Next Article in Journal
Einstein, Barcelona, Symmetry & Cosmology: The Birth of an Equation for the Universe
Next Article in Special Issue
Geraghty–Pata–Suzuki-Type Proximal Contractions and Related Coincidence Best Proximity Point Results
Previous Article in Journal
The Landweber Iterative Regularization Method for Identifying the Unknown Source of Caputo-Fabrizio Time Fractional Diffusion Equation on Spherically Symmetric Domain
Previous Article in Special Issue
Bipolar b-Metric Spaces in Graph Setting and Related Fixed Points
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Best Proximity Point Results for n-Cyclic and Regular-n-Noncyclic Fisher Quasi-Contractions in Metric Spaces

Department of Mathematics, Payame Noor University, Tehran P.O. Box 19395-4697, Iran
*
Authors to whom correspondence should be addressed.
Symmetry 2023, 15(7), 1469; https://doi.org/10.3390/sym15071469
Submission received: 20 June 2023 / Revised: 14 July 2023 / Accepted: 21 July 2023 / Published: 24 July 2023
(This article belongs to the Special Issue New Trends in Fixed Point Theory with Emphasis on Symmetry)

Abstract

:
In this work, we introduce some new concepts such as n-cyclic Fisher quasi-contraction mappings, full-n-noncyclic and regular-n-noncyclic Fisher quasi-contraction mappings in metric spaces. We then generalize the results by Safari-Hafshejani, Amini-Harandi and Fakhar. Meanwhile, we answer the question “under what conditions does a full-n-noncyclic Fisher quasi-contraction mapping have n ( n 1 ) / 2 unique optimal pairs of fixed points?”. Further, to support the main results, we highlight all of the new concepts via non-trivial examples.

1. Introduction

In 1974, Ćirić [1] proved a fixed point (fp) result for a novel class of contractive mappings, which is called quasi-contraction mapping. In fact, he showed that quasi-contraction is a real generalization of some well-known linear contractions, and his result was an expansion of the Banach contraction principle. Then, he proved the existence and uniqueness of fp for T-orbitally single-valued mappings and F-orbitally multi-valued mappings in complete metric spaces.
A self-mapping S on a metric space Y is named a generalized contraction if nonnegative functions exist m ( i , j ) , n ( i , j ) , o ( i , j ) and p ( i , j ) for every i , j Y so that
sup i , j Y { m ( i , j ) + n ( i , j ) + o ( i , j ) + 2 p ( i , j ) } < 1
and
d ( S i , S j ) m ( i , j ) d ( i , j ) + n ( i , j ) d ( i , S i ) + o ( i , j ) d ( j , S j ) + 2 p ( i , j ) [ d ( i , S j ) + d ( j , S i ) ] ,
and is named a quasi-contraction if there is 0 ω < 1 so that
d ( S i , S j ) ω max { d ( i , j ) , d ( i , S i ) , d ( j , S j ) , d ( i , S j ) , d ( j , S i ) }
for all i , j Y [1].
In 1977, Rhoades [2] compared various contractive mappings in metric spaces and showed that Ćirić contractive mapping is one of the most total contractive mappings in metric spaces as it contains many different versions of contractions. Thus, many authors became interested in studying quasi-contractions and extended the Ćirić’s fp results in various aspects. One of these results was introduced by Fisher [3] as follows:
Theorem 1. 
Let ( Y , d ) be a complete metric space and S : Y Y be a continuous mapping. Assume that there exists m , n N and some λ [ 0 , 1 ) provided that
d ( S m i , S n j ) λ max { d ( S α i , S β j ) , d ( S α i , S α i ) , d ( S β j , S β j ) : 0 α , α m   a n d   0 β , β n }
for all i , j Y . Then S has a unique fp.
For more details, see [4,5,6] and references therein.
Although the fp theory is a significant tool for solving fp equations for mappings T defined on a subset A of a metric space ( X , d ) , a non-self mapping T : A B does not necessarily have an fp. Hence, one may attempt to find an element x that is, in some sense, closest to T x . The best approximation theorems and best proximity point (bpp) theorems became famous in this viewpoint. Let ( X , d ) be a metric space, A , B X , d ( A , B ) = inf { d ( x , y ) ; x A , y B } and T : A B be a non-self mapping. The bpp(s) of T is the set of all points x A so that d ( x , T x ) = d ( A , B ) . The main goal of the bpp theory is to provide enough conditions that vouch for the existence of such points. Hence, this theory for various mappings has been considered by many researchers (for example, see [7,8,9,10,11,12]). On the other hand, in 2003, Kirk et al. [13] formulated and defined cyclic mappings as follows:
A mapping T : A B A B is said to be cyclic if T ( A ) B and T ( B ) A . Note that if T ( A ) A and T ( B ) B , then T is called a noncyclic mapping.
In [7], Eldred et al. proved the existence of an optimal pair of fp(s) of noncyclic mappings. After that, Eldred and Veeremani [8] studied the existence of the bpp of cyclic contraction mappings on uniformly convex Banach spaces. Moreover, Suzuki et al. [9] and Espínola et al. [11] established the existence of the bpp for cyclic contraction mappings in metric spaces by applying the properties: unconditionally Cauchy and weakly unconditionally Cauchy, respectively. In fact, the researchers mentioned above have fused cyclical and noncyclic concepts of mappings with the bpp theory to solve some problems in the approximation and optimization theories. Hence, many authors are working on finding the bpp for cyclic and noncyclic mappings in various spaces in [14,15,16,17] and the references therein. Ultimately, Safari-Hafshjani et al. [18] defined a Fisher quasi-contraction and studied the existence of fp(s) and bpp(s) for noncyclic and cyclic Fisher quasi-contraction mappings.
In this work, we define the concepts of n-cyclic Fisher quasi-contraction mappings, as well as full-n-noncyclic and regular-n-noncyclic Fisher quasi-contraction mappings in metric spaces. Next, we generalize the results by Safari-Hafshejan et al. [18] and prove the existence of n ( n 1 ) / 2 , the unique optimal pair of fp(s) for full-n-noncyclic Fisher quasi-contraction mappings.
Let us start with some well-known definitions and notions, which are required in the following sections.
For two nonempty sets, A and B in X, we denote δ [ A , B ] by δ [ A , B ] = sup { d ( x , y ) : x A , y B } . Note that δ [ A , B ] exhibits symmetry.
Definition 1 
([9,14]). Let A and B be two nonempty subsets of a metric space ( X , d ) . Then,
1. 
The pair ( A , B ) has the unconditionally Cauchy (UC) property if for two sequences { x n } and { x n } in A and a sequence { y n } in B, lim n d ( x n , y n ) = lim n d ( x n , y n ) = d ( A , B ) implies lim n d ( x n , x n ) = 0 .
2. 
The pair ( A , B ) has the weakly unconditionally Cauchy (WUC) property if for each { x n } A and ϵ > 0 , there exists y B so that d ( x n , y ) d ( A , B ) + ϵ for n n 0 implies { x n } is Cauchy.
Proposition 1 
([11]). Let A and B be two nonempty subsets of a metric space ( X , d ) provided that A is complete and ( A , B ) has the UC property. Then ( A , B ) has the WUC property.
Definition 2 
([19]). The mapping T on A 1 A 2 A n is called n-cyclic when T ( A 1 ) A 2 , T ( A 2 ) A 3 , …, T ( A n ) A 1 .

2. Results of the n-Cyclic Fisher Quasi-Contraction

Inspired by the findings of the study about cyclic Fisher quasi-contractions [18], we introduce the n-cyclic Fisher quasi-contraction as follows:
Definition 3. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be n-cyclic mapping on A 1 A 2 A n . Point x * A 1 A 2 A n is called the bpp for T if there exist m N and 1 m n provided that
x * A m a n d d ( x * , T x * ) = d ( A m , A m + 1 ) 1 m n 1 d ( A n , A 1 ) m = n .
It is obvious that if d ( A m , A m + 1 ) = 0 or d ( A n , A 1 ) = 0 , then the above problem finds an fp of T.
Remark 1. 
From now on, whenever the term ( A m , A m + 1 ) is observed, it refers to one of the pairs of consecutive sets like ( A 1 , A 2 ) , ( A 2 , A 3 ) , …, ( A n 1 , A n ) and ( A n , A 1 ) .
Notations. Let A 1 , A 2 , , A n be nonempty subsets of a metric space X, p , q N and T be a n-cyclic mapping on A 1 A 2 A n . Then,
A p n , q n x , y = { T i n x , T j n 1 y ; 0 i p , 1 j q } A m , B p n , q n x , y = { T j n + 1 x , T i n y ; 0 j p 1 , 0 i q } A m + 1 ,
for all x A m and y A m + 1 . Moreover, for x A m and r N , consider two sets A r x and B r x as follows:
A r x = { x , T n x , T 2 n x , T 3 n x , , T r n x } A m , B r x = { T x , T n + 1 x , T 2 n + 1 x , T 3 n + 1 x , , T r n + 1 x } A m + 1 .
Definition 4. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be n-cyclic mapping on A 1 A 2 A n . The mapping T is called the n-cyclic Fisher quasi-contraction for some 1 m n if there exist p , q N and 0 c < 1 so that
max { d ( T p n x , T q n y ) , d ( T p n + 1 x , T q n 1 y ) } c δ [ A p n , q n x , y , B p n , q n x , y ] + ( 1 c ) d ( A m , A m + 1 )
for all x A m and y A m + 1 .
Example 1. 
Consider R with the Euclidean metric, A 1 = [ 0 , 2 ] and A 2 = [ 2 , 0 ] . Assume that T : A 1 A 2 A 1 A 2 is defined by
T x = x x A 1 x 2 x A 2 .
For p = q = 1 and c = 1 2 , we have
max { d ( T 2 x , T 2 y ) , d ( T 3 x , T 1 y ) } = | x y | 2 1 2 δ [ { x , T 2 x , T y } , { y , T 2 y , T x } ] + 1 2 d ( A 1 , A 2 ) .
Thus, T is a 2-cyclic Fisher quasi-contraction for every x A 1 and y A 2 . Since
d ( T 2 x , T 2 y ) max { d ( T 2 x , T 2 y ) , d ( T 3 x , T 1 y ) } ,
then
d ( T 2 x , T 2 y ) 1 2 δ [ { x , T 2 x , T y } , { y , T 2 y , T x } ] + 1 2 d ( A 1 , A 2 ) .
This is the same example from Safari et al. [18] for n = 2 .
One of the fundamental steps in the literature on fp(s) is to find a Picard iteration sequence. Here, for x 0 A m and y 0 A m + 1 , we define our iteration sequence as follows:
x i = T i n 2 x 0 A m , i is   even T ( i 1 ) n 2 + 1 x 0 A m + 1 , i   is   odd and y i = T i n 2 y 0 A m + 1 , i   is   even T ( i + 1 ) n 2 1 y 0 A m , i   is   odd .
Lemma 1. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be the n-cyclic Fisher quasi-contraction mapping on A 1 A 2 A n with the quantities p n , q n , and 0 c < 1 . Also, for x 0 A m , consider sequence { x i } to be the same as in (1). Then, δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) for some s , s N , where s n < p n or s n < q n .
Proof. 
For simplicity, assume that q n p n . Since A r x 0 and B r x 0 are finite sets, we have δ [ A r x 0 , B r x 0 ] = sup { d ( α , β ) ; α A r x 0 , β B r x 0 ) } = d ( T s n x 0 , T s n + 1 x 0 ) for some s , s N . On the contrary, suppose that p s r and q s r . Then, by the definition of { x i } , T s n p n x 0 A m and T s n q n + 1 x 0 A m + 1 . Thus, we have
δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) = d ( T p n ( T s n p n x 0 ) , T q n ( T s n q n + 1 x 0 ) ) max { d ( T p n ( T s n p n x 0 ) , T q n ( T s n q n + 1 x 0 ) ) , d ( T p n + 1 ( T s n p n x 0 ) , T q n 1 ( T s n q n + 1 x 0 ) ) } c δ [ A p n , q n T ( s p ) n x 0 , T ( s q ) n + 1 x 0 , B p n , q n T ( s p ) n x 0 , T ( s q ) n + 1 x 0 ] + ( 1 c ) d ( A m , A m + 1 ) ,
which implies that δ [ A r x 0 , B r x 0 ] c δ [ A r x 0 , B r x 0 ] + ( 1 c ) d ( A m , A m + 1 ) and since 0 c < 1 , this is impossible. □
Lemma 2. 
Assume that all the conditions of Lemma 1 are met. Then, for k , k N with k k q p , we have
lim k , k d ( x 2 k , x 2 k + 1 ) = d ( A m , A m + 1 ) .
Proof. 
From Lemma 1, we have δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) for some s , s N , where s n < p n or s n < q n . At the first, we show that δ [ A r x 0 , B r x 0 ] is bounded from above; that is,
δ [ A r x 0 , B r x 0 ] M x 0 c ,
where
M x 0 c = 1 1 c max { d ( T i n + 1 x 0 , T j n + 1 x 0 ) ; 0 i , j p + q + 1 } + d ( A m , A m + 1 )
is the same upper bound. For this, we consider the three following cases.
Case 1: Suppose that s n < p n and s n < q n . Then, we have the following:
( 1 c ) d ( T s n x 0 , T s n + 1 x 0 ) d ( T s n x 0 , T s n + 1 x 0 ) max { d ( T i n + 1 x 0 , T j n + 1 x 0 ) ; 0 i , j p + q + 1 } .
Thus,
d ( T s n x 0 , T s n + 1 x 0 ) 1 1 c max { d ( T s n x 0 , T s n + 1 x 0 ) ; 0 i , j p + q + 1 } 1 1 c max { d ( T s n x 0 , T s n + 1 x 0 ) ; 0 i , j p + q + 1 } + d ( A m , A m + 1 ) ,
which implies that δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) M x 0 c .
Case 2: Suppose that s n < p n and q n s n . Then, T s n q n + 1 A m + 1 and
d ( T s n x 0 , T s n + 1 x 0 ) d ( T s n x 0 , T p n x 0 ) + d ( T p n x 0 , T s n + 1 x 0 ) = d ( T s n x 0 , T p n x 0 ) + d ( T p n x 0 , T q n ( T s n q n + 1 x 0 ) ) d ( T s n x 0 , T p n x 0 ) + ( c δ [ A r x 0 , B r x 0 ] + ( 1 c ) d ( A m , A m + 1 ) ) .
Therefore, ( 1 c ) δ [ A r x 0 , B r x 0 ] d ( T s n x 0 , T p n x 0 ) + ( 1 c ) d ( A m , A m + 1 ) , which concludes that δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) M x 0 c .
Case 3: Suppose that p n s n and s n < q n . Then, T s n p n x 0 A m and
d ( T s n x 0 , T s n + 1 x 0 ) = d ( T p n ( T s n p n x 0 ) , T s n + 1 x 0 ) d ( T p n ( T s n p n x 0 ) , T q n + 1 x 0 ) + d ( T q n + 1 x 0 , T s n + 1 x 0 ) d ( T q n + 1 x 0 , T s n + 1 x 0 ) + ( c δ [ A r x 0 , B r x 0 ] + ( 1 c ) d ( A m , A m + 1 ) ) .
Therefore, ( 1 c ) δ [ A r x 0 , B r x 0 ] d ( T q n + 1 x 0 , T s n + 1 x 0 ) + ( 1 c ) d ( A m , A m + 1 ) , which concludes that δ [ A r x 0 , B r x 0 ] = d ( T s n x 0 , T s n + 1 x 0 ) M x 0 c .
Now, for the optional λ N and k q p , we show that
δ [ A λ x 2 k , B λ x 2 k ] c δ [ A λ + q 2 x 2 k q , B λ + q 2 x 2 k q ] + ( 1 c ) d ( A m , A m + 1 ) ,
in which
A λ + q 2 x 2 k q = { x 2 k q , T n x 2 k q , T 2 n x 2 k q , , T ( λ + q 2 ) n x 2 k q } = { x 2 k q , x 2 k q + 2 , , x 2 k + 2 λ } , B λ + q 2 x 2 k q = { T 1 x 2 k q , T n + 1 x 2 k q , T 2 n + 1 x 2 k q , , T ( λ + q 2 ) n + 1 x 2 k q } = { x 2 k q + 1 , x 2 k q + 3 , , x 2 k + 2 λ + 1 } .
Since k q p , then x 2 k + 2 s 2 p A m and x 2 k + 2 s 2 q + 1 A m + 1 . Also, T is n-cyclic Fisher quasi-contraction mapping. Thus,
δ [ A λ x 2 k , B λ x 2 k ] = d ( T s n x 2 k , T s n + 1 x 2 k ) = d ( T p n x 2 k + 2 s 2 p , T q n x 2 k + 2 s 2 q + 1 ) c δ [ A p n , q n x 2 k + 2 s 2 p , x 2 k + 2 s 2 q + 1 , B p n , q n x 2 k + 2 s 2 p , x 2 k + 2 s 2 q + 1 ] + ( 1 c ) d ( A m , A m + 1 ) = c δ [ { T i n x 2 k + 2 s 2 p , T j n 1 x 2 k + 2 s 2 q + 1 : 0 i p , 1 j q } , { T j n + 1 x 2 k + 2 s 2 p , T i n x 2 k + 2 s 2 q + 1 : 0 j p 1 , 0 i q } ] + ( 1 c ) d ( A m , A m + 1 ) = c δ [ { x 2 k + 2 s 2 p + 2 i , x 2 k + 2 s 2 q + 2 j : 0 i p , 1 j q } , { x 2 k + 2 s 2 p + 2 j + 1 , x 2 k + 2 s 2 q + 2 i + 1 : 0 j p 1 , 0 i q } ] + ( 1 c ) d ( A m , A m + 1 ) ,
which induces that (4) holds.
Also, for k , k N with k k q p , we show that
d ( x 2 k , x 2 k + 1 ) c δ [ A k k + q 2 x 2 k q , B k k + q 2 x 2 k q ] + ( 1 c ) d ( A m , A m + 1 ) ,
in which
A k k + q 2 x 2 k q = { x 2 k q , T n x 2 k q , T 2 n x 2 k q , , T ( k k + q 2 ) n x 2 k q } = { x 2 k q , x 2 k q + 2 , , x 2 k } , B k k + q 2 x 2 k q = { T 1 x 2 k q , T n + 1 x 2 k q , T 2 n + 1 x 2 k q , , T ( k k + q 2 ) n + 1 x 2 k q } = { x 2 k q + 1 , x 2 k q + 3 , , x 2 k + 1 } .
Since k k q p , then x 2 k 2 p A m and x 2 k 2 q + 1 A m + 1 . Also, the mapping T is an n-cyclic Fisher quasi-contraction. Thus, we have
d ( x 2 k , x 2 k + 1 ) = d ( T p n x 2 k 2 p , T q n x 2 k 2 q + 1 ) c δ [ A p n , q n x 2 k 2 p , x 2 k 2 q + 1 , B p n , q n x 2 k 2 p , x 2 k 2 q + 1 ] + ( 1 c ) d ( A m , A m + 1 ) = c δ [ { T i n x 2 k 2 p , T j n 1 x 2 k 2 q + 1 : 0 i p , 1 j q } , { T j n + 1 x 2 k 2 p , T i n x 2 k 2 q + 1 : 0 j p 1 , 0 i q } ] + ( 1 c ) d ( A m , A m + 1 ) = c δ [ { x 2 k 2 p + 2 i , x 2 k 2 q + 2 j : : 0 i p , 1 j q } , { x 2 k 2 p + 2 j + 1 , x 2 k 2 q + 2 i + 1 : 0 j p 1 , 0 i q } ] + ( 1 c ) d ( A m , A m + 1 ) ,
which induces that (5) holds.
Using (4) and (5), we have
d ( x 2 k , x 2 k + 1 ) c ( c δ [ A k k + q x 2 k 2 q , B k k + q x 2 k 2 q ] + ( 1 c ) d ( A m , A m + 1 ) ) + ( 1 c ) d ( A m , A m + 1 ) = c 2 δ [ A k k + q x 2 k 2 q , B k k + q x 2 k 2 q ] + ( 1 c 2 ) d ( A m , A m + 1 ) .
Continuing this procedure, we have
d ( x 2 k , x 2 k + 1 ) c [ 2 k q ] δ [ A k k + q 2 [ 2 k q ] x 2 k [ 2 k q ] q , B k k + q 2 [ 2 k q ] x 2 k [ 2 k q ] q ] + ( 1 c [ 2 k q ] ) d ( A m , A m + 1 ) c [ 2 k q ] δ [ A k x 0 , B k x 0 ] + ( 1 c [ 2 k q ] ) d ( A m , A m + 1 ) .
Using (3), we have
d ( A m , A m + 1 ) d ( x 2 k , x 2 k + 1 ) c [ 2 k q ] M x 0 c + ( 1 c [ 2 k q ] ) d ( A m , A m + 1 ) .
If k in (6), then k and c [ 2 k q ] 0 , and so (2) is established. □
Example 2. 
Consider R with the Euclidean metric, A 1 = [ 0 , 1 ] , A 2 = [ 2 , 3 ) and A 3 = [ 3 , 4 ] . We define T : A 1 A 2 A 3 A 1 A 2 A 3 as follows:
T x = 2 , x A 1 3 , x A 2 x + 4 , x A 3 .
Then, for p = q = 1 , c = 1 2 and for any x A 1 and y A 2 , we have
max { d ( 1 , 2 ) , d ( 2 , 1 ) } = max { d ( T 3 x , T 3 y ) , d ( T 4 x , T 2 y ) } 1 2 δ [ { x , 1 } , { y , 2 } ] + ( 1 1 2 ) d ( A 1 , A 2 ) 1 2 sup { d ( x , y ) , d ( x , 2 ) , d ( 1 , y ) , d ( 1 , 2 ) } + 1 2 .
Hence, the mapping T is a 3-cyclic Fisher quasi-contraction. On the other hand, by a simple calculation, we have { x 2 k } = { 1 } and { x 2 k + 1 } = { 2 } . This shows that
lim k , k d ( x 2 k , x 2 k + 1 ) = d ( A 1 , A 2 ) = 1 .
Lemma 3. 
Consider a metric space ( X , d ) with the subsets A 1 , A 2 , , A n such that A m and A m + 1 have the WUC property. Also, suppose that T is an n-cyclic Fisher quasi-contraction mapping on A 1 A 2 A n . Further, for x 0 A m , consider the sequence { x i } to be the same as in (1). Then, two sequences { x 2 k } = { T k n x 0 } and { x 2 k + 1 } = { T k n + 1 x 0 } are Cauchy.
Proof. 
Using Lemma 2, for any { x 2 k } A m and for every ϵ > 0 , there exists y A m + 1 so that d ( x 2 k , y ) d ( A m , A m + 1 ) + ϵ for n n 0 . Since the pair ( A m , A m + 1 ) has the WUC property, then { x 2 k } is a Cauchy sequence. Analogously, { x 2 k + 1 } is Cauchy. □
Now, we find the bpp for the n-cyclic Fisher quasi-contraction mapping.
Theorem 2. 
Assume that T is the n-cyclic Fisher quasi-contraction mapping on A 1 A 2 A n with the quantities p n , q n , 1 m n and 0 c < 1 , where A 1 , A 2 , , A n are nonempty subsets of a metric space ( X , d ) and A m is complete. If the mapping T is continuous at each point of set
S = { z A m : z = lim k T k n x   for   some   x A m }
and the pair ( A m , A m + 1 ) has the UC property, then
1. 
T has at least one bpp z A m ;
2. 
T n has at most n fp(s).
Proof. 
Using Lemma 3, for any x 0 A m , { x 2 k } = { T k n x 0 } is a Cauchy sequence in A m . Since A m is complete, we have lim k T k n x 0 = z for some z A m . Thus, z S . Now, by the continuity of T and d, and using Lemma 2, we have
d ( z , T z ) = d ( lim k T k n x 0 , T ( lim k T k n x 0 ) ) = d ( lim k T k n x 0 , lim k T k n + 1 x 0 ) = lim k d ( x 2 k , x 2 k + 1 ) = d ( A m , A m + 1 ) .
This displays that z A m is a bpp of T.
Now, we establish T n has at most n fp(s). Suppose that { x 2 k } = { T k n x 0 } , { x k } = { T ( k + 1 ) n x 0 } and { y k } = { T k n + 1 x 0 } . Then, by Lemma 2, we gain
lim k d ( x 2 k , y k ) = lim k d ( x k , y k ) = d ( A m , A m + 1 ) .
Since the pair ( A m , A m + 1 ) has the UC property, then lim k d ( x 2 k , x k ) = 0 . Using the continuity of d, we have
0 = lim k d ( x 2 k , x k ) = d ( lim k x 2 k , lim k x k ) = d ( z , T n ( lim k T k n x 0 ) ) = d ( z , T n z ) ,
which induces that T n z = z ; that is, T n has an fp. Now, we prove this fp is unique. Similar to the above argument, for each x A m , assume that z A m provided that
lim k T k n x = z , d ( z , T z ) = d ( A m , A m + 1 ) and T n z = z .
Now, without the loss of generality, consider d ( z , T z ) d ( z , T z ) . Then, we have
d ( z , T z ) = d ( T p n z , T ( T q n z ) ) = d ( T p n z , T q n ( T z ) ) max { d ( T p n z , T q n ( T z ) ) , d ( T p n + 1 z , T q n 1 ( T z ) ) } c δ [ A p n , q n z , T z , B p n , q n z , T z ] + ( 1 c ) d ( A m , A m + 1 ) = c δ [ { z , z } , { T z , T z } ] + ( 1 c ) d ( A m , A m + 1 ) = c sup { d ( z , T z ) , d ( z , T z ) , d ( z , T z ) , d ( z , T z ) } + ( 1 c ) d ( A m , A m + 1 ) c d ( z , T z ) + ( 1 c ) d ( A m , A m + 1 ) ,
which implies that d ( z , T z ) = d ( A m , A m + 1 ) . Thus, d ( z , T z ) = d ( z , T z ) = d ( A m , A m + 1 ) . Let z = lim k T k n x and z = lim k T k n x 0 . Then, we have
lim k d ( T k n x 0 , T z ) = lim k d ( T k n x , T z ) = d ( A m , A m + 1 ) .
Since the pair ( A m , A m + 1 ) has the UC property, we have z = z . Therefore, for each x A m , { T k n x } converges to z. Since n ordered pairs exist ( A 1 , A 2 ) , ( A 2 , A 3 ) , …, ( A n , A 1 ) , then T n has at most n fp(s). □
Corollary 1. 
Assume that all the conditions of Theorem 2 are met. Further, suppose that
max { d ( T n u , T v ) , d ( T u , T n v ) } c min { d ( T u , v ) , d ( u , T v ) } + ( 1 c ) d ( A m , A m + 1 )
for all u , v A m . Then, T has a unique bpp z A m .
Proof. 
Let z 1 and z 2 be the bpp(s) of the mapping T. Then, z 1 and z 2 are the fp(s) of the mapping T n . Now, without loss of generality, assume d ( T z 1 , z 2 ) d ( z 1 , T z 2 ) . Then,
d ( z 1 , T z 2 ) = d ( T n z 1 , T z 2 ) c d ( T z 1 , z 2 ) + ( 1 c ) d ( A m , A m + 1 ) c d ( z 1 , T z 2 ) + ( 1 c ) d ( A m , A m + 1 ) .
Hence, d ( z 1 , T z 2 ) = d ( z 2 , T z 2 ) = d ( A m , A m + 1 ) . Since ( A m , A m + 1 ) has the UC property, then z 1 = z 2 . □
Example 3. 
Consider d, A 1 , A 2 , A 3 , and T to be the same as in Example 2. Clearly, T has no fp(s). Here, we find the bpp of the mapping T and the fp of the mapping T 3 .
By the definition of the bpp: If x * A 1 is the bpp, then d ( x * , T x * ) = d ( A 1 , A 2 ) . Thus, d ( x * , 2 ) = 1 and x * = 1 A 1 (note that x * = 3 A 1 ). If x * A 2 is the bpp, then d ( x * , T x * ) = d ( A 2 , A 3 ) . Thus, d ( x * , 3 ) = 0 , which induces that x * = 3 A 2 . Similarly, if x * A 3 is the bpp, then x * = 3 A 3 .
Note that all the assumptions of Theorem 2 are held. Thus, we can check the validity of the assertion of this theorem.
1. 
By using ( A 1 , A 2 ) : For z = 1 in A 1 , { x 2 k } = { T 3 k x 0 } = { 1 } . Thus, z = lim k T 3 k x 0 = 1 is the bpp of T on A 1 . Also, T 3 1 = 1 and z = 1 are unique fps of T 3 .
2. 
By using ( A 2 , A 3 ) : A 2 is not complete. We cannot apply Theorem 2 to this case.
3. 
By using ( A 3 , A 1 ) : For z = 3 in A 3 , { x 2 k } = { T 3 k x 0 } = { 3 } . Thus, z = lim k T 3 k x 0 = 3 is the bpp of T on A 3 . Also, T 3 3 = 3 and z = 3 are unique fps of T 3 .
Consequently, z = 1 and z = 3 are bpp(s) for the mapping T. Also, we have T 3 1 = 1 , T 3 2 = 2 , and T 3 3 = 3 . Thus, T 3 has three fp(s).

3. Regular-n-Noncyclic and Full-n-Noncyclic Fisher Quasi-Contractions

Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) . A self-mapping T on A 1 A 2 A n is called noncyclic if T ( A i ) A i for 1 i n . Also, the pair ( x i , x j ) A i × A j for 0 i , j n with i j is denoted as an optimal pair of fp(s) of the noncyclic mapping T if T x i = x i , T x j = x j , and d ( x i , x j ) = d ( A i , A j ) .
It is obvious that if x 0 A i , y 0 A j and T is the noncyclic mapping, then x k + 1 = T x k A i and y k + 1 = T y k A j for k 0 .
Notations. Let m N { 0 } . Define the set C m u by C m u = { u , T u , , T m u } for u A 1 A 2 A n . Clearly, if u A i for i = 1 , 2 , , n , then C m u A i .
Now, we define the notion of regular-n-noncyclic and full-n-noncyclic Fisher quasi-contractions in metric spaces. Then, we obtain the main outcomes of this part.
Definition 5. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be a noncyclic mapping on A 1 A 2 A n . Then, T is said to be
1. 
A regular-n-noncyclic Fisher quasi-contraction if there exist two sets A i and A j for 1 i , j n with i j and some p i , p j N so that
d ( T p i x , T p j y ) c δ [ C p i x , C p j y ] + ( 1 c ) d ( A i , A j )
for each x A i and y A j , where 0 c < 1 ;
2. 
A full-n-noncyclic Fisher quasi-contraction if for all A i and A j , where 1 i , j n with i j , there exist some p i , p j N so that
d ( T p i x , T p j y ) c δ [ C p i x , C p j y ] + ( 1 c ) d ( A i , A j )
for each x A i and y A j , where 0 c < 1 .
Lemma 4. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be a regular-n-noncyclic Fisher quasi-contraction mapping on A 1 A 2 A n . Then,
δ [ C k x 0 , C l y 0 ] M x 0 , y 0
for each k , l N , where
M x 0 , y 0 = 1 1 c max { d ( T i x 0 , T j y 0 ) , d ( T i x 0 , T j x 0 ) , d ( T i y 0 , T j y 0 ) ; 0 i , j max { p i , p j } } + d ( A i , A j ) .
Proof. 
Since the mapping T is a regular-n-noncyclic Fisher quasi-contraction, there exist two sets A i and A j for 1 i , j n with i j and some p i , p j N such that
d ( T p i x , T p j y ) c δ [ C p i x , C p j y ] + ( 1 c ) d ( A i , A j ) .
First, we show that
δ [ C k x 0 , C l y 0 ] = d ( T r x 0 , T r y 0 ) where r < p i or r < p j .
On the contrary, suppose that δ [ C k x 0 , C l y 0 ] = d ( T u x 0 , T v y 0 ) , where p i u k and p j v l . Then, u p i 0 and v p j 0 , and x u p i = T u p i x 0 and y v p j = T v p j y 0 , respectively. It follows from (8) that
δ [ C k x 0 , C l y 0 ] = d ( T u x 0 , T v y 0 ) = d ( T p i T u p i x 0 , T p j T v p j y 0 ) c δ [ C p i x u p i , C p j y v p j ] + ( 1 c ) d ( A i , A j ) c δ [ C k x 0 , C l y 0 ] + ( 1 c ) d ( A i , A j ) ,
which, by c [ 0 , 1 ) , implies that δ [ C k x 0 , C l y 0 ] d ( A i , A j ) . Consequently, δ [ C k x 0 , C l y 0 ] = d ( A i , A j ) and (9) holds.
Now, we prove (7) by applying (9) and consider the three following cases.
Case 1: Suppose that r < p i and r < p j . Then,
( 1 c ) d ( T r x 0 , T r y 0 ) d ( T r x 0 , T r y 0 ) { d ( T i x 0 , T j y 0 ) , d ( T i x 0 , T j x 0 ) , d ( T i y 0 , T j y 0 ) ; 0 i , j max { p i , p j } } + ( 1 c ) d ( A i , A j ) ,
which concludes that d ( T r x 0 , T r y 0 ) M x 0 , y 0 . Thus, (7) holds.
Case 2: Assume that 0 r < p i and p j r l . Then,
δ [ C k x 0 , C l y 0 ] = d ( T r x 0 , T r y 0 ) d ( T r x 0 , T p i x 0 ) + d ( T p i x 0 , T r y 0 ) d ( T r x 0 , T p i x 0 ) + d ( T p i x 0 , T p j T r p j y 0 ) d ( T r x 0 , T p i x 0 ) + ( c δ [ C k x 0 , C l y 0 ] + ( 1 c ) d ( A i , A j ) ) ,
which implies that ( 1 c ) δ [ C k x 0 , C l y 0 ] d ( T r x 0 , T p i x 0 ) + ( 1 c ) d ( A i , A j ) . Thus, (7) holds.
Case 3: Similarly, if p i r k and r < p j , then (7) holds. □
Lemma 5. 
Assume that all the conditions of Lemma 4 are met. Further, suppose that x t + 1 = T x t and y t + 1 = T y t for t , t N { 0 } . Then,
lim t , t d ( x t , y t ) = d ( A i , A j ) .
Proof. 
Since t , t , without loss of generality, we can suppose that t , t max { p i , p j } . Hence, t p i 0 and t p j 0 . On the other hand, δ [ C k x t , C l y t ] = s u p { d ( x , y ) : x C k x t , y C l y t } . Thus, for 0 r k and 0 s l , we obtain
δ [ C k x t , C l y t ] = d ( T r x t , T s y t ) = d ( T p i + r x t p i , T p j + s y t p j ) = d ( T p i T r x t p i , T p j T s y t p j ) c δ [ C k + p i x t p i , C l + p j y t p j ] + ( 1 c ) d ( A i , A j ) .
Consequently,
δ [ C k x t , C l y t ] c δ [ C k + p i x t p i , C l + p j y t p j ] + ( 1 c ) d ( A i , A j ) .
Now, by using (11) and by putting k = l = 0 , we have
d ( x t , y t ) = δ [ C 0 x t , C 0 y t ] c δ [ C p i x t p i , C p j y t p j ] + ( 1 c ) d ( A i , A j ) c ( c δ [ C 2 p i x t 2 p i , C 2 p j y t 2 p j ] + ( 1 c ) d ( A i , A j ) ) + ( 1 c ) d ( A i , A j ) = c 2 δ [ C 2 p i x t 2 p i , C 2 p j y t 2 p j ] + ( 1 c 2 ) d ( A i , A j ) c 2 ( c δ [ C 3 p i x t 3 p i , C 3 p j y t 3 p j ] + ( 1 c ) d ( A i , A j ) ) + ( 1 c 2 ) d ( A i , A j ) = c 3 δ [ C 3 p i x t 3 p i , C 3 p j y t 3 p j ] + ( 1 c 3 ) d ( A i , A j )
for t , t max { 2 p i , 2 p j } .
Continuing this process, using Lemma 4 and setting α ( t , t ) = min { [ t p i ] , [ t p j ] } , we have
d ( A i , A j ) d ( x t , y t ) c α ( t , t ) δ [ C α ( t , t ) . p i x t α ( t , t ) . p i , C α ( t , t ) . p j y t α ( t , t ) . p j ] + ( 1 c α ( t , t ) ) d ( A i , A j ) c α ( t , t ) δ [ C t x 0 , C t y 0 ] + ( 1 c α ( t , t ) ) d ( A i , A j ) c α ( t , t ) M x 0 , y 0 + ( 1 c α ( t , t ) ) d ( A i , A j ) .
Now, by taking the limit as t , t , (10) is established. □
Lemma 6. 
Let A 1 , A 2 , , A n be nonempty subsets of a metric space ( X , d ) and T be a regular-n-noncyclic Fisher quasi-contraction mapping on A 1 A 2 A n . Further, suppose that ( A i , A j ) has the WUC property, and for x 0 A i , consider x t + 1 = T x t for any t 0 . Then, the sequence { x t } is Cauchy.
Proof. 
By Lemma 5, { x t } is Cauchy. □
Theorem 3. 
Let A 1 , A 2 , , A n be subsets of ( X , d ) and T be a regular-n-noncyclic Fisher quasi-contraction mapping on A 1 A n . Also, let A i and A j for 1 i , j n be complete subsets of X such that ( A i , A j ) and ( A j , A i ) have the UC property. Further, assume that T : A i A i and T : A j A j are continuous. Then, T has a unique optimal pair of fp(s) ( x i * , y i * ) provided that { T n x 0 } and { T n y 0 } converge to x i * and y j * for each x 0 A i and y 0 A j , respectively.
Proof. 
From Lemma 6, { T n x 0 } is Cauchy. Since A i is complete, { T n x 0 } converges to a certain x i * A i . Since T is the continuous mapping on A i , we deduce that T x i * = x i * ; that is, x i * A i is an fp of T. For uniqueness, assume that x i * * A i is another fp of T. Also, let y 0 A j . Using Lemma 5, we have
lim n d ( x i * , T n y 0 ) = lim n d ( T n x i * , T n y 0 ) = d ( A i , A j ) = lim n d ( T n x i * * , T n y 0 ) = lim n d ( x i * * , T n y 0 ) .
Since the pair ( A i , A j ) has the UC property, then x i * = x i * * . Similarly, T has a unique fp y j * A j such that { T n y 0 } converges to a certain y i * A i . Also, by Lemma 5, we have
d ( x i * , y j * ) = lim n d ( T n x i * , T n y j * ) = d ( A i , A j ) .
Hence, ( x i * , y j * ) A i × A j is a unique optimal pair of fp of T. □
As an application, in the following corollary, we show that a full-n-noncyclic Fisher quasi-contraction mapping has n ( n 1 ) / 2 unique optimal pairs of fixed points.
Corollary 2. 
Let A 1 , A 2 , , A n be nonempty and complete subsets of a metric space ( X , d ) and T be a full-n-noncyclic Fisher quasi-contraction mapping on A 1 A 2 A n . Also, assume that the pairs ( A i , A j ) and ( A j , A i ) have the UC property for each 1 i , j n with i j . Further, suppose that the mapping T : A i A i is continuous. Then, T has n ( n 1 ) 2 unique optimal pairs of fp(s) ( x i * , y i * ) provided that { T n x 0 } and { T n y 0 } converge to x i * and y j * for each x 0 A i and y 0 A j , respectively.
Proof. 
By Theorem 3, ( x i * , y j * ) is a unique optimal pair of fp of T for some i , j = 1 , , n with i j . Since there exist n ( n 1 ) 2 cases of the different pairs ( A i , A j ) for any i , j , then the assertion holds. □
Example 4. 
Consider R 2 with the metric d ( ( α , β ) , ( γ , δ ) ) = ( α γ ) 2 + ( β δ ) 2 for each ( α , β ) , ( γ , δ ) R 2 . Also, assume that A 1 = { ( a , 0 ) : 1 a 2 } , A 2 = { ( 0 , b ) : 1 b 2 } and A 3 = { ( a , 0 ) : 2 a 1 } are three subsets of R 2 . Moreover, suppose that T : A 1 A 2 A 3 A 1 A 2 A 3 is defined as follows:
T x = ( 1 , 0 ) , x A 1 ( 0 , 1 ) , x A 2 ( 2 , 0 ) , x A 3 .
Clearly, T is a noncyclic mapping. For the pair ( A 1 , A 2 ) , let p 1 = p 2 = 1 and c be a fixed number in [ 0 , 1 ) . For each x A 1 and y A 2 , we have δ [ C 1 x , C 1 y ] 2 . Thus, we conclude that
d ( T 1 x , T 1 y ) = d ( ( 1 , 0 ) , ( 0 , 1 ) ) = 2 c δ [ C 1 x , C 1 y ] + 2 ( 1 c ) ;
that is, the mapping T is a regular-3-noncyclic Fisher quasi-contraction. Further, since
lim t , t d ( x t , y t ) = d ( ( 1 , 0 ) , ( 1 , 0 ) ) = d ( A 1 , A 2 ) = 2 ,
the assertion of Lemma 5 holds. Furthermore,
T ( 1 , 0 ) = ( 1 , 0 ) , T ( 0 , 1 ) = ( 0 , 1 ) a n d d ( ( 1 , 0 ) , ( 0 , 1 ) ) = d ( A 1 , A 2 ) .
Thus, ((1,0),(0,1)) is an optimal pair of fp of T. Moreover, look at the pair ( A 1 , A 3 ) . For every p 1 , p 2 N and x = ( 1 , 0 ) A 1 and y = ( 1 , 0 ) A 3 , we obtain
d ( T p 1 x , T p 2 y ) = d ( ( 1 , 0 ) , ( 2 , 0 ) ) = 3 c δ [ C p 1 x , C p 2 y ] + 2 ( 1 c ) .
Note that δ [ C p 1 x , C p 2 y ] = δ [ C p 1 ( 1 , 0 ) , C p 2 ( 1 , 0 ) ] = δ [ d ( ( 1 , 0 ) , ( 1 , 0 ) ) , d ( ( 1 , 0 ) , ( 2 , 0 ) ) ] = 3 . So, T is not the full-3-noncyclic Fisher quasi-contraction mapping. In addition,
lim t , t d ( x t , y t ) = d ( ( 1 , 0 ) , ( 2 , 0 ) ) = 3 2 = d ( A 1 , A 3 ) .
Example 5. 
Let ( R 2 , d ) be the same metric space in Example 4. Also, suppose that A 1 = { ( a , 0 ) : 1 a 2 } , A 2 = { ( a , 0 ) : 2 a 1 } , A 3 = { ( 0 , b ) : 1 b 2 } and A 4 = { ( 0 , b ) : 2 b 1 } are four arbitrary subsets of R 2 . We define a mapping T on A 1 A 2 A 3 A 4 as follows:
T x = ( 1 , 0 ) , x A 1 ( 1 , 0 ) , x A 2 ( 0 , 1 ) , x A 3 ( 0 , 1 ) , x A 4 .
For p = q = 1 and for each x A 1 and y A 2 , we have
C 1 x = { x , ( 1 , 0 ) } , C 1 y = { y , ( 1 , 0 ) } , δ [ C 1 x , C 1 y ] = 4 a n d d ( A 1 , A 2 ) = 2 .
Thus, there exists a 0 c < 1 so that
d ( T x , T y ) = d ( ( 1 , 0 ) , ( 1 , 0 ) ) = 2 c δ [ C 1 x , C 1 y ] + ( 1 c ) d ( A 1 , A 2 ) .
Also,
lim t , t d ( x t , y t ) = d ( ( 1 , 0 ) , ( 1 , 0 ) ) = d ( A 1 , A 2 ) = 2 .
Similarly, we can use the above discussion for all 1 i , j 4 with i j . Hence, mapping T is a full-4-noncyclic Fisher quasi-contraction. In addition, T has six unique optimal pairs of fp(s) as follows:
( ( 1 , 0 ) , ( 0 , 1 ) ) , ( ( 1 , 0 ) , ( 1 , 0 ) ) , ( ( 1 , 0 ) , ( 0 , 1 ) ) ( ( 0 , 1 ) , ( 1 , 0 ) ) , ( ( 1 , 0 ) , ( 0 , 1 ) ) , ( ( 0 , 1 ) , ( 0 , 1 ) ) .

4. Conclusions

In the present paper, we introduced the concepts of n-cyclic Fisher quasi-contraction mappings, as well as full-n-noncyclic and regular-n-noncyclic Fisher quasi-contraction mappings in metric spaces. Then we stated and proved several bpp theorems regarding these contractions. Moreover, we solved an open problem about the number of optimal pairs of fp(s) for full-n-noncyclic Fisher quasi-contraction mappings. Due to the generalization of this paper, unlike the other articles, we found more than one bpp. In future studies, readers may concentrate on specific aspects of these points. For example, they could explore the optimum of a bpp or discuss the unique conditions of these points. Also, they may obtain similar results in various metric spaces.

Author Contributions

Conceptualization, K.F., M.A. and G.S.R.; methodology, K.F., M.A. and G.S.R.; writing—original draft preparation, K.F., M.A. and G.S.R.; writing—review and editing, K.F., M.A. and G.S.R. All authors contributed equally and significantly to the writing of this paper. All authors read and approved the final manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to the academic editor and anonymous referees for their accurate reading and their helpful suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ćirić, L.B. A generalization of Banach’s contraction principle. Proc. Amer. Math. Soc. 1974, 45, 267–273. [Google Scholar] [CrossRef] [Green Version]
  2. Rhoades, B.E. A comparison of various definitions of contractive mappings. Trans. Amer. Math. Soc. 1977, 226, 257–290. [Google Scholar] [CrossRef]
  3. Fisher, B. Quasicontractions on metric spaces. Proc. Am. Math. Soc. 1979, 75, 321–325. [Google Scholar] [CrossRef] [Green Version]
  4. Alqahtani, B.; Fulga, A.; Karapınar, E. Sehgal type contractions on b-metric space. Symmetry 2018, 10, 560. [Google Scholar] [CrossRef] [Green Version]
  5. Gajić, L.; Ilić, D.; Rakočević, V. On Ćirić maps with a generalized contractive iterate at a point and Fisher’s quasi-contractions in cone metric spaces. Appl. Math. Comput. 2010, 216, 2240–2247. [Google Scholar] [CrossRef]
  6. Yousefi, F.; Rahimi, H.; Soleimani Rad, G. A discussion on fixed point theorems of Ćirić and Fisher on w-distance. J. Nonlinear Convex Anal. 2022, 23, 1409–1418. [Google Scholar]
  7. Eldred, A.A.; Kirk, W.A.; Veeramani, P. Proximal normal structure and relatively nonexpansive mappings. Studia Math. 2005, 171, 283–293. [Google Scholar] [CrossRef] [Green Version]
  8. Eldred, A.A.; Veeramani, P. Existence and convergence of best proximity points. J. Math. Anal. Appl. 2006, 323, 1001–1006. [Google Scholar] [CrossRef] [Green Version]
  9. Suzuki, T.; Kikkawa, M.; Vetro, C. The existence of the best proximity points in metric spaces with the UC property. Nonlinear Anal. 2009, 71, 2918–2926. [Google Scholar] [CrossRef]
  10. Sadiq Basha, S. Best proximity point theorems generalizing the contraction principle. Nonlinear Anal. 2011, 74, 5844–5850. [Google Scholar] [CrossRef]
  11. Espínola, R.; Fernández-León, A. On best proximity points in metric and Banach space. Can. J. Math. 2011, 63, 533–550. [Google Scholar] [CrossRef] [Green Version]
  12. Fallahi, K.; Fulga, A.; Soleimani Rad, G. Best proximity points for (φψ)-weak contractions and some applications. Filomat 2023, 37, 1835–1842. [Google Scholar]
  13. Kirk, W.A.; Srinivasan, P.S.; Veeramani, P. Fixed points for mappings satisfying cyclical contractive conditions. Fixed Point Theory. 2003, 4, 79–89. [Google Scholar]
  14. Espínola, R.; Gabeleh, M. On the structure of minimal sets of relatively nonexpansive mappings. Numer. Func. Anal. Optim. 2013, 34, 845–860. [Google Scholar] [CrossRef]
  15. Karapınar, E.; Karpagam, S.; Magadevan, P.; Zlatanov, B. On Ω class of mappings in a p-cyclic complete metric space. Symmetry 2019, 11, 534. [Google Scholar] [CrossRef] [Green Version]
  16. De la Sen, M.; Ibeas, A. Generalized cyclic p-contractions and p-contraction pairs some properties of asymptotic regularity best proximity points, fixed foints. Symmetry 2022, 14, 2247. [Google Scholar] [CrossRef]
  17. Aydi, H.; Lakzian, H.; Mitrović, Z.D.; Radenović, S. Best proximity points of MT-cyclic contractions with UC property. Numer. Func. Anal. Optim. 2020, 41, 871–882. [Google Scholar] [CrossRef]
  18. Safari-Hafshejani, A.; Amini-Harandi, A.; Fakhar, M. Best proximity points and fixed points results for noncyclic and cyclic Fisher quasi-contractions. Numer. Func. Anal. Optim. 2019, 40, 603–619. [Google Scholar] [CrossRef]
  19. Nashine, H.K.; Romaguera, S. Fixed point theorems for cyclic self-maps involving weaker Meir-Keeler functions in complete metric spaces and applications. Fixed Point Theory Appl. 2013, 2013, 224. [Google Scholar] [CrossRef] [Green Version]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fallahi, K.; Ayobian, M.; Soleimani Rad, G. Best Proximity Point Results for n-Cyclic and Regular-n-Noncyclic Fisher Quasi-Contractions in Metric Spaces. Symmetry 2023, 15, 1469. https://doi.org/10.3390/sym15071469

AMA Style

Fallahi K, Ayobian M, Soleimani Rad G. Best Proximity Point Results for n-Cyclic and Regular-n-Noncyclic Fisher Quasi-Contractions in Metric Spaces. Symmetry. 2023; 15(7):1469. https://doi.org/10.3390/sym15071469

Chicago/Turabian Style

Fallahi, Kamal, Morteza Ayobian, and Ghasem Soleimani Rad. 2023. "Best Proximity Point Results for n-Cyclic and Regular-n-Noncyclic Fisher Quasi-Contractions in Metric Spaces" Symmetry 15, no. 7: 1469. https://doi.org/10.3390/sym15071469

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