Hostname: page-component-7bb8b95d7b-fmk2r Total loading time: 0 Render date: 2024-09-28T17:02:30.089Z Has data issue: false hasContentIssue false

A strongly convergent algorithm for solving multiple set split equality equilibrium and fixed point problems in Banach spaces

Published online by Cambridge University Press:  15 June 2023

E.C. Godwin
Affiliation:
School of Mathematics, Statistics and Computer Science, University of KwaZulu-Natal, Durban, South Africa ([email protected]; [email protected]; [email protected]; [email protected]; [email protected])
O.T. Mewomo
Affiliation:
School of Mathematics, Statistics and Computer Science, University of KwaZulu-Natal, Durban, South Africa ([email protected]; [email protected]; [email protected]; [email protected]; [email protected])
T.O. Alakoya
Affiliation:
School of Mathematics, Statistics and Computer Science, University of KwaZulu-Natal, Durban, South Africa ([email protected]; [email protected]; [email protected]; [email protected]; [email protected])
Rights & Permissions [Opens in a new window]

Abstract

In this article, using an Halpern extragradient method, we study a new iterative scheme for finding a common element of the set of solutions of multiple set split equality equilibrium problems consisting of pseudomonotone bifunctions and the set of fixed points for two finite families of Bregman quasi-nonexpansive mappings in the framework of p-uniformly convex Banach spaces, which are also uniformly smooth. For this purpose, we design an algorithm so that it does not depend on prior estimates of the Lipschitz-type constants for the pseudomonotone bifunctions. Furthermore, we present an application of our study for finding a common element of the set of solutions of multiple set split equality variational inequality problems and fixed point sets for two finite families of Bregman quasi-nonexpansive mappings. Finally, we conclude with two numerical experiments to support our proposed algorithm.

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on Behalf of The Edinburgh Mathematical Society.

1. Introduction

Let E 1 and E 2 be two real Banach spaces with duals $E_{1}^{*}$ and $E_{2}^{*},$ respectively, and let C and Q be nonempty closed and convex subsets of E 1 and $E_2,$ respectively. Let $A:E_{1}\to E_{2}$ be a bounded linear operator. Censor and Elving [Reference Censor and Elfving7] introduced the concept of split feasibility problem (SFP), which is formulated as

(1.1)\begin{align} \mbox{find } x^* \in C \quad \mbox{such that} \ Ax^* \in Q. \end{align}

The SFP has been found useful in solving numerous real-life problems, including medical image reconstruction, phase retrieval, signal processing, radiation therapy treatment planning, among others. See, for example [Reference Godwin, Izuchukwu and Mewomo17, Reference Godwin, Izuchukwu and Mewomo18, Reference Kassay, Reich and Sabach24] and the references therein.

For solving the SFP (1.1) using Bregman projection, Schöpfer et al. [Reference Schöpfer, Schuster and Louis40] proposed the following method in the framework of p-uniformly convex real Banach spaces: for $x_1 \in E_1$, set

(1.2)\begin{align} x_{n+1}=\Pi_{C}J_{q}^{E_1^*}\left[J_{E_1}^{p}(x_n)-\gamma_{n}A^*J_{E_2}^{p}\left(Ax_{n}-P_{Q}(Ax_n)\right)\right], \quad n\geq 1, \end{align}

where $\Pi_{C}$ denotes the Bregman projection from E 1 onto C and $J_{E}^{p}$ is the duality mapping.

Closely related to the SFP (1.1) is the following split variational inequality problem (SVIP) introduced by Censor et al. [Reference Censor, Gibali and Reich9] in the framework of real Hilbert spaces as follows: find $x^* \in C$, which satisfies the inequality

(1.3)\begin{align} \langle F(x^*), x-x^*\rangle \geq0 \quad \forall x\in C, \end{align}

such that $y^*=Ax^* \in Q$ solves the inequality

(1.4)\begin{align} \langle G(y^*), y-y^*\rangle \geq 0 \quad \forall y \in Q, \end{align}

where C and Q are nonempty, closed and convex subsets of real Hilbert spaces H 1 and $H_2,$ respectively, $F:H_1 \to H_1$ and $G:H_2 \to H_2$ are two given operators, $A:H_1 \to H_2$ is a bounded linear operator.

Recall that when problems (1.3) and (1.4) are viewed separately, then Equation (1.3) is the classical variational inequality problem (VIP) in H 1 with its solution set ${\rm VIP}(C, F)$ and Equation (1.4) is another VIP in H 2 with its solution set ${\rm VIP}(Q, G).$ To solve the SVIP (1.3) and (1.4), Censor et al. [Reference Censor, Gibali and Reich9] put forward the following algorithm. Let $x_1 \in H_1,$ the sequence $\{x_n\}$ is generated by

(1.5)\begin{align} x_{n+1}= P_C(I-\lambda U)(x_n + \gamma A^*(P_Q(I-\lambda V)-I)Ax_n), \quad n\geq1, \end{align}

where $\gamma \in (0, {1}/{L})$ and L is the spectral radius of the operator $A^*A$. They proved that the above Algorithm (1.5) converges weakly to a solution of the SVIP under the assumption that $U, V$ are $\alpha_1,\alpha_2$-inverse strongly monotone operators and $\lambda\in(0,2\alpha)$ (where $\alpha:=\min\{\alpha_1,\alpha_2\}$).

Let C be a nonempty, closed and convex subset of a real Banach space E with dual $E^*$. Let $f:C\times C\to \mathbb{R}$ be a bifunction. The equilibrium problem (EP) studied by Blum and Oettli [Reference Blum and Oettli5] is to locate a point $x^*\in C$ such that

(1.6)\begin{align} f(x^*, x)\geq 0, \quad \forall ~x\in C. \end{align}

We denote by EP(f) the set of solutions of Equation (1.6). The EP was formerly introduced as the Ky Fan inequality [Reference Fan and Shisha15]. This class of problem has been extensively studied by numerous scholars because of its several applications. It is well known that many problems arising in economics, optimization and physics can be reduced to problem (1.6). Moreover, several iterative algorithms have been proposed to solve the EP (1.6) and related optimization problems in both Hilbert and Banach spaces (see [Reference Alakoya and Mewomo1, Reference Alakoya, Uzor and Mewomo3, Reference Blum and Oettli5, Reference Combettes and Hilstoaga13, Reference Godwin, Alakoya, Mewomo and Yao16, Reference Godwin, Izuchukwu and Mewomo17, Reference Ma, Wang and Cho28, Reference Moudafi31, Reference Owolabi, Alakoya, Taiwo and Mewomo36, Reference Ogwo, Izuchukwu and Mewomo34, Reference Reich and Sabach39, Reference Uzor, Alakoya and Mewomo45] and other references contained therein). However, most of the existing results on the EP are of the monotone type.

In 2011, Moudafi [Reference Moudafi30] extended the SVIP Equations (1.3) and (1.4) to split equilibrium problem (shortly, SEP), which is defined as

(1.7)\begin{align} \text{find }x^* \in C \quad \text{such that }f_1(x^*,y)\geq 0, \quad \forall~y \in C\nonumber\\ \text{and }y^*=Ax^* \in Q \quad \text{such that }f_2(y^*, z)\geq 0, \quad \forall~z \in Q, \end{align}

where C and Q are nonempty, closed and convex subsets of real Hilbert spaces H 1 and $H_2,$ respectively, $f_1:C \times C \rightarrow \mathbb{R}$ and $f_2:Q \times Q \rightarrow \mathbb{R}$ are bifunctions with a bounded linear operator $A: H_1 \rightarrow H_2$.

In addition, for solving the SEP (1.7), the author in [Reference He20] put forward a proximal technique without product space formulation as follows:

\begin{align*} &x^{0}\in C; \quad \{\rho_n\}\subset (0, \infty); \quad \mu \gt 0\\ &f_{1}(y_n, y)+\frac{1}{\rho_n}\langle y-y_n, y_n-x_n\rangle \geq 0 \quad \forall y\in C,\\ &f_{2}(u_n, v)+\frac{1}{\rho_n}\langle v-u_n, u_n-Ay_n\rangle \geq 0 \quad \forall v\in Q,\\ &x_{n+1}=P_{C}(y_n+ \mu A^*(u_n-Ay_n)),\quad \forall n\geq 0, \end{align*}

where $A^*$ is the adjoint of A. The author in [Reference He20] obtained a weak convergence result when the bifunctions f 1 and f 2 are monotone on C and Q, respectively. Since then, several iterative schemes have been proposed when the bifunctions are either monotone or pseudomonotone, see for example [Reference Anh4, Reference Eskandani, Raeisi and Rassias14, Reference Hieu21Reference Hieu23, Reference Kim and Dinh25] and check also the references therein.

In Section 6.1 of Censor et al. [Reference Censor, Gibali and Reich9], the authors proposed an improvement of the SVIP (1.3) and (1.4), which they called multiple set SVIP (MSSVIP), which is formulated as follows:

\begin{align*} \text{find }x^*\in C: & =\bigcap_{i=1}^{N}C_{i} \quad \text{such that }\langle F_{i}(x^*), y-x^*\rangle \geq 0, \quad \forall y\in C_{i},\\ & \qquad i=1,2,\ldots,N\\ \text{and such that }y^*=Ax^*\in Q: & =\bigcap_{j=1}^{M} Q_{j} \ \text{solves} \ \langle G_{j}(y^*), z-y^*\rangle \geq 0,\quad \forall z\in Q_{j},\\ & \qquad j=1,2,\ldots,M, \end{align*}

where $A:H_{1}\to H_{2}$ is a bounded linear operator, $F_{i}:H_1\to H_1$, $i=1,2,\ldots,N$ and $G_j:H_2\to H_2$, $j=1,2,\ldots,M$ are given operators and $C_i, \ i=1,2,\ldots,N$ and $Q_j,\ j=1,2,\ldots,M$ are nonempty, closed and convex subsets of real Hilbert spaces H 1 and $H_2,$ respectively. They proposed an algorithm for solving the problem and proved that the sequence generated by the proposed iterative scheme converges weakly to the solution set of MSSVIP when $F_{i},\ i=1,2,\ldots,N$ and $G_j, \ j=1,2,\ldots,M$ are inverse strongly monotone operators.

Moudafi [Reference Moudafi31] proposed a new SFP, which he called split equality problem (SEqP). Let H 1, H 2 and H 3 be real Hilbert spaces. $A: H_1\rightarrow H_3$ and $B:H_2 \rightarrow H_3$ be two bounded linear operators, and $C \subset H_1$ and $Q \subset H_2$ be two nonempty, closed and convex sets. The SEqP is formulated as follows:

(1.8)\begin{align} \text{find }x^* \in C, \qquad y^* \in Q \quad \text{such that } Ax^*=By^*. \end{align}

Let $T:E\to E$ be a mapping. We denote the fixed point set of T by $\mbox{F}(T)$; that is $\mbox{F}(T):=\{x\in E: Tx=x\}$. The Fixed Point Problem has application in various fields, such as optimization theory, economics, game theory, as well as in establishing the existence of solutions of several physical problems arising in differential and integral equations [Reference Godwin, Taiwo and Mewomo19, Reference Ogbuisi and Mewomo32, Reference Ogwo, Izuchukwu and Mewomo34, Reference Taiwo and Mewomo42].

If $C:=F(S)$ and $Q:=F(T)$ in Equation (1.8), where $S:H_1\rightarrow H_1$ and $T:H_2\rightarrow H_2$ are two nonlinear mappings, then the SEqP becomes the split equality fixed point problem.

Motivated and inspired by the above mentioned results, we introduce and study in the framework of p-uniformly convex Banach space an extension of the MSSVIP to multiple set split equality equilibrium and common fixed points problem of Bregman quasi-nonexpansive mappings. Using Bregman distance, we make use of the Halpern extragradient technique for solving the pseudomonotone EP, which guarantees strong convergence. We design our algorithm in such a way that it does not depend on the prior estimates of the Lipschitz-like constants.

We organize the rest of this article as follows: Section 2 presents preliminaries and some existing results, $\S$ 3 is the design of our iterative method, whereas $\S$ 4 focuses on the convergence analysis of the proposed algorithm. In $\S$ 5, we apply our result to solve a certain class of variational inequality problems. We present some numerical experiments in $\S$ 6 and conclude with some final remarks in $\S$ 7.

2. Preliminaries

In this section, we call up some important definitions and existing results, which will be needed in the proof of our main result. We denote strong and weak convergence of the sequence $\{x_{n}\}$ to a point x by ‘$\rightarrow$’ and ‘$\rightharpoonup$’, respectively.

Let E be a real Banach space and $1 \lt q\leq 2 \lt p \lt \infty$ with $\frac{1}{p}+\frac{1}{q}=1.$ The modulus of smoothness of E is the function $\rho_E: \mathbb{R}^{+}:=[0, \infty)\rightarrow \mathbb{R}^{+}$ defined by

\begin{align*} \rho_E(\tau)=\sup\left\{\frac{\|\bar{x}+\tau \bar{y}\| +\|\bar{x}-\tau \bar{y}\|}{2}-1: \bar{x}=1=\bar{y}\right\}. \end{align*}

The space E is called uniformly smooth if and only if $\frac{\rho_E(\tau)}{\tau}\rightarrow 0$ as $\tau \rightarrow 0$. Let q > 1, E is said to be q-uniformly smooth if there exists $\kappa_q \gt 0$ such that $\rho_E(\tau) \leq \kappa_q \tau^{q}$ for all τ > 0. The modulus of convexity of E is defined as

\begin{align*} \beta_{E}(\epsilon)=\inf \left\{1-\frac{\|\bar{x}+\bar{y}\|}{2}:\|\bar{x}\|=\|\bar{y}\|=1;\quad \epsilon=\|\bar{x}-\bar{y}\|\right\}. \end{align*}

The Banach space E is called uniformly convex if and only if $\beta_E(\epsilon) \gt 0$ for every $\epsilon \in (0,2]$. Now, suppose p > 1, then E is called p-uniformly convex if there exists a constant $C_p \gt 0$ such that $\beta_E(\epsilon) \geq C_p\epsilon^{p}$ for all $\epsilon \in (0,2]$.

Remark 2.1. It is well known that every p-uniformly convex space is also strictly convex and reflexive. In addition, if a Banach space E is p-uniformly convex and uniformly smooth, then its dual space $E^*$ is q-uniformly smooth and uniformly convex (see [Reference Cholamjiak and Sunthrayuth11]).

Definition 2.2. see [Reference Cioranescu12]

Let p > 1 be a real number, the generalized duality mapping $J_{E}^{p}:E \rightarrow 2^{E^*}$ is defined by

\begin{align*} J_{E}^{p}(x)=\{x^*\in E^*: \langle x, x^*\rangle=\|x\|^p, \|x^*\|=\|x\|^{p-1}\}, \end{align*}

where $\langle.,.\rangle$ denotes the duality pairing between elements of E and $E^*$. In particular, $J_E^{p}=J_E^{2}$ is called the normalized duality mapping. If E is p-uniformly convex and uniformly smooth, then $E^*$ is q-uniformly smooth and uniformly convex. In this case, the generalized duality mapping $J_{E}^{p}$ is one-to-one, single-valued and satisfies $J_E^{p}=(J_{E^*}^{q})^{-1}$, where $J_{E^*}^{q}$ is the generalized duality mapping of $E^*$. Furthermore, if E is uniformly smooth, then the duality mapping $J_{E}^{p}$ is norm-to-norm uniformly continuous on bounded subsets of E, and E is smooth if and only if $J_{E}^{p}$ is single valued.

Let $f: E \rightarrow (-\infty, + \infty]$ be a proper, lower semicontinuous and convex function, then the Fenchel conjugate of f denoted as $f^*: E^* \rightarrow (-\infty,+ \infty]$ is defined as

\begin{align*} f^*(x^*)=\sup\{\langle x^*, x\rangle-f(x): x \in E,\ x^*\in E^*\}. \end{align*}

See [Reference Simons and Reich41] for more information about Fenchel conjugate.

Let the domain of f be denoted by $({\rm dom}\, f)=\{x \in E: f(x) \lt +\infty\}$; hence, for any $x \in {\rm int}({\rm dom}\, f)$ and $y \in E$, we define the right-hand derivative of f at x in the direction y by

\begin{align*} f^{0}(x, y)=\lim_{t \rightarrow 0^+}\frac{f(x+ty)-f(x)}{t}. \end{align*}

The function f is said to be Gâteaux differentiable at x if $\lim_{t \rightarrow 0^+}\frac{f(x+ty)-f(x)}{t}$ exists for any $y.$ In this case, $f^0(x,y)$ coincides with $\bigtriangledown f(x)$ (the value of the gradient $\bigtriangledown f$ of f at x). The function f is said to be Gâteaux differentiable if it is Gâteaux differentiable for any $x \in {\rm int}({\rm dom}\, f)$. The function f is said to be Fréchet differentiable at x if its limit is attained uniformly in $||y||=1$. Moreover, f is said to be uniformly Fréchet differentiable on a subset C of E if the above limit is attained uniformly for $x \in C$ and $||y||=1$.

A function f is said to be Legendre if it satisfies the following conditions:

  1. (1) The interior of the domain of f, ${\rm int}({\rm dom}\, f)$ is nonempty, f is Gâteaux differentiable on ${\rm int}({\rm dom}\, f)$ and ${\rm dom}\bigtriangledown f={\rm int}({\rm dom}\, f)$.

  2. (2) The interior of the domain of $f^*$, ${\rm int}({\rm dom}\, f^*)$ is nonempty, $f^*$ is Gâteaux differentiable on ${\rm int}({\rm dom}\, f^*)$ and ${\rm dom} \bigtriangledown f^*={\rm int} ({\rm dom}\, f)$.

Definition 2.3 [Reference Bregman6] Let $f:E \rightarrow (-\infty,+ \infty]$ be a convex and Gâteaux differentiable function. The function $\Delta_f: E \times E \rightarrow [0, + \infty)$ defined by

\begin{align*} \Delta_f(x, y):=f(y)-f(x)- \langle \bigtriangledown f(x), y-x\rangle \end{align*}

is called the Bregman distance with respect to f.

We highlight the following interesting properties of Bregman distance (see [Reference Bregman6, Reference Reem, Reich and De Pierro37]):

  • (i) $\Delta_{f}(x,x)=0,$ but $\Delta_{f}(x,y)=0$ does not necessarily imply that x = y,

  • (ii) for $x\in {\rm dom}\, f$ and $y, z\in {\rm int}({\rm dom}\, f),$ we have

    (2.1)\begin{align} \Delta_{f}(x,y)+\Delta_{f}(y,z)-\Delta_{f}(x,z)= \langle \bigtriangledown g(z)-\bigtriangledown g(y), x-y\rangle, \end{align}
  • (iii) for each $z\in E,$ $\{x_i\}_{i=1}^{N}\subset E$ and $\{\alpha_n\}_{i=1}^{N} \subset (0,1)$ with $\sum\limits_{i=1}^{N}\alpha_i=1,$ we have

    \begin{equation*}\Delta_f\left(z,\nabla g^{*}\left(\sum\limits_{i=1}^{N}\alpha_i\nabla g(x_i) \right)\right) \leq \sum\limits_{i=1}^{N}\alpha_i\Delta_f(z,x_i). \end{equation*}

It is well-known that in general the Bregman distance $\Delta_f$ is not a metric because it fails to satisfy the symmetric and triangle inequality properties. Moreover, it is well known that the duality mapping $J_{E}^{p}$ is the sub-differential of the functional $f_p(.)= \frac{1}{p}||.||^{p}$ for p > 1, see [Reference Chidume10]. Then, the Bregman distance $\Delta_p$ is defined with respect to fp as follows:

(2.2)\begin{align} \Delta_p(x,y)& = \frac{1}{p}\|y\|^p-\frac{1}{p}\|x\|^p-\langle J_{E}^{p}x, y-x \rangle\nonumber\\ &=\frac{1}{q}\|x\|^p-\langle J_{E}^{p}x, y \rangle + \frac{1}{p}\|y\|^p\nonumber\\ &=\frac{1}{q}\|x\|^p-\frac{1}{q}\|y\|^p-\langle J_{E}^{p}x-J_{E}^{p}y, y \rangle. \end{align}

Definition 2.4. Let C be a nonempty, closed and convex subset of a real Banach space E and let $T:C\to C$ be a nonlinear map. The mapping IT is said to be demiclosed at zero if for any sequence $\{x_n\}\subset C,$ the following implication holds: $x_{n}\rightharpoonup x$ and $(I-T)x_n\to 0\implies x\in F(T).$

Definition 2.5. [Reference Censor and Reich8, Reference Martin-Marquez, Reich and Sabach29]

Let $T: C \rightarrow {\rm int}({\rm dom}f)$ be a mapping. Then,

  1. (i) a point $p \in C$ is called an asymptotic fixed point of T if C contains a sequence $\{x_n\}$, which converges weakly to p such that $\lim_{n \rightarrow \infty}\|Tx_n-x_n\|=0$. We denote by $\hat{F}(T)$ the set of asymptotic fixed points of T;

  2. (ii) T is called Bregman firmly nonexpansive if

    \begin{align*} \langle \Delta_f (Tx)-\Delta_f (Ty), Tx-Ty \rangle\leq \langle\Delta_{f}(x)-\Delta_{f}(y), Tx-Ty\rangle, \quad \forall x,y\in C. \end{align*}
  3. (iii) T is called Bregman strongly nonexpansive if $\hat{F}(T)\ne\emptyset$ and

    \begin{align*} \Delta_{f}(p, Tx)\leq \Delta_{f} (p, x), \qquad \forall p\in \hat{F}(T) \quad \mbox{and} \quad x\in C. \end{align*}
  4. (iv) T is said to be Bregman quasi-nonexpansive if

    \begin{align*} F(T) \neq \emptyset\quad \text {and} \quad \Delta_f(p, Tx) \leq \Delta_f(p, x),\quad \forall x \in C,\ p \in F(T). \end{align*}

Recall that the metric projection PC from E onto C satisfies the following property:

\begin{align*} \|x-P_Cx\| \leq \inf_{y \in C}\|x-y\|,\quad \forall x \in E. \end{align*}

It is well known that PC is the unique minimizer of the norm distance. Moreover, PC is characterized by the following property:

(2.3)\begin{align} \langle J_{E}^{p}x-J_{E}^{p}(P_Cx), y-P_C x\rangle \leq 0, \quad \forall y \in C. \end{align}

The Bregman projection from E onto C denoted by $\Pi_{C}$ also satisfies the property

(2.4)\begin{align} \Delta_p(x, \Pi_C(x))=\inf_{y \in C}\Delta_p(x, y),\quad \forall x \in E. \end{align}

Also, if C is a nonempty, closed and convex subset of a p-uniformly convex and uniformly smooth Banach space E and $x \in E$, then the following assertions hold (see [Reference Cholamjiak and Sunthrayuth11]):

  1. (i) $z=\Pi_{C}x$ if and only if

    (2.5)\begin{align} \langle J_{E}^{p}(x)-J_{E}^{p}(z), y-z\rangle \leq 0, \quad \forall y \in C; \end{align}
  2. (ii)

    (2.6)\begin{align} \Delta_p(\Pi_{C}x, y)+ \Delta_p(x, \Pi_Cx)\leq \Delta_p(x,y),\quad \forall y \in C. \end{align}

Lemma 2.6. [ Reference Tiel43] Let C be a nonempty convex subset of a Banach space E. Let $g:C\to \mathbb{R}$ be a convex, subdifferentiable function on C. Then g attains its minimum at $x\in C$ if and only if $0\in \partial g(x)+ N_{C}(x),$ where $N_{C}(x)$ is the normal cone of C at x, that is

\begin{align*} N_{C}(x):=\{\hat{x}\in E^{*}:\langle x-\varphi, \hat{x}\rangle\geq 0, \quad \forall \varphi\in C\}. \end{align*}

Lemma 2.7. [ Reference Chidume10] Let E be a Banach space and $x, y \in E$. If E is q-uniformly smooth, then there exists $C_q \gt 0$ such that

\begin{align*} \|x-y\|^q \leq \|x\|^q - q\langle J_{q}^{E}(x),y \rangle + C_{q}\|y\|^q. \end{align*}

Lemma 2.8. [ Reference Cioranescu12] Let f and g be two convex functions on E, such that $x_{0}\in {\rm dom}\, f\cap {\rm dom} g$, where f is continuous, then

\begin{align*} \partial (f+g)(x)=\partial f(x)+\partial g(x),\quad \forall x\in E. \end{align*}

Lemma 2.9. [ Reference Kuo and Sahu27] Let E be a real p-uniformly convex and uniformly smooth Banach space. Let $z, x_k \in E (k=1,2,\ldots,N)$ and $\alpha_k \in (0,1)$ with $\sum_{k=1}^{N}\alpha_k=1$. Then, we have

\begin{align*} \Delta_p\left(J_{q}^{E^*}\left(\sum_{k=1}^{N}\alpha_k J_{p}^{E}(x_k)\right), z\right) \leq \sum_{k=1}^{N}\alpha_k \Delta_p(x_k,z) - \alpha_i \alpha_jg_r^*\left(\|J_{p}^{E}(x_i)-J_{p}^{E}(x_j)\|\right), \end{align*}

for all $i, j \in \{1,2,\dots,N\}$ and $g^*_r:\mathbb{R}^+ \rightarrow \mathbb{R}^+$ being a strictly increasing function such that $g^*_r(0)=0$.

Lemma 2.10. [ Reference Schöpfer, Schuster and Louis40] Let E be a real p-uniformly convex and uniformly smooth Banach space. Let $V_p: E^* \times E \rightarrow [0, +\infty)$ be defined by

\begin{align*} V_p(x, x^*) = \frac{1}{p}\|x\|^p- \langle x, x^*\rangle + \frac{1}{q}\|x^*\|^q,\quad \forall x \in E,\ x^* \in E^*. \end{align*}

Then the following assertions hold:

  1. (i) Vp is nonnegative and convex in the first variable.

  2. (ii) $\Delta_p \left(x, J_{q}^{E^*}(x^*)\right)= V_p( x, x^*),\quad \forall x \in E,\ x^* \in E^*$.

  3. (iii) $V_p( x, x^*) + \langle J_{q}^{E^*}(x^*)-x, y^*\rangle \leq V_p(x, x^*+y^*),\quad \forall x \in E,\ x^*, y^* \in E^*$.

Lemma 2.11. [ Reference Cholamjiak and Sunthrayuth11] Let E be a real p-uniformly convex and uniformly smooth Banach space. Suppose that $\{x_n\}$ and $\{y_n\}$ are bounded sequences in E. Then the following assertions are equivalent:

  1. (i) $\lim_{n\rightarrow \infty} \Delta_p(x_n, y_n)=0;$

  2. (ii) $\lim_{n\rightarrow \infty}||x_n-y_n||=0$.

Lemma 2.12. [ Reference Xu and Roach46] Let $q\geq 1$ and r > 0 be two fixed real numbers. Then, a Banach space E is uniformly convex if and only if there exists a continuous, strictly increasing and convex function $g:\mathbb{R}^{+}\to \mathbb{R}^{*}$, $g(0)=0$ such that for all $x, y \in B_r$ and $0\leq\alpha \lt 1,$

\begin{align*} \|\alpha x + (1-\alpha)y\|^q\leq\alpha \|x\|^q +(1-\alpha)\|y\|^q-W_q(\alpha)g(\|x-y\|), \end{align*}

where $W_q(\alpha):=\alpha^q(1-\alpha)+\alpha(1-\alpha)^q$ and $B_r:=\{x\in E:\|x\|\leq r\}.$

Lemma 2.13. [ Reference Reich and Sabach38] Let E be a real Banach space and let $f:E\to \mathbb{R}$ be a Gâteaux differentiable and totally convex function. If $x_0\in E$ and the sequence $\{\Delta_f(x_n, x_0)\}$ is bounded, then the sequence $\{x_n\}$ is also bounded.

Lemma 2.14. [ Reference Kimura and Saejung26] Let $\{a_n\}\subset \mathbb{R}_{+}$, $\{\varepsilon_{n}\}\subset (0,1)$ be a sequence such that $\sum_{n=1}^{\infty}\varepsilon_{n}=\infty$ and $\{b_n\}\subset\mathbb{R}$. Assume that

\begin{align*} a_{n+1}\leq (1-\varepsilon_{n})a_n + \varepsilon_{n}b_n,\quad \forall n\geq 0. \end{align*}

If $\limsup\limits_{k\to\infty}b_{n_k}\le 0$ for every subsequence $\{a_{n_k}\}$ of $\{a_n\}$ satisfying the condition $\liminf\limits_{k\to \infty}(a_{n_k+1}-a_{n_k})\ge0,$ then $\lim\limits_{k\to \infty}a_n=0.$

To solve the EP (1.6), the following assumptions are needed:

Assumption A:

  1. (C1) f is pseudomonotone, that is, for all $x, y \in C,$ $f(x, y)\geq 0\Rightarrow f(y, x)\leq 0$ and $f(x, x)=0,$ for all $x\in C$.

  2. (C2) f satisfies the Bregman–Lipschitz type condition on C, that is, there exists two positive constants c 1 and c 2 such that

    \begin{align*} f(x,y)+ f(y,z)\geq f(x,z)-c_1\Delta_p(y, x)-c_2\Delta_p(y, z),\quad \forall x, y,z \in C, \end{align*}
    where $p:E\to (-\infty, +\infty]$ is a Legendre function. The constants c 1 and c 2 are called Bregman–Lipschitz coefficients with respect to p.
  3. (C3) $f(x, .)$ is convex, lower semicontinuous and subdifferentiable on C for all $x\in C$.

  4. (C4) f is jointly weakly continuous on C × C in the sense that if $x, y \in C$ and $\{x_n\}$ and $\{y_n\}$ converges weakly to x and y, respectively, then $f(x_n, y_n)\to f(x, y)$ as $n\to \infty$.

3. Proposed method

In this section, we present our method and discuss some of its features. We begin with the following assumptions under which our strong convergence result is obtained.

Assumption 3.1. We assume that the following conditions hold:

  • (1) (a) E 1, E 2 and E 3 are three p-uniformly convex and uniformly smooth real Banach spaces.

    (b) Ci and Qj are nonempty closed and convex subsets of E 1 and $E_2,$ respectively, for $i=1,2,\ldots,N$ and $j=1,2,\ldots,M.$

    (c) $A:E_{1}\to E_{3}$ and $B:E_2\to E_3$ are bounded linear operators.

    (d) $f_{i}:C_{i}\times C_{i}\to \mathbb{R}$ and $g_{j}:Q_{j}\times Q_{j}\to \mathbb{R}$ are bifunctions satisfying conditions $C_{1}-C_{4}$ of Assumption A.

    (e) $D_{s}:E_{1}\to E_{1}$ and $G_{t}:E_{2}\to E_{2}$ are Bregman quasi-nonexpansive mappings such that $I-D_s$ and $I-G_t$ are demiclosed at zero for each $s=1,2,\ldots,l$ and $t=1,2,\ldots,m.$

    (f) Assume that the solution set

    $\Upsilon:=\{\bar{x}\in \bigcap_{s=1}^l F(D_{s})\cap\bigcap_{i=1}^N \mbox{EP}(C_{i}, f_{i}),\quad \bar{y}\in \bigcap_{t=1}^mF(G_{t})\cap \bigcap_{j=1}^M\mbox{EP}(Q_{j}, g_{j}):A\bar{x}=B\bar{y}\}\neq 0$.

  • (2) $\{\beta_n\}^{\infty}_{n=1}$, $\{\alpha_{n, s}\}_{s=0}^{l}$, $\{\eta_{n, t}\}_{t=0}^{m}$ are positive sequences satisfying the following conditions:

    (a)$\{\beta_n\}\subset (0,1),\ \lim\limits_{n\to\infty} \beta_n=0,\ \sum_{n=1}^{\infty}\beta_n= \infty,\ \tau_0 \gt 0,\ \lambda_0 \gt 0,\ \kappa\in(0,1),\ \epsilon\in(0,1).$

    (b) $\{\alpha_{n,s}\}\subset (0,1),\ \sum_{s=0}^{l}\alpha_{n,s}=1 \ \mbox{and}\ \liminf\limits_{n\to\infty}\alpha_{n,0}\alpha_{n,s} \gt 0.$

    (c) $\{\eta_{n,t}\}\subset (0,1),\ \sum_{t=0}^{m}\eta_{n,t}=1 \ \mbox{and}\ \liminf\limits_{n\to\infty}\eta_{n,0}\eta_{n,t} \gt 0.$

We now present the proposed method of this paper.

Algorithm 3.2.

For fixed $\mu\in E_1$ and $\vartheta\in E_2$, choose an initial guess $(x_0, y_0)\in E_1\times E_2$. Suppose that the nth iterate $(x_n, y_n)\subset E_1\times E_2$ has been constructed; then we compute the $(n+1)$th iterate $(x_{n+1}, y_{n+1})$ via the iteration

(3.1)\begin{align} \begin{cases} s_n=J_{E_1^*}^{q}\left(J_{E_1}^{p}(x_n)-\rho_{n}A^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ a_{n}^{i}=\arg\min\left\{f_{i}(s_n, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n):\quad \sigma\in C_{i}\right\},\\ z_{n}^{i}=\arg\min\left\{f_{i}(a_{n}^{i}, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n):~~\sigma\in C_{i}\right\}.\\ \mbox{Obtain the farthest element of}\ z_{n}^{i}\ \mbox{from}\ s_n,\ \mbox{i.e.,}\\ i_{n}\in \arg\max \{\Delta_{p}(s_n, z_{n}^{i}):i=1,\ldots,N\}.\\ \mbox{Set}~ z_{n}^{i_n}=\bar{z}_{n} \\ u_n=J_{q}^{E_1^*}\left(\alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{s=1}^{l}\alpha_{n,s}J_{E_1}^{p}(D_s\bar{z}_n)\right)\\ x_{n+1}=J_{E_1^*}^{q}\left(\beta_{n}J_{E_1}^{p}(\mu)+(1-\beta_{n})J_{E_1}^{p}(u_n)\right),\\ t_n=J_{E_2^*}^{q}\left(J_{E_2}^{p}(y_n)+\rho_nB^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ b_{n}^{j}=\arg\min\left\{g_j(t_n, \varphi)+ \frac{1}{\lambda_n}\Delta_p(\varphi, t_n):\quad \varphi\in Q_{j}\right\},\\ h_{n}^{j}=\arg\min\left\{g_j(b_{n}^{j}, \varphi)+ \frac{1}{\lambda_n}\Delta_p(\varphi, t_n):\quad \varphi\in Q_{j}\right\}.\\ \mbox{Obtain the farthest element of}\ h_{n}^{j}\ \mbox{from}\ t_n,\ \mbox{i.e.,}\\ j_{n}\in \arg\max \left\{\Delta_{p}(t_n, h_{n}^{j}):j=1,\ldots,M\right\}.\\ \mbox{Set}\ h_{n}^{j_n}=\bar{\theta}_{n} \\ v_n=J_{E_1^*}^{q}\left(\eta_{n,0}J_{E_2}^{p}(\bar{\theta}_n)+\sum_{t=1}^{m}\eta_{n,t}J_{E_2}^{p}(G_t\bar{\theta}_n)\right)\\ y_{n+1}=J_{E_2^*}^{q}\left(\beta_{n}J_{E_2}^{p}(\vartheta)+(1-\beta_{n})J_{p}^{E_2}(v_n)\right), \end{cases} \end{align}

where

(3.2)\begin{align} \rho_{n}\in \left(\zeta, \left(\frac{q\|Ax_n-By_n\|^p}{C_q\|A^*J_{E_3}^{p}(Ax_n-By_n)\|^q+ Q_q\|B^*J_{E_3}^{p}(Ax_n-By_n)\|^q}-\zeta\right)^\frac{1}{q-1}\right),\quad n\in \Omega \end{align}

for small enough ζ; Cq and Qq are constants of smoothness of E1 and E2, respectively. Otherwise, $\rho_{n}=\rho$ (ρ being any nonnegative value), where the set of indexes $\Omega=\{n:Ax_n-By_n\neq 0\}$.

(3.3)\begin{align} \tau_{n+1}=\begin{cases} \min\left\{\tau_n, \min\limits _{1\leq i\leq N} \left\{\frac{\kappa(\Delta_p(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i}, a_{n}^{i}))}{f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})-f_{i}(a_{n}^{i},z_{n}^{i})}\right\}\right\},\quad &\mbox{if}~f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})\\ & \quad -f_{i}(a_{n}^{i},z_{n}^{i}) \gt 0,\\ \tau_n, \quad &\mbox{otherwise}.\end{cases}\end{align}

and

(3.4)\begin{align} \lambda_{n+1}=\begin{cases} \min\left\{\lambda_n, \min\limits _{1\leq j\leq M} \left\{\frac{\epsilon(\Delta_p(b_{n}^{j},t_n)+\Delta_{p}(h_{n}^{j}, b_{n}^{j}))}{g_{j}(t_n, h_{n}^{j})-g_{j}(t_n, b_{n}^{j})-g_{j}(b_{n}^{j},h_{n}^{j})}\right\}\right\},\quad &\mbox{if}~g_{j}(t_n, h_{n}^{j})-g_{j}(t_n, b_{n}^{j})\\ & \quad-g_{j}(b_{n}^{j},h_{n}^{j}) \gt 0,\\ \lambda_n, \quad &\mbox{otherwise}. \end{cases} \end{align}
Remark 3.3.

  1. (a) Algorithm 3.2 solves split equality EP consisting of two strongly convex optimization problems in parallel for $i=1,2,\ldots,N$, as well as another two strongly convex optimization problems in parallel for $j=1,2,\ldots,M$ under bounded linear operators.

  2. (b) The step size $\{\rho_n\}$ given by Equation (3.2) is generated at each iteration by some simple computations. Thus, $\{\rho_n\}$ is easily implemented without the prior knowledge of the operator norms $\|A\|$ and $\|B\|$. Similarly, the step size $\{\tau_n\}$ given by Equation (3.3) and step size $\{\lambda_n\}$ given by Equation (3.4) do not depend on the prior estimates of the Lipschitz-like constants of the pseudomonotone bifunctions $f_{i},~i=1,2,\ldots,N$, and $g_{j},~j=1,2,\ldots,M$, unlike the step sizes used in [Reference Eskandani, Raeisi and Rassias14, Reference Hieu, Muu and Anh22], which require finding the prior estimates of the Lipschitz-like constants of the pseudomonotone bifunctions, which is known to be computationally expensive.

  3. (c) Moreover, our result in this paper extends the results in [Reference Hieu, Muu and Anh22, Reference Kim and Dinh25] from the framework of Hilbert spaces to Banach spaces.

4. Convergence analysis

Lemma 4.1. The sequences $\{\tau_{n}\}$ and $\{\lambda_{n}\}$ of step sizes generated by Algorithm 3.2 are well defined and bounded.

Proof. Clearly, from Equations (3.3) and (3.4), we have $\tau_{n+1}\leq \tau_n\ \forall n\in \mathbb{N}$ and $\lambda_{n+1}\leq \lambda_n\ \forall n\in \mathbb{N}$. This implies that $\{\tau_n\}$ and $\{\lambda_n\}$ are monotonically decreasing sequences. Moreover, it follows from condition C 2 of Assumption A that

\begin{align*} f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})-f_{i}(a_{n}^{i},z_{n}^{i})\leq k_{1,i}\Delta_{p}(a_{n}^{i}, s_{n})+ k_{2,i}\Delta_{p}(z_{n}^{i},a_{n}^{i}),\quad \forall i=1,2,\ldots,N. \end{align*}

Hence, we obtain for all $i=1,2,\dots,N$

\begin{align*} \frac{\kappa\left(\Delta_{p}(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i}, a_{n}^{i})\right)}{f_i(s_n,z_{n}^{i})-f_i(s_n,a_{n}^{i})-f_i(a_{n}^{i},z_{n}^{i})}&\geq \frac{\kappa\left(\Delta_{p}(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i}, a_{n}^{i})\right)}{k_{1,i}\Delta_{p}(a_{n}^{i},s_n)+k_{2,i}\Delta_{p}(z_{n}^{i},a_{n}^{i})}\\ &\ge\frac{\kappa\left(\Delta_{p}(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i}, a_{n}^{i})\right)}{\max\{k_{1,i}, k_{2,i}\}(\Delta_{p}(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i},a_{n}^{i}))}\\ &\geq \frac{\kappa}{\max\{k_{1,i}, k_{2,i}\}}. \end{align*}

Similarly, we obtain

\begin{align*} \frac{\epsilon\left(\Delta_{p}(b_{n}^{j},t_n)+\Delta_{p}(h_{n}^{j}, b_{n}^{j})\right)}{g_j(t_n,h_{n}^{j})-g_j(t_n,b_{n}^{j})-g_j(b_{n}^{j},h_{n}^{j})}&\geq\frac{\epsilon}{\max\{c_{1,j}, c_{2,j}\}},\quad \forall j=1,2,\ldots,M. \end{align*}

Hence, we conclude that $\{\tau_n\}$ has lower bound $\min\left\{\tau_0, \frac{\kappa}{\max\limits_{1\le i\le N}\{k_{1,i},k_{2,i}\}}\right\} \gt 0$ and $\{\lambda_n\}$ has lower bound

$\min\left\{\lambda_0, \frac{\epsilon}{\max\limits_{1\le j\le M}\{c_{1,j},c_{2,j}\}}\right\} \gt 0.$ It then follows that $\lim\limits_{n\to \infty}\tau_n=\tau \gt 0$ and $\lim\limits_{n\to \infty}\lambda_n=\lambda \gt 0.$

Lemma 4.2. Let $C_i,\ i=1,2,\ldots,N$ and $Q_j, j=1,2,\ldots,M$ be nonempty, closed and convex subsets of E1 and E2, respectively. Suppose that $f_{i}:C_{i}\times C_{i}\to \mathbb{R},~i=1,2,\ldots, N$ and $g_j:Q_j\times Q_j\to \mathbb{R},~j=1,2,\ldots, M$ are bifunctions satisfying conditions $C_{1}-C_{4}$. Then, for all $(\bar{x},\bar{y})\in \Upsilon$, we have

(4.1)\begin{align} \Delta_{p}(\bar{x}, z_{n}^{i})\leq \Delta_{p}(\bar{x}, s_n)-\left(1-\kappa\frac{\tau_{n}}{\tau_{n+1}}\right)\left(\Delta_{p}(a_{n}^{i}, s_n) + \Delta_{p}(z_{n}^{i}, a_{n}^{i})\right),\quad \forall i=1,2,\ldots,N \end{align}

and

(4.2)\begin{align} \Delta_{p}(\bar{y}, h_{n}^j)\leq \Delta_{p}(\bar{y}, t_n)-\left(1-\epsilon\frac{\lambda_{n}}{\lambda_{n+1}}\right)\left(\Delta_{p}(b_{n}^j, t_n) + \Delta_{p}(h_{n}^j, b_{n}^j)\right),\quad \forall j=1,2,\ldots,M. \end{align}

Proof. Since $z_{n}^{i}=\arg\min\left\{f_{i}(a_{n}^{i}, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n):\sigma\in C_{i}\right\},$ then from Lemma 2.6, we get

\begin{align*} 0\in \partial_{2}(\tau_{n}f_{i}(a_{n}^{i}, z_{n}^{i})+\Delta_{p}(z_{n}^{i}, s_{n}))+ N_{C_i}(z_{n}^{i}). \end{align*}

Then, there exists $\xi\in \partial_{2}f_{i}(a_{n}^{i}, z_{n}^{i}),$ $\bar{\xi}\in N_{C_i}(z_{n}^{i}),$ such that

(4.3)\begin{align} \tau_{n}\xi + J_{E_1}^{p}(z_{n}^{i})-J_{E_1}^{p}(s_n)+\bar{\xi}=0. \end{align}

Also, by the definition of $\partial_{2}f_{i}(a_{n}^{i}, z_{n}^{i})$, we obtain

\begin{align*} f_{i}(a_{n}^{i},\sigma)-f_{i}(a_{n}^{i}, z_{n}^{i})\geq \langle \sigma-z_{n}^{i}, \xi \rangle, \quad \forall \sigma\in C_{i}. \end{align*}

If we replace σ with $\bar{x}$ in the inequality above, we have

(4.4)\begin{align} f_{i}(a_{n}^{i},\bar{x})-f_{i}(a_{n}^{i}, z_{n}^{i})\geq \langle \bar{x}-z_{n}^{i}, \xi \rangle, \quad \forall \bar{x}\in \Upsilon. \end{align}

Using the definition of $N_{C_i}(z_{n}^{i})$ together with Equation (4.3), we have

(4.5)\begin{align} \langle \sigma-z_{n}^{i}, J_{E_1}^{p}(z_{n}^{i})- J_{E_1}^{p}(s_n)\rangle \geq \tau_{n}\langle z_{n}^{i}-\sigma, \xi \rangle, \quad \forall \sigma\in C_{i}. \end{align}

Again, if we let $\sigma=\bar{x}$ in Equation (4.5), we get

(4.6)\begin{align} \langle \bar{x}-z_{n}^{i}, J_{E_1}^{p}(z_{n}^{i})- J_{E_1}^{p}(s_n)\rangle \geq \tau_{n}\langle z_{n}^{i}-\bar{x}, \xi \rangle, \quad \forall \bar{x}\in \Upsilon. \end{align}

The combination of Equations (4.4) and (4.6) gives

(4.7)\begin{align} \langle \bar{x}-z_{n}^{i}, J_{E_1}^{p}(z_{n}^{i})-J_{E_1}^{p}(s_{n})\rangle &\geq \tau_{n}\langle f_{i}(a_{n}^{i}, z_{n}^{i})-f_{i}(a_{n}^{i}, \bar{x})\rangle\\ &\geq \tau_{n}f_{i}(a_{n}^{i}, z_{n}^{i}),\nonumber \end{align}

because $f_{i}(\bar{x}, a_{n}^{i})\geq 0$ and fi is pseudomonotone on $C_{i},\ \forall i=1,2,\ldots,N.$ Similarly, since $a_{n}^{i}=\arg\min\{f_{i}(s_{n}, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n): \sigma\in C_{i}\},$ we obtain

(4.8)\begin{align} \langle a_{n}^{i}-z_{n}^{i}, J_{E_1}^{p}(a_{n}^{i})-J_{E_1}^{p}(s_{n})\rangle \geq \tau_{n}\left[ f_{i}(s_{n}, z_{n}^{i})-f_{i}(s_{n}, a_{n}^{i})\right]. \end{align}

Using Equations (4.7) and (4.8) together, we get

\begin{align*} \langle \bar{x}-z_{n}^{i}, J_{E_1}^{p}(z_{n}^{i})-J_{E_1}^{p}(s_{n})\rangle &+ \langle a_{n}^{i}-z_{n}^{i}, J_{E_1}^{p}(a_{n}^{i})-J_{E_1}^{p}(s_{n})\rangle\\ &\geq \tau_{n}\left[ f_{i}(s_{n}, z_{n}^{i})-f_{i}(s_{n}, a_{n}^{i})+f_{i}(a_{n}^{i}, z_{n}^{i})\right]. \end{align*}

Applying Bregman three-point identity Equation (2.1), we obtain

\begin{align*} \Delta_{p}(\bar{x}, z_{n}^{i})\leq \Delta_{p}(\bar{x}, s_n)-\Delta_{p}(a_{n}^{i}, s_n)-\Delta_{p}(z_{n}^{i}, a_{n}^{i})+\tau_{n}\{f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})-f(a_{n}^{i}, z_{n}^{i})\}. \end{align*}

Furthermore, by the definition of τn, we obtain

(4.9)\begin{align} \Delta_{p}(\bar{x}, z_{n}^{i})& \leq \Delta_{p}(\bar{x}, s_{n})-\Delta_{p}(a_{n}^{i}, s_n)-\Delta_{p}(z_{n}^{i}, a_{n}^{i})\nonumber\\ & \quad+\frac{\tau_{n}}{\tau_{n+1}}\tau_{n+1}\{f_{i}(s_n,z_{n}^{i})-f_{i}(s_n, a_{n}^{i})-f_{i}(a_{n}^{i}, z_{n}^{i})\}\nonumber\\ & \leq \Delta_{p}(\bar{x}, s_{n})-\Delta_{p}(a_{n}^{i}, s_n)-\Delta_{p}(z_{n}^{i}, a_{n}^{i})+\frac{\tau_{n}}{\tau_{n+1}}\kappa \left(\Delta_{p}(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i},a_{n}^{i})\right)\nonumber\\ & =\Delta_{p}(\bar{x}, s_n)-\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right)\Big(\Delta_p(a_{n}^{i}, s_n)+\Delta_p(z_{n}^{i}, a_{n}^{i})\Big). \end{align}

Following similar procedure, we obtain

(4.10)\begin{align} \Delta_{p}(\bar{y}, h_{n}^j)\leq \Delta_{p}(\bar{y}, t_n)-\left(1-\epsilon\frac{\lambda_{n}}{\lambda_{n+1}}\right)\left(\Delta_{p}(b_{n}^j, t_n) + \Delta_{p}(h_{n}^j, b_{n}^j)\right). \end{align}

Observe that since $\lim\limits_{n\to \infty}\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right)=1-\kappa \gt 0,$ then there exists $K\in\mathbb{N}$ such that

\begin{equation*}\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right) \gt 0,\quad \forall n\ge K.\end{equation*}

Hence, from Equation (4.9), we get

(4.11)\begin{align} \Delta_{p}(\bar{x}, z_n^i)\leq \Delta_{p}(\bar{x}, s_n),\quad\forall i=1,2\ldots,N,~ n\ge K. \end{align}

Similarly, from Equation (4.10), we obtain

(4.12)\begin{align} \Delta_{p}(\bar{y}, h_n^j)\leq \Delta_{p}(\bar{y}, t_n),\quad\forall n\ge L\in\mathbb{N}. \end{align}

Lemma 4.3. Suppose $\{x_n\}$ and $\{y_n\}$ are iterative sequences generated by Algorithm 3.2 under Assumption 3.1. Then, the sequences $\{x_n\}$ and $\{y_n\}$ are bounded.

Proof. Let $(\bar{x}, \bar{y})\in \Upsilon.$ Since Ds is Bregman quasi-nonexpansive for each $s=1,2,\ldots,l,$ we obtain from Equation (3.1) that

(4.13)\begin{align} \Delta_{p}(\bar{x}, u_n)&=\Delta_{p}\left(\bar{x}, J_{q}^{E_1^*}\left(\alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{s=1}^{l}\alpha_{n,s}J_{E_1}^{p}(D_s\bar{z}_{n})\right)\right)\nonumber\\ &\leq\alpha_{n,0}\Delta_{p}(\bar{x}, \bar{z}_{n})+\sum_{s=1}^{l}\alpha_{n,s}\Delta_{p}(\bar{x}, D_s\bar{z}_{n})\nonumber\\ &\leq \alpha_{n,0}\Delta_{p}(\bar{x}, \bar{z}_{n})+\sum_{s=1}^{l}\alpha_{n,s}\Delta_{p}(\bar{x}, \bar{z}_{n})\nonumber\\ &=\Delta_{p}(\bar{x}, \bar{z}_{n}). \end{align}

Similarly, we obtain

(4.14)\begin{align} \Delta_{p}(\bar{y}, v_n)\le\Delta_{p}(\bar{y}, \bar{\theta}_{n}). \end{align}

Furthermore, from Equation (3.1), Lemma 2.7 and Lemma 2.10, we obtain

(4.15)\begin{align} \Delta_p(\bar{x}, s_n)&=\Delta_p\left(\bar{x}, J_{E_1^*}^{q}\left(J_{E_1}^{p}(x_n)-\rho_n A^*J_{E_3}^{p}(Ax_n-By_n)\right)\right)\nonumber\\ &=V_p\left(\bar{x}, J_{E_1}^{p}(x_n)-\rho_n A^*J_{E_3}^{p}(Ax_n-By_n)\right)\nonumber\\ &= \frac{1}{p}\|\bar{x}\|^p -\langle \bar{x}, J_{E_1}^{p}(x_n)\rangle + \rho_n\langle \bar{x}, A^*J_{E_3}^{p}(Ax_n-By_n)\rangle\nonumber\\ &+\frac{1}{q}\|J_{E_1}^{p}(x_n)-\rho_n A^*J_{E_3}^{p}(Ax_n-By_n)\|^q\nonumber\\ &\leq \frac{1}{p}\|\bar{x}\|^p -\langle \bar{x}, J_{E_1}^{p}(x_n)\rangle + \rho_{n}\langle A\bar{x}, J_{E_3}^{p}(Ax_n-By_n)\rangle \nonumber\\ &+\frac{1}{q}\|J_{E_1}^{p}(x_n)\|^q-\rho_{n}\langle J_{E_3}^{p}(Ax_n-By_n), Ax_n\rangle +\frac{C_q}{q}\rho_{n}^q\|A^*J_{E_3}^{p}(Ax_n-By_n)\|^q\nonumber\\ &= \frac{1}{p}\|\bar{x}\|^p-\langle \bar{x}, J_{E_1}^{p}(x_n)\rangle +\frac{1}{q}\|J_{E_1}^{p}(x_n)\|^q -\rho_{n}\langle J_{E_3}^{p}(Ax_n-By_n), Ax_n-A\bar{x}\rangle\nonumber\\ &+\frac{C_q}{q}\rho_{n}^q\|A^*J_{p}^{E_3}(Ax_n-By_n)\|^q\nonumber\\ &=\Delta_p(\bar{x}, x_n)-\rho_{n}\langle J_{E_3}^{p}(Ax_n-By_n), Ax_n-A\bar{x}\rangle\nonumber\\ &+ \frac{C_q}{q}\rho_{n}^q\|A^*J_{p}^{E_3}(Ax_n-By_n)\|^q. \end{align}

Similarly, we have

(4.16)\begin{align} \Delta_p(\bar{y}, t_n)\leq\Delta_p(\bar{y}, y_n)-\rho_{n}\langle J_{E_3}^{p}(Ax_n-By_n), B\bar{y}-By_n\rangle + \frac{Q_q}{q}\rho_{n}^q\|B^*J_{E_3}^{p}(Ax_n-By_n)\|^q. \end{align}

Combining Equations (4.15) and (4.16) and noting that $A\bar{x}=B\bar{y}$, we have

(4.17)\begin{align} \Delta_p(\bar{x}, s_n) & +\Delta_p( \bar{y}, t_n)\leq\Delta_p(\bar{x}, x_n) +\Delta_p(\bar{y}, y_n)-\rho_{n}\left[\|Ax_n-By_n\|^p-\frac{\rho_{n}^{q-1}}{q}\right.\nonumber\\ & \times\left.\left(C_q\|A^*J_{E_3}^{p}(Ax_n-By_n)\|^q+Q_q\|B^*J_{E_3}^{p}(Ax_n-By_n)\|^q\right)\right]. \end{align}

Hence,

(4.18)\begin{align} \Delta_p( \bar{x}, s_n) + \Delta_p(\bar{y}, t_n)\leq\Delta_p(\bar{x}, x_n) +\Delta_p(\bar{y}, y_n). \end{align}

Also, from Equation (3.1) and applying Equation (4.11), we obtain

(4.19)\begin{align} \Delta_{p}(\bar{x}, x_{n+1})&=\Delta_{p}\left(\bar{x}, J_{E_1^*}^{q}\left(\beta_{n} J_{E_1}^{p}(\mu)+(1-\beta_{n})J_{E_1}^{p}(u_n)\right)\right)\nonumber\\ &\leq \beta_{n} \Delta_p(\bar{x}, \mu)+(1-\beta_{n})\Delta_p(\bar{x}, u_{n})\nonumber\\ &\leq \beta_{n} \Delta_p(\bar{x}, \mu)+(1-\beta_{n})\Delta_p(\bar{x}, \bar{z}_{n})\nonumber\\ &\leq \beta_n\Delta_p(\bar{x}, \mu)+(1-\beta_n)\Delta_{p}(\bar{x}, s_n). \end{align}

In like manner, we have

(4.20)\begin{align} \Delta_{p}(\bar{y}, y_{n+1})\leq \beta_n\Delta_p(\bar{y}, \vartheta)+(1-\beta_n)\Delta_{p}(\bar{y}, t_n). \end{align}

It follows from Equations (4.18), (4.19) and (4.20) that

(4.21)\begin{align} \Delta_p(\bar{x}, x_{n+1})+ \Delta_p(\bar{y}, y_{n+1})&\leq\beta_{n}\left(\Delta_p(\bar{x},\mu) +\Delta_p(\bar{y}, \vartheta)\right)+(1-\beta_{n})\left(\Delta_p(\bar{x}, s_n)+\Delta_p(\bar{y}, t_n)\right)\nonumber\\ &\leq\beta_{n}\left(\Delta_p(\bar{x},\mu) +\Delta_p(\bar{y}, \vartheta)\right)+(1-\beta_{n})\left(\Delta_p(\bar{x}, x_n)+\Delta_p(\bar{y}, y_n)\right)\nonumber\\ &\leq \max\{\Delta_p(\bar{x}, \mu) +\Delta_p(\bar{y}, \vartheta),\Delta_p(\bar{x}, x_n)+\Delta_p(\bar{y}, y_n)\}\nonumber\\ &\vdots\nonumber\\ &\leq \max\{\Delta_p(\bar{x}, \mu) +\Delta_p(\bar{y}, \vartheta),\Delta_p(\bar{x}, x_{\bar{N}})+\Delta_p(\bar{y}, y_{\bar{N}} )\},\quad\nonumber\\ \bar{N} & =\max\{K,L\}. \end{align}

Therefore, $\{\Delta_p(\bar{x}, x_{n})+ \Delta_p(\bar{y}, y_{n})\}$ is bounded, and consequently $\{\Delta_p(\bar{x}, x_{n})\}$ and $\{\Delta_p(\bar{y}, y_{n})\}$ are bounded. Hence, by Lemma 2.13, the sequences $\{x_n\}$ and $\{y_n\}$ are bounded. Therefore, $\{s_n\}$, $\{a_{n}^{i}\}$, $\{z_{n}^{i}\}$, $\{u_n\}$, $\{t_n\}$, $\{b_{n}^{j}\}$, $\{h_{n}^{j}\}$ and $\{v_n\}$ are all bounded.

Lemma 4.4. Assume that $r=\sup\{\|J_{E_1}^{p}(\bar{z}_{n})\|,~\|J_{E_1}^{p}(D_s\bar{z}_{n})\|\}$ and let $(\bar{x}, \bar{y})\in \Upsilon.$ Then, the following inequality holds:

(4.22)\begin{align} \Delta_{p}(\bar{x}, x_{n+1}) & +\Delta_{p}(\bar{y}, y_{n+1}) \leq \beta_{n}\left[\Delta_{p}(\bar{x}, \mu)+\Delta_{p}(\bar{y}, \vartheta)\right]+ (1-\beta_n)[\Delta_{p}(\bar{x}, x_n)+\Delta_{p}(\bar{y}, y_n)]\nonumber\\ & -(1-\beta_n)\left(\frac{W_q(\alpha_{n,s})}{q} g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\right.\nonumber\\ & \left.+ \frac{W_q(\eta_{n,t})}{q} g\left(\|J_{E_2}^{p}(\bar{\theta}_n)-J_{E_2}^{p}(G_t\bar{\theta}_n)\|\right)\right), \end{align}

where $W_q(\alpha_{n,s})= (\alpha_{n,0})^q\sum_{s=1}^{l}\alpha_{n,s} + \alpha_{n,0}(\sum_{s=1}^{l}\alpha_{n,s})^q$ and $W_q(\eta_{n,t})= (\eta_{n,0})^q\sum_{t=1}^{m}\eta_{n,t} + \eta_{n,0}(\sum_{t=1}^{m}\eta_{n,t})^q.$

Proof. Let $(\bar{x}, \bar{y})\in \Upsilon.$ Then, from Equation (3.1), Lemma 2.10 and Lemma 2.12, we obtain

\begin{align*} \Delta_{p}(\bar{x}, u_n)&=\Delta_{p}\left(\bar{x}, J_{E_1^*}^{q}\left(\alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{s=1}^{l}\alpha_{n,s}J_{E_1}^{p}(D_s \bar{z}_n)\right)\right)\\ &=V_{p}\left(\bar{x}, \alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{s=1}^{l}\alpha_{n,s}J_{E_1}^{p}(D_s \bar{z}_n)\right)\\ &=\frac{1}{p}\|\bar{x}\|^p-\alpha_{n,0}\langle \bar{x}, J_{E_1}^{p}(\bar{z}_{n})\rangle-\sum_{s=1}^{l}\alpha_{n,s}\langle \bar{x}, J_{E_1}^{p}(D_{s}\bar{z}_n)\rangle\nonumber\\ &+\frac{1}{q}\|\alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{s=1}^{l}\alpha_{n,s}J_{E_1}^{p}(D_s \bar{z}_n)\|^q\nonumber\\ &\leq\frac{1}{p}\|\bar{x}\|^p-\alpha_{n,0}\langle \bar{x},J_{E_1}^{p}(\bar{z}_n)\rangle-\sum_{s=1}^{l}\alpha_{n,s}\langle \bar{x}, J_{E_1}^{p}(D_{s}\bar{z}_n)\rangle\\ &+ \frac{1}{q}\alpha_{n,0}\|J_{E_1}^{p}(\bar{z}_n)\|^p + \frac{1}{q}\sum_{s=1}^{l}\alpha_{n,s}\|J_{E_1}^{p}(D_s\bar{z}_{n})\|^P\\ & -\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\nonumber\\ &=\frac{1}{p}\alpha_{n,0}\|\bar{x}\|^p +\sum_{s=1}^{l}\alpha_{n,s}\frac{1}{p}\|\bar{x}\|^{p}-\alpha_{n,0}\langle \bar{x},J_{E_1}^{p}(\bar{z}_n)\rangle-\sum_{s=1}^{l}\alpha_{n,s}\langle \bar{x}, J_{E_1}^{p}(D_{s}\bar{z}_n)\rangle\\ &+ \frac{1}{q}\alpha_{n,0}\|J_{E_1}^{p}(\bar{z}_n)\|^p + \frac{1}{q}\sum_{s=1}^{l}\alpha_{n,s}\|J_{E_1}^{p}(D_s\bar{z}_n)\|^P\\ & -\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\nonumber\\ &=\alpha_{n,0}\left\{\frac{1}{p}\|\bar{x}\|^p-\langle \bar{x},J_{E_1}^{p}(\bar{z}_n)\rangle +\frac{1}{q}\|J_{E_1}^{p}(\bar{z}_n)\|^p\right\} \nonumber\\ &+\sum_{s=1}^{l}\alpha_{n,s}\left\{\frac{1}{p}\|\bar{x}\|^p -\langle \bar{x}, J_{E_1}^{p}(D_s\bar{z}_n)\rangle +\frac{1}{q}\|J_{E_1}^{p}(D_s\bar{z}_n)\|^p\right\}\nonumber\\ &-\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\nonumber\\ &=\alpha_{n,0}\Delta_p(\bar{x},\bar{z}_n) + \sum_{s=1}^{l}\alpha_{n,s}\Delta_p(\bar{x},D_{s}\bar{z}_n)\\ & -\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right). \end{align*}

By the Bregman quasi-nonexpansivity of Ds for $s=1,2,\ldots,l$, we get

(4.23)\begin{align} \Delta_{p}(\bar{x}, u_n)&\leq \alpha_{n,0}\Delta_p(\bar{x},\bar{z}_n) + \sum_{s=1}^{l}\alpha_{n,s}\Delta_p(\bar{x},\bar{z}_n)-\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\nonumber\\ &=\Delta_p(\bar{x},\bar{z}_n) -\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right). \end{align}

By the definition of $x_{n+1}$ and applying Equations (4.11) and (4.13) from Equation (4.23), we obtain

(4.24)\begin{align} \Delta_{p}(\bar{x}, x_{n+1})&=\Delta_{p}\left(\bar{x}, J_{E_1^*}^{q}\left(\beta_{n} J_{E_1}^{p}(\mu)+(1-\beta_{n})J_{E_1}^{p}(u_n)\right)\right)\nonumber\\ &\leq \beta_{n} \Delta_p(\bar{x}, \mu)+(1-\beta_{n})\Delta_p(\bar{x}, u_{n}) \le \beta_{n}\Delta_{p}(\bar{x}, \mu)+ (1-\beta_n)\nonumber\\ &\quad\times\left(\Delta_p(\bar{x},\bar{z}_n) -\frac{W_q(\alpha_{n,s})}{q}g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\right)\nonumber\\ &\leq \beta_{n}\Delta_{p}(\bar{x}, \mu)+ (1-\beta_n)\Delta_{p}(\bar{x}, s_n)-(1-\beta_n)\frac{W_q(\alpha_{n,s})}{q}g\nonumber\\ & \quad\times\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right). \end{align}

Following similar argument, we have

(4.25)\begin{align} \Delta_{p}(\bar{y}, y_{n+1}) & \leq \beta_{n}\Delta_{p}(\bar{y}, \vartheta)+ (1-\beta_n)\Delta_{p}(\bar{y}, t_n)-(1-\beta_n)\frac{W_q(\eta_{n,t})}{q}g\nonumber\\ & \quad\times\left(\|J_{E_2}^{p}(\bar{\theta}_n)-J_{E_2}^{p}(G_t\bar{\theta}_n)\|\right). \end{align}

By adding Equations (4.24) and (4.25) and applying Equation (4.18), we get

\begin{align*} \Delta_{p}(\bar{x}, x_{n+1})& +\Delta_{p}(\bar{y}, y_{n+1}) \leq \beta_{n}\left[\Delta_{p}(\bar{x}, \mu)+\Delta_{p}(\bar{y}, \vartheta)\right]+ (1-\beta_n)[\Delta_{p}(\bar{x}, s_n)+\Delta_{p}(\bar{y}, t_n)]\nonumber\\ &\quad-(1-\beta_n)\left(\frac{W_q(\alpha_{n,s})}{q} g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\right.\nonumber\\ &\quad\left.+ \frac{W_q(\eta_{n,t})}{q} g\left(\|J_{E_2}^{p}(\bar{\theta}_n)-J_{E_2}^{p}(G_t\bar{\theta}_n)\|\right)\right)\nonumber\\ &\quad\leq \beta_{n}\left[\Delta_{p}(\bar{x}, \mu)+\Delta_{p}(\bar{y}, \vartheta)\right]+ (1-\beta_n)[\Delta_{p}(\bar{x}, x_n)+\Delta_{p}(\bar{y}, y_n)]\nonumber\\ &\quad-(1-\beta_n)\left(\frac{W_q(\alpha_{n,s})}{q} g\left(\|J_{E_1}^{p}(\bar{z}_n)-J_{E_1}^{p}(D_s\bar{z}_n)\|\right)\right.\nonumber\\ &\quad\left.+ \frac{W_q(\eta_{n,t})}{q} g\left(\|J_{E_2}^{p}(\bar{\theta}_n)-J_{E_2}^{p}(G_t\bar{\theta}_n)\|\right)\right), \end{align*}

which is the required inequality.

We now present the main theorem for our proposed algorithm as follows.

Theorem 4.5 Suppose $\{(x_n, y_n)\}$ is a sequence generated by Algorithm 3.2 under Assumption 3.1. Then $\{(x_n, y_n)\}$ converges strongly to $(\bar{x}, \bar{y})\in \Upsilon,$ where $\bar{x}=\Pi_\Upsilon(\mu)$ and $\bar{y}=\Pi_\Upsilon(\vartheta).$

Proof. Let $(\bar{x}, \bar{y})=(\Pi_\Upsilon(\mu),\Pi_\Upsilon(\vartheta)).$ It follows from Algorithm 3.2 and by applying Lemma 2.10 (iii) that

(4.26)\begin{align} \Delta_{p}(\bar{x}, x_{n+1})& =\Delta_{p}\left(\bar{x}, J_{E_1^*}^{q}\left(\beta_{n}J_{E_1}^{p}(\mu)+(1-\beta_n)J_{E_1}^{p}(u_n)\right)\right)\nonumber\\ &=V_{p}\left(\bar{x}, \beta_n J_{E_1}^{p}(\mu)+(1-\beta_n)J_{E_1}^{p}(u_n)\right)\nonumber\\ &\leq V_{p}\left(\bar{x}, \beta_{n}J_{E_1}^{p}(\mu)+(1-\beta_n)J_{E_1}^{p}(u_n)-\beta_{n}\left(J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x})\right)\right)\nonumber\\ &\quad +\beta_{n}\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle\nonumber\\ &=V_{p}\left(\bar{x}, \beta_nJ_{E_1}^{p}(\bar{x})+(1-\beta_n)J_{E_1}^{p}(u_n)\right)+\beta_{n}\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle\nonumber\\ &\leq \beta_{n}\Delta_{p}(\bar{x},\bar{x})+(1-\beta_n)\Delta_{p}(\bar{x}, u_n)+\beta_{n}\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle\nonumber\\ &\leq(1-\beta_n)\Delta_{p}(\bar{x}, s_n)+\beta_{n}\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle. \end{align}

In the same vein, we have

(4.27)\begin{align} \Delta_{p}(\bar{y}, y_{n+1})\leq(1-\beta_n)\Delta_{p}(\bar{y}, t_n)+\beta_{n}\langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n+1}-\bar{y}\rangle. \end{align}

Hence, by adding Equations (4.26) and (4.27) and applying Equation (4.18), we get

(4.28)\begin{align} \Delta_{p}(\bar{x}, x_{n+1})&+\Delta_{p}(\bar{y}, y_{n+1})\leq(1-\beta_n)[\Delta_{p}(\bar{x}, s_n)+\Delta_{p}(\bar{y}, t_n)]\nonumber\\ &+\beta_{n}\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n+1}-\bar{y}\rangle\right)\nonumber\\ &\leq (1-\beta_n)[\Delta_{p}(\bar{x}, x_n)+\Delta_{p}(\bar{y}, y_n)]\nonumber\\ &+\beta_{n}\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n+1}-\bar{y}\rangle\right)\nonumber\\ &=(1-\beta_n)[\Delta_{p}(\bar{x}, x_n)+\Delta_{p}(\bar{y}, y_n)]+\beta_n\chi_{n},\quad \forall n\geq 1, \end{align}

where $\chi_{n}:=\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n+1}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n+1}-\bar{y}\rangle\right).$

In order to show that $\{(x_n, y_n)\}$ converges strongly to $(\bar{x}, \bar{y})$ by Lemma 2.14, we only need to show that $\limsup_{k\to \infty}\chi_{n_k}\leq 0$ for every subsequence $\{\Delta_{p}(x_{n_k}, \bar{x})\}$ of $\{\Delta_{p}(x_{n}, \bar{x})\}$ and $\{\Delta_{p}(y_{n_k}, \bar{y})\}$ of $\{\Delta_{p}(y_{n}, \bar{y})\}$ satisfy the inequality

(4.29)\begin{align} \liminf_{k \rightarrow \infty}\left(\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]-\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p( \bar{y}, y_{n_k})\right]\right)\geq0. \end{align}

Now, from Algorithm 3.2 and Lemma 4.2, we obtain

(4.30)\begin{align} \Delta_{p}(\bar{x}, x_{n+1})&=\Delta_{p}\left(\bar{x}, J_{E_1^*}^{q}(\beta_{n}J_{E_1}^p(\mu)+(1-\beta_{n})J_{E_1}^p(u_n))\right)\nonumber\\ &\leq \beta_n\Delta_p(\mu, \bar{x})+(1-\beta_n)\Delta_p(\bar{x}, u_n)\nonumber\\ &\leq \beta_n\Delta_p(\mu, \bar{x})+(1-\beta_n)\Delta_p(\bar{x},\bar{z}_{n})\nonumber\\ &\leq \beta_n\Delta_p(\mu, \bar{x})+(1-\beta_n)\Delta_{p}(\bar{x}, s_n)-(1-\beta_n)\nonumber\\ &\quad\times\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right)\left(\Delta_{p}(a_{n}^{i_n}, s_n) + \Delta_{p}(z_{n}^{i_n}, a_{n}^{i_n})\right). \end{align}

In the same vein, we obtain

(4.31)\begin{align} \Delta_{p}(\bar{y}, y_{n+1})& \leq \beta_n\Delta_p(\vartheta, \bar{y})+(1-\beta_n)\Delta_{p}(\bar{y}, t_n)-(1-\beta_n)\nonumber\\ &\quad\times\left(1-\frac{\lambda_{n}}{\lambda_{n+1}}\eta \right) \left(\Delta_{p}(b_{n}^{j_n}, t_n) + \Delta_{p}(h_{n}^{j_n}, b_{n}^{j_n})\right). \end{align}

Adding Equations (4.30) and (4.31) together, we obtain

(4.32)\begin{align} &\Delta_{p}(\bar{x}, x_{n+1})+\Delta_{p}(\bar{y}, y_{n+1})\leq \beta_{n}[\Delta_{p}(\mu, \bar{x})+\Delta_p(\vartheta,\bar{y})]+(1-\beta_n)[\Delta_{p}(\bar{x}, s_n)+\Delta_{p}(\bar{y}, t_n)]\nonumber\\ &\quad-(1-\beta_n)\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right)\left(\Delta_{p}(a_{n}^{i_n}, s_n) + \Delta_{p}(z_{n}^{i_n}, a_{n}^{i_n})\right)\nonumber\\ &\quad-(1-\beta_n)\left(1-\frac{\lambda_{n}}{\lambda_{n+1}}\eta \right) \left(\Delta_{p}(b_{n}^{j_n}, t_n) + \Delta_{p}(h_{n}^{j_n}, b_{n}^{j_n})\right). \end{align}

Applying Equation (4.17) in Equation (4.32), we obtain

(4.33)\begin{align} & \Delta_{p}(\bar{x}, x_{n+1})+\Delta_{p}(\bar{y}, y_{n+1})\leq \beta_{n}\left[\Delta_{p}(\mu, \bar{x})+\Delta_p(\vartheta,\bar{y})\right]+ (1-\beta_n)\left[\Delta_p(\bar{x}, x_n) +\Delta_p(\bar{y}, y_n)\right]\nonumber\\ &\quad-(1-\beta_n)\rho_{n}\left[\|Ax_n-By_n\|^p-\frac{\rho_{n}^{q-1}}{q}\left(C_q\|A^*J_{E_3}^{p}(Ax_n-By_n)\|^q\right.\right.\nonumber\\ &\quad\left.\left.+Q_q\|B^*J_{E_3}^{p}(Ax_n-By_n)\|^q\right)\right]-(1-\beta_n)\left(1-\frac{\tau_{n}}{\tau_{n+1}}\kappa\right)\left(\Delta_{p}(a_{n}^{i_n}, s_n) + \Delta_{p}(z_{n}^{i_n}, a_{n}^{i_n})\right)\nonumber\\ &\quad-(1-\beta_n)\left(1-\frac{\lambda_{n}}{\lambda_{n+1}}\eta \right) \left(\Delta_{p}(b_{n}^{j_n}, t_n) + \Delta_{p}(h_{n}^{j_n}, b_{n}^{j_n})\right). \end{align}

By Equation (4.29), Assumption 3.1(2)(a) and (4.33), we obtain

(4.34)\begin{align} & \limsup_{k\rightarrow \infty}\left((1-\beta_{n_k})\rho_{n_k}\left[\|Ax_{n_k}-By_{n_k}\|^p-\frac{\rho_{n_k}^{q-1}}{q}\left(C_q\|A^*J_{p}^{E_3}(Ax_{n_k}-By_{n_k})\|^q\right.\right.\right.\nonumber\\ &\left.\left.\left.+Q_q\|B^*J_{p}^{E_3}(Ax_{n_k}-By_{n_k})\|^q\right)\right]\right)\nonumber\\ &\leq\limsup_{k\rightarrow \infty}\left(\beta_{n_k}\left[\Delta_p(\mu,\bar{x})+\Delta_p(\vartheta,\bar{y})\right]+(1-\beta_{n_k})\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right.\nonumber\\ &\left.-\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]\right)\nonumber\\ &=-\liminf_{k\rightarrow \infty}\left(\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]-\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right)\nonumber\\ &\leq 0. \end{align}

In the same vein as in Equation (4.34), we get using Equation (4.29), Assumption 3.1(2)(a) and Equation (4.33) that

(4.35)\begin{align} \limsup_{k\to \infty}&\left((1-\beta_{n_k})\left(1-\frac{\tau_{n_k}}{\tau_{{n_k}+1}}\kappa\right)\left(\Delta_{p}(a_{n_k}^{i_{n_k}}, s_{n_k}) + \Delta_{p}(z_{n_k}^{i_{n_k}}, a_{n_k}^{i_{n_k}})\right)\right.\nonumber\\ &\left.+\left(1-\frac{\lambda_{n_k}}{\lambda_{{n_k}+1}}\eta \right) \left(\Delta_{p}(b_{n_k}^{j_{n_k}}, t_{n_k}) + \Delta_{p}(h_{n_k}^{j_{n_k}}, b_{n_k}^{j_{n_k}})\right)\right)\nonumber\\ &\leq\limsup_{k\rightarrow \infty}\left(\beta_{n_k}\left[\Delta_p(\mu,\bar{x})+\Delta_p(\vartheta,\bar{y})\right]+(1-\beta_{n_k})\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right.\nonumber\\ &\left.-\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]\right)\nonumber\\ &=-\liminf_{k\rightarrow \infty}\left(\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]-\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right)\nonumber\\ &\leq 0. \end{align}

Now, suppose we let $\varrho_{n_k}=C_q\|A^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q+Q_q\|B^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q.$ Using the condition we placed on our step size $\rho_{n_k},$ we have that

\begin{align*} \rho_{n_k}^{q-1} \lt \frac{q\|Ax_{n_k}-By_{n_k}\|^p}{\varrho_{n_k}}-\zeta, \end{align*}

it follows that

(4.36)\begin{align} \rho_{n_k}^{q-1}{\varrho_{n_k}} \lt q\|Ax_{n_k}-By_{n_k}\|^p-\zeta\varrho_{n_k}, \end{align}

Hence, by Equations (4.34) and (4.36), we have

\begin{equation*} \frac{\zeta\varrho_{n_k}}{q} \lt \left(\|Ax_{n_k}-By_{n_k}\|^p-\frac{\rho_{n_k}^{q-1}}{q}\varrho_{n_k}\right)\to 0,\quad \mbox{as}~k\to\infty. \end{equation*}

Thus, $C_q\|A^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q+Q_q\|B^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q\to 0\quad \mbox{as}~ k\to \infty,$

which implies that

(4.37)\begin{align} \lim_{k\rightarrow \infty}\|A^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q=0 \end{align}

and

(4.38)\begin{align} \lim_{k\rightarrow \infty}\|B^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|^q=0. \end{align}

Additionally, we obtain from Equation (4.34) that

(4.39)\begin{align} & \limsup_{k\rightarrow \infty}\left((1-\beta_{n_k})\rho_{n_k}\left[\|Ax_{n_k}-By_{n_k}\|^p\right]\right)\nonumber\\ &\leq\limsup_{k\rightarrow \infty}\left(\beta_{n_k}\left[\Delta_p(\mu, \bar{x})+\Delta_p(\vartheta,\bar{y})\right]+(1-\beta_{n_k})\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right.\nonumber\\ &\left.-\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]\right)+\limsup_{k\rightarrow \infty}(1-\beta_{n_k})\frac{\rho_{n_k}^q}{q}\varrho_{n_k}\nonumber\\ &=-\liminf_{k\rightarrow \infty}\left(\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]-\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right)\nonumber\\ &\leq 0. \end{align}

Thus, we conclude from Equations (4.34), (4.35) and (4.39) that

(4.40)\begin{align} \lim_{k\rightarrow \infty}\|Ax_{n_k}-By_{n_k}\|=0, \end{align}

(4.41)\begin{align} \begin{cases} \lim_{k\rightarrow \infty}\Delta_p(a_{n_k}^{i_{n_k}}, s_{n_k})=0,\\\\ \lim_{k\rightarrow \infty}\Delta_p(z_{n_k}^{i_{n_k}}, a_{n_k}^{i_{n_k}})=0,\\\\ \lim_{k\rightarrow \infty}\Delta_p(b_{n_k}^{j_{n_k}}, t_{n_k})=0,\\\\ \lim_{k\rightarrow \infty}\Delta_p(h_{n_k}^{j_{n_k}}, b_{n_k}^{j_{n_k}})=0. \end{cases} \end{align}

Therefore, by Lemma 2.11, we obtain

(4.42)\begin{align} \begin{cases} \lim_{k\rightarrow \infty}\|a_{n_k}^{i_{n_k}}-s_{n_k}\|=0,\\\\ \lim_{k\rightarrow \infty}\|z_{n_k}^{i_{n_k}}-a_{n_k}^{i_{n_k}}\|=0,\\\\ \lim_{k\rightarrow \infty}\|b_{n_k}^{j_{n_k}}-t_{n_k}\|=0,\\\\ \lim_{k\rightarrow \infty}\|h_{n_k}^{j_{n_k}}-b_{n_k}^{j_{n_k}}\|=0. \end{cases} \end{align}

Observe that by Equation (4.42) and Lemma 2.11, we have

(4.43)\begin{align} \lim_{k \rightarrow \infty}\|s_{n_k}-z_{n_k}^{i_{n_k}}\|=0,\qquad \lim_{k \rightarrow \infty}\Delta_p(s_{n_k},z_{n_k}^{i_{n_k}})=0. \end{align}

In like manner, we have

(4.44)\begin{align} \lim_{k \rightarrow \infty}\|t_{n_k}-h_{n_k}^{j_{n_k}}\|=0,\qquad \lim_{k \rightarrow \infty}\Delta_p(t_{n_k},h_{n_k}^{j_{n_k}})=0. \end{align}

By the definitions of in and $j_n,$ it follows that

(4.45)\begin{align} \lim_{k \rightarrow \infty}\Delta_p(s_{n_k},z_{n_k}^i)=0,~i=1,2,\ldots,N\quad \text{and}\quad \lim_{k \rightarrow \infty}\Delta_p(t_{n_k},h_{n_k}^j)=0,~j=1,2,\ldots,M. \end{align}

Consequently, we have

(4.46)\begin{align} \lim_{k \rightarrow \infty}\|s_{n_k}-z_{n_k}^i\|=0,~i=1,2,\ldots,N\quad \text{and}\quad \lim_{k \rightarrow \infty}\|t_{n_k},h_{n_k}^j\|=0,~j=1,2,\ldots,M. \end{align}

From Equation (4.1) and by applying the three-point identity (2.1) and (4.46), we have

\begin{align*} \left(1-\kappa\frac{\tau_{n_k}}{\tau_{{n_k}+1}}\right)\Delta_{p}(a_{n_k}^{i}, s_{n_k})&\leq \Delta_{p}(\bar{x}, s_n) - \Delta_{p}(\bar{x}, z_{n_k}^{i})\\ &\leq \Delta_{p}(\bar{x}, s_{n_k}) - \Delta_{p}(\bar{x}, z_{n_k}^{i}) + \Delta_{p}(s_{n_k}, z_{n_k}^{i})\\ &=\langle \bar{x}-s_{n_k}, J_{E_1}^p(z_{n_k}^{i})-J_{E_1}^p(s_{n_k}) \rangle\to 0,\quad k\to\infty. \end{align*}

Hence, we have

\begin{equation*} \Delta_{p}(a_{n_k}^{i}, s_{n_k})\to 0,\quad k\to\infty,\quad i=1,2,\ldots,N. \end{equation*}

Consequently, we obtain

(4.47)\begin{equation} \|a_{n_k}^{i}- s_{n_k}\|\to 0,\quad k\to\infty,\quad i=1,2,\ldots,N. \end{equation}

Following similar procedure, we have

(4.48)\begin{equation} \|b_{n_k}^{j}- t_{n_k}\|\to 0,\quad k\to\infty,\quad j=1,2,\ldots,M. \end{equation}

Furthermore, using Equations (4.22) and (4.29), we have

(4.49)\begin{align} \limsup_{k\to \infty}& (1-\beta_{n_k})\left(\frac{W_q(\alpha_{{n_k},s})}{q} g\left(\|J_{E_1}^{p}(\bar{z}_{n_k})-J_{E_1}^{p}(D_s\bar{z}_{n_k})\|\right)\right.\nonumber\\ & \left.+ \frac{W_q(\eta_{{n_k},t})}{q} g\left(\|J_{E_2}^{p}(\bar{\theta}_{n_k})-J_{E_2}^{p}(G_t\bar{\theta}_{n_k})\|\right)\right)\nonumber\\ &\leq\limsup_{k\rightarrow \infty}\left(\beta_{n_k}\left[\Delta_p(\mu, \bar{x})+\Delta_p(\vartheta,\bar{y})\right]+(1-\beta_{n_k})\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right.\nonumber\\ &\left.-\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]\right)\nonumber\\ &=-\liminf_{k\rightarrow \infty}\left(\left[\Delta_p(\bar{x}, x_{n_k+1})+\Delta_p(\bar{y}, y_{n_k+1})\right]-\left[\Delta_p(\bar{x}, x_{n_k})+\Delta_p(\bar{y}, y_{n_k})\right]\right)\nonumber\\ &\leq 0. \end{align}

Thus,

(4.49)\begin{align*} &\lim_{k\to \infty}\left(\frac{W_q(\alpha_{{n_k},s})}{q} g\left(\|J_{E_1}^{p}(\bar{z}_{n_k})-J_{E_1}^{p}(D_s\bar{z}_{n_k})\|\right) \right.\\ & \quad\left. + \frac{W_q(\eta_{{n_k},t})}{q} g\left(\|J_{E_2}^{p}(\bar{\theta}_{n_k})-J_{E_2}^{p}(G_t\bar{\theta}_{n_k})\|\right)\right)=0. \end{align*}

Hence, we have

\begin{align*} \lim_{k\to \infty}g\left(\|J_{E_1}^{p}(\bar{z}_{n_k})-J_{E_1}^{p}(D_s\bar{z}_{n_k})\|\right)& =0,\ s=1,2,\ldots,l,\quad\\ \lim_{k\to \infty}g\left(\|J_{E_2}^{p}(\bar{\theta}_{n_k})-J_{E_2}^{p}(G_t\bar{\theta}_{n_k})\|\right) & =0,\ t=1,2,\ldots,m. \end{align*}

By the property of $g,$ and sine $J_{E_1^*}^{q}$ and $J_{E_2^*}^{q}$ are norm-to-norm uniformly continuous on bounded subsets of E 1 and E 2, respectively, then we obtain

(4.50)\begin{align} \lim_{k\to \infty}\|D_s\bar{z}_{n_k}-\bar{z}_{n_k}\|=0,\quad \forall s=1,2,\ldots, l \end{align}

and

(4.51)\begin{align} \lim_{k\to \infty}\|G_t\bar{\theta}_{n_k}-\bar{\theta}_{n_k}\|=0,\quad \forall t=1,2,\ldots,m. \end{align}

Observe that from Equation (3.1) and by Equation (4.40), we obtain

(4.52)\begin{align} \|J_{E_1}^{p}(s_{n_k})-J_{E_1}^{p}(x_{n_k})\|&=\|J_{E_1}^{p}(x_{n_k})-\rho_{n_k} A^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})-J_{E_1}^{p}(x_{n_k})\|\nonumber\\ &=\rho_{n_k}\|A^*J_{E_3}^{p}(Ax_{n_k}-By_{n_k})\|\to 0\quad \mbox{as}\ k\to \infty. \end{align}

Also, because E 1 is uniformly smooth, $J_{E_1^*}^{q}$ is norm-to-norm uniformly continuous on bounded subsets of E 1, then we have

(4.53)\begin{align} \lim_{k\to \infty}\|s_{n_k}-x_{n_k}\|=0. \end{align}

In the same vein, we get

(4.54)\begin{align} \lim_{k\to \infty}\|t_{n_k}-y_{n_k}\|=0. \end{align}

Moreover, it is easy to see from Equations (4.46) and (4.53)

(4.55)\begin{align} \lim_{k \rightarrow \infty}\|z_{n_k}^{i}-x_{n_k}\|\leq \lim_{k\to \infty}\|z_{n_k}^{i}-s_{n_k}\|+\lim_{k\to \infty}\|s_{n_k}-x_{n_k}\|=0,\quad \forall i=1,2,\ldots,N. \end{align}

In the same way, we obtain from Equations (4.46) and (4.54) that

(4.56)\begin{align} \|h_{n_k}^{j}-y_{n_k}\|\leq \|h_{n_k}^{j}-t_{n_k}\|+\|t_{n_k}-y_{n_k}\|\to 0 \quad \mbox{as}~k\to \infty,~\forall j=1,2,\dots,M. \end{align}

Moreover, we obtain from Equations (3.1) and (4.50) that

\begin{align*} \lim_{k\to \infty}& \|J_{E_1}^{p}(u_{n_k})-J_{E_1}^{p}(\bar{z}_{n_k})\|=\|\alpha_{n_k,0}J_{E_1}^{p}(\bar{z}_{n_k})+\sum_{s=1}^{l}\alpha_{n_k,s}J_{E_1}^{p}(D_s\bar{z}_{n_k})-J_{E_1}^{p}(\bar{z}_{n_k})\|\nonumber\\ &\leq \alpha_{n_k,0}\|J_{E_1}^{p}(\bar{z}_{n_k})-J(\bar{z}_{n_k})\|+\sum_{s=1}^{l}\alpha_{n_k,s}\|J_{E_1}^{p}(D_s\bar{z}_{n_k})-J_{E_1}^{p}(\bar{z}_{n_k})\|, \end{align*}

which implies that

\begin{align*} \lim_{k\to \infty}\|J_{E_1}^{p}(u_{n_k})-J_{E_1}^{p}\bar{z}_{n_k}\|=0. \end{align*}

By the uniform continuity of $J_{E_1^*}^{q}$ on bounded subsets of $E_{1}^{*}$, we have

(4.57)\begin{align} \lim_{k\to \infty}\|u_{n_k}-\bar{z}_{n_k}\|=0. \end{align}

Hence, from Equations (4.55) and (4.57), we obtain

(4.58)\begin{align} \lim_{k\to \infty}\|u_{n_k}-{z}_{n_k}^{i}\|=0,\quad \forall i=1,2,\ldots,N. \end{align}

Similarly, we obtain

(4.59)\begin{align} \lim_{k\to \infty}\|v_{n_k}-{h}_{n_k}^{j}\|=0,\quad \forall j=1,2,\ldots,M. \end{align}

It is easy to see from Equations (4.55) and (4.58) that

(4.60)\begin{align} \|u_{n_k}-x_{n_k}\|\leq \|u_{n_k}-z_{n_k}^{i}\|+\|z_{n_k}^{i}-x_{n_k}\| \to 0\quad \mbox{as}\ k\to \infty. \end{align}

Similarly, we obtain from Equations (4.56) and (4.59) that

(4.61)\begin{align} \|v_{n_k}-y_{n_k}\|\leq \|v_{n_k}-h_{n_k}^{j}\|+\|h_{n_k}^{j}-y_{n_k}\| \to 0\quad \mbox{as}\ k\to \infty. \end{align}

Furthermore, from Equation (3.1) and the fact that $\lim\limits_{k\to \infty}\beta_{n_k}=0$, we obtain

\begin{align*} \lim_{k\to \infty}\|J_{E_1}^{p}(x_{n_k+1})-J_{E_1}^{p}(u_{n_k})\|=0. \end{align*}

In the same way, we get

\begin{align*} \lim_{k\to \infty}\|J_{E_2}^{p}(y_{n_k+1})-J_{E_2}^{p}(v_{n_k})\|=0. \end{align*}

Since $J_{E_1^*}^{p}$ is norm-to-norm uniformly continuous on bounded subsets of E 1, we obtain

(4.62)\begin{align} \lim_{k\to\infty} \|x_{n_k+1}-u_{n_k}\|=0. \end{align}

Similarly, we get

(4.63)\begin{align} \lim_{k\to\infty} \|y_{n_k+1}-v_{n_k}\|=0. \end{align}

Hence, from Equations (4.60) and (4.62), we obtain

(4.64)\begin{align} \lim_{k\to\infty}\|x_{n_k+1}-x_{n_k}\|=0. \end{align}

In the same vein, from Equations (4.61) and (4.63), we get

(4.65)\begin{align} \lim_{k\to\infty}\|y_{n_k+1}-y_{n_k}\|=0. \end{align}

Since $\{x_n\}$ and $\{y_n\}$ are bounded, then $w_\omega(x_n)$ and $w_{\omega}(y_n)$ are nonempty. Now, let $(x^*, y^*)\in w_\omega(x_n, y_n)$ be arbitrary elements. Then, there exists subsequences $\{x_{n_{k}}\}$ of $\{x_n\}$ and $\{y_{n_{k}}\}$ of $\{y_n\}$ that converge weakly to $x^*\in E_{1}$ and $y^*\in E_{2}$, respectively. Also, from Equations (4.55) and (4.56), $\{z_{n_{k}}^{i}\}$ converges weakly to $x^*\in C_{i}$ for each $i=1,2,\ldots,N$ and $\{h_{n_{k}}^{j}\}$ converges weakly to $y^*\in Q_{j}$ for each $j=1,2,\ldots,M.$ Using Equations (4.50) and (4.51) and by the demiclosedness of $I-D_s$ and $I-G_t$, we obtain

(4.66)\begin{align} x^*\in F(D_{s}),\ \forall s=1,2,\ldots,l\quad\text{and}\quad y^*\in F(G_t),\ \forall t=1,2,\ldots,m, \end{align}

which implies that

(4.67)\begin{align} x^*\in\bigcap_{s=1}^l F(D_{s})\quad\text{and}\quad y^*\in\bigcap_{t=1}^m F(G_t). \end{align}

Next, recall that

\begin{align*} a_{n_k}^{i}=\arg\min_{\sigma\in C_{i}}\{f_{i}(s_{n_k}, \sigma)+ \frac{1}{\tau_{n}}\Delta_{p}(\sigma,s_{n_k})\}. \end{align*}

Using Lemma 2.6 and applying condition (C4), we get

\begin{align*} 0\in \partial_{2} (\tau_{n_k} f_{i}(s_{n_k}, a_{n_k}^{i})+\Delta_{p}(a_{n_k}^{i}, s_{n_k}))+ N_{C_i}(a_{n_k}^{i}). \end{align*}

Hence, there exists $\varsigma_{n_k}^{i}\in \partial_{2}f_i(s_{n_k}, a_{n_k}^{i})$ and $\bar{\varsigma}_{n_k}^{i}\in N_{C_i}(a_{n_k}^{i})$ such that

(4.68)\begin{align} \tau_{n_k} \varsigma_{n_k}^{i} + J_{E_1}^{p}(a_{n_k}^{i})-J_{E_1}^{p}(s_{n_k})+ \bar{\varsigma}_{n_k}^{i}=0. \end{align}

Since $\bar{\varsigma}_{n_k}^{i}\in N_{C_i}(a_{n_k}^{i})$, $\langle \omega-a_{n_k}^{i}, \bar{\varsigma}_{n_k}^{i} \rangle \leq 0$ for all $\omega\in C_{i},$ then this together with Equation (4.68) gives

(4.69)\begin{align} \tau_{n_k}\langle \omega-a_{n_k}^{i}, {\varsigma}_{n_k}^{i} \rangle \geq \langle a_{n_k}^{i}-\omega, J_{E_1}^{p}(a_{n_k}^{i})-J_{E_1}^{p}(s_{n_k})\rangle,\quad \forall \omega \in C_{i}. \end{align}

Again, since $\varsigma_{n_k}^{i}\in \partial_{2}f_i(s_{n_k}, a_{n_k}^{i})$, we obtain

(4.70)\begin{align} f_{i}(s_{n_k}, \omega)-f_{i}(s_{n_k}, a_{n_k}^{i})\geq \langle w-a_{n_k}^{i}, \varsigma_{n_k}^{i}\rangle\quad \forall \omega \in C_{i}. \end{align}

Combining Equations (4.69) and (4.70), we obtain

\begin{align*} \tau_{n_k}\left[f_{i}(s_{n_k}, \omega)-f_{i}(s_{n_k}, a_{n_k}^{i})\right]\geq \langle a_{n_k}^{i}-\omega, J_{E_1}^{p}(a_{n_k}^{i})-J_{E_1}^{p}(s_{n_k})\rangle,\quad \forall \omega \in C_{i}, \end{align*}

which implies that

(4.71)\begin{align} \tau_{n_k}\left[f_{i}(s_{n_k}, a_{n_k}^{i})-f_{i}(s_{n_k}, \omega)\right]&\leq \langle J_{E_1}^{p}(s_{n_k})-J_{E_1}^{p}(a_{n_k}^{i}), a_{n_k}^{i}-\omega\rangle\nonumber\\ &\leq \|J_{E_1}^{p}(s_{n_k})-J_{E_1}^{p}(a_{n_k}^{i})\|\|a_{n_k}^{i}-\omega\|. \end{align}

Since $J_{E_1}^{p}$ is uniformly continuous, applying Equation (4.47) to Equation (4.71) and using Equation (4.53) together with the fact that $x_{n_k}\rightharpoonup x^*,$ we get

\begin{align*} -f_{i}(x^*, \omega)\leq 0,\quad \forall \omega\in C_{i},\ i=1,2,\ldots,N, \end{align*}

which implies that

\begin{align*} f_i(x^*, \omega)\geq 0,\quad \forall \omega\in C_{i},\ i=1,2,\ldots,N. \end{align*}

Hence, we have

\begin{equation*} x^*\in\bigcap_{i=1}^N {\rm EP}(C_i,f_i). \end{equation*}

Similarly, we obtain

\begin{align*} g_j(y^*, z)\geq 0 ,\quad \forall z\in Q_{j},\ j=1,2,\ldots,M, \end{align*}

which implies that

\begin{equation*} y^*\in\bigcap_{j=1}^M {\rm EP}(Q_j,g_j). \end{equation*}

Next, recall that $\{x_{n_k}\}$ and $\{y_{n_k}\}$ converges to $x^*$ and $y^*,$ respectively, where $A:E_{1}\to E_{3}$ and $B:E_{2}\to E_{3}$ are bounded linear operators. Then, by Equation (4.40) and the weakly lower semi-continuity of the norm, we have

\begin{align*} \|Ax^*-By^*\|\leq \liminf_{k\to \infty}\|Ax_{n_k}-By_{n_k}\|=0, \end{align*}

which implies that

\begin{equation*}Ax^*=By^*\end{equation*}

Since $(x^*, y^*)\in w_\omega(x_n, y_n)$ is an arbitrary element, then it follows that

\begin{equation*} w_\omega(x_n, y_n)\subset \Upsilon. \end{equation*}

Next, by the boundedness of $\{x_{n_k}\}$ and $\{y_{n_k}\},$ there exist subsequences $\{x_{n_{k_j}}\}$ of $\{x_{n_k}\}$ and $\{y_{n_{k_j}}\}$ of $\{y_{n_k}\}$ such that $x_{n_{k_j}}\rightharpoonup \hat{x}\in E_1$ and $y_{n_{k_j}}\rightharpoonup \hat{y}\in E_2$ and

\begin{align*} \lim_{j\to \infty}&\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n_{k_j}}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_{k_j}}-\bar{y}\rangle\right)\\ &=\limsup_{k\to \infty}\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n_{k}}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_{k}}-\bar{y}\rangle\right) \end{align*}

Since $\bar{x}=\Pi_\Upsilon(\mu)$ and $\bar{y}=\Pi_\Upsilon(\vartheta),$ then by Equation (2.5), (4.64) and (4.65), we have

(4.72)\begin{align} \limsup_{k\to \infty}&\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n_{k+1}}-\bar{x}\rangle + \langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_{k+1}}-\bar{y}\rangle\right)\nonumber\\ &= \limsup_{k\to \infty}\left(\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n_k}-\bar{x}\rangle +\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), x_{n_{k+1}-x_{n_k}}\rangle\right)\nonumber\\ &+ \limsup_{k\to \infty}\left(\langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_k}-\bar{y}\rangle +\langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_{k+1}-y_{n_k}}\rangle\right)\nonumber\\ &=\lim_{j\to \infty}\left(\langle J_{E_1}^{p}(u)-J_{E_1}^{p}(\bar{x}), x_{n_{k_j}}-\bar{x}\rangle +\langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), y_{n_{k_j}}-\bar{y}\rangle\right)\nonumber\\ &=\langle J_{E_1}^{p}(\mu)-J_{E_1}^{p}(\bar{x}), \hat{x}-\bar{x}\rangle +\langle J_{E_2}^{p}(\vartheta)-J_{E_2}^{p}(\bar{y}), \hat{y}-\bar{y}\rangle\nonumber\\ &\leq 0. \end{align}

Hence, by Equation (4.72), we have $\limsup\limits_{k\to \infty}\chi_{n_k}\leq 0.$ Therefore, by applying Lemma 2.14 to Equation (4.28), it follows that $\{(x_n, y_n)\}$ converges strongly to $(\bar{x}, \bar{y})\in \Upsilon$ as required.

Some corollaries

The following consequent result can easily be obtained from Theorem 4.5 by setting $l=m=N=M=1$.

Corollary. Let E1, E2 and E3 be three p-uniformly convex Banach space and $C, Q$ be nonempty, closed and convex subsets of E1 and E2, respectively. Suppose $f:C \times C \to \mathbb{R}$ and $g:Q \times Q \to \mathbb{R}$ be bifunctions satisfying (C1)–(C4) of Assumption A. Let $A:E_1\to E_3$ and $B:E_2\to E_3$ be bounded linear operators and let $D:E_1\to E_1$ and $G:E_2\to E_2$ be Bregman quasi-nonexpansive mappings such that I ‒ D and I ‒ G are demiclosed at zero and $\Upsilon:=\{\bar{x}\in F(D)\cap \mbox{EP}(C, f)$, $\bar{y}\in F(G)\cap \mbox{EP}(Q, g):A\bar{x}=B\bar{y}\}\neq \emptyset$. Suppose other conditions of Theorem 4.5 hold. For fixed $\mu\in E_1$ and $\vartheta\in E_2$ and initial point $(x_0, y_0)\in E_1\times E_2,$ let $\{(x_n,y_n)\}$ be a sequence generated as follows:

(4.73)\begin{align} \begin{cases} s_n=J_{E_1^*}^{q}\left(J_{E_1}^{p}(x_n)-\rho_{n}A^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ y_{n}=\arg\min\{f(s_n, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n):\quad \sigma\in C\},\\ z_{n}=\arg\min\{f(y_{n}, \sigma)+ \frac{1}{\tau_n}\Delta_p(\sigma, s_n):\quad \sigma\in C\}.\\ u_n=J_{q}^{E_1^*}\left(\alpha_{n,0}J_{E_1}^{p}({z}_n)+\alpha_{n,1}J_{E_1}^{p}(D{z}_n)\right)\\ x_{n+1}=J_{E_1^*}^{q}\left(\beta_{n}J_{E_1}^{p}(\mu)+(1-\beta_{n})J_{E_1}^{p}(u_n)\right),\\ t_n=J_{E_2^*}^{q}\left(J_{E_2}^{p}(y_n)+\rho_nB^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ b_{n}=\arg\min\{g(t_n, \varphi)+ \frac{1}{\lambda_n}\Delta_p(\varphi, t_n):\quad \varphi\in Q\},\\ h_{n}=\arg\min\{g(b_{n}, \varphi)+ \frac{1}{\lambda_n}\Delta_p(\varphi, t_n):\quad \varphi\in Q\}.\\ v_n=J_{E_1^*}^{q}\left(\eta_{n,0}J_{E_2}^{p}(h_n)+\eta_{n,1}J_{E_2}^{p}(Gh_n)\right)\\ y_{n+1}=J_{E_2^*}^{q}\left(\beta_{n}J_{E_2}^{p}(\vartheta)+(1-\beta_{n})J_{p}^{E_2}(v_n)\right), \end{cases} \end{align}

where $\rho_{n}\in \left(\zeta, \left(\frac{q\|Ax_n-By_n\|^p}{C_q\|A^*J_{E_3}^{p}(Ax_n-By_n)\|^q+ Q_q\|B^*J_{E_3}^{p}(Ax_n-By_n)\|^q}-\zeta\right)^\frac{1}{q-1}\right),\ n\in \Omega,$ for small enough ζ; Cq and Qq are constants of smoothness of E 1 and E 2, respectively. Otherwise, $\rho_{n}=\rho$ (ρ being any nonnegative value), where the set of indexes $\Omega=\{n:Ax_n-By_n\neq 0\}$.

(4.74)\begin{align} \tau_{n+1}=\begin{cases} \min\left\{\tau_n, \min \left\{\frac{\kappa(\Delta_p(y_{n},s_n)+\Delta_{p}(z_{n}, y_{n}))}{f(s_n, z_{n})-f(s_n, y_{n})-f(y_{n},z_{n})}\right\}\right\},\quad&\mbox{if}\ f(s_n, z_{n})-f(s_n, y_{n})\\ & \quad-f(y_{n},z_{n}) \gt 0,\\ \tau_n, \quad&\mbox{otherwise}. \end{cases} \end{align}

and

(4.75)\begin{align} \lambda_{n+1}=\begin{cases} \min\left\{\lambda_n, \min \left\{\frac{\epsilon(\Delta_p(b_{n},t_n)+\Delta_{p}(h_{n}, b_{n}))}{g(t_n, h_{n})-g(t_n, b_{n})-g(b_{n},h_{n})}\right\}\right\},\quad&\mbox{if}\ g(t_n, h_{n})-g(t_n, b_{n})\\ & \quad-g(b_{n},h_{n}) \gt 0,\\ \lambda_n, \quad &\mbox{otherwise}. \end{cases} \end{align}

Then, the sequence $\{(x_n, y_n)\}$ generated by Equation (4.73) converges strongly to $(\bar{x}, \bar{y})\in \Upsilon$.

Let $E_r=H_r$, $r=1,2,3$ be real Hilbert spaces, then we obtain the following consequent result for approximating a common solution of multiple sets split equality pseudomonotone EP and common fixed point problems of quasi-nonexpansive mappings in real Hilbert spaces.

Corollary 4.7. Let H 1, H 2 and H 3 be three real Hilbert spaces, and let Ci and Qj be nonempty closed and convex subsets of H 1 and $H_2,$ respectively, for $i=1,2,\ldots,N$ and $j=1,2,\ldots,M.$ Suppose $f_{i}:C_{i} \times C_{i} \to \mathbb{R}$ and $g_{j}:Q_{j} \times Q_{j} \to \mathbb{R}$ are bifunctions satisfying (C1)–(C4) of Assumption A. Let $A:H_1\to H_3$ and $B:H_2\to H_3$ be bounded linear operators. Let $D_{s}:H_1\to H_1$ and $G_{t}:H_2\to H_2$ be quasi-nonexpansive mappings such that $\Upsilon:=\{\bar{x}\in \bigcap_{s=1}^l F(D_s)\cap \bigcap_{i=1}^{N}\mbox{EP}(C_{i}, f_{i})$, $\bar{y}\in \bigcap_{t=1}^mF(G_t)\cap \bigcap_{j=1}^{M}\mbox{EP}(Q_j, g_j):A\bar{x}=B\bar{y}\}\neq \emptyset$. Suppose other conditions of Theorem 4.5 hold. For fixed $\mu\in H_1$ and $\vartheta\in H_2$ and initial point $(x_0, y_0)\in H_1\times H_2,$ let $\{(x_n,y_n)\}$ be a sequence generated as follows:

(4.76)\begin{align} \begin{cases} s_n=\left(x_n-\rho_{n}A^*(Ax_n-By_n)\right),\\ a_{n}^{i}=\arg\min\left\{\tau_n f_{i}(s_n, \sigma)+ \frac{1}{2} \|\sigma-s_n\|^{2}:\quad \sigma\in C_{i}\right\},\\ z_{n}^{i}=\arg\min\left\{{\tau_n}f_{i}(a_{n}^{i}, \sigma)+ \frac{1}{2}\|\sigma-s_n\|^2:\quad \sigma\in C_{i}\right\}.\\ \mbox{Obtain the farthest element of}\ z_{n}^{i}\ \mbox{from}\ s_n,\ \mbox{i.e.,}\\ i_{n}\in \arg\max \left\{\frac{1}{2} \|s_n-z_{n}^{i}\|^2 :i=1,\ldots,N\right\}.\\ \mbox{Set}\ z_{n}^{i_n}=\bar{z}_{n} \\ u_n=\alpha_{n,0}\bar{z}_n+\sum_{s=1}^{l}\alpha_{n,s}(D_s\bar{z}_n)\\ x_{n+1}=\beta_{n}(\mu)+(1-\beta_{n})(u_n),\\ t_n=\left(y_n+\rho_nB^*(Ax_n-By_n)\right),\\ b_{n}^{j}=\arg\min\left\{{\lambda_n} g_j(t_n, \varphi)+ \frac{1}{2} \|\varphi-t_n\|^2:\quad \varphi\in Q_{j}\right\},\\ h_{n}^{j}=\arg\min\left\{{\lambda_n} g_j(b_{n}^{j}, \varphi)+ \frac{1}{2} \|\varphi-t_n\|^2:\quad \varphi\in Q_{j}\right\}.\\ \mbox{Obtain the farthest element of}\ h_{n}^{j}\ \mbox{from}\ t_n,\ \mbox{i.e.,}\\ j_{n}\in \arg\max\left\{\frac{1}{2}\|t_n, h_{n}^{j}\|^2:j=1,\ldots,M\right\}.\\ \mbox{Set}\ h_{n}^{j_n}=\bar{\theta}_{n} \\ v_n=\eta_{n,0}\bar{\theta}_n+\sum_{t=1}^{m}\eta_{n,t}(G_t\bar{\theta}_n)\\ y_{n+1}=\beta_{n}(\vartheta)+(1-\beta_{n})(v_n), \end{cases} \end{align}

where $\rho_{n}\in \left(\zeta, \left(\frac{2\|Ax_n-By_n\|^2}{\|A^*(Ax_n-By_n)\|^2+ \|B^*(Ax_n-By_n)\|^2}-\zeta\right)\right),\ n\in \Omega,$ for small enough $\zeta.$ Otherwise, $\rho_{n}=\rho$ (ρ being any nonnegative value), where the set of indexes $\Omega=\{n:Ax_n-By_n\neq 0\}$.

(4.77)\begin{align} \tau_{n+1}=\begin{cases} \min\left\{\tau_n, \min\limits _{1\leq i\leq N} \left\{\frac{\kappa}{2}\frac{\|a_{n}^{i}-s_n\|^2+\|z_{n}^{i}-a_{n}^{i}\|^2}{f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})-f_{i}(a_{n}^{i},z_{n}^{i})}\right\}\right\},\quad &\mbox{if}\ f_{i}(s_n, z_{n}^{i})-f_{i}(s_n, a_{n}^{i})\\ & \quad-f_{i}(a_{n}^{i},z_{n}^{i}) \gt 0,\\ \tau_n, \quad &\mbox{otherwise} \end{cases} \end{align}

and

(4.78)\begin{align} \lambda_{n+1}=\begin{cases} \min\left\{\lambda_n, \min\limits _{1\leq j\leq M} \left\{\frac{\epsilon}{2}\frac{\|b_{n}^{j}-t_n\|^2+\|h_{n}^{j}- b_{n}^{j}\|^2}{g_{j}(t_n, h_{n}^{j})-g_{j}(t_n, b_{n}^{j})-g_{j}(b_{n}^{j},h_{n}^{j})}\right\}\right\},\quad &\mbox{if}\ g_{j}(t_n, h_{n}^{j})-g_{j}(t_n, b_{n}^{j})\\ & \quad-g_{j}(b_{n}^{j},h_{n}^{j}) \gt 0,\\ \lambda_n, \quad &\mbox{otherwise}. \end{cases} \end{align}

Then the sequence $\{(x_n, y_n)\}$ generated by Equation (4.76) converges strongly to $(\bar{x}, \bar{y})\in \Upsilon$.

5. Application

5.1. Multiple set split equality variational inequality problem

In this section, we apply our result to study the multiple set split equality variational inequality problem (MSSEVIP).

Let $U:C\to E^{*}$ be a nonlinear mapping. The classical VIP is formulated as locating a point

(5.1)\begin{align} x^{*}\in C\quad \mbox{such that}\ \langle \bar{x}-x^{*}, U(x^{*})\rangle \geq 0,\ \forall \bar{x}\in C. \end{align}

The solution set of VIP (5.1) is denoted by $\mbox{VI}(C, U)$. Variational inequalities have been found very applicable in several real-world problems such as optimization problems, minimax theorems, differential equations and in certain applications to economic theory and mechanics. For more details on variational inequalities, see [Reference Alakoya, Mewomo and Shehu2, Reference Ogwo, Izuchukwu, Shehu and Mewomo35, Reference Owolabi, Alakoya, Taiwo and Mewomo36, Reference Uzor, Alakoya and Mewomo44] and the references therein.

Now we consider the MSSEVIP defined as follows:

(5.2)\begin{align} \mbox{find}\ \bar{x}\in C_{i}\quad \mbox{such that}\ \langle U_i\bar{x}, x-\bar{x}\rangle \geq 0, \ \forall x\in C_{i},\ i=1,2,\ldots,N\nonumber\\ \mbox{and}\ \bar{y}\in Q_{j}\quad\mbox{such that}\ \langle V_j\bar{y}, y-\bar{y}\rangle \geq 0, \ \forall y\in Q_{j},\ j=1,2,\ldots,M\nonumber\\ \mbox{such that}\ A\bar{x}=B\bar{y}, \end{align}

where $U_i:E_1\to E_1$ and $V_j:E_2\to E_2$ are two nonlinear mappings, and $A:E_1\to E_3$ and $B:E_2\to E_3$ are two bounded linear operators. When viewed separately, Equation (5.2) consists of two classical multiple sets variational inequality problem (MSVIP) whose solution sets are denoted by $\mbox{VI}(C_{i},U_i)$ and $\mbox{VI}(Q_{j},V_j),$ respectively.

Let $U:C\to E^{*}$ be a nonlinear mapping. Then, U is said to be

  • (D1) pseudomonotone; if for any $x, y \in C$, we have

    \begin{align*} \langle Ux, y-x \rangle \geq 0 \Longrightarrow\langle Uy, y-x\rangle \geq 0, \end{align*}
  • (D2) K-Lipschitz continuous, if there exists a constant K > 0 such that

    \begin{align*} \|Ux-Uy\|\leq K\|x-y\|,\quad \forall x, y \in C, \end{align*}
  • (D3) sequentially weakly continuous, if for any sequence $\{x_n\}\subset C$, we have $x_n\rightharpoonup x\in C$ implying that $Ux_n\rightharpoonup Ux\in E^*.$

We need the following lemma to establish our next result.

Lemma 5.1. [ Reference Eskandani, Raeisi and Rassias14] Let C be a nonempty, closed convex subset of a reflexive, smooth and strictly convex Banach space E, $U:C\to E^*$ be a nonlinear mapping. Then

(5.3)\begin{align} \Pi_{C}\left(J_{E^*}^{q}[J_E^{p}(x)-\lambda U(y)]\right)=\arg\min_{\omega\in C} \left\{\lambda\langle \omega-y, U(y)\rangle + \Delta_p(\omega, x)\right\} \end{align}

for all $x\in E$, $y\in C$ and $\lambda \in (0, +\infty).$

Setting $f_{i}(x,y)=\langle U_ix, y-x\rangle,\ \forall x, y \in C_{i},\ i=1,2,\ldots,N$ and $g_{j}(x, y)=\langle V_jx, y-x\rangle,\ \forall x, y \in Q_{j},\ j=1,2,\ldots,M$ in Algorithm (3.2), then the bifunctions fi and gj satisfy conditions (C1)–(C4) of Assumption A (see [Reference Eskandani, Raeisi and Rassias14]).

Hence, by applying Theorem 4.5 and Lemma 5.1, we obtain the following consequent result for approximating a common solution of MSSEVIP and common fixed point problem for finite families of Bregman quasi-nonexpansive mappings in p-uniformly convex real Banach spaces, which are also uniformly smooth.

Theorem 5.2. Let E1, E2 and E3 be three p-uniformly convex and uniformly smooth real Banach spaces. Let $C_i,\ i=1,2,\ldots,N$ and $Q_j,\ j=1,2,\ldots,M$ be nonempty, closed and convex subsets of E1 and E2, respectively. Let $U_i:C_i\to E^{*}$ and $V_j:Q_j\to E^{*}$ be two nonlinear mappings satisfying conditions (D1)–(D3) above. Let $D_s:E_1\to E_1,\ s=1,2,\ldots,l$ and $G_t:E_2\to E_2,\ t=1,2,\ldots,m$ be two finite families of Bregman quasi-nonexpansive mappings such that $I-D_s$ and $I-G_t$ are demiclosed at zero for each s and t, respectively. Suppose that Assumption 3.12(a)–2(c) holds and the solution set $\Upsilon:=\{\bar{x}\in F(D_s)\cap\mbox{VI}(C_{i},U_i),\ \bar{y}\in F(G_t)\cap\mbox{VI}(Q_j,V_j): A\bar{x}=B\bar{y}\}\neq \emptyset$. Then, the sequence $\{x_n, y_n\}$ generated by Algorithm (5.3) below converges strongly to $(\bar{x}, \bar{y})\in \Upsilon,$ where $\bar{x}=\Pi_\Upsilon(\mu)$ and $\bar{y}=\Pi_\Upsilon(\vartheta).$

Algorithm 5.3.

For fixed $\mu\in E_1$ and $\vartheta\in E_2$, choose an initial guess $(x_0, y_0)\in E_1\times E_2$. Suppose that the nth iterate $(x_n, y_n)\subset E_1\times E_2$ has been constructed, then we compute the $(n+1)$th iterate $(x_{n+1}, y_{n+1})$ via the iteration

\begin{align*} \begin{cases} s_n=J_{E_1^*}^{q}\left(J_{E_1}^{p}(x_n)-\rho_{n}A^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ a_{n}^{i}=\Pi_{C_i}\left[J_{E_1^*}^{p}(J_{E_1}^{p}(s_n)-\tau_n U_i(s_n))\right],\quad i=1,2,\ldots,N\\ z_{n}^{i}=\Pi_{C_i}\left[J_{E_1^*}^{p}(J_{E_1}^{p}(a_{n}^{i})-\tau_n U_i(s_n))\right],\quad i=1,2,\ldots,N\\ i_{n}\in \arg\max \left\{\Delta_{p}(s_n, z_{n}^{i}):i=1,\ldots,N\right\}, \quad z_{n}^{i_n}=\bar{z}_{n}\\ u_n=J_{q}^{E_1^*}\left(\alpha_{n,0}J_{E_1}^{p}(\bar{z}_n)+\sum_{i=1}^{N}\alpha_{n,i}J_{E_1}^{p}(D_s\bar{z}_n)\right)\\ x_{n+1}=J_{E_1^*}^{q}\left(\beta_{n}J_{E_1}^{p}(\mu)+(1-\beta_{n})J_{E_1}^{p}(u_n)\right),\\ t_n=J_{E_2^*}^{q}\left(J_{E_2}^{p}(y_n)+\rho_nB^*J_{E_3}^{p}(Ax_n-By_n)\right),\\ b_{n}^{j}=\Pi_{Q_j}\left[J_{p}^{E_2^*}(J_{p}^{E_2}(t_n)-\lambda_n V_j(t_n))\right]\quad j=1,2,\ldots,M\\ h_{n}^{j}=\Pi_{Q_j}\left[J_{p}^{E_2^*}(J_{p}^{E_2}(t_n)-\lambda_n V_j(b_{n}^{j}))\right]\quad j=1,2,\ldots,M\\ j_{n}\in \arg\max \left\{\Delta_{p}(t_n, h_{n}^{j}):j=1,\ldots,M\right\},\quad h_{n}^{j_n}=\bar{\theta}_{n}\\ v_n=J_{E_1^*}^{q}\left(\eta_{n,0}J_{E_2}^{p}(\bar{\theta}_n)+\sum_{t=1}^{m}\eta_{n,t}J_{E_2}^{p}(G_t\bar{\theta}_n)\right)\\ y_{n+1}=J_{E_2^*}^{q}\left(\beta_{n}J_{E_2}^{p}(\vartheta)+(1-\beta_{n})J_{p}^{E_2}(v_n)\right). \end{cases} \end{align*}

where $\rho_{n}\in \left(\zeta, \left(\frac{q\|Ax_n-By_n\|^p}{C_q\|A^*J_{p}^{E_3}(Ax_n-By_n)\|^q+ Q_q\|B^*J_{p}^{E_3}(Ax_n-By_n)\|^q}-\zeta\right)^\frac{1}{q-1}\right)\ n\in \Omega,$ for small enough ζ, Cq and Qq are constants of smoothness of E1 and E2, respectively. Otherwise, $\rho_{n}=\rho$ (ρ being any nonnegative value), where the set of indexes $\Omega=\{n:Ax_n-By_n\neq 0\}.$

\begin{align*} \tau_{n+1}=\begin{cases} \min\left\{\tau_n, \min\limits _{1\leq i\leq N} \left\{\frac{\kappa(\Delta_p(a_{n}^{i},s_n)+\Delta_{p}(z_{n}^{i}, a_{n}^{i}))}{\langle U_{i}s_n-U_{i}a_{n}^{i}, z_{n}^{i}-a_{n}^{i}\rangle}\right\}\right\},\quad &\mbox{if}\ \langle U_{i}s_n-U_{i}a_{n}^{i}, z_{n}^{i}-a_{n}^{i}\rangle \gt 0,\\ \tau_n, \quad&\mbox{otherwise} \end{cases} \end{align*}

and

\begin{align*} \lambda_{n+1}=\begin{cases} \min\left\{\lambda_n, \min\limits _{1\leq j\leq M} \left\{\frac{\epsilon(\Delta_p(b_{n}^{j},t_n)+\Delta_{p}(h_{n}^{j}, b_{n}^{j}))}{\langle V_{j}t_n-V_{j}b_{n}^{j}, h_{n}^{j}-b_{n}^{j}\rangle}\right\}\right\},\quad &\mbox{if}\ \langle V_{j}t_n-V_{j}b_{n}^{j}, h_{n}^{j}-b_{n}^{j}\rangle \gt 0,\\ \lambda_n, \quad&\mbox{otherwise}. \end{cases} \end{align*}

6. Computational Experiments

In this section, we demonstrate the efficiency and applicability of our proposed method with two numerical examples. In all the experiments, we consider the case when $l=m=N=M=5.$

Example 6.1. Let $E_r=\mathbb{R}^{m},\ r=1,2,3$, equipped with induced norm $\|x\|=\sqrt{\sum_{i=1}^{m}|x_{i}|}$ and the inner product $\langle x,y\rangle = \sum_{i=1}^{m}x_{i}y_{i}$, for all $x=(x_{1}, x_{2},\ldots, x_{m})\in \mathbb{R}^{m}$ and $y=(y_{1}, y_{2},\ldots, y_{m})\in \mathbb{R}^{m}$. Let $C_{i}=Q_{j}=C,$ where the feasible set C has the form

\begin{align*} C=\{(x_{1}, x_2, \ldots,x_m)\in \mathbb{R}_{+}^{m}: |x_{k}|\leq 1,\quad k=1,2,\ldots,m\}. \end{align*}

Consider the following problem:

\begin{align*} \text{Find}\ (\bar{x},\bar{y})\in\Upsilon &:=\left\{\bar{x}\in F(D_s)\cap \bigcap_{i=1}^{N}\mbox{EP}(C_{i}, f_{i}),\right.\\ &\qquad\qquad\qquad\qquad\left.\bar{y}\in F(G_{t})\cap \bigcap_{j=1}^{M}\mbox{EP}(Q_{j}, g_{j}):A\bar{x}=B\bar{y}\right\}, \end{align*}

where $f_{i}:\mathbb{R}^{m}\times \mathbb{R}^{m}\to \mathbb{R}$ is given by

\begin{align*} f_{i}(x,y)=\sum_{k=1}^{n}(q_{ik}y_{k}^{2}-q_{ik}x_{k}^{2}),\quad i=1,2,\ldots,N, \end{align*}

where $q_{ik}\in (0,1)$ is randomly selected $\forall i=1,2,\ldots,N,\ k=1,2,\ldots,m$ and $D_{s}:\mathbb{R}^{m}\to \mathbb{R}^{m}$ is defined by

\begin{align*} D_{s}(x)=\frac{x}{s+1},\quad \forall s=1,2,\ldots,l. \end{align*}

In the same vein, let $g_{j}:\mathbb{R}^{m}\times \mathbb{R}^{m}\to \mathbb{R}$ is given by

\begin{align*} g_{j}(x,y)=\sum_{k=1}^{m}(q_{jk}y_{k}^{2}-q_{jk}x_{k}^{2}),\quad j=1,2,\ldots,M, \end{align*}

where $q_{jk}\in (0,1)$ is randomly selected $\forall j=1,2,\ldots,M,\ k=1,2,\ldots,5$ and $G_{t}:\mathbb{R}^{m}\to \mathbb{R}^{m}$ is defined by

\begin{align*} G_{t}(x)=\frac{x}{t+1},\quad \forall t=1,2,\ldots,5. \end{align*}

It is easy to see that conditions (C1)–(C4) of Assumption A are satisfied and Ds and Gt are Bregman quasi-nonexpansive mappings for $s=1,2,\ldots,l$ and $t=1,2,\ldots,m$, respectively, $I-D_s$ and $I-G_t$ are demiclosed at zero. Moreover, we define $A(x)= \frac{x}{2}$ and $B(x)= \frac{x}{3},$ then A and B are bounded linear operators. Furthermore, $\Upsilon=\{0\}$. In this example, we choose $\beta_{n}=\frac{3}{2n+3}$, κ = 0.36, $\tau_{0}=0.24$, ϵ = 0.5, $\lambda_{0}=0.4$, $\alpha_{n,0}=\frac{3n}{8n+11}$, $\alpha_{n,s}=\frac{1}{5}(1-\frac{3n}{8n+11})$, $s=1,2,\ldots,5$, $\eta_{n,0}=\frac{2n}{4n+7}$, $\eta_{n,t}=\frac{1}{5}(1-\frac{2n}{4n+7})$, $t=1,2,\ldots,5$. Using $\frac{\|x_{n+1}-x_{n}\|}{\|x_2-x_1\|} \lt 10^{-4}$ as our stopping criterion, we generate randomly different starting points $(\mu, \vartheta),(x_0, y_0)\in E_1\times E_2$ for different cases of $m=20,\ 50,\ 100\ {\rm and}\ 500.$

We plot the graphs $\|x_{n+1}-x_{n}\|$ against the number of iterations. The numerical results can be seen in Figure 1.

Figure 1. Top left: m = 20; top right: m = 50; bottom left: m = 100; bottom right: m = 500.

The next example is presented in an infinite dimensional space setting.

Example 6.2. Let $E_r=L^{2}\left([0,1]\right)$, $r=1,2,3$, with the induced norm given by $\|x\|_{L}=\int_{0}^{1}|x(s)|^{2}\,{\rm d}s$ and the corresponding inner product $\langle x,y\rangle=\int_{0}^{1}x(s)y(s)\,{\rm d}s.$ Let the feasible sets Ci and Qj be defined as follows:

\begin{align*} C_{i}:=\{x\in H:\|x\|_{L}\leq 1\}\ i=1,2,\ldots,5\quad \mbox{and}\quad Q_{j}:=\{x\in H:\|x\|_{L}\leq 1\}\ j=1,2,\ldots,5. \end{align*}

Let $f_{i}(x,y)=\langle S_{i}x, y-x\rangle$ and $g_{j}(x,y)=\langle T_{j}x, y-x\rangle$ with the operators $(S_ix)(t)=\max\left\{0, \frac{x(t)}{i}\right\}$ for $i=1,2,\ldots,5$ and $(T_jx)(t)=\max\left\{0, \frac{x(t)}{j}\right\}$ for $j=1,2,\ldots,5.$ Then, it is easy to see that each fi is monotone (and by implication, pseudomonotone) on $C_{i}.$ Similarly, gj is pseudomonotone on Qj. Furthermore, let $D_{s}:H\to H$ and $G_{t}:H\to H$ be defined by $D_{s}(x)(t)=\frac{x(t)}{2s}$ and $G_{t}(x)(t)= \frac{x(t)}{2t},$ then the mappings Ds and Gt are quasi-nonexpansive $\forall s=1,2,\ldots,5$ and $t=1,2,\ldots,5.$ Moreover, we define $A(x)(t)= \frac{x(t)}{3}$ and $B(x)(t)= \frac{x(t)}{5},$ then A and B are bounded linear operators. The solution set $\Upsilon=\{0\}$. We choose $\beta_{n}=\frac{1}{n+2}$, $\alpha_{n,0}=\frac{n+1}{2n+3}$, $\alpha_{n,s}=\frac{1}{5}(1-\frac{n+1}{2n+3})$, $s=1,2,\ldots,5$, $\eta_{n,0}=\frac{n+2}{2n+5}$, $\eta_{n,t}=\frac{1}{5}(1-\frac{n+2}{2n+5})$, $t=1,2,\ldots,5$, κ = 0.54, $\tau_{0}=0.63$, ϵ = 0.75, $\lambda_{0}=0.83,$ and using $\frac{\|x_{n+1}-x_{n}\|}{\|x_2-x_1\|} \lt 10^{-4}$ as stopping criterion.

We choose fixed points $\mu=t^2+2$, $\vartheta=4t^3+3$ and different starting points as follows:

Case I: $x_0=t^2 + 4$, $y_0=t^3 + 2t + 1$

Case II: $x_0=t^4 + 5$, $y_0=t^2 + t + 3$

Case III: $ x_0= \sin(2t),\quad y_0=\cos(5t)$

Case IV: $x_0=\exp(t), \quad y_0=\exp(2t)$.

We plot the graph of errors against the number of iterations in each case. The numerical results can be found in Figure 2.

7. Conclusion

In this article, using Bregman distance, we proposed and studied a new algorithm for approximating the common solution of multiple set split equality pseudomonotone EP and fixed points of Bregman quasi-nonexpansive mappings in real p-uniformly convex Banach spaces, which are also uniformly smooth. The algorithm under discourse is designed in such a way that its convergence does not rely on prior estimates of the Lipschitz constants of the pseudomonotone bifunctions as well as the prior knowledge of the norm of the bounded linear operators. We proved a strongly convergent theorem under some mild conditions on our parameters. We gave a theoretical application of our result and finally present two numerical examples to show the efficiency and applicability of our method. The result presented in this article extends numerous results in the literature in this direction of research.

Figure 2. Top left: Case I; top right: Case II; bottom left: Case III; bottom right: Case IV.

Acknowledgements

The authors sincerely thank the anonymous referee for his careful reading, constructive comments and useful suggestions that improved the manuscript. Opinions expressed and conclusions arrived are those of the authors and are not necessarily to be attributed to the National Research Foundation.

Funding Statement

The second author is supported by the National Research Foundation of South Africa Incentive Funding for Rated Researchers (grant number 119903). The third author is funded by University of KwaZulu-Natal, Durban, South Africa, Postdoctoral Fellowship. He is grateful for the funding and financial support.

Competing Interests

The authors declare that they have no conflict of interest.

References

Alakoya, T. O. and Mewomo, O. T., S-Iteration inertial subgradient extragradient method for variational inequality and fixed point problems, Optimization (2023). doi: 10.1080/02331934.2023.2168482.CrossRefGoogle Scholar
Alakoya, T. O., Mewomo, O. T. and Shehu, Y., Strong convergence results for quasimonotone variational inequalities, Math. Methods Oper. Res. 95(2) (2022), 249279.10.1007/s00186-022-00780-2CrossRefGoogle Scholar
Alakoya, T. O., Uzor, V. A. and Mewomo, O. T., A new projection and contraction method for solving split monotone variational inclusion, pseudomonotone variational inequality, and common fixed point problems, Comput. Appl. Math. 42(1) (2023), Paper No. 3, .10.1007/s40314-022-02138-0CrossRefGoogle Scholar
Anh, P. N., A hybrid extragradient method extended to fixed point problems and equilibrium problems, Optimization 62 (2013), 271283.10.1080/02331934.2011.607497CrossRefGoogle Scholar
Blum, E. and Oettli, W., From optimization and variational inequalities to equilibrium problems, Math. Stud. 63 (1994), 123145.Google Scholar
Bregman, L. M., The relaxation method for finding the common point of convex sets and its application to solution of problems in convex programming, U.S.S.R Comput. Math. Phys. 7 (1967), 200217.10.1016/0041-5553(67)90040-7CrossRefGoogle Scholar
Censor, Y. and Elfving, T., A multi-projection algorithms using Bregman projections in a product space, Numer. Algorithms 8 (1994), 221239.10.1007/BF02142692CrossRefGoogle Scholar
Censor, Y. and Reich, S., Iterations of paracontractions and firmly nonexpansive operators with applications to feasibility and optimization, Optimization 37 (1996), 323339.10.1080/02331939608844225CrossRefGoogle Scholar
Censor, Y., Gibali, A. and Reich, S., Algorithms for the split variational inequality problem, Numer. Algorithms 59 (2012), 301323.10.1007/s11075-011-9490-5CrossRefGoogle Scholar
Chidume, C. E., Geometric Properties of Banach Spaces and Nonlinear Iterations, Springer Verlag Series, Lecture Notes in Mathematics 2009.10.1007/978-1-84882-190-3CrossRefGoogle Scholar
Cholamjiak, P. and Sunthrayuth, P., A Halpern-type iteration for solving the split feasibility problem and fixed point problem of Bregman relatively nonexpansive semigroup in Banach spaces, Filomat 32 (9) (2018), 32113227.10.2298/FIL1809211CCrossRefGoogle Scholar
Cioranescu, I., Geometry of Banach Spaces, Duality Mappings and Nonlinear Problems (Kluwer Academic, Dordrecht, 1990).10.1007/978-94-009-2121-4CrossRefGoogle Scholar
Combettes, P. L. and Hilstoaga, S. A., Equilibrium programming in Hilbert spaces, J. Nonlinear Convex Anal. 6 (2005), 117136.Google Scholar
Eskandani, G. Z., Raeisi, M. and Rassias, T. M., A hybrid extragradient method for solving pseudomonotone equilibrium problems using Bregman distance, J. Fixed Point Theory Appl. 20(3) (2018), Paper No. 132, .10.1007/s11784-018-0611-9CrossRefGoogle Scholar
Fan, K., A minimax inequality and applications, in Inequalities (ed. Shisha, O.), , Volume III (Academic Press, New York, 1972).Google Scholar
Godwin, E. C., Alakoya, T. O., Mewomo, O. T. and Yao, J. -C., Relaxed inertial Tseng extragradient method for variational inequality and fixed point problems, Appl. Anal. (2022). doi: 10.1080/00036811.2022.2107913.Google Scholar
Godwin, E. C., Izuchukwu, C. and Mewomo, O. T., An inertial extrapolation method for solving generalized split feasibility problems in real Hilbert spaces, Boll. Unione Mat. Ital. 14(2) (2021), 379401.10.1007/s40574-020-00272-3CrossRefGoogle Scholar
Godwin, E. C., Izuchukwu, C. and Mewomo, O. T., Image restoration using a modified relaxed inertial method for generalized split feasibility problems, Math. Methods Appl. Sci. 46(5) (2023), 55215544.10.1002/mma.8849CrossRefGoogle Scholar
Godwin, E. C., Taiwo, A. and Mewomo, O. T., Iterative method for solving split common fixed point problem of asymptotically demicontractive mappings in Hilbert spaces, Numer. Algebra Control Optim. 13(2) (2023), 239257.10.3934/naco.2022005CrossRefGoogle Scholar
He, Z., The split equilibrium problem and its convergence algorithms, J. Inequal. Appl. 2012 (2012), . doi: 10.1186/1029-242X-2012-162.CrossRefGoogle Scholar
Hieu, D. V., Common solutions to pseudomonotone equilibrium problems, Bull. Iranian Math. Soc. 42 (2016), 12071219.Google Scholar
Hieu, D. V., Muu, L. D. and Anh, P. K., Parallel hybrid extragradient methods for pseudmonotone equilibrium problems and nonexpansive mappings, Numer. Algorithms 73 (2016), 197217.10.1007/s11075-015-0092-5CrossRefGoogle Scholar
Hieu, D. V. and , P. K., Quy, Accelerated hybrid methods for solving pseudomonotone equilibrium problems, Adv. Comput. Math. 46 (2020), 124.Google Scholar
Kassay, G., Reich, S. and Sabach, S., Iterative methods for solving systems of variational inequalities in reflexive Banach spaces, SIAM J. Optim. 21 (2011), 13191344.10.1137/110820002CrossRefGoogle Scholar
Kim, D. S., and Dinh, B. V., Parallel extragradient algorithms for multiple set split equilibrium problems in Hilbert spaces, Numer. Algorithms 77 (2018), 741761. doi: 10.1007/S11075-017-0338-5CrossRefGoogle Scholar
Kimura, Y. and Saejung, S., Strong convergence for a common fixed points of two different generalizations of cutter operators, Linear Nonlinear Anal. 1 (2015), 5365.Google Scholar
Kuo, L. W. and Sahu, D. R., Bregman distance and strong convergence of proximal-type algorithms, Abstr. Appl. Anal. 2013 (2013), .10.1155/2013/590519CrossRefGoogle Scholar
Ma, Z., Wang, L., and Cho, Y. J., Some results for split equality equilibrium problems in Banach spaces, Symmetry 11 (2019), . doi: 10.3390/sym11020194.CrossRefGoogle Scholar
Martin-Marquez, V., Reich, S. and Sabach, S., Bregman strongly nonexpansive operators in reflexive Banach spaces, J. Math. Anal. Appl. 400 (2013), 597614.10.1016/j.jmaa.2012.11.059CrossRefGoogle Scholar
Moudafi, A., Split monotone variational inclusions, J. Optim. Theory Appl. 150 (2011), 275283.10.1007/s10957-011-9814-6CrossRefGoogle Scholar
Moudafi, A., A second order differential proximal methods for equilibrium problems, J. Inequal. Pure Appl. Math. 2013 (2013), .Google Scholar
Ogbuisi, F. U. and Mewomo, O. T., Convergence analysis of common solution of certain nonlinear problems, Fixed Point Theory 19(1) (2018), 335358.10.24193/fpt-ro.2018.1.26CrossRefGoogle Scholar
Ogwo, G. N., Alakoya, T. O. and Mewomo, O. T., Iterative algorithm with self-adaptive step size for approximating the common solution of variational inequality and fixed point problems, Optimization 72(3) (2023), .10.1080/02331934.2021.1981897CrossRefGoogle Scholar
Ogwo, G. N., Izuchukwu, C. and Mewomo, O. T., Relaxed inertial methods for solving split variational inequality problems without product space formulation, Acta Math. Sci. Ser. B (Engl. Ed.) 42(5) (2022), 17011733.Google Scholar
Ogwo, G. N., Izuchukwu, C., Shehu, Y. and Mewomo, O. T., Convergence of relaxed inertial subgradient extragradient methods for quasimonotone variational inequality problems, J. Sci. Comput. 90(1) (2022), .10.1007/s10915-021-01670-1CrossRefGoogle Scholar
Owolabi, A. O. -E., Alakoya, T. O., Taiwo, A. and Mewomo, O. T., A new inertial-projection algorithm for approximating common solution of variational inequality and fixed point problems of multivalued mappings, Numer. Algebra Control Optim. 12(2) (2022), 255278.10.3934/naco.2021004CrossRefGoogle Scholar
Reem, D., Reich, S. and De Pierro, A., Re-examination of Bregman functions and new properties of their divergences, Optimization 68 (2019), 279348.10.1080/02331934.2018.1543295CrossRefGoogle Scholar
Reich, S. and Sabach, S., A strong convergence theorem for a proximal-type algorithm in reflexive Banach space, J. Nonlinear Convex Anal. 10 (2009), 471485.Google Scholar
Reich, S. and Sabach, S., Three strong convergence theorems regarding iterative methods for solving equilibrium problems in reflelxive Banach spaces, Contemp. Mathematicians 568 (2012), 225240.10.1090/conm/568/11285CrossRefGoogle Scholar
Schöpfer, F., Schuster, T. and Louis, A. K., An iterative regularization method for solving the split feasibility problem in Banach spaces, Inverse Problems 24 (2008), .10.1088/0266-5611/24/5/055008CrossRefGoogle Scholar
Simons, S. and Reich, S., Fenchel duality, Fitzpatrick functions and the Kirszbraun extension theorem, Proc. Amer. Math. Soc. 133 (2005), 26572660.Google Scholar
Taiwo, A. and Mewomo, O. T., Inertial-viscosity-type algorithms for solving generalized equilibrium and fixed point problems in Hilbert spaces, Vietnam J. Math. 50(1) (2022), 125149.10.1007/s10013-021-00485-9CrossRefGoogle Scholar
Tiel, J. V., Convex Analysis: An Introductory Text (Wiley, New York, NY, 1984).Google Scholar
Uzor, V. A., Alakoya, T. O. and Mewomo, O. T., Strong convergence of a self-adaptive inertial Tseng’s extragradient method for pseudomonotone variational inequalities and fixed point problems, Open Math. 20 (2022), 234257.10.1515/math-2022-0030CrossRefGoogle Scholar
Uzor, V. A., Alakoya, T. O. and Mewomo, O. T., On split monotone variational inclusion problem with multiple output sets with fixed point constraints, Comput. Methods Appl. Math. (2022). doi: 10.1515/cmam-2022-0199.Google Scholar
Xu, Z. B. and Roach, G. F., Characteristics inequalities of uniformly convex and uniformly smooth Banach spaces, J. Math. Anal. Appl. 157(1) (1991), 189210.10.1016/0022-247X(91)90144-OCrossRefGoogle Scholar
Figure 0

Figure 1. Top left: m = 20; top right: m = 50; bottom left: m = 100; bottom right: m = 500.

Figure 1

Figure 2. Top left: Case I; top right: Case II; bottom left: Case III; bottom right: Case IV.