Loading [MathJax]/jax/element/mml/optable/SpacingModLetters.js
Research article Special Issues

Fixed point theorem on CATp(0) metric spaces with applications in solving matrix equations and fractional differential equations

  • This paper aimed to explore fixed point theorems for CMJ generalized mappings in CATp(0) metric spaces. To strengthen the established results, we presented a positive example. In applications, we found the existence of the solution to nonlinear matrix equations, and unique solutions of two scale fractal hybrid fractional differential equations in CATp(0).

    Citation: Mohammad Sajid, Lucas Wangwe, Hemanta Kalita, Santosh Kumar. Fixed point theorem on CATp(0) metric spaces with applications in solving matrix equations and fractional differential equations[J]. AIMS Mathematics, 2025, 10(5): 11131-11158. doi: 10.3934/math.2025505

    Related Papers:

    [1] Ahmed Alamer, Faizan Ahmad Khan . Boyd-Wong type functional contractions under locally transitive binary relation with applications to boundary value problems. AIMS Mathematics, 2024, 9(3): 6266-6280. doi: 10.3934/math.2024305
    [2] Rajagopalan Ramaswamy, Gunaseelan Mani . Application of fixed point result to solve integral equation in the setting of graphical Branciari -metric spaces. AIMS Mathematics, 2024, 9(11): 32945-32961. doi: 10.3934/math.20241576
    [3] 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
    [4] Fatima M. Azmi, Nabil Mlaiki, Salma Haque, Wasfi Shatanawi . Complex-valued controlled rectangular metric type spaces and application to linear systems. AIMS Mathematics, 2023, 8(7): 16584-16598. doi: 10.3934/math.2023848
    [5] Kanagaraj Muthuselvan, Baskar Sundaravadivoo, Kottakkaran Sooppy Nisar, Suliman Alsaeed . Discussion on iterative process of nonlocal controllability exploration for Hilfer neutral impulsive fractional integro-differential equation. AIMS Mathematics, 2023, 8(7): 16846-16863. doi: 10.3934/math.2023861
    [6] 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
    [7] Nizar Souayah, Nabil Mlaiki, Salma Haque, Doaa Rizk, Amani S. Baazeem, Wasfi Shatanawi . A new type of three dimensional metric spaces with applications to fractional differential equations. AIMS Mathematics, 2022, 7(10): 17802-17814. doi: 10.3934/math.2022980
    [8] Saif Ur Rehman, Iqra Shamas, Shamoona Jabeen, Hassen Aydi, Manuel De La Sen . A novel approach of multi-valued contraction results on cone metric spaces with an application. AIMS Mathematics, 2023, 8(5): 12540-12558. doi: 10.3934/math.2023630
    [9] Yujun Cui, Chunyu Liang, Yumei Zou . Existence and uniqueness of solutions for a class of fractional differential equation with lower-order derivative dependence. AIMS Mathematics, 2025, 10(2): 3797-3818. doi: 10.3934/math.2025176
    [10] Hasanen A. Hammad, Hassen Aydi, Choonkil Park . Fixed point results for a new contraction mapping with integral and fractional applications. AIMS Mathematics, 2022, 7(8): 13856-13873. doi: 10.3934/math.2022765
  • This paper aimed to explore fixed point theorems for CMJ generalized mappings in CATp(0) metric spaces. To strengthen the established results, we presented a positive example. In applications, we found the existence of the solution to nonlinear matrix equations, and unique solutions of two scale fractal hybrid fractional differential equations in CATp(0).



    Meir and Keeler [1] introduced the concept of a weak contractive condition for mappings, which guarantees the existence of a fixed point in a complete metric space. In 1981, Ćirić [2] used the continuity feature to expand on the findings established by Meir and Keeler [1] in a complete metric space. Later, Matkowski [3] extended the fixed point theorems of Banach and Kannan as well as some results of Boyd and Wong [4], Meir and Keeler [1], Reich [5], and Wong [6] in a complete metric space. Subsequently, Jachymski [7] demonstrated the equivalent requirements to those of Meir and Keeler [1]. The work of Ćirić [2], Matkowski [3], and Jachymski [7] is collectively abbreviated as CMJ.

    Gromov [8] introduced the concept of CAT(0) spaces, which was later expanded upon with fixed point results in R-trees and CAT(0) spaces by Kirk [9]. Goebel and Reich [10] contributed significant findings on non-expansive mappings, hyperbolic geometry, and uniform convexity. Reich and Shafrir [11] demonstrated results on nonexpansive iterations in hyperbolic spaces. The -convergence theorems in CAT(0) spaces were given by Dhompongsa and Panyanak [12]. Khamsi and Shukri provided generalized CAT(0) spaces in [13]. Results on monotone non-expansive mappings in CATp(0) spaces were provided by Shukri [14]. In CATp(0) metric, the fixed points of Suzuki-generalized nonexpansive mappings were established by Darweesh and Shukri [15].

    Sun and Agarwal [16] investigated functionals associated with nonlinear differential operators in order to determine whether boundary value issues have solutions. They demonstrated that identifying a fixed point of the corresponding nonlinear operator is comparable to solving such issues. They provided significant theoretical contributions about fixed point existence in the context of partially ordered metric spaces, where their study was conducted. Xie et al. [17] used the Banach contraction principle to show that there are solutions to multi-order nonlinear fractional differential equations over the unbounded interval [0,). Additionally, they developed Ulam-Hyers and Ulam-Hyers-Rassias stability, among other forms of stability, for the comparable initial value problems.

    Further, Zhou et al. [18] focused on nonlinear ψ-Hilfer fractional integrodifferential coupled systems defined over a bounded domain. By employing the contraction mapping principle, they proved both the existence and uniqueness of solutions. Additionally, they examined various notions of stability, such as Ulam-Hyers, Ulam-Hyers-Rassias, and semi-Ulam-Hyers-Rassias, within the framework of generalized complete metric spaces. Imran [19] applied the fractal-fractional derivative with a power-law kernel, denoted as FFP0Dα,βx, to analyze magnetohydrodynamic (MHD) viscous fluid flow between two parallel plates. The study also delved into the chaotic dynamics exhibited by the system.

    In modeling complex systems with irregular or hierarchical structures, fractal derivatives offer a robust mathematical tool. Traditionally, hybrid fractional differential equations (HFDEs) utilize standard fractal derivatives characterized by a single scaling exponent, assuming uniform self-similarity. However, many real-world systems display multiscale or heterogeneous behavior, which a single exponent cannot fully capture. To address this, the two-scale fractal derivative extends the standard model by incorporating two distinct fractal dimensions. This approach is especially useful for phenomena like anomalous diffusion, porous media transport, or biological tissue dynamics, where local and global scaling behaviors differ. The two-scale model effectively represents systems with varying fractal dynamics across scales, enabling a more accurate description of localized irregularities in transport or energy processes.

    Recent studies, such as the work by He et al. [20], have highlighted the connection between fractional calculus and fractal geometry and establish the fractal Fick law, the fractal Darcy law, and the fractal Richards equation. According to these theories, the fractional order in differential equations can be interpreted in terms of two-scale fractal dimensions, capturing both the local and global scaling behaviors of heterogeneous materials or systems. This perspective is particularly relevant when modeling media with complex microstructures or hierarchical properties, where classical integer-order models may fall short. By integrating this view, the fractional order becomes more than a fitting parameter; it acquires a geometric and physical interpretation, thereby enriching the modeling framework and enhancing the descriptive power of fractional differential equations where a coupled transport equation can be modeled. Also, it has a wider supplication in MHD (magnetohydrodynamic) fluid flows, where memory and spatial non-locality due to electromagnetic interactions are effectively described through fractional operators. Furthermore, it is applied in MRI (magnetic resonance imaging).

    On the other hand, Bini and Meini [21] focused on solving a class of nonlinear matrix equations that arise in queueing theory, offering analytical insights into the structure and solvability of such equations. Lim [22] presented a solution methodology for the nonlinear matrix equation X=Q+mi=1MiXδiMi using a contraction principle, further extending the applicability of fixed point theory to nonlinear algebraic systems.

    In this study, we employ the two-scale derivative to enhance the descriptive power of HFDEs in modeling multiscale transport phenomena. Its integration provides a more nuanced approach to fractional modeling, especially in media exhibiting dual fractality, such as fractured rock formations, hybrid organic materials, or anomalous diffusion in biological tissues.

    This work's novelty lies in establishing fixed point results for CMJ-type mappings in CATp(0) metric spaces, with an application to matrix equations and the two-scale fractal fractional hybrid differential equation. It is inspired by the contributions of Darweesh and Shukri [15], Kirk [23], Nanjaras et al. [24], Ćirić [2], Matkowski [3], Jachymski [7], and others.

    The structure of the manuscript is as follows. The Introduction provides the motivation and significance of these spaces and mappings in the mathematical analysis in Section 1. Materials and methods covers essential definitions, notations, and auxiliary results necessary for understanding the main findings in Section 2. Section 3 contains the main results section that introduces and proves key fixed point theorems for CMJ mappings, highlighting their generality and improvements over existing results. In Section 4, as an application we investigate the existence of the solution to nonlinear matrix equations in CATp(0). In Section 5, we study unique solutions for two-scale fractal hybrid fractional differential equations by utilizing the theorems of Section 3. The paper concludes by summarizing findings and suggesting directions for future research, supported by a comprehensive list of references.

    Definitions, lemmas, and some preliminary findings are provided in this section to aid in the development of the primary findings.

    According to Bridson and Haefliger [25] and Gromov [8], CAT(0) and CAT(k) spaces have the following properties:

    Definition 2.1. Let (E,.) be a normed vector space and (X,d) be a geodesic metric space. When a comparison triangle ˉ exists in E such that the comparison axioms are satisfied for all c,s and the comparison points ˉc,ˉsˉ, then X is considered a generalized CAT(0) space.

    d(c,s)ˉcˉs.

    If every triangle in X is at least as "thin" as its comparison triangle in the Euclidean plane and the metric space is geodesically connected, it is referred to as a CAT(0) space.

    Definition 2.2. Let k be a real number and (X,d) be a metric space. Let ˉM2k be a comparison triangle of , and let be a geodesic triangle in X. When all c,s and the comparison points ˉc,ˉsˉ are satisfied, then is said to meet the CAT(k) inequality.

    d(c,s)d(ˉc,ˉs).

    Mnk-metric space is a complete, simply linked, Riemannian n-manifold with constant section curvature kR. For every integer n, the space Mnk can be divided into three qualitative classes based on whether k is zero, positive or negative. In order to simplify the notation, we have En=Mn0, Sn=Mn1, and Hn=Mn1. The law of cosines and the triangle inequality are shown to be closely related in each instance. The Euclidean space n-space En is a vector space in Rn that has a scalar product metric.

    (c/s)=ni=1cisi,

    where c=(cicn) and s=(sisn). En is a uniquely geodesic space and the geodesic segments in En are the subset of the form

    [c,s]={λ1s+λ2c | 0λ11},

    where λ2=1λ1. The n-sphere Sn belongs to the set {c=(cicn+1)Rn+1|(c|c)=1}, where (.|.) represents the Euclidean scalar product. Assume that d:Sn×SnR is a function that allocates the unique real number d(C1,C2)[0,π] to each pair (C1,C2)Sn×Sn. For example, cosd(C1,C2)=(C1|C2). The metric is thus d.

    Let d:Hn×HnR be a function that allocates the unique nonnegative number d(C1,C2)0 to each pair (C1,C2)Hn×Hn in such a way that coshd(C1,C2)=(C1|C2). The metric is thus d.

    Definition 2.3. [8,25] For the given real number k, the model spaces Mnk are defined as:

    (ⅰ) Mn0 is the Euclidean space En if k=0;

    (ⅱ) If k>0 then Mnk is obtained from the sphere Sn by multiplying the distance function by the constant 1k.

    Standard n-sphere Sn

    The standard n-sphere of radius 1, Sn, is the set of points in Rn+1 at the unit distance from the origin.

    It comes equipped with the standard round metric, with sectional curvature equal to 1.

    When you multiply distances by 1k, you are scaling the Riemannian metric g by: gk=1kg. This operation does the following: It multiplies all lengths by 1k. It multiplies all areas by 1k. It multiplies all sectional curvatures by k. So, if you start with Sn, which has constant sectional curvature 1, and rescale the metric as described, you get a new Riemannian manifold: Mnk=(Sn,1kg), which is still a sphere (same topology), but now has constant sectional curvature k.

    (ⅲ) If k<0, then Mnk can be obtained from hyperbolic Hn by multiplying the distance function by 1k.

    The hyperbolic space Hn, which has constant negative sectional curvature.

    Standard hyperbolic space Hn

    Hn is the simply connected, complete Riemannian manifold of constant sectional curvature 1.

    It can be modeled in several ways (Poincaré disk, upper half-space, hyperboloid model), but they all describe the same geometric structure.

    If we multiply all distances in Hn by 1k (where k<0), we are scaling the metric g by:

    gk=1kg,

    which has these effects:

    Lengths scale by 1k.

    Sectional curvatures scale by k.

    So if the original curvature was 1, the new curvature is k.

    So, for k<0, we define:

    Mnk=(Hn,1kg),

    which gives a manifold of constant sectional curvature k<0, the same underlying topology as Hn, and a distance function scaled by 1k.

    Proposition 2.1. Mnk is a geodesic metric space.

    (i) If k0, then Mnk is uniquely geodesic and all balls in Mnk are convex.

    (ii) If k>0, then there is a uniquely geodesic segment joining c,sMnk if and only if d(c,s)<πk.

    (iii) If k>0, closed balls in Mnk of radius <πk are convex.

    Gromov's study [8] of CAT(0) spaces was an extensive exploration of CATp(0) spaces, which was initially addressed by Khamsi and Shukri [13], taking into account that the comparison triangle pertains to a generic Banach space. Specifically, the situation where lp, p2 is the Banach space. In a geodesic metric space (X,d), a geodesic triangle (c1,c2,c3) in Figure 1, consists of the three vertices c1,c2,c3 in X along with the geodesic segments connecting each pair of vertices, which form the edges of . A comparison triangle for the geodesics triangle c1,c2,c3=△∈(X,d) is a triangle ˉ(c1,c2,c3)=△(¯c1,¯c2,¯c3) for p2 in the Banach space lp, such that ¯ci¯cj=d(ci,cj), i,j1,2,3. A point ˉc[¯c1,¯c2] is called a comparison point for c[c1,c2] if d(c1,c)=¯c1ˉc.

    Figure 1.  Geodesic triangle.

    Definition 2.4. [13] Let (X,d) be a geodesic metric space and (E,.) be a normed vector space. When a comparison triangle ˉ in lp exists for any geodesic triangle in X, and the comparison axioms are satisfied (that is, for all c,s∈△ and comparison points ˉc,ˉsˉ), d(c,s)ˉcˉs and then X is said to be a CATP(0) space.

    The following is an important theorem of a CATp(0) space [13].

    Theorem 2.1. If s1s22 is the midpoint of geodesic [s1,s2] and c,s1,s2 are in X, then for p2, the comparison axiom entails that

    dp(c,s1s22)12dp(c,s1)+12dp(c,s2)12pdp(s1,s2). (2.1)

    The inequality is known as (CNp), established by Khamis and Shukra [13]. As for lp, for p>2, the (CNp) inequality implies that (CNp)(0) metric spaces are uniformly convex with δ(r,ϵ)1(1ϵp2p)1p, for each ϵ>0 and each r>0. The traditional (CN) inequality of Bruhat and Tits [26] is reduced to the (CNp) inequality for p=2. The CAT(0) inequality means that if z=s1s22 is the midpoint of the segment [s1,s2] and c,s1,s2 are points in CAT(0) space, then

    d(c,z)212d(c,s1)2+12d(c,s2)214d(s1,s2)2. (2.2)

    Recall uniform convexity of a Banach space X as follows:

    Definition 2.5. [27] For each ϵ, 0<ϵ2, the inequalities c1,s1,csϵ imply there exists δ=δ(ϵ)>0 such that c+s21δ. The midpoint of c and s lies inside the unit ball BX at a distance of at least δ from the unit sphere δX. This indicates that c and s are in the closed ball BX:={cX:c1} with csϵ>0.

    Example 2.1. [27] A uniformly convex space is any Hilbert space H. The parallelogram law provides us with c+s2=2(c2+s2)cs2, for all c,sH. Suppose c,sBH with c  s and csϵ. Then cs24ϵ2. So it follows that c+s21δ(ϵ), where δ(ϵ)=11ϵ24. Therefore H is uniformly convex.

    Let R be a set of real numbers. A metric space (X,d) defined by a mapping γ:RX in a metric embedding of R into X are the definitions of CATp(0) metric spaces with d(γ(c),γ(s))=|cs|,c,sR. The image R under a metric embedding is called a metric line. The image γ([a,b])X is called a metric segment.

    Definition 2.6. [28,29] Let X be a space with metrics. In X, a geodesic path is a path γ:[a,b]X.

    Assume c,sX. If γ(b)=s and γ(a)=c, then γ([c,s]) is said to link c and s. This demonstrates the hyperbolic type of (X,d).

    Lemma 2.1. [30] Let γ:RX be a metric embedding abR and t[0,1]. Then

    (i) d(γ(a),γ(λ2aλ1b))=λ1d(γ(a),γ(b)),

    (ii) d(γ(b),γ(λ2aλ1b))=λ2d(γ(a),γ(b)).

    Khamis et al. [13] introduced a property of a CATp metric space. The CAT(0) metric space (X,d) is said to be convex whenever [c,s]K, for any c,sK. Consider map γ:XR is a type function if there exists a bounded sequence {cn} in X such that γ(c)=limsupid(c,ci).

    Theorem 2.2. [13] Let p2. Consider (X,d) as a complete CATp(0) metric space, and K is any closed, bounded, convex, non-empty subset of X. Assume that a type function specified on K is γ. Any γ minimizing sequence is then convergent. Its limit z satisfies and is the unique minimum of γ and

    γp(z)+12p1dp(z,c)γp(c), (2.3)

    for any cK.

    Definition 2.7. [29] Let E be a vector space. An affinely convex subset XE is defined as follows: for any c,sX and λ1,λ2(0,1) with 1λ1=λ2, the affine segment [c,s]:={λ2c+λ1s:λ[0,1]} is contained in X.

    If X has a family of metric segments, then there exists a unique metric line joining c and s for every pair of distinct points c and s in X. The unique metric segment connecting the two points c and s from X is indicated by the notation [c,s] or [s,c]. This demonstrates that, for all cX, [c,c]={c}.

    Proposition 2.2. [30] Let (X,d) be a hyperbolic metric space. Assume c,sX. There exists a unique point z[c,s] for every λ[0,1], such that d(c,z)=λ1d(c,s), and d(s,z)=λ2d(c,s)-such points will be denoted by z=λ2cλ1s. For z[c,s], this imply that d(c,z)+d(z,s)=d(c,s).

    Definition 2.8. [10] (X,d) is a hyperbolic metric space if

    (ⅰ) d(λ2cλ1s,(λ2cλ1z)λ1d(s,z),

    (ⅱ) d(λ2cλ1s,λ2zλw)λ2d(c,z)+λ1d(s,w),

    for any λ1,λ2[0,1] and all c,s,z,wX. Note that every hyperbolic space is a space of hyperbolic type [30].

    Next, we recall that Hadamard manifolds [10], the Hilbert open unit ball [23], and CATp(0) metric spaces [13] are some examples of normed spaces which are hyperbolic metric spaces. Opial [31] introduced an inequality for a well-convergent sequence characterizing its limits as follows:

    Definition 2.9. Suppose there are Banach spaces X. If every sequence {cn} weakly converges to c for every c in X, then X meets Opial's condition:

    liminfncns>liminfncnc, (2.4)

    which holds for sc.

    Browder [32] obtained an equivalent definition by replacing (2.4) by

    limsupncns>limsupncnc. (2.5)

    The following is the extension of p-uniform convexity to the set of geodesic space by Naor and Silberman [33]:

    Definition 2.10. Let 1<p< be fixed. If (X,d) is geodesic, it is called p-uniformly convex if there exist a constant μ>0 such that c,s,zX with every λi[0,1], and for all i=1,2, we have:

    dp(λ2cλ1s,z)λ2dp(c,z)+λ1dp(s,z)μ2λ1λ2dp(c,s). (2.6)

    One should bear in mind that the space X is confirmed to be uniquely geodesic based on the inequality stated above. Darweesh et al. [15] proved the following theorem on CATp(0) space.

    Theorem 2.3. [15] Given a complete CATp(0) metric space X with p2, let K be a non-empty bounded, closed, convex subset of that space. Given a Suzuki generalized non-expansive mapping T:KK, T has a fixed point.

    Lemma 2.2. [34] Let K be a non-empty subset of a metric space X. Suppose T:KK is a Suzuki generalized non-expansive mapping. Then d(c,Ts)3d(Tc,c)+d(c,s), for all c,sK.

    Lemma 2.3. [34] Assume that the generalized metric space is (X,d). If we have TcnTz such that T:XX is continuous at zX, then limnd(Tcn,Tz)=0, for any sequence {cn} in X, converges to zX. That is cnz.

    The following is the relationship between the aforementioned lemma and the Ćirić [2], Jachymski [7], and Matkowski [3] (CMJ) fixed point theorem.

    Theorem 2.4. [35] Let X be a CMJ contraction on X, and let (X,d) be a complete v-generalized metric space. This means that the following is true:

    (ⅰ) For every ϵ>0, there exists δ>0 such that d(c,s)<ϵ+δ implies d(Tc,Ts)ϵ for any c,sX.

    (ⅱ) cs implies that d(Tc,Ts)<d(c,s) for any c,sX.

    Then T has a unique fixed point z of T. Moreover limnd(Tnc,z)=0 for any c,sX.

    A weak contractive condition that grants the existence of a fixed point in all metric space was developed by Meir et al. [1]. In their investigation, they considered ϵ>0, where there exists δ>0 such that ϵd(c,s)<ϵ+δ implies d(Tc,Ts)<ϵ.

    Theorem 2.5. [1] Assume that T is a mapping from X into itself and that (X,d) is a full metric space. Then, there is a single fixed point ζ for T. Additionally, limnTnc=ζ for any cX.

    The following theorem is proved in metric space by Ćirić [2] using the preceding principles for contractive mappings with the continuity property.

    Theorem 2.6. [2] Given ϵ>0, there exists δ>0 such that, given (X,d), a full metric space, and T a self mapping of X into itself, satisfying the condition ϵ<d(c,s)<ϵ+δ implies d(Tc,Ts)ϵ. Then T has a unique fixed point ζX and limnTnc=ζ for each x˚.

    Additionally, Matkowski [3] established the subsequent theorem.

    Theorem 2.7. [3] Let \mathfrak{T}:\mathfrak{X}\to \mathfrak{X} be a full metric space, and let (\mathfrak{X}, d) be its metric space. Assume that for each c, s \in \mathfrak{X} and \epsilon > 0 ,

    0 \leq \max\Big\{d(\mathfrak{T}c,s),d(s,\mathfrak{T}s),d(c,s),\frac{d(\mathfrak{T}c,s)+d(c,\mathfrak{T}s)}{2}\Big\}\leq\epsilon \Rightarrow \ d(\mathfrak{T}c,\mathfrak{T}s) < \epsilon.

    If for every \epsilon > 0 there is a \delta > 0 such that for c, s \in \mathfrak{X}:

    \left. \begin{array}{l} \epsilon\leq d(c, s) < \epsilon+\delta\nonumber \\ 0 < \max\Big\{d(\mathfrak{T}c, s), \frac{d(\mathfrak{T}c, s)+d(c, \mathfrak{T}s)}{2}\Big\}\leq \epsilon \\ d(s, \mathfrak{T}s) < \epsilon+\delta \end{array} \right\}\Rightarrow \ d(\mathfrak{T}c, \mathfrak{T}s) < \epsilon, then for every c\in\mathfrak{X} , the sequence \{\mathfrak{T}^{n}c\} converges. Moreover, if \mathfrak{T} is continuous, or, given \epsilon > 0 , there is a, \mu , 0 < \mu < \epsilon such that for every c, s \in \mathfrak{X},

    \left. \begin{array}{l} 0 < \max\Bigg\{d(\mathfrak{T}c, c), \frac{d(\mathfrak{T}c, s)+d(c, \mathfrak{T}s)}{2}\Bigg\}\leq \epsilon \\ 0 < { d(c, s), d(s, \mathfrak{T}s)} < \mu \end{array} \right\}\Rightarrow \ d(\mathfrak{T}c, \mathfrak{T}s) < \epsilon, and then \mathfrak{T} has a unique fixed point \zeta \in \mathfrak{X} and \lim\limits_{n\to\infty}\mathfrak{T}^{n}c = \zeta for each c \in \mathfrak{X} .

    Furthermore, the following Mier Keeler-type theorem was demonstrated by Jachymski [7].

    Theorem 2.8. [7] Let (\mathfrak{X}, d) be a complete metric space and let \mathfrak{T}:\mathfrak{X}\to \mathfrak{X} be a self map. If \mathfrak{T} satisfies

    (ⅰ) for every \epsilon > 0 and \delta > 0 , for any c, s \in \mathfrak{X} ,

    \begin{eqnarray} \epsilon& < & \max\Big\{d(c,s),d(c,\mathfrak{T}c),d(s,\mathfrak{T}s),\frac{d(c,\mathfrak{T}s)+d(s,\mathfrak{T}c)}{2}\Big\} < \epsilon+\delta, \Rightarrow d(\mathfrak{T}c,\mathfrak{T}s)\leq \epsilon, \end{eqnarray}

    (ⅱ) for every \epsilon > 0 and \delta > 0 , for any c, s \in \mathfrak{X} ,

    \begin{eqnarray} d(\mathfrak{T}c,\mathfrak{T}s)& < & \max\Big\{d(c,s),d(c,\mathfrak{T}s),d(s,\mathfrak{T}s),\frac{d(c,\mathfrak{T}s)+d(c,\mathfrak{T}c)}{2}\Big\} \end{eqnarray}

    then \mathfrak{T} has a unique fixed point \zeta \in \mathfrak{X} and \lim\limits_{n\to\infty}\mathfrak{T}^{n}c = \zeta for each c \in \mathfrak{X} .

    In this section, we shall cover our main results. We start the section with the following theorem.

    Theorem 3.1. Let K\in\mathfrak{X} be a non-empty, bounded, closed, weakly uniformly convex subset of a CAT_{p}(0) metric space, where p\geq2 , \lambda_{1}, \lambda_{2}\in[0, 1] , and let \mathfrak{T}:\mathfrak{X}\to \mathfrak{X} be a CMJ-type mapping. Suppose that for every \epsilon > 0 and c, s \in \mathfrak{X} ,

    \begin{eqnarray} 0 < \max\Big\{d^{p}(\mathfrak{T}c,s),d^{p}(s,\mathfrak{T}s),d^{p}(c,s),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\}\leq\epsilon \Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon, \end{eqnarray} (3.1)

    then for every c\in\mathfrak{X} , the sequence \{\mathfrak{T}^{n}c\} converges. Furthermore, if \mathfrak{T} is continuous, or if, for any c, s \in \mathfrak{X} , there exists \mu , \; 0 < \mu < \epsilon such that

    \begin{eqnarray} \left. \begin{array}{l} 0 < \max\Big\{d^{p}(\mathfrak{T}c,c),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\}\leq \epsilon \\ 0 < \max\Big\{d^{p}(c,s), d^{p}(s,\mathfrak{T}s)\Big\} < \mu \end{array} \right\}\Rightarrow \ d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon-\mu. \end{eqnarray}

    Then \mathfrak{T} has a unique fixed point z \in \mathfrak{X} and for each c \in \mathfrak{X} , \lim\limits_{i\to\infty}\mathfrak{T}^{i}c = z.

    Proof. Let K be a bounded, closed convex subset of \mathfrak{X} . For simplicity, let \mathfrak{X} represnted CAT_{p}(0) -metric spaces. Assume c_{0}\in K . Now, for all i\geq0 , \{c_{i}\} is a Picard sequence defined by c_{i} = \mathfrak{T}^{i}c_{0} . Clearly for c_{i} = c_{i+1} , we can have \mathfrak{T}^{i}c_{0} be a fixed point of \mathfrak{T}. This means that for some i, \; \mathfrak{T}^{i}c_{0} = \mathfrak{T}^{i+1}c_{0}. If possible, let \mathfrak{T}^{i}c_{0}\neq\mathfrak{T}^{i+1}c_{0} . If c = c_{0} and s = \mathfrak{T}c_{0} , then from the inequality (3.1), we have

    \begin{eqnarray} 0&\leq& \max\Big\{d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0}),d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}),d(c_{0},\mathfrak{T}c_{0}),\frac{d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0})+d^{p}(c_{0},\mathfrak{T}^{2}c_{0})}{2}\Big\}\leq\epsilon\\&& \Rightarrow d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}) < \epsilon. \end{eqnarray} (3.2)

    Now using Lemma 2.1, we can compute all metrics of the above inequality as follows.

    \begin{eqnarray} d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0})& = & d^{p}(\gamma(\mathfrak{T}c_{0}), \gamma(\lambda_{2}\mathfrak{T}c_{0}\oplus\lambda_{1} \mathfrak{T}c_{0})) = \lambda_{1} d^{p}(\gamma(\mathfrak{T}c_{0}),\gamma(\mathfrak{T}c_{0}))\\& = &\lambda_{1} d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0}) = 0. \\ d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0})& = & d^{p}(\gamma(\mathfrak{T}c_{0}), \gamma(\lambda_{2}\mathfrak{T}c_{0}\oplus\lambda_{1} \mathfrak{T}^{2}c_{0})) = \lambda_{1} d^{p}(\gamma(\mathfrak{T}c_{0}),\gamma(\mathfrak{T}^{2}c_{0}))\\& = &\lambda_{1} d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}). \\ d^{p}(c_{0},\mathfrak{T}c_{0})& = & d^{p}(\gamma(c_{0}), \gamma(\lambda_{2}c_{0}\oplus\lambda_{1} \mathfrak{T}c_{0})) = \lambda_{1} d^{p}(\gamma(c_{0}),\gamma(\mathfrak{T}c_{0}))\\& = &\lambda_{1} d^{p}(c_{0},\mathfrak{T}c_{0}). \\ d^{p}(c_{0},\mathfrak{T}^{2}c_{0})& = & d^{p}(\gamma(c_{0}), \gamma(\lambda_{2}c_{0}\oplus\lambda_{1} \mathfrak{T}^{2}c_{0})) = \lambda_{1} d^{p}(\gamma(c_{0}),\gamma(\mathfrak{T}^{2}c_{0}))\\& = &\lambda_{1} d^{p}(c_{0},\mathfrak{T}^{2}c_{0}). \end{eqnarray}

    Substituting the derived metrics in (3.2), we obtain

    \begin{eqnarray} 0& < & \max\Big\{\lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0}),\lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}),\lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0}),\\&&\frac{d^{p}(\mathfrak{T}c_{0},\mathfrak{T}c_{0})+\lambda_{1}d^{p}(c_{0},\mathfrak{T}^{2}c_{0})}{2}\Big\}\leq\epsilon\Rightarrow \lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}) < \epsilon, \\ 0& < & \max\Big\{0,\lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}),\lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0}),\frac{\lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0})+\lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0})}{2}\Big\}\leq\epsilon\\&& \Rightarrow \lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}) < \epsilon. \end{eqnarray}

    Suppose \frac{\lambda_{1}d^{p}(c_{0}, \mathfrak{T}c_{0})+\lambda_{1}d^{p}(\mathfrak{T}c_{0}, \mathfrak{T}^{2}c_{0})}{2}\leq \lambda_{1}d^{p}(c_{0}, \mathfrak{T}c_{0}) , and we get

    \begin{eqnarray} 0& < & \max\Big\{0,\lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}),\lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0}),\lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0})\Big\}\leq\epsilon \Rightarrow \lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}) < \epsilon, \\ 0& < & \lambda_{1}d^{p}(c_{0},\mathfrak{T}c_{0})\Rightarrow \lambda_{1}d^{p}(\mathfrak{T}c_{0},\mathfrak{T}^{2}c_{0}) < \epsilon. \end{eqnarray}

    Next, by using mathematical induction, we have

    \begin{eqnarray} 0& < & \lambda_{1}\lim\limits_{i\to\infty}d^{p}(\mathfrak{T}^{n}c_{0},\mathfrak{T}^{n+1}c_{0})\Rightarrow \lambda_{1}\lim\limits_{i\to\infty}d^{p}(\mathfrak{T}^{i}c_{0},\mathfrak{T}^{i+1}c_{0}) < \epsilon. \end{eqnarray}

    Hence, we get

    \begin{eqnarray} 0& < & \lim\limits_{i\to\infty}d^{p}(\mathfrak{T}^{i}c_{0},\mathfrak{T}^{i+1}c_{0}) < \epsilon. \end{eqnarray}

    This is a contradiction to our assumption. So the sequences \{d^{p}(\mathfrak{T}^{i}c_{0}, \mathfrak{T}^{i+1}c_{0})\} converge.

    Further, let (\mathfrak{X}, d) be a complete CAT_{p}(0) metric space and \mathfrak{T} is closed, bounded, and uniformly convex. We need to prove \{\mathfrak{T}^{i}c_{0}\} is a Cauchy sequenc. Let us construct a sequence \{\mathfrak{T}^{i}c_{0}\} in K in such a way that \mathfrak{T}c = \lim\limits_{i\to\infty}d(c_{i}, c) . Let \gamma_{0} = \inf\{\gamma(c):c\in K \} and \{s_{i}\} be a \gamma minimized sequence. Since K is bounded, then for any c, s \in K , there exists an R > 0 such that d(c, s)\leq R . Since (\mathfrak{X}, d) is a CAT_{p}(0) metric space, it is enough to show that the inequality (2.6) of Definition 2.10 is satisfied.

    \begin{eqnarray} d^{p}(c_{i},\lambda_{2}s_{i}\oplus \lambda_{1} s_{j})&\leq&\lambda_{2}d^{p}(c_{i},s_{i})+\lambda_{1} d^{p}(s_{j},c_{i})-\lambda_{1}\lambda_{2}d^{p}(s_{i},s_{j}). \end{eqnarray} (3.3)

    By applying inequality (2.3) from Theorem 2.2, we can derive the inequality (3.3), which yields

    \begin{eqnarray} \gamma^{p}(c_{i},\lambda_{2}s_{i}\oplus \lambda_{1} s_{j})&\leq&\lambda_{2}\gamma^{p}s_{i}+\lambda_{1} \gamma^{p}s_{j}-\lambda_{1}\lambda_{2}\gamma^{p}(s_{i},s_{j}). \end{eqnarray}

    This gives a conclusion that for all s_{i, j}\geq1 , we have

    \begin{eqnarray} \gamma^{p}_{0} \leq \lambda_{2}\gamma^{p}s_{i}+\lambda_{1} \gamma^{p}s_{j}-\lambda_{1}\lambda_{2}\gamma^{p}(s_{i},s_{j}). \end{eqnarray}

    As \{s_i\} is a minimizing sequence for \gamma , we have \lim\limits_{i, j \to \infty} d(s_i, s_j) = 0 , implying that \{s_i\} is Cauchy. Consequently, \{s_i\} converges to some z \in K with \gamma_0 = \gamma(z).

    Since \gamma is continuous and \mathfrak{X} is complete, Lemma 2.2 gives us

    \begin{eqnarray} d^{p}(z,\mathfrak{T}^{i} z)&\leq&3d^{p}(\mathfrak{T}^{i}z,z)+d^{p}(z,z), \\ \lim\sup\limits_{i\to \infty} d^{p}(z,\mathfrak{T}^{i} z)&\leq&3 \lim\sup\limits_{i\to \infty}d^{p}(\mathfrak{T}^{i} z,z)+ \lim\sup\limits_{i\to \infty}d^{p}(z,z), \\ \lim\sup\limits_{i\to \infty} d^{p}(z,\mathfrak{T} z)&\leq&\frac{1}{2} \lim\sup\limits_{i\to \infty}d^{p}(z,z). \end{eqnarray}

    Therefore, d^{p}(z, \mathfrak{T}z) = 0 , which implies that z is a fixed point of \mathfrak{T} .

    Case 1: If \mathfrak{T} is continuous, then taking limits yields:

    z = \lim\limits_{i \to \infty} \mathfrak{T}^i c = \lim\limits_{i \to \infty} \mathfrak{T}c_{0} = \mathfrak{T}z,

    and hence z is a fixed point.

    Case 2: Suppose instead that the second condition in the theorem holds. Assume, for contradiction, that \mathfrak{T}z \neq z . Then d^p(\mathfrak{T}z, z) > 0 . Since c_i \to z , for sufficiently large i , we have

    \begin{aligned} &0 < \max\left\{ d^p(\mathfrak{T}z, z), \frac{d^p(\mathfrak{T}z, c_i) + d^p(z, \mathfrak{T}c_i)}{2} \right\} \leq \epsilon, \\ &0 < \max\left\{ d^p(\mathfrak{T}z, z), \frac{d^p(\mathfrak{T}z, z) + d^p(z, \mathfrak{T}z}{2} \right\} \leq \epsilon, \\ &0 < \max\left\{ d^p(\mathfrak{T}z, z), d^p(z, \mathfrak{T}z \right\} \leq \epsilon, \\ &0 < d^p(z, \mathfrak{T}z ) \leq \epsilon \end{aligned}

    and

    \begin{aligned} &0 < \max\left\{ d^p(z, c_i), d^p(c_i, \mathfrak{T}c_i) \right\} < \mu, \\ &0 < \max\left\{ d^p(z, z), d^p(z, \mathfrak{T}z) \right\} < \mu, \\ &0 < d^p(z, \mathfrak{T}z) < \mu, \end{aligned}

    for some 0 < \mu < \epsilon . Then the hypothesis implies

    d^p(\mathfrak{T}z, \mathfrak{T}c_i) < \epsilon - \mu,
    d^p(\mathfrak{T}z, \mathfrak{T}z) < \epsilon - \mu,
    0 < \epsilon - \mu.
    0 < \epsilon - \mu < d^p(z, \mathfrak{T}z ) \leq \epsilon.

    Letting i \to \infty , the right-hand side remains positive while the left-hand side tends to d^p(\mathfrak{T}z, z) , a contradiction. Hence, \mathfrak{T}z = z .

    Next for uniqueness: let \gamma(\mathfrak{T} z)\leq \gamma(z) . According to the asymptotic center's uniqueness, \mathfrak{T} z = z .

    Given s_{i } = \zeta and s_{j} = z , we may use inequality (2.6) to determine the uniqueness of \mathfrak{T} . Since,

    \begin{eqnarray} d^{p}(\zeta_{i},\lambda_{2}\zeta\oplus \lambda_{1} z)&\leq&\lambda_{2}d^{p}(\zeta_{i},\zeta)+\lambda_{1} d^{p}(z,\zeta_{i})-\lambda_{1}\lambda_{2}d^{p}(\zeta,z), \end{eqnarray}

    by taking the limit on each side of the inequality shown above, we arrive at

    \begin{eqnarray} \lim\sup\limits_{i\to \infty} d^{p}(\zeta_{i},\lambda_{2}\zeta\oplus \lambda_{1} z)&\leq& \lim\sup\limits_{i\to \infty}\lambda_{2}d^{p}(\zeta_{i},\zeta)+\\&& \lim\sup\limits_{i\to \infty}\lambda_{1} d^{p}(z,\zeta_{i})- \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z). \end{eqnarray}

    Since, \lim\sup\limits_{i\to \infty} d^{p}(\zeta_{i}, z)\leq \lim\sup\limits_{i\to \infty} d^{p}(\zeta_{i}\lambda_{2}\zeta\oplus \lambda_{1} z), then we have:

    \begin{eqnarray} \lim\sup\limits_{i\to \infty} d^{p}(\zeta_{i}, z)&\leq& \lim\sup\limits_{i\to \infty}\lambda_{2}d^{p}(\zeta_{i},\zeta)+\\&& \lim\sup\limits_{i\to \infty}\lambda_{1} d^{p}(z,\zeta_{i})- \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z), \\ \lim\sup\limits_{i\to \infty} d^{p}(z, z)&\leq& \lim\sup\limits_{i\to \infty}\lambda_{2}d^{p}(z,\zeta)+\\&& \lim\sup\limits_{i\to \infty}\lambda_{1} d^{p}(z,z)- \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z), \\ 0&\leq& \lim\sup\limits_{i\to \infty}\lambda_{2}d^{p}(z,\zeta)- \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z), \\ 0&\leq& \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z). \\ \lim\sup\limits_{i\to \infty}\lambda_{1}\lambda_{2}d^{p}(\zeta,z)&\geq&0, \\ \lim\sup\limits_{i\to \infty}d^{p}(\zeta,z)&\geq&0, \end{eqnarray}

    which is a conflict. Thus, \zeta = z . In conclusion, z represents a single fixed point in \mathfrak{T} and Opial's condition is equal to this argument.

    The second important work extends Theorem 2.6 to consider an analog of the theorem in the context of Meir and Keeler-type uniformly convex contractive mappings in CAT_{p}(0) space.

    Theorem 3.2. Given \epsilon > 0 , there exists \delta > 0 such that, given (\mathfrak{X}, d) , a full CAT_{p}(0) , where p\geq2 , \lambda_{1}, \lambda_{2}\in[0, 1] , and let K be a non-empty, bounded, closed, convex subset of a CAT_{p}(0) metric space and \mathfrak{T}:K \to K is a convex contractive mapping, satisfying the condition

    \begin{eqnarray} \epsilon& < & d^{p}(c,s) < \epsilon+\delta \implies d^{p}(\mathfrak{T}c,\mathfrak{T}s)\leq \epsilon. \end{eqnarray} (3.4)

    Then \mathfrak{T} has a unique fixed point \zeta \in \mathfrak{X} and \lim\limits_{n\to\infty}\mathfrak{T}^{i}c_{0} = \zeta for each c \in \mathfrak{X} .

    Proof. Let c_{0}\in \mathfrak{X} . Let us construct a series \{c_{i}\} by c_{i} = \mathfrak{T}^{i}c_{0} for each i\in \mathbb{N} . Then c_{i} is a fixed point on \mathfrak{T} if there exists i such that c_{i} = c_{i+1} . The proof is hereby finalized. Alternatively, suppose that for every i\geq0 , c_{i}\neq c_{i+1} . As c = c_{0} and s = \mathfrak{T}^{i}c_{0} in (3.4) implies that \mathfrak{T} is contractive, therefore real sequence d^{p}(c_{0}, \mathfrak{T}^{i}c_{0}) is non-growing and has a limit \epsilon\geq 0. Based on monotonicity, we obtain

    \begin{eqnarray} \lambda_{1}d^{p}(c_{0},\mathfrak{T}^{i}c_{0}) > \epsilon, for\ i = 0,1,2,\dots \end{eqnarray} (3.5)

    Let us assume \epsilon > 0 . Then, for all i such that \delta = \delta(\epsilon) > 0 ,

    \begin{eqnarray} \epsilon < \lambda_{1}d^{p}(c_{0},\mathfrak{T}^{i}c_{0}) < \epsilon +\delta. \end{eqnarray} (3.6)

    By (3.4),

    \begin{eqnarray} \lambda_{1}d^{p}(\mathfrak{T}^{i}c_{0},\mathfrak{T}^{i+1}c_{0})\leq \epsilon, \end{eqnarray} (3.7)

    which is a contradiction with (3.5). Therefore \epsilon = 0 and

    \begin{eqnarray} \lambda_{1}d^{p}(\mathfrak{T}^{i}c_{0},\mathfrak{T}^{i+1}c_{0}) = 0, \\ d^{p}(\mathfrak{T}^{i}c_{0},\mathfrak{T}^{i+1}c_{0}) = 0. \end{eqnarray} (3.8)

    Let us assume c_{0}\in K . Using the argument of induction, we will create a sequence \{c_{i}\} in K such that, for any i\geq0 , the sequence \{\mathfrak{T}c_{i}\} has a point z linked with it in Theorem 2.2. For any integer i \geq0 ,

    \begin{eqnarray} \epsilon = r_{i}& = &\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(c_{i},\mathfrak{T}^{j}c_{i}), \end{eqnarray} (3.9)
    \begin{eqnarray} \delta = R_{i}& = &\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(\mathfrak{T}^{j}c_{i},\mathfrak{T}^{j+1}c_{i}). \end{eqnarray} (3.10)

    By (3.9) and (3.10), we have

    \begin{eqnarray} r_{i}& < &\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(c_{i},\mathfrak{T}^{j+1}c_{i}) < \lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(c_{i},\mathfrak{T}^{j}c_{i})+\\&&\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(\mathfrak{T}^{j}c_{i},\mathfrak{T}^{j+1}c_{i}), \\ r_{i}& < &\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(c_{i},\mathfrak{T}^{j+1}c_{i}) < r_{i}+R_{i}, \\ r_{i}& < &\lambda_{1}\lim\sup\limits_{j\to \infty} d^{p}(c_{i},\mathfrak{T}^{j+1}c_{i}) < 2R_{i} < r_{i}. \end{eqnarray} (3.11)

    Hence by (3.4) for i\geq 1 , the series \sum_{i = j = 1}^{\infty}\lambda_{i}d^{p}(c_{i}, \mathfrak{T}^{j+1}c_{i}) is also convergent and therefore \{c_{i}\} is Cauchy. The remaining steps to demonstrate that \{c_{i}\} , the fixed point's existence and uniqueness, follow Theorem 3.1's proof in a similar manner. This concludes the proof.

    Utilizing the concepts derived from CMJ-type contractive mappings within the context of CAT_{p}(0) space, we establish the subsequent corollary that supports Theorem 3.1.

    Corollary 3.1. Let K\in\mathfrak{X} be a non-empty, bounded, closed, weakly uniformly convex subset of a CAT_{p}(0) metric space since p\geq2 , \lambda_{1}, \lambda_{2}\in[0, 1] and we specify \mathfrak{T}:\mathfrak{X} \to \mathfrak{X} as a CMJ-type mapping. For all c, s \in \mathfrak{X} and \epsilon > 0 , let us assume

    0 < \max\Big\{d^{p}(c,s), d^{p}(s,\mathfrak{T}s)\Big\} < \epsilon+\mu < \Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon.

    Then \mathfrak{T} has a unique fixed point \zeta \in \mathfrak{X} and \lim\limits_{i\to\infty}\mathfrak{T}^{i}c = \zeta for each c \in \mathfrak{X} .

    Proof. The similar evidence for Theorem 3.1 is applied in the proofs of the remaining stages. Therefore, the proofs have been concluded.

    To support the established results mentioned above, an example is illustrated below.

    Example 3.1. Let \mathfrak{X} = [0, 1] , K = [1, \frac{1}{3}, \dots, \frac{1}{2n-1}]\in {X} for n\in \mathbb{N} , and (\mathfrak{X}, d) be a complete CAT_{p}(0) metric space, p\geq 2 . Define a metric by d^{p}(c, s) = \|c-s\|^{p}, and a mapping \mathfrak{T}: \mathfrak{X} \to \mathfrak{X} given by

    \mathfrak{T}c = \left\{ \begin{array}{ll} \frac{1}{2c}, & \ for \ n \in \mathbb{N}\\ \nonumber 0, & \ otherwise. \\ \end{array} \right.

    We show that \mathfrak{T} satisfies inequality (3.1).

    First we calculate the following metrics for c = \frac{1}{2n-1}, s = \frac{1}{2n+1} .

    \begin{eqnarray} d^{p}(c,s) = d^{p}(\frac{1}{2n-1},\frac{1}{2n+1})& = &\lambda_{1}\Big|\Big|\frac{2}{(2n-1)(2n+1)}\Big|\Big|^{p}, \\ d^{p}(\mathfrak{T}c,s) = d^{p}\Big(\frac{1}{2c},s\Big) = \lambda_{1}\Big|\Big|\frac{(2n-1)}{2}-\frac{1}{2n+1}\Big|\Big|^{p}& = &\lambda_{1}\Big|\Big|\frac{4n^{2}}{(2n+1)}\Big|\Big|^{p}, \\ d^{p}(s,\mathfrak{T}s) = d^{p}\Big(s,\frac{1}{2s}\Big) = \lambda_{1}\Big|\Big|\frac{1}{2n+1}-\frac{2n+1}{2}\Big|\Big|^{p}& = &\lambda_{1}\Big|\Big|\frac{1-(4n^{2}+4n)}{2(2n+1)}\Big|\Big|^{p}, \\ d^{p}(c,\mathfrak{T}s) = d^{p}\Big(c,\frac{1}{2s}\Big) = \lambda_{1}\Big|\Big|\frac{1}{2n-1}-\frac{2n+1}{2}\Big|\Big|^{p}& = &\lambda_{1}\Big|\Big|\frac{3-4n^{2}}{2(2n-1)}\Big|\Big|^{p}, \\ d^{p}(\mathfrak{T}c,c) = d^{p}\Big(\frac{1}{2c},\frac{1}{2n-1}\Big) = \lambda_{1}\Big|\Big|\frac{(2n-1)}{2}-\frac{1}{2n-1}\Big|\Big|^{p}& = &\lambda_{1}\Big|\Big|\frac{(2n-1)^{2}-2}{(2(2n-1))}\Big|\Big|^{p}, \\ d^{p}(\mathfrak{T}c,\mathfrak{T}s) = d^{p}\Big(\frac{1}{2c},\frac{1}{2s}\Big) = \lambda_{1}\Big|\Big|\frac{(2n-1)}{2}-\frac{2n+1}{2}\Big|\Big|^{p}& = &\lambda_{1}\Big|\Big|\frac{2n-1-2n-1}{2}\Big|\Big|^{p}. \end{eqnarray}

    By applying all of the above equalities in (3.1), for n = 2 \dots , \lambda_{1} = \frac{1}{2} and p\geq 2 , we get

    \begin{eqnarray} 0& < & \max\Big\{\lambda_{1}\Big|\Big|\frac{4n^{2}}{(2n+1)}\Big|\Big|^{p},\lambda_{1}\Big|\Big|\frac{1-(4n^{2}+4n)}{2(2n+1)}\Big|\Big|^{p},\lambda_{1}\Big|\Big|\frac{2}{(2n-1)(2n+1)}\Big|\Big|^{p},\\&&\frac{\lambda_{1}\Big|\Big|\frac{4n^{2}}{(2n+1)}\Big|\Big|^{p}+\lambda_{1}\Big|\Big|\frac{3-4n^{2}}{2(2n-1)}\Big|\Big|^{p}}{2}\Big\}\leq\epsilon\\&& \Rightarrow \lambda_{1}\Big|\Big|\frac{2n-1}{2}-\frac{2n+1}{2}\Big|\Big|^{p} < \epsilon. \\ 0& < & \max\Big\{5.12,2.645,0.008,\frac{5.12+2.3464}{2}\Big\}\leq\epsilon \Rightarrow 1 < \epsilon. \\ 0& < & 5.12\leq\epsilon \Rightarrow 0.5 < \epsilon. \end{eqnarray}

    Furthermore, since \mathfrak{T} is continuous, it follows that

    \begin{eqnarray} \left. \begin{array}{l} 0 < \max\Big\{0.6805,3.733\Big\}\leq \epsilon\nonumber \\ 0 < \max\Big\{0.0088,2.645 \Big\} < \mu\nonumber \end{array} \right\} \Rightarrow 0.5 < \epsilon-\mu. \end{eqnarray}

    Consequently

    \begin{eqnarray} \left. \begin{array}{l} 0 < 3.733\leq \epsilon\nonumber \\ 0 < 2.645, < \mu\nonumber \end{array} \right\} \Rightarrow \ 0.5 < \epsilon-\mu. \end{eqnarray}

    This shows that the conditions given in Theorem 3.1 are satisfied.

    Similarly, assume that \mathfrak{T} does not satisfy the CN_{p} inequality:

    \begin{eqnarray} d^{p}(\lambda_{2}c\oplus\lambda_{1}s,z)&\leq&\lambda_{2}d^{p}(c,z)+\lambda_{1} d^{p}(s,z)-\lambda_{1}\lambda_{2}d^{p}(c,s), \end{eqnarray} (3.12)

    where

    \begin{eqnarray} d^{p}(\lambda_{2}c\oplus\lambda_{1}s,z)&\leq& \lambda_{2}d^{p}(c,z)+ \lambda_{1} d^{p}(s,z). \end{eqnarray} (3.13)

    Using (3.13) in (3.12), we get

    \begin{eqnarray} \lambda_{2}d^{p}(\kappa,z)+ \lambda_{1} d^{p}(s,z)&\leq&\lambda_{2}d^{p}(c,z)+\lambda_{1} d^{p}(s,z)-\lambda_{1}\lambda_{2}d^{p}(c,s), \\ 0&\leq&-\lambda_{1}\lambda_{2}d^{p}(c,s), \\ 0&\leq&\lambda_{1}\lambda_{2}\|c-s\|^{p}, \\ 0&\leq& \lambda_{1}\lambda_{2}\Big|\Big|\frac{2}{(2n-1)(2n+1)}\Big|\Big|^{p}, \\ 0&\leq&\frac{1}{2}.\frac{1}{2}\Big|\Big|\frac{2}{15}\Big|\Big|^{p}, \\ 0&\leq&\frac{1}{4}\Big|\Big|\frac{2}{15}\Big|\Big|^{2}, \\ 0&\leq&\frac{1}{225}. \end{eqnarray}

    Thus, \mathfrak{T} satisfies the CN_p inequality and possesses a unique fixed point at c = \frac{\sqrt{2}}{2}, which leads to a contradiction.

    For CAT(0) space, we show that our results are a real extension by showing that the example given above does not satisfy the CN inequality, which is an example of a CAT(0) space. Let c = 1, s_{1} = \frac{1}{3}, s_{2} = \frac{1}{5} be the three points in \mathfrak{X} and z = \frac{s_{1} \oplus s_{2}}{2} = \frac{8}{15} :

    \begin{eqnarray} d(c,z)^{2}&\leq& \frac{1}{2}d(c,s_{1})^{2}+\frac{1}{2}d(c,s_{2})^{2}-\frac{1}{4}d(s_{1},s_{2})^{2}, \\ 0.54&\leq&0.222+0.32-0.0088, \\ 0.54&\leq&0.53, \end{eqnarray}

    which is a contradiction when you compare with a CAT(0)_{P} space.

    Theorem 3.1 extends the scope of fixed point theory by considering CMJ-type mappings in CAT_p(0) metric space with p \geq 2 and weak uniform convexity. We now compare this result with existing literature across various settings:

    (1) Banach spaces

    In normed linear spaces, especially Banach spaces, the foundational result is the Banach contraction principle [36], which requires:

    d(\mathfrak{T}c, \mathfrak{T}s) \leq \lambda d(c, s), \quad \text{for } \lambda \in [0,1).

    By taking c = 1, s = \frac{1}{3} with \lambda = \frac{1}{2} , we have

    0.5 \leq 0.066,

    which is a contradiction.

    The Mankowski-type conditions in Theorem 2.7 serve as a generalization of the Banach contraction principle, particularly in cases where Example 3.1 fails to satisfy the Banach conditions. This highlights that the results established in the theorem are indeed a genuine extension.

    Theorem 3.1, on the other hand:

    ● Uses d^p instead of d , extending the metric to include nonlinear or weighted behaviors.

    ● Requires only weak uniform convexity, thus covering a broader class of metric spaces.

    ● Applies to CAT_p(0) spaces, which generalize Hilbert spaces to nonlinear, non-positive curvature settings.

    (2) CAT(0) and CAT( \kappa ) spaces

    Previous works by Kirk [9], Gromov [8], and others studied fixed points of nonexpansive mappings in CAT(0) spaces. These studies typically rely on:

    ● Convexity of the metric (midpoint inequality).

    ● Nonexpansive or asymptotically nonexpansive conditions.

    Theorem 3.1 strengthens such results by:

    ● Introducing CMJ-type conditions with d^p terms.

    ● Applying to CAT_p(0) spaces, which accommodate generalized distances and stronger nonlinear structures.

    ● Allowing weakly uniformly convex subsets instead of requiring global convexity.

    We present a comparison of fixed point results in various metric and geometric settings in the following table (see Table 1).

    Table 1.  Comparison of fixed point results in various metric and geometric settings.
    Reference / Approach Space Considered Type of Mapping Conditions Imposed Limitation / Novelty
    Banach (1922) [36] Metric space Contraction mapping Constant Lipschitz condition k < 1 Limited to linear contraction in complete metric spaces
    Meir-Keeler (1969) [1] Metric space Generalized contraction Convergence-type condition on distance functions Fails in geometric or non-Euclidean spaces
    CMJ Mappings Metric / b-metric space Cyclic or multi-valued mappings Use of cyclical and intermediate contractive conditions Not always extendable to non-linear geometric spaces
    Results in CAT(0) Spaces Kirk [9] \mathrm{CAT}(0) (non-positively curved) spaces Nonexpansive mappings Convexity + geodesic properties Limited to single-valued and nonexpansive cases
    Results in \mathrm{CAT}(0)_{P} Spaces Khamsi and Shukri [13]. Generalized geodesic spaces with projection properties Generalized mappings under convexity constraints Geometric assumptions can be too restrictive for hybrid mappings
    Present Work (This Paper) \mathrm{CAT}(0)_{P} spaces / generalized metric spaces CMJ-type + Meir–Keeler + hybrid/self/non-self mappings Weak cyclic contraction, convexity, and non-linear projection structure Unifies multiple mapping types; applicable in abstract geometric settings with greater flexibility

     | Show Table
    DownLoad: CSV

    In this section, we explore the existence of solutions to the non-linear matrix equation in the context of the CAT_{p}(0) metric space, which is intended to demonstrate Theorem 3.1. Consider the following a non-linear matrix equation:

    \begin{eqnarray} \mathfrak{X}+\sum\limits_{1\leq i\leq d}^{}A_{i}\mathfrak{X}^{-1}D_{i}& = &C. \end{eqnarray} (5.1)

    The solution pertaining to a tree-like stochastic process, which acts as a generalization of the quasi-birth-and-death (QBD) process, is fundamental to this equation. The expression for S is regarded as minimal, specifically S = \mathfrak{T}-I , where \mathfrak{T} is identified as a sub-stochastic matrix, as demonstrated by Bini et al. [37].

    Lemma 5.1. [37] The matrix \mathfrak{T} = S+I is the minimum non-negative solution of the equation

    \begin{eqnarray} \mathfrak{X}& = &B+\sum\limits_{1\leq i\leq d}^{}A_{i}(A-I)^{-1}D_{i}, \end{eqnarray} (5.2)

    where

    (i) C = B-I , B is sub-stochastic,

    (ii) A_{i} and D_{i} have non-negative entries,

    (iii) the matrices I+C+D_{i}+A_{i}+\dots +A_{d}, i = 1 \dots d, are stochastic.

    The matrices G_{i} are defined as G_{i} = (-S)^{-1}D_{i} , where (G_{i})_{k, k'} denotes the probability of the tree-like process starting at state (i, k) in N_{i} and eventually reaching the root node, with k' being the first state visited in this process. If the process exhibits positive recruitment, then G_{i} is stochastic for every i . Furthermore, Eq 5.1 can be expressed as a system of coupled equations.

    \begin{eqnarray} S& = &C+\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i}, \end{eqnarray} (5.3)

    where G_{i} = (-S)^{-1}D_{i}, for 1\leq i\leq d , from which S is obtained by fixed point iteration. Then (\mathfrak{X}, d) is a complete CAT_{p}(0) metric space. We hereby introduce the subsequent theorem.

    Theorem 5.1. Suppose the following hypotheses hold:

    (ⅰ) The sequences \{S_{n}:n\geq 0\} and \{G_{i, n}:n\geq 0\} , 1\leq i\leq d , defined by S_{n} = C+\sum_{1\leq i\leq d} A_{i}G_{i, n}, where G_{i, n+1} = (-S_{n})^{-1}D_{i}\; \forall\; 1\leq i\leq d, n\geq 0 with G_{i, 0} = \dots = G_{d, 0} = 0, monotonically converge to S and G_{i} , respectively. S is a non-singular matrix.

    (ⅱ) There exists a matrix G such that

    \begin{equation} \|\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n}-\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n+1}\| \leq \|c-s\|^{p},\nonumber \end{equation}

    where

    \begin{equation} \|c-s\|^{p} = \max\Big\{d^{p}(\mathfrak{T}c,s),d^{p}(s,\mathfrak{T}s),d^{p}(c,s),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\}\leq \epsilon \Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon.\nonumber \end{equation}

    Then, the stochastic matrix equation (5.1) has a minimum unique at c .

    Proof. It is known that the stability of the positive steady state of the stochastic matrix equation can be ascertained from conditions (i) and (ii) . We now demonstrate that a minimal unique positive point can be reached by the stochastic matrix equation (5.1), and that point is c .

    Define a mapping \mathfrak{T}:C^{1}([0, T], \mathfrak{X}) \to C^{1}([0, T], \mathfrak{X}) by \mathfrak{T}S_{n} = C+\sum_{1\leq i\leq d}^{}A_{i}G_{i, n} . By the continuity property, for 0\leq t\leq T , we have \mathfrak{T} c, \mathfrak{T} s \in \mathfrak{X} and \mathfrak{T}\ c\neq \mathfrak{T} s . Let c, s \in C^{1}([0, T], \mathfrak{X}) for s\leq c and c = S_{n}, s = S_{n+1} , and we get

    \begin{eqnarray} \label{eq 4.3} \| \mathfrak{T}S_{n} - \mathfrak{T}S_{n+1}\|^{p}& = &\Big|\Big|C+\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n}-\Bigg(C+\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n+1}\Bigg)\Big|\Big|^{p}, \\ \| \mathfrak{T}S_{n} - \mathfrak{T}S_{n+1}\|^{p}&\leq&\Big|\Big|\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n}-\sum\limits_{1\leq i\leq d}^{}A_{i}G_{i,n+1}\Big|\Big|^{p}, \\ \| \mathfrak{T}S_{n} - \mathfrak{T}S_{n+1}\|^{p}&\leq&\Big|\Big|S_{n}-S_{n+1}\Big|\Big|^{p}, \\ \| \mathfrak{T}c - \mathfrak{T}s\|^{p}&\leq&\|c-s\|^{p}, \end{eqnarray}

    which implies

    \begin{eqnarray} d^{p}(\mathfrak{T}c, \mathfrak{T}s) &\leq& d^{p}(c,s) , \\ d^{p}(\mathfrak{T}c, \mathfrak{T}s) &\leq& \max\Big\{d^{p}(\mathfrak{T}c,s),d^{p}(s,\mathfrak{T}s),d^{p}(c,s),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\},\\&\leq&\epsilon \Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon , \end{eqnarray}

    which is a contradiction. As a result, there is a minimal unique positive solution for the stochastic matrix Equation (5.1) at c .

    In this section, we delve into the results associated with two-scale fractal fractional hybrid fractional differential equations (HFDEs) in CAT_{p}(0) metric space, which leads to Theorem 3.1. One can see [38] and the references therein for several hybrid differential equations. In addition to the commonly used Caputo and Riemann-Liouville definitions, several alternative formulations of fractional derivatives have been introduced to address specific modeling needs. For instance, the He [39] fractional derivative, based on an invariant transformation method, is particularly useful in solving nonlinear problems analytically. Similarly, the Atangana-Baleanu [40] fractional derivative, defined with a non-singular and non-local kernel based on the Mittag-Leffler function, provides advantages in modeling processes with fading memory. These definitions, alongside the Caputo-Fabrizio and other generalized operators, offer diverse tools for fractional-order modeling in various disciplines, including physics, engineering, and biology.

    Atangana and Baleanu [40] gave the following definition on fractional derivatives:

    Definition 6.1. [40] The Atangana-Baleanu fractional differential equation is defined as follows with the left and right in Caputo sense:

    \begin{eqnarray} ^{ABC}_aD^{\sigma}f(t)& = &\frac{M(\sigma)}{1-\sigma} \int^{t}_{0}f'(s)E_{\sigma}\Big(\frac{-\sigma}{1-\sigma}(t-s)^{\sigma}\Big)ds, t \in I = [0,T], \\ ^{ABC}_bD^{\sigma}f(t)& = &-\frac{M(\sigma)}{1-\sigma} \int^{b}_{t}f'(s)E_{\sigma}\Big(\frac{-\sigma}{1-\sigma}(t-s)^{\sigma}\Big)ds, \ t \in I = [0,T]. \end{eqnarray}

    Definition 6.2. [40] The Atangana-Baleanu fractional differential equation is defined as follows, with the left and right in Riemann-Lioville sense:

    \begin{eqnarray} ^{ABR}_aD^{\sigma}f(t)& = &\frac{M(\sigma)}{1-\sigma}\frac{d}{dt} \int^{t}_{0}f'(s)E_{\sigma}\Big(\frac{-\sigma}{1-\sigma}(t-s)^{\sigma}\Big)ds, \\ ^{ABR}_bD^{\sigma}f(t)& = &-\frac{M(\sigma)}{1-\sigma}\frac{d}{dt} \int^{b}_{t}f'(s)E_{\sigma}\Big(\frac{-\sigma}{1-\sigma}(t-s)^{\sigma}\Big)ds, \end{eqnarray}

    where f\in H'(a, b) , a < b , \sigma \in [0, 1], and M(\sigma) is a normalization function with M(0) = M(1) = 1.

    According to He [39,41], the fractional derivative is defined in the following form.

    Definition 6.3. [39]

    \begin{eqnarray} D_{t}^{\sigma}f(t)& = &\frac{1}{\Gamma(n-\sigma)}\frac{d^{n}}{dt^{n}} \int^{t}_{t_{0}}(s-t)^{n-\sigma-1}\Big[\sum\limits_{i = 0}^{n-1}\frac{1}{i!}(s-t)^{i}f^{i}(t_{0})-f(s)\Big]ds, \end{eqnarray}

    for the continuous and differentiable case.

    \begin{eqnarray} D_{t}^{\sigma}f(t)& = &\frac{1}{\Gamma(n-\sigma)}\frac{d^{n}}{dt^{n}} \int^{t}_{t_{0}}(s-t)^{n-\sigma-1}\Big[\sum\limits_{i = 0}^{n-1}\frac{(s-t_{0})^{i}}{\Gamma(1+i\sigma)}f^{i\sigma}(t_{0})-f(s)\Big]ds, \end{eqnarray}

    for continuous and not differentiable case.

    Shafiullah and his team [42] have recently addressed this topic, employing fixed point analysis alongside numerical approaches to derive significant theoretical and computational insights regarding a class of fractal HFDEs that utilize a power law kernel. The following challenge discussed by Shafiullah et al. [42] is presented as follows:

    \begin{eqnarray} \left\{ \begin{array}{ll} ^{FABC}_{0}D^{\sigma,\varepsilon}_{\vartheta}[\Psi(\vartheta)-K(\vartheta,c{\theta})] = F(\vartheta,\Psi(\lambda\theta)), \ \vartheta \in I = (0,T), \\\ \Psi(0) = \Psi_{0}+H(\vartheta,\Psi(\vartheta)), \end{array} \right. \end{eqnarray} (6.1)

    where ^{FABC}_{0}D^{\sigma, \varepsilon}_{\theta} denotes the fractal fractional Atangana-Baleanu-Caputo (ABC) derivative fractal order \lambda \in (0, 1) , K, F\in C[I\times \mathbb{R}, \mathbb{R}] is a given function, and \psi \in C and \lambda are the proportional delay terms.

    Differential equations featuring proportional delay terms constitute the pantograph equations. These equations are applicable in both pure and applied mathematics, encompassing fields such as control systems, probability, and electrodynamics. In the case where \lambda = 1 , Equation 6.1 yields the standard HFDE fractals. However, when \lambda > 1 , Equation 6.1 is classified as ill-posed. The fractal dimension is denoted by the symbol \varepsilon , and a value of \varepsilon = 1 indicates the consideration of the commonly employed fractional differential operator.

    Some examples of fractional derivatives are as follows: The symbol for Riemann Liouville is RL. CP -is the derivative of Caputo. CFD -is the derivative of Caputo Fabrizio. The exponential kernel in the CFD was replaced with a Mittag-Leffler by Atagana and Baleanu [40], and the resulting definition was dubbed ABC. It has been applied to the Klein-Gordan equation using the HIV-1 model and the Atangana-Baleanu-Caputo (ABC) differential operator in [42,45].

    The following definitions and lemma are taken from [39,40,43,46].

    Definition 6.4. The following is the definition of the RL fractional integral of order 0 < \sigma\leq 1 :

    \begin{eqnarray} ^{RL}I^{\sigma}\Psi(\vartheta) = \int^{\theta}_{0}(\vartheta-r)^{\sigma-1}\Psi(r)\frac{1}{\Gamma(\sigma)}dr, \ \vartheta \in I = [0,T]. \end{eqnarray} (6.2)

    Definition 6.5. The following is the definition of an ABC -type arbitrary-order integral with order \sigma > 0 :

    \begin{eqnarray} ^{ABC}I^{\sigma}_{0}c(\vartheta) = \frac{1-\sigma}{N(\sigma)} \Psi(\vartheta) + \frac{\sigma}{\Gamma(\sigma)N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}\Psi(r)d\tau \end{eqnarray} (6.3)

    such that the right side exists.

    Additionally, the definition of Mittag-Leffler is as follows:

    Definition 6.6. The definition of the arbitrary-order derivative of ABC -type with order 0 < \sigma \leq 1 is as follows:

    \begin{eqnarray} ^{ABC}D^{\sigma}_{0}v(\vartheta) = \frac{N(\sigma)}{1-\sigma } \Psi(\vartheta) \int^{\vartheta}_{0}E_{\sigma}[-\frac{\sigma}{1-\sigma}(\vartheta-r)^{\sigma-1}]\Psi'(r)d\tau. \end{eqnarray} (6.4)

    In the definition, E_{\sigma} is the Mittag-Leffler function, and the function N(\sigma) , which is called normalization, obeys N(0) = N(1) = 1 .

    Moreover, the definitions of the fractal fractional ABC integral and derivatives are as follows:

    Definition 6.7. The following is the definition of the fractal fractional ABC integral with fractional order \sigma \in (0, 1) and fractals \varepsilon \in (0, 1) :

    \begin{eqnarray} ^{FABC}I^{\sigma,\varepsilon}_{0^{+}}v(\vartheta) = \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} \Psi(\vartheta) +\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}\Psi(r)dr. \end{eqnarray} (6.5)

    Definition 6.8. The following is the definition of the fractal fractional ABC derivative with fractional order \sigma \in (0, 1) and fractals \varepsilon \in (0, 1) :

    \begin{eqnarray} ^{FABC}D^{\sigma,\varepsilon}_{0}\Psi(\vartheta) = \frac{N(\sigma)}{1-\sigma } \Psi(\vartheta) \int^{\vartheta}_{0}E_{\sigma}[-\frac{\sigma}{1-\sigma}(\vartheta-r)^{\sigma-1}]\frac{du(r)}{dr^{\varepsilon}}d(r). \end{eqnarray} (6.6)

    Lemma 6.1. If c\in L[0, T] and c(0) = 0 , the solution of

    \begin{eqnarray} \label{equation 5.6} ^{FABC}D^{\sigma,\varepsilon}_{0}\Psi(\vartheta) = c(\vartheta), with \ \sigma \in (0,1], \\ \Psi(0) = \Psi_{0}, \end{eqnarray}

    is described as follows:

    \begin{eqnarray} \Psi(\vartheta) = \Psi_{0}+ \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} c(\vartheta) +\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}c(r)dr. \end{eqnarray} (6.7)

    The continuous function defined on I has a space I = \mathfrak{X} . Assume that all continuous functions from I into \mathfrak{X} have a norm of \Vert c \Vert : = \sup{|c(t)|} , \vartheta \in I for c \in C(I, \mathfrak{X}) . Let (\mathfrak{X}, | · |) be a Banach space.

    This space defines the path metric by

    \begin{equation} d^{p}(c,s) = \sup\limits_{\vartheta\in I}\{|c(\vartheta)-s(\vartheta)|^{p}\} \end{equation} (6.8)

    where \forall c, s \in \mathfrak{X} . This is a complete CAT_{p}(0) metric space.

    It is possible to define problem (6.1) as a fixed point problem by using Lemma 6.1.

    \begin{eqnarray} \Psi(\vartheta)& = &\Psi_{0}+ H(\vartheta,\Psi(\vartheta))+ K(\vartheta,\Psi(\vartheta))+ \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} c(\vartheta) +\\&&\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}c(r)dr, \end{eqnarray} (6.9)

    where

    c(\vartheta) = F(\vartheta,\Psi(\lambda \vartheta))

    and

    c(r) = F(r,\Psi(\lambda r)).

    Referencing [44], the following image in Figure 2 depicts the fractal fractional behavior of Eq (6.1), which is identical to Eq (6.9).

    Figure 2.  FABC- CAT_{p}(0) metric spaces.

    From the above information, we prove the following theorem.

    Theorem 6.1. For this investigation, the following presumptions are required.

    (ⅰ) There exists a constant \zeta > 0 , such that

    \begin{eqnarray} \vert H(\vartheta,\Psi(\vartheta))- H(\vartheta,\bar{\Psi}(\vartheta))\vert &\leq& \zeta\vert \Psi(\vartheta)-\bar{\Psi}(\vartheta)\vert = \zeta \vert c(\vartheta)-s(\vartheta)\vert, \end{eqnarray}

    (ⅱ) a constant \eta > 0 , such that

    \begin{eqnarray} \vert K(\vartheta,\Psi(\vartheta))- K(\vartheta,\bar{\Psi}(\vartheta))\vert &\leq& \eta\vert \Psi(\vartheta)-\bar{\Psi}(\vartheta)\vert = \eta \vert c(\vartheta)-s(\vartheta)\vert, \end{eqnarray}

    (ⅲ) a constant \gamma > 0 given by

    \begin{eqnarray} \vert F(\vartheta,\Psi(\vartheta))- F(\lambda\vartheta,\bar{\Psi}(\lambda\vartheta))\vert &\leq& \gamma\vert \Psi(\vartheta)-\bar{\Psi}(\vartheta)\vert = \gamma\vert c(\vartheta)-s(\vartheta)\vert, \end{eqnarray}

    where \gamma = \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} ,

    (ⅳ) a constant \rho > 0 is given by

    \begin{eqnarray} \vert F(\vartheta,\Psi(\lambda r))- F(\vartheta,\bar{\Psi}(\lambda r))\vert &\leq& \rho\vert \Psi(\vartheta)-\bar{\Psi}(\vartheta)\vert = \rho\vert c(\vartheta)-s(\vartheta)\vert, \end{eqnarray}

    imply that \rho = \frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)}\vartheta^{\sigma + \varepsilon - 1} B(\sigma, \varepsilon) , and

    (ⅳ) k' = \zeta+\eta+\gamma+\rho < \epsilon < 1 , where

    \begin{eqnarray} k' < \max\Big\{d^{p}(\mathfrak{T}c,s),d^{p}(s,\mathfrak{T}s),d^{p}(c,s),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\}\leq\epsilon\Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon. \end{eqnarray}

    Then Eq (6.9) has a unique solution which is also a solution to problem (6.1).

    Proof. We define an operator \mathfrak{T}:\mathfrak{X} \to \mathfrak{X} by

    \begin{eqnarray} \mathfrak{T}\Psi(\vartheta)& = &\Psi_{0}+ H(\vartheta,\Psi(\vartheta))+ K(\vartheta,\Psi(\vartheta))+ \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} c(\vartheta) +\\&&\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}c(r)dr. \end{eqnarray}

    The following uses the aforementioned conditions (i)-(v) to demonstrate the existence of a fixed point:

    \begin{eqnarray} \vert\vert \mathfrak{T}\Psi(\vartheta)- \mathfrak{T}\bar{\Psi}(\vartheta)\vert\vert^{p}& = &\Big|\Big|\Bigg(\Psi_{0}+ H(\vartheta,\Psi(\vartheta))+ K(\vartheta,\Psi(\vartheta))+ \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} c(\vartheta)\\&& +\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}c(r)dr\Bigg)-\\&&\Bigg(\Psi_{0}+ H(\vartheta,\bar{\Psi}(\vartheta))+ K(\vartheta,\bar{\Psi}(\vartheta))+ \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} s(\vartheta)\\&& +\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}s(r)dr\Bigg)\Big|\Big|^{p},\\ &\leq& \vert \vert H(\vartheta,\Psi(\vartheta))-H(\vartheta,\bar{\Psi}(\vartheta))\vert\vert^{p}+ \\&& \vert \vert K(\vartheta,\Psi(\vartheta))-K(\vartheta,\bar{\Psi}(\vartheta)) \vert \vert^{p}+\\&& \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} \vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p} +\\&&\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)} \int^{\vartheta}_{0}(\vartheta-r)^{\sigma-1}r^{\varepsilon-1}\vert \vert c(r)-s(r) \vert \vert^{p} dr, \end{eqnarray}
    \begin{eqnarray} &\leq& \zeta\vert \vert \Psi(\vartheta)-\bar{\Psi}(\vartheta)\vert\vert^{p}+ \eta\vert \vert \Psi(\vartheta)-\bar{\Psi}(\vartheta) \vert \vert^{p}+\\&& \frac{\varepsilon(1-\sigma)\vartheta^{\varepsilon-1}}{N(\sigma)} \vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p} +\\&&\frac{\sigma \varepsilon}{\Gamma(\sigma) N(\sigma)}\vartheta^{\sigma + \varepsilon - 1} B(\sigma, \varepsilon)\vert \vert c(r)-s(r) \vert \vert^{p},\\ &\leq& \zeta\vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p}+ \eta\vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p}+\\&& \gamma \vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p} +\rho\vert \vert c(r)-s(r) \vert \vert^{p}, \\ \vert\vert \mathfrak{T}\Psi(\vartheta)- \mathfrak{T}\bar{\Psi}(\vartheta)\vert\vert^{p}&\leq& (\zeta+ \eta+ \gamma +\rho)\vert \vert c(\vartheta)-s(\vartheta)\vert \vert^{p}, \\ d^{p}(\mathfrak{T}c(\vartheta), \mathfrak{T}s(\vartheta))&\leq& (\zeta+ \eta+ \gamma +\rho) d^{p}(c(\vartheta),s(\vartheta)), \\ d^{p}(\mathfrak{T}c(\vartheta), \mathfrak{T}s(\vartheta))&\leq& k' d^{p}(c(\vartheta),s(\vartheta)). \end{eqnarray}

    This suggests that

    \begin{eqnarray} k'd^{p}(c(\vartheta),s(\vartheta)) < \max\Big\{d^{p}(\mathfrak{T}c,s),d^{p}(s,\mathfrak{T}s),d^{p}(c,s),\frac{d^{p}(\mathfrak{T}c,s)+d^{p}(c,\mathfrak{T}s)}{2}\Big\}\leq\epsilon \Rightarrow d^{p}(\mathfrak{T}c,\mathfrak{T}s) < \epsilon. \end{eqnarray}

    Consequently, the fractal fractional hybrid differential Eq (6.1) has a unique solution, which is also the solution to Eq (6.9). Thus, our evidence is finished.

    This model captures the fractal time evolution of MHD flow velocity, reflecting the memory and multiscale effects inherent in porous or heterogeneous conducting media. Such models have relevance in:

    ● Geophysical flows (e.g., lava or saline water),

    ● Plasma confinement and

    ● Blood flow in biological tissues with magnetic therapy.

    The two-scale derivative introduces control over scale effects (\alpha(s)) and temporal memory (\beta(s)), while VIM ensures analytical tractability and rapid convergence.

    The current study demonstrates that the fractal derivative model can be used to describe the anomalous diffusion process as captured by diffusion-weighted MRI in a fixed mouse brain. In the model, the fractal dimension of the diffusion trajectory is directly expressed in terms of the Hausdorff fractal dimension, and the spectral entropy is used to measure the uncertainty in the heterogeneous and multi-scale system of biological tissues. Interestingly, the fractal derivative order influences the diffusion behavior significantly, especially in systems with anomalous transport properties.

    In this manuscript, we establish and validate fundamental fixed point theorems for CMJ-type mappings, highlighting their broad applicability and improvements over existing results in the literature. Our investigation addresses the existence of solutions to a nonlinear matrix equation within the framework of a CAT_p(0) metric space, thereby reinforcing the claims presented in Theorem 3.1.

    Additionally, we examine the behavior of an iterative matrix process governed by a recursive scheme, which guarantees convergence. This approach is particularly relevant in the context of solving stochastic matrix equations that frequently arise in image reconstruction problems, such as those encountered in tomography and signal processing—most notably in X-ray computed tomography (CT). In these applications, the goal is to reconstruct internal images, such as anatomical cross-sections, from projection data—often requiring the resolution of large systems of equations affected by noise and uncertainty.

    Our study further extends to hybrid fractional differential equations (HFDEs) within the CAT_p(0) setting, leading to the formulation and conclusion of Theorem 3.1. We demonstrate that the fractal derivative framework in CAT_p(0) spaces effectively models anomalous diffusion phenomena, as observed in diffusion-weighted MRI scans of fixed mouse brains. In this model, the fractal nature of the diffusion pathway is directly quantified via the Hausdorff fractal dimension, while spectral entropy serves as a metric for the uncertainty inherent in the complex, multi-scale biological tissue environment. Notably, the order of the fractal derivative is intricately linked to the underlying diffusion characteristics.

    A promising direction for future work involves establishing fixed point theorems for extended (p, q)-F-interpolative mappings within CAT(0) spaces. To support the theoretical development, one may construct concrete examples that validate the proposed results. These fixed point formulations can further be applied to models in economic dynamics such as economic growth, market equilibrium, and bifurcation scenarios in energy capital accumulation by incorporating fractional differential equations to demonstrate their practical utility. Additionally, the structure of CAT(0) spaces can be integrated with generalized two-scale fractional derivatives within the framework of the variational iteration method (VIM), particularly for modeling heat transfer in fractal porous media and magnetohydrodynamic (MHD) flow in fractal environments.

    Conceptualization: M. Sajid, L. Wangwe; Methodology: M. Sajid, H. Kalita; Formal analysis: L. Wangwe, H. Kalita; Investigation: M. Sajid, H. Kalita; Writing-original draft: M. Sajid, S. Kumar; Writing-review and editing: L. Wangwe, S. Kumar; Visualization: H. Kalita; Supervision: S. Kumar; Project administration: S. Kumar; Funding acquisition: S. Kumar.

    The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The Researchers would like to thank the Deanship of Graduate Studies and Scientific Research at Qassim University for financial support (QU-APC-2025).

    All authors declare that they have no conflicts of interest in this paper.



    [1] A. Meir, E. Keeler, A theorem on contraction mappings, J. Math. Anal. Appl., 28 (1969), 326–329. https://doi.org/10.1016/0022-247X(69)90031-6 doi: 10.1016/0022-247X(69)90031-6
    [2] L. B. Ćirić, A new fixed-point theorem for contractive mapping, Publ. I. Math., 30 (1981), 25–27.
    [3] J. Matkowski, Fixed point theorems for contractive mappings in metric spaces, \check{C}asopis Pro P\check{e}stování Matematiky, 105 (1980), 341–344. https://doi.org/10.21136/CPM.1980.108246
    [4] D. W. Boyd, J. S. W. Wong, On nonlinear contraction, Proc. Amer. Math. Soc., 20 (1969), 458–464. https://doi.org/10.2307/2035677
    [5] S. Reich, Fixed points of contractive functions, Boll. UMI, 5 (1972), 26–42.
    [6] C. S. Wong, Generalized contraction and fixed point theorems, Proc. Amer. Math. Soc., 42 (1974), 409–417. https://doi.org/10.1090/S0002-9939-1974-0331358-4 doi: 10.1090/S0002-9939-1974-0331358-4
    [7] j. Jachymski, Equivalent conditions and the Meir-Keeler type theorems, J. Math. Anal. Appl., 194 (1995), 293–303. https://doi.org/10.1006/jmaa.1995.1299 doi: 10.1006/jmaa.1995.1299
    [8] M. Gromov, Metric structure for Riemannian and Non-Riemannian spaces, Boston: Birkhauser, 1984. https://doi.org/10.1007/978-0-8176-4583-0
    [9] W. A. Kirk, Fixed point theorems in CAT(0) spaces and R-trees, Fixed Point Theory Appl., 2004 (2004), 738084. https://doi.org/10.1155/S1687182004406081 doi: 10.1155/S1687182004406081
    [10] K. Goebel, S. Reich, Uniform convexity, hyperbolic geometry, and non-expansive mappings, 1984.
    [11] S. Reich, I. Shafrir, Non-expansive iterations in hyperbolic spaces, Nonlinear Anal. Theor., 15 (1990), 537–558. https://doi.org/10.1016/0362-546X(90)90058-O doi: 10.1016/0362-546X(90)90058-O
    [12] S. Dhompongsa, B. Panyanak, On \bigtriangleup-convergence theorems in CAT (0) spaces, Comput. Math. Appl., 56 (2008), 2572–2579. https://doi.org/10.1016/j.camwa.2008.05.036 doi: 10.1016/j.camwa.2008.05.036
    [13] M. A. Khamsi, S. Shukri, Generalized CAT(0) spaces, Bull. Belg. Math. Soc. Simon Stevin, 24 (2017), 417–426. https://doi.org/10.36045/bbms/1506477690
    [14] S. Shukri, On monotone nonexpansive mappings in C AT_{p}(0) spaces, Fixed Point Theory Appl., 8 (2020). https://doi.org/10.1186/s13663-020-00675-z
    [15] A. A. Darweesh, S. Shukri, Fixed points of Suzuki-generalized nonexpansive mappings in CAT_p (0) metric spaces, Arab. J. Math., 13 (2024), 227–236. https://doi.org/10.1007/s40065-024-00455-2 doi: 10.1007/s40065-024-00455-2
    [16] Y. Sun, R. P. Agarwal, Existence of fixed point for nonlinear operator in partially ordered metric space, Adv. Differ. Equ. Contr., 30 (2023), 97–116. https://doi.org/10.17654/0974324323007 doi: 10.17654/0974324323007
    [17] L. Xie, J. Zhou, H. Deng, Y. He, Existence and stability of solution for multi-order nonlinear fractional differential equations, AIMS Math., 7 (2022), 16440–16448. https://doi.org/10.3934/math.2022899 doi: 10.3934/math.2022899
    [18] J. L. Zhou, Y. B. He, S. Q. Zhang, H. Y. Deng, X. Y. Lin, Existence and stability results for nonlinear fractional integrodifferential coupled systems, Bound. Value Probl., 2023 (2023), 10. https://doi.org/10.1186/s13661-023-01698-2 doi: 10.1186/s13661-023-01698-2
    [19] M. A. Imran, Application of fractal fractional derivative of power law kernel ^{FFP}_{0}D^{\alpha, \beta}_{x} to MHD viscous fluid flow between two plates, Chaos Soliton. Fract., 134 (2020), 10969. https://doi.org/10.1016/j.chaos.2020.109691 doi: 10.1016/j.chaos.2020.109691
    [20] C. H. He, H. W. Liu, C. Liu, A fractal-based approach to the mechanical properties of recycled aggregate concretes, Fact. Univ. Ser. Mech., 22 (2024), 329–342. https://doi.org/10.22190/FUME240605035H doi: 10.22190/FUME240605035H
    [21] D. Bini, B. Meini, On the solution of a nonlinear matrix equation arising in queueing problems, SIAM J. Matrix Anal. A., 17 (1996), 906–926. https://doi.org/10.1137/S0895479895284804 doi: 10.1137/S0895479895284804
    [22] Y. Lim, Solving the nonlinear matrix equation X = Q+\sum_{i = 1}^{m} M_{i}X^{\delta_{i}}M_{i}^{*} via a contraction principle, Linear Algebra Appl., 430 (2009), 1380–1383. https://doi.org/10.1016/j.laa.2008.10.034 doi: 10.1016/j.laa.2008.10.034
    [23] W. A. Kirk, A fixed point theorem for mappings which do not increases distances, Amer. Math. Mon., 72 (1965), 1004–1006. https://doi.org/10.2307/2313345 doi: 10.2307/2313345
    [24] B. Nanjaras, B. Panyanaka, W. Phuengrattanab, Fixed point theorems and convergence theorems for Suzuki-generalized non-expansive mappings in C AT (0) spaces, Nonlinear Anal. Hybri., 4 (2010), 25–31. https://doi.org/10.1016/j.nahs.2009.07.003 doi: 10.1016/j.nahs.2009.07.003
    [25] M. R. Bridson, A. Haefliger, Metric spaces of non-positive curvature, Berlin: Springer, 1999. https://doi.org/10.1007/978-3-662-12494-9
    [26] F. Bruhat, J. Tits, Groupes réductifs sur un corps local: I. Données radicielles valuées, In: Publications Mathématiques de l'IHÉS, 41 (1972), 5–251.
    [27] R. P. Agarwal, J. Mohamed, B. Samet, Fixed point theory in metric spaces, Singapore: Springer, 2018. https://doi.org/10.1007/978-981-13-2913-5
    [28] K. Goebel, W. A. Kirk, Iteration processes for non-expansive mappings, Contemp. Math., 21 (1983), 115–123.
    [29] A. Papadopoulos, Metric spaces, convexity and nonpositive curvature, 2004.
    [30] U. Kohlenbach, L. Leustean, Mann iterates of directionally non-expansive mappings in hyperbolic spaces, Abstr. Appl. Anal., 8 (2003), 449–477. https://doi.org/10.1155/S1085337503212021 doi: 10.1155/S1085337503212021
    [31] Z. Opial, Weak convergence of the sequence of successive approximations for non-expansive mappings, Bull. Amer. Math. Soc., 73 (1967), 591–597.
    [32] F. E. Browder, Nonlinear operators and nonlinear equations of evolution in Banach spaces, Bull. Amer. Math. Soc., 73 (1967), 867–874. https://doi.org/10.1090/S0002-9904-1967-11820-2 doi: 10.1090/S0002-9904-1967-11820-2
    [33] A. Naor, L. Silberman Poincare inequalities, embeddings, and wild groups, Compos. Math., 147 (2011), 1546–1572. https://doi.org/10.1112/S0010437X11005343 doi: 10.1112/S0010437X11005343
    [34] T. Suzuki, Fixed point theorems and convergence theorems for some generalized non-expansive mapping, J. Math. Anal. Appl., 340 (2008), 1088–1095.
    [35] T. Suzuki, Generalized metric spaces do not have the compatible topology, Abstr. Appl. Anal., 2014 (2014), 458098. https://doi.org/10.1155/2014/458098 doi: 10.1155/2014/458098
    [36] S. Banach, On operations in abstract sets and their application to integral equations, Fund. Math., 3 (1922), 133–181.
    [37] D. A. Bini, G. Latouche, B. Meini, Solving nonlinear matrix equations arising in tree-like stochastic processes, Linear Algebra Appl., 366 (2003), 39–64. https://doi.org/10.1016/S0024-3795(02)00593-1 doi: 10.1016/S0024-3795(02)00593-1
    [38] M. L. Maheswari, K. S. Keerthana Shri, M. Sajid, Analysis on existence of system of coupled multi fractional nonlinear hybrid differential equations with coupled boundary conditions, AIMS Math., 9 (2024), 13642–13658. https://doi.org/10.3934/math.2024666 doi: 10.3934/math.2024666
    [39] J. H. He, Fractal calculus and its geometrical explanation, Res. Phys., 10 (2018), 272–276. https://doi.org/10.1016/j.rinp.2018.06.011 doi: 10.1016/j.rinp.2018.06.011
    [40] A. Atangana, D. Baleanu, New fractional derivatives with nonlocal and non-singular kernel: theory and application to heat transfer model, 2016. https://doi.org/10.48550/arXiv.1602.03408
    [41] J. H. He, A tutorial review on fractal space time and fractional calculus, Int. J. Theor. Phys., 53 (2014), 3698–3718. https://doi.org/10.1007/s10773-014-2123-8 doi: 10.1007/s10773-014-2123-8
    [42] S. K. Shafiullah, M. Sarwar, T. Abdeljawad, On theoretical and numerical analysis of fractal-fractional non-linear hybrid differential equations, Nonlinear Eng., 13 (2024), 20220372. https://doi.org/10.1515/nleng-2022-0372 doi: 10.1515/nleng-2022-0372
    [43] S. Qureshi, A. Atangana, Fractal-fractional diferentiation for the modeling and mathematical analysis of nonlinear diarrhea transmission dynamics under the use of real data, Chaos Soliton. Fract., 136 (2020), 109812. https://doi.org/10.1016/j.chaos.2020.109812 doi: 10.1016/j.chaos.2020.109812
    [44] T. Abdeljawad, M. Sher, K. Shah, M. Sarwar, I. Amacha, M. Alqudah, et al., Analysis of a class of fractal hybrid fractional differential equation with application to a biological model, Sci. Rep., 14 (2024), 18937. https://doi.org/10.1038/s41598-024-67158-8 doi: 10.1038/s41598-024-67158-8
    [45] C. Xu, Z. Liu, Y. Pang, S. Saifullah, M. Inc, Oscillatory, crossover behavior and chaos analysis of HIV-1 infection model using piece-wise Atangana-Baleanu fractional operator: Real data approach, Chaos Soliton. Fract., 164 (2022), 112662. https://doi.org/10.1016/j.chaos.2022.112662 doi: 10.1016/j.chaos.2022.112662
    [46] G. S. Teodoro, J. T. Machado, E. C. De Oliveira, A review of defnitions of fractional derivatives and other operators, J. Comput. Phys., 388 (2019), 195–208. https://doi.org/10.1016/j.jcp.2019.03.008 doi: 10.1016/j.jcp.2019.03.008
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(160) PDF downloads(42) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog