Next Article in Journal
Exploring the Mutual Influence Relationships of International Airport Resilience Factors from the Perspective of Aviation Safety: Using Fermatean Fuzzy DEMATEL Approach
Next Article in Special Issue
On Local Unique Solvability for a Class of Nonlinear Identification Problems
Previous Article in Journal
Probabilistic Interval Ordering Prioritized Averaging Operator and Its Application in Bank Investment Decision Making
Previous Article in Special Issue
On the Study of Pseudo 𝒮-Asymptotically Periodic Mild Solutions for a Class of Neutral Fractional Delayed Evolution Equations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Common Fixed Point of (ψ, β, L)-Generalized Contractive Mapping in Partially Ordered b-Metric Spaces

1
School of Mathematics and Statistics, Hubei Normal University, Huangshi 435002, China
2
School of Mathematics and Statistics, Chongqing Three Gorges University, Chongqing 404020, China
3
Faculty of Mechanical Engineering, University of Belgrade, Kraljice Marije 16, 11120 Beograd, Serbia
*
Author to whom correspondence should be addressed.
Axioms 2023, 12(11), 1008; https://doi.org/10.3390/axioms12111008
Submission received: 23 September 2023 / Revised: 21 October 2023 / Accepted: 23 October 2023 / Published: 26 October 2023

Abstract

:
The purpose of this paper is to attain the existence of coincidences and common fixed points in four mappings satisfying ( ψ , β , L ) -generalized contractive conditions in the framework of partially ordered b-metric spaces. The main results presented in this paper generalize some recent results in the existing literature. Furthermore, a nontrivial example is presented to support the obtained results.

1. Introduction

It is world-renowned that the Banach contraction principle (for short, BCP) (see [1]) occupies a significant role in different fields of basic mathematics, applied mathematics and other subjects, and it has been generalized and improved in many aspects. Numerous generalizations and improvements come forth by changing metric spaces into general abstract metric spaces (see [2,3]). In [4], Vulpe et al. introduced b-metric space (sometimes so-called metric-type space, see [5]) as a new generalization of usual metric space. He provided the generalized BCP in b-metric space. From then on, a large number of papers have considered fixed point theory and its applications or variational methods for single-valued and multi-valued operators in b-metric spaces (the reader may see [5,6,7,8,9,10] and the related references therein). The mappings satisfying certain contractive conditions can be utilized to establish the existence of solutions to all kinds of operator equations such as integral equations, differential equations and fractional differential equations. Beg and Abbas (see [11]) obtained common fixed point theorems by extending a weakly contractive condition into two mappings. Abbas et al. (see [12]) investigated common fixed points for four mappings satisfying generalized weakly contractive conditions in complete partially ordered metric spaces. Esmaicy et al. (see [13]) initiated coincidence point results for four mappings in partially ordered metric spaces and used their results to seek the common solution of two integral equations. Recently, Abbas et al. (see [14]) acquired coincidence and common fixed points for four mappings satisfying generalized ( ψ , β ) -contractive conditions in complete partially ordered metric spaces.
Based on the previous work, throughout this paper, we aim to established coincidence and common fixed points for four mappings under generalized ( ψ , β , L ) -contractive conditions in complete partially ordered b-metric spaces. Our results make great progress in extending, unifying and generalizing the corresponding results in [14,15,16].

2. Preliminaries

In this paper, unless there are special statements, we always denote R , R + , N and N * as the set of all real numbers, the set of all non-negative real numbers, the set of all non-negative integers and the set of all positive integers, respectively.
First, we give some basic definitions and results which will be needed in what follows.
Definition 1
([17]). Let X be a nonempty set and d : X × X R + a mapping satisfying
(1) 
d ( ξ , η ) = 0 if and only if ξ = η ;
(2) 
d ( ξ , η ) = d ( η , ξ ) for all ξ , η X ;
(3) 
d ( ξ , η ) s [ d ( ξ , ζ ) + d ( ζ , η ) ] for all ξ , η , ζ X ,
where s 1 is a given real number, d is then said to be a b-metric on X, and ( X , d ) is said to be a b-metric space. If ( X , ) is still a partially ordered set, then ( X , , d ) is said to be a partially ordered b-metric space.
Otherwise, for more notions such as b-convergence, b-completeness, b-Cauchy sequence and b-continuity in b-metric spaces, the reader may refer to [1,4,5,6,7,8,9,10,11,13,15,16,17,18,19,20,21,22,23,24,25,26,27,28] and the references mentioned therein.
In general, b-metric is not continuous; kindly see the following examples.
Example 1
([16]). Let X = N * { } ; define a mapping d : X × X R + using
d ( ξ , η ) = 0 , when ξ = η ; | 1 ξ 1 η | , when one of ξ , η is even and another is distinctly even or infinity ; 5 , when one of ξ , η is odd and another is distinctly odd or infinity ; 2 , otherwise .
It is easy to see that when
d ( ξ , ζ ) 5 2 [ d ( ξ , η ) + d ( η , ζ ) ] ( ξ , η , ζ X ) ,
then ( X , d ) is a b-metric space with coefficient s = 5 2 . Put ξ n = 2 n ( n N ) , then
d ( ξ n , ) = 1 2 n 0 ( n ) ,
so ξ n ( n ), but d ( ξ n , 1 ) = 2 5 = d ( , 1 ) ( n ) . That is to say, the b-metric is not continuous.
Example 2
([6]). Let X = R and α > 1 be a constant. Define a mapping d : X × X R + using
d ( ξ , η ) = | ξ η | , ξ η 0 , α | ξ η | , ξ η = 0 , for all ξ , η X .
Then, ( X , d ) is a b-metric space with coefficient s = α , but the b-metric d is not continuous.
Definition 2
([20,27,28]). Let ( X , , d ) be a partially ordered b-metric space and f , g , h be self-mappings on X such that f ( X ) g ( X ) h ( X ) .
(1) 
If ξ , η X , ξ η or η ξ holds, then the elements ξ , η are called comparable;
(2) 
If f ξ g f ξ for all ξ X , then the pair ( f , g ) is called partially weakly increasing;
(3) 
If f ξ g η for all η h 1 ( f ξ ) , then the pair ( f , g ) is called partially weakly increasing with respect to h;
(4) 
If lim n d ( f g ξ n , g f ξ n ) = 0 , whenever { ξ n } is a sequence in X such that lim n f ξ n = lim n g ξ n = t for some t X , then the pair ( f , g ) is called compatible;
(5) 
If w = f ξ = g ξ for some ξ in X, then ξ is called a coincidence point of f and g, and w is called a point of coincidence of f and g.
(6) 
If f and g commute at their coincidence points, i.e., f g ξ = g f ξ , where f ξ = g ξ , then the pair ( f , g ) is called weakly compatible;
(7) 
If ξ f ξ for each ξ X , then f is called dominating. If f ξ ξ for each ξ X , then f is called dominated.
Example 3
([14]). Let f : X X be a mapping defined by f ξ = ξ 1 4 , where X = [ 0 , 1 ] is endowed with usual ordering. Clearly, ξ ξ 1 4 = f ξ for all ξ X ; thus, f is a dominating map.
Example 4
([14]). Let f : X X be a mapping defined by f ξ = 1 ξ + 1 , where X = [ 1 , + ) is endowed with usual ordering. Clearly, f ξ = 1 ξ + 1 ξ for all ξ X ; hence, f is a dominated map.
An assertion similar to the following lemma was used (and proved) in the course of proofs of several fixed point results in various articles.
Lemma 1
([26]). Every sequence { ξ n } n N from a b-metric space ( X , d ) with the property that there exists γ [ 0 , 1 ) such that
d ( ξ n , ξ n + 1 ) γ d ( ξ n 1 , ξ n )
for every n N is b-Cauchy.
Lemma 2.
Let ( X , d ) be a b-metric space with s 1 and { η n } a sequence in X such that lim n d ( η n + 1 , η n ) = 0 . If { η 2 n } or { η 2 n 1 } is a b-Cauchy sequence, then { η n } is a b-Cauchy sequence in X.
Proof. 
We only prove the case that { η 2 n } is a b-Cauchy sequence. The another case can be proved similarly.
In view of lim n d ( η n + 1 , η n ) = 0 , then for every ε > 0 , there is a natural number N 1 such that, for all n N 1 ,
d ( η n + 1 , η n ) < ε 3 s 2 .
Since { η 2 n } is a b-Cauchy sequence, then for the above ε > 0 there is a natural number N 2 such that, for all n N 2 and any p N , one has
d ( η 2 n , η 2 n + 2 p ) < ε 3 s 2 .
Now, let N = max { N 1 , N 2 } . We shall claim that, for all n > N , it satisfies d ( η n , η n + p ) < ε . We complete the proof with four cases.
(c1)
If n and p are even numbers, then, by (2), it follows that
d ( η n , η n + p ) < ε 3 s 2 < ε .
(c2)
If n is an odd number and n + p is an even number, then, by (1) and (2), one has
d ( η n , η n + p ) s [ d ( η n , η n + 1 ) + d ( η n + 1 , η n + p ) ] < s ε 3 s 2 + ε 3 s 2 < ε .
(c3)
If n is an even number and n + p is an odd number, then, by (1) and (2), it is easy to see that
d ( η n , η n + p ) s [ d ( η n , η n + p + 1 ) + d ( η n + p , η n + p + 1 ) ] < s ε 3 s 2 + ε 3 s 2 < ε .
(c4)
If n and n + p all are odd numbers, then, by (1) and (2), we have
d ( η n , η n + p ) s [ d ( η n , η n + 1 ) + d ( η n + 1 , η n + p ) ] s d ( η n , η n + 1 ) + s 2 [ d ( η n + 1 , η n + p + 1 ) + d ( η n + p , η n + p + 1 ) ] s · ε 3 s 2 + s 2 · ε 3 s 2 + ε 3 s 2 < ε .
Hence, { η n } is a b-Cauchy sequence. □

3. Main Results

In this section, we improve and generalize some common fixed point theorems from several references in several sides.
Throughout this paper, let P be the family of all functions β : [ 0 , + ) [ 0 , 1 ) . Let Ψ be the family of all functions ψ : R + R + satisfying the condition that ψ is continuous, nondecreasing and ψ ( t ) = 0 if and only if t = 0 . In this case, ψ is called altering distance function.
Let ( X , , d ) be a partially ordered b-metric space with s > 1 and f , g , S , T : X X be mappings. If there exist β P and ψ Ψ and a constant L 0 such that for every two comparable elements ξ , η X , it satisfies
ψ ( s ε d ( f ξ , g η ) ) β ( M ( ξ , η ) ) ψ max d ( S ξ , T η ) , d ( S ξ , f ξ ) , d ( T η , g η ) + L ψ ( N ( ξ , η ) ) ,
where
M ( ξ , η ) = ψ max d ( S ξ , T η ) , d ( S ξ , f ξ ) , d ( T η , g η ) , d ( S ξ , g η ) + d ( T η , f ξ ) 2 s ,
N ( ξ , η ) = min d ( S ξ , T η ) , d ( S ξ , g η ) , d ( T η , f ξ ) , d ( g η , T η ) , d ( ξ , g η ) ,
and ε > 0 is a constant, then ( f , g ) is said to be a ( ψ , β , L ) -ordered contractive pair with respect to S and T.
Theorem 1.
Let f , g , S and T be self-mappings on a partially ordered b-complete b-metric space ( X , , d ) with s > 1 . Let f X T X and g X S X . Suppose that ( f , g ) of dominating maps is a ( ψ , β , L ) -ordered contractive pair with respect to dominated maps S and T. If { ξ n } is a nondecreasing sequence with ξ n η n for all n and η n u implies ξ n u and either
(i) 
f or S is b-continuous, ( f , S ) are compatible and ( g , T ) are weakly compatible;
(ii) 
g or T is b-continuous, ( g , T ) are compatible and ( f , S ) are weakly compatible, then the mappings f , g , S and T possess a common fixed point in X. Moreover, the set of common points of f , g , S and T is well ordered if f , g , S and T have a unique common fixed point.
Proof. 
Let ξ 0 be an arbitrary point of X. Similar to [14], we construct a sequence { η n } in X such that η 2 n 1 = T ξ 2 n 1 = f ξ 2 n 2 and η 2 n = S ξ 2 n = g ξ 2 n 1 . Since f , g are dominating maps and S , T are dominated maps, it follows that ξ 2 n 2 f ξ 2 n 2 = T ξ 2 n 1 ξ 2 n 1 and ξ 2 n 1 g ξ 2 n 1 = S ξ 2 n ξ 2 n . Thus, for all n 1 , we have ξ n ξ n + 1 .
Without loss of generality, we assume d ( η 2 n , η 2 n + 1 ) > 0 for every n. If not, then η 2 n 0 = η 2 n 0 + 1 for some n 0 N . From (3) to (5), we obtain
ψ ( s ε d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) ) = ψ ( s ε d ( f ξ 2 n 0 , g ξ 2 n 0 + 1 ) ) β ( M ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) × ψ ( max { d ( S ξ 2 n 0 , T ξ 2 n 0 + 1 ) , d ( S ξ 2 n 0 , f ξ 2 n 0 ) , d ( T ξ 2 n 0 + 1 , g ξ 2 n 0 + 1 ) } ) + L ψ ( N ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) = β ( M ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) × ψ ( max { d ( η 2 n 0 , η 2 n 0 + 1 ) , d ( η 2 n 0 , η 2 n 0 + 1 ) , d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) } ) + L ψ ( N ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) = β ( M ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) × ψ ( max { 0 , 0 , d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) } ) = β ( M ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) ψ ( d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) ) ψ ( d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) ) ,
where
N ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) = min { d ( S ξ 2 n 0 , T ξ 2 n 0 + 1 ) , d ( S ξ 2 n 0 , g ξ 2 n 0 + 1 ) , d ( T ξ 2 n 0 + 1 , f ξ 2 n 0 ) , d ( g ξ 2 n 0 + 1 , T ξ 2 n 0 + 1 ) , d ( ξ 2 n 0 , g ξ 2 n 0 + 1 ) } = min { d ( η 2 n 0 , η 2 n 0 + 1 ) , d ( η 2 n 0 , η 2 n 0 + 2 ) , d ( η 2 n 0 + 1 , η 2 n 0 + 1 ) , d ( η 2 n 0 + 2 , η 2 n 0 + 1 ) , d ( ξ 2 n 0 , η 2 n 0 + 2 ) } = min { 0 , d ( η 2 n 0 , η 2 n 0 + 2 ) , 0 , d ( η 2 n 0 + 2 , η 2 n 0 + 1 ) , d ( ξ 2 n 0 , η 2 n 0 + 2 ) } = 0 ,
then
ψ ( N ( ξ 2 n 0 , ξ 2 n 0 + 1 ) ) = 0 .
Thus, by the monotonicity of function ψ , it is valid from (6) that
s ε d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) ,
that is,
d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) 1 s ε d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) ,
so d ( η 2 n 0 + 1 , η 2 n 0 + 2 ) = 0 (because s ε > 1 ). Hence, η 2 n 0 + 1 = η 2 n 0 + 2 .
Following similar arguments, we obtain η 2 n 0 + 2 = η 2 n 0 + 3 . Thus, { η n } becomes a constant sequence, and η 2 n 0 is the common fixed point of f , g , S and T. In this case, the conclusion we need to prove is clear.
Now, we take d ( η 2 n , η 2 n + 1 ) > 0 for each n. As ξ 2 n and ξ 2 n + 1 are comparable, by inequality (3) we have
ψ ( s ε d ( η 2 n + 1 , η 2 n + 2 ) ) = ψ ( s ε d ( f ξ 2 n , g ξ 2 n + 1 ) ) β ( M ( ξ 2 n , ξ 2 n + 1 ) ) × ψ ( max { d ( S ξ 2 n , T ξ 2 n + 1 ) , d ( S ξ 2 n , f ξ 2 n ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) + L ψ ( N ( ξ 2 n , ξ 2 n + 1 ) ) = β ( M ( ξ 2 n , ξ 2 n + 1 ) ) × ψ ( max { d ( η 2 n , η 2 n + 1 ) , d ( η 2 n , η 2 n + 1 ) , d ( η 2 n + 1 , η 2 n + 2 ) } ) + L ψ ( N ( ξ 2 n , ξ 2 n + 1 ) ) = β ( M ( ξ 2 n , ξ 2 n + 1 ) ) × ψ ( max { d ( η 2 n , η 2 n + 1 ) , d ( η 2 n + 1 , η 2 n + 2 ) } ) , < ψ ( max { d ( η 2 n , η 2 n + 1 ) , d ( η 2 n + 1 , η 2 n + 2 ) } ) ,
which follows immediately from the monotonicity of function ψ that
s ε d ( η 2 n + 1 , η 2 n + 2 ) < max { d ( η 2 n , η 2 n + 1 ) , d ( η 2 n + 1 , η 2 n + 2 ) } .
Now, if
d ( η 2 n , η 2 n + 1 ) d ( η 2 n + 1 , η 2 n + 2 ) ,
then by using (7) we have
s ε d ( η 2 n + 1 , η 2 n + 2 ) < d ( η 2 n + 1 , η 2 n + 2 ) ,
which leads to a contradiction because of s ε > 1 . Therefore, d ( η 2 n , η 2 n + 1 ) > d ( η 2 n + 1 , η 2 n + 2 ) . In this case, we have
s ε d ( η 2 n + 1 , η 2 n + 2 ) < d ( η 2 n , η 2 n + 1 ) ,
which implies that
d ( η 2 n + 1 , η 2 n + 2 ) < 1 s ε d ( η 2 n , η 2 n + 1 ) .
Again, by inequality (3) we have
ψ ( s ε d ( η 2 n + 1 , η 2 n ) ) = ψ ( s ε d ( f ξ 2 n , g ξ 2 n 1 ) ) β ( M ( ξ 2 n , ξ 2 n 1 ) ) × ψ ( max { d ( S ξ 2 n , T ξ 2 n 1 ) , d ( S ξ 2 n , f ξ 2 n ) , d ( T ξ 2 n 1 , g ξ 2 n 1 ) } ) + L ψ ( N ( ξ 2 n , ξ 2 n 1 ) ) = β ( M ( ξ 2 n , ξ 2 n 1 ) ) × ψ ( max { d ( η 2 n , η 2 n 1 ) , d ( η 2 n , η 2 n + 1 ) , d ( η 2 n 1 , η 2 n ) } ) + L ψ ( N ( ξ 2 n , ξ 2 n 1 ) ) = β ( M ( ξ 2 n , ξ 2 n 1 ) ) × ψ ( max { d ( η 2 n , η 2 n 1 ) , d ( η 2 n , η 2 n + 1 ) } ) < ψ ( max { d ( η 2 n , η 2 n 1 ) , d ( η 2 n , η 2 n + 1 ) } ) ,
where
N ( ξ 2 n , ξ 2 n 1 ) ) = min { d ( S ξ 2 n , T ξ 2 n 1 ) , d ( S ξ 2 n , g ξ 2 n 1 ) , d ( T ξ 2 n 1 , f ξ 2 n ) , d ( g ξ 2 n 1 , T ξ 2 n 1 ) , d ( ξ 2 n , g ξ 2 n 1 ) } = min { d ( η 2 n , η 2 n 1 ) , d ( η 2 n , η 2 n ) , d ( η 2 n 1 , η 2 n + 1 ) , d ( η 2 n , η 2 n 1 ) , d ( ξ 2 n , η 2 n ) } = min { d ( η 2 n , η 2 n 1 ) , 0 , d ( η 2 n 1 , η 2 n + 1 ) , d ( η 2 n , η 2 n 1 ) , d ( ξ 2 n , η 2 n ) } = 0 .
By (9) and the monotonicity of function ψ , we have
s ε d ( η 2 n + 1 , η 2 n ) < max { d ( η 2 n , η 2 n 1 ) , d ( η 2 n , η 2 n + 1 ) } .
Now, if
d ( η 2 n , η 2 n 1 ) d ( η 2 n , η 2 n + 1 ) ,
then by using (10) we have
s ε d ( η 2 n , η 2 n + 1 ) < d ( η 2 n , η 2 n + 1 ) ,
which leads to a contradiction because of s ε > 1 . Therefore, d ( η 2 n , η 2 n 1 ) > d ( η 2 n , η 2 n + 1 ) . In this case, we have
s ε d ( η 2 n , η 2 n + 1 ) < d ( η 2 n 1 , η 2 n ) ,
which implies that
d ( η 2 n , η 2 n + 1 ) < 1 s ε d ( η 2 n 1 , η 2 n ) .
Making full use of (8) and (11), we have
d ( η n , η n + 1 ) < 1 s ε d ( η n 1 , η n ) .
Accordingly, by Lemma 1, we claim that { η n } is a b-Cauchy sequence. Since ( X , , d ) is b-complete, then there exists a point ζ in X such that { η n } converges to ζ .
We first suppose that (i) holds. Assume that S is b-continuous. Because ( f , S ) are compatible, we have
lim n d ( f S ξ 2 n + 2 , S f ξ 2 n + 2 ) = 0 ,
that is,
lim n d ( f η 2 n + 2 , S η 2 n + 3 ) = 0 .
As a result, by the b-continuity of S, we speculate that
d ( f η 2 n + 2 , S ζ ) s [ d ( f η 2 n + 2 , S η 2 n + 3 ) + d ( S η 2 n + 3 , S ζ ) ] 0 ( n ) ,
which implies that
lim n d ( f η 2 n + 2 , S ζ ) = 0 .
Now, we show that ζ = S ζ . If not, that is, d ( S ζ , ζ ) > 0 . As ξ 2 n + 1 g ξ 2 n + 1 = S ξ 2 n + 2 , from inequality (3), we have
ψ ( s ε d ( f η 2 n + 2 , η 2 n + 2 ) ) = ψ ( s ε d ( f S ξ 2 n + 2 , g ξ 2 n + 1 ) ) β ( M ( S ξ 2 n + 2 , ξ 2 n + 1 ) ) × ψ ( max { d ( S S ξ 2 n + 2 , T ξ 2 n + 1 ) , d ( S S ξ 2 n + 2 , f S ξ 2 n + 2 ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } + L ψ ( N ( S ξ 2 n + 2 , ξ 2 n + 1 ) ) < ψ ( max { d ( S S ξ 2 n + 2 , T ξ 2 n + 1 ) , d ( S S ξ 2 n + 2 , f S ξ 2 n + 2 ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) = ψ ( max { d ( S η 2 n + 2 , η 2 n + 1 ) , d ( S η 2 n + 2 , f η 2 n + 2 ) , d ( η 2 n + 1 , η 2 n + 2 ) } ) ,
where
N ( S ξ 2 n + 2 , ξ 2 n + 1 ) ) = min { d ( S S ξ 2 n + 2 , T ξ 2 n + 1 ) , d ( S S ξ 2 n + 2 , g ξ 2 n + 1 ) , d ( T ξ 2 n + 1 , f S ξ 2 n + 2 ) , d ( g ξ 2 n + 1 , T ξ 2 n + 1 ) , d ( S ξ 2 n + 2 , g ξ 2 n + 1 ) } = min { d ( S η 2 n + 2 , η 2 n + 1 ) , d ( S η 2 n + 2 , η 2 n + 2 ) , d ( η 2 n + 1 , f η 2 n + 2 ) , d ( η 2 n + 2 , η 2 n + 1 ) , d ( η 2 n + 2 , η 2 n + 2 ) } = 0 .
By (13) and the monotonicity of function ψ , we have
s ε d ( f η 2 n + 2 , η 2 n + 2 ) < max { d ( S η 2 n + 2 , η 2 n + 1 ) , d ( S η 2 n + 2 , f η 2 n + 2 ) , d ( η 2 n + 1 , η 2 n + 2 ) } .
Combing (12) and the b-continuity of function S, we obtain
d ( S η 2 n + 2 , f η 2 n + 2 ) s [ d ( S η 2 n + 2 , S ζ ) + d ( S ζ , f η 2 n + 2 ) ] s ( 0 + 0 ) = 0 ( n ) ,
which means that
lim n d ( S η 2 n + 2 , f η 2 n + 2 ) = 0 .
Moreover, since { η n } is a b-Cauchy sequence, we have
lim n d ( η 2 n + 1 , η 2 n + 2 ) = 0 .
By (15) and (16), there exists N 1 N such that, for all n > N 1 , one has
d ( S η 2 n + 2 , η 2 n + 1 ) > d ( S η 2 n + 2 , f η 2 n + 2 ) , d ( S η 2 n + 2 , η 2 n + 1 ) > d ( η 2 n + 1 , η 2 n + 2 ) .
Via (14) and (17), it is easy to see that
s ε d ( f η 2 n + 2 , η 2 n + 2 ) < d ( S η 2 n + 2 , η 2 n + 1 ) ( n > N 1 ) .
Using the triangular inequality of the b-metric (12) and the b-continuity of function S, we obtain
| d ( f η 2 n + 2 , η 2 n + 2 ) d ( S η 2 n + 2 , η 2 n + 1 ) | = | [ d ( f η 2 n + 2 , η 2 n + 2 ) s d ( η 2 n + 2 , S ζ ) ] + [ s d ( η 2 n + 2 , S ζ ) d ( S ζ , ζ ) ] + [ d ( S ζ , ζ ) s d ( S η n + 2 , ζ ) ] + [ s d ( S η n + 2 , ζ ) d ( S η n + 2 , η 2 n + 1 ) ] | | d ( f η 2 n + 2 , η 2 n + 2 ) s d ( η 2 n + 2 , S ζ ) | + | s d ( η 2 n + 2 , S ζ ) d ( S ζ , ζ ) | + | d ( S ζ , ζ ) s d ( S η n + 2 , ζ ) | + | s d ( S η n + 2 , ζ ) d ( S η n + 2 , η 2 n + 1 ) | s d ( f η 2 n + 2 , S ζ ) + s d ( η 2 n + 2 , ζ ) + s d ( S ζ , S η n + 2 ) + s d ( ζ , η 2 n + 1 ) 0 ( n ) ,
which establishes that
lim   sup n   d ( f η 2 n + 2 , η 2 n + 2 ) = lim   sup n   d ( S η 2 n + 2 , η 2 n + 1 ) .
Consider (18) and (19), they lead to a contradiction because of s ε > 1 . Consequently, S ζ = ζ .
Now, by virtue of ξ 2 n + 1 g ξ 2 n + 1 and g ξ 2 n + 1 ζ as n , ξ 2 n + 1 ζ . It suffices to prove ζ = f ζ . As a matter of fact, firstly, notice that
| d ( f ζ , η n + 2 ) d ( ζ , f ζ ) | = | [ d ( f ζ , η n + 2 ) s d ( f ζ , η n + 1 ) ] + [ s d ( f ζ , η n + 1 ) d ( ζ , f ζ ) ] | | d ( f ζ , η n + 2 ) s d ( f ζ , η n + 1 ) | + | s d ( f ζ , η n + 1 ) d ( ζ , f ζ ) | s d ( η n + 2 , η n + 1 ) + s d ( η n + 1 , ζ ) 0 ( n ) ,
which follows that
lim   sup n   d ( f ζ , η n + 2 ) = d ( ζ , f ζ ) .
Secondly, by inequality (3), we have
ψ ( s ε d ( f ζ , η 2 n + 2 ) ) = ψ ( s ε d ( f ζ , g ξ 2 n + 1 ) ) β ( M ( ζ , ξ 2 n + 1 ) ) × ψ ( max { d ( S ζ , T ξ 2 n + 1 ) , d ( S ζ , f ζ ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) + L ψ ( N ( ζ , ξ 2 n + 1 ) ) = β ( M ( ζ , ξ 2 n + 1 ) ) × ψ ( max { d ( ζ , T ξ 2 n + 1 ) , d ( ζ , f ζ ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) + L ψ ( N ( ζ , ξ 2 n + 1 ) ) < ψ ( max { d ( ζ , η 2 n + 1 ) , d ( ζ , f ζ ) , d ( η 2 n + 1 , η 2 n + 2 ) } ) + L ψ ( N ( ζ , ξ 2 n + 1 ) ) ,
where
ψ ( N ( ζ , ξ 2 n + 1 ) ) = ψ ( min { d ( S ζ , T ξ 2 n + 1 ) , d ( S ζ , g ξ 2 n + 1 ) , d ( T ξ 2 n + 1 , f ζ ) , d ( g ξ 2 n + 1 , T ξ 2 n + 1 ) , d ( ζ , g ξ 2 n + 1 ) } )
tends to
ψ min lim   sup n   d ( S ζ , T ξ 2 n + 1 ) , lim   sup n   d ( S ζ , g ξ 2 n + 1 ) , lim   sup n   d ( T ξ 2 n + 1 , f ζ ) , 0 , 0 = ψ ( 0 ) = 0
as n . By taking the upper limit as n from (21), we obtain
s ε   lim   sup n   d ( f ζ , η 2 n + 2 ) d ( f ζ , ζ ) ,
which is a contradiction with s ε > 1 and (20) unless lim   sup n   d ( f ζ , η 2 n + 2 ) = 0 . As a result, by
1 s d ( f ζ , ζ ) d ( f ζ , η 2 n + 2 ) + d ( η 2 n + 2 , ζ ) ,
we have
1 s d ( f ζ , ζ ) lim   sup n   [ d ( f ζ , η 2 n + 2 ) + d ( η 2 n + 2 , ζ ) ] = 0 ,
which implies that d ( f ζ , ζ ) = 0 . Hence, f ζ = ζ .
In view of f ( X ) T ( X ) , then there exists a point w X such that ζ = f ζ = T w . Now, we show d ( T w , g w ) = 0 . If not, i.e., d ( T w , g w ) > 0 . Since ζ f ζ = T w w implies ζ w , from inequality (3), we have
ψ ( s ε d ( T w , g w ) ) = ψ ( s ε d ( f ζ , g w ) ) β ( M ( ζ , w ) ) × ψ ( max { d ( S ζ , T w ) , d ( S ζ , f ζ ) , d ( T w , g w ) } ) + L ψ ( N ( ζ , w ) ) = β ( M ( ζ , w ) ) × ψ ( max { 0 , 0 , d ( T w , g w ) } ) + L ψ ( min { d ( S ζ , T w ) , d ( S ζ , g w ) , d ( T w , f ζ ) , d ( g w , T w ) , d ( ζ , g w ) } ) ) = β ( M ( ζ , w ) ) × ψ ( d ( T w , g w ) ) + L ψ ( min { 0 , d ( ζ , g w ) , 0 , d ( g w , ζ ) , d ( ζ , g w ) } ) ) = β ( M ( ζ , w ) ) ψ ( d ( T w , g w ) ) < ψ ( d ( T w , g w ) ) .
Hence, by the monotonicity of function ψ , we have
s ε d ( T w , g w ) < d ( T w , g w ) ,
which leads to a contradiction with s ε > 1 . As a result, T w = g w . On account of the fact that g and T are weakly compatible, we obtain
g ζ = g f ζ = g T w = T g w = T f ζ = T ζ .
Thus, ζ is a coincidence point of g and T. Next, we show that ζ = g ζ . As ξ 2 n f ξ 2 n and f ξ 2 n ζ ( n ) implies that ξ 2 n ζ , from (3) we have
ψ ( s ε d ( η 2 n + 1 , g ζ ) ) = ψ ( s ε d ( f ξ 2 n , g ζ ) ) β ( M ( ξ 2 n , ζ ) ) × ψ ( max { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , f ξ 2 n ) , d ( T ζ , g ζ ) } ) + L ψ ( N ( ξ 2 n , ζ ) ) < ψ ( max { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , f ξ 2 n ) , d ( T ζ , g ζ ) } ) + L ψ ( min { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , g ζ ) , d ( T ζ , f ξ 2 n ) , d ( g ζ , T ζ ) , d ( ξ 2 n , g ζ ) } ) = ψ ( max { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , f ξ 2 n ) , 0 } ) + L ψ ( min { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , g ζ ) , d ( T ζ , f ξ 2 n ) , 0 , d ( ξ 2 n , g ζ ) } ) = ψ ( max { d ( S ξ 2 n , T ζ ) , d ( S ξ 2 n , f ξ 2 n ) } ) + L ψ ( 0 ) = ψ ( max { d ( η 2 n , g ζ ) , d ( η 2 n , η 2 n + 1 ) } ) ,
which follows immediately from the monotonicity of function ψ that
s ε d ( f η 2 n + 1 , g ζ ) < max { d ( η 2 n , g ζ ) , d ( η 2 n , η 2 n + 1 ) } .
Since lim n d ( η 2 n , η 2 n + 1 ) = 0 , then there exists N 2 N such that, for all n > N 2 , one has
d ( η 2 n , g ζ ) > d ( η 2 n , η 2 n + 1 ) .
Considering (22) and (23), we acquire
s ε d ( η 2 n + 1 , g ζ ) < d ( η 2 n , g ζ ) ( n > N 2 ) .
Thus, it implies
s ε   lim   sup n   d ( η 2 n + 1 , g ζ ) lim   sup n   d ( η 2 n , g ζ ) .
By virtue of
| d ( η 2 n + 1 , g ζ ) d ( η 2 n , g ζ ) | = | [ d ( η 2 n + 1 , g ζ ) s d ( g ζ , η 2 n 1 ) ] + [ s d ( g ζ , η 2 n 1 ) d ( η 2 n , g ζ ) ] | | d ( η 2 n + 1 , g ζ ) s d ( g ζ , η 2 n 1 ) | + | s d ( g ζ , η 2 n 1 ) d ( η 2 n , g ζ ) | s d ( η 2 n + 1 , η 2 n 1 ) + s d ( η 2 n 1 , η 2 n ) 0 ( n ) ,
which follows that
lim   sup n   d ( η 2 n + 1 , g ζ ) = lim   sup n   d ( η 2 n , g ζ ) .
Consequently, (24) is a contradiction with s ε > 1 unless lim   sup n   d ( η 2 n + 1 , g ζ ) = 0 . That is to say, lim   sup n   d ( η 2 n + 1 , g ζ ) = 0 . Now, by
1 s d ( g ζ , ζ ) d ( g ζ , η 2 n + 1 ) + d ( η 2 n + 1 , ζ ) ,
we claim that
1 s d ( g ζ , ζ ) lim   sup n   d ( g ζ , η 2 n + 1 ) + lim   sup n   d ( η 2 n + 1 , ζ ) = 0 ,
which follows that d ( g ζ , ζ ) = 0 . Thus, g ζ = ζ .
To sum up, f ζ = g ζ = S ζ = T ζ = ζ . In other words, ζ is a common fixed point of f, g, S and T. The proof is similar when f is b-continuous.
Similarly, the result follows when (ii) holds. Now, suppose that the set of common fixed points of f , g , S and T is well ordered; we will show that the common fixed point of f , g , S and T is unique. Indeed, assume, on the contrary, that f q = g q = S q = T q = q and f r = g r = S r = T r = r but d ( q , r ) > 0 . By the given assumption, we replace ξ with q and η with r in (3). Then,
ψ ( s ε d ( q , r ) ) = ψ ( s ε d ( f q , g r ) ) β ( M ( q , r ) ) × ψ ( max { d ( S q , T r ) , d ( S q , f q ) , d ( T r , g r ) } ) + L ψ ( N ( q , r ) ) < ψ ( d ( q , r ) ) .
Consequently, we claim that s ε d ( q , r ) < d ( q , r ) , a contradiction. As a result, q = r (because s ε > 1 ). In reverse, if f , g , S and T are single, then it is well ordered. □
Remark 1.
Theorem 1 greatly generalizes Theorem 12 of [14] from several sides. On the one hand, Theorem 1 refers to the conclusion in the setting of b-metric spaces, whereas Theorem 12 of [14] considered the result in usual metric space. It is well-known that b-metric space is a sharp generalization of usual metric space since the given b-metric usually is not necessarily continuous, but the usual metric must be a continuous function. Therefore, Theorem 1 is more important than Theorem 12 of [14]. On the other hand, as compared with Theorem 12 of [14], function β from Theorem 1 satisfies the simpler condition. In addition, the proof of Theorem 1 is more straightforward than the one of Theorem 12 of [14].
Remark 2.
We use Lemma 2 to prove that the constructed sequence is a b-Cauchy sequence instead of using Lemma 11 from [14] or Lemma 1 from [16]. This is a straightforward improvement because Lemma 2 is easily understood for most of our readers. Otherwise, by Lemma 2, we also can prove that the sequence is a b-Cauchy sequence but, due to the complicated process, we omit it in this paper. In addition, in order to overcome the discontinuity of b-metric, we use a new method to prove our theorem. This is a great innovation.
Example 5.
Let X = { 1 , 2 , 3 , 4 } be a partially ordered set defined as ξ η if and only if ξ η and d ( ξ , η ) = | ξ η | 2 for all ξ , η X . Then, ( X , d ) is a b-complete b-metric space with s = 9 5 . Define ordered self-mappings f , g , S , and T on X using
f = 1 2 3 4 1 1 2 2 , g = 1 2 3 4 1 1 1 2 ,
S = 1 2 3 4 1 3 4 4 , T = 1 2 3 4 1 4 3 4 .
It is easy to verify that the mappings f and g are dominating and S and T are dominated. Take ψ ( ξ ) = ln ( ξ + 1 ) and β = ψ ( ξ ) ξ ; then, the mappings f , g , S and T satisfy all the conditions given in Theorem 1 with ε = 1 4 . Moreover, one is the unique common fixed point of f , g , S , and T.
Theorem 2.
Let f , g , S and T be self-mappings on a partially ordered b-complete b-metric space ( X , , d ) with s > 1 . Let f ( X ) T ( X ) and g ( X ) S ( X ) . Suppose that the mappings f , g , S and T are b-continuous, the pairs ( f , S ) and ( g , T ) are compatible, and the pairs ( f , g ) and ( g , f ) are partially weakly increasing with respect to T and S, respectively. Assume that ( f , g ) is a ( ψ , β , L ) -ordered contractive pair with respect to S and T for each ξ , η X , for which S ξ and T η are comparable. Then, the pairs ( f , S ) and ( g , T ) have a coincidence point ζ X . Moreover, ζ is a coincidence point of the mappings f , g , S and T provided that S ζ and T ζ are comparable.
Proof. 
Choose ξ 0 X . Define a sequence { η n } in X that satisfies η 2 n = f ξ 2 n = T ξ 2 n + 1 and η 2 n + 1 = g ξ 2 n + 1 = S ξ 2 n + 2 for all n N . By the hypothesis, it is easy to see that η n η n + 1 for all n 1 . We start the proof with two steps.
Step 1. We prove that
d ( η n + 1 , η n + 2 ) λ d ( η n , η n + 1 ) ,
for each n N , where λ [ 0 , 1 ) is a constant.
Firstly, we assume η n η n + 1 for each n N . Since S ξ 2 n = η 2 n 1 = g ξ 2 n 1 and T ξ 2 n + 1 = η 2 n = f ξ 2 n are comparable, then, via (3), it has
ψ ( s ε d ( η 2 n , η 2 n + 1 ) ) = ψ ( s ε d ( f ξ 2 n , g ξ 2 n + 1 ) ) β ( M ( ξ 2 n , ξ 2 n + 1 ) ) × ψ ( max { d ( S ξ 2 n , T ξ 2 n + 1 ) , d ( S ξ 2 n , f ξ 2 n ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) + L ψ ( min { d ( S ξ 2 n , T ξ 2 n + 1 ) , d ( S ξ 2 n , g ξ 2 n + 1 ) , d ( T ξ 2 n + 1 , f ξ 2 n ) , d ( g ξ 2 n + 1 , T ξ 2 n + 1 ) , d ( ξ 2 n , g ξ 2 n + 1 ) } ) < ψ ( max { d ( η 2 n 1 , η 2 n ) , d ( η 2 n 1 , η 2 n ) , d ( η 2 n , η 2 n + 1 ) } ) + L ψ ( min { d ( η 2 n 1 , η 2 n ) , d ( η 2 n 1 , η 2 n + 1 ) , d ( η 2 n , η 2 n ) , d ( η 2 n + 1 , η 2 n ) , d ( ξ 2 n , η 2 n + 1 ) } ) ) = ψ ( max { d ( η 2 n 1 , η 2 n ) , d ( η 2 n , η 2 n + 1 ) } ) ,
which follows from the monotonicity of function ψ that
s ε d ( η 2 n , η 2 n + 1 ) < max { d ( η 2 n 1 , η 2 n ) , d ( η 2 n , η 2 n + 1 ) } .
If d ( η 2 n 1 , η 2 n ) d ( η 2 n , η 2 n + 1 ) , then by (26) it means that
s ε d ( η 2 n , η 2 n + 1 ) < d ( η 2 n , η 2 n + 1 ) ,
which leads to a contradiction (because s ε > 1 ) . Then,
s ε d ( η 2 n , η 2 n + 1 ) < d ( η 2 n 1 , η 2 n ) .
Again, since S ξ 2 n + 2 = η 2 n + 1 = g ξ 2 n + 1 and T ξ 2 n + 1 = η 2 n = f ξ 2 n are comparable, then, by (3), it implies that
ψ ( s ε d ( η 2 n + 1 , η 2 n + 2 ) ) = ψ ( s ε d ( f ξ 2 n + 2 , g ξ 2 n + 1 ) ) β ( M ( ξ 2 n + 2 , ξ 2 n + 1 ) ) × ψ ( max { d ( S ξ 2 n + 2 , T ξ 2 n + 1 ) , d ( S ξ 2 n + 2 , f ξ 2 n + 2 ) , d ( T ξ 2 n + 1 , g ξ 2 n + 1 ) } ) + L ψ ( min { d ( S ξ 2 n + 2 , T ξ 2 n + 1 ) , d ( S ξ 2 n + 2 , g ξ 2 n + 1 ) , d ( T ξ 2 n + 1 , f ξ 2 n + 2 ) , d ( g ξ 2 n + 1 , T ξ 2 n + 1 ) , d ( ξ 2 n + 2 , g ξ 2 n + 1 ) } ) < ψ ( max { d ( η 2 n + 1 , η 2 n ) , d ( η 2 n + 1 , η 2 n + 2 ) , d ( η 2 n , η 2 n + 1 ) } ) + L ψ ( min { d ( η 2 n + 1 , η 2 n ) , d ( η 2 n + 1 , η 2 n + 1 ) , d ( η 2 n , η 2 n + 2 ) , d ( η 2 n + 1 , η 2 n ) , d ( ξ 2 n + 2 , η 2 n + 1 ) } ) ) = ψ ( max { d ( η 2 n + 1 , η 2 n + 2 ) , d ( η 2 n , η 2 n + 1 ) } ) ,
which follows from the monotonicity of function ψ that
s ε d ( η 2 n + 1 , η 2 n + 2 ) < max { d ( η 2 n + 1 , η 2 n + 2 ) , d ( η 2 n , η 2 n + 1 ) } .
If d ( η 2 n , η 2 n + 1 ) d ( η 2 n + 1 , η 2 n + 2 ) , then, by means of (28), it establishes
s ε d ( η 2 n + 1 , η 2 n + 2 ) < d ( η 2 n + 1 , η 2 n + 2 ) ,
which leads to a contradiction (because s ε > 1 ). Thus,
s ε d ( η 2 n + 1 , η 2 n + 2 ) < d ( η 2 n , η 2 n + 1 ) .
Now, by using (27) and (29), we obtain (25), where λ = 1 s ε [ 0 , 1 ) .
We now assume η n 0 = η n 0 + 1 for some n 0 N . If n 0 = 2 k 1 , then η 2 k 1 = η 2 k implies η 2 k = η 2 k + 1 . As a matter of fact, if η 2 k η 2 k + 1 , i.e., d ( η 2 k , η 2 k + 1 ) > 0 , then, by the fact that η 2 k = f ξ 2 k = T ξ 2 k + 1 and η 2 k 1 = g ξ 2 k 1 = S ξ 2 k are comparable, then, via (26), we speculate that
s ε d ( η 2 k , η 2 k + 1 ) < max { d ( η 2 k 1 , η 2 k ) , d ( η 2 k , η 2 k + 1 ) } = d ( η 2 k , η 2 k + 1 ) .
This is a contradiction (because s ε > 1 ). Hence, d ( η 2 k , η 2 k + 1 ) = 0 , i.e., η 2 k = η 2 k + 1 . If n 0 = 2 k , then η 2 k = η 2 k + 1 leads to η 2 k + 1 = η 2 k + 2 . Actually, if η 2 k + 1 η 2 k + 2 , then, i.e., d ( η 2 k + 1 , η 2 k + 2 ) > 0 . Since η 2 k + 1 = g ξ 2 k + 1 = S ξ 2 k + 2 and η 2 k = f ξ 2 k = T ξ 2 k + 1 are comparable, then, by (28), we claim that
s ε d ( η 2 k + 1 , η 2 k + 2 ) < max { d ( η 2 k , η 2 k + 1 ) , d ( η 2 k + 1 , η 2 k + 2 ) } = d ( η 2 k + 1 , η 2 k + 2 ) .
This is a contradiction (because s ε > 1 ). Thus, d ( η 2 k + 1 , η 2 k + 2 ) = 0 , i.e., η 2 k + 1 = η 2 k + 2 . Therefore, the sequence { η n } in both cases is equal to a constant for n n 0 and so (25) holds.
Step 2. We prove that f , g , S and T have a coincidence point. Taking advantage of (25) and Lemma 1, we say that { η n } is a b-Cauchy sequence. Since ( X , d ) is b-complete, then there is a ζ X satisfying that lim n η n = ζ . We obtain
lim n d ( S ξ 2 n , ζ ) = lim n d ( f ξ 2 n , ζ ) = lim n d ( S ξ 2 n + 2 , ζ ) = lim n d ( T ξ 2 n + 1 , ζ ) = lim n d ( g ξ 2 n + 1 , ζ ) = 0 .
Since the pairs ( f , S ) and ( g , T ) are compatible, then
lim n d ( S f ξ 2 n , f S ξ 2 n ) = lim n d ( T g ξ 2 n + 1 , g T ξ 2 n + 1 ) = 0 .
On the other hand, owing to the b-continuity of f , g , S and T, we obtain
lim n d ( S f ξ 2 n , S ζ ) = lim n d ( f S ξ 2 n , f ζ ) = 0 ,
lim n d ( T g ξ 2 n + 1 , T ζ ) = lim n d ( g T ξ 2 n + 1 , g ζ ) = 0 .
We now acquire that
1 s d ( S ζ , f ζ ) d ( S ζ , S f ξ 2 n ) + d ( S f ξ 2 n , f ζ ) d ( S ζ , S f ξ 2 n ) + s [ d ( S f ξ 2 n , f S ξ 2 n ) + d ( f S ξ 2 n , f ζ ) ] ,
and
1 s d ( T ζ , g ζ ) d ( T ζ , T g ξ 2 n + 1 ) + d ( T g ξ 2 n + 1 , g ζ ) d ( T ζ , T g ξ 2 n + 1 ) + s [ d ( T g ξ 2 n + 1 , g T ξ 2 n + 1 ) + d ( g T ξ 2 n + 1 , g ζ ) ] .
On taking the limit as n from both sides of (30) and (31), we obtain 1 s d ( S ζ , f ζ ) 0 and 1 s d ( T ζ , g ζ ) 0 , that is, f ζ = S ζ , g ζ = T ζ .
Since S ζ and T ζ are comparable, we prove f ζ = g ζ . Suppose the contrary, then, by (3), it is valid that
ψ ( s ε d ( f ζ , g ζ ) β ( M ( ζ , ζ ) ) × ψ ( max { d ( S ζ , T ζ ) , d ( S ζ , f ζ ) , d ( T ζ , g ζ ) } ) + L ψ ( N ( ζ , ζ ) ) < ψ ( d ( S ζ , T ζ ) ) = ψ ( d ( f ζ , g ζ ) ) ,
that is to say,
s ε d ( f ζ , g ζ ) < d ( f ζ , g ζ ) .
This is a contradiction (because s ε > 1 ). Accordingly, f ζ = g ζ = S ζ = T ζ . □
Remark 3.
According to the main results of [14,15], Theorem 2 makes a general generalization. That is to say, it generalizes Theorem 15 in [14] and Theorem 2.1 in [15]. Similar superiority is discussed in Remarks 1 and 2.
The following result is a straightforward outcome of Theorem 2.
Corollary 1.
Let f and T be self-mappings on a partially ordered b-complete b-metric space ( X , , d ) with s > 1 . Assume that f ( X ) T ( X ) and the pair ( f , T ) is compatible, f and T are b-continuous, and f is partially weakly increasing with respect to T. Assume that f satisfies the following inequality
ψ ( s ε d ( f ξ , f η ) ) β ( M ( ξ , η ) ) ψ max d ( T ξ , T η ) , d ( T ξ , f ξ ) , d ( T η , f η ) + L ψ ( N ( ξ , η ) ) ,
for every ξ , η X , for which T ξ and T η are comparable, where
M ( ξ , η ) = ψ max d ( T ξ , T η ) , d ( T ξ , f ξ ) , d ( T η , f η ) , d ( T ξ , f η ) + d ( T η , f ξ ) 2 s ,
N ( ξ , η ) = min d ( T ξ , T η ) , d ( T ξ , f η ) , d ( T η , f ξ ) , d ( f η , T η ) , d ( ξ , f η ) ,
and ε > 0 is a constant. Then, the pair ( f , T ) has a coincidence point ζ X . Moreover, ζ is a coincidence point of f and T.

Author Contributions

B.J. offered the draft preparation and gave the support of funding acquisition. H.H. designed the research and made revisions to the paper. S.R. gave the methodology and checked the final version before submission. All authors have read and agreed to the published version of the manuscript.

Funding

The second author is partially supported by Grant No. NSTC 111-2115-M-017-002 of the National Science and Technology Council of the Republic of China.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The authors thank the Editor and the Referees for their valuable comments and suggestions, which improved greatly the quality of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Banach, S. Sur les opérations dans les ensembles abstraits et leur application auxéquations intégrales. Fundam. Math. 1922, 3, 133–181. [Google Scholar] [CrossRef]
  2. Eshraghisamani, M.; Vaezpour, S.M.; Asadi, M. New fixed point results on Branciari metric spaces. J. Math. Appl. 2017, 8, 132–141. [Google Scholar]
  3. Gabeleh, M. Best proximity points for weak proximal contractions. Bull. Malays. Math. Sci. Soc. 2015, 38, 143–154. [Google Scholar] [CrossRef]
  4. Vulpe, I.M.; Ostrajkh, D.; Khojman, F. The topological structure of a quasi-metric space. In Investigations in Functional Analysis and Differential Equations; Math. Sci., Interuniv. Work Collect., Kishinev; 1981; pp. 14–19, Prosevechtchénié, St. Petersburg. (In Russian) [Google Scholar]
  5. Jovanović, M.; Kadelburg, Z.; Radenović, S. Common fixed point results in metric-type spaces. Fixed Point Theory Appl. 2010, 2010, 978121. [Google Scholar] [CrossRef]
  6. Lu, N.; He, F.; Du, W.-S. Fundamental questions and new counterexamples for b-metric spaces and Fatou property. Mathematics 2019, 7, 1107. [Google Scholar] [CrossRef]
  7. Boriceanu, M.; Bota, M.; Petrusel, A. Mutivalued fractals in b-metric spaces. Cent. Eur. J. Math. 2010, 8, 367–377. [Google Scholar] [CrossRef]
  8. Lu, N.; He, F.; Du, W.-S. On the best areas for Kannan system and Chatterjea system in b-metric spaces. Optimization 2021, 70, 973–986. [Google Scholar] [CrossRef]
  9. Czerwik, S. Nonlinear set-valued contraction mappings in b-metric spaces. Atti Semin. Mat. Fis. Univ. Modena 1998, 46, 263–276. [Google Scholar]
  10. Khamsi, M.A.; Hussain, N. KKM mappings in metric type spaces. Nonlinear Anal. 2010, 73, 3123–3129. [Google Scholar] [CrossRef]
  11. Beg, I.; Abbas, M. Coincidence point and invariant approximation for mappings satisfying generalized weak contractive condition. Fixed Point Theory Appl. 2006, 2006, 978121. [Google Scholar] [CrossRef]
  12. Abbas, M.; Nazir, T.; Radenović, S. Common fixed points of four maps in partially ordered metric spaces. Appl. Math. Lett. 2011, 24, 1520–1526. [Google Scholar] [CrossRef]
  13. Esmaily, J.; Vaezpour, S.M.; Rhoades, B.E. Coincidence point theorem for generalized weakly contractions in ordered metric spaces. Appl. Math. Comput. 2012, 219, 1536–1548. [Google Scholar] [CrossRef]
  14. Abbas, M.; Zulfaqar, I.; Radenović, S. Common fixed point of (ψ,β)-generalized contractive mappings in partially ordered metric spaces. Chin. J. Math. 2014, 9, 5497–5509. [Google Scholar] [CrossRef]
  15. Huang, H.; Stojan, R.; Jelena, V. On some recent coincidence and immediate consequences in partially ordered b-metric spaces. Fixed Point Theory Appl. 2015, 2015, 63. [Google Scholar] [CrossRef]
  16. Hussain, N.; Parvaneh, V.; Roshan, J.R.; Kadelburg, Z. Fixed points of cyclic weakly (ψ,φ,L,A,B)-contractive mappings in ordered b-metric spaces with applications. Fixed Point Theory Appl. 2013, 2013, 256. [Google Scholar] [CrossRef]
  17. Czerwik, S. Contraction mappings in b-metric spaces. Acta Math. Inform. Univ. Ostrav. 1993, 1, 5–11. [Google Scholar]
  18. Berinde, V.; Pǎcurar, M. The early developments in fixed point theory on b-metric spaces: A brief survey and some important related aspects. Carpathian J. Math. 2022, 38, 523–538. [Google Scholar] [CrossRef]
  19. Aydi, H.; Bota, M.-F.; Karapınar, E.; Moradi, S. A common fixed point for weak φ-contractions on b-metric spaces. Fixed Point Theory 2012, 13, 337–346. [Google Scholar]
  20. Berinde, V.; Borcut, M. Tripled fixed point theorems for contractive type mappings in partially ordered metric spaces. Nonlinear Anal. 2011, 74, 4889–4897. [Google Scholar] [CrossRef]
  21. Bota, M.-F.; Karapınar, E. A note on “ Some results on multi-valued weakly Jungck mappings in b-metric spaces”. Cent. Eur. J. Math. 2013, 11, 1711–1712. [Google Scholar] [CrossRef]
  22. Bota, M.-F.; Karapınar, E.; Mlesnite, O. Ulam-Hyers stability results for fixed point problems via (α,ψ)-contractive mappings in b-metric spaces. Abstr. Appl. Anal. 2013, 2013, 825293. [Google Scholar] [CrossRef]
  23. Romaguera, S. On the correlation between Banach contraction principle and Caristi’s fixed point theorem in b-metric spaces. Mathematics 2022, 10, 136. [Google Scholar] [CrossRef]
  24. Kutbi, M.-A.; Karapınar, E.; Ahmad, J.; Azam, A. Some fixed point results for multi-valued mappings in b-metric spaces. J. Inequal. Appl. 2014, 2014, 126. [Google Scholar] [CrossRef]
  25. Radenović, S.; Kadelburg, Z.; Jandrlic, D.; Jandrlic, A. Some results on weakly contractive maps. Bull. Iranian Math. Soc. 2012, 38, 625–645. [Google Scholar]
  26. Radu, M.; Alexandru, M. New fixed point theorems for set-valued contractions in b-metric spaces. J. Fixed Point Theory Appl. 2017, 10, 2153–2163. [Google Scholar]
  27. Roshan, J.R.; Parvaneh, V.; Altun, I. Some coincidence point results in ordered b-metric spaces and applications in a system of integral equations. Appl. Math. Comput. 2014, 226, 725–737. [Google Scholar] [CrossRef]
  28. Roshan, J.R.; Parvaneh, V.; Kadelburg, Z. Common fixed point theorems for weakly isotone increasing mappings in ordered b-metric spaces. J. Nonlinear Sci. Appl. 2014, 7, 229–245. [Google Scholar] [CrossRef]
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

Jiang, B.; Huang, H.; Radenović, S. Common Fixed Point of (ψ, β, L)-Generalized Contractive Mapping in Partially Ordered b-Metric Spaces. Axioms 2023, 12, 1008. https://doi.org/10.3390/axioms12111008

AMA Style

Jiang B, Huang H, Radenović S. Common Fixed Point of (ψ, β, L)-Generalized Contractive Mapping in Partially Ordered b-Metric Spaces. Axioms. 2023; 12(11):1008. https://doi.org/10.3390/axioms12111008

Chicago/Turabian Style

Jiang, Binghua, Huaping Huang, and Stojan Radenović. 2023. "Common Fixed Point of (ψ, β, L)-Generalized Contractive Mapping in Partially Ordered b-Metric Spaces" Axioms 12, no. 11: 1008. https://doi.org/10.3390/axioms12111008

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