximity point.
Proof. 
Note that ( A 0 , B 0 ) M T ( A , B ) . Put:
A n : = con ¯ ( T ( A n 1 ) ) , B n : = con ¯ ( T ( B n 1 ) ) .
By induction, we show that T is cyclic on A n B n for all n N . Since A 1 = con ¯ ( T ( A 0 ) ) B 0 ,
T ( A 1 ) T ( B 0 ) con ¯ ( T ( B 0 ) ) = B 1 .
Similarly, we can see that T ( B 1 ) A 1 , that is, T is cyclic on A 1 B 1 . Now, suppose T is cyclic on A k B k for some k N . Then T ( A k ) B k and so A k + 1 = con ¯ ( T ( A k ) ) B k which implies that:
T ( A k + 1 ) T ( B k ) con ¯ ( T ( B k ) ) = B k + 1 .
Equivalently, we can see that T ( B k + 1 ) A k + 1 , which ensures that T is cyclic on A k + 1 B k + 1 . Besides,
A n + 1 = con ¯ ( T ( A n ) ) B n = con ¯ ( T ( B n 1 ) ) A n 1 , n N ,
which concludes that the sequence { A 2 n } n N { 0 } is decreasing and, similarly, we can see that the sequence { B 2 n } n N { 0 } is also decreasing. Now, let ( x 0 , y 0 ) A 0 × B 0 be such that x 0 y 0 = dist ( A , B ) . Since T is cyclic relatively nonexpansive, ( T n x 0 , T n y 0 ) A n × B n and:
dist ( A n , B n ) T n x 0 T n y 0 x 0 y 0 = dist ( A , B ) , n N .
Thus, dist ( A n , B n ) = dist ( A , B ) for all n N { 0 } . Again, by mathematical induction, we assert that any pair ( A n , B n ) is proximinal. We note that the pair ( A 0 , B 0 ) is proximinal. Let ( A k , B k ) be a proximinal pair. We consider the following observations:
The pair ( con ( T ( A k ) ) , con ( T ( B k ) ) ) is proximinal.
Proof. 
Let x con ( T ( A k ) ) be an arbitrary element. Then, x = j = 1 m λ j T ( a j ) for some m N , where a j A k for all 1 j m . Since ( A k , B k ) is proximinal, for all 1 j m there exists an element b j B k for which a j b j = dist ( A k , B k ) ( = dist ( A , B ) ) . Put y : = j = 1 m λ j T ( b j ) . Clearly, y con ( T ( B k ) ) and we have:
x y = j = 1 m λ j T ( a j ) j = 1 m λ j T ( b j ) j = 1 m λ j T ( a j ) T ( b j ) j = 1 m λ j a j b j = dist ( A , B ) ,
and the result follows. □
The pair ( con ¯ ( T ( A k ) ) , con ¯ ( T ( B k ) ) ) is proximinal.
Proof. 
Let u con ¯ ( T ( A k ) ) . Then there is a sequence { w n } in con ( T ( A k ) ) such that w n u . Since ( con ( T ( A k ) ) , con ( T ( B k ) ) ) is proximinal, for any n N there exists a point z n con ( T ( B k ) ) such that:
w n z n = dist ( con ( T ( A k ) ) , con ( T ( B k ) ) ) = dist ( A , B ) .
By the fact that con ¯ ( T ( B k ) ) is weakly compact, there exists a subsequence { z n j } of the sequence { z n } , which converges weakly to a point v con ¯ ( T ( B k ) ) . It now follows from the weakly lower semi-continuity of the norm that:
u v lim inf j w n j z n j = dist ( A , B ) .
So, the pair ( A k + 1 , B k + 1 ) is proximinal. □
Therefore, { ( A 2 n , B 2 n ) } n N { 0 } is a descending sequence in M T ( A , B ) . Set:
r : = lim n μ ( A 2 n B 2 n ) .
Since T is a cyclic generalized condensing operator, there exist ψ Ψ and l 1 N such that μ ( A 2 l 1 B 2 l 1 ) ψ ( μ ( A B ) ) . Note that ( A 2 l 1 , B 2 l 1 ) is a weakly compact, convex, and proximinal pair and that T : A 2 l 1 B 2 l 1 A 2 l 1 B 2 l 1 is cyclic. From the above arguments, we can find a positive integer l 2 such that:
μ ( A 2 ( l 1 + l 2 ) B 2 ( l 1 + l 2 ) ) ψ ( μ ( A 2 l 1 B 2 l 1 ) ) ψ 2 ( μ ( A B ) ) .
Continuing this process, there exists l i N such that:
μ ( A 2 ( j = 1 i l j ) B 2 ( j = 1 i l j ) ) ψ i ( μ ( A B ) ) , i N .
In view of the fact that ψ i ( μ ( A B ) ) 0 , we must have r = 0 . Now, if we set:
( A , B ) = ( j = 1 A 2 j , j = 1 B 2 j ) ,
then ( A , B ) is a nonempty, convex, compact, and T-invariant pair with dist ( A , B ) = dist ( A , B ) . Hence, from Theorem 42, we obtain the existence of a best proximity point for the map T, and this completes the proof.
The noncyclic version of Theorem 47 can be reformulated as below.
Theorem 48.
Let ( A , B ) be a weakly compact and convex pair in a strictly convex Banach space X and μ be an MNC on X. Let T : A B A B be a noncyclic relatively nonexpansive map, which is a generalized condensing operator in the sense of Definition 21. Then, T has a best proximity pair.
Proof. 
As in the proof of Theorem 47, let A n = con ¯ ( T ( A n 1 ) ) and B n = con ¯ ( T ( B n 1 ) ) for all n N . Since T is noncyclic, A 1 = con ¯ ( T ( A 0 ) ) A 0 , and so:
T ( A 1 ) T ( A 0 ) con ¯ ( T ( A 0 ) ) = A 1 .
Similarly, T ( B 1 ) B 1 , that is, T is noncyclic on A 1 B 1 . Continuing this process, and by induction, we can see that T is noncyclic on A n B n for all n N . For all n N we have:
A n + 1 = con ¯ ( T ( A n ) ) A n , B n + 1 = con ¯ ( T ( B n ) ) B n .
Moreover, by an equivalent discussion of Theorem 42, we conclude that ( A n , B n ) is a proximinal pair with dist ( A n , B n ) = dist ( A , B ) for all n N . Hence, { ( A n , B n ) } is a descending sequence of nonempty, weakly compact, convex, T-invariant, and proximinal pairs and so its even subsequence, that is, { ( A 2 n , B 2 n ) } is a member of M T ( A , B ) . By a similar manner of Theorem 42 if we define:
( A , B ) = ( j = 1 A 2 j , j = 1 B 2 j ) ,
then ( A , B ) is a nonempty, compact, convex, and T-invariant pair in a strictly convex Banach space X and so Theorem 43 guarantees the existence of a best proximity pair for the map T. □
Theorem 49
([87]). Let ( A , B ) be a bounded, closed, and convex pair in a Banach space X such that A 0 is nonempty and μ is an MNC on X. Let T : A B A B be a cyclic relatively nonexpansive map, which is φ–condensing in the sense of Definition 22 for some φ Φ . Then, T has a best proximity point.
Proof. 
For all n N define:
C n = con ¯ ( T ( C n 1 ) ) , D n = con ¯ ( T ( D n 1 ) ) ,
where, C 0 : = A 0 and D 0 : = B 0 . Then we have:
C 1 = con ¯ ( T ( C 0 ) ) = con ¯ ( T ( A 0 ) ) B 0 = D 0 ,
and so, T ( C 1 ) T ( D 0 ) which implies that:
C 2 = con ¯ ( T ( C 1 ) ) con ¯ ( T ( D 0 ) ) = D 1 .
Continuing this process, we obtain: C n + 1 D n . We also have:
D 1 = con ¯ ( T ( D 0 ) ) = con ¯ ( T ( B 0 ) ) A 0 = C 0 ,
and hence, T ( D 1 ) T ( C 0 ) . Thus,
D 2 = con ¯ ( T ( D 1 ) ) con ¯ ( T ( C 0 ) ) = C 1 .
Then by induction we conclude that: D n C n 1 for all n N . Therefore,
C n + 2 D n + 1 C n D n 1 , for all n N .
Thereby, { ( C 2 n , D 2 n ) } n 0 is a decreasing sequence consisting of closed and convex pairs in A 0 × B 0 . Furthermore, for all n N { 0 } we have:
T ( D 2 n ) T ( C 2 n 1 ) con ¯ ( T ( C 2 n 1 ) ) = C 2 n ,
T ( C 2 n ) T ( D 2 n 1 ) con ¯ ( T ( D 2 n 1 ) ) = D 2 n .
So, we deduce that ( C 2 n , D 2 n ) is T-invariant. Let ( x , y ) C 0 × D 0 be such that x y = dist ( A , B ) . Then, ( T 2 n x , T 2 n y ) C 2 n × D 2 n and by the fact that T is relatively nonexpansive, we have:
dist ( C 2 n , D 2 n ) T 2 n x T 2 n y x y = dist ( A , B ) .
We can see that ( C 2 n , D 2 n ) is also proximinal for all n N . Notice that if:
max { { μ ( C 2 k ) , μ ( D 2 k ) } = 0 f o r s o m e k N ,
then the result follows from Theorem 42. So, we assume that max { { μ ( C 2 n ) , μ ( D 2 n ) } > 0 for all n N . Then, we obtain min { { μ ( C 2 n ) , μ ( D 2 n ) } > 0 for all n N . Since T is cyclic φ -condensing, for all n N { 0 } we have:
μ ( C 2 n + 2 D 2 n + 2 ) = max { μ ( C 2 n + 2 ) , μ ( D 2 n + 2 ) } max { μ ( D 2 n + 1 ) , μ ( C 2 n + 1 ) } = max { μ ( con ¯ ( T ( D 2 n ) ) ) , μ ( con ¯ ( T ( C 2 n ) ) ) } = max { μ ( ( T ( C 2 n ) ) ) , μ ( ( T ( D 2 n ) ) ) } = μ ( T ( C 2 n ) T ( D 2 n ) ) φ ( μ ( C 2 n D 2 n ) ) μ ( C 2 n D 2 n ) μ ( C 2 n D 2 n ) .
Consequently,
lim n μ ( C 2 n D 2 n ) = max { lim n μ ( C 2 n ) , lim n μ ( D 2 n ) } = 0 .
If we set C = n = 0 C 2 n and D = n = 0 D 2 n then ( C , D ) is nonempty, closed, convex, and T–invariant with dist ( A , B ) = dist ( C , D ) for which we have max { μ ( C ) , μ ( ( D ) ) }   = 0 . Hence, T has a best proximity point. □
Theorem 50
([87]). Let ( A , B ) be a bounded, closed, and convex pair in a strictly convex Banach space X such that A 0 is nonempty and μ is an MNC on X. Let T : A B A B be a noncyclic, relatively nonexpansive map, which is φ-condensing in the sense of Definition 22. Then, T has a best proximity pair.
Proof. 
Note that ( A 0 , B 0 ) is closed, convex, and proximinal. Let x A 0 . Then, there exists y B 0 such that x y = dist ( A , B ) . Since T is relatively nonexpansive, T x T y = dist ( A , B ) , and so T x A 0 . Thus, T ( A 0 ) A 0 . Similarly, T ( B 0 ) B 0 , which implies that ( A 0 , B 0 ) is T-invariant. Set C 0 = A 0 and D 0 = B 0 and for all n N define:
C n = con ¯ ( T ( C n 1 ) ) , D n = con ¯ ( T ( D n 1 ) ) .
Then, we have:
C 1 = con ¯ ( T ( C 0 ) ) = con ¯ ( T ( A 0 ) ) A 0 = C 0 .
Continuing this process and by induction we obtain C n 1 C n for all n N . Equivalently, D n 1 D n for all n N . Suppose that there exists k N for which max { μ ( C k ) , μ ( D k ) } = 0 . Then, ( C k , D k ) is a compact pair. Moreover, we have:
T ( C k ) con ¯ ( T ( C k ) ) = C k + 1 C k .
A similar argument implies that T ( D k ) D k and so, T is noncyclic relatively nonexpansive on C k D k , where ( C k , D k ) is a compact and convex pair in a strictly convex Banach space X. Thus, from Theorem 43, T has a best proximity pair and we are finished.
So, we assume that: max { μ ( C n ) , μ ( D n ) } > 0 for any n N . If there exist l 1 , l 2 N with l 1 < l 2 such that μ ( C l 1 ) = μ ( D l 2 ) = 0 then, by the fact that the sequence { C n } n N { 0 } is a decreasing sequence, we have C l 2 C l 1 and so, μ ( C l 2 ) μ ( C l 1 ) which leads to μ ( C l 2 ) = 0 . Hence max { μ ( C l 2 ) , μ ( D l 2 ) } = 0 which is a contradiction, and so
min { μ ( C n ) , μ ( D n ) } > 0 , for all n N { 0 } .
Also, for the pair ( x , y ) C 0 × D 0 with x y = dist ( A , B ) we have T n x T n y = dist ( A , B ) for all n N , because of the fact that T is noncyclic relatively nonexpansive. >From the definition of the pair ( C n , D n ) we obtain ( T n x , T n y ) C n × D n which implies that
dist ( C n , D n ) = dist ( A , B ) , for all n N .
Now suppose that u C 1 = con ¯ ( T ( C 0 ) ) . Then u = j = 1 m c j T ( u j ) where u j C 0 for all 1 j m such that c j 0 and j = 1 m c j = 1 . Since ( C 0 , D 0 ) is proximinal, for all 1 j m there exists v j D 0 such that u j v j = dist ( C 0 , D 0 ) ( = dist ( A , B ) ) and so T u j T v j = dist ( A , B ) . Put v : = j = 1 m c j T ( v j ) . Then v D 1 and:
u v = j = 1 m c j T ( u j ) j = 1 m c j T ( v j ) j = 1 m T ( u j ) T ( v j ) = dist ( A , B ) .
Therefore, the pair ( C 1 , D 1 ) is proximinal. Using a similar discussion, we can see that the pair ( C n , D n ) is proximinal for all n N { 0 } . Thus, ( C n , D n ) is a nonempty, bounded, closed, convex, and proximinal pair, which is T-invariant. Since T is noncyclic φ -condensing, for all n N { 0 } we have:
μ ( C n + 1 D n + 1 ) = max { μ ( C n + 1 ) , μ ( D n + 1 ) } = max { μ ( con ¯ ( T ( C n ) ) ) , μ ( con ¯ ( T ( D n ) ) ) } = max { μ ( ( T ( C n ) ) ) , μ ( ( T ( D n ) ) ) } = μ ( T ( C n ) T ( D n ) ) φ ( μ ( C n D n ) ) μ ( C n D n ) μ ( C n D n ) .
Then, { μ ( C n D n ) } is a decreasing sequence and bounded below, so there exists a real number r 0 such that lim n μ ( C n D n ) = r . We claim that r = 0 . Suppose the contrary. Thus for all n N we have:
μ ( C n + 1 D n + 1 ) μ ( C n D n ) φ ( μ ( C n D n ) ) .
The above inequality yields lim n φ ( μ ( C n D n ) ) = 1 . In view of the fact that φ Φ , we conclude that r = 0 which is impossible. Hence,
lim n μ ( C n D n ) = max { lim n μ ( C n ) , lim n μ ( D n ) } = 0 .
So the pair ( C , D ) is nonempty, closed, and convex, which is T-invariant, where C = n = 0 C n and D = n = 0 D n . Furthermore, dist ( C , D ) = dist ( A , B ) and it is easy to check that ( C , D ) is proximinal. On the other hand, max { μ ( C ) , μ ( ( D ) ) } = 0 , which ensures that the pair ( C , D ) is compact. Finally, the result follows from Theorem 43. □
At the end of this section, we give the following existence theorems which were recently presented in [89] as generalizations of Sadovskii’s fixed point problem.
Theorem 51
([89]). Let ( A , B ) be a bounded, closed, and convex pair in a Banach space X such that A 0 and μ be an MNC on X. Let T : A B A B be a cyclic relatively nonexpansive map such that for any ( H 1 , H 2 ) M T ( A , B ) we have:
μ ( T ( H 1 ) T ( H 2 ) ) μ ( H 1 H 2 ) .
Then, T has a best proximity point.
Proof. 
Let F denote a family of all nonempty, closed, convex proximinal and T-invariant pairs ( C , D ) ( A , B ) . Then ( A 0 , B 0 ) F . Set:
δ : = inf { μ ( C D ) : ( C , D ) F } ,
and assume that ( K 1 , K 2 ) : = ( C , D ) F ( C , D ) . Then, clearly, ( K 1 , K 2 ) F is a nonempty pair for which μ ( K 1 K 2 ) = δ .
Note that if δ = 0 , then μ ( K 1 K 2 ) = 0 and so by Theorem 42, T has a best proximity point in K 1 K 2 . Suppose that μ ( K 1 K 2 ) = δ > 0 . This follows that μ ( T ( K 1 ) T ( K 2 ) ) μ ( K 1 K 2 ) . Since T ( K 1 ) K 2 and T ( K 2 ) K 1 , we have:
μ ( T ( K 1 ) T ( K 2 ) ) < μ ( K 1 K 2 ) .
Let us now define the sets N : = con ¯ ( T ( K 1 ) { y 0 } ) and M : = con ¯ ( T ( K 2 ) { x 0 } ) . Thus, ( x 0 , y 0 ) M × N and ( M , N ) ( K 1 , K 2 ) . Moreover, T ( M ) T ( K 1 ) N and T ( N ) T ( K 2 ) M , that means T is cyclic on M N . Furthermore, if x M , then, x = j = 1 n 1 c j T ( y j ) + c n x 0 , where y j K 2 for all j { 1 , 2 , n 1 } for which c i 0 , j = 1 n c j = 1 . Since ( K 1 , K 2 ) is proximinal, there exists x j K 1 so that x j y j = dist ( A , B ) for all j { 1 , 2 , , n 1 } . Now, if y = j = 1 n 1 c j T ( x j ) + c n y 0 , then, y N and we have x y = dist ( A , B ) . Therefore, M 0 = M . Similarly, N 0 = N and so, ( M , N ) is a proximinal pair. Hence, ( M , N ) F . Considering the definition of ( K 1 , K 2 ) , it follows that M = K 1 and N = K 2 . Therefore,
μ ( M N ) = max { μ ( M ) , μ ( N ) } = max { μ ( con ¯ ( T ( K 2 ) { x 0 } ) ) , μ ( con ¯ ( T ( K 1 ) { y 0 } ) ) } = max { μ ( T ( K 1 ) ) , μ ( T ( K 2 ) ) } = μ ( T ( K 1 ) T ( K 2 ) ) < μ ( K 1 K 2 ) = δ ,
which is a contradiction. □
Theorem 52
([89]). Let ( A , B ) be a bounded, closed, and convex pair in a strictly convex Banach space X, such that A 0 is nonempty and μ be an MNC on X. Let T : A B A B be a noncyclic relatively nonexpansive map such that for any ( H 1 , H 2 ) M T ( A , B ) we have:
μ ( T ( H 1 ) T ( H 2 ) ) μ ( H 1 H 2 ) .
Then, T has a best proximity pair.
Proof. 
Let ( u , v ) A 0 × B 0 such that u v = d i s t ( A , B ) and G denote the family of all nonempty, closed, convex, proximinal and T–invariant pairs ( E , F ) ( A , B ) such that ( u , v ) E × F and T ( E ) E and T ( F ) F . Then ( A 0 , B 0 ) G . Let:
δ : = inf { μ ( E F ) : ( E , F ) G } ,
and define ( K 1 , K 2 ) = ( E , F ) G ( E , F ) . Then, clearly, ( K 1 , K 2 ) G is a nonempty pair such that μ ( K 1 K 2 ) = δ . If δ = 0 then μ ( K 1 K 2 ) = 0 and the result follows from Theorem 43.
Suppose that μ ( K 1 K 2 ) = δ > 0 . It follows that μ ( T ( K 1 ) T ( K 2 ) ) μ ( K 1 K 2 ) . Since T ( K 1 ) K 1 and T ( K 2 ) K 2 , we have:
μ ( T ( K 1 ) T ( K 2 ) ) < μ ( K 1 K 2 ) .
Set H 1 : = con ¯ ( T ( K 1 ) { u } ) and H 2 : = con ¯ ( T ( K 2 ) { v } ) . Thus, ( u , v ) H 1 × H 2 and ( H 1 , H 2 ) ( K 1 , K 2 ) . Moreover, T ( H 1 ) T ( K 1 ) H 1 , T ( H 2 ) T ( K 2 ) H 2 . Therefore, T is noncyclic on H 1 H 2 . Thus, if x H 1 , then x = j = 1 n 1 c j T ( u j ) + c n u , where u j K 1 for all j { 1 , 2 , , n 1 } for which c j 0 and j = 1 n c j = 1 . From the fact that ( K 1 , K 2 ) is proximinal, there exists v j K 2 such that u j v j = dist ( A , B ) for all j { 1 , 2 , , n 1 } . Now, if we define y = j = 1 n 1 c j T ( v j ) + c n v , then y H 2 and x y = dist ( A , B ) . Hence, ( H 1 ) 0 = H 1 . By similar argument, ( H 2 ) 0 = H 2 and hence, ( H 1 , H 2 ) is a proximinal pair. Further, from the definition of ( K 1 , K 2 ) , we have H 1 = K 1 and H 2 = K 2 . Therefore, we have ( H 1 , H 2 ) G . Thus:
μ ( H 1 H 2 ) = μ ( T ( K 1 ) T ( K 2 ) ) < μ ( K 1 K 2 ) = δ ,
That is, μ ( H 1 H 2 ) < δ which is contradiction. □

10. Application to a System of Differential Equations

In this section, we present some applications of the existence results of best proximity points in order to establish the optimal solutions for various systems of differential equations.
Application A.
We begin with the following extension of the Mean-Value Theorem.
Theorem 53
([8]). Let J be a real interval, X be a Banach space, and f : J X be a differentiable map. Let a , b J with a < b . Then:
f ( b ) f ( a ) ( b a ) con ¯ ( { f ( t ) : t [ a , b ] } ) .
Now, we apply the existence theorems of best proximity points to solve the systems of initial-value problems in Banach spaces. To this end, we introduce the following notion.
Definition 23.
Let a and b be real positive numbers, I be the real interval [ t 0 a , t 0 + a ] and V 1 = B ( x 0 ; b ) , V 2 = B ( x 1 ; b ) be closed balls in a Banach space X, where t 0 is a real number and x 0 , x 1 X . Assume that f : I × V 1 X and g : I × V 2 X are continuous maps. Consider the following system of differential equations:
x ( t ) = f ( t , x ( t ) ) ; x ( t 0 ) = x 0 ,
y ( t ) = g ( t , y ( t ) ) ; y ( t 0 ) = x 1 ,
defined on a closed real interval J = [ t 0 h , t 0 + h ] for some real positive number h. Let us consider the Banach space C ( J , X ) of continuous maps from J into X with the supremum norm and define C ( J , V 1 ) = { x C ( J , X ) : x ( t 0 ) = x 0 } and C ( J , V 2 ) = { y C ( J , X ) : y ( t 0 ) = x 1 } . In this case, for any ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) we have:
x y = sup t J x ( t ) y ( t ) x 0 x 1 ,
and so, dist ( C ( J , V 1 ) , C ( J , V 2 ) ) = x 0 x 1 . Let:
T : C ( J , V 1 ) C ( J , V 1 ) C ( J , X ) ,
be an operator defined as:
T x ( t ) = x 1 + t 0 t g ( s , x ( s ) ) d s ; x C ( J , V 1 ) ,
T y ( t ) = x 0 + t 0 t f ( s , y ( s ) ) d s ; y C ( J , V 2 ) .
We say that z C ( J , V 1 ) C ( J , V 2 ) is an optimal solution for the system of differential equations given in (62) and (63) provided that:
z T z = dist ( A , B ) .
Here, we state the following existence theorem.
Theorem 54
([83]). Under the assumptions of Definition 23 if,
α ( f ( I × W 2 ) g ( I × W 1 ) ) r α ( W 1 W 2 ) ,
f ( t , x ) g ( t , y ) 1 h ( x ( t ) y ( t ) x 1 x 0 ) ,
for some r ] 0 , 1 [ and for any ( W 1 , W 2 ) ( V 1 , V 2 ) and h < min { a , b M 1 , b M 2 , 1 r } , where M 1 = sup { f ( t , x ) : ( t , x ) I × V 1 } and M 2 = sup { g ( t , y ) : ( t , y ) I × V 2 } , then the systems (62) and (63) have an optimum solution.
Proof. 
Clearly, ( C ( J , V 1 ) , C ( J , V 2 ) ) is a bounded, closed, and convex pair in C ( J , X ) and T is cyclic on C ( J , V 1 ) C ( J , V 2 ) . We now prove that T ( C ( J , V 1 ) ) is a bounded and equicontinuous subset of C ( J , V 2 ) . Suppose t , t J and x C ( J , V 1 ) . Then we have:
T x ( t ) T x ( t ) = t 0 t g ( s , x ( s ) ) d s t 0 t g ( s , x ( s ) ) d s t t g ( s , x ( s ) ) d s M 2 | t t | ,
that is, T ( C ( J , V 1 ) ) is equicontinuous. Equivalently, we can see that T ( C ( J , V 2 ) ) is also bounded and equicontinuous. Now, from the Arzela–Ascoli theorem, we conclude that the pair ( C ( J , V 1 ) , C ( J , V 2 ) ) is relatively compact. In the following, we verify that T is a condensing operator. Let ( K 1 , K 2 ) ( C ( J , V 1 ) , C ( J , V 2 ) ) be nonempty, closed, convex, and proximinal pair, which is T-invariant such that dist ( K 1 , K 2 ) = dist ( C ( J , V 1 ) , C ( J , V 2 ) )   ( = x 0 x 1 ) . From ([7], Theorem 2.11) we deduce that:
α ( T ( K 1 ) , T ( K 2 ) ) = max { α ( T ( K 1 ) ) , α ( T ( K 2 ) ) } = max sup t J { α ( { T x ( t ) : x K 1 } ) } , sup t J { α ( { T y ( t ) : y K 2 } ) } = max sup t J { α ( { x 1 + t 0 t g ( s , x ( s ) ) d s : x K 1 } ) } , sup t J { α ( { x 0 + t 0 t f ( s , y ( s ) ) d s : y K 2 } ) } } .
On the other hand, using Theorem 53 we obtain:
x 1 + t 0 t g ( s , x ( s ) ) d s x 1 + ( t t 0 ) con ¯ ( { g ( s , x ( s ) ) : s [ t 0 , t ] } ) ,
x 0 + t 0 t f ( s , y ( s ) ) d s x 0 + ( t t 0 ) con ¯ ( { f ( s , y ( s ) ) : s [ t 0 , t ] } ) ,
and thus,
α ( T ( K 1 ) , T ( K 2 ) ) max { sup t J { α ( { x 1 + ( t t 0 ) con ¯ ( { g ( s , x ( s ) ) : s [ t 0 , t ] } ) } ) } ,
sup t J { α ( { x 0 + ( t t 0 ) con ¯ ( { f ( s , y ( s ) ) : s [ t 0 , t ] } ) } ) } } .
max { sup 0 λ h { α ( { x 1 + λ con ¯ ( { g ( J × K 1 ) } ) } ) } , sup 0 λ h { α ( { x 0 + λ con ¯ ( { f ( J × K 2 ) } ) } ) } } .
= h α ( { g ( J × K 1 ) f ( J × K 2 ) } ) h r α ( K 1 K 2 ) .
Since h r < 1 , we conclude that T is a condensing operator. Finally, we show that T is cyclic relatively nonexpansive. From the assumptions of theorem, for any ( x , y ) C ( J , V 1 ) × C ( J , V 1 ) we have:
T x ( t ) T y ( t ) = ( x 1 + t 0 t g ( s , x ( s ) ) d s ) ( x 0 + t 0 t f ( s , y ( s ) ) d s ) x 1 x 0 + t 0 t g ( s , x ( s ) ) f ( s , y ( s ) ) ) d s x 1 x 0 + 1 h t 0 t ( x ( s ) y ( s ) x 1 x 0 ) d s x 1 x 0 + ( x y x 1 x 0 ) = x y ,
and thereby, T x T y x y . Now the result follows from Theorem 44. □
Application B.
In what follows, let a , b , h be positive real numbers with h < a . For a given real number t 0 and a Banach space X, we consider the Banach space C ( I , X ) of continuous maps from I = [ t 0 a , t 0 + a ] into X, endowed with the supremum norm. Furthermore, let V 1 = B ( x 1 ; b ) and V 2 = B ( x 2 ; b ) be closed balls in X, where x 1 , x 2 X . Assume that k i : I × I × V i X and f i : I × V i × V i X , with i = 1 , 2 , continuous maps, and k i is k i -invariant. Here, we consider the problem:
x ( t ) = f 1 ( t , x ( t ) , t 0 t k 1 ( t , s , x ( s ) ) d s ) , x ( t 0 ) = x 1 , y ( t ) = f 2 ( t , y ( t ) , t 0 t k 2 ( t , s , y ( s ) ) d s ) , y ( t 0 ) = x 2 ,
where the integral is the Bochner integral. Let J = [ t 0 h , t 0 + h ] and define C ( J , V 1 ) = { x C ( J , X ) : x ( t 0 ) = x 1 } and C ( J , V 2 ) = { y C ( J , X ) : y ( t 0 ) = x 2 } . Clearly, ( C ( J , V 1 ) , C ( J , V 2 ) ) is a bounded, closed, and convex pair in C ( J , X ) . Thus, for any ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) , we have x 1 x 2 sup t J x ( t ) y ( t ) = x y , and so,
dist ( C ( J , V 1 ) , C ( J , V 2 ) ) = x 1 x 2 .
Now, let T : C ( J , V 1 ) C ( J , V 2 ) C ( J , X ) be the operator defined as:
T x ( t ) = x 2 + t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ , x C ( J , V 1 ) , x 1 + t 0 t f 2 ( σ , x ( σ ) , t 0 σ k 2 ( σ , s , x ( s ) ) d s ) d σ , x C ( J , V 2 ) .
We show that T is a cyclic operator. Indeed, for x C ( J , V 1 ) we have:
T x ( t ) x 2 = t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ | t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ | M 1 h ,
where M i = sup { f i ( t , x ( t ) , t 0 t k i ( t , s , x ( s ) ) d s ) : ( t , x ) I × V i } , i = 1 , 2 . Now, if we assume h < b max i { 1 , 2 } M i , we get T x ( t ) x 2 b for all t J and so T x C ( J , V 2 ) . The same argument shows that x C ( J , V 2 ) implies T x C ( J , V 1 ) .
Taking into account the above notions and notation, for 0 < h < b max i { 1 , 2 } M i , the hypotheses are as follows:
( H 1 )
Let μ be an MNC on C ( J , X ) such that for any r > 0 there exists δ ( r ) > 0 such that r μ ( W 1 W 2 ) < r + δ ( r ) for any bounded ( W 1 , W 2 ) ( V 1 , V 2 ) implies:
μ ( f 1 ( I × W 1 × W 1 ) f 2 ( I × W 2 × W 2 ) ) < r h ;
( H 2 )
also,
f 1 ( t , x ( t ) , t 0 t k 1 ( t , s , x ( s ) ) d s ) f 2 ( t , y ( t ) , t 0 t k 2 ( t , s , y ( s ) ) d s ) 1 h ( x ( t ) y ( t ) x 1 x 0 ) , for all ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) .
We recall another extension of the Mean–Value Theorem, which we arrange according to our notation and further use.
Theorem 55.
Let I, J, X, V i , k i : I × I × V i X and f i : I × V i × V i X with i = 1 , 2 be given as above. Let t 0 , t J with t 0 < t . Then:
x j + t 0 t f i ( σ , x ( σ ) , t 0 σ k i ( σ , s , x ( s ) ) d s ) d σ x j + ( t t 0 ) c o n ¯ ( { f i ( σ , x ( σ ) , t 0 σ k i ( σ , s , x ( s ) ) d s ) : σ [ t 0 , t ] } ) ,
with ( i , j ) { ( 1 , 2 ) , ( 2 , 1 ) } .
We say that z C ( J , V 1 ) C ( J , V 2 ) is an optimal solution for the system (64) provided that z T z = dist ( C ( J , V 1 ) , C ( J , V 2 ) ) , that is, z is a best proximity point of the operator T in (65). Then we give the following result.
Theorem 56
([86]). If the hypotheses ( H 1 ) , ( H 2 ) and h < b max i { 1 , 2 } M i are satisfied, then the problem (64) has an optimal solution.
Proof. 
Since T is a cyclic operator, it follows trivially that T ( C ( J , V 1 ) ) is a bounded subset of C ( J , V 2 ) . So, we prove that T ( C ( J , V 1 ) ) is also an equicontinuous subset of C ( J , V 2 ) . Suppose t , t J and x C ( J , V 1 ) . We observe that:
T x ( t ) T x ( t ) = t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ | t t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ | M 1 | t t | ,
that is, T ( C ( J , V 1 ) ) is equicontinuous. The same argument is valid for T ( C ( J , V 2 ) ) and hence, to avoid repetition, we omit the details. Moreover, by use of the Arzelà–Ascoli theorem, it follows that the pair ( C ( J , V 1 ) , C ( J , V 2 ) ) is relatively compact. Here, we show that T is a Meir–Keeler condensing operator. Let ( K 1 , K 2 ) ( C ( J , V 1 ) , C ( J , V 2 ) ) be a closed, convex, and proximinal pair, which is T-invariant and such that dist ( K 1 , K 2 ) = dist ( C ( J , V 1 ) , C ( J , V 2 ) ) ( = x 1 x 2 ) . Using a generalized version of the Arzelà–Ascoli theorem (see Ambrosetti [90]) and hypothesis ( H 1 ) , we get:
μ ( T ( K 1 ) T ( K 2 ) ) = max { μ ( T ( K 1 ) ) , μ ( T ( K 2 ) ) } = max { sup t J { μ ( { T x ( t ) : x K 1 } ) } , sup t J { μ ( { T y ( t ) : y K 2 } ) } } = max { sup t J { μ ( { x 2 + t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ : x K 1 } ) } , sup t J { μ ( { x 1 + t 0 t f 2 ( σ , y ( σ ) , t 0 σ k 2 ( σ , s , y ( s ) ) d s ) d σ : y K 2 } ) } } .
So, in view of (66), it follows that:
μ ( T ( K 1 ) T ( K 2 ) ) max { sup t J { μ ( { x 2 + ( t t 0 ) c o n ¯ ( { f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) : σ [ t 0 , t ] } ) } ) } , sup t J { μ ( { x 1 + ( t t 0 ) c o n ¯ ( { f 2 ( σ , x ( σ ) , t 0 σ k 2 ( σ , s , x ( s ) ) d s ) : σ [ t 0 , t ] } ) } ) } } max { sup 0 λ h { μ ( { x 2 + λ c o n ¯ ( { f 1 ( J × K 1 × K 1 ) } ) } ) } , sup 0 λ h { μ ( { x 1 + λ c o n ¯ ( { f 2 ( J × K 2 × K 2 ) } ) } ) } } = max { h μ ( f 1 ( J × K 1 × K 1 ) ) , h μ ( f 2 ( J × K 2 × K 2 ) ) } = h μ ( { f 1 ( J × K 1 × K 1 ) f 2 ( J × K 2 × K 2 ) } ) < h r h = r .
We conclude that T is a Meir–Keeler condensing operator. The last step of the proof is to show that T is cyclic relatively nonexpansive. Indeed, for any ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) we have:
T x ( t ) T y ( t ) = ( x 2 + t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) d σ ) ( x 1 + t 0 t f 2 ( σ , x ( σ ) , t 0 σ k 2 ( σ , s , x ( s ) ) d s ) d σ x 1 x 2 + | t 0 t f 1 ( σ , x ( σ ) , t 0 σ k 1 ( σ , s , x ( s ) ) d s ) f 1 ( σ , x ( σ ) , t 0 σ k 2 ( σ , s , y ( s ) ) d s ) d σ | x 1 x 2 + 1 h | t 0 t ( x ( s ) y ( s ) x 1 x 2 ) d s | ( by hypothesis ( H 2 ) ) x 1 x 2 + ( x y x 1 x 2 ) = x y ,
and thereby, T x T y x y . All the hypotheses of Theorem 46 hold and so the operator T has a best proximity point z C ( J , V 1 ) C ( J , V 2 ) , which is an optimal solution for the system (64). □
Application C.
Let a , b , h be positive real numbers with h < a . For a given real number t 0 and a Banach space X, we consider the Banach space C ( I , X ) of continuous maps from I = [ t 0 a , t 0 + a ] into X, endowed with the supremum norm. Furthermore, let V 1 = B ( x * ; b ) and V 2 = B ( x * * ; b ) be closed balls in X, where x * , x * * X . Assume that f : I × V 1 X and g : I × V 2 X are continuous maps. So, we recall the problem:
x ( t ) = f ( t x ( t ) ) , x ( t 0 ) = x * , y ( t ) = g ( t , y ( t ) ) , y ( t 0 ) = x * * .
Let J = [ t 0 h , t 0 + h ] and define C ( J , V 1 ) = { x C ( J , X ) : x ( t 0 ) = x * } , C ( J , V 2 ) = { y C ( J , X ) : y ( t 0 ) = x * * } . Clearly, ( C ( J , V 1 ) , C ( J , V 2 ) ) is a bounded, closed, and convex pair in C ( J , X ) . Moreover, for any ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) we have x 1 x 2 sup t J x ( t ) y ( t ) = x y , and so, dist ( C ( J , V 1 ) , C ( J , V 2 ) ) = x 1 x 2 .
Now, let T : C ( J , V 1 ) C ( J , V 2 ) C ( J , X ) be the operator defined as:
T x ( t ) = x * * + t 0 t g ( σ , x ( σ ) ) d σ , x C ( J , V 1 ) , x * + t 0 t f ( σ , x ( σ ) ) d σ , x C ( J , V 2 ) .
We show that T is a cyclic operator. Indeed, for x C ( J , V 1 ) we have:
T x ( t ) x * * = t 0 t g ( σ , x ( σ ) ) d σ t 0 t g ( σ , x ( σ ) ) d σ M 1 h ,
where M 1 = sup { g ( t , x ( t ) ) : ( t , x ) I × V 2 } (analogously, M 2 = sup { f ( t , x ( t ) ) : ( t , x ) I × V 1 } ). Now, if we assume h < min { b max i { 1 , 2 } M i , 1 2 b } , we get T x ( t ) x * * b for all t J , and so, T x C ( J , V 2 ) . The same argument shows that x C ( J , V 2 ) implies T x C ( J , V 1 ) .
Taking into account the above notions and notation, for 0 < h < min { b max i { 1 , 2 } M i , 1 2 b } , the hypotheses are as follows:
( H 1 )
There exists ψ Ψ such that α ( f ( I × W 2 ) g ( I × W 1 ) ) 2 b ψ ( α ( W 1 W 2 ) ) for any ( H 1 , H 2 ) ( V 1 , V 2 ) ;
( H 2 )
f ( t , x ( t ) ) g ( t , y ( t ) ) 1 h ( x ( t ) y ( t ) x * * x * ) , for all ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) .
We recall the following extension of the Mean–Value Theorem, which we arrange according to our notation and further use.
Theorem 57.
Let I, J, X, f : I × V 1 X , g : I × V 2 X be given as above. Let t 0 , t J with t 0 < t . Then:
x * + t 0 t f ( σ , x ( σ ) ) d σ x * + ( t t 0 ) c o n ¯ ( { f ( σ , x ( σ ) ) : σ [ t 0 , t ] } ) ,
x * * + t 0 t g ( σ , x ( σ ) ) d σ x * * + ( t t 0 ) c o n ¯ ( { g ( σ , x ( σ ) ) : σ [ t 0 , t ] } ) .
Furthermore, we need the next generalization of the Arzela–Ascoli theorem.
Theorem 58
([90]). Let X be a Banach space, D R n compact and B C ( D , X ) a bounded and equicontinuous set. Then α ( B ) = sup t D α ( { x ( t ) : x B } ) .
We say that z C ( J , V 1 ) C ( J , V 2 ) is an optimal solution for the system (67) provided that z T z = dist ( C ( J , V 1 ) , C ( J , V 2 ) ) , that is, z is a best proximity point of the operator T in (68). Then, we give the following result.
Theorem 59
([86]). If the hypotheses ( H 1 ) , ( H 2 ) and h < min { b max i { 1 , 2 } M i , 1 2 b } are satisfied, then the problem (67) has an optimal solution.
Proof. 
Since T is a cyclic operator, it follows trivially that T ( C ( J , V 1 ) ) is a bounded subset of C ( J , V 2 ) . So, we prove that T ( C ( J , V 1 ) ) is also an equicontinuous subset of C ( J , V 2 ) . Suppose t , t J and x C ( J , V 1 ) . We observe that:
T x ( t ) T x ( t ) = t 0 t g ( σ , x ( σ ) ) d σ t 0 t g ( σ , x ( σ ) ) d σ t t g ( σ , x ( σ ) ) d σ M 1 | t t | ,
that is, T ( C ( J , V 1 ) ) is equicontinuous. The same argument is valid for T ( C ( J , V 2 ) ) and hence, to avoid repetition, we omit the details. Here, we show that T is a generalized condensing operator. Let ( K 1 , K 2 ) ( C ( J , V 1 ) , C ( J , V 2 ) ) be a closed, convex, and proximinal pair, which is T-invariant and such that dist ( K 1 , K 2 ) = dist ( C ( J , V 1 ) , C ( J , V 2 ) ) ( = x * * x * ) . By Theorem 58 and hypothesis ( H 1 ) , we obtain:
α ( T ( K 1 ) T ( K 2 ) ) = max { α ( T ( K 1 ) ) , α ( T ( K 2 ) ) } = max { sup t J { α ( { T x ( t ) : x K 1 } ) } , sup t J { α ( { T y ( t ) : y K 2 } ) } } = max { sup t J { α ( { x * * + t 0 t g ( σ , x ( σ ) ) d σ : x K 1 } ) } , sup t J { α ( { x * + t 0 t f ( σ , y ( σ ) ) d σ : y K 2 } ) } } .
So, in view of (69) and (70), it follows that:
α ( T ( K 1 ) T ( K 2 ) ) max { sup t J { α ( { x * * + ( t t 0 ) c o n ¯ ( { g ( σ , x ( t ) ) : σ [ t 0 , t ] } ) } ) } , sup t J { α ( { x * + ( t t 0 ) c o n ¯ ( { f ( t , x ( t ) ) : σ [ t 0 , t ] } ) } ) } } max { sup 0 λ h { α ( { x * * + λ c o n ¯ ( { g ( J × K 1 ) } ) } ) } , sup 0 λ h { α ( { x * + λ c o n ¯ ( { f ( J × K 2 ) } ) } ) } } = max { h α ( g ( J × K 1 ) ) , h α ( f ( J × K 2 ) ) } = h α ( { g ( J × K 1 ) f ( J × K 2 ) } ) 1 2 b 2 b ψ ( α ( K 1 K 2 ) ) = ψ ( α ( K 1 K 2 ) ) .
We conclude that T is a generalized condensing operator. The last step of the proof is to show that T is cyclic relatively nonexpansive. Indeed, for any ( x , y ) C ( J , V 1 ) × C ( J , V 2 ) we have:
T x ( t ) T y ( t ) = ( x * * + t 0 t g ( σ , x ( σ ) ) d σ ) ( x * + t 0 t f ( σ , x ( σ ) ) d σ x * * x * + t 0 t ( g ( σ , x ( σ ) ) f ( σ , x ( σ ) ) ) d σ x * * x * + 1 h t 0 t ( x ( s ) y ( s ) x * * x * ) d s ( by hypothesis ( H 2 ) ) x * * x * + ( x y x * * x * ) = x y ,
and thereby, T x T y x y . All the hypotheses of Theorem 47 hold and so the operator T has a best proximity point z C ( J , V 1 ) C ( J , V 2 ) , which is an optimal solution for the system (63). □
An application of a coupled measure of noncompactness can be found in the recent paper [91].

11. Concluding Remarks

We gave a survey of measures of noncompactness and their most important properties. Furthermore, we discussed some fixed point theorems of Darbo type.
First, we applied measures of noncompactness in characterizing classes of compact operators between certain sequence spaces, and in solving infinite systems of integral equations in some sequence and function spaces.
Second, we included some recent results related to the existence of best proximity points (pairs) for some classes of cyclic and noncyclic condensing operators in Banach spaces equipped with a suitable measure of noncompactness.
Finally, we discussed the existence of an optimal solution for systems of integro-differentials.
It is worth mentioning that measures of noncompactness play an important role in nonlinear functional analysis. They are important tools in metric fixed point theory, the theory of operator equations in Banach spaces, and the characterizations of classes of compact operators. They are also applied in the studies of various kinds of differential and integral equations.

Author Contributions

All authors made equal contributions to the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research did not receive any funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kuratowski, K. Sur les espaces complets. Fund. Math. 1930, 15, 301–309. [Google Scholar] [CrossRef] [Green Version]
  2. Darbo, G. Punti uniti in transformazioni a condominio non compatto. Rend. Sem. Math. Univ. Padova 1955, 24, 84–92. [Google Scholar]
  3. Гольденштейн, Л.С.; Гохберг, И.Ц.; Маркус, А.С. Исследование некоторых свойств линейных ограниченных операторов в связи с их q-нормой. Уч. заn. Кишиневского гос. ун-та 1957, 29, 29–36. [Google Scholar]
  4. Гольденштейн, Л.С.; Маркус, А.С. О мере некомпактности ограниченных множеств и линейных операторов. В кн. Исследование по алгебре и математическому анализу Кищинев Картя Молдавеняске 1965, 45–54. [Google Scholar]
  5. Banaś, J.; Goebel, K. Measures of Noncompactness in Banach Spaces; Lecture Notes in Pure and Applied Mathematics; Marcel Dekker Inc.: New York, NY, USA; Basel, Switzerland, 1980; Volume 60. [Google Scholar]
  6. Akhmerov, R.R.; Kamenskii, M.I.; Potapov, A.S.; Rodkina, A.E.; Sadovskii, B.N. Measures of Noncompactness and Condensing Operators; Birkhäuser Verlag: Basel, Switzerland, 1992. [Google Scholar]
  7. Toledano, J.M.A.; Benavides, T.D.; Acedo, G.L. Measures of Noncompactness in Metric Fixed Point Theory; Vol. 99, Operator Theory Advances and Applications; Birkhäuser Verlag: Basel, Switzerland; Boston, MA, USA; Berlin, Germany, 1997. [Google Scholar]
  8. Istrǎţescu, V. Fixed Point Theory, An Introduction; Reidel Publishing Company: Dordrecht, The Netherlands; Boston, MA, USA; London, UK, 1981. [Google Scholar]
  9. Malkowsky, E.; Rakočević, V. Advanced Functional Analysis; Taylor and Francis: Boca Raton, FL, USA, 2019. [Google Scholar]
  10. Hadžić, O. Fixed Point Theory in Topological Vector Spaces; University of Novi Sad, Institute of Mathematics: Novi Sad, Serbia, 1984. [Google Scholar]
  11. Boos, J. Classical and Modern Methods in Summability; Oxford University Press: Oxford, UK, 2000. [Google Scholar]
  12. Wilansky, A. Functional Analysis; Blaisdell Publishing Company: New York, NY, USA, 1964. [Google Scholar]
  13. Wilansky, A. Summability through Functional Analysis; Mathematical Studies; North–Holland: Amsterdam, The Netherlands, 1984; Volume 85. [Google Scholar]
  14. Wilansky, A. Modern Methods in Topological Vector Spaces; McGraw Hill: New York, NY, USA, 1978. [Google Scholar]
  15. Kamthan, P.K.; Gupta, M. Sequence Spaces and Series; Marcel Dekker: New York, NY, USA, 1981. [Google Scholar]
  16. Zeller, K. Allgemeine Eigenschaften von Limitierungsverfahren. Math. Z. 1951, 53, 463–487. [Google Scholar] [CrossRef]
  17. Zeller, K. Abschnittskonvergenz in FK–Räumen. Math. Z. 1951, 55, 55–70. [Google Scholar] [CrossRef]
  18. Zeller, K. Matrixtransformationen von Folgenräumen. Univ. Rend. Mat. 1954, 12, 340–346. [Google Scholar]
  19. Benavides, T.D.; M. Ayerbe, J. Set–contractions and ball contractions in Lp spaces. J. Math. Anal. Appl. 1991, 159, 500–506. [Google Scholar] [CrossRef] [Green Version]
  20. Webb, J.R.L.; Zhao, W. On connections between set and ball measures of noncompactness. Bull. London Math. Soc. 1990, 22, 471–477. [Google Scholar] [CrossRef]
  21. Ахкмеров, П.П.; Каменнский, М.И.; Потапов и др, А.С. Меры некомпактности и упломняющце орера-моры; Наука: Новосибирск, Russia, 1986. [Google Scholar]
  22. Mallet-Paret, J.; Nussbaum, R.D. Inequivalent measures of nonccompactness and the radius of the essential spectrum. Proc. AMS 2011, 193, 917–930. [Google Scholar] [CrossRef]
  23. Mallet-Paret, J.; Nussbaum, R.D. Inequivalent measures of nonccompactness. Ann. Mat. 2011, 190, 453–488. [Google Scholar] [CrossRef] [Green Version]
  24. Smart, D.R. Fixed Point Theorems; Cambridge University Press: Cambridge, UK, 1980. [Google Scholar]
  25. Agarwal, R.P. Fixed Point Theory and Applications; Cambridge University Press: Cambridge, UK, 2001. [Google Scholar]
  26. Rus, I.A. Metrical Fixed Point Theorems; Univ. of Cluj-Napoca: Cluj-Napoca, Romania, 1979. [Google Scholar]
  27. Khamsi, M.; Kirk, W. An Introduction to Metric Spaces and Fixed Point Theory; John Wiley & Sons, Inc.: New York, MY, USA, 2000. [Google Scholar]
  28. Kirk, W.; Sims, B. Handbook of Metric Fixed Point Theory; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2001. [Google Scholar]
  29. Appell, J. Measures of noncompactness, condensing operators and fixed points: An application oriented survey. Fixed Point Theory 2005, 6, 157–229. [Google Scholar]
  30. Schauder, J. Der Fixpunktsatz in Funktionalräumen. Stud. Math. 1930, 2, 171–180. [Google Scholar] [CrossRef] [Green Version]
  31. Sadovskii, B.N. A fixed point principle. Funct. Anal. 1967, 1, 74–76. [Google Scholar] [CrossRef]
  32. Садовский, Б.Н. Об одном принципе неподвижной точки. Функцион. анализ и эго прил. 1967, 1, 74–76. [Google Scholar]
  33. Kirişci, M. The Hahn sequence space defined by the Cesàro mean. Abstr. Appl. Anal. 2013, 2013. [Google Scholar] [CrossRef]
  34. Kirişci, M. A survey of the Hahn sequence space. Gen. Math. Notes 2013, 19, 37–58. [Google Scholar]
  35. Das, R. On the fine spectrum of the lower triangular matrix B(r;s) over the Hahn sequence space. Kyungpook Math. J. 2017, 57, 441–455. [Google Scholar]
  36. Hahn, H. Über Folgen linearer Operationen. Monatsh. Math. Phys. 1922, 32, 3–88. [Google Scholar] [CrossRef]
  37. Rao, K.C. The Hahn sequence space. Bull. Cal. Math. Soc. 1990, 82, 72–78. [Google Scholar]
  38. Goes, G. Sequences of bounded variation and sequences of Fourier coefficients II. J. Math. Anal. Appl. 1972, 39, 477–494. [Google Scholar] [CrossRef] [Green Version]
  39. Maddox, I.J. On Kuttner’s theorem. London J. Math. Soc. 1968, 43, 285–298. [Google Scholar] [CrossRef]
  40. Kuttner, B.; Thorpe, B. Strong convergence. J. Reine Angew. Math. 1979, 311/312, 42–55. [Google Scholar]
  41. Mòricz, F. On Λ–strong convergence of numerical sequences and Fourier series. Acta Math. Hungar. 1989, 54, 319–327. [Google Scholar] [CrossRef]
  42. Malkowsky, E. The continuous duals of the sequence spaces c0(Λ) and c(Λ) for exponentially bounded sequences Λ. Acta Sci. Math. Szeged 1995, 61, 241–250. [Google Scholar]
  43. Malkowsky, E.; Rakočević, V.; Tuǧ, O. Compact operators on the Hahn space. Monatsh. Math. 2021, 196, 519–551. [Google Scholar] [CrossRef]
  44. Dolićanin-Djekić, D.; Gilić, E. Characterisations of bounded linear and compact operators on the generalised Hahn space. Filomat 2022, 36, 497–505. [Google Scholar] [CrossRef]
  45. Malkowsky, E.; Rakočević, V.; Veličković, V. Bounded linear and compact operators between the Hahn space and spaces of strongly summable and bounded sequences. Bull. Sci. Math. Nat. Sci. Math. 2020, 45, 25–41. [Google Scholar]
  46. Malkowsky, E. Some compact operators on the Hahn space. Sci. Res. Comm. 2021, 1, 1–14. [Google Scholar] [CrossRef]
  47. Sawano, Y.; El-Shabrawy, S.R. Fine spectra of the discrete generalized Cesàro operator on Banach sequence spaces. Monatshefte Math. 2020, 192, 185–224. [Google Scholar] [CrossRef]
  48. Rhaly, H.C. Discrete generalized Cesàro operators. Proc. Amer. Math. Soc. 1982, 86, 405–409. [Google Scholar]
  49. Goldenstein, L.S.; Gohberg, I.C.; Markus, A.S. Investigation of some properties of bounded linear operators in connection with their q–norms. Učen. Zap. Kishinevsk. Univ. 1957, 29, 29–36. [Google Scholar]
  50. Aronszajn, N.; Panitchpakdi, P. Extension of uniformly continuous transformations and hyperconvex metric spaces. Pacific J. Math. 1956, 6, 405–439. [Google Scholar] [CrossRef] [Green Version]
  51. Espínola, R.; Khamsi, M.A. Introduction to Hyperconvex Spaces, Handbook of Metric Fixed Point Theory; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2001; pp. 391–435. [Google Scholar]
  52. Bugajewski, D.; Grzelaczyk, E. On the measures of noncompactness in some metric spaces. N. Z. J. Math. 1998, 27, 177–182. [Google Scholar]
  53. Samadi, A. Applications of measure of noncompactness to coupled fixed points and systems of integral equations. Miskolc Math. Notes 2018, 119, 537–553. [Google Scholar] [CrossRef]
  54. Meir, A.; Keeler, E. A theorem on contraction mappings. J. Math. Anal. Appl. 1969, 28, 326–329. [Google Scholar] [CrossRef] [Green Version]
  55. Aghajani, A.; Mursaleen, M.; Haghighi, A.S. Fixed point theorems for Meir–Keeler condensing operator via measure of noncompactness. Acta Math. Sci. 2015, 35B, 552–566. [Google Scholar] [CrossRef]
  56. Lim, T.C. On characterizations of Meir–Keeler contractive maps. Nonlinear Anal. 2001, 46, 113–120. [Google Scholar] [CrossRef]
  57. Suzuki, T. Fixed point theorem for asymptotic contraction of Meir–Keeler type in complete metric spaces. Nonlinear Anal. 2006, 64, 971–978. [Google Scholar] [CrossRef]
  58. Edelstein, M. On fixed and periodic points under contractive mappings. J. Lond. Math. Soc. 1962, 37, 74–79. [Google Scholar] [CrossRef]
  59. Hajji, A. A generalization of Darbo’s fixed point and common solutions in Banach spaces. Fixed Point Theory Appl. 2013, 62, 1–9. [Google Scholar] [CrossRef] [Green Version]
  60. Hajji, A.; Hanebaly, E. Commuting mappings and α–compact type fixed point theorems in locally convex spaces. Int. J. Math. Anal. 2007, 1, 661–680. [Google Scholar]
  61. Samadi, A.; Avini, M.M.; Mursaleen, M. Solutions of an infinite system of integral equations of Volterra–Stieltjes type in the sequence spaces p(1<p<1) and c0. AIMS Math. 2020, 5, 3791–3808. [Google Scholar]
  62. Malik, I.A.; Jalal, T. Infinite system of integral equations in two variables of Hammerstein type in c0 and 1 spaces. Filomat 2019, 33, 3441–3455. [Google Scholar] [CrossRef]
  63. Banaś, J.; Woś, W. Solvability of an infinite system of integral equations on the real half–axis. Adv. Nonlinear Anal. 2021, 10, 202–216. [Google Scholar] [CrossRef]
  64. Banaś, J.; Chlebowicz, A. On solutions of an infinite system of nonlinear integral equations on the real half–axis. Banach J. Math. Anal. 2019, 13, 944–968. [Google Scholar] [CrossRef]
  65. Banaś, J.; Chlebowicz, A.; Woś, W. On measures of noncompactness in the space of functions defined on the half-axis with values in a Banach space. J. Math. Anal. Appl. 2020, 489. [Google Scholar] [CrossRef]
  66. Banaś, J.; Rzepka, B. An application of a measure of noncompactness in the study of asymptotic stability. Appl. Math. Lett. 2003, 16, 1–6. [Google Scholar] [CrossRef] [Green Version]
  67. Salem, A.; Alshehri, H.M.; Almaghamsi, L. Measure of noncompactness for an infinite system of fractional Langevin equation in a sequence space. Adv. Differ. Equ. 2021, 2021, 1–21. [Google Scholar] [CrossRef]
  68. Abbas, S.; Arifi, N.A.; Benchohra, M.; Graef, J. Periodic mild solutions of infinite delay evolution equations with non–instantaneous impulses. J. Nonlinear Funct. Anal. 2020, 2020, 7. [Google Scholar] [CrossRef]
  69. Delfani, M.; Farajzadeh, A.; Wen, C.F. Some fixed point theorems of generalized Ft–contraction mappings in b–metric spaces. J. Nonlinear Var. Anal. 2021, 5, 515–625. [Google Scholar]
  70. Aghajani, A.; Banas, J.; Sabzali, N. Some generalizations of Darbo’s fixed point theorem and applications. Bull. Belg. Math. Soc. Simon Stevin 2013, 20, 345–358. [Google Scholar] [CrossRef]
  71. Banaś, J. On measures of noncompactness in Banach spaces. Comment. Math. Univ. Carolin. 1980, 21, 131–143. [Google Scholar]
  72. Banaś, J.; ORegan, D.; Agarwal, R.P. Measures of noncompactness and asymptotic stability of solutions of a quadratic Hammerstein integral equation. Rocky Mountain J. Math. 2011, 6, 1769–1792. [Google Scholar] [CrossRef]
  73. Mohiuddine, S.A.; Srivastava, H.M.; Alotaibi, A. Application of measures of noncompactness to the infinite system of second-order differential equations in p spaces. Adv. Differ. Equ. 2016, 2016, 317. [Google Scholar] [CrossRef] [Green Version]
  74. Szegö, G. Orthogonal Polynomials; Amer. Math. Soc.: New York, NY, USA, 1959. [Google Scholar]
  75. Fan, K. Extensions of two fixed point theorems of F. E. Browder. Math. Z. 1969, 122, 234–240. [Google Scholar] [CrossRef]
  76. Abkar, A.; Gabeleh, M. Best proximity points for cyclic mappings in ordered metric spaces. J. Optim. Theory Appl. 2011, 150, 188–193. [Google Scholar] [CrossRef]
  77. Gabeleh, M. Best proximity points and fixed point results for certain maps in Banach spaces. Numer. Funct. Anal. Optim. 2015, 36, 1013–1028. [Google Scholar] [CrossRef]
  78. Markin, J.; Shahzad, N. Best approximation theorems for nonexpansive and condensing mappings in hyperconvex spaces. Nonlin. Anal. 2009, 70, 2435–2441. [Google Scholar] [CrossRef]
  79. Markin, J.; Shahzad, N. Best proximity points for relatively u–continuous mappings in Banach and hyperconvex spaces. Abstract Appl. Anal. 2013, 2013, 1–6. [Google Scholar] [CrossRef] [Green Version]
  80. Eldred, A.A.; Kirk, W.; Veeramani, P. Proximal normal structure and relatively nonexpansive mappings. Studia Math. 2005, 171, 283–293. [Google Scholar] [CrossRef] [Green Version]
  81. Gabeleh, M. A characterization of proximal normal structure via proximal diametral sequences. J. Fixed Point Theory Appl. 2017, 19, 2909–2925. [Google Scholar] [CrossRef]
  82. Dunford, N.; Schwartz, J.T. Linear Operators, Part III, Spectral Operators; Interscience: New York, NY, USA, 1971. [Google Scholar]
  83. Gabeleh, M.; Markin, J. Optimum solutions for a system of differential equations via measure of noncompactness. Indag. Math. (N.S.) 2018, 29, 895–906. [Google Scholar] [CrossRef]
  84. Gabeleh, M. Remarks on minimal sets for cyclic mappings in uniformly convex Banach spaces. Numer. Funct. Anal. Optim. 2017, 38, 360–375. [Google Scholar] [CrossRef]
  85. Gabeleh, M.; Kunzy, H.P. Min-max property in metric spaces with convex structure. Acta Math. Hung. 2019, 157, 173–190. [Google Scholar] [CrossRef]
  86. Gabeleh, M.; Vetro, C. A new extension of Darbo’s fixed point theorem using relatively Meir–Keeler condensing operators. Bull. Aust. Math. Soc. 2018, 98, 286–297. [Google Scholar] [CrossRef]
  87. Gabeleh, M.; Vetro, C. A best proximity point approach to existence of solutions for a system of ordinary differential equations. Bull. Belg. Math. Soc. Simon Stevin 2019, 26, 493–503. [Google Scholar] [CrossRef]
  88. Gabeleh, M. Minimal sets of noncyclic relatively nonexpansive mappings in convex metric spaces. Fixed Point Theory 2015, 16, 313–322. [Google Scholar]
  89. Patle, P.R.; Gabeleh, M.; Rakočević, V. Sadovskii type best proximity point (pair) theorems with an application to fractional differential equations. Mediterr. J. Math. 2022, 19, 141. [Google Scholar] [CrossRef]
  90. Ambrosetti, A. Un teorema di esistenza per le equazioni differenziali negli spazi di Banach. Rend. Semin. Mat. Univ. Padova 1967, 39, 349–361. [Google Scholar]
  91. Hosseinzadeh, H.; Işik, H.; Bonab, S.; George, R. Coupled measure of noncompactness and functional integral equations. Gruyter Open Math. 2022, 20, 38–49. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gabeleh, M.; Malkowsky, E.; Mursaleen, M.; Rakočević, V. A New Survey of Measures of Noncompactness and Their Applications. Axioms 2022, 11, 299. https://doi.org/10.3390/axioms11060299

AMA Style

Gabeleh M, Malkowsky E, Mursaleen M, Rakočević V. A New Survey of Measures of Noncompactness and Their Applications. Axioms. 2022; 11(6):299. https://doi.org/10.3390/axioms11060299

Chicago/Turabian Style

Gabeleh, Moosa, Eberhard Malkowsky, Mohammad Mursaleen, and Vladimir Rakočević. 2022. "A New Survey of Measures of Noncompactness and Their Applications" Axioms 11, no. 6: 299. https://doi.org/10.3390/axioms11060299

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop