In this study for examining the fractional Michaelis-Menten enzymatic reaction (FMMER) model, we suggested a computational method by using an operational matrix of Jacobi polynomials (JPs) as its foundation. We obtain an operational matrix for the arbitrary order derivative in the Caputo sense. The fractional differential equations (FDEs) are then reduced to a set of algebraic equations by using attained operational matrix and the collocation method. The approach which utilized in this study is quicker and more effective compared to other schemes. We also compared the suggested method with the Vieta-Lukas collocation technique (VLCM) and we obtain excellent results. A comparison between numerical outcomes is shown by figures and tables. Error analysis of the recommended methods is also presented.
Citation: Devendra Kumar, Hunney Nama, Dumitru Baleanu. Computational analysis of fractional Michaelis-Menten enzymatic reaction model[J]. AIMS Mathematics, 2024, 9(1): 625-641. doi: 10.3934/math.2024033
[1] | Saima Rashid, Abdulaziz Garba Ahmad, Fahd Jarad, Ateq Alsaadi . Nonlinear fractional differential equations and their existence via fixed point theory concerning to Hilfer generalized proportional fractional derivative. AIMS Mathematics, 2023, 8(1): 382-403. doi: 10.3934/math.2023018 |
[2] | Muhammad Tariq, Sotiris K. Ntouyas, Hijaz Ahmad, Asif Ali Shaikh, Bandar Almohsen, Evren Hincal . A comprehensive review of Grüss-type fractional integral inequality. AIMS Mathematics, 2024, 9(1): 2244-2281. doi: 10.3934/math.2024112 |
[3] | Ahmed Alsaedi, Bashir Ahmad, Afrah Assolami, Sotiris K. Ntouyas . On a nonlinear coupled system of differential equations involving Hilfer fractional derivative and Riemann-Liouville mixed operators with nonlocal integro-multi-point boundary conditions. AIMS Mathematics, 2022, 7(7): 12718-12741. doi: 10.3934/math.2022704 |
[4] | Pinghua Yang, Caixia Yang . The new general solution for a class of fractional-order impulsive differential equations involving the Riemann-Liouville type Hadamard fractional derivative. AIMS Mathematics, 2023, 8(5): 11837-11850. doi: 10.3934/math.2023599 |
[5] | Ugyen Samdrup Tshering, Ekkarath Thailert, Sotiris K. Ntouyas . Existence and stability results for a coupled system of Hilfer-Hadamard sequential fractional differential equations with multi-point fractional integral boundary conditions. AIMS Mathematics, 2024, 9(9): 25849-25878. doi: 10.3934/math.20241263 |
[6] | Donny Passary, Sotiris K. Ntouyas, Jessada Tariboon . Hilfer fractional quantum system with Riemann-Liouville fractional derivatives and integrals in boundary conditions. AIMS Mathematics, 2024, 9(1): 218-239. doi: 10.3934/math.2024013 |
[7] | Ravi Agarwal, Snezhana Hristova, Donal O'Regan . Integral presentations of the solution of a boundary value problem for impulsive fractional integro-differential equations with Riemann-Liouville derivatives. AIMS Mathematics, 2022, 7(2): 2973-2988. doi: 10.3934/math.2022164 |
[8] | Bashir Ahmad, Manal Alnahdi, Sotiris K. Ntouyas, Ahmed Alsaedi . On a mixed nonlinear boundary value problem with the right Caputo fractional derivative and multipoint closed boundary conditions. AIMS Mathematics, 2023, 8(5): 11709-11726. doi: 10.3934/math.2023593 |
[9] | Asghar Ahmadkhanlu, Hojjat Afshari, Jehad Alzabut . A new fixed point approach for solutions of a p-Laplacian fractional q-difference boundary value problem with an integral boundary condition. AIMS Mathematics, 2024, 9(9): 23770-23785. doi: 10.3934/math.20241155 |
[10] | Muath Awadalla, Manigandan Murugesan, Subramanian Muthaiah, Bundit Unyong, Ria H Egami . Existence results for a system of sequential differential equations with varying fractional orders via Hilfer-Hadamard sense. AIMS Mathematics, 2024, 9(4): 9926-9950. doi: 10.3934/math.2024486 |
In this study for examining the fractional Michaelis-Menten enzymatic reaction (FMMER) model, we suggested a computational method by using an operational matrix of Jacobi polynomials (JPs) as its foundation. We obtain an operational matrix for the arbitrary order derivative in the Caputo sense. The fractional differential equations (FDEs) are then reduced to a set of algebraic equations by using attained operational matrix and the collocation method. The approach which utilized in this study is quicker and more effective compared to other schemes. We also compared the suggested method with the Vieta-Lukas collocation technique (VLCM) and we obtain excellent results. A comparison between numerical outcomes is shown by figures and tables. Error analysis of the recommended methods is also presented.
Segre [1] made a pioneering attempt in the development of special algebra. He conceptualized the commutative generalization of complex numbers, bicomplex numbers, tricomplex numbers, etc. as elements of an infinite set of algebras. Subsequently, in the 1930s, researchers contributed in this area [2,3,4]. The next fifty years failed to witness any advancement in this field. Later, Price [5] developed the bicomplex algebra and function theory. Recent works in this subject [6,7] find some significant applications in different fields of mathematical sciences as well as other branches of science and technology. An impressive body of work has been developed by a number of researchers. Among these works, an important work on elementary functions of bicomplex numbers has been done by Luna-Elizaarrarˊas et al. [8]. Choi et al. [9] proved some common fixed point theorems in connection with two weakly compatible mappings in bicomplex valued metric spaces. Jebril [10] proved some common fixed point theorems under rational contractions for a pair of mappings in bicomplex valued metric spaces. In 2017, Dhivya and Marudai [11] introduced the concept of a complex partial metric space, suggested a plan to expand the results and proved some common fixed point theorems under a rational expression contraction condition. In 2019, Mani and Mishra [12] proved coupled fixed point theorems on a complex partial metric space using different types of contractive conditions. In 2021, Gunaseelan et al. [13] proved common fixed point theorems on a complex partial metric space. In 2021, Beg et al.[14] proved fixed point theorems on a bicomplex valued metric space. In 2021, Zhaohui et al. [15] proved common fixed theorems on a bicomplex partial metric space. In this paper, we prove coupled fixed point theorems on a bicomplex partial metric space. An example is provided to verify the effectiveness and applicability of our main results. An application of these results to Fredholm integral equations and nonlinear integral equations is given.
Throughout this paper, we denote the set of real, complex and bicomplex numbers, respectively, as C0, C1 and C2. Segre [1] defined the complex number as follows:
z=ϑ1+ϑ2i1, |
where ϑ1,ϑ2∈C0, i21=−1. We denote the set ofcomplex numbers C1 as:
C1={z:z=ϑ1+ϑ2i1,ϑ1,ϑ2∈C0}. |
Let z∈C1; then, |z|=(ϑ21+ϑ22)12. The norm ||.|| of an element in C1 is the positive real valued function ||.||:C1→C+0 defined by
||z||=(ϑ21+ϑ22)12. |
Segre [1] defined the bicomplex number as follows:
ς=ϑ1+ϑ2i1+ϑ3i2+ϑ4i1i2, |
where ϑ1,ϑ2,ϑ3,ϑ4∈C0, and independent units i1,i2 are such that i21=i22=−1 and i1i2=i2i1. We denote the set of bicomplex numbers C2 as:
C2={ς:ς=ϑ1+ϑ2i1+ϑ3i2+ϑ4i1i2,ϑ1,ϑ2,ϑ3,ϑ4∈C0}, |
i.e.,
C2={ς:ς=z1+i2z2,z1,z2∈C1}, |
where z1=ϑ1+ϑ2i1∈C1 and z2=ϑ3+ϑ4i1∈C1. If ς=z1+i2z2 and η=ω1+i2ω2 are any two bicomplex numbers, then the sum is ς±η=(z1+i2z2)±(ω1+i2ω2)=z1±ω1+i2(z2±ω2), and the product is ς.η=(z1+i2z2)(ω1+i2ω2)=(z1ω1−z2ω2)+i2(z1ω2+z2ω1).
There are four idempotent elements in C2: They are 0,1,e1=1+i1i22,e2=1−i1i22 of which e1 and e2 are nontrivial, such that e1+e2=1 and e1e2=0. Every bicomplex number z1+i2z2 can be uniquely expressed as the combination of e1 and e2, namely
ς=z1+i2z2=(z1−i1z2)e1+(z1+i1z2)e2. |
This representation of ς is known as the idempotent representation of a bicomplex number, and the complex coefficients ς1=(z1−i1z2) and ς2=(z1+i1z2) are known as the idempotent components of the bicomplex number ς.
An element ς=z1+i2z2∈C2 is said to be invertible if there exists another element η in C2 such that ςη=1, and η is said to be inverse (multiplicative) of ς. Consequently, ς is said to be the inverse(multiplicative) of η. An element which has an inverse in C2 is said to be a non-singular element of C2, and an element which does not have an inverse in C2 is said to be a singular element of C2.
An element ς=z1+i2z2∈C2 is non-singular if and only if ||z21+z22||≠0 and singular if and only if ||z21+z22||=0. When it exists, the inverse of ς is as follows.
ς−1=η=z1−i2z2z21+z22. |
Zero is the only element in C0 which does not have a multiplicative inverse, and in C1, 0=0+i10 is the only element which does not have a multiplicative inverse. We denote the set of singular elements of C0 and C1 by O0 and O1, respectively. However, there is more than one element in C2 which does not have a multiplicative inverse: for example, e1 and e2. We denote this set by O2, and clearly O0={0}=O1⊂O2.
A bicomplex number ς=ϑ1+ϑ2i1+ϑ3i2+ϑ4i1i2∈C2 is said to be degenerated (or singular) if the matrix
(ϑ1ϑ2ϑ3ϑ4) |
is degenerated (or singular). The norm ||.|| of an element in C2 is the positive real valued function ||.||:C2→C+0 defined by
||ς||=||z1+i2z2||={||z21||+||z22||}12=[|z1−i1z2|2+|z1+i1z2|22]12=(ϑ21+ϑ22+ϑ23+ϑ24)12, |
where ς=ϑ1+ϑ2i1+ϑ3i2+ϑ4i1i2=z1+i2z2∈C2.
The linear space C2 with respect to a defined norm is a normed linear space, and C2 is complete. Therefore, C2 is a Banach space. If ς,η∈C2, then ||ςη||≤√2||ς||||η|| holds instead of ||ςη||≤||ς||||η||, and therefore C2 is not a Banach algebra. For any two bicomplex numbers ς,η∈C2, we can verify the following:
1. ς⪯i2η⟺||ς||≤||η||,
2. ||ς+η||≤||ς||+||η||,
3. ||ϑς||=|ϑ|||ς||, where ϑ is a real number,
4. ||ςη||≤√2||ς||||η||, and the equality holds only when at least one of ς and η is degenerated,
5. ||ς−1||=||ς||−1 if ς is a degenerated bicomplex number with 0≺ς,
6. ||ςη||=||ς||||η||, if η is a degenerated bicomplex number.
The partial order relation ⪯i2 on C2 is defined as follows. Let C2 be the set of bicomplex numbers and ς=z1+i2z2, η=ω1+i2ω2∈C2. Then, ς⪯i2η if and only if z1⪯ω1 and z2⪯ω2, i.e., ς⪯i2η if one of the following conditions is satisfied:
1. z1=ω1, z2=ω2,
2. z1≺ω1, z2=ω2,
3. z1=ω1, z2≺ω2,
4. z1≺ω1, z2≺ω2.
In particular, we can write ς⋦i2η if ς⪯i2η and ς≠η, i.e., one of 2, 3 and 4 is satisfied, and we will write ς≺i2η if only 4 is satisfied.
Now, let us recall some basic concepts and notations, which will be used in the sequel.
Definition 2.1. [15] A bicomplex partial metric on a non-void set U is a function ρbcpms:U×U→C+2, where C+2={ς:ς=ϑ1+ϑ2i1+ϑ3i2+ϑ4i1i2,ϑ1,ϑ2,ϑ3,ϑ4∈C+0} and C+0={ϑ1∈C0|ϑ1≥0} such that for all φ,ζ,z∈U:
1. 0⪯i2ρbcpms(φ,φ)⪯i2ρbcpms(φ,ζ) (small self-distances),
2. ρbcpms(φ,ζ)=ρbcpms(ζ,φ) (symmetry),
3. ρbcpms(φ,φ)=ρbcpms(φ,ζ)=ρbcpms(ζ,ζ) if and only if φ=ζ (equality),
4. ρbcpms(φ,ζ)⪯i2ρbcpms(φ,z)+ρbcpms(z,ζ)−ρbcpms(z,z) (triangularity) .
A bicomplex partial metric space is a pair (U,ρbcpms) such that U is a non-void set and ρbcpms is a bicomplex partial metric on U.
Example 2.2. Let U=[0,∞) be endowed with bicomplex partial metric space ρbcpms:U×U→C+2 with ρbcpms(φ,ζ)=max, where e^{i_{2}\theta} = \cos \theta +i_{2}\sin \theta , for all \varphi, \zeta\in \mathcal{U} and 0\leq \theta\leq \frac{\pi}{2} . Obviously, (\mathcal{U}, \rho_{bcpms}) is a bicomplex partial metric space.
Definition 2.3. [15] A bicomplex partial metric space \mathcal{U} is said to be a T_{0} space if for any pair of distinct points of \mathcal{U} , there exists at least one open set which contains one of them but not the other.
Theorem 2.4. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space; then, (\mathcal{U}, \rho_{bcpms}) is T_{0} .
Definition 2.5. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space. A sequence \{\varphi_{\tau}\} in \mathcal{U} is said to be convergent and converges to \varphi\in\mathcal{U} if for every 0\prec_{i_{2}}\epsilon\in \mathscr{C}^{+}_{2} there exists \mathcal{N}\in \mathbb{N} such that \varphi_{\tau}\in \mathfrak{B}_{ \rho_{bcpms}}(\varphi, \epsilon) = \{\omega\in \mathcal{U}:\rho_{bcpms}(\varphi, \omega) < \epsilon+\rho_{bcpms}(\varphi, \varphi)\} for all \tau\geq \mathcal{N} , and it is denoted by \lim\limits_{\tau\rightarrow \infty} \varphi_{\tau} = \varphi .
Lemma 2.6. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space. A sequence \{\varphi_{\tau}\}\in \mathcal{U} is converges to \varphi\in \mathcal{U} iff \rho_{bcpms}(\varphi, \varphi) = \lim\limits_{\tau \to \infty} \rho_{bcpms}(\varphi, \varphi_{\tau}) .
Definition 2.7. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space. A sequence \{\varphi_{\tau}\} in \mathcal{U} is said to be a Cauchy sequence in (\mathcal{U}, \rho_{bcpms}) if for any \epsilon > 0 there exist \vartheta\in \mathscr{C}^{+}_{2} and \mathcal{N}\in \mathbb{N} such that || \rho_{bcpms}(\varphi_{\tau}, \varphi_{\upsilon})-\vartheta|| < \epsilon for all \tau, \upsilon\geq\mathcal{N} .
Definition 2.8. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space. Let \{\varphi_{\tau}\} be any sequence in \mathcal{U} . Then,
1. If every Cauchy sequence in \mathcal{U} is convergent in \mathcal{U} , then (\mathcal{U}, \rho_{bcpms}) is said to be a complete bicomplex partial metric space.
2. A mapping \mathcal{S}:\mathcal{U} \to \mathcal{U} is said to be continuous at \varphi_{0}\in \mathcal{U} if for every \epsilon > 0 , there exists \delta > 0 such that \mathcal{S}(\mathfrak{B}_{ \rho_{bcpms}}(\varphi_{0}, \delta))\subset \mathfrak{B}_{ \rho_{bcpms}}(\mathcal{S}(\varphi_{0}, \epsilon)) .
Lemma 2.9. [15] Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space and \{\varphi_{\tau}\} be a sequence in \mathcal{U} . Then, \{\varphi_{\tau}\} is a Cauchy sequence in \mathcal{U} iff \lim\limits_{\tau, \upsilon\to \infty} \rho_{bcpms}(\varphi_{\tau}, \varphi_{\upsilon}) = \rho_{bcpms}(\varphi, \varphi) .
Definition 2.10. Let (\mathcal{U}, \rho_{bcpms}) be a bicomplex partial metric space. Then, an element (\varphi, \zeta)\in \mathcal{U}\times \mathcal{U} is said to be a coupled fixed point of the mapping \mathcal{S}: \mathcal{U}\times \mathcal{U}\to \mathcal{U} if \mathcal{S}(\varphi, \zeta) = \varphi and \mathcal{S}(\zeta, \varphi) = \zeta .
Theorem 2.11. [15] Let (\mathcal{U}, \rho_{bcpms}) be a complete bicomplex partial metric space and \mathcal{S}, \mathcal{T} \colon \mathcal{ U} \rightarrow \mathcal{U} be two continuous mappings such that
\begin{align*} \rho_{bcpms}(\mathcal{S} \varphi, \mathcal{T} \zeta) &\preceq_{i_{2}} \mathfrak{l} \max\{ \rho_{bcpms}(\varphi, \zeta), \rho_{bcpms}(\varphi, \mathcal{S} \varphi), \rho_{bcpms}(\zeta, \mathcal{T} \zeta), \notag \\ &\; \; \; \; \dfrac{1}{2}( \rho_{bcpms}(\varphi, \mathcal{T} \zeta)+ \rho_{bcpms}(\zeta, \mathcal{S} \varphi))\}, \label{e1} \end{align*} |
for all \varphi, \zeta \in \mathcal{U} , where 0\leq \mathfrak{l} < 1 . Then, the pair (\mathcal{S}, \mathcal{T}) has a unique common fixed point, and \rho_{bcpms}(\varphi^{*}, \varphi^{*}) = 0 .
Inspired by Theorem 2.11, here we prove coupled fixed point theorems on a bicomplex partial metric space with an application.
Theorem 3.1. Let (\mathcal{U}, \rho_{bcpms}) be a complete bicomplex partial metric space. Suppose that the mapping \mathcal{S}:\mathcal{U}\times \mathcal{U}\to \mathcal{U} satisfies the following contractive condition:
\begin{equation*} \rho_{bcpms}(\mathcal{S}(\varphi , \zeta), \mathcal{S}(\nu, \mu)) \preceq_{i_{2}} \lambda\rho_{bcpms}(\mathcal{S}(\varphi, \zeta) , \varphi )+ \mathfrak{l}\rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu), \end{equation*} |
for all \varphi, \zeta, \nu, \mu\in \mathcal{U} , where \lambda, \mathfrak{l} are nonnegative constants with \lambda+\mathfrak{l} < 1 . Then, \mathcal{S} has a unique coupled fixed point.
Proof. Choose \nu_{0}, \mu_{0}\in \mathcal{U} and set \nu_{1} = \mathcal{S}(\nu_{0}, \mu_{0}) and \mu_{1} = \mathcal{S}(\mu_{0}, \nu_{0}) . Continuing this process, set \nu_{\tau+1} = \mathcal{S}(\nu_{\tau}, \mu_{\tau}) and \mu_{\tau+1} = \mathcal{S}(\mu_{\tau}, \nu_{\tau}) . Then,
\begin{align*} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})& = \rho_{bcpms}(\mathcal{S}(\nu_{\tau-1}, \mu_{\tau-1}), \mathcal{S}(\nu_{\tau}, \mu_{\tau}))\\ &\preceq_{i_{2}} \lambda\rho_{bcpms}(\mathcal{S}(\nu_{\tau-1}, \mu_{\tau-1}), \nu_{\tau-1})+\mathfrak{l}\rho_{bcpms}(\mathcal{S}(\nu_{\tau}, \mu_{\tau}), \nu_{\tau})\\ & = \lambda\rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})+\mathfrak{l}\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau})\\ \rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})&\preceq_{i_{2}} \frac{\lambda}{1-\mathfrak{l}}\rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1}), \end{align*} |
which implies that
\begin{align} \lvert\lvert\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})\rvert \rvert\leq \mathfrak{z} \lvert\lvert\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})\rvert\rvert \end{align} | (3.1) |
where \mathfrak{z} = \frac{\lambda}{1-\mathfrak{l}} < 1 . Similarly, one can prove that
\begin{align} ||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||\leq \mathfrak{z}\lvert \lvert\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})\rvert\rvert. \end{align} | (3.2) |
From (3.1) and (3.2), we get
\begin{align*} \lvert\lvert\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})\rvert\rvert+ ||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||&\leq \mathfrak{z} (\lvert\lvert\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})\rvert\rvert\\ &+\lvert\lvert \rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||), \end{align*} |
where \mathfrak{z} < 1 .
Also,
\begin{align} ||\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau+2})||\leq \mathfrak{z}||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})|| \end{align} | (3.3) |
\begin{align} ||\rho_{bcpms}(\mu_{\tau+1}, \mu_{\tau+2})||\leq \mathfrak{z}||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||. \end{align} | (3.4) |
From (3.3) and (3.4), we get
\begin{align*} ||\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau+2})||+||\rho_{bcpms}(\mu_{\tau+1}, \mu_{\tau+2})||&\leq \mathfrak{z}(||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||\\ &+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||). \end{align*} |
Repeating this way, we get
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||&\leq \mathfrak{z}(||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||)\\ &\leq \mathfrak{z}^{2}(||\rho_{bcpms}(\mu_{\tau-2}, \mu_{\tau-1})||\\ &+||\rho_{bcpms}(\nu_{\tau-2}, \nu_{\tau-1})||)\\ &\leq \dots \leq \mathfrak{z}^{\tau}(||\rho_{bcpms}(\mu_{0}, \mu_{1})||\\ &+||\rho_{bcpms}(\nu_{0}, \nu_{1})||). \end{align*} |
Now, if ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})|| = \gamma_{\tau} , then
\begin{align} \gamma_{\tau}\leq \mathfrak{z} \gamma_{\tau-1} \leq \dots \leq \mathfrak{z}^\tau\gamma_{0}. \end{align} | (3.5) |
If \gamma_{0} = 0 , then ||\rho_{bcpms}(\nu_{0}, \nu_{1})||+||\rho_{bcpms}(\mu_{0}, \mu_{1})|| = 0 . Hence, \nu_{0} = \nu_{1} = \mathcal{S}(\nu_{0}, \mu_{0}) and \mu_{0} = \mu_{1} = \mathcal{S}(\mu_{0}, \mu_{0}) , which implies that (\nu_{0}, \mu_{0}) is a coupled fixed point of \mathcal{S} . Let \gamma_{0} > 0 . For each \tau\geq \upsilon , we have
\begin{align*} \rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})&\preceq_{i_{2}} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})+\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})-\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-1})\\ &+\rho_{bcpms}(\nu_{\tau-2}, \nu_{\tau-3})+\rho_{bcpms}(\nu_{\tau-3}, \nu_{\tau-4})-\rho_{bcpms}(\nu_{\tau-3}, \nu_{\tau-3})\\ &+\dots +\rho_{bcpms}(\nu_{\upsilon+2}, \nu_{\upsilon+1})+\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon})-\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon+1})\\ &\preceq_{i_{2}} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})+\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})+\dots+\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon}), \end{align*} |
which implies that
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})||&\leq ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})||+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})||\\ &+\dots+||\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon})||. \end{align*} |
Similarly, one can prove that
\begin{align*} ||\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon})||&\leq ||\rho_{bcpms}({\mu_{\tau}, \mu_{\tau-1}})||+||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau-2})||\\ &+\dots +||\rho_{bcpms}(\mu_{\upsilon+1}, \mu_{\upsilon})||. \end{align*} |
Thus,
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon})||&\leq \gamma_{\tau-1}+\gamma_{\tau-2}+\gamma_{\tau-3}+\dots +\gamma_{\upsilon}\\ &\leq (\mathfrak{z}^{\tau-1}+\mathfrak{z}^{\tau-2}+\dots +\mathfrak{z}^{\upsilon})\gamma_{0}\\ &\leq \frac{\mathfrak{z}^{\upsilon}}{1-\mathfrak{z}}\gamma_{0}\rightarrow 0\, \, \text{as}\, \, \upsilon\rightarrow \infty, \end{align*} |
which implies that \{\nu_{\tau}\} and \{\mu_{\tau}\} are Cauchy sequences in (\mathcal{U}, \rho_{bcpms}) . Since the bicomplex partial metric space (\mathcal{U}, \rho_{bcpms}) is complete, there exist \nu, \mu\in \mathcal{U} such that \{\nu_{\tau}\}\rightarrow \nu and \{\mu_{\tau}\}\rightarrow \mu as \tau \rightarrow \infty , and
\begin{align*} \rho_{bcpms}(\nu, \nu) = \lim\limits_{\tau \rightarrow \infty }\rho_{bcpms}(\nu, \nu_{\tau}) = \lim\limits_{\tau, \upsilon \rightarrow \infty }\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon}) = 0, \\ \rho_{bcpms}(\mu, \mu) = \lim\limits_{\tau \rightarrow \infty }\rho_{bcpms}(\mu, \mu_{\tau}) = \lim\limits_{\tau, \upsilon \rightarrow \infty }\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon}) = 0. \end{align*} |
We now show that \nu = \mathcal{S}(\nu, \mu) . We suppose on the contrary that \nu\neq \mathcal{S}(\nu, \mu) and \mu\neq \mathcal{S} (\mu, \nu) , so that 0\prec_{i_{2}} \rho_{bcpms}(\nu, \mathcal{S}(\nu, \mu)) = \mathfrak{l}_{1} and 0\prec_{i_{2}}\rho_{bcpms}(\mu, \mathcal{S}(\mu, \nu)) = \mathfrak{l}_{2} . Then,
\begin{align*} \mathfrak{l}_{1} = \rho_{bcpms}(\nu, \mathcal{S}(\nu, \mu))&\preceq_{i_{2}} \rho_{bcpms}(\nu, \nu_{\tau+1})+\rho_{bcpms}(\nu_{\tau+1}, \mathcal{S}(\nu, \mu))\\ & = \rho_{bcpms}(\nu, \nu_{\tau+1})+\rho_{bcpms}(\mathcal{S}(\nu_{\tau}, \mu_{\tau}), \mathcal{S}(\nu, \mu))\\ &\preceq_{i_{2}} \rho_{bcpms}(\nu, \nu_{\tau+1})+\lambda\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})+\mathfrak{l}\rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu)\\ &\preceq_{i_{2}}\frac{1}{1-\mathfrak{l}} \rho_{bcpms}(\nu, \nu_{\tau+1})+\frac{\lambda}{1-\mathfrak{l}}\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau}), \end{align*} |
which implies that
\begin{align*} \lvert \lvert\mathfrak{l}_{1}\rvert\rvert\leq \frac{1}{1-\mathfrak{l}} \lvert\lvert\rho_{bcpms}(\nu, \nu_{\tau+1})\rvert\rvert+\frac{\lambda}{1-\mathfrak{l}}\lvert\lvert\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})\rvert\rvert. \end{align*} |
As \tau\rightarrow \infty , \lvert\lvert \mathfrak{l}_{1}\rvert\rvert\leq 0 . This is a contradiction, and therefore \lvert\lvert\rho_{bcpms}(\nu, \mathcal{S}(\nu, \mu))\rvert\rvert = 0 implies \nu = \mathcal{S}(\nu, \mu) . Similarly, we can prove that \mu = \mathcal{S}(\mu, \nu) . Thus (\nu, \mu) is a coupled fixed point of \mathcal{S} . Now, if (\mathfrak{g}, \mathfrak{h}) is another coupled fixed point of \mathcal{S} , then
\begin{align*} \rho_{bcpms}(\nu, \mathfrak{g}) = \rho_{bcpms}(\mathcal{S}(\nu, \mu), \mathcal{S}(\mathfrak{g}, \mathfrak{h})) &\preceq_{i_{2}} \lambda\rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu)+\mathfrak{l}\rho_{bcpms}(\mathcal{S}(\mathfrak{g}, \mathfrak{h}), \mathfrak{g})\\ & = \lambda\rho_{bcpms}(\nu, \nu)+\mathfrak{l}\rho_{bcpms}(\mathfrak{g}, \mathfrak{g}) = 0. \end{align*} |
Thus, we have \mathfrak{g} = \nu . Similarly, we get \mathfrak{h} = \mu . Therefore \mathcal{S} has a unique coupled fixed point.
Corollary 3.2. Let (\mathcal{U}, \rho_{bcpms}) be a complete bicomplex partial metric space. Suppose that the mapping \mathcal{S}:\mathcal{U}\times \mathcal{U} \to \mathcal{U} satisfies the following contractive condition:
\begin{equation} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \mu))\preceq_{i_{2}} \lambda(\rho_{bcpms}(\mathcal{S}(\varphi, \zeta) , \varphi)+\rho_{bcpms}( \mathcal{S}(\nu, \mu), \nu)), \end{equation} | (3.6) |
for all \varphi, \zeta, \nu, \mu\in \mathcal{U} , where 0\leq \lambda < \frac{1 }{2} . Then, \mathcal{S} has a unique coupled fixed point.
Theorem 3.3. Let (\mathcal{U}, \rho_{bcpms}) be a complete complex partial metric space. Suppose that the mapping \mathcal{S}:\mathcal{U}\times \mathcal{U}\to \mathcal{U} satisfies the following contractive condition:
\begin{equation*} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \mu))\preceq_{i_{2}} \lambda\rho_{bcpms}(\varphi, \nu)+\mathfrak{l}\rho_{bcpms}(\zeta, \mu), \end{equation*} |
for all \varphi, \zeta, \nu, \mu\in \mathcal{U} , where \lambda, \mathfrak{l} are nonnegative constants with \lambda+\mathfrak{l} < 1 . Then, \mathcal{S} has a unique coupled fixed point.
Proof. Choose \nu_{0}, \mu_{0}\in \mathcal{U} and set \nu_{1} = \mathcal{S}(\nu_{0}, \mu_{0}) and \mu_{1} = \mathcal{S}(\mu_{0}, \nu_{0}) . Continuing this process, set \nu_{\tau+1} = \mathcal{S}(\nu_{\tau}, \mu_{\tau}) and \mu_{\tau+1} = \mathcal{S}(\mu_{\tau}, \nu_{\tau}) . Then,
\begin{align*} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})& = \rho_{bcpms}(\mathcal{S}(\nu_{\tau-1}, \mu_{\tau-1}), \mathcal{S}(\nu_{\tau}, \mu_{\tau}))\\ &\preceq_{i_{2}} \lambda\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})+\mathfrak{l}\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau}), \end{align*} |
which implies that
\begin{align} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||\leq \lambda||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||+\mathfrak{l}||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||. \end{align} | (3.7) |
Similarly, one can prove that
\begin{align} ||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||\leq \lambda||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||+\mathfrak{l}||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||. \end{align} | (3.8) |
From (3.7) and (3.8), we get
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||&\leq (\lambda+\mathfrak{l})(||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||\\ &+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||)\\ & = \alpha(||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||\\ &+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||), \end{align*} |
where \alpha = \lambda+\mathfrak{l} < 1 . Also,
\begin{align} ||\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau+2})||\leq \lambda||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||+\mathfrak{l}||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})|| \end{align} | (3.9) |
\begin{align} ||\rho_{bcpms}(\mu_{\tau+1}, \mu_{\tau+2})||\leq \lambda||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||+\mathfrak{l}||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||. \end{align} | (3.10) |
From (3.9) and (3.10), we get
\begin{align*} ||\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau+2})||+||\rho_{bcpms}(\mu_{\tau+1}, \mu_{\tau+2})||&\leq (\lambda+\mathfrak{l})(||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||\\ &+||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||)\\ & = \alpha(||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||\\ &+||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||). \end{align*} |
Repeating this way, we get
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{n+1})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})||&\leq\alpha(||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau})||\\ &+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau})||)\\ &\leq \alpha^{2}(||\rho_{bcpms}(\mu_{\tau-2}, \mu_{\tau-1})||\\ &+||\rho_{bcpms}(\nu_{\tau-2}, \nu_{\tau-1})||)\\ &\leq \dots \leq \alpha^{\tau}(||\rho_{bcpms}(\mu_{0}, \mu_{1})||\\ &+||\rho_{bcpms}(\nu_{0}, \nu_{1})||). \end{align*} |
Now, if ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau+1})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\tau+1})|| = \gamma_{\tau} , then
\begin{align} \gamma_{\tau}\leq \alpha \gamma_{\tau-1} \leq \dots \leq \alpha^{\tau}\gamma_{0}. \end{align} | (3.11) |
If \gamma_{0} = 0 , then ||\rho_{bcpms}(\nu_{0}, \nu_{1})||+||\rho_{bcpms}(\mu_{0}, \mu_{1})|| = 0 . Hence, \nu_{0} = \nu_{1} = \mathcal{S}(\nu_{0}, \mu_{0}) and \mu_{0} = \mu_{1} = \mathcal{S}(\mu_{0}, \nu_{0}) , which implies that (\nu_{0}, \mu_{0}) is a coupled fixed point of \mathcal{S} . Let \gamma_{0} > 0 . For each \tau\geq \upsilon , we have
\begin{align*} \rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})&\preceq_{i_{2}} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})+\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})-\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-1})\\ &+\rho_{bcpms}(\nu_{\tau-2}, \nu_{\tau-3})+\rho_{bcpms}(\nu_{\tau-3}, \nu_{\tau-4})-\rho_{bcpms}(\nu_{\tau-3}, \nu_{\tau-3})\\ &+\dots +\rho_{bcpms}(\nu_{\upsilon+2}, \nu_{\upsilon+1})+\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon})-\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon+1})\\ &\preceq_{i_{2}} \rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})+\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})+\dots+\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon}), \end{align*} |
which implies that
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})||&\leq ||\rho_{bcpms}(\nu_{\tau}, \nu_{\tau-1})||+||\rho_{bcpms}(\nu_{\tau-1}, \nu_{\tau-2})||\\ &+\dots+||\rho_{bcpms}(\nu_{\upsilon+1}, \nu_{\upsilon})||. \end{align*} |
Similarly, one can prove that
\begin{align*} ||\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon})||&\leq ||\rho_{bcpms}({\mu_{\tau}, \mu_{\tau-1}})||+||\rho_{bcpms}(\mu_{\tau-1}, \mu_{\tau-2})||\\ &+\dots+||\rho_{bcpms}(\mu_{\upsilon+1}, \mu_{\upsilon})||. \end{align*} |
Thus,
\begin{align*} ||\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon})||+||\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon})||&\leq \gamma_{\tau-1}+\gamma_{\tau-2}+\gamma_{\tau-3}+\dots +\gamma_{\upsilon}\\ &\leq (\alpha^{\tau-1}+\alpha^{\tau-2}+\dots +\alpha^{\upsilon})\gamma_{0}\\ &\leq \frac{\alpha^{\upsilon}}{1-\alpha}\gamma_{0}\, \, {\rm{as}}\, \, \tau\to \infty, \end{align*} |
which implies that \{\nu_{\tau}\} and \{\mu_{\tau}\} are Cauchy sequences in (\mathcal{U}, \rho_{bcpms}) . Since the bicomplex partial metric space (\mathcal{U}, \rho_{bcpms}) is complete, there exist \nu, \mu\in \mathcal{U} such that \{\nu_{\tau}\}\rightarrow \nu and \{\mu_{\tau}\}\rightarrow \mu as \tau \rightarrow \infty , and
\begin{align*} \rho_{bcpms}(\nu, \nu) = \lim\limits_{\tau \rightarrow \infty }\rho_{bcpms}(\nu, \nu_{\tau}) = \lim\limits_{\tau, \upsilon \rightarrow \infty }\rho_{bcpms}(\nu_{\tau}, \nu_{\upsilon}) = 0, \\ \rho_{bcpms}(\mu, \mu) = \lim\limits_{\tau \rightarrow \infty }\rho_{bcpms}(\mu, \mu_{\tau}) = \lim\limits_{\tau, \upsilon \rightarrow \infty }\rho_{bcpms}(\mu_{\tau}, \mu_{\upsilon}) = 0. \end{align*} |
Therefore,
\begin{align*} \rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu)&\leq \rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu_{\tau+1})+\rho_{bcpms}(\nu_{\tau+1}, \nu)-\rho_{bcpms}(\nu_{\tau+1}, \nu_{\tau+1}), \\ &\leq \rho_{bcpms}(\mathcal{S}(\nu, \mu)), \mathcal{S}(\nu_{\tau}, \mu_{\tau})+\rho_{bcpms}(\nu_{\tau+1}, \nu)\\ &\leq \lambda\rho_{bcpms}(\nu_{\tau}, \nu)+\mathfrak{l}\rho_{bcpms}(\mu_{\tau}, \mu)+\rho_{bcpms}(\nu_{\tau+1}, \nu). \end{align*} |
As \tau \rightarrow \infty , from (3.6) and (3.12) we obtain \rho_{bcpms}(\mathcal{S}(\nu, \mu), \nu) = 0 . Therefore \mathcal{S}(\nu, \mu) = \nu . Similarly, we can prove \mathcal{S}(\mu, \nu) = \mu , which implies that (\nu, \mu) is a coupled fixed point of \mathcal{S} . Now, if (\mathfrak{g}_{1}, \mathfrak{h}_{1}) is another coupled fixed point of \mathcal{S} , then
\begin{align*} \rho_{bcpms}(\mathfrak{g}_{1}, \nu) = \rho_{bcpms}(\mathcal{S}(\mathfrak{g}_{1}, \mathfrak{h}_{1}), \mathcal{S}(\nu, \mu)) &\preceq_{i_{2}} \lambda\rho_{bcpms}(\mathfrak{g}_{1}, \nu)+\mathfrak{l}\rho_{bcpms}(\mathfrak{h}_{1}, \mu), \\ \rho_{bcpms}(\mathfrak{h}_{1}, \mu) = \rho_{bcpms}(\mathcal{S}(\mathfrak{h}_{1}, \mathfrak{g}_{1}), \mathcal{S}(\mu, \nu))&\preceq_{i_{2}} \lambda\rho_{bcpms}(\mathfrak{h}_{1}, \mu)+\mathfrak{l}\rho_{bcpms}(\mathfrak{g}_{1}, \nu), \end{align*} |
which implies that
\begin{align} \lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert&\leq \lambda\lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert+\mathfrak{l}\lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert, \end{align} | (3.12) |
\begin{align} \lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert&\leq \lambda\lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert+\mathfrak{l}\lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert. \end{align} | (3.13) |
From (3.12) and (3.13), we get
\begin{align*} \lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert+\lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert\leq (\lambda+\mathfrak{l})[\lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert+\lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert]. \end{align*} |
Since \lambda+\mathfrak{l} < 1 , this implies that \lvert\lvert\rho_{bcpms}(\mathfrak{g}_{1}, \nu)\rvert\rvert+\lvert\lvert\rho_{bcpms}(\mathfrak{h}_{1}, \mu)\rvert\rvert = 0 . Therefore, \nu = \mathfrak{g}_{1} and \mu = \mathfrak{h}_{1} . Thus, \mathcal{S} has a unique coupled fixed point.
Corollary 3.4. Let (\mathcal{U}, \rho_{bcpms}) be a complete bicomplex partial metric space. Suppose that the mapping \mathcal{S}:\mathcal{U}\times \mathcal{U} \to \mathcal{U} satisfies the following contractive condition:
\begin{equation} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \mu))\preceq_{i_{2}} \lambda(\rho_{bcpms}(\varphi, \nu)+\rho_{bcpms}(\zeta, \mu)), \end{equation} | (3.14) |
for all \varphi, \zeta, \nu, \mu\in \mathcal{U} , where 0\leq \lambda < \frac{1 }{2} . Then, \mathcal{S} has a unique coupled fixed point.
Example 3.5. Let \mathcal{U} = [0, \infty) and define the bicomplex partial metric \rho_{bcpms} :\mathcal{U}\times\mathcal{U}\to \mathscr{C}_{2}^{+} defined by
\begin{align*} \rho_{bcpms}(\varphi, \zeta) = \max\{\varphi, \zeta\}e^{i_{2}\theta}, \, \, 0\leq\theta\leq\frac{\pi}{2}. \end{align*} |
We define a partial order \preceq in \mathscr{C}_{2}^{+} as \varphi \preceq \zeta iff \varphi \leq \zeta . Clearly, (\mathcal{U}, \rho_{bcpms}) is a complete bicomplex partial metric space.
Consider the mapping \mathcal{S}:\mathcal{U}\times \mathcal{U}\to \mathcal{U } defined by
\begin{align*} \mathcal{S}(\varphi, \zeta) = \frac{\varphi+\zeta}{4}\, \, \, \forall \varphi, \zeta\in \mathcal{U}. \end{align*} |
Now,
\begin{align*} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \mu))& = \rho_{bcpms} \bigg(\frac{\varphi+\zeta}{4}, \frac{\nu+\mu}{4}\bigg) \\ & = \frac{1}{4}\max\{\varphi+\zeta, \nu+\mu\} e^{i_{2}\theta} \\ &\preceq_{i_{2}}\frac{1}{4}\bigg[\max\{\varphi, \nu\}+\max\{\zeta, \mu\}\bigg] e^{i_{2}\theta} \\ & = \frac{1}{4}\bigg[\rho_{bcpms}(\varphi, \nu)+\rho_{bcpms}(\zeta, \mu)\bigg] \\ & = \lambda\bigg(\rho_{bcpms}(\varphi, \nu)+\rho_{bcpms}(\zeta, \mu)\bigg), \end{align*} |
for all \varphi, \zeta, \nu, \mu\in \mathcal{U} , where 0\leq\lambda = \frac{1 }{4} < \frac{1}{2} . Therefore, all the conditions of Corollary 3.4 are satisfied, then the mapping \mathcal{S} has a unique coupled fixed point (0, 0) in \mathcal{U} .
As an application of Theorem 3.3, we find an existence and uniqueness result for a type of the following system of nonlinear integral equations:
\begin{align} \varphi (\mu )& = \int_{0}^{\mathcal{M}}\kappa (\mu , \mathfrak{p})[\mathcal{G} _{1}(\mathfrak{p}, \varphi (\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \zeta (\mathfrak{p}))]d\mathfrak{p}+\delta (\mu ), \\ \zeta (\mu )& = \int_{0}^{\mathcal{M}}\kappa (\mu , \mathfrak{p})[\mathcal{G} _{1}(\mathfrak{p}, \zeta (\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \varphi (\mathfrak{p}))]d\mathfrak{p}+\delta (\mu ), \, \, \mu , \in \lbrack 0, \mathcal{M }], \mathcal{M}\geq1. \end{align} | (4.1) |
Let \mathcal{U} = C([0, \mathcal{M}], \mathbb{R}) be the class of all real valued continuous functions on [0, \mathcal{M}] . We define a partial order \preceq in \mathscr{C}_{2}^{+} as x\preceq y iff x \leq y . Define \mathcal{S}:\mathcal{U}\times\mathcal{U}\to \mathcal{U} by
\begin{align*} \mathcal{S}(\varphi, \zeta)(\mu) = \int^{\mathcal{M}}_{0}\kappa(\mu, \mathfrak{p} )[\mathcal{G}_{1}(\mathfrak{p}, \varphi(\mathfrak{p}))+\mathcal{G}_{2}( \mathfrak{p}, \zeta(\mathfrak{p}))] d\mathfrak{p}+\delta(\mu). \end{align*} |
Obviously, (\varphi(\mu), \zeta(\mu)) is a solution of system of nonlinear integral equations (4.1) iff (\varphi(\mu), \zeta(\mu)) is a coupled fixed point of \mathcal{S} . Define \rho _{bcpms}:\mathcal{U} \times \mathcal{U}\rightarrow \mathscr{C}_{2} by
\begin{equation*} \rho _{bcpms}(\varphi , \zeta ) = (|\varphi -\zeta |+1)e^{i_{2}\theta }, \end{equation*} |
for all \varphi, \zeta \in \mathcal{U} , where 0\leq \theta \leq \frac{\pi }{2} . Now, we state and prove our result as follows.
Theorem 4.1. Suppose the following:
1. The mappings \mathcal{G}_{1}:[0, \mathcal{M}]\times\mathbb{R}\to \mathbb{R} , \mathcal{G}_{2}:[0, \mathcal{M}]\times\mathbb{R}\to\mathbb{R} , \delta:[0, \mathcal{M}]\to \mathbb{R} and \kappa:[0, \mathcal{M}]\times \mathbb{R}\to [0, \infty) are continuous.
2. There exists \eta > 0 , and \lambda, \mathfrak{l} are nonnegative constants with \lambda+ \mathfrak{l} < 1 , such that
\begin{align*} |\mathcal{G}_{1}(\mathfrak{p}, \varphi(\mathfrak{p}))-\mathcal{G}_{1}( \mathfrak{p}, \zeta(\mathfrak{p}))|&\preceq_{i_{2}}\eta\lambda (|\varphi-\zeta|+1)-\frac{1}{2}, \\ |\mathcal{G}_{2}(\mathfrak{p}, \zeta(\mathfrak{p}))-\mathcal{G}_{2}(\mathfrak{ p}, \varphi(\mathfrak{p}))|&\preceq_{i_{2}}\eta\mathfrak{l} (|\zeta-\varphi|+1)-\frac{1}{2}. \end{align*} |
3. \int^{\mathcal{M}}_{0}\eta|\kappa(\mu, \mathfrak{p})|d\mathfrak{p} \preceq_{i_{2}}1 .
Then, the integral equation (4.1) has a unique solution in \mathcal{U} .
Proof. Consider
\begin{align*} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \varPhi))& = (|\mathcal{S}(\varphi, \zeta)-\mathcal{S}(\nu, \varPhi)|+1)e^{i_{2}\theta}\\ & = \bigg(|\int^{\mathcal{M}}_{0}\kappa(\mu, \mathfrak{p})[\mathcal{G}_{1}(\mathfrak{p}, \varphi(\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \zeta(\mathfrak{p}))] d\mathfrak{p}+\delta(\mu)\\ &-\bigg(\int^{\mathcal{M}}_{0}\kappa(\mu, \mathfrak{p})[\mathcal{G}_{1}(\mathfrak{p}, \nu(\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \varPhi(\mathfrak{p}))] d\mathfrak{p}+\delta(\mu)\bigg)|+1\bigg)e^{i_{2}\theta}\\ & = \bigg(|\int^{\mathcal{M}}_{0}\kappa(\mu, \mathfrak{p})[\mathcal{G}_{1}(\mathfrak{p}, \varphi(\mathfrak{p}))-\mathcal{G}_{1}(\mathfrak{p}, \nu(\mathfrak{p}))\\ &+\mathcal{G}_{2}(\mathfrak{p}, \zeta(\mathfrak{p}))-\mathcal{G}_{2}(\mathfrak{p}, \varPhi(\mathfrak{p}))] d\mathfrak{p}|+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}}\bigg(\int^{\mathcal{M}}_{0}|\kappa(\mu, \mathfrak{p})|[|\mathcal{G}_{1}(\mathfrak{p}, \varphi(\mathfrak{p}))-\mathcal{G}_{1}(\mathfrak{p}, \nu(\mathfrak{p}))|\\ &+|\mathcal{G}_{2}(\mathfrak{p}, \zeta(\mathfrak{p}))-\mathcal{G}_{2}(\mathfrak{p}, \varPhi(\mathfrak{p}))|] d\mathfrak{p}+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}}\bigg(\int^{\mathcal{M}}_{0}|\kappa(\mu, \mathfrak{p})|d\mathfrak{p}(\eta\lambda (|\varphi-\nu|+1)-\frac{1}{2}\\ &+\eta\mathfrak{l} (|\zeta-\varPhi|+1)-\frac{1}{2})+1\bigg)e^{i_{2}\theta}\\ & = \bigg(\int^{\mathcal{M}}_{0}\eta|\kappa(\mu, \mathfrak{p})|d\mathfrak{p}(\lambda (|\varphi-\nu|+1)\\ &+\mathfrak{l} (|\zeta-\varPhi|+1))\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}} \bigg(\lambda (|\varphi-\nu|+1)+\mathfrak{l} (|\zeta-\varPhi|+1)\bigg)e^{i_{2}\theta}\\ & = \lambda\rho_{bcpms}(\varphi, \nu)+\mathfrak{l}\rho_{bcpms}(\zeta, \varPhi) \end{align*} |
for all \varphi, \zeta, \nu, \varPhi\in \mathcal{U} . Hence, all the hypotheses of Theorem 3.3 are verified, and consequently, the integral equation (4.1) has a unique solution.
Example 4.2. Let \mathcal{U} = C([0, 1], \mathbb{R}) . Now, consider the integral equation in \mathcal{U} as
\begin{align} \varphi(\mu) = \int_{0}^{1}\frac{\mathfrak{\mu p}}{23(\mu+5)}\bigg[\frac{1}{ 1+\varphi(\mathfrak{p})}+\frac{1}{2+\zeta(\mathfrak{p})}\bigg]d\mathfrak{p}+ \frac{6\mu^{2}}{5} \\ \zeta(\mu) = \int_{0}^{1}\frac{ \mathfrak{\mu p}}{23(\mu+5)}\bigg[\frac{1}{ 1+\zeta(\mathfrak{p})}+\frac{1}{ 2+\varphi(\mathfrak{p})}\bigg]d\mathfrak{p}+ \frac{6\mu^{2}}{5}. \end{align} | (4.2) |
Then, clearly the above equation is in the form of the following equation:
\begin{align} \varphi (\mu )& = \int_{0}^{\mathcal{M}}\kappa (\mu , \mathfrak{p})[\mathcal{G} _{1}(\mathfrak{p}, \varphi (\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \zeta (\mathfrak{p}))]d\mathfrak{p}+\delta (\mu ), \\ \zeta (\mu )& = \int_{0}^{\mathcal{M}}\kappa (\mu , \mathfrak{p})[\mathcal{G} _{1}(\mathfrak{p}, \zeta (\mathfrak{p}))+\mathcal{G}_{2}(\mathfrak{p}, \varphi (\mathfrak{p}))]d\mathfrak{p}+\delta (\mu ), \, \, \mu , \in \lbrack 0, \mathcal{M }], \end{align} | (4.3) |
where \delta(\mu) = \frac{6\mu^{2}}{5} , \kappa(\mu, \mathfrak{p}) = \frac{ \mathfrak{\mu p}}{23(\mu+5)} , \mathcal{G}_{1}(\mathfrak{p}, \mu) = \frac{1}{ 1+\mu} , \mathcal{G}_{2}(\mathfrak{p}, \mu) = \frac{1}{2+\mu} and \mathcal{M} = 1 . That is, (4.2) is a special case of (4.1) in Theorem 4.1. Here, it is easy to verify that the functions \delta(\mu) , \kappa(\mu, \mathfrak{p}) , \mathcal{G}_{1}(\mathfrak{p}, \mu) and \mathcal{ G}_{2}(\mathfrak{p}, \mu) are continuous. Moreover, there exist \eta = 10 , \lambda = \frac{1}{3} and \mathfrak{l} = \frac{1}{4} with \lambda+\mathfrak{l } < 1 such that
\begin{align*} |\mathcal{G}_{1}(\mathfrak{p}, \varphi)-\mathcal{G}_{1}(\mathfrak{p} , \zeta)|&\leq\eta\lambda (|\varphi-\zeta|+1)-\frac{1}{2}, \\ |\mathcal{G}_{2}(\mathfrak{p}, \zeta)-\mathcal{G}_{2}(\mathfrak{p} , \varphi)|&\leq\eta\mathfrak{l} (|\zeta-\varphi|+1)-\frac{1}{2} \end{align*} |
and \int^{\mathcal{M}}_{0}\eta|\kappa(\mu, \mathfrak{p})|d\mathfrak{p} = \int^{1}_{0}\frac{\eta\mu\mathfrak{p}}{23(\mu+5)}d\mathfrak{p} = \frac{\mu\eta }{23(\mu+5)} < 1 . Therefore, all the conditions of Theorem 3.3 are satisfied. Hence, system (4.2) has a unique solution (\varphi^{*}, \zeta^{*}) in \mathcal{U}\times\mathcal{U} .
As an application of Corollary 3.4, we find an existence and uniqueness result for a type of the following system of Fredholm integral equations:
\begin{align} \varphi (\mu )& = \int_{\mathcal{E}}\mathcal{G}(\mu , \mathfrak{p}, \varphi ( \mathfrak{p}), \zeta (\mathfrak{p}))d\mathfrak{p}+\delta (\mu ), \, \, \mu , \mathfrak{p}\in \mathcal{E}, \\ \zeta (\mu )& = \int_{\mathcal{E}}\mathcal{G}(\mu , \mathfrak{p}, \zeta ( \mathfrak{p}), \varphi (\mathfrak{p}))d\mathfrak{p}+\delta (\mu ), \, \, \mu , \mathfrak{p}\in \mathcal{E}, \end{align} | (4.4) |
where \mathcal{E} is a measurable, \mathcal{G}:\mathcal{E}\times \mathcal{ E}\times \mathbb{R}\times \mathbb{R}\rightarrow \mathbb{R} , and \delta \in \mathcal{L}^{\infty }(\mathcal{E}) . Let \mathcal{U} = \mathcal{L}^{\infty }(\mathcal{E}) . We define a partial order \preceq in \mathscr{C}_{2}^{+} as x\preceq y iff x\leq y. Define \mathcal{S}:\mathcal{U}\times \mathcal{U}\to \mathcal{U} by
\begin{align*} \mathcal{S}(\varphi, \zeta)(\mu) = \int_{\mathcal{E}}\mathcal{G}(\mu , \mathfrak{ p}, \varphi (\mathfrak{p}), \zeta (\mathfrak{p}))d\mathfrak{p}+\delta (\mu ). \end{align*} |
Obviously, (\varphi(\mu), \zeta(\mu)) is a solution of the system of Fredholm integral equations (4.4) iff (\varphi(\mu), \zeta(\mu)) is a coupled fixed point of \mathcal{S} . Define \rho _{bcpms}:\mathcal{U}\times \mathcal{U}\rightarrow \mathscr{C}_{2} by
\begin{equation*} \rho _{bcpms}(\varphi , \zeta ) = (|\varphi -\zeta |+1)e^{i_{2}\theta }, \end{equation*} |
for all \varphi, \zeta \in \mathcal{U} , where 0\leq \theta \leq \frac{\pi }{2} . Now, we state and prove our result as follows.
Theorem 4.3. Suppose the following:
1. There exists a continuous function \kappa:\mathcal{E}\times\mathcal{E} \to \mathbb{R} such that
\begin{align*} |\mathcal{G}(\mu, \mathfrak{p}, \varphi(\mathfrak{p}), \zeta(\mathfrak{p}))- \mathcal{G}(\mu, \mathfrak{p}, \nu(\mathfrak{p}), \varPhi(\mathfrak{p} ))|&\preceq_{i_{2}} |\kappa(\mu, \mathfrak{p})|(|\varphi(\mathfrak{p})-\nu( \mathfrak{p})| \\ &+|\zeta(\mathfrak{p})-\varPhi(\mathfrak{p})|-2), \end{align*} |
for all \varphi, \zeta, \nu, \varPhi\in \mathcal{U} , \mu, \mathfrak{p}\in \mathcal{E} .
2. \int_{\mathcal{E}}|\kappa(\mu, \mathfrak{p})|d\mathfrak{p} \preceq_{i_{2}} \frac{1}{4}\preceq_{i_{2}}1 .
Then, the integral equation (4.4) has a unique solution in \mathcal{U} .
Proof. Consider
\begin{align*} \rho_{bcpms}(\mathcal{S}(\varphi, \zeta), \mathcal{S}(\nu, \varPhi))& = (|\mathcal{S}(\varphi, \zeta)-\mathcal{S}(\nu, \varPhi)|+1)e^{i_{2}\theta}\\ & = \bigg(|\int_{\mathcal{E}}\mathcal{G}(\mu, \mathfrak{p}, \varphi(\mathfrak{p}), \zeta(\mathfrak{p}))d\mathfrak{p}+\delta(\mu)\\ &-\bigg(\int_{\mathcal{E}}\mathcal{G}(\mu, \mathfrak{p}, \nu(\mathfrak{p}), \varPhi(\mathfrak{p}))d\mathfrak{p}+\delta(\mu)\bigg)|+1\bigg)e^{i_{2}\theta}\\ & = \bigg(|\int_{\mathcal{E}}\bigg(\mathcal{G}(\mu, \mathfrak{p}, \varphi(\mathfrak{p}), \zeta(\mathfrak{p}))\\ &-\mathcal{G}(\mu, \mathfrak{p}, \nu(\mathfrak{p}), \varPhi(\mathfrak{p}))\bigg)d\mathfrak{p}|+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}}\bigg(\int_{\mathcal{E}}|\mathcal{G}(\mu, \mathfrak{p}, \varphi(\mathfrak{p}), \zeta(\mathfrak{p}))-\mathcal{G}(\mu, \mathfrak{p}, \nu(\mathfrak{p}), \varPhi(\mathfrak{p}))|d\mathfrak{p}+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}}\bigg(\int_{\mathcal{E}}|\kappa(\mu, \mathfrak{p})|(|\varphi(\mathfrak{p})-\nu(\mathfrak{p})|+|\zeta(\mathfrak{p})-\varPhi(\mathfrak{p})|-2)d\mathfrak{p}+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}}\bigg(\int_{\mathcal{E}}|\kappa(\mu, \mathfrak{p})|d\mathfrak{p}(|\varphi(\mathfrak{p})-\nu(\mathfrak{p})|+|\zeta(\mathfrak{p})-\varPhi(\mathfrak{p})|-2)+1\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}} \frac{1}{4}\bigg(|\varphi(\mathfrak{p})-\nu(\mathfrak{p})|+|\zeta(\mathfrak{p})-\varPhi(\mathfrak{p})|-2+4\bigg)e^{i_{2}\theta}\\ &\preceq_{i_{2}} \frac{1}{4}(\rho_{bcpms}(\varphi, \nu)+\rho_{bcpms}(\zeta, \varPhi))\\ & = \lambda(\rho_{bcpms}(\varphi, \nu)+\rho_{bcpms}(\zeta, \varPhi)), \end{align*} |
for all \varphi, \zeta, \nu, \varPhi\in \mathcal{U} , where 0\leq\lambda = \frac{1}{4} < \frac{1}{2} . Hence, all the hypotheses of Corollary 3.4 are verified, and consequently, the integral equation (4.4) has a unique solution.
In this paper, we proved coupled fixed point theorems on a bicomplex partial metric space. An illustrative example and an application on a bicomplex partial metric space were given.
The authors declare no conflict of interest.
[1] |
K. Diethelm, N. J. Ford, Analysis of fractional differential equations, J. Math. Anal. Appl., 265 (2002), 229–248. https://doi.org/10.1006/jmaa.2000.7194 doi: 10.1006/jmaa.2000.7194
![]() |
[2] |
M. Eckert, M. Kupper, S. Hohmann, Functional fractional calculus for system identification of battery cells, Automatisierungstechnik, 62 (2014), 272–281. https://doi.org/10.1515/auto-2014-1083 doi: 10.1515/auto-2014-1083
![]() |
[3] | A. A. Kilbas, H. M. Srivastava, J. J. Trujillo, Theory and applications of fractional differential equations, Elsevier, 2006. |
[4] | K. S. Miller, B. Ross, An introduction to the fractional calculus and fractional differential equations, John Wiley & Sons, 1993. |
[5] |
M. N. Alam, I. Talib, O. Bazighifan, D. N. Chalishajar, B. Almarri, An analytical technique implemented in the fractional clannish random Walker's parabolic equation with nonlinear physical phenomena, Mathematics, 9 (2021), 801. https://doi.org/10.3390/math9080801 doi: 10.3390/math9080801
![]() |
[6] | K. B. Oldham, J. Spanier, The fractional calculus, Academic Press, 1974. |
[7] |
H. Zhang, X. Jiang, X. Yang, A time-space spectral method for the time-space fractional Fokker-Planck equation and its inverse problem, Appl. Math. Comput., 320 (2018), 302–318. https://doi.org/10.1016/j.amc.2017.09.040 doi: 10.1016/j.amc.2017.09.040
![]() |
[8] |
H. Zhang, X. Jiang, F. Zeng, G. E. Karniadakis, A stabilized semi-implicit Fourier spectral method for nonlinear space-fractional reaction-diffusion equations, J. Comput. Phys., 405 (2020), 109141. https://doi.org/10.1016/j.jcp.2019.109141 doi: 10.1016/j.jcp.2019.109141
![]() |
[9] |
X. Zheng, H. Wang, Optimal-order error estimates of finite element approximations to variable-order time-fractional diffusion equations without regularity assumptions of the true solutions, IMA J. Numer. Anal., 41 (2021), 1522–1545. https://doi.org/10.1093/imanum/draa013 doi: 10.1093/imanum/draa013
![]() |
[10] |
S. Shateyi, S. S. Motsa, Y. Khan, A new piecewise spectral homotopy analysis of the Michaelis-Menten enzymatic reactions model, Numer. Algorithms, 66 (2014), 495–510. https://doi.org/10.1007/s11075-013-9745-4 doi: 10.1007/s11075-013-9745-4
![]() |
[11] |
I. M. Abu-Reesh, Optimal design of continuously stirred membrane reactors in series using Michaelis-Menten kinetics with competitive product inhibition: theoretical analysis, Desalination, 180 (2005), 119–132. https://doi.org/10.1016/j.desal.2004.12.033 doi: 10.1016/j.desal.2004.12.033
![]() |
[12] |
M. Goličnik, Explicit reformulations of time-dependent solution for a Michaelis-Menten enzyme reaction model, Anal. Biochem., 406 (2010), 94–96. https://doi.org/10.1016/j.ab.2010.06.041 doi: 10.1016/j.ab.2010.06.041
![]() |
[13] |
H. Alrabaiah, A. Ali, F. Haq, K. Shah, Existence of fractional order semianalytical results for enzyme kinetics model, Adv. Differ. Equations, 2020 (2020), 443. https://doi.org/10.1186/s13662-020-02897-2 doi: 10.1186/s13662-020-02897-2
![]() |
[14] |
M. Alqhtani, J. F. Gómez-Aguilar, K. M. Saad, Z. Sabir, E. Pérez-Careta, A scale conjugate neural network learning process for the nonlinear malaria disease model, AIMS Math., 8 (2023), 21106–21122. https://doi.org/10.3934/math.20231075 doi: 10.3934/math.20231075
![]() |
[15] |
S. Liao, On the homotopy analysis method for nonlinear problems, Appl. Math. Comput., 147 (2004), 499–513. https://doi.org/10.1016/S0096-3003(02)00790-7 doi: 10.1016/S0096-3003(02)00790-7
![]() |
[16] |
X. C. Shi, L. L. Huang, Y. Zeng, Fast Adomian decomposition method for the Cauchy problem of the time-fractional reaction diffusion equation, Adv. Mech. Eng., 8 (2016), 1687814016629898. https://doi.org/10.1177/16878140166298 doi: 10.1177/16878140166298
![]() |
[17] |
H. M. Srivastava, K. M. Saad, Some new and modified fractional analysis of the time-fractional Drinfeld-Sokolov-Wilson system, Chaos, 30 (2020), 113104. https://doi.org/10.1063/5.0009646 doi: 10.1063/5.0009646
![]() |
[18] |
Y. Takeuchi, Y. Yoshimoto, R. Suda, Second order accuracy finite difference methods for space-fractional partial differential equations, J. Comput. Appl. Math., 320 (2017), 101–119. https://doi.org/10.1016/j.cam.2017.01.013 doi: 10.1016/j.cam.2017.01.013
![]() |
[19] |
K. M. Saad, H. M. Srivastava, J. F. Gómez-Aguilar, A fractional quadratic autocatalysis associated with chemical clock reactions involving linear inhibition, Chaos Solitons Fract., 132 (2020), 109557. https://doi.org/10.1016/j.chaos.2019.109557 doi: 10.1016/j.chaos.2019.109557
![]() |
[20] |
A. Atangana, Fractal-fractional differentiation and integration: connecting fractal calculus and fractional calculus to predict complex system, Chaos Solitons Fract., 102 (2017), 396–406. https://doi.org/10.1016/j.chaos.2017.04.027 doi: 10.1016/j.chaos.2017.04.027
![]() |
[21] |
M. H. Heydari, A. Atangana, Z. Avazzadeh, Y. Yang, Numerical treatment of the strongly coupled nonlinear fractal-fractional Schrödinger equations through the shifted Chebyshev cardinal functions, Alex. Eng. J., 59 (2020), 2037–2052. https://doi.org/10.1016/j.aej.2019.12.039 doi: 10.1016/j.aej.2019.12.039
![]() |
[22] |
M. Alqhtani, K. M. Saad, Fractal-fractional Michaelis-menten enzymatic reaction model via different kernels, Fractal Fract., 6 (2022), 13. https://doi.org/10.3390/fractalfract6010013 doi: 10.3390/fractalfract6010013
![]() |
[23] |
M. M. Alsuyuti, E. H. Doha, S. S. Ezz-Eldien, Galerkin operational approach for multi-dimensions fractional differential equations, Commun. Nonlinear Sci. Numer. Simul., 114 (2022), 106608. https://doi.org/10.1016/j.cnsns.2022.106608 doi: 10.1016/j.cnsns.2022.106608
![]() |
[24] |
M. M. Alsuyuti, E. H. Doha, S. S. Ezz-Eldien, I. K. Youssef, Spectral Galerkin schemes for a class of multi-order fractional pantograph equations, J. Comput. Appl. Math., 384 (2021), 113157. https://doi.org/10.1016/j.cam.2020.113157 doi: 10.1016/j.cam.2020.113157
![]() |
[25] |
A. H. Bhrawy, M. A. Abdelkawy, S. S. Ezz-Eldien, Efficient spectral collocation algorithm for a two-sided space fractional Boussinesq equation with non-local conditions, Mediterr. J. Math., 13 (2016), 2483–2506. https://doi.org/10.1007/s00009-015-0635-y doi: 10.1007/s00009-015-0635-y
![]() |
[26] | L. Michaelis, M. L. Menten, Die kinetik der invertinwirkung, Biochem. Z., 49 (1913), 333–369. |
[27] | I. Podlubny, Fractional differential equations, Elsevier, 1998. |
[28] | S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional integrals and derivatives: theory and applications, Gordon and Breach Science Publishers, 1993. |
[29] |
E. H. Doha, A. H. Bhrawy, D. Baleanu, S. S. Ezz-Eldien, The operational matrix formulation of the Jacobi tau approximation for space fractional diffusion equation, Adv. Differ. Equations, 2014 (2014), 231. https://doi.org/10.1186/1687-1847-2014-231 doi: 10.1186/1687-1847-2014-231
![]() |
[30] |
A. Ahmadian, M. Suleiman, S. Salahshour, D. Baleanu, A Jacobi operational matrix for solving a fuzzy linear fractional differential equation, Adv. Differ. Equations, 2013 (2013), 104. https://doi.org/10.1186/1687-1847-2013-104 doi: 10.1186/1687-1847-2013-104
![]() |
[31] | A. H. Bhrawy, M. M. Tharwat, M. A. Alghamdi, A new operational matrix of fractional integration for shifted Jacobi polynomials, Bull. Malays. Math. Sci. Soc., 37 (2013), 983–995. |
[32] |
Z. A. Noor, I. Talib, T. Abdeljawad, M. A. Alqudah, Numerical study of Caputo fractional-order differential equations by developing new operational matrices of Vieta-Lucas polynomials, Fractal Fract., 6 (2022), 79. https://doi.org/10.3390/fractalfract6020079 doi: 10.3390/fractalfract6020079
![]() |
[33] |
H. Singh, Approximate solution of fractional vibration equation using Jacobi polynomials, Appl. Math. Comput., 317 (2018), 85–100. https://doi.org/10.1016/j.amc.2017.08.057 doi: 10.1016/j.amc.2017.08.057
![]() |
[34] |
C. S. Singh, H. Singh, V. K. Singh, O. P. Singh, Fractional order operational matrix methods for fractional singular integro-differential equation, Appl. Math. Modell., 40 (2016), 10705–10718. https://doi.org/10.1016/j.apm.2016.08.011 doi: 10.1016/j.apm.2016.08.011
![]() |
[35] |
H. Singh, A new numerical algorithm for fractional model of Bloch equation in nuclear magnetic resonance, Alex. Eng. J., 55 (2016), 2863–2869. https://doi.org/10.1016/j.aej.2016.06.032 doi: 10.1016/j.aej.2016.06.032
![]() |
[36] | T. J. Rivlin, An introduction to the approximation of functions, Courier Corporation, 1981. |
[37] | E. Kreyszig, Introductory functional analysis with applications, John Wiley & Sons, 1991 |
[38] |
M. Behroozifar, A. Sazmand, An approximate solution based on Jacobi polynomials for time-fractional convection-diffusion equation, Appl. Math. Comput., 296 (2017), 1–17. https://doi.org/10.1016/j.amc.2016.09.028 doi: 10.1016/j.amc.2016.09.028
![]() |
[39] |
S. S. Ezz-Eldien, A. A. El-Kalaawy, Numerical simulation and convergence analysis of fractional optimization problems with right-sided Caputo fractional derivative, J. Comput. Nonlinear Dyn., 13 (2017), 011010. https://doi.org/10.1115/1.4037597 doi: 10.1115/1.4037597
![]() |
[40] |
H. Singh, H. M. Srivastava, Numerical investigation of the fractional-order Liénard and Duffing equations arising in oscillating circuit theory, Front. Phys., 8 (2020), 120. https://doi.org/10.3389/fphy.2020.00120 doi: 10.3389/fphy.2020.00120
![]() |
[41] |
S. S. Ezz-Eldien, New quadrature approach based on operational matrix for solving a class of fractional variational problems, J. Comput. Phys., 317 (2016), 362–381. https://doi.org/10.1016/j.jcp.2016.04.045 doi: 10.1016/j.jcp.2016.04.045
![]() |
1. | Sunisa Theswan, Sotiris K. Ntouyas, Bashir Ahmad, Jessada Tariboon, Existence Results for Nonlinear Coupled Hilfer Fractional Differential Equations with Nonlocal Riemann–Liouville and Hadamard-Type Iterated Integral Boundary Conditions, 2022, 14, 2073-8994, 1948, 10.3390/sym14091948 |