This paper deals with the concept of α-ˆv-A-B-C-Meir-Keeler type nonlinear contractions, a new class mappings within the of modular extended b-metric spaces. We establish common unique fixed-point theorems that generalize, unify, and extend several key results in modular b-metric and modular extended b-metric spaces. These theorems bridge the gap between classical and contemporary fixed-point theories, showcasing broader applicability in nonlinear analysis. To ensure clarity and practical relevance, a detailed example is presented, further validating the theoretical findings. This work provides some level of understanding of the space under investigation and sets the stage for future developments in this evolving domain.
Citation: Daniel Francis, Godwin Amechi Okeke, Aviv Gibali. Another Meir-Keeler-type nonlinear contractions[J]. AIMS Mathematics, 2025, 10(4): 7591-7635. doi: 10.3934/math.2025349
[1] | Abdolsattar Gholidahneh, Shaban Sedghi, Ozgur Ege, Zoran D. Mitrovic, Manuel de la Sen . The Meir-Keeler type contractions in extended modular b-metric spaces with an application. AIMS Mathematics, 2021, 6(2): 1781-1799. doi: 10.3934/math.2021107 |
[2] | Erdal Karapınar, Andreea Fulga . Discussion on the hybrid Jaggi-Meir-Keeler type contractions. AIMS Mathematics, 2022, 7(7): 12702-12717. doi: 10.3934/math.2022703 |
[3] | Ma'mon Abu Hammad, Oualid Zentar, Shameseddin Alshorm, Mohamed Ziane, Ismail Zitouni . Theoretical analysis of a class of φ-Caputo fractional differential equations in Banach space. AIMS Mathematics, 2024, 9(3): 6411-6423. doi: 10.3934/math.2024312 |
[4] | Pragati Gautam, Vishnu Narayan Mishra, Rifaqat Ali, Swapnil Verma . Interpolative Chatterjea and cyclic Chatterjea contraction on quasi-partial b-metric space. AIMS Mathematics, 2021, 6(2): 1727-1742. doi: 10.3934/math.2021103 |
[5] | Shaoyuan Xu, Yan Han, Suzana Aleksić, Stojan Radenović . Fixed point results for nonlinear contractions of Perov type in abstract metric spaces with applications. AIMS Mathematics, 2022, 7(8): 14895-14921. doi: 10.3934/math.2022817 |
[6] | Hongyan Guan, Jinze Gou, Yan Hao . On some weak contractive mappings of integral type and fixed point results in b-metric spaces. AIMS Mathematics, 2024, 9(2): 4729-4748. doi: 10.3934/math.2024228 |
[7] | Muhammad Sarwar, Aiman Mukheimer, Syed Khayyam Shah, Arshad Khan . Existence of solutions of fractal fractional partial differential equations through different contractions. AIMS Mathematics, 2024, 9(5): 12399-12411. doi: 10.3934/math.2024606 |
[8] | Monica-Felicia Bota, Liliana Guran . Existence of a solution of fractional differential equations using the fixed point technique in extended b-metric spaces. AIMS Mathematics, 2022, 7(1): 518-535. doi: 10.3934/math.2022033 |
[9] | Monairah Alansari, Mohammed Shehu Shagari, Akbar Azam, Nawab Hussain . Admissible multivalued hybrid Z-contractions with applications. AIMS Mathematics, 2021, 6(1): 420-441. doi: 10.3934/math.2021026 |
[10] | Afrah Ahmad Noman Abdou . Chatterjea type theorems for complex valued extended b-metric spaces with applications. AIMS Mathematics, 2023, 8(8): 19142-19160. doi: 10.3934/math.2023977 |
This paper deals with the concept of α-ˆv-A-B-C-Meir-Keeler type nonlinear contractions, a new class mappings within the of modular extended b-metric spaces. We establish common unique fixed-point theorems that generalize, unify, and extend several key results in modular b-metric and modular extended b-metric spaces. These theorems bridge the gap between classical and contemporary fixed-point theories, showcasing broader applicability in nonlinear analysis. To ensure clarity and practical relevance, a detailed example is presented, further validating the theoretical findings. This work provides some level of understanding of the space under investigation and sets the stage for future developments in this evolving domain.
A major development in fixed-point theory was achieved by Meir and Keeler [32], who established a pivotal theorem that extends the renowned Banach contraction mapping principle.
Let (X,d) be a complete metric space and T:X→X be a mapping such that, for each ϵ>0, ∃ a δ(ϵ)>0 such that
ϵ≤d(x,y)<ϵ+δ(ϵ)⟹d(Tx,Ty)<ϵ |
for all x,y∈X. Then the fixed point of T in X is unique.
Meir and Keeler's fixed-point method [32] has become a fundamental topic in fixed-point theory. It has inspired extensive research, with numerous authors contributing new ideas and methods to further develop their work.
Maiti and Pal [31] introduced the following contraction mapping and provided it's validity.
For every ϵ>0, δ>0 exists such that
ϵ≤max{d(x,y),d(x,Ty),d(y,Ty)}<ϵ+δ⟹d(Tx,Ty)<ϵ. |
However, other researchers [36,39] enhanced the above result with the following condition:
ϵ≤max{d(Sx,Sy),d(Sx,Tx),d(Sy,Ty),12(d(Sx,Ty)+d(Sy,Tx))}<ϵ+δ⟹d(Tx,Ty)<ϵ,whereS,Tareselfmappingsinmetricspace(X,d). |
In 1976, Jungck [25] demonstrated a shared fixed-point result for commute mappings, assuming that one of them is continuous. Subsequently, in 1982, Sessa [41] introduced the concept of weakly commuting pairs of self-mappings and established a fixed-point theorem in complete metric spaces. Sharma [42] later presented novel results for weakly commuting mappings in such spaces, extending several related findings. Kumam et al. [30] explored fixed point results under generalized contractive conditions in b-metric spaces, providing an example to highlight the practical implications of their work. Moreover, He et al. [23] examined the existence and uniqueness of fixed-points for weakly commutative mappings within the framework of complete multiplicative metric spaces. Alsulami et al. [7] introduced α-admissible and generalized α-admissible Meir-Keeler contractions in quasi-metric spaces and extended their findings to G-metric spaces, proving fixed-point theorems. Abtahi [2] provided a criterion for sequences in metric spaces to be Cauchy, simplifying proofs of fixed-point results for Meir-Keeler-type contractions. Canzoneri and Vetroa [11] studied asymptotic contractions of the integral Meir-Keeler-type and established a fixed-point theorem ensuring existence and uniqueness. Gholamian and Khanehgir [20] introduced generalized Meir-Keeler contractions in b-metric-like spaces, proving fixed-point theorems with illustrative examples. Gulyaz et al. [22] extended the above work to α-Meir-Keeler and generalized α-Meir-Keeler contractions in Branciari b-metric spaces, establishing the existence and uniqueness of fixed-points. Karapinar et al. [26] studied (α-ψ)-Meir-Keeler contractions in generalized b-metric spaces, while Barootkoob et al. [10] introduced (α-ψ-p)-Meir-Keeler contractions, extending fixed point results via w-distance and applying them to nonlinear Fredholm integral equations. Panthi [35] proved common fixed-point theorems for compatible mappings in metric and dislocated metric spaces, and Koti et al. [29] established fixed-point results using Gupta-Saxena's rational expression. Additional results are documented in [15,16] and some other results on common fixed-point theory can be found in [17,18,33]. Further results in this area can also be found in [1,9,13]. Aksoya et al. [3] studied Meir-Keeler type contractions in modular metric spaces, proving fixed point theorems with examples. Further results on such contractions in partial Hausdorff and JS-metric spaces are found in [12,27]. Aksoy et al. [4] extended the fixed-point results to a broader class of contractive and non-expansive mappings in modular metric spaces. Jleli et al. [24] introduced proximal quasi-contractions in modular spaces with the Fatou property, establishing the best proximity point theorems. Aksoy et al. [5] explored α-admissible contractions in b-metric spaces, proving the fixed-point results and applying them to differential equations. Arshad et al. [8] generalized Jleli et al.'s work using triangular α-orbital admissible mappings [38], while Alharbi et al. [6] combined α-orbital admissibility with simulation functions to establish fixed-point results in b-metric spaces. Gholidahneh et al. [21] introduced modular extended b-metric spaces and established fixed point theorems for α ˆv Meir-Keeler contractions, extending their results to various settings, including graph structured and partially ordered spaces. They also linked fuzzy b metric spaces with modular extended b metric spaces and provided examples and applications to illustrate their findings. Modular extended b-metric spaces provide a powerful and flexible framework that extends classical metric spaces and enables the study of nonlinear contractions, multi-mapping systems, and function dependent fixed point problems. Their broad applicability, as seen in [21], makes them an essential tool in modern mathematical analysis. There are examples of extended b-modular metric spaces which are not classical, b-metric, modular, or modular b-metric spaces. Example 2.2 in [21] is an extended modular b-metric space which is not a classical metric or b-metric space.
We analyze whether the function introduced in [21, Example 2.2], namely
ˆνλ(x,y)=sinh(νλ(x,y)), |
defines a modular extended b-metric space and whether it satisfies the conditions of a classical metric or b-metric spaces.
1) Verification as an extended modular b-metric space
A modular extended b-metric space satisfies the following conditions:
● Non-negativity and identity of indiscernibles: Since sinh(t)≥0 for all t≥0 and sinh(0)=0, we have
ˆνλ(x,y)=0⟺νλ(x,y)=0⟺x=y. |
● Symmetry: Since νλ(x,y)=νλ(y,x), we obtain
ˆνλ(x,y)=sinh(νλ(x,y))=sinh(νλ(y,x))=ˆνλ(y,x). |
● Extended modular triangle inequality: Since νλ satisfies the modular b-metric inequality
νλ+μ(x,y)≤s(νλ(x,z)+νμ(z,y)), |
applying sinh gives
ˆνλ+μ(x,y)=sinh(νλ+μ(x,y))≤sinh(s(νλ(x,z)+νμ(z,y))). |
Defining Ω(t)=sinh(st), which is strictly increasing, we obtain
ˆνλ+μ(x,y)≤Ω(ˆνλ(x,z)+ˆνμ(z,y)). |
Thus, ˆνλ satisfies the modular extended b-metric conditions.
2) Not a classical metric space
A classical metric satisfies the triangle inequality
d(x,z)≤d(x,y)+d(y,z). |
However, the function ˆνλ satisfies
sinh(a+b)≠sinh(a)+sinh(b), |
which means that the standard triangle inequality does not hold. Hence, ˆνλ is not a classical metric space.
3) Not a b-metric space
A b-metric satisfies the inequality
d(x,z)≤s(d(x,y)+d(y,z)). |
However, for ˆνλ, we have
ˆνλ(x,z)≤sinh(s(νλ(x,y)+νμ(y,z))), |
which does not match the linear form of the b-metric inequality. Since sinh is non-linear, ˆνλ does not satisfy the b-metric condition.
● ˆνλ is an extended modular b-metric space, since it satisfies the extended modular triangle inequality.
● ˆνλ is not a classical metric space, as it violates the standard triangle inequality.
● ˆνλ is not a b-metric space, as it does not satisfy the linear b-metric triangle inequality.
Okeke et al. [34] further generalized these concepts by defining new types of α-ˆv-Meir-Keeler-type contractions in modular extended b-metric spaces, supported by examples that validated their results.
This paper builds on these works by presenting the concept of α-ˆv-A-B-C-Meir-Keeler-type nonlinear contractions, a new class of mappings within modular extended b-metric spaces. We establish common unique fixed-point theorems that generalize, unify, and extend several key results in modular b-metric and modular extended b-metric spaces. These theorems bridge the gap between classical and contemporary fixed-point theories, showcasing broader applicability in nonlinear analysis. An example is provided to support the findings.
The structure of this work is organized as follows: Initially, we review the fundamental definitions and results in Section 2. Subsequently, the core findings are introduced and examined in Section 3.
Definition 1. [30,41] Let f and g be mappings from a b-metric space (X,d) onto itself. The mappings f and g are called weakly commuting if d(fgx,gfx)≤d(fx,gx) for each x∈X.
Definition 2. [30] Let f and g be mappings from a b-metric space (X,d) onto itself. The mappings f and g are termed R-weakly commuting if a positive real number R exists such that d(fgx,gfx)≤Rd(fx,gx) for each x∈X.
Definition 3. [40] Let T be a self-mapping on X and α:Xˆv×Xˆv→[0,+∞) be a function. Then T is called an α-admissible mapping if, for all x,y∈X,
α(x,y)≥1⟹α(Tx,Ty)≥1. |
Definition 4. [28] Let α:Xˆv×Xˆv→[0,+∞) be a mapping. Then the self-mapping f:X→X is said to be triangular α-admissible if
(1) for all x,y∈X, α(x,y)≥1⟹α(fx,fy)≥1;
(2) for all x,y,z∈X, α(x,z)≥1andα(z,y)≥1⟹α(x,y)≥1.
Lemma 1. [28] Let f be a triangular α-admissible mapping. Assume that x0∈X exists such that α(x0,fx0)≥1. Define a sequence {xn}n≥1 as xn=fnx0. Then α(xm,xn)≥1 ∀ distinct n,m∈N.
Lemma 2. [38] Let f be a triangular α-orbital admissible mapping. Assume that x1∈X exists such that α(x1,fx1)≥1. Define a sequence {xn}n≥1 by xn+1=fxn. Then we have α(xm,xn)≥1 ∀ distinct n,m∈N.
Definition 5. [38] (a) Let h:X→X be a mapping and α:X×X→R be a function. Then h is said to be α-orbital admissible if α(x,hx)≥1⟹α(hx,h2x)≥1.
(b) Let h:X→X be a map and α:x×X→R be a function. Then h is termed a triangular α-orbital admissible mapping if α(x,hx)≥1⟹α(hx,h2x)≥1, α(x,y)≥1, and α(y,hy)≥1⟹α(x,hy)≥1.
It is evident that every mapping that is α-admissible also qualifies as an α-orbital admissible mapping. Additionally, any triangular mapping that is α-admissible is inherently a triangular α-orbital admissible mapping as well. However, an instance of a triangular α-orbital admissible mapping exists that does not conform to the criteria of being triangular α-admissible. An example of this can be found in [38]. A b-metric space serves as a natural extension of both classical metric space by modifying the well-known triangle inequality to the form d(x,z)≤s(d(x,y)+d(y,z)) for all points x,y,z∈X and a fixed parameter s≥1. Recently, the concept of a b-metric space has been further extended to encompass p-metric spaces, as elaborated in [37]. However, the p above is not a partial metric space.
Definition 6. [37] Let X be a non-empty set. A mapping d:X×X→R+ is called a p-metric if a strictly increasing continuous function Ω:[0,∞)→[0,∞) exists with t≤Ω(t) for all t∈[0,∞) such that, for all x,y,z∈X, the following conditions hold:
(1) d(x,y)=0 if and only if x=y;
(2) d(x,y)=d(y,x);
(3) d(x,z)≤Ω(d(x,y)+d(y,z)).
The pair (X,d) is called a p-metric space or an extended b-metric space.
Definition 7. [14] Let X be a non-empty set. A mapping ω:(0,+∞)×X×X→R+∪{∞} is called a metric modular on X if, for all x,y,z∈X and λ>0, the following conditions hold:
(1) ωλ(x,y)=0 for all λ>0 if and only if x=y;
(2) ωλ(x,y)=ωλ(y,x) for all λ>0;
(3) ωλ+μ(x,y)≤ωλ(x,z)+ωμ(z,y) for al λ, μ>0.
The pair (X,ω) is called a modular metric space.
Definition 8. [14] Let (X,ω) be a modular metric space. Fix x0∈X and set
Xω=Xω(x0)={x∈X:ωλ(x,x0)=0,asλ→∞} |
and
X∗ω=X∗ω(x0)={x∈X:ωλ(x,x0)<∞,forallλ>0}. |
In this context, the sets Xω and X∗ω are referred to as modular spaces that are based on the point x0.
Definition 9. [19] Let X be a non-empty set and s≥1 be a real number. A mapping ω:(0,+∞)×X×X→R+∪{∞} is called a modular b-metric on X if the following statements hold: for all x,y,z∈X:
(1) ωλ(x,y)=0 for all λ>0 if and only if x=y;
(2) ωλ(x,y)=ωλ(y,x) for all λ>0;
(3) ωλ+μ(x,y)≤s(ωλ(x,z)+ωμ(z,y)) for all λ,μ>0.
The pair (X,ω) is called a modular b-metric space.
In the paper, take Xˆv=X∗ˆv(x0)={x∈X:ˆvλ(x,x0)<∞,forallλ>0}.
Definition 10. [21] Let X be a non-empty set. A mapping ˆvλ:(0,∞)×X×X→[0,∞] is called a modular extended b-metric if a strictly increasing continuous function A:[0,∞)→[0,∞) with A−1(t)≤t≤A(t) for all t∈[0,∞) exists such that, for all x,y,z∈X, the following conditions hold:
(1) ˆvλ(x,y)=0 if and only if x=y for all λ>0;
(2) ˆvλ(x,y)=ˆvλ(y,x) for all λ>0;
(3) ˆvλ+μ(x,y)≤A(ˆvλ(x,z)+ˆvμ(z,y)) for all λ,μ>0.
Then the pair (X,ˆv) is called a modular extended b-metric space.
The class of modular extended b-metric spaces is bigger than the class of b-metric spaces, since a b-metric space is modular extended b-metric space whenever A(t)=st and a metric space if it is a modular extended b-metric space with A(t)=t.
Example 1. [21] Consider Xω to be a modular space equipped with a b-metric, where the coefficient satisfies s≥1 and ˆvλ(x,y)=sinh(ωλ(x,y)). Then ˆvλ is a modular extended b-metric space with A(t)=sinh(st) for all t≥1 and A−1(r)=1ssinh−1(r) for all r≥0.
Definition 11. [21] Let Xˆv be a modular extended b-metric space. Then a sequence {xn}n≥1⊆Xˆv is called
(1) a ˆv-Cauchy sequence if, for all ϵ>0, n(ϵ)∈N exists such that, for each n,m≥n(ϵ) and λ>0, ˆvλ(xn,xm)<ϵ;
(2) ˆv-convergent to x∗∈Xˆv if ˆvλ(xn,x∗)⟶0 as n→∞ for all λ>0;
(3) ˆv-complete if each ˆv-Cauchy sequence in Xˆv is ˆv-convergent and its limit is in Xˆv.
Definition 12. [34] Let Xˆv be a modular extended b-metric space, T:Xˆv→Xˆv be a mapping, and x0∈Xˆv. T is said to be orbitally continuous at x0 whenever limk→∞ˆvλ(Tnku,x0)=0 implies that limk→∞ˆvλ(TTnku,Tx0)=0 whenever u∈Xˆv and {nk}⊆N is a strictly increasing sequence of non-negative integer numbers.
Definition 13. Let Xˆv be a modular extended b-metric space, T1,T2:Xˆv→Xˆv be mappings, and x∈Xˆv. Then T1,T2 are called modular extended b-weakly commuting mappings in Xˆv if
ˆvλ(T1T2x,T2T1x)≤ˆvλ(T2x,T1x). |
Example 2. Consider the set X=(R+∖{0})∪{∞} equipped with the modular extended b-metric defined by
ˆvλ(x,y):=11+λmaxx,y∈Xˆv{||x−y||}, |
which is complete on Xˆv for all λ>0. Define the mappings T1,T2:(R+∖{0})∪{∞}→(R+∖{0})∪{∞} as T1x=log32x5, T2x=log8x3 for all x∈(R+∖{0})∪{∞} and λ>0, respectively. Then T1 and T2 are ˆv-weakly commuting mappings.
In fact, it suffices to show that T1 and T2 satisfy Definition 13. For all λ>0 and x∈(R+∖{0}), we show that
ˆvλ(T2T4x,T4T2x)≤ˆvλ(T4x,T2x). |
Using the above definition, we get,
11+λmaxx∈Xˆv{||T1T2x−T2T1x||}=11+λmaxx∈Xˆv{||log32(log8x3)5−log8(log32x5)3||}=11+λmaxx∈Xˆv{||5log32(log8x3)−3log8(log32x5)||}=11+λmaxx∈Xˆv||ln(log8x3)ln(2)−ln(log32x5)ln(2)||=1ln(2)11+λmaxx∈Xˆv||ln(log8x3)−ln(log32x5)||=1ln(2)11+λmaxx∈Xˆv||ln(log8x3log32x5)||=1ln(2)11+λmaxx∈Xˆv||ln(3log8x5log32x)||=0. |
Again, we have
ˆvλ(T2x,T1x)=11+λmaxx∈Xˆv{||T2x−T1x||}=11+λmaxx∈Xˆv{||log8x3−log32x5||}=11+λ1ln(2)maxx∈Xˆv{||ln(x)−ln(x)||}=0. |
Thus we have
ˆvλ(T1T2x,T2T1x)≤ˆvλ(T2x,T1x), |
which shows that T1 and T2 are weakly commuting.
We construct an example of the main result below related to Example 2 above.
Remark 1. For all λ>0 and x,y∈Xˆv, we have
ˆvλ(T1x,T2y)≤FAλ(T1x,T2y),ˆvλ(T1x,T3y)≤FAλ(T1x,T3y),ˆvλ(T2x,T3y)≤FAλ(T2x,T3y). |
For the details of this remark, see [34, Remark 3.3].
The definition below is essential throughout this paper.
Definition 14. Ti:Xˆv⟶Xˆv be six continuous mappings. We say that Ti for i=1,2,⋯,6 follows an α-ˆv-A-B-C -Meir-Keeler-type contraction, and {T2,T4}, {T3,T5} and {T1,T6} are weakly commuting pairs of self-mappings such that T3(Xˆv)⊆T4(Xˆv),T2(Xˆv)⊆T5(Xˆv),T1(Xˆv)⊆T6(Xˆv), if a function α:Xˆv×Xˆv→[0,+∞), a≠0, a<1, b≠0, acb<1 and x0∈Xˆv exists such that α(x0,Tx0)≥1, and, for each i=1,⋯,6, the mapping Ti qualifies as triangular α-orbital admissible function for every λ,ϵ,δ>0 satisfying the following conditions:
α(x,y)ˆvλ(T1x,T2y)<A(ϵ)⟹ϵ≤A−1(FAλ(T1x,T2y))<A(ϵ)+A(δ(ϵ)), | (2.1) |
α(x,y)ˆvλ(T1x,T3y)<A(ϵ)⟹ϵ≤A−1(FBλ(T1x,T3y))<A(ϵ)+A(δ(ϵ)), | (2.2) |
α(x,y)ˆvλ(T2x,T3y)<A(ϵ)⟹ϵ≤A−1(FCλ(T2x,T3y))<A(ϵ)+A(δ(ϵ)), | (2.3) |
where
FAλ(T1x,T2y):=amax{ˆvλ(T6x,T4y),ˆvλ(T1x,T6y),ˆvλ(T3x,T4y),ˆvλ(T2x,T5y)}, | (2.4) |
FBλ(T1x,T3y):=bmax{ˆvλ(T5x,T4y),ˆvλ(T2x,T5y),ˆvλ(T1x,T26x),ˆvλ(T3y,T24y)}, | (2.5) |
FCλ(T2x,T3y):=cmax{ˆvλ(T6x,T2x),ˆvλ(T2x,T5y),ˆvλ(T3x,T4y),ˆvλ(T1x,T6y)}. | (2.6) |
Remark 2. ∙ Following Definitions 10, A is a strictly increasing continuous function A:[0,∞)→[0,∞) with A−1(ϵ)≤ϵ≤A(ϵ) for all ϵ∈[0,∞).
∙ FAλ(T1x,T2y), FBλ(T1x,T3y), and FCλ(T2x,T3y) are functions containing some mixed-metric extended modular space at the Ath, Bth, and Cth levels, respectively, for λ>0.
∙ An example of Definition 14 will be given after the proof of Theorem 1 below.
Definition 14 introduces the concept of α-ˆv-A-B-C-Meir-Keeler-type contraction mappings in a modular extended b-metric space, where six self- mappings interact under certain admissibility and commutativity conditions. We analyze how it builds upon Definitions 1–12.
Definitions 1 and 2 introduce weakly commuting mappings in b-metric spaces, which are crucial for Definition 14 because the contraction conditions require specific pairs of mappings {T2,T4},{T3,T5},{T1,T6} to be weakly commuting.
● Definition 1 (weakly commuting mappings) states that two mappings f,g are weakly commuting if d(fgx,gfx)≤d(fx,gx). This ensures some level of interaction between the mappings, which is a necessary condition in Definition 14.
● Definition 2 (R-weakly commuting mappings) extends this concept by introducing a control parameter R, leading to d(fgx,gfx)≤Rd(fx,gx). This generalization helps establish the contraction conditions in Definition 14.
● Definition 3 (α-admissibility of a mapping) defines an auxiliary function α(x,y) that controls the behavior of a mapping T in relation to the fixed-point process. Definition 14 relies on this property, since the contraction conditions depend on triangular α-admissibility.
● Definition 4 (triangular α-admissible mappings) extends the concept of admissibility to triangular structures, ensuring that if α(x,z) and α(z,y) hold, then so does α(x,y). This property is crucial in Definition 14 for handling sequences within the modular extended b-metric space.
● Definitions 5 (orbital and triangular orbital admissibility) provides further generalizations that influence the structure of Definition 14. Since the mappings in Definition 14 must satisfy orbital admissibility, these preliminary definitions establish the conditions necessary for the contraction framework.
The transition from metric and b-metric spaces to a modular extended b-metric space is key in Definition 14.
● Definition 6 (extended b metric space) introduces the concept of an extended b-distance function d, which satisfies d(x,y)≤Ω(d(x,z)+d(z,y)). This idea is extended in Definition 14 to work with six mappings instead of one, for a strictly increasing continuous function, Ω:[0,∞)→[0,∞) with t≤Ω(t) for every t∈[0,∞).
● Definition 7 (modular metric space) introduces the concept of a modular distance function ω, which satisfies ωλ+μ(x,y)≤ωλ(x,z)+ωμ(z,y). This idea is extended in Definition 14 to work with six mappings instead of one.
● Definition 8 (modular spaces Xω and X∗ω) introduces two types of modular metric spaces based on growth conditions. These modular structures are embedded into Definition 14 to ensure the existence of fixed points in a well-defined modular space.
● Definition 9 (modular b-metric space) extends the modular metric concept to b-metric spaces, introducing a relaxation factor s similar to the classic b-metric condition. This generalization is needed in Definition 14 because the contraction inequalities involve a max operation, requiring a structure that supports modular-b-metric behavior.
● Definition 10 (modular extended b-metric space) is one of the most crucial precursors to Definition 14. It defines the function ˆvλ that satisfies: ˆvλ+μ(x,y)≤A(ˆvλ(x,z)+ˆvμ(z,y)), where A is a strictly increasing function. Definition 14 applies this framework to analyze contractions under modular extended b-metric settings.
Definitions 11 and 12 introduce key sequence properties that Definition 14 relies upon to ensure convergence and the fixed points existence.
● Definition 11 (ˆv-Cauchy sequences and ˆv-convergence) formalizes when a sequence is Cauchy and convergent in modular extended b-metric spaces. In Definition 14, the iterative sequences
ξn=T1xn=T6xn+1,ξn+1=T2xn+1=T5xn+2,ξn+2=T3xn+2=T4xn+3, | (2.7) |
must be ˆv-Cauchy to ensure the fixed points results hold.
● Definition 12 (orbital continuity in modular extended b-metric spaces) establishes the continuity conditions needed for taking the limits in Definition 14. Since the proof of Theorem 3 relies on passing those limits, these conditions ensure the stability of the mappings under iteration.
Definition 14 unifies and extends all previous definitions by combining the following:
● Weakly commuting mappings (Definitions 1, 2, and 13).
● α-admissibility conditions (Definitions 3–5).
● Extended b-metric space and modular extended b-metric properties (Definitions 6–10).
● Convergence and continuity results (Definitions 11–12).
Now, we give the main results in this paper. We start with the following lemma.
Lemma 3. Let Xˆv be a ˆv-regular ˆv-complete modular extended b-metric space and let Ti:Xˆv⟶Xˆv be six orbitally continuous mappings satisfying the α-ˆv-A-B-C-Meir-Keeler-type contraction for i=1,2,⋯,6, {T2,T4}, {T3,T5} and {T1,T6} be weakly commuting pairs of self-mappings such that T3(Xˆv)⊆T4(Xˆv),T2(Xˆv)⊆T5(Xˆv),T1(Xˆv)⊆T6(Xˆv). We then have a function α:Xˆv×Xˆv→[0,+∞), a≠0, a<1, b≠0, acb<1 and x0∈Xˆv such that α(x0,x1)≥1,ϵ,δ>0 and, for each i=1,⋯,6, Ti remains a triangular α-orbital admissible mapping for every λ>0, provided the following conditions are met:
α(x,y)ˆvλ(T1x,T2y)<A(ϵ)⟹ϵ≤A−1(FAλ(T1x,T2y))<A(ϵ)+A(δ(ϵ)), | (3.1) |
α(x,y)ˆvλ(T1x,T3y)<A(ϵ)⟹ϵ≤A−1(FBλ(T1x,T3y))<A(ϵ)+A(δ(ϵ)), | (3.2) |
α(x,y)ˆvλ(T2x,T3y)<A(ϵ)⟹ϵ≤A−1(FCλ(T2x,T3y))<A(ϵ)+A(δ(ϵ)), | (3.3) |
where
FAλ(T1x,T2y):=amax{ˆvλ(T6x,T4y),ˆvλ(T1x,T6y),ˆvλ(T3x,T4y),ˆvλ(T2x,T5y)}, | (3.4) |
FBλ(T1x,T3y):=bmax{ˆvλ(T5x,T4y),ˆvλ(T2x,T5y),ˆvλ(T1x,T26x),ˆvλ(T3y,T24y)}, | (3.5) |
FCλ(T2x,T3y):=cmax{ˆvλ(T6x,T2x),ˆvλ(T2x,T5y),ˆvλ(T3x,T4y),ˆvλ(T1x,T6y)}. | (3.6) |
Suppose that {xn}n∈N and {ξn}n∈N are sequences in Xˆv so that for xn in Xˆv, we choose xn+1 such that ξn=T1xn=T6xn+1; again for xn+1 in Xˆv, we choose xn+2 such that ξn+1=T2xn+1=T5xn+2 and, for a point xn+2 in Xˆv, we choose xn+3 such that ξn+2=T3xn+2=T4xn+3 for n=0,1,2,⋯. Then ˆvλ(ξn,ξm)=0∀λ>0.
Proof. Suppose that Xˆv is empty. In that case, there is nothing to prove. We now assume that Xˆv≠∅. Suppose that the mappings, Ti for i=1,⋯,6 satisfy the inequalities in (3.1)–(3.6). Since x0,x1, and x2 are points in Xˆv and T1(Xˆv)⊆T6(Xˆv), we can find a point x1 in Xˆv such that ξ0=T1x0=T6x1. For T2(Xˆv)⊆T5(Xˆv), we can find a point x2∈Xˆv such that ξ1=T2x1=T5x2 and for T3(Xˆv)⊆T4(Xˆv), we can find a point x3 in Xˆv such that ξ2=T3x2=T4x3. Now for all λ>0, we induce on n, so that there are sequences {xn}n∈N and {ξn}n∈N within Xˆv that satisfy the following equations:
ξn=T1xn=T6xn+1,ξn+1=T2xn+1=T5xn+2,ξn+2=T3xn+2=T4xn+3. | (3.7) |
If n0∈N exists such that ξn0=ξn0+1, T3(Xˆv)⊆T4(Xˆv), T2(Xˆv)⊆T5(Xˆv), T1(Xˆv)⊆T6(Xˆv) holds. In fact, if m∈N exists such that ξm+2=ξm+3, then T1u=T6u, where u=xm+3. Therefore, the pair {T1,T6} has a coincidence point u∈Xˆv. If ξm=ξm+1, then T2u=T4u, where u=xm+1. Therefore, the pair {T2,T4} has a coincidence point u∈Xˆv. If ξm+1=ξm+3, then T3u=T5u, where u=xm+2. Thus, the pair {T3,T5} has a coincidence point u∈Xˆv. Again, if there is an n0∈N such that ξn0=ξn0+1=ξn0+2, then ξn=ξn0 for any n≥n0. This implies that {ξn} is a modular ˆv Cauchy sequence in Xˆv. Actually, if η∈N exists such that (1) ξη=ξη+1=ξη+2, (2) ξη≠ξη+1=ξη+2, (3) ξη≠ξη+2=ξη+1, and (4) ξη≠ξη+1≠ξη+2 hold. In fact the Case (1) is easy, and Case (3) is similar to Case (2); then from inequality (3.1)–(3.6), we can set x=ξη+2 and y=ξη+3. Thus Case (2) stipulates that;
α(ξη+2,ξη+3)ˆvλ(T1ξη+2,T2ξη+3)<A(ϵ)⟹ϵ≤A−1(FAλ(T1ξη+2,T2ξη+3))<A(ϵ)+A(δ(ϵ)), | (3.8) |
α(ξη+2,ξη+3)ˆvλ(T1ξη+2,T3ξη+3)<A(ϵ)⟹ϵ≤A−1(FBλ(T1ξη+2,T3ξη+3))<A(ϵ)+A(δ(ϵ)), | (3.9) |
α(ξη+2,ξη+3)ˆvλ(T2ξη+2,T3ξη+3)<A(ϵ)⟹ϵ≤A−1(FCλ(T2ξη+2,T3ξη+3))<A(ϵ)+A(δ(ϵ)), | (3.10) |
where
FAλ(T1ξη+2,T2ξη+3):=amax{ˆvλ(T6ξη+2,T4ξη+3),ˆvλ(T1ξη+2,T6ξη+3),ˆvλ(T3ξη+2,T4ξη+3),ˆvλ(T2ξη+2,T5ξη+3)}, | (3.11) |
FBλ(T1ξη+2,T3ξη+3):=bmax{ˆvλ(T5ξη+2,T4ξη+3),ˆvλ(T2ξη+2,T5ξη+3),ˆvλ(T1ξη+2,T26ξη+2),ˆvλ(T3ξη+3,T24ξη+3)}, | (3.12) |
FCλ(T2ξη+2,T3ξη+3):=cmax{ˆvλ(T6ξη+2,T2ξη+2),ˆvλ(T2ξη+2,T5ξη+3),ˆvλ(T3ξη+2,T4ξη+3),ˆvλ(T1ξη+2,T6ξη+3)}. | (3.13) |
Using inequality (3.8), Eqs (3.11) and (3.7), we get
ˆvλ(T1ξη+2,T2ξη+3)>0,ˆvλ(T1ξη+1,T3ξη+2)>0,ˆvλ(T2ξη,T3ξη+1)>0 |
for all λ>0. According to Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, we get 1≤α(η,η+1) for all η,η+1∈N with η<η+1.
Consider ˆvλ(T1ξη+2,T2ξη+3)>0 for all η∈N∪{0}. Since T1 and T2 are triangular α-orbital admissible mappings, it follows from ˆv-regularity that, for all λ>0, ˆvλ(T1ξη+2,T2ξη+3)>0 and so, by Remark 1, we get
A−1(FAλ(T1ξη+2,T2ξη+3)≥ˆvλ(T1ξη+2,T2ξη+3), |
A−1(FBλ(T1ξη+1,T3ξη+2)≥ˆvλ(T1ξη+1,T3ξη+2), |
and
A−1(FCλ(T2ξη,T3ξη+1)≥ˆvλ(T2ξη,T3ξη+1), |
for all η∈N∪{0}. Therefore, we have
A−1(FAλ(T1ξη+2,T2ξη+3))=amax{ˆvλ(T6ξη+2,T4ξη+3),ˆvλ(T1ξη+2,T6ξη+3),ˆvλ(T3ξη+2,T4ξη+3),ˆvλ(T2ξη+2,T5ξη+3)}=amax{ˆvλ(T1ξη+1,T3ξη+2),ˆvλ(T1ξη+2,T1ξη+2),ˆvλ(T3ξη+2,T3ξη+2),ˆvλ(T2ξη+2,T2ξη+2)}=aˆvλ(T1ξη+1,T3ξη+2). | (3.14) |
Again, from Eq (3.14) and for all λ>0, according to Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, 1≤α(η+1,η+2) for all η+1,η+2∈N with η+1<η+2, using inequality (3.2) and Eqs (3.5) and (3.7), we get
α(ξη+1,ξη+2)ˆvλ(T1ξη+1,T3ξη+2)=bamax{ˆvλ(T5ξη+1,T4ξη+2),ˆvλ(T2ξη+1,T5ξη+2),ˆvλ(T1ξη+1,T26ξη+1),ˆvλ(T3ξη+2,T24ξη+2)}=bamax{ˆvλ(T2ξη,T3ξη+1),ˆvλ(T2ξη+1,T2ξη+1),ˆvλ(T1ξη+1,T1ξη+1),ˆvλ(T3ξη+2,T3ξη+2)}=baˆvλ(T2ξη,T3ξη+1). | (3.15) |
Finally, from Eq (3.15) and for all λ>0, according to Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, 1≤α(η,η+1) for all η,η+1∈N with η<η+1, using inequality (3.3) and Eqs (3.6) and (3.7), we get
α(ξη,ξη+1)ˆvλ(T2ξη,T3ξη+1)=acbmax{ˆvλ(T6ξη,T2ξη),ˆvλ(T2ξη,T5ξη+1),ˆvλ(T3ξη,T4ξη+1),ˆvλ(T1ξη,T6ξη+1)}=acbmax{ˆvλ(T1ξη−1,T2ξη),ˆvλ(T2ξη,T2ξη),ˆvλ(T3ξη,T3ξη),ˆvλ(T1ξη,T1ξη)}=acbˆvλ(T1ξη−1,T2ξη)=acbˆvλ(ξη−1,ξη). | (3.16) |
So, it follows from (3.1)–(3.3) and (3.14)–(3.16) that, for all λ>0,
α(ξη+2,ξη+3)ˆvλ(T1ξη+2,T2ξη+3)≤aα(ξη+1,ξη+2)ˆvλ(T1ξη+1,T3ξη+2)≤baα(ξη,ξη+1)ˆvλ(T2ξη,T3ξη+1)≤acbα(ξη,ξη+1)ˆvλ(T1ξη−1,T2ξη), | (3.17) |
and hence, using Eq (3.7), we get
α(ξη+2,ξη+3)ˆvλ(ξη+2,ξη+3)≤aα(ξη+1,ξη+2)ˆvλ(ξη+1,ξη+2)≤baα(ξη,ξη+1)ˆvλ(ξη,ξη+1)≤acbα(ξη,ξη+1)ˆvλ(ξη−1,ξη)=hα(ξη,ξη+1)ˆvλ(ξη−1,ξη). | (3.18) |
Therefore
α(ξη+2,ξη+3)ˆvλ(ξη+2,ξη+3)≤aα(ξη+1,ξη+2)ˆvλ(ξη+1,ξη+2)≤baα(ξη,ξη+1)ˆvλ(ξη,ξη+1)≤acbα(ξη,ξη+1)ˆvλ(ξη−1,ξη)⋮≤hn+1α(ξ0,ξ1)ˆvλ(ξη−1,ξη)=hn+1α(ξ0,ξ1)ˆvλ(ξ0,ξ1). | (3.19) |
Therefore, after some algebra and using condition (3) of Definition 10 we get
ˆvλ(xn,xm)=0∀λ>0,n≥m. | (3.20) |
This is a modular extended ˆv-Cauchy sequence in ˆv-complete modular extended b-metric space.
Again, for ξη≠ξη+1≠ξη+2, take x=ξη and y=ξη+1, from Case (4). We then have
α(ξη,ξη+1)ˆvλ(T1ξη,T2ξη+1)<A(ϵ)⟹ϵ≤A−1(FAλ(T1ξη,T2ξη+1))<A(ϵ)+A(δ(ϵ)), | (3.21) |
α(ξη,ξη+1)ˆvλ(T1ξη,T3ξη+1)<A(ϵ)⟹ϵ≤A−1(FBλ(T1ξη,T3ξη+1))<A(ϵ)+A(δ(ϵ)), | (3.22) |
α(ξη,ξη+1)ˆvλ(T2ξη,T3ξη+1)<A(ϵ)⟹ϵ≤A−1(FCλ(T2ξη,T3ξη+1))<A(ϵ)+A(δ(ϵ)), | (3.23) |
where
FAλ(T1ξη,T2ξη+1):=amax{ˆvλ(T6ξη,T4ξη+1),ˆvλ(T1ξη,T6ξη+1),ˆvλ(T3ξη,T4ξη+1),ˆvλ(T2ξη,T5ξη+1)}, | (3.24) |
FBλ(T1ξη,T3ξη+1):=bmax{ˆvλ(T5ξη,T4ξη+1),ˆvλ(T2ξη,T5ξη+1),ˆvλ(T1ξη,T26ξη),ˆvλ(T3ξη+1,T24ξη+1)}, | (3.25) |
FCλ(T2ξη,T3ξη+1):=cmax{ˆvλ(T6ξη,T2ξη),ˆvλ(T2ξη,T5ξη+1),ˆvλ(T3ξη,T4ξη+1),ˆvλ(T1ξη,T6ξη+1)}. | (3.26) |
Using inequality (3.21), Eqs (3.24) and (3.7), we get
ˆvλ(T1ξη,T2ξη+1)>0,ˆvλ(T1ξη−1,T3ξη)>0,ˆvλ(T2ξη−2,T3ξη−1)>0 |
for all λ>0. According to Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, we get 1≤α(η,η+1) for all η,η+1∈N with η≠η+1.
Consider ˆvλ(T1ξη,T2ξη+1)>0 for all η∈N∪{0}. Since T1 and T2 are triangular α-orbital admissible mappings, it follows from ˆv-regularity that, for all λ>0, ˆvλ(T1ξη,T2ξη+1)>0 and so, by Remark 1, we get
A−1(FAλ(T1ξη,T2ξη+1)≥ˆvλ(T1ξη,T2ξη+1), |
A−1(FBλ(T1ξη−1,T3ξη)≥ˆvλ(T1ξη−1,T3ξη), |
and
A−1(FCλ(T2ξη−2,T3ξη−1)≥ˆvλ(T2ξη−2,T3ξη−1), |
for all η∈N∪{0}. Therefore, we have
A−1(FAλ(T1ξη,T2ξη+1))=amax{ˆvλ(T6ξη,T4ξη+1),ˆvλ(T1ξη,T6ξη+1),ˆvλ(T3ξη,T4ξη+1),ˆvλ(T2ξη,T5ξη+1)}=amax{ˆvλ(T1ξη−1,T3ξη),ˆvλ(T1ξη,T1ξη),ˆvλ(T3ξη,T3ξη),ˆvλ(T2ξη,T2ξη)}=aˆvλ(T1ξη−1,T3ξη). | (3.27) |
Again, from Eq (3.27) and for all λ>0, and Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, 1≤α(η−1,η) for all η−1,η∈N with η−1≠η, using inequality (3.2) and Eqs (3.5) and (3.7), we get
α(ξη−1,ξη)ˆvλ(T1ξη−1,T3ξη)=bamax{ˆvλ(T5ξη−1,T4ξη),ˆvλ(T2ξη−1,T5ξη),ˆvλ(T1ξη−1,T26ξη),ˆvλ(T3ξη,T24ξη)}=bamax{ˆvλ(T2ξη−2,T3ξη−1),ˆvλ(T2ξη−1,T2ξη−1),ˆvλ(T1ξη−1,T1ξη−1),ˆvλ(T3ξη,T3ξη)}=baˆvλ(T2ξη−2,T3ξη−1). | (3.28) |
Finally, from Eq (3.28) and for all λ>0, and Lemma 2, given that Ti functions as triangular α-orbital admissible mapping, for each i=1,⋯,6, 1≤α(η−1,η−2) for all η−1,η−2∈N with η−1≠η−2, using inequality (3.3) and Eqs (3.6) and (3.7), we get
α(ξη−2,ξη−1)ˆvλ(T2ξη−2,T3ξη−1)=acbmax{ˆvλ(T6ξη−2,T2ξη−2),ˆvλ(T2ξη−2,T5ξη−1),ˆvλ(T3ξη−2,T4ξη−1),ˆvλ(T1ξη−2,T6ξη−1)}=acbmax{ˆvλ(T1ξη−3,T2ξη−2),ˆvλ(T2ξη−2,T2ξη−2),ˆvλ(T3ξη−2,T3ξη−2),ˆvλ(T1ξη−2,T1ξη−2)}=acbˆvλ(T1ξη−3,T2ξη−2)=acbˆvλ(ξη−3,ξη−2). | (3.29) |
So, it follows from (3.1)–(3.3) and (3.27)–(3.29) that, for all λ>0,
α(ξη,ξη+1)ˆvλ(T1ξη,T2ξη+1)≤aα(ξη−1,ξη)ˆvλ(T1ξη−1,T3ξη)≤baα(ξη−2,ξη−1)ˆvλ(T2ξη−2,T3ξη−1)≤acbα(ξη−3,ξη−2)ˆvλ(T1ξη−3,T2ξη−2), | (3.30) |
Hence, using Eq (3.7), we get
α(ξη,ξη+1)ˆvλ(ξη,ξη+1)≤aα(ξη−1,ξη)ˆvλ(ξη−1,ξη)≤baα(ξη−2,ξη−1)ˆvλ(ξη−2,ξη−1)≤acbα(ξη−3,ξη−2)ˆvλ(ξη−3,ξη−2)=hα(ξη−3,ξη−2)ˆvλ(ξη−3,ξη−2). | (3.31) |
Therefore
α(ξη,ξη+1)ˆvλ(ξη,ξη+1)≤aα(ξη−1,ξη)ˆvλ(ξη−1,ξη)≤baα(ξη−2,ξη−1)ˆvλ(ξη−2,ξη−1)≤acbα(ξη−3,ξη−2)ˆvλ(ξη−3,ξη−2)=hα(ξη−3,ξη−2)ˆvλ(ξη−3,ξη−2)⋮≤hn+1α(ξ0,ξ1)ˆvλ(ξ0,ξ1). | (3.32) |
Therefore, after some algebra and condition (3) of Definition 10, we get the result. Hence,
ˆvλ(ξn,ξm)=0∀λ>0,n≥m. | (3.33) |
Remark 3. Suppose that u is the common fixed point of Ti for i=1,2,⋯,6 when either T2 or T4 is ˆv-continuous and the pair {T2,T4} is weakly commuting. Again suppose that u is true for Ti, i=1,2,⋯,6 when T1 is ˆv-continuous, it is also true when T2 or T6 is ˆv continuous and the pair {T1,T6} is weakly commuting. Furthermore, T3 or T5 is ˆv-continuous and the pair {T3,T5} is weakly commuting.
Theorem 1. Suppose that Lemma 3 holds. Then Ti has a fixed point in Xˆv. Moreover, if α(x∗,y∗)≥1 for all x∗,y∗∈⋂i=1Fix(Ti), then Ti has a unique common fixed point in ⋂i=1Fix(Ti) for each i=1,⋯,6.
Proof. Suppose that Xˆv is empty, in which case, there is nothing to prove. We now assume that Xˆv≠∅. Then a function α:Xˆv×Xˆv→[0,+∞), a≠0, a<1, b≠0, acb<1 and x0∈Xˆv exists such that α(x0,x1)≥1,ϵ,δ>0, and, for each i=1,⋯,6, Ti remain triangular α-orbital admissible mappings for every λ>0. Therefore the mappings, Ti for i=1,⋯,6 satisfy the inequalities (3.1)–(3.6). Since x0,x1 and x2 are points in Xˆv and T1(Xˆv)⊆T6(Xˆv), we can find a point x1 in Xˆv such that ξ0=T1x0=T6x1. For T2(Xˆv)⊆T5(Xˆv), we can find a point x2∈Xˆv such that ξ1=T2x1=T5x2 and for T3(Xˆv)⊆T4(Xˆv); we can find a point x3 in Xˆv such that ξ2=T3x2=T4x3. Now for all λ>0, we induce on n, so that there are sequences {xn}n∈N and {ξn}n∈N within Xˆv that satisfy the subsequent Eq (3.7). If n0∈N exists such that ξn0=ξn0+1, T3(Xˆv)⊆T4(Xˆv), T2(Xˆv)⊆T5(Xˆv), T1(Xˆv)⊆T6(Xˆv) hold. In fact, if m∈N exists such that ξm+2=ξm+3, then T1u=T6u, where u=xm+3. Therefore, the pair {T1,T6} has a coincidence point u∈Xˆv. If ξm=ξm+1, then T2u=T4u, where u=xm+1. Therefore, the pair {T2,T4} has a coincidence point u∈Xˆv. If ξm+1=ξm+3, then T3u=T5u, where u=xm+2. Thus, the pair {T3,T5} has a coincidence point u∈Xˆv. Again, if there is an n0∈N such that ξn0=ξn0+1=ξn0+2, then ξn=ξn0 for any n≥n0. This implies that {ξn} is a modular ˆv Cauchy sequence in Xˆv. Actually, if η∈N exists such that (1) ξη=ξη+1=ξη+2, (2) ξη≠ξη+1=ξη+2, (3) ξη≠ξη+2=ξη+1, and (4) ξη≠ξη+1≠ξη+2 hold. In fact Case (1) is easy, and Case (3) is similar to Case (2); then from inequality (3.1)–(3.6), we get the result by setting x=ξη+2 and y=ξη+3. By Lemma 3, we conclude that
ˆvλ(ξn,ξm)=0∀λ>0,n≥m. | (3.34) |
Consequently, Eq (3.34) indicates that {ξn}n∈N forms a modular extended ˆv-Cauchy sequence within a ˆv-complete modular extended b-metric space. Therefore, a point u∈Xˆv exists such that ξn converges to u as n approaches infinity. Furthermore, since the sequences {T1xn}={T6xn+1}, {T2xn+1}={T5xn+2}, and {T3xn−1}={T4xn} for all n∈N are all subsequences of {ξn}, it follows that all subsequences of a convergent sequence converge to the same limit. Thus, we conclude that
limn→∞T1xn=limn→∞T6xn+1=limn→∞ξn=u, |
limn→∞T2xn+1=limn→∞T5xn+2=limn→∞ξn+1=u, |
limn→∞T3xn−1=limn→∞T4xn=limn→∞ξn+1=u. |
Since {T2,T4} is weakly commuting mappings, we have, for all λ>0,
ˆvλ(T2T4xn+1,T4T2xn+1)≤ˆvλ(T4xn+1,T2xn+1). | (3.35) |
Taking the limit of inequality (3.35) as n→∞ and noticing that T2 or T4 are ˆv-continuous mappings, we get
ˆvλ(T2T4xn+1,T4T2xn+1)≤ˆvλ(T4xn+1,T2xn+1)⟶0. | (3.36) |
We know that T4 is ˆv continuous, then T24xn+1→T4u as n→∞, T4T2xn+1→T4u as n→∞. But we can clearly see from inequality (3.35) that T2T4xn+1→T4u as n→∞. Since T2,T4 are weakly commuting mappings, for each i=1,…,6, where Ti represents a triangular α-orbital admissible mapping. and 1≤α(xm,xn) for all n,m∈N with n<m, it follows that, for all λ>0,
α(xn,xn+1)ˆvλ(T2T4xn+1,T4T2xn+1)≤α(xn,xn+1)ˆvλ(T4xn+1,T2xn+1), |
and hence
limn→∞α(xn,xn+1)ˆvλ(T2T4xn+1,T4T2xn+1)≤limn→∞α(xn,xn+1)ˆvλ(T4xn+1,T2xn+1). |
Since T2,T4 are weakly commuting mappings and orbitally continuous, we have
α(xn,xn+1)ˆvλ(limn→∞T2T4xn+1,limn→∞T4T2xn+1)≤α(xn,xn+1)ˆvλ(limn→∞T4xn+1,limn→∞T2xn+1), |
such that
α(xn,xn+1)ˆvλ(T2limn→∞T4xn+1,T4limn→∞T2xn+1)≤α(xn,xn+1)ˆvλ(limn→∞T4xn+1,limn→∞T2xn+1). |
Therefore, we have
ˆvλ(T2u,T4u)≤ˆvλ(u,u)=0, |
which implies that T2u=T4u for all λ>0.
Additionally, given that T3 and T5 are weakly commuting mappings, it follows that for every λ>0,
ˆvλ(T3T5xn+2,T5T3xn+2)≤ˆvλ(T5xn+2,T3xn+2). | (3.37) |
Taking the limit of inequality (3.35) as n→∞ and noticing that T3 or T4 is a ˆv-continuous mapping, then we get
ˆvλ(T3T5xn+2,T5T3xn+2)≤ˆvλ(T5xn+2,T3xn+2)⟶0. | (3.38) |
Since T5 is ˆv continuous, T25xn+2→T5u as n→∞, and T5T3xn+2→T5u as n→∞. But we can clearly see from inequality (3.37) that T3T5xn+2→T5u as n→∞.
Now we know that T3,T5 are weakly commuting mappings and orbitally continuous, Ti remains a triangular α-orbital admissible mapping for each i=1,⋯,6, and 1≤α(xm,xn) for all n,m∈N with n<m, it follows that, for all λ>0
α(xn,xn+2)ˆvλ(limn→∞T3T5xn+2,limn→∞T5T3xn+2)≤α(xn,xn+2)ˆvλ(limn→∞T5xn+2,limn→∞T3xn+2), |
thus,
α(xn,xn+2)ˆvλ(T3limn→∞T5xn+2,T5limn→∞T3xn+2)≤α(xn,xn+2)ˆvλ(limn→∞T5xn+2,limn→∞T3xn+2). |
Therefore, we have
ˆvλ(T3u,T5u)≤ˆvλ(u,u)=0, |
which implies that T3u=T5u for all λ>0.
Lastly, since T1,T6 are weakly commuting mappings and orbitally continuous, Ti remains a triangular α-orbital admissible mapping for each i=1,⋯,6, and 1≤α(xm,xn) for all n,m∈N with n<m, it follow that, for all λ>0
α(xn,xn+1)ˆvλ(limn→∞supT1T6xn,limn→∞T6T2xn)≤α(xn,xn+1)ˆvλ(limn→∞supT6xn,limn→∞T1xn), |
and so
α(xn,xn+1)ˆvλ(T1limn→∞T6xn,T6limn→∞T1xn)≤α(xn,xn+1)ˆvλ(limn→∞T6xn,limn→∞T1xn). |
Therefore, we have
ˆvλ(T1u,T6u)≤ˆvλ(u,u)=0, |
which implies that T1u=T6u for all λ>0.
Now, we claim that, for all u∈Xˆv, T1u=T2u=T3u. If we suppose the contrary, then T1u≠T2u≠T3u. So, for all λ>0, the following cases emerge:
Case 1a. T1u≠T2u⟹α(u,u)ˆvλ(T1u,T2u)>0.
Case 1b. T1u≠T3u⟹α(u,u)ˆvλ(T1u,T3u)>0.
Case 1c. T2u≠T3u⟹α(u,u)ˆvλ(T2u,T3u)>0.
Indeed, since Ti remains a triangular α-orbital admissible mapping for each i=1,⋯,6, and α(u,u)≠0, it follows that, for all λ>0
α(x,y)ˆvλ(T1x,T2y)<A(ϵ)⟹ϵ≤A−1(FAλ(T1x,T2y))<A(ϵ)+A(δ(ϵ)),α(x,y)ˆvλ(T1x,T3y)<A(ϵ)⟹ϵ≤A−1(FBλ(T1x,T3y))<A(ϵ)+A(δ(ϵ)),α(x,y)ˆvλ(T2x,T3y)<A(ϵ)⟹ϵ≤A−1(FCλ(T2x,T3y))<A(ϵ)+A(δ(ϵ)), |
where
FAλ(T1x,T2y):=amax{ˆvλ(T6x,T4y),ˆvλ(T1x,T6y),ˆvλ(T3x,T4y),ˆvλ(T2x,T5y)},FBλ(T1x,T3y):=bmax{ˆvλ(T5x,T4y),ˆvλ(T2x,T5y),ˆvλ(T1x,T26x),ˆvλ(T3y,T24y)},FCλ(T2x,T3y):=cmax{ˆvλ(T6x,T2x),ˆvλ(T2x,T5y),ˆvλ(T3x,T4y),ˆvλ(T1x,T6y)}. |
Now, we consider Case 1a. Since T1u≠T2u, for all λ>0, we have α(u,u)ˆvλ(T1u,T2u)>0 and so, from (3.28), (3.29), and (3.31), we have
α(u,u)ˆvλ(T1u,T2u)≤aα(u,u)ˆvλ(T1u,T3u)≤baα(u,u)ˆvλ(T2u,T3u)≤acbα(u,u)ˆvλ(T1u,T2u), | (3.39) |
which implies α(u,u)≠0. Thus we have
α(u,u)ˆvλ(T1u,T2u)≤acbα(u,u)ˆvλ(T1u,T2u), |
which implies that
α(u,u)(1−acb)ˆvλ(T1u,T2u)=0⟹T1u=T2u |
for all λ>0 since acb<1 and b≠0.
For Case 1c, using Case 1a above, that is, T1u=T2u, it follows from (3.39) that
0=α(u,u)ˆvλ(T1u,T1u)≤aα(u,u)ˆvλ(T2u,T3u)≤baα(u,u)ˆvλ(T2u,T3u)≤acbα(u,u)ˆvλ(T1u,T1u). | (3.40) |
So, by using the condition (1) of Definition 10,
0≤aα(u,u)ˆvλ(T2u,T3u)≤baα(u,u)ˆvλ(T2u,T3u)≤0, | (3.41) |
which implies that
α(u,u)baˆvλ(T2u,T3u)=0. |
Hence T2u=T3u, since ba<1, a2≠0, and α(u,u)≠0.
For Case 1b, using Cases 1a and 1c in (3.39), we get 0<aα(u,u)ˆvλ(T1u,T3u)≤0. Since a≠0 and α(u,u)≠0, we have T1u=T3u.
So, in all the cases above, we have T1u=T2u=T3u. However, since T1u=T6u, T2u=T4u and T3u=T5u, it follows that
T1u=T2u=T3u=T4u=T5u=T6u, |
which implies that u∈Xˆv is the coincidence point of Ti for each i=1,2,⋯,6.
We demonstrate that if a point is a fixed point of T1, it is also a fixed point for T2,T3,T4,T5, and T6. Assume that a point p∈Xˆv exists such that p satisfies p=T1p. We assert that p=T2p=T3p. Indeed, suppose that this is not true. Then p≠T2p and p≠T3p, and so we have the following cases for all λ>0,
Case 1. p≠T2p⟹T1p≠T2p⟹α(p,p)ˆvλ(T1p,T2p)>0;
Case 2. p≠T3p⟹T1p≠T3p⟹α(p,p)ˆvλ(T1p,T3p)>0.
Indeed, note that α(p,p)≠0 and Ti remain a triangular α-orbital admissible mapping for each i=1,⋯,6 such that
A−1(FAλ(T1p,T2p))≥ˆvλ(T1p,T2p)>0,A−1(FBλ(T1p,T3p))≥ˆvλ(T1p,T3p)>0. |
For Case 1, it follows from (3.39) that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p). | (3.42) |
From p∈Fix(T1), we have
α(T1p,T1p)ˆvλ(p,T2p)≤aα(T1p,T1p)ˆvλ(p,T3p)≤baα(T1p,T1p)ˆvλ(T2p,T3p)≤acbα(T1p,T1p)ˆvλ(p,T2p), | (3.43) |
which implies that α(T1p,T1p)(1−acb)ˆvλ(p,T2p)≤0, for all λ>0 since α(T1p,T1p)≠0 and acb<1. This is a contradiction. Therefore, p=T2p and hence p=T1p=T2p.
For Case 2, if p≠T3p, then T1p≠T3p and therefore
α(p,p)ˆvλ(T1p,T3p)>0 |
for all λ>0. Thus, it follows from p=T1p=T2p, Case 1, and (3.39) that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p), | (3.44) |
which implies
0≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(p,T3p)≤0 | (3.45) |
for all λ>0. Since b≠0 and α(p,p)≠0, we have
aα(p,p)baˆvλ(T1p,T3p)=0, |
and so ˆvλ(T1p,T3p)=0; that is, ˆvλ(p,T3p)=0. Therefore, p=T3p and so p=T1p=T3p. Hence, from Cases 1 and 2, p=T1p=T2p=T3p.
Again, suppose that p∈Xˆv exists such that p∈Fix(T2), i.e., p=T2p. We claim that p=T1p=T3p. Indeed, suppose that this is not true. Then p≠T1p and p≠T3p, so we have the following cases for all λ>0:
Case 3. p≠T1p⟹T2p≠T1p⟹α(p,p)ˆvλ(T1p,T2p)>0;
Case 4. p≠T3p⟹T2p≠T3p⟹α(p,p)ˆvλ(T2p,T3p)>0.
Indeed, note that α(p,p)≠0 and Ti remains a triangular α-orbital admissible mapping for each i=1,⋯,6 such that
A−1(FAλ(T1p,T2p))≥ˆvλ(T1p,T2p)>0,A−1(FCλ(T2p,T3p))≥ˆvλ(T2p,T3p)>0. |
For Case 3, it follows from (3.39) that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p). | (3.46) |
Since p∈Fix(T2), we have
α(T2p,T2p)ˆvλ(p,T1p)≤aα(T2p,T2p)ˆvλ(T1p,T3p)≤baα(T2p,T2p)ˆvλ(p,T3p)≤acbα(T2p,T2p)ˆvλ(p,T1p), | (3.47) |
which implies that
α(T2p,T2p)(1−acb)ˆvλ(p,T1p)≤0 |
for all λ>0, since α(T2p,T2p)≠0 and acb<1, which is a contradiction. Therefore, p=T1p and hence p=T1p=T2p.
For Case 4, if p≠T3p, then T2p≠T3p and therefore
α(p,p)ˆvλ(T2p,T3p)>0 |
for all λ>0. Thus it follows from p=T2p=T1p and (3.39) that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p), | (3.48) |
which implies
0≤aα(p,p)ˆvλ(p,T3p)≤baα(p,p)ˆvλ(p,T3p)≤0 | (3.49) |
for all λ>0. Since b≠0 and α(p,p)≠0, we have
aα(p,p)baˆvλ(p,T3p)=0, |
and so ˆvλ(p,T3p)=0; that is, p=T3p. Therefore, p=T2p=T3p. Hence, from Cases 3 and 4, p=T1p=T2p=T3p.
Lastly, suppose that p∈Xˆv exists such that p∈Fix(T3), i.e., p=T3p. We claim that p=T1p=T2p. Indeed, suppose that this is not true. Then p≠T1p and p≠T2p, and therefore we have the following cases for all λ>0:
Case 5. p≠T1p⟹T3p≠T1p⟹α(p,p)ˆvλ(T1p,T3p)>0.
Case 6. p≠T2p⟹T3p≠T2p⟹α(p,p)ˆvλ(T2p,T3p)>0.
Indeed, note that α(p,p)≠0 and Ti remains a triangular α-orbital admissible mapping for each i=1,⋯,6 such that
A−1(FBλ(T1p,T3p))≥ˆvλ(T1p,T3p)>0,A−1(FCλ(T2p,T3p))≥ˆvλ(T2p,T3p)>0. |
For Case 5, if p≠T1p, T3p≠T1p, and so
α(p,p)ˆvλ(T1p,T3p)>0 |
for all λ>0. Thus it follows from p=T3p, (3.39), and Case 2 that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p), | (3.50) |
which implies
α(p,p)ˆvλ(T1p,p)≤acbα(p,p)ˆvλ(T1p,p) | (3.51) |
for all λ>0. Since b≠0 and α(p,p)≠0, we have
aα(p,p)(1−acb)ˆvλ(T1p,p)=0, |
which implies ˆvλ(T1p,p)=0; that is, p=T1p. Therefore, p=T1p=T2p=T3p.
For Case 6, if p≠T2p, then T2p≠T3p, and therefore
α(p,p)ˆvλ(T2p,T3p)>0 |
for all λ>0. Thus, using p=T3p=T1p, it follows from (3.39) that
α(p,p)ˆvλ(T1p,T2p)≤aα(p,p)ˆvλ(T1p,T3p)≤baα(p,p)ˆvλ(T2p,T3p)≤acbα(p,p)ˆvλ(T1p,T2p), | (3.52) |
which implies
α(p,p)ˆvλ(T1p,T2p)≤acbα(p,p)ˆvλ(T1p,T2p), |
and so
α(T1p,T1p)ˆvλ(p,T2p)≤acbα(T1p,T1p)ˆvλ(p,T2p) |
for all λ>0. Since b≠0 and α(T1p,T1p)≠0, we have
α(T1p,T1p)(1−acb)ˆvλ(p,T2p)=0, |
which implies ˆvλ(p,T2p)=0, that is, p=T2p. Therefore, p=T1p=T2p. Hence, it follows from Cases 5 and 6 that p=T1p=T2p=T3p. In all the cases above, p=T1p=T2p=T3p.
Now, using Cases 1a–1c above, we have
p=T1p=T6p,p=T2p=T4p,p=T3p=T5p. |
Therefore, we have p=T1p=T2p=T3p=T4p=T5p=T6p or p∈Fix(Ti) for each i=1,2,⋯,6, which shows that p is a common fixed point of the mappings T1, T2, T3, T4, T5,andT6.
To prove the uniqueness of the common fixed point p, suppose that another common fixed point x∗ (p≠x∗) of T1,T2,T3,T4,T5,andT6 exists, namely
x∗=T1x∗=T2x∗=T3x∗=T4x∗=T5x∗=T6x∗. |
Since α(x∗,p)≥1, we have the following cases:
Case 1a*. α(x∗,y∗)≥1, x∗≠y∗⟹α(x∗,y∗)ˆvλ(T1x∗,T2y∗)>0.
Case 1b*. α(x∗,y∗)≥1, x∗≠y∗⟹α(x∗,y∗)ˆvλ(T1x∗,T3y∗)>0.
Case 1c*. α(x∗,y∗)≥1, x∗≠y∗⟹α(x∗,y∗)ˆvλ(T2x∗,T3y∗)>0.
We can see that Case 1a* follows from Case 1a above and hence, x∗=y∗. Again, Case 1b* follows from Case 1b above and hence, x∗=y∗. Finally, Case 1c* follows from Case 1c above and therefore, x∗=y∗. Therefore, we have x∗=y∗. Hence, the common fixed-point p is a unique common fixed point Ti for each i=1,⋯,6. We are now done with the proof.
Remark 4. Theorem 1 is a generalization of results in Karapinar et al. [26], Theorem 2.8 in Gholidahneh et al. [21], and Theorem 3.7 in Okeke et al. [34].
Our result have made the following progress over the classical results in the following ways:
(a) This study extends classical metric and b-metric spaces by introducing modular extended b-metric spaces, which allow function-controlled distances.
(b) It defines the α-ˆv-A-B-C-Meir-Keeler-type contractions, a function-dependent contraction that generalizes Banach, Kannan, and Meir-Keeler contractions.
(c) Unlike traditional single-mapping results, this work establishes common fixed-point results for six self-mappings, significantly broadening the applicability of fixed-point theorems.
(d) The proposed contraction conditions use function-dependent inequalities instead of fixed contraction constants, making them adaptable to nonlinear and dynamic systems.
(e) This work bridges classical and modern fixed-point results by incorporating modular functions, b-metric spaces, and multi-mapping interactions, leading to a more generalized framework.
Now, we establish the following example to solidify Theorem 1.
Example 3. Let X=(R∖{0})∪{∞} with the modular extended b-metric defined by
ln(2)2ˆvλ(x,y):=11+λmaxx,y∈Xˆv{||x−y||}, |
which is complete in Xˆv for all λ>0. Define the ˆv-weakly commuting mappings T1,T2,T3,T4,T5,T6:(R∖{0})∪{∞}→(R∖{0})∪{∞} as follows:
T1x=log64x6,T2x=log32x5,T3x=log16x4, |
T4x=log8x3,T5x=log4x2,T6x=log2x, |
for all x∈(R∖{0})∪{∞} and λ>0, for each i=1,2,⋯,6 and also for x,y∈(R∖{0})∪{∞}, α(x,Tix)≥1⟹α(Tix,T2ix)≥1, α(x,y)≥1 and α(y,Tiy)≥1⟹α(x,Tiy)≥1. Then the mappings T1,T2,T3,T4,T5, and T6 satisfy the inequalities (3.1)–(3.3) of Theorem 1.
In fact, let Ti:Xˆv⟶Xˆv be six orbitally continuous α-ˆv-A-B-C-Meir-Keeler-type contraction mappings for i=1,2,⋯,6, and let {T2,T4}, {T3,T5} and {T1,T6} be weakly commuting pairs of self-mappings. Indeed, consider {T2,T4}. Now for all λ>0, and x∈(R∖{0}), we show that ˆvλ(T2T4x,T4T2x)≤ˆvλ(T4x,T2x). Then, by the definition of a ˆv-modular extended b-metric, we have
11+λmaxx∈Xˆv{||T2T4x−T4T2x||}=11+λmaxx∈Xˆv{||log32(log8x3)5−log8(log32x5)3||}=11+λmaxx∈Xˆv{||5log32(log8x3)−3log8(log32x5)||}=11+λmaxx∈Xˆv||ln(log8x3)ln(2)−ln(log32x5)ln(2)||=1ln(2)11+λmaxx∈Xˆv||ln(log8x3)−ln(log32x5)||=1ln(2)11+λmaxx∈Xˆv||ln(log8x3log32x5)||=0. |
Again,
ˆvλ(T4x,T2x)=11+λmaxx∈Xˆv{||T4x−T2x||}=11+λmaxx∈Xˆv{||log8x3−log32x5||}=11+λ1ln(2)maxx∈Xˆv{||ln(x)−ln(x)||}=0. |
Thus ˆvλ(T2T4x,T4T2x)≤ˆvλ(T4x,T2x), showing that {T2,T4} is weakly commuting pair of self-mappings, and {T3,T5} and {T1,T6} are weakly commuting pairs of self-mappings according to the above mentioned procedure. It is clear that T3(Xˆv)⊆T4(Xˆv), T2(Xˆv)⊆T5(Xˆv), T1(Xˆv)⊆T6(Xˆv), so that there exists a function α:Xˆv×Xˆv→[0,+∞), a≠0,a<1, b≠0, acb<1 and x0∈Xˆv such that α(x0,Tix0)≥1. Therefore by definition, we have
ˆvλ(T1x,T2y)=11+λmaxx,y∈Xˆv{||T1x−T2y||}=11+λmaxx,y∈Xˆv{||log64x6−log32y5||}=11+λmaxx,y∈Xˆv{||6log64x−5log32y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.53) |
ˆvλ(T1x,T3y)=11+λmaxx,y∈Xˆv{||T1x−T3y||}=11+λmaxx,y∈Xˆv{||log64x6−log16y4||}=11+λmaxx,y∈Xˆv{||6log64x−4log16y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.54) |
ˆvλ(T2x,T3y)=11+λmaxx,y∈Xˆv{||T2x−T3y||}=11+λmaxx,y∈Xˆv{||log32x5−log16y4||}=11+λmaxx,y∈Xˆv{||5log32x−4log16y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.55) |
Now,
ˆvλ(T6x,T4y)=11+λmaxx,y∈Xˆv{||T6x−T4y||}=11+λmaxx,y∈Xˆv{||log2x−log8y3||}=11+λmaxx,y∈Xˆv{||log2x−3log8y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.56) |
ˆvλ(T1x,T6y)=11+λmaxx,y∈Xˆv{||T1x−T6y||}=11+λmaxx,y∈Xˆv{||log64x6−log2y||}=11+λmaxx,y∈Xˆv{||6log64x−log2y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.57) |
ˆvλ(T3x,T4y)=11+λmaxx,y∈Xˆv{||T3x−T4y||}=11+λmaxx,y∈Xˆv{||log16x4−log8y3||}=11+λmaxx,y∈Xˆv{||4log16x−3log8y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.58) |
ˆvλ(T2x,T5y)=11+λmaxx,y∈Xˆv{||T2x−T5y||}=11+λmaxx,y∈Xˆv{||log32x5−log4y2||}=11+λmaxx,y∈Xˆv{||5log32x−2log4y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.59) |
Again,
ˆvλ(T5x,T4y)=11+λmaxx,y∈Xˆv{||T5x−T4y||}=11+λmaxx,y∈Xˆv{||log4x2−log8y3||}=11+λmaxx,y∈Xˆv{||2log4x−3log8y||}=11+λ1ln(2)maxx,y∈Xˆv{||ln(x)−ln(y)||}=11+λ1ln(2)maxx,y∈Xˆv{ln||xy||}. | (3.60) |
ˆvλ(T1x,T26x)=11+λmaxx∈Xˆv{||T1x−T26x||}=11+λmaxx∈Xˆv{||log64x6−(log2x)2||}=11+λmaxx∈Xˆv{||6log64x−(log2x)2||}=11+λ1ln(2)maxx∈Xˆv{||ln(x)−ln(x)2ln(2)||}≤11+λ1ln(2)maxx∈Xˆv{ln||x||}. | (3.61) |
ˆvλ(T3y,T24y)=11+λmaxy∈Xˆv{||T3y−T24y||}=11+λmaxy∈Xˆv{||log16y4−(log8y3)2||}=11+λmaxy∈Xˆv{||4log16y−(3log8y)2||}=11+λ1ln(2)maxy∈Xˆv{||ln(y)−ln(y)2ln(2)||}≤11+λ1ln(2)maxy∈Xˆv{ln||y||}. | (3.62) |
ˆvλ(T6x,T2x)=11+λmaxx∈Xˆv{||T6x−T2x||}=11+λmaxx∈Xˆv{||log2x−log32x5||}=11+λmaxx∈Xˆv{||log2x−5log32x||}=11+λ1ln(2)maxx∈Xˆv{||ln(x)−ln(x)||}=11+λ1ln(2)maxx∈Xˆv{||0||}=0. | (3.63) |
Therefore, we have FAλ(T1x,T2y)=amaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||xy||,ln||xy||}, where a=12. Given 12ln(ϵ), we choose ln(δ(ϵ))=12ln(ϵ). If ϵ≤exp{ln(FAλ(T1x,T2y))}<12ln(ϵ)+ln(δ(ϵ))=ln(ϵ)>ϵ, therefore, we have, as ϵ>0,
α(x,y)ˆvλ(T1x,T2y)≤a11+λmaxx,y∈Xˆv{ln||xy||}≤aexp{ln{11+λmaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||xy||,ln||xy||}}}=aexpln(FAλ(T1x,T2y))<ln(ϵ)<ϵ. |
Again, FBλ(T1x,T3y)=bmaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||x||,ln||y||}, where b=12. Given 12ln(ϵ), we choose ln(δ(ϵ))=12ln(ϵ). If ϵ≤expln(FBλ(T1x,T3y))<12ln(ϵ)+ln(δ(ϵ))=ln(ϵ)<ϵ, therefore, we have ϵ>0 and
α(x,y)ˆvλ(T1x,T3y)≤b11+λmaxx,y∈Xˆv{ln||xy||}≤bexp{ln{11+λmaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||x||,ln||y||}}}=bexpln(FBλ(T1x,T3y))<ln(ϵ)<ϵ. |
Lastly, FCλ(T2x,T3y)=cmaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||xy||}, where c=12. Given 12ln(ϵ), we choose ln(δ(ϵ))=12ln(ϵ). If ϵ≤expln(FCλ(T2x,T3y))<12ln(ϵ)+ln(δ(ϵ))=ln(ϵ)<ϵ, Therefore, we have
α(x,y)ˆvλ(T2x,T3y)≤c11+λmaxx,y∈Xˆv{ln||xy||}≤cexp{ln{11+λmaxx,y∈Xˆv{ln||xy||,ln||xy||,ln||xy||}}}=cexpln(FCλ(T2x,T3y))<ln(ϵ)<ϵ. |
Thus we have
α(x,y)ˆvλ(T1x,T2y)<A(ϵ)⟹ϵ≤A−1(FAλ(T1x,T2y))<A(ϵ)+A(δ(ϵ)); | (3.64) |
α(x,y)ˆvλ(T1x,T3y)<A(ϵ)⟹ϵ≤A−1(FBλ(T1x,T3y))<A(ϵ)+A(δ(ϵ)); | (3.65) |
α(x,y)ˆvλ(T2x,T3y)<A(ϵ)⟹ϵ≤A−1(FCλ(T2x,T3y))<A(ϵ)+A(δ(ϵ)), | (3.66) |
where
FAλ(T1x,T2y):=amax{ˆvλ(T6x,T4y),ˆvλ(T1x,T6y),ˆvλ(T3x,T4y),ˆvλ(T2x,T5y)}; | (3.67) |
FBλ(T1x,T3y):=bmax{ˆvλ(T5x,T4y),ˆvλ(T2x,T5y),ˆvλ(T1x,T26x),ˆvλ(T3y,T24y)}; | (3.68) |
FCλ(T2x,T3y):=cmax{ˆvλ(T6x,T2x),ˆvλ(T2x,T5y),ˆvλ(T3x,T4y),ˆvλ(T1x,T6y)}. | (3.69) |
Since Ti is orbitally continuous, all the conditions of Theorem 1 are fulfilled.
Corollary 1. Let Xˆv be a ˆv-regular and ˆv-complete modular extended b-metric space. Consider six mappings Ti:Xˆv→Xˆv for i=1,2,…,6, which are orbitally continuous and satisfy a specific type of contraction known as the α-ˆv-A-B-C-Meir-Keeler condition. The pairs {T2,T4}, {T3,T5}, and {T1,T6} are weakly commuting self-mappings, with the following inclusions holding:
T3(Xˆv)⊆T4(Xˆv),T2(Xˆv)⊆T5(Xˆv),T1(Xˆv)⊆T6(Xˆv). |
Additionally, a function α:Xˆv×Xˆv→[0,+∞) exists with the parameters a≠0, a<1, b≠0, and the condition acb<1. Let x0∈Xˆv be given such that
α(x0,x1)≥1,ϵ,δ>0. |
For each mapping Ti (where i=1,…,6), we find that Ti remains triangular α-orbital admissible mapping for all λ>0. The following conditions are satisfied for some positive integer m≥1:
α(x,y)ˆvλ(Tm1x,Tm2y)<A(ϵ)⟹ϵ≤A−1(FAλ(Tm1x,Tm2y))<A(ϵ)+A(δ(ϵ)); | (3.70) |
α(x,y)ˆvλ(Tm1x,Tm3y)<A(ϵ)⟹ϵ≤A−1(FBλ(Tm1x,Tm3y))<A(ϵ)+A(δ(ϵ)); | (3.71) |
α(x,y)ˆvλ(Tm2x,Tm3y)<A(ϵ)⟹ϵ≤A−1(FCλ(Tm2x,Tm3y))<A(ϵ)+A(δ(ϵ)), | (3.72) |
where
FAλ(Tm1x,Tm2y):=amax{ˆvλ(Tm6x,Tm4y),ˆvλ(Tm1x,Tm6y),ˆvλ(Tm3x,Tm4y),ˆvλ(Tm2x,Tm5y)}; | (3.73) |
FBλ(Tm1x,Tm3y):=bmax{ˆvλ(Tm5x,Tm4y),ˆvλ(Tm2x,Tm5y),ˆvλ(Tm1x,T2m6x),ˆvλ(Tm3y,T2m4y)}; | (3.74) |
FCλ(Tm2x,Tm3y):=cmax{ˆvλ(Tm6x,Tm2x),ˆvλ(Tm2x,Tm5y),ˆvλ(Tm3x,Tm4y),ˆvλ(Tm1x,Tm6y)}. | (3.75) |
Let the sequences {xn}n∈N and {ξn}n∈N be in Xˆv so that for xn in Xˆv, we choose xn+1 such that ξn=T1xn=T6xn+1; again, for xn+1 in Xˆv, we choose xn+2 such that ξn+1=T2xn+1=T5xn+2 and, for a point xn+2 in Xˆv, we choose xn+3 such that ξn+2=T3xn+2=T4xn+3 for n=0,1,2,⋯. Then Ti has a fixed point in Xˆv. Moreover, if α(x∗,y∗)≥1 for all x∗,y∗∈⋂i=1Fix(Ti), then Ti has common unique fixed-point in ⋂i=1Fix(Ti) for i=1,⋯,6 for some positive integer, m≥1.
Proof. According to Theorem 1, for a certain positive integer m≥1, we have the following equalities: p=Tm1p=Tm6p, p=Tm2p=Tm4p, and p=Tm3p=Tm5p. This implies that p can be expressed as:
p=Tm1p=Tm2p=Tm3p=Tm4p=Tm5p=Tm6p, |
indicating that p lies in the fixed-point set Fix(Ti) for each i=1,2,…,6 and the positive integer m≥1. Consequently, p serves as a fixed point for each mapping Tm1,Tm2,Tm3,Tm4,Tm5,Tm6. The uniqueness of this point can be derived similarly to the argument presented in Theorem 1. Thus, it follows that the mappings Ti possess a unique common fixed point located in the intersection ⋂i=1Fix(Ti) for i=1,…,6 and for some positive integer m≥1.
Corollary 2. Let Xˆv be a ˆv-regular and ˆv-complete modular extended b-metric space. Consider six orbitally continuous mappings Ti:Xˆv→Xˆv (i=1,2,…,6) that satisfy the α-ˆv-A-B-C-Meir-Keeler-type contraction condition. Additionally, assume that the pairs {T2,T4}, {T3,T5}, and {T1,T6} are weakly commuting, with the following inclusions holding:
T3(Xˆv)⊆T4(Xˆv),T2(Xˆv)⊆T5(Xˆv),T1(Xˆv)⊆T6(Xˆv). |
Assume that a function α:Xˆv×Xˆv→[0,+∞) exists, along with the constants a≠0, a<1, b≠0, and acb<1, as well as a point x0∈Xˆv, such that:
α(x0,x1)≥1,ϵ,δ>0. |
Furthermore, for all λ>0, each Ti is assumed to be a triangular α-orbital admissible mapping and satisfies the following conditions:
α(x,y)ˆvλ(T1x,T2y)<A(ϵ)⟹ϵ≤A−1(FAλ(T1x,T2y))<A(ϵ)+A(δ(ϵ)); | (3.76) |
α(x,y)ˆvλ(T1x,T3y)<A(ϵ)⟹ϵ≤A−1(FBλ(T1x,T3y))<A(ϵ)+A(δ(ϵ)); | (3.77) |
α(x,y)ˆvλ(T2x,T3y)<A(ϵ)⟹ϵ≤A−1(FCλ(T2x,T3y))<A(ϵ)+A(δ(ϵ)), | (3.78) |
where
FAλ(T1x,T2y):=amax{max{ˆvλ(T6x,T4y),ˆvλ(T1x,T6y)},min{ˆvλ(T3x,T4y),ˆvλ(T2x,T5y)}}; | (3.79) |
FBλ(T1x,T3y):=bmax{max{ˆvλ(T5x,T4y),ˆvλ(T2x,T5y)},min{ˆvλ(T1x,T26x),ˆvλ(T3y,T24y)}}; | (3.80) |
FCλ(T2x,T3y):=cmax{max{ˆvλ(T6x,T2x),ˆvλ(T2x,T5y)},min{ˆvλ(T3x,T4y),ˆvλ(T1x,T6y)}}. | (3.81) |
Let the sequences {xn}n∈N and {ξn}n∈N be in Xˆv so that for xn be defined in Xˆv, we choose xn+1 such that ξn=T1xn=T6xn+1; again for xn+1 in Xˆv, we choose xn+2 such that ξn+1=T2xn+1=T5xn+2 and, for a point xn+2 in Xˆv, we choose xn+3 such that ξn+2=T3xn+2=T4xn+3 for n=0,1,2,⋯. Then Ti has a fixed point in Xˆv. Moreover, if α(x∗,y∗)≥1 for all x∗,y∗∈⋂i=1Fix(Ti), then Ti has a unique common fixed point in ⋂i=1Fix(Ti) for each i=1,⋯,6.
Proof. If we take Xˆv to be empty, then there is nothing to prove. Henceforth, we assume that Xˆv≠∅. Then a function α:Xˆv×Xˆv→[0,+∞), a≠0, a<1, b≠0, acb<1, and x0∈Xˆv exists such that α(x0,x1)≥1,ϵ,δ>0, and, for each i=1,⋯,6, Ti remains a triangular α-orbital admissible mappings for every λ>0. Suppose that the mappings, Ti for i=1,⋯,6 satisfy theinequalities (3.76)–(3.78). Since x0,x1, and x2 are points in Xˆv and T1(Xˆv)⊆T6(Xˆv), we can find a point x1 in Xˆv such that ξ0=T1x0=T6x1. For T2(Xˆv)⊆T5(Xˆv), we can find a point x2∈Xˆv such that ξ1=T2x1=T5x2, and for T3(Xˆv)⊆T4(Xˆv), we can find a point x3 in Xˆv such that ξ2=T3x2=T4x3. Now for all λ>0, in general, one can find sequences {xn}n∈N and {ξn}n∈N residing in the space Xˆv that fulfill the relationships defined in Eq (3.7). If there is an integer n0∈N such that ξn0=ξn0+1, it follows that T3(Xˆv)⊆T4(Xˆv), T2(Xˆv)⊆T5(Xˆv), and T1(Xˆv)⊆T6(Xˆv) hold. In fact, if m∈N exists such that ξm+2=ξm+3, then T1u=T6u, where u=xm+3. Therefore, the pair {T1,T6} has a coincidence point u∈Xˆv. If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is an n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, if \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2); therefore, from the inequalities (3.76)–(3.78), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . Following the proof of Theorem 1, the conclusion is now evident.
Remark 5. Corollary 2 is a generalization of [34, Corollaries 3.16 and 3.18].
Corollary 3. Let X_{\hat{v}} be a modular extended b -metric space that is both \hat{v} -regular and \hat{v} -complete. Consider six mappings T_i: X_{\hat{v}} \rightarrow X_{\hat{v}} for i = 1, 2, \ldots, 6 that exhibit orbital continuity and adhere to the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction conditions. The pairs \{T_2, T_4\} , \{T_3, T_5\} , and \{T_1, T_6\} are considered weakly commuting self-mappings, satisfying the following inclusions:
T_3(X_{\hat{v}}) \subseteq T_4(X_{\hat{v}}), \quad T_2(X_{\hat{v}}) \subseteq T_5(X_{\hat{v}}), \quad T_1(X_{\hat{v}}) \subseteq T_6(X_{\hat{v}}). |
A function \alpha: X_{\hat{v}} \times X_{\hat{v}} \rightarrow [0, +\infty) exists such that a \neq 0 , a < 1 , b \neq 0 , and \frac{ac}{b} < 1 . Additionally, let x_0 \in X_{\hat{v}} satisfy \alpha(x_0, x_1) \geq 1 for some \epsilon, \delta > 0 . For each mapping T_i with i = 1, \ldots, 6 , it is required that T_i qualifies as triangular \alpha -orbital admissible mapping for all \lambda > 0 , and the following conditions hold for some positive integer m \geq 1 :
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.82) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.83) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.84) |
where
\begin{align} F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y): & = a\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{6}y)\},\\ & \min\{\hat{v}_{\lambda}(T^{m}_{3}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\}\biggr\}; \end{align} | (3.85) |
\begin{align} F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y): & = b\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{5}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\}, \\&\min\{\hat{v}_{\lambda}(T^{m}_{1}x,T^{2m}_{6}x), \hat{v}_{\lambda}(T^{m}_{3}y,T^{2m}_{4}y)\}\biggr\}; \end{align} | (3.86) |
\begin{align} F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y): & = c\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{2}x), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\}, \\&\min\{\hat{v}_{\lambda}(T^{m}_{3}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{6}y)\}\biggr\}. \end{align} | (3.87) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be defined in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 for some positive integer, where m\ge 1 .
Proof. According to Corollary 2, for a certain positive integer m \geq 1 , we have the equalities p = T_1^m p = T_6^m p , p = T_2^m p = T_4^m p , and p = T_3^m p = T_5^m p . Consequently, it follows that
p = T_1^m p = T_2^m p = T_3^m p = T_4^m p = T_5^m p = T_6^m p, |
indicating that p is a fixed point for each mapping T_i^m where i = 1, 2, \ldots, 6 for the specified positive integer m \geq 1 . This establishes that p is, in fact, a fixed point of T_1^m as well as T_2^m, T_3^m, T_4^m, T_5^m, and T_6^m individually. Hence, the uniqueness of this fixed point can be derived as shown in Theorem 1 above. Therefore, the mappings T_i possess a unique common fixed point within the intersection \bigcap_{i = 1} \text{Fix}(T_i) for i = 1, \ldots, 6 for some positive integer m \geq 1 .
Remark 6. Corollary 3 is a generalization of [34, Corollary 3.19].
Corollary 4. Consider the space X_{\hat{v}} , which is characterized as a \hat{v} -regular and \hat{v} -complete modular extended b -metric space. Within this framework, let T_i: X_{\hat{v}} \to X_{\hat{v}} for i = 1, 2, \ldots, 6 denote six mappings that are orbitally continuous and adhere to the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction conditions. The pairs \{T_2, T_4\} , \{T_3, T_5\} , and \{T_1, T_6\} are weakly commuting self-mappings, and the following inclusions are satisfied:
T_3(X_{\hat{v}}) \subseteq T_4(X_{\hat{v}}), \quad T_2(X_{\hat{v}}) \subseteq T_5(X_{\hat{v}}), \quad T_1(X_{\hat{v}}) \subseteq T_6(X_{\hat{v}}). |
A function \alpha: X_{\hat{v}} \times X_{\hat{v}} \to [0, +\infty) exists with parameters such that a \neq 0 , a < 1 , b \neq 0 , and \frac{ac}{b} < 1 . Additionally, let x_0 \in X_{\hat{v}} satisfy \alpha(x_0, x_1) \geq 1 for some \epsilon, \delta > 0 . Each mapping T_i for i = 1, \ldots, 6 is required to be triangular \alpha -orbital admissible mapping for all \lambda > 0 , fulfilling the following specific conditions:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.88) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.89) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.90) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): = a\max\biggl\{\max\{\hat{v}_{\lambda}(T_{6}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}+ \min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}\biggr\}; \end{align} | (3.91) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): = b\max\biggl\{\max\{\hat{v}_{\lambda}(T_{5}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}+ \min\{\hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x), \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y)\}\biggr\}; \end{align} | (3.92) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): = c\max\biggl\{\max\{\hat{v}_{\lambda}(T_{6}x,T_{2}x), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}+ \min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}\biggr\} . \end{align} | (3.93) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} , and for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. Suppose that X_{\hat{v}} is empty; then there is nothing to prove. We now assume that X_{\hat{v}}\neq\emptyset . Then a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge\; 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mappings for every \lambda > 0 . Suppose that the mappings, T_{i} for i = 1, \cdots, 6 satisfy the inequalities (3.72)–(3.90). Since x_{0}, x_{1} and x_{2} are points in X_{\hat{v}} and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) , we can find a point x_{1} in X_{\hat{v}} such that \xi_{0} = T_{1}x_{0} = T_{6}x_{1} . For T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , we can find a point x_{2}\in X_{\hat{v}} such that \xi_{1} = T_{2}x_{1} = T_{5}x_{2} ; for T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , we can find a point x_{3} in X_{\hat{v}} such that \xi_{2} = T_{3}x_{2} = T_{4}x_{3} . Now for all \lambda > 0 , in general, sequences \{x_n\}_{n \in \mathbb{N}} and \{\xi_n\}_{n \in \mathbb{N}} exist within X_{\hat{v}} that satisfy the conditions outlined in Eq (3.7). If a natural number n_0 exists for which \xi_{n_0} = \xi_{n_0 + 1} , then it can be inferred that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) hold. In fact, m\in\mathbb{N} exists such that \xi_{m+2} = \xi_{m+3} , then T_{1}u = T_{6}u , where u = x_{m+3} . Therefore, the pair \{T_{1}, T_{6}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2). Therefore, from the inequalities (3.72)–(3.90), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . By Theorem 1, T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Remark 7. Corollary 4 is a generalization of [34, Corollary 3.14].
Corollary 5. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , the pairs \{T_2, T_4\} , \{T_3, T_5\} , and \{T_1, T_6\} are considered weakly commuting self-mappings that satisfy the condition that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). Moreover, a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge\; 1, \, \, \, \epsilon, \delta > 0 , and, for each i = 1, \cdots, 6 . For all \lambda > 0 , T_i is classified as a triangular \alpha -orbital admissible mapping, provided that the following conditions are met for some positive integer m \geq 1 :
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.94) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.95) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.96) |
where
\begin{align} F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y): = &a\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{6}y)\} \\&+ \min\{\hat{v}_{\lambda}(T^{m}_{3}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\}\biggr\}; \end{align} | (3.97) |
\begin{align} F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y): = &b\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{5}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\} \\ &+\min\{\hat{v}_{\lambda}(T^{m}_{1}x,T^{2m}_{6}x), \hat{v}_{\lambda}(T^{m}_{3}y,T^{2m}_{4}y)\}\biggr\}; \end{align} | (3.98) |
\begin{align} F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y): = &c\max\biggl\{\max\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{2}x), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\} \\&+\min\{\hat{v}_{\lambda}(T^{m}_{3}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{6}y)\}\biggr\}. \end{align} | (3.99) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be defined in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 for some positive integer, m\ge 1 .
Proof. By Corollary 4, p = T^{m}_{1}p = T^{m}_{6}p , p = T^{m}_{2}p = T^{m}_{4}p , and p = T^{m}_{3}p = T^{m}_{5}p . Thus, p = T^{m}_{1}p = T^{m}_{2}p = T^{m}_{3}p = T^{m}_{4}p = T^{m}_{5}p = T^{m}_{6}p or p\in Fix(T^{m}_{i}) for i = 1, 2, \cdots, 6 and some positive integer m\ge 1 , showing that p is a fixed point of T^{m}_{1} and also a fixed point of T^{m}_{2}, T^{m}_{3}, T^{m}_{4}, T^{m}_{5}, \; and\; T^{m}_{6} respectively. Therefore, the uniqueness follows as in Theorem 1 above. Hence, T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 and some positive integer, m\ge 1 .
Corollary 6. Consider X_{\hat{v}} to be a \hat{v} -regular and \hat{v} -complete modular extended b -metric space. Let T_i: X_{\hat{v}} \rightarrow X_{\hat{v}} represent six orbitally continuous mappings that adhere to the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \ldots, 6 . The pairs \{T_2, T_4\} , \{T_3, T_5\} , and \{T_1, T_6\} are identified as weakly commuting self-mappings, satisfying the inclusions:
T_3(X_{\hat{v}}) \subseteq T_4(X_{\hat{v}}), \quad T_2(X_{\hat{v}}) \subseteq T_5(X_{\hat{v}}), \quad T_1(X_{\hat{v}}) \subseteq T_6(X_{\hat{v}}). |
Additionally, a function \alpha: X_{\hat{v}} \times X_{\hat{v}} \rightarrow [0, +\infty) exist with the parameters a \neq 0 , a < 1 , b \neq 0 , and \frac{ac}{b} < 1 . Let x_0 \in X_{\hat{v}} be such that \alpha(x_0, x_1) \geq 1 and \epsilon, \delta > 0 . For each i = 1, \ldots, 6 , it follows that T_i qualifies as triangular \alpha -orbital admissible mapping for every \lambda > 0 , with the following specific conditions being satisfied:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.100) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.101) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.102) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): = a\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\biggr\}; \end{align} | (3.103) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): = b\max\biggl\{\hat{v}_{\lambda}(T_{5}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\biggr\}; \end{align} | (3.104) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): = c\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{2}x), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\biggr\} . \end{align} | (3.105) |
Let the sequences \{x_n\}_{n \in \mathbb{N}} and \{\xi_n\}_{n \in \mathbb{N}} be defined in X_{\hat{v}} as follows. For a given x_n in X_{\hat{v}} , select x_{n+1} such that \xi_n = T_1 x_n = T_6 x_{n+1} . Next, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} so that \xi_{n+1} = T_2 x_{n+1} = T_5 x_{n+2} . Furthermore, for the point x_{n+2} in X_{\hat{v}} , we determine x_{n+3} such that \xi_{n+2} = T_3 x_{n+2} = T_4 x_{n+3} , for n = 0, 1, 2, \ldots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. If we take X_{\hat{v}} to be empty, then there is nothing to prove. Henceforth, we assume that X_{\hat{v}}\neq\emptyset . Then a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mapping for every \lambda > 0 . Suppose that the mappings, T_{i} for i = 1, \cdots, 6 satisfies the inequalities (3.73)–(3.102). Since x_{0}, x_{1} , and x_{2} are points in X_{\hat{v}} and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) , we can find a point x_{1} in X_{\hat{v}} such that \xi_{0} = T_{1}x_{0} = T_{6}x_{1} . For T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , we can find a point x_{2}\in X_{\hat{v}} such that \xi_{1} = T_{2}x_{1} = T_{5}x_{2} ; for T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , we can find a point x_{3} in X_{\hat{v}} such that \xi_{2} = T_{3}x_{2} = T_{4}x_{3} . Now for all \lambda > 0 , in general, there are sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} in X_{\hat{v}} such that Eq (3.7) holds. For any n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} , and T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) hold, if, m\in\mathbb{N} such that \xi_{m+2} = \xi_{m+3} , then T_{1}u = T_{6}u , where u = x_{m+3} . Therefore, the pair \{T_{1}, T_{6}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is an n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, if \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2); therefore, from inequalities (3.73)–(3.102), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . By Theorem 1, T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Remark 8. Corollary 6 is a generalization of [34, Corollary 3.12] and the results in Karapinar et al. [26].
Corollary 7. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , where \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} are weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , the mapping T_{i} is classified as a triangular \alpha -orbital admissible function for every \lambda > 0 , provided that certain conditions are satisfied for a positive integer m \ge 1 .
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.106) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.107) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.108) |
where
\begin{align} F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y): = a\max\biggl\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{6}y)\biggr\}; \end{align} | (3.109) |
\begin{align} F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y): = b\max\biggl\{\hat{v}_{\lambda}(T^{m}_{5}x,T^{m}_{4}y), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\biggr\}; \end{align} | (3.110) |
\begin{align} F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y): = c\max\biggl\{\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{2}x), \hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y)\biggr\} . \end{align} | (3.111) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 for some positive integer, m\ge 1 .
Proof. By Corollary 6, we get p = T^{m}_{1}p = T^{m}_{6}p , p = T^{m}_{2}p = T^{m}_{4}p , and p = T^{m}_{3}p = T^{m}_{5}p . Thus, p = T^{m}_{1}p = T^{m}_{2}p = T^{m}_{3}p = T^{m}_{4}p = T^{m}_{5}p = T^{m}_{6}p , or p\in Fix(T^{m}_{i}) for i = 1, 2, \cdots, 6 and some positive integer, m\ge 1 , showing that p is a fixed point of T^{m}_{1} and also a fixed point of T^{m}_{2}, T^{m}_{3}, T^{m}_{4}, T^{m}_{5}, T^{m}_{6} respectively. Therefore, the uniqueness follows as in Theorem 1 above. Hence, T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 and some positive integer m\ge 1 .
Remark 9. Corollary 7 is a generalization of [34, Corollary 3.13].
Corollary 8. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , where \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mapping for all \lambda > 0 satisfying the following conditions:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.112) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.113) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.114) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): = a\hat{v}_{\lambda}(T_{6}x,T_{4}y); \end{align} | (3.115) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): = b\hat{v}_{\lambda}(T_{5}x,T_{4}y); \end{align} | (3.116) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): = c\hat{v}_{\lambda}(T_{6}x,T_{2}x) . \end{align} | (3.117) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. If we take X_{\hat{v}} to be empty, then there is nothing to prove. Now, we assume that X_{\hat{v}}\neq\emptyset . Then a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge\; 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mappings for every \lambda > 0 . Suppose that the mappings, T_{i} for i = 1, \cdots, 6 satisfy inequalities the (3.74)–(3.114). Since x_{0}, x_{1} , and x_{2} are points in X_{\hat{v}} and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) , we can find a point x_{1} in X_{\hat{v}} such that \xi_{0} = T_{1}x_{0} = T_{6}x_{1} . For T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , we can find a point x_{2}\in X_{\hat{v}} such that \xi_{1} = T_{2}x_{1} = T_{5}x_{2} ; for T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , we can find a point x_{3} in X_{\hat{v}} such that \xi_{2} = T_{3}x_{2} = T_{4}x_{3} . Now for all \lambda > 0 , in general, the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} are in X_{\hat{v}} such that Eq (3.7) hold. If n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} , then T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) hold. In fact, if there is an m\in\mathbb{N} such that \xi_{m+2} = \xi_{m+3} , then T_{1}u = T_{6}u , where u = x_{m+3} . Therefore, the pair \{T_{1}, T_{6}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is an n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, if \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2); therefore, from the inequalities (3.74)–(3.114), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . By Theorem 1, T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Remark 10. Corollary 8 is a generalization of [20, Theorem 2.8] and results in Karapinar et al. [26].
Corollary 9. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 , and, for each i = 1, \cdots, 6 , T_{i} remain a triangular \alpha -orbital admissible mapping for all \lambda > 0 satisfying the following conditions for some positive integer m\ge 1 :
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.118) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.119) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.120) |
where,
\begin{align} F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y): = a\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{4}y); \end{align} | (3.121) |
\begin{align} F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y): = b\hat{v}_{\lambda}(T^{m}_{5}x,T^{m}_{4}y); \end{align} | (3.122) |
\begin{align} F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y): = c\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{2}x) . \end{align} | (3.123) |
Let the two sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 for some positive integer, m\ge 1 .
Proof. By Corollary 8, p = T^{m}_{1}p = T^{m}_{6}p , p = T^{m}_{2}p = T^{m}_{4}p and p = T^{m}_{3}p = T^{m}_{5}p . Thus, p = T^{m}_{1}p = T^{m}_{2}p = T^{m}_{3}p = T^{m}_{4}p = T^{m}_{5}p = T^{m}_{6}p or p\in Fix(T^{m}_{i}) for i = 1, 2, \cdots, 6 and some positive integer, m\ge 1 ; showing that p is a fixed point of T^{m}_{1} and also a fixed point of T^{m}_{2}, T^{m}_{3}, T^{m}_{4}, T^{m}_{5}, \; and\; T^{m}_{6} respectively. Therefore, the uniqueness follows as in Theorem 1 above. Hence, T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 and some positive integer, m\ge 1 .
Corollary 10. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mappings for all \lambda > 0 for which the following hold:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.124) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.125) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.126) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): = a\hat{v}_{\lambda}(T_{6}x,T_{4}y); \end{align} | (3.127) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): = b\hat{v}_{\lambda}(T_{5}x,T_{4}y); \end{align} | (3.128) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): = c\hat{v}_{\lambda}(T_{2}x,T_{5}y) . \end{align} | (3.129) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again. for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. Suppose that X_{\hat{v}} is empty; then there is nothing to prove. We now assume that X_{\hat{v}}\neq\emptyset . Then a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge\; 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mappings for every \lambda > 0 . Suppose that the mappings T_{i} for i = 1, \cdots, 6 satisfy the inequalities (3.124)–(3.126). Since x_{0}, x_{1} , and x_{2} are points in X_{\hat{v}} and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) , we can find a point x_{1} in X_{\hat{v}} such that \xi_{0} = T_{1}x_{0} = T_{6}x_{1} . For T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , we can find a point x_{2}\in X_{\hat{v}} such that \xi_{1} = T_{2}x_{1} = T_{5}x_{2} , and for T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , we can find a point x_{3} in X_{\hat{v}} such that \xi_{2} = T_{3}x_{2} = T_{4}x_{3} . Now for all \lambda > 0 , induce on n so that the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} are in X_{\hat{v}} such that Eq (3.7) hold. For any n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} , then T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) holds. In fact, if m\in\mathbb{N} exists such that \xi_{m+2} = \xi_{m+3} , then T_{1}u = T_{6}u , where u = x_{m+3} . Therefore, the pair \{T_{1}, T_{6}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is an n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, if \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2); therefore, from the inequalities (3.124)–(3.126), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . By Theorem 1, T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Remark 11. (1) If T = T_{1} = \cdots = T_{6} , then the inequalities (3.124)–(3.126) and Eqs (3.127)–(3.129) of Corollary 10 coincides, which is a modification of [20, Theorem 2.8].
(2) T_{1} = T_{2} = T_{3} = T and T_{4} = T_{5} = T_{6} = I , c = 0 , and a+b = 1 , then the inequalities (3.124)–(3.126) and Eqs (3.127)–(3.129) of Corollary 10 coincide with [20, Theorem 2.8].
Corollary 11. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 , and, for each i = 1, \cdots, 6 , T_{i} is triangular \alpha -orbital admissible mapping for all \lambda > 0 and for some positive integer m\ge 1 , for which:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.130) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.131) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.132) |
where
\begin{align} F^{A}_{\lambda}(T^{m}_{1}x,T^{m}_{2}y): = a\hat{v}_{\lambda}(T^{m}_{6}x,T^{m}_{4}y); \end{align} | (3.133) |
\begin{align} F^{B}_{\lambda}(T^{m}_{1}x,T^{m}_{3}y): = b\hat{v}_{\lambda}(T^{m}_{5}x,T^{m}_{4}y); \end{align} | (3.134) |
\begin{align} F^{C}_{\lambda}(T^{m}_{2}x,T^{m}_{3}y): = c\hat{v}_{\lambda}(T^{m}_{2}x,T^{m}_{5}y) . \end{align} | (3.135) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 for some positive integer, m\ge 1 .
Proof. By Corollary 10, p = T^{m}_{1}p = T^{m}_{6}p , p = T^{m}_{2}p = T^{m}_{4}p , and p = T^{m}_{3}p = T^{m}_{5}p . Thus, p = T^{m}_{1}p = T^{m}_{2}p = T^{m}_{3}p = T^{m}_{4}p = T^{m}_{5}p = T^{m}_{6}p , or p\in Fix(T^{m}_{i}) for i = 1, 2, \cdots, 6 and some positive integer m\ge 1 , showing that p is a fixed point of T^{m}_{1} and also a fixed point of T^{m}_{2}, T^{m}_{3}, T^{m}_{4}, T^{m}_{5}, \; and\; T^{m}_{6} respectively. Therefore, the uniqueness follows as in Theorem 1 above. Hence, T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 and some positive integer, m\ge 1 .
Remark 12. Corollary 11 is a generalization of [20, Theorem 2.8].
Corollary 12. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 , and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mapping for all \lambda > 0 satisfying the following conditions:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.136) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.137) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.138) |
where,
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): = a\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{4}y), \frac{1}{3}\{\hat{v}_{\lambda}(T_{1}x,T_{6}y)+ \hat{v}_{\lambda}(T_{3}x,T_{4}y)+ \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}\biggr\}; \end{align} | (3.139) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): = b\max\biggl\{\hat{v}_{\lambda}(T_{5}x,T_{4}y), \frac{1}{3}\{\hat{v}_{\lambda}(T_{2}x,T_{5}y)+ \hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x)+ \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y)\}\biggr\}; \end{align} | (3.140) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): = c\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{2}x), \frac{1}{3}\{\hat{v}_{\lambda}(T_{2}x,T_{5}y)+ \hat{v}_{\lambda}(T_{3}x,T_{4}y)+ \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}\biggr\} . \end{align} | (3.141) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. Upon appeal to Theorem 1, we can see that T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Corollary 13. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mapping for all \lambda > 0 satisfying the following conditions:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.142) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.143) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.144) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): & = a^{2}\max\biggl\{\dfrac{\max\{\hat{v}_{\lambda}(T_{6}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}}{a+\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}},\\ &\dfrac{\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}}{a+\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}}\biggr\}; \end{align} | (3.145) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): & = b^{2}\max\biggl\{\dfrac{\max\{\hat{v}_{\lambda}(T_{5}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}}{b+\min\{\hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x), \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y)\}},\\ &\dfrac{\min\{\hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x), \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y)\}}{b+\min\{\hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x), \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y)\}}\biggr\}; \end{align} | (3.146) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): & = c^{2}\max\biggl\{\dfrac{\max\{\hat{v}_{\lambda}(T_{6}x,T_{2}x), \hat{v}_{\lambda}(T_{2}x,T_{5}y)\}}{c+\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}},\\ &\dfrac{\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}}{c+\min\{\hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y)\}}\biggr\} . \end{align} | (3.147) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. Thanks to Corollary 2, we can see that T_{i} has a common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Corollary 14. Let X_{\hat{v}} be a \hat{v} -regular \hat{v} -complete modular extended b -metric space and let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous mappings satisfying the \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings such that T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}), \, \, \, T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}), \, \, \, T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}). A function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mapping for all \lambda > 0 satisfying the following conditions:
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{2}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{A}_{\lambda}(T_{1}x,T_{2}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.148) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{1}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}( F^{B}_{\lambda}(T_{1}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}; \end{align} | (3.149) |
\begin{align} \alpha(x,y)\hat{v}_{\lambda}(T_{2}x,T_{3}y) < \mathfrak{A(\epsilon)}\implies\epsilon \le\mathfrak{A}^{-1}(F^{C}_{\lambda}(T_{2}x,T_{3}y)) < \mathfrak{A(\epsilon)}+\mathfrak{A(\delta(\epsilon))}, \end{align} | (3.150) |
where
\begin{align} F^{A}_{\lambda}(T_{1}x,T_{2}y): & = a\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y), \hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y),\\ &\frac{\hat{v}_{\lambda}(T_{3}x,T_{4}y)+ \hat{v}_{\lambda}(T_{2}x,T_{5}y)}{2}\biggr\}; \end{align} | (3.151) |
\begin{align} F^{B}_{\lambda}(T_{1}x,T_{3}y): & = b\max\biggl\{\hat{v}_{\lambda}(T_{5}x,T_{4}y), \hat{v}_{\lambda}(T_{2}x,T_{5}y), \hat{v}_{\lambda}(T_{1}x,T^{2}_{6}x), \hat{v}_{\lambda}(T_{3}y,T^{2}_{4}y),\\ &\frac{\hat{v}_{\lambda}(T_{5}x,T_{4}y)+ \hat{v}_{\lambda}(T_{2}x,T_{5}y)}{2}\biggr\}; \end{align} | (3.152) |
\begin{align} F^{C}_{\lambda}(T_{2}x,T_{3}y): & = c\max\biggl\{\hat{v}_{\lambda}(T_{6}x,T_{2}x), \hat{v}_{\lambda}(T_{2}x,T_{5}y), \hat{v}_{\lambda}(T_{3}x,T_{4}y), \hat{v}_{\lambda}(T_{1}x,T_{6}y),\\ &\frac{\hat{v}_{\lambda}(T_{3}x,T_{4}y)+ \hat{v}_{\lambda}(T_{1}x,T_{6}y)}{2}\biggr\} . \end{align} | (3.153) |
Let the sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} be in X_{\hat{v}} so that for x_{n} in X_{\hat{v}} , we choose x_{n+1} such that \xi_{n} = T_{1}x_{n} = T_{6}x_{n+1} ; again, for x_{n+1} in X_{\hat{v}} , we choose x_{n+2} such that \xi_{n+1} = T_{2}x_{n+1} = T_{5}x_{n+2} and, for a point x_{n+2} in X_{\hat{v}} , we choose x_{n+3} such that \xi_{n+2} = T_{3}x_{n+2} = T_{4}x_{n+3} for n = 0, 1, 2, \cdots . Then T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has a unique common fixed point in \bigcap_{i = 1}Fix(T_{i}) for each i = 1, \cdots, 6 .
Proof. If we take X_{\hat{v}} to be empty, then there is nothing to prove. Now, we assume that X_{\hat{v}}\neq\emptyset . Then a function \alpha:X_{\hat{v}}\times X_{\hat{v}} \rightarrow [0, +\infty) , a\neq 0 , a < 1 , b\neq 0 , \frac{ac}{b} < 1 , and x_{0}\in X_{\hat{v}} exist such that \alpha(x_{0}, x_{1})\ge\; 1, \, \, \, \epsilon, \delta > 0 and, for each i = 1, \cdots, 6 , T_{i} remains a triangular \alpha -orbital admissible mappings for every \lambda > 0 . Suppose that the mappings, T_{i} for i = 1, \cdots, 6 satisfy inequalities the (3.148)–(3.150). Since x_{0}, x_{1} , and x_{2} are points in X_{\hat{v}} and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) , we can find a point x_{1} in X_{\hat{v}} such that \xi_{0} = T_{1}x_{0} = T_{6}x_{1} . For T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , we can find a point x_{2}\in X_{\hat{v}} such that \xi_{1} = T_{2}x_{1} = T_{5}x_{2} ; for T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , we can find a point x_{3} in X_{\hat{v}} such that \xi_{2} = T_{3}x_{2} = T_{4}x_{3} . Now for all \lambda > 0 , inductively, sequences \{x_{n}\}_{n\in\mathbb{N}} and \{\xi_{n}\}_{n\in\mathbb{N}} in X_{\hat{v}} exist such that Eq (3.7) hold. For any integer, n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} , then T_{3}(X_{\hat{v}})\subseteq T_{4}(X_{\hat{v}}) , T_{2}(X_{\hat{v}})\subseteq T_{5}(X_{\hat{v}}) , and T_{1}(X_{\hat{v}})\subseteq T_{6}(X_{\hat{v}}) hold. In fact, if m\in\mathbb{N} exists such that \xi_{m+2} = \xi_{m+3} , then T_{1}u = T_{6}u , where u = x_{m+3} . Therefore, the pair \{T_{1}, T_{6}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m} = \xi_{m+1} , then T_{2}u = T_{4}u , where u = x_{m+1} . Therefore, the pair \{T_{2}, T_{4}\} has a coincidence point u\in X_{\hat{v}} . If \xi_{m+1} = \xi_{m+3} , then T_{3}u = T_{5}u , where u = x_{m+2} . Thus, the pair \{T_{3}, T_{5}\} has a coincidence point u\in X_{\hat{v}} . Again, if there is an n_{0}\in\mathbb{N} such that \xi_{n_{0}} = \xi_{n_{0}+1} = \xi_{n_{0}+2} , then \xi_{n} = \xi_{n_{0}} for any n\geq n_{0} . This implies that \{\xi_{n}\} is a modular \hat{v} Cauchy sequence in X_{\hat{v}} . Actually, if \eta\in\mathbb{N} exists such that (1) \xi_{\eta} = \xi_{\eta+1} = \xi_{\eta+2} , (2) \xi_{\eta}\neq\xi_{\eta+1} = \xi_{\eta+2} , (3) \xi_{\eta}\neq\xi_{\eta+2} = \xi_{\eta+1} , and (4) \xi_{\eta}\neq\xi_{\eta+1}\neq\xi_{\eta+2} hold. In fact, Case (1) is easy, and Case (3) is similar to Case (2); therefore, from the inequalities (3.148)–(3.150), we get the result by setting x = \xi_{\eta+2} and y = \xi_{\eta+3} . By Theorem 1, T_{i} has a fixed point in X_{\hat{v}} . Moreover, if \alpha(x^{\ast}, y^{\ast})\ge 1 for all x^{\ast}, y^{\ast}\in \bigcap_{i = 1} Fix(T_{i}) , then T_{i} has common unique fixed point in \bigcap_{i = 1}Fix(T_{i}) for i = 1, \cdots, 6 .
Example 4. Suppose that X_{\hat{v}} = (\mathbb{R}\setminus\{0\})\cup\{\infty\} with the modular extended b -metric as defined in Example 3, which is complete in X_{\hat{v}} for all \lambda > 0 . Define the \hat{v} -weakly commuting mappings T_{1}, T_{2}, T_{3}, T_{4}, T_{5}, T_{6}:(\mathbb{R}\setminus\{0\})\cup\{\infty\}\rightarrow (\mathbb{R}\setminus\{0\})\cup\{\infty\} as follows: T_{1}x = \log_{64}x^{6}, \, \, \, \, T_{2}x = \log_{32}x^{5}, \, \, \, \, T_{3}x = \log_{16}x^{4}, T_{4}x = \log_{8}x^{3}, \, \, \, \, T_{5}x = \log_{4}x^{2}, \, \, \, \, T_{6}x = \log_{2}x for all x\in(\mathbb{R}\setminus\{0\})\cup\{\infty\} and \lambda > 0 , for each i = 1, 2, \cdots, 6 , and also for x, y\in(\mathbb{R}\setminus\{0\})\cup\{\infty\} , \alpha(x, T_{i}x)\geq 1\implies \alpha(T_{i}x, T_{i}^{2}x)\geq1 , and \alpha(x, y)\geq 1 and \alpha(y, T_{i}y)\geq 1\implies\alpha(x, T_{i}y)\geq 1 . Then the mappings T_{1}, T_{2}, T_{3}, T_{4}, T_{5} , and T_{6} satisfy the inequalities (3.151)–(3.153) of Corollary 14. In fact, let T_{i}:X_{\hat{v}}\longrightarrow X_{\hat{v}} be six orbitally continuous \alpha - \hat{v} - A - B - C -Meir-Keeler-type contraction mappings for i = 1, 2, \cdots, 6 , and let \{T_{2}, T_{4}\} , \{T_{3}, T_{5}\} , and \{T_{1}, T_{6}\} be weakly commuting pairs of self-mappings. Actually, it suffices to show that the inequalities (3.1)–(3.3) coincide with inequalities (3.148)–(3.150). Now observe that from Example 3, \hat{v}_{\lambda}(T_{3}x, T_{4}y) = \frac{1}{1+\lambda}\frac{1}{\ln(2)}\max_{x, y\in X_{\hat{v}}}\biggl\{\ln||\frac{x}{y}||\biggr\} and \hat{v}_{\lambda}(T_{2}x, T_{5}y) = \frac{1}{1+\lambda}\frac{1}{\ln(2)}\max_{x, y\in X_{\hat{v}}}\biggl\{\ln||\frac{x}{y}||\biggr\} , and from inequality (3.151), we can see that \frac{\hat{v}_{\lambda}(T_{3}x, T_{4}y)+\hat{v}_{\lambda}(T_{2}x, T_{5}y)}{2} = \frac{1}{1+\lambda}\frac{1}{\ln(2)}\max_{x, y\in X_{\hat{v}}}\biggl\{\ln||\frac{x}{y}||\biggr\} , which is either \hat{v}_{\lambda}(T_{3}x, T_{4}y) or \hat{v}_{\lambda}(T_{2}x, T_{5}y) . Similarly, the inequalities (3.152) and (3.153) hold. It is now evident that the inequalities (3.1)–(3.3) and (3.148)–(3.150) coincides.
In this paper, we introduced the concept of \alpha - \hat{v} -A-B-C-Meir-Keeler-type nonlinear contractions in modular extended b -metric spaces. We established common unique fixed-point theorems that unify and extend existing results in modular b -metric and modular extended b -metric spaces. These findings provide a broader perspective on contraction mappings and deepen the understanding of fixed-point theory in modular settings. The results presented here build on and generalize classical fixed-point theorems, demonstrating the richness of modular extended b -metric spaces.
Daniel Francis: Conceptualization, methodology, validation, formal analysis, investigation, writing-original draft preparation, writing-review and editing, visualization; Godwin Amechi Okeke: Methodology, validation, resources, writing-original draft preparation, Aviv Gibali: Formal analysis, investigation; writing-original draft preparation, and editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We are very grateful to the editor and to the anonymous reviewers for their insightful and careful comments and suggestions, and for taking the trouble to read the long proofs in our manuscript and for their constructive report, which has been very useful for improving our paper.
The authors declare no conflict of interest.
[1] |
A. A. N. Abdou, Fixed points of Kannan maps in modular metric spaces, AIMS Math., 5 (2020), 6395–6403. https://doi.org/10.3934/math.2020411 doi: 10.3934/math.2020411
![]() |
[2] | M. Abtahi, Fixed point theorems for Meir-Keeler type contractions in metric spaces, arXiv Preprint, 2016, 1–11. https://doi.org/10.48550/arXiv.1604.01296 |
[3] |
U. Aksoya, E. Karapinara, I. M. Erhana, V. Rakocevic, Meir-Keeler type contractions on modular metric spaces, Filomat, 32 (2018), 3697–3707. https://doi.org/10.2298/FIL1810697A doi: 10.2298/FIL1810697A
![]() |
[4] |
U. Aksoy, E. Karapinar, I. M. Erhan, Fixed point theorems in complete modular metric spaces and an application to anti-periodic boundary value problems, Filomat, 31 (2017), 5475–5488. https://doi.org/10.2298/FIL1717475A doi: 10.2298/FIL1717475A
![]() |
[5] | U. Aksoy, E. Karapinar, I. M. Erhan, Fixed points of generalized alpha-admissible contractions on b-metric spaces with an application to boundary value problems, J. Nonlinear Convex A., 17 (2016), 1095–1108. |
[6] | A. S. S. Alharbi, H. H. Alsulami, E. Karapinar, On the power of simulation and admissible functions in metric fixed point theory, J. Funct. Space., 2017 (2017). https://doi.org/10.1155/2017/2068163 |
[7] | H. H. Alsulami, S. Gulyaz, I. M. Erhan, Fixed points of \alpha-admissible Meir-Keeler contraction mappings on quasi-metric spaces, J. Inequal. Appl., 84 (2015). https://doi.org/10.1186/s13660-015-0604-9 |
[8] |
M. Arshad, E. Ameer, E. Karapinar, Generalized contractions with triangular alpha-orbital admissible mapping on Branciari metric spaces, J. Inequal. Appl., 63 (2016), 22–34. https://doi.org/10.1186/s13660-016-1010-7 doi: 10.1186/s13660-016-1010-7
![]() |
[9] |
B. Azadifar, G. Sadeghi, R. Saadati, C. Park, Integral type contractions in modular metric spaces, J. Inequal. Appl., 483 (2013), 65–79. https://doi.org/10.1186/1029-242X-2013-483 doi: 10.1186/1029-242X-2013-483
![]() |
[10] | S. Barootkoob, E. Karapinar, H. Lakzian, A. Chanda, Extensions of Meir-Keeler contraction via w-distances with an application, Kragujev. J. Math., 46 (2022), 533–547. |
[11] | E. Canzoneri, P. Vetroa, Fixed points for asymptotic contractions of integral Meir-Keeler type, J. Nonlinear Sci. Appl., 5 (2012), 126–132. |
[12] | C. M. Chen, E. Karapinar, D. O'regan, On (\alpha-\phi)-Meir-Keeler contractions on partial Hausdorff metric spaces, U. Politeh. Buch. Ser. A, 80 (2018), 101–110. |
[13] | V. V. Chistyakov, A fixed point theorem for contractions in metric modular spaces, arXiv Preprint, 2011, 65–92. |
[14] | V. V. Chistyakov, Metric modular spaces, I basic concepts, Nonlinear Anal.-Theor., 72 (2010), 1–14. https://doi.org/10.1016/j.na.2009.04.057 |
[15] | L. B. Ciric, N. Cakic, M. Rajovic, J. S. Ume, Monotone generalized nonlinear contractions in partially ordered metric spaces, Fixed Point Theory A., 11 (2008), 131294. |
[16] | L. B. Ciric, On contraction type mappings, Math. Balk., 1 (1971), 52–57. |
[17] |
P. Debnath, Z. D. Mitrović, S. Y. Cho, Common fixed points of Kannan, Chatterjea and Reich type pairs of self-maps in a complete metric space, São Paulo J. Math. Sci., 15 (2021), 383–391. https://doi.org/10.1007/s40863-020-00196-y doi: 10.1007/s40863-020-00196-y
![]() |
[18] |
P. Debnath, Banach, Kannan, Chatterjea, and Reich-type contractive inequalities for multivalued mappings and their common fixed points, Math. Method. Appl. Sci., 45 (2021), 1587–1596. https://doi.org/10.1002/mma.7875 doi: 10.1002/mma.7875
![]() |
[19] | M. E. Ege, C. Alaca, Some results for modular b-metric spaces and an application to system of linear equations, Azerbaijan J. Math., 8 (2018), 3–14. |
[20] | N. Gholamian, M. Khanehgir, Fixed points of generalized Meir-Keeler contraction mappings in b-metric-like spaces, Fixed Point Theory A., 34 (2016). https://doi.org/10.1186/s13663-016-0507-6 |
[21] |
A. Gholidahneh, S. Sedghi, O. Ege, Z. D. Mitrovic, M. de la Sen, The Meir-Keeler type contractions in extended modular b-metric spaces with an application, AIMS Math., 6 (2020), 1781–1799. https://doi.org/10.3934/math.2021107 doi: 10.3934/math.2021107
![]() |
[22] |
S. Gulyaz, E. Karapinar, I. M. Erhan, Generalized \alpha-Meir-Keeler contraction mappings on Branciari b-metric spaces, Filomat, 17 (2017), 5445–5456. https://doi.org/10.2298/FIL1717445G doi: 10.2298/FIL1717445G
![]() |
[23] |
X. He, M. Song, D. Chen, Common fixed points for weak commutative mappings on a multiplicative metric space, Fixed Point Theory A., 48 (2014), 1–23. https://doi.org/10.1186/1687-1812-2014-48 doi: 10.1186/1687-1812-2014-48
![]() |
[24] | M. Jleli, E. Karapinar, B. Samet, A best proximity point result in modular spaces with the Fatou property, Abstr. Appl. Anal., 2013, 329451. https://doi.org/10.1155/2013/329451 |
[25] |
G. Jungck, Commuting mappings and fixed points, Am. Math. Mon., 83 (1976), 261–263. https://doi.org/10.1080/00029890.1976.11994093 doi: 10.1080/00029890.1976.11994093
![]() |
[26] |
E. Karapinar, S. Czerwik, H. Aydi, (\alpha-\psi)-Meir-Keeler contraction mappings in generalized b-metric spaces, J. Funct. Space., 2018 (2018), 3264620. https://doi.org/10.1155/2018/3264620 doi: 10.1155/2018/3264620
![]() |
[27] |
E. Karapinar, B. Samet, D. Zhang, Meir-Keeler type contractions on JS-(metric) spaces and related fixed point theorems, J. Fixed Point Theory A., 20 (2018), 1–18. https://doi.org/10.1007/s11784-018-0544-3 doi: 10.1007/s11784-018-0544-3
![]() |
[28] | E. Karapinar, P. Kumam, P. Salimi, On \alpha-\psi-Meir-Keeler contractive mappings, Fixed Point Theory A., 2013, 1–12. |
[29] | K. N. V. V. V. Prasad, A. K. Singh, Meir-Keeler type contraction via rational expression, Acta Math. Univ. Comen., 1 (2020), 19–25. |
[30] |
P. Kumam, W. Sintunavarat, S. Sedghi, N. Shobkolaei, Common fixed point of two R-weakly commuting mappings in b-metric spaces, J. Funct. Space., 2015 (2015), 350840. https://doi.org/10.1155/2015/350840 doi: 10.1155/2015/350840
![]() |
[31] | M. Maiti, T. K. Pal, Generalization of two fixed point theorems, B. Calcutta Math. Soc., 70 (1978), 57–61. |
[32] | A. Meir, E. Keeler, A theorem on contraction mapping, J. Math. Anal. Appl., 28 (1969), 326–329. |
[33] |
M. Neog, P. Debnath, S. Radenović, New extension of some common fixed point theorems in complete metric spaces, Fixed Point Theory, 20 (2019), 567–580. https://doi.org/10.24193/fpt-ro.2019.2.37 doi: 10.24193/fpt-ro.2019.2.37
![]() |
[34] |
G. A. Okeke, D. Francis, A. Gibali, On fixed point theorems for a class of \alpha-\hat{v}-Meir-Keeler-type contraction mapping in modular extended b-metric spaces, J. Anal., 2022 (2022), 1–22. https://doi.org/10.1007/s41478-022-00403-3 doi: 10.1007/s41478-022-00403-3
![]() |
[35] |
D. Panthi, Some common fixed point theorems satisfying Meir-Keeler type contractive conditions, Open J. Discrete Math., 8 (2018), 35–47. https://doi.org/10.4236/ojdm.2018.82004 doi: 10.4236/ojdm.2018.82004
![]() |
[36] | S. Park, B. E. Rhoades, Meir-Keeler type contractive conditions, Math. Japonica, 26 (1981), 13–20. |
[37] | V. Parvaneh, S. J. H. Ghoncheh, Fixed points of (\psi, \varphi)_{\Omega}-contractive mappings in ordered p-metric spaces, Global Anal. Discrete Math., 4 (2020), 15–29. |
[38] | O. Popescu, Some new fixed point theorems for \alpha-Geraghty contraction type maps in metric spaces, Fixed Point Theory A., 2014 (2014). https://doi.org/10.1186/1687-1812-2014-190 |
[39] | I. H. N. Rao, K. P. R. Rao, Generalization of fixed point theorems of Meir-Keleer type, Indian J. Pure Ap. Mat., 16 (1985), 1249–1262. |
[40] |
B. Samet, C. Vetro, P. Vetro, Fixed point theorems for \alpha-\psi-contractive type mapping, Nonlinear Anal., 75 (2012), 2154–2165. https://doi.org/10.1016/j.na.2011.10.014 doi: 10.1016/j.na.2011.10.014
![]() |
[41] | S. Sessa, On a weak commutativity condition of mappings in fixed point considerations, Publ. I. Math., 32 (1982), 149–153. |
[42] | Y. R. Sharma, Fixed point and weak commuting mapping, Int. J. Res. Eng. Technol., 2 (2013), 1–4. |