1 Introduction

In this paper, we shall assume that H is a real Hilbert space with inner product \(\langle .,.\rangle \) and norm ||.||. Let I denote the identity operator on H. Now, let us recall the definitions of some operators that will be used in this paper.

Let \(T:H\rightarrow H\) be a mapping. A point \(x \in H\) is called a fixed point of T if \(Tx=x\). The set of fixed points of T is denoted by F(T). A point \(x^*\in H\) is called a minimum norm fixed point of T if and only if \(x^* \in F(T)\) and \(||x^*|| =\min \{ ||x||: x \in F(T)\}\).

Definition 1.1

The mapping \(T:H\rightarrow H\) is said to be

  1. (a)

    nonexpansive if

    $$\begin{aligned} ||Tx-Ty||\le ||x-y||, \forall x, y \in H. \end{aligned}$$
  2. (b)

    asymptotically nonexpansive mapping if there exists a sequence \(\{\mu _n\}\) of real positive numbers such that \(\lim \mu _n=0\) and

    $$\begin{aligned} ||T^nx-T^ny||\le (1+\mu _n)||x-y||, \forall x,y \in H, ~~\forall n\ge 1. \end{aligned}$$
  3. (c)

    k-strictly pseudocontractive (see, [2]) if for \(0\le k <1\),

    $$\begin{aligned} || Tx-Ty||^{2} \le || x-y||^{2}+k||(I-T)x-(I-T)y||^{2}, ~~ \forall ~~x,y\in H. \end{aligned}$$
    (1.1)
  4. (d)

    asymptotically k-strict pseudo-contraction mapping in the intermediate sense if there exist a constant \(k\in [0,1)\) and sequences \(\{\mu _n\}\subset [0, \infty ), \{\xi _n\}\subset [0, \infty )\) with \(\mu _n\rightarrow 0\) and \(\xi _n\rightarrow 0\) as \(n\rightarrow \infty \) such that for all \(n\ge 1,\)

    $$\begin{aligned} ||T^nx-T^ny||^2\le (1+\mu _n)||x-y||^2+k||(x-y)-(Tx-Ty)||^2+\xi _n, \forall x,y \in H. \end{aligned}$$
  5. (e)

    \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping [6], if there exist a constant \(k\in [0,1)\) and sequences \(\{\mu _n\}\subset [0, \infty ), \{\xi _n\}\subset [0, \infty )\) with \(\mu _n\rightarrow 0\) and \(\xi _n\rightarrow 0\) as \(n\rightarrow \infty \), and a continuous and strictly increasing function \(\phi :[0,\infty )\rightarrow [0,\infty )\) with \(\phi (0)=0\) such that for all \(n\ge 1,\)

    $$\begin{aligned} ||T^nx-T^ny||^2\le & {} ||x-y||^2+k||(x-y)-(Tx-Ty)||^2\\&+\mu _n\phi (||x-y||)+\xi _n, \forall x,y \in H. \end{aligned}$$

For an example of a total asymptotically strict pseudocontractive mapping, we refer the reader to Chang et al. [6].

In this paper, we shall assume that H is a real Hilbert space with inner product \(\langle .,.\rangle \) and norm ||.||. Let I denote the identity operator on H. Let C and Q be nonempty, closed and convex subsets of real Hilbert spaces \(H_1\) and \(H_2\), respectively. The split feasibility problem (SFP) is to find a point

$$\begin{aligned} x \in C~~ \mathrm{such ~~that~~} Ax\in Q, \end{aligned}$$
(1.2)

where \(A:H_1\rightarrow H_2\) is a bounded linear operator. The SFP in finite-dimensional Hilbert spaces was first introduced by Censor and Elfving [5] for modeling inverse problems which arise from phase retrievals and in medical image reconstruction [3]. The SFP attracts the attention of many authors due to its application in signal processing. Various algorithms have been invented to solve it (see, for example, [4, 12, 15, 17, 19, 21, 23, 26] and references therein).

Note that the split feasibility problem (1.2) can be formulated as a fixed-point equation by using the fact

$$\begin{aligned} P_C(I-\gamma A^*(I-P_Q)A)x^*=x^*, \end{aligned}$$
(1.3)

where \(P_C\) is the metric projection from \(H_1\) onto C [defined as \(||x-P_Cx||=\min _{y\in C}||x-y||\)]. Hence, \(x^*\) solves the SFP (1.2) if and only if \(x^*\) solves the fixed point equation (1.3) (see [18] for the details). This implies that we can use fixed-point algorithms (see [24, 25, 27]) to solve SFP. A popular algorithm that solves the SFP (1.2) is due to Byrne’s CQ algorithm [3] which is found to be a gradient-projection method (GPM) in convex minimization. Subsequently, Byrne [4] applied Krasnoselskii-Mann iteration to the CQ algorithm, and Zhao and Yang [28] applied Krasnoselskii-Mann iteration to the perturbed CQ algorithm to solve the SFP. It is well known that the CQ algorithm and the Krasnoselskii-Mann algorithm for a split feasibility problem do not necessarily converge strongly in the infinite-dimensional Hilbert spaces.

Let \(h: H\rightarrow \mathbb {R}\cup \{+\infty \}\) be a functional on a real Hilbert space H.

Definition 1.2

  1. 1.

    The Moreau-Yosida approximate of the function h of parameter \(\lambda >0\) is defined as \(h_\lambda (y):=\min _{u\in H} \{h(u) + \frac{1}{2 \lambda }||u-y||^2\}\).

  2. 2.

    \(\mathrm{argmin} f:= \{ \bar{x} \in H: f(\bar{x})\le f(x), \forall x \in H\}\).

  3. 3.

    The proximal mapping of f is defined as \(\mathrm{prox}_{\lambda f}(y) = \mathrm{argmin}_{u\in H} \{f(u) + \frac{1}{2 \lambda }||u-y||^2\}\).

  4. 4.

    The subdifferential of f at x is the set

    $$\begin{aligned} \partial f(x):=\{u \in H:f(y)\ge f(x)+\langle u, y-x\rangle , \forall y \in H \}. \end{aligned}$$

Let us consider the following problem: find a solution \(x^*\in H_1\) such that

$$\begin{aligned} \underset{x \in H_1}{\min }\{f(x)+g_\lambda (Ax)\}, \end{aligned}$$
(1.4)

where \(H_1, H_2\) are two real Hilbert spaces, \(f:H_1\rightarrow \mathbb {R}\cup \{+\infty \}, g:H_2 \rightarrow \mathbb {R}\cup \{+\infty \}\) two proper, convex, lower semicontinuous functions and \(A:H_1\rightarrow H_2\) a bounded linear operator, \(g_\lambda (y) = \min _{u\in H_2} \{g(u) + \frac{1}{2 \lambda }||u-y||^2\}\) stands for the Moreau-Yosida approximate of the function g of parameter \(\lambda \).

Observe that by taking \(f =\delta _C\) [defined as \(\delta _C(x) = 0\) if \(x\in C\) and \(+\infty \) otherwise], \(g =\delta _Q\) the indicator functions of two nonempty, closed and convex sets C, Q of \(H_1\) and \(H_2\) respectively, Problem (1.4) reduces to

$$\begin{aligned} \underset{x \in H_1}{\min }\{\delta _C(x)+(\delta _{Q})_\lambda (Ax)\}\Leftrightarrow \underset{x \in C}{\min }\left\{ \frac{1}{2\lambda }||(I-P_Q)(Ax)||^2\right\} \end{aligned}$$
(1.5)

which, when \(C\cap A^{-1}(Q)\ne \emptyset \), is equivalent to (1.2).

By the differentiability of the Moreau-Yosida approximate \(g_\lambda \), see for instance [16], we have the additivity of the subdifferentials and thus we can write

$$\begin{aligned} \partial (f(x)+g_\lambda (Ax))=\partial f(x)+A^*\nabla g_\lambda (Ax)=\partial f(x)+A^*\Big (\frac{I-prox_{\lambda g}}{\lambda } \Big )(Ax). \end{aligned}$$

This implies that the optimality condition of (1.4) can be then written as

$$\begin{aligned} 0 \in \lambda \partial f(x)+A^*(I-prox_{\lambda g})(Ax). \end{aligned}$$
(1.6)

Inclusion (1.6) in turn yields to the following equivalent fixed point formulation

$$\begin{aligned} \mathrm{prox}_{\mu \lambda f}\left( x^*-\mu A^*(I-\mathrm{prox}_{\lambda g})Ax^*\right) =x^*. \end{aligned}$$
(1.7)

To solve (1.4), relation (1.7) suggests to consider the following split proximal algorithm

$$\begin{aligned} x_{n+1}=\mathrm{prox}_{\mu _n \lambda f}\left( x_n-\mu _n A^*(I-\mathrm{prox}_{\lambda g})Ax_n\right) . \end{aligned}$$
(1.8)

Based on an idea introduced in Lopez et al. [11], Moudafi and Thakur [13] recently proved weak convergence results for solving (1.4) in the case \(\mathrm{argmin} f\cap A^{-1}(\mathrm{argmin} g)\ne \emptyset \), or in other words: in finding a minimizer \(x^*\) of f such that \(Ax^*\) minimizes g, namely

$$\begin{aligned} x^* \in \mathrm{argmin} f~~\mathrm{such~~ that}~~ Ax^*\in \mathrm{argmin}~~ g, \end{aligned}$$
(1.9)

f, g being two proper, lower semicontinuous convex functions. We will denote the solution set of (1.9) by \(\Gamma \). Concerning problem (1.9), Moudafi and Thakur [13] introduced a new way of selecting the step-sizes: Set \(\theta (x_n):=\sqrt{||\nabla h(x_n)||^2+||\nabla l(x_n)||^2}\) with \(h(x_n)=\frac{1}{2}||(I-\mathrm{prox}_{\lambda g})Ax_n||^2, l(x_n)=\frac{1}{2}||(I-\mathrm{prox}_{\lambda \mu _n f})x_n||^2\) and introduced the following split proximal algorithm:

Split Proximal Algorithm: Given an initial point \(x_1\in H_1\). Assume that \(x_n\) has been constructed and \(\theta (x_n) \ne 0\), then compute \(x_{n+1}\) via the rule

$$\begin{aligned} x_{n+1} =\mathrm{prox}_{\lambda \mu _n f}( x_n-\mu _n A^*(I-\mathrm{prox}_{\lambda g})Ax_n),~~n \ge 1, \end{aligned}$$
(1.10)

where stepsize \(\mu _n:=\rho _n\frac{h(x_n)+l(x_n)}{\theta ^2(x_n)}\) with \(0<\rho _n<4\). If \(\theta (x_n)=0\), then \(x_{n+1}=x_n\) is a solution of (1.4) and the iterative process stops, otherwise, we set \(n:= n + 1\) and go to (1.10).

Using the split proximal algorithm (1.10), Moudafi and Thakur [13] proved the following weak convergence theorem for approximating a solution of (1.9).

Theorem 1.3

Assume that f and g are two proper convex lower-semicontinuous functions and that (1.9) is consistent (i.e., \(\Gamma \ne \emptyset \)). If the parameters satisfy the following conditions \(\epsilon \le \rho _n \le \frac{4 h(x_n)}{h(x_n)+l(x_n)}-\epsilon \) (for some \(\epsilon >0\) small enough), then the sequence \(\{x_n\}\) generated by (1.10) weakly converges to a solution of (1.9).

We remark here that it is quite usual to seek a particular solution of a given nonlinear problem, in particular, the minimum-norm solution. For instance, given a nonempty, closed and convex subset C of a Hilbert space \(H_1\) and a bounded linear operator \(A:H_1\rightarrow H_2\), where \(H_2\) is another Hilbert space. The C-constrained pseudoinverse of A, \(A^{\dag }_C\) is then defined as the minimum-norm solution of the constrained minimization problem

$$\begin{aligned} A^{\dag }_C(b):=\mathrm{argmin}_{x \in C}||Ax-b|| \end{aligned}$$

which is equivalent to the fixed point problem

$$\begin{aligned} x=P_C(x-\lambda A^*(Ax-b)), \end{aligned}$$

where \(P_C\) is the metric projection from \(H_1\) onto C, \(A^*\) is the adjoint of A, \(\lambda >0\) is a constant, and \(b\in H_2\) is such that \(P_{\overline{A(C)}}(b)\in A(C)\). It is therefore our aim in this paper to introduce an iterative algorithm that can generate sequence which converges strongly to the minimum-norm solution of a given convex proximal split feasibility problem and fixed point problems for total asymptotically strict pseudocontractive mapping in real Hilbert spaces.

2 Preliminaries

Definition 2.1

A mapping \(T:H\rightarrow H\) is said to be uniformly L-Lipschitzian continuous if there exists a constant \(L > 0\) such that \(||T^nx - T^ny||\le L||x-y||, \forall x, y \in H, n\ge 1.\)

Lemma 2.2

Let K be a nonempty, closed and convex subset of a real Hilbert space H and let \(T:K\rightarrow K\) be a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping such that \(F(T)\ne \emptyset \). Suppose there exist constants \(M_0>0, K_1 > 0\) such that \(\phi (\lambda )\le M_0 \lambda ^2, \forall \lambda >K_1\). Then F(T) is closed and convex.

Proof

Since T is a uniformly L-Lipschitzian continuous, F(T) is closed. Next, we show that F(T) is convex. Next, we show that F(T) is convex. For \(t\in [0,1]\) and \(x,y \in F(T)\), put \(z:= tx + (1- t)y\), we show that \(z=Tz\). Since \(\phi \) is continuous, it follows that \(\phi \) attains maximum (say M) in \([0,K_1]\) and by our assumption, \(\phi (t)\le M_0 t^2, \forall t >K_1\). In either case, we have that

$$\begin{aligned} \phi (t)\le M+M_0 t^2, \forall t \in [0,\infty ). \end{aligned}$$

Using Lemma 2.3 (iii), we obtain

$$\begin{aligned} ||T^nz-z||^2= & {} ||t(T^nz-x)+(1-t)(T^nz-y)||^2\nonumber \\= & {} t||T^nz-x||^2+(1-t)||T^nz-y||^2-t(1-t)||x-y||^2\nonumber \\\le & {} t\Big (||z-x||^2+k||T^nz-z||^2+\mu _n\phi (||z-x||)+\xi _n\Big ) \nonumber \\&+\,(1-t)\Big (||z-y||^2+k||T^nz-z||^2+\mu _n\phi (||z-y||)+\xi _n\Big )\nonumber \\&-\,t(1-t)||x-y||^2 \nonumber \\\le & {} t\Big (||z-x||^2+k||T^nz-z||^2+\mu _n(M+M_0||z-x||^2)+\xi _n\Big ) \nonumber \\&+\,(1-t)\Big (||z-y||^2+k||T^nz-z||^2+\mu _n(M+M_0||z-y||^2)+\xi _n\Big )\nonumber \\&-\,t(1-t)||x-y||^2 \nonumber \\= & {} t\Big ((1\,{-}\,t)^2||x\,{-}\,y||^2+k||T^nz\,{-}\,z||^2+\mu _n(M+M_0||z\,{-}\,x||^2)+\xi _n\Big )(1\,{-}\,t) \nonumber \\&\times \,\Big (t^2||x-y||^2+k||T^nz-z||^2+\mu _n(M+M_0||z-y||^2)+\xi _n\Big )\nonumber \\&-\,t(1-t)||x-y||^2 \nonumber \\= & {} k||T^nz-z||^2+\mu _n(M+M_0||z-y||^2)+\xi _n. \end{aligned}$$
(2.1)

This implies from (2.1) that

$$\begin{aligned} (1-k)||T^nz-z||^2\le \mu _n(M+M_0||z-y||^2)+\xi _n. \end{aligned}$$

Thus, \(\lim \nolimits _{n\rightarrow \infty } ||T^nz-z||=0\), which implies that \(\lim \nolimits _{n\rightarrow \infty } T^nz=z\). By continuity of T, we obtain that

$$\begin{aligned} z=\underset{n\rightarrow \infty }{\lim }T^nz=\underset{n\rightarrow \infty }{\lim }T(T^{n-1}z)= T(\underset{n\rightarrow \infty }{\lim }T^{n-1}z)=Tz. \end{aligned}$$

Hence, \(z \in F(T)\), that F(T) is convex. \(\square \)

We state the following well-known lemmas which will be used in the sequel.

Lemma 2.3

Let H be a real Hilbert space. Then there holds the following well-known results:

$$\begin{aligned} (i)~~||x+y||^2=||x||^2+2\langle x,y\rangle +||y||^2, \forall x, y \in H. \end{aligned}$$
$$\begin{aligned} (ii)~~||x+y||^2 \le ||x||^2+ 2\langle y, x+y \rangle , \forall x, y \in H. \end{aligned}$$
$$\begin{aligned} (iii)~~||tx+(1\!-\!t)y||^2=t||x||^2+(1\!-\!t)||y||^2\!-\!t(1\!-\!t)||x\!-\!y||^2, \forall x, y \in H, \forall t \in [0,1]. \end{aligned}$$

Lemma 2.4

(Chang et al. [6]) Let \(T:H\rightarrow H\) be a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping, then \(I-T\) is demiclosed at 0, i.e., if \(x_n\rightharpoonup x\in H\) and \(x_{n}-Tx_{n}\rightarrow 0\), then \(x=Tx.\)

Lemma 2.5

(Alber et al. [1]) Let \(\{\lambda _n\}\) and \(\{\gamma _n\}\) be nonnegative, \(\{\alpha _n\}\) be positive real numbers such that

$$\begin{aligned} \lambda _{n+1} \le \lambda _n-\alpha _n\lambda _n +\gamma _n, n\ge 1. \end{aligned}$$

Let for all \(n > 1\),

$$\begin{aligned} \frac{\gamma _n}{\alpha _n}\le c_1~~\mathrm{and}~~\alpha _n \le \alpha . \end{aligned}$$

Then \(\lambda _n \le \max \{\lambda _1,K_{*}\},\) where \(K_{*}=(1 + \alpha )c_1\).

Lemma 2.6

(Xu, [20]) Let \(\{a_n\}\) be a sequence of nonnegative real numbers satisfying the following relation:

$$\begin{aligned} a_{n+1}\le (1-\alpha _n)a_n+\alpha _n\sigma _n+\gamma _n,~~n \ge 1, \end{aligned}$$

where

  1. (i)

    \(\{a_n\}\subset [0,1],~~\sum \alpha _n=\infty ;\)

  2. (ii)

    \(\limsup \sigma _n \le 0\);

  3. (iii)

    \(\gamma _n \ge 0;~~(n \ge 1),~~\Sigma \gamma _n <\infty .\)

Then, \(a_n\rightarrow 0\) as \(n\rightarrow \infty \).

Now, our interest is in studying the convergence properties of the following algorithm: Given an initial point \(x_1\in H_1\), then compute \(u_n\) using \(u_n=(1-\alpha _n)x_n\). Set \(\theta (u_n):=\sqrt{||\nabla h(u_n)||^2+||\nabla l(u_n)||^2}\) with \(h(u_n)=\frac{1}{2}||(I-\mathrm{prox}_{\lambda g})Au_n||^2, l(u_n)=\frac{1}{2}||(I-\mathrm{prox}_{\lambda \tau _n f})u_n||^2\) and introduce the following algorithm:

Algorithm: Let \(T:H_1\rightarrow H_1\) is a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping such that \(F(T)\ne \emptyset \). Given an initial point \(x_1\in H_1\), the compute \(u_n\) using \(u_n=(1-\alpha _n)x_n\) and \(\theta (u_n) \ne 0\), then compute \(x_{n+1}\) via the rule

$$\begin{aligned} \left\{ \begin{array}{l} u_n=(1-\alpha _n)x_n,\\ y_n =\mathrm{prox}_{\lambda \tau _n f}(u_n-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Au_n),\\ x_{n+1} =(1-\beta _n)y_n+ \beta _nT^ny_n,~~n \ge 1, \end{array} \right. \end{aligned}$$
(2.2)

where stepsize \(\tau _n:=\rho _n\frac{h(u_n)+l(u_n)}{\theta ^2(u_n)}\) with \(0<\rho _n<4\). If \(\theta (u_n)=0\), then \(x_{n+1}=x_n\) is a solution of (1.9) which is also a fixed point of a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping T and the iterative process stops, otherwise, we set \(n:= n + 1\) and go to (2.2).

3 Main results

Theorem 3.1

Let \(T:H_1\rightarrow H_1\) is a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping such that \(F(T)\ne \emptyset \). Assume that \(f:H_1\rightarrow \mathbb {R}\cup \{+\infty \}\) is a proper convex lower-semicontinuous function, \(g:H_2 \rightarrow \mathbb {R}\cup \{+\infty \}\) is a proper convex lower-semicontinuous function such that (1.9) is consistent (i.e., \(\Gamma \ne \emptyset \)) and \(\Gamma \cap F(T)\ne \emptyset \). Let \(\{\alpha _n\}\) and \(\{\beta _n\}\) be sequences in (0, 1). If the parameters satisfy the following conditions

  1. (a)

    \(\sum \limits _{n=1}^{\infty }\mu _n<\infty ; \sum \limits _{n=1}^{\infty } \xi _n<\infty ;\)

  2. (b)

    \(\mu _n=o(\alpha _n); \xi _n=o(\alpha _n); \underset{n\rightarrow \infty }{\lim }\alpha _n =0; \sum \limits _{n=1}^{\infty } \alpha _n =\infty ;\)

  3. (c)

    \(\epsilon \le \rho _n \le \frac{4 h(u_n)}{h(u_n)+l(u_n)}-\epsilon \) for some \(\epsilon >0\);

  4. (d)

    \(0<\underset{n\rightarrow \infty }{\liminf }\beta _n \le \underset{n\rightarrow \infty }{\limsup }\beta _n<1-k\);

  5. (e)

    there exist constants \(M_0>0, K_1 > 0\) such that \(\phi (t)\le M_0 t^2, \forall t >K_1\);

then sequence \(\{x_n\}\) generated by (2.2) converges strongly to \(x^*\in \Gamma \cap F(T)\) which is also the minimum-norm solution (i.e., \(x^*\in \Gamma \cap F(T)\) and \(||x^*|| =\min \{ ||x||: x \in \Gamma \cap F(T)\}\)).

Proof

Let \(x^{*}\in \Gamma \). Observe that \(\nabla h(u_n)= A^*(I-\mathrm{prox}_{\lambda g})Au_n, \nabla l(u_n) = (I-\mathrm{prox}_{\tau _n \lambda f})x\). Using the fact that \(\mathrm{prox}_{\tau _n \lambda f}\) is nonexpansive, \(x^{*}\) verifies (1.9) (since minimizers of any function are exactly fixed-points of its proximal mapping) and having in hand

$$\begin{aligned} \langle \nabla h(u_n), u_n-x^{*} \rangle =\langle (I-\mathrm{prox}_{\lambda g})Au_n, Au_n-Ax^{*}\rangle \ge ||(I-\mathrm{prox}_{\lambda g})Au_n||^2=2h(u_n), \end{aligned}$$

thanks to the fact that \(I-\mathrm{prox}_{\lambda g}\) is firmly nonexpansive, we can write

$$\begin{aligned} ||y_n-x^{*}||^2= & {} ||u_n-x^{*}||^2+\tau _n^2||\nabla h(u_n)||^2-2\tau _n\langle \nabla h(u_n), u_n-x^{*}\rangle \nonumber \\\le & {} ||u_n-x^{*}||^2+\tau _n^2||\nabla h(u_n)||^2-4\tau _nh(u_n) \nonumber \\= & {} ||u_n-x^{*}||^2+\rho ^2_n\frac{(h(u_n)+l(u_n))^2}{(\theta ^2(u_n))^2}||\nabla h(u_n)||^2-4\rho _n\frac{h(u_n)+l(u_n)}{\theta ^2(u_n)}h(u_n) \nonumber \\\le & {} ||u_n-x^{*}||^2+\rho ^2_n\frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)}-4\rho _n\frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)}\frac{h(u_n)}{h(u_n)+l(u_n)} \nonumber \\= & {} ||u_n-x^{*}||^2-\rho _n\Big (\frac{4h(u_n)}{h(u_n)+l(u_n)}-\rho _n\Big )\frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)}. \end{aligned}$$
(3.1)

Using (3.1) in (2.2), we obtain that

$$\begin{aligned} || x_{n+1}-x^{*}||^2= & {} ||(1-\beta _n)y_n+\beta _n T^ny_n-x^{*}||^2 \nonumber \\= & {} || (1-\beta _n)(y_n-x^{*})+\beta _n(T^ny_n-x^{*})||^2 \nonumber \\= & {} (1-\beta _n)|| y_n-x^{*}||^2 +\beta _n|| T^ny_n-x^{*}||^2-\beta _n(1-\beta _n)||T^ny_n-y_n||^2 \nonumber \\\le & {} (1-\beta _n)|| y_n-x^{*}||^2 +\beta _n\Big [|| y_n-x^{*}||^2 +k||T^ny_n-y_n||^2 \nonumber \\&+\,\mu _n \phi (||y_n-x^{*}||)+\xi _n\Big ]-\beta _n(1-\beta _n)||T^ny_n-y_n||^2 \nonumber \\= & {} || y_n-x^{*}||^2+ \beta _n(\beta _n-(1-k))||T^ny_n-y_n||^2+\beta _n\mu _n\phi (||y_n-x^{*}||) \nonumber \\&+\,\beta _n \xi _n. \end{aligned}$$
(3.2)

Since \(\phi \) is continuous, it follows that \(\phi \) attains maximum (say M) in \([0,K_1]\) and by our assumption, \(\phi (t)\le M_0 t^2, \forall t >K_1\). In either case, we have that

$$\begin{aligned} \phi (t)\le M+M_0 t^2, \forall t \in [0,\infty ). \end{aligned}$$
(3.3)

Using (3.2) and (3.3), we have that

$$\begin{aligned} || x_{n+1}-x^{*}||^2\le & {} || y_n-x^{*}||^2+ \beta _n(\beta _n-(1-k))||T^ny_n-y_n||^2+\beta _n\mu _n M_0||y_n-x^{*}||^2 \nonumber \\&+\,\beta _n \xi _n+\beta _n\mu _n M_1\nonumber \\\le & {} || y_n-x^{*}||^2+\beta _n\mu _n M_0||y_n-x^{*}||^2 +\beta _n \xi _n+\beta _n\mu _n M_1\nonumber \\\le & {} ||u_n-x^{*}||^2+\beta _n\mu _n M_0 ||u_n-x^{*}||^2 +\beta _n \xi _n+\beta _n\mu _n M_1\nonumber \\\le & {} \Big [(1-\alpha _n)||x_n-x^{*}||^2+\alpha _n||x^{*}||^2\Big ]\nonumber \\&+\,\beta _n\mu _n M_0\Big [(1-\alpha _n)||x_n-x^{*}||^2+\alpha _n||x^{*}||^2\Big ] +\beta _n \xi _n+\beta _n\mu _n M_1\nonumber \\\le & {} \Big [(1-\alpha _n)||x_n-x^{*}||^2+\alpha _n||x^{*}||^2\Big ]+\beta _n\mu _n M_0||x_n-x^{*}||^2\nonumber \\&+\,\beta _n\mu _n M_0||x^{*}||^2+\beta _n \xi _n+\beta _n\mu _n M_1\nonumber \\= & {} ||x_n-x^{*}||^2-\Big [\alpha _n-\beta _n\mu _n M_0\Big ]||x_n-x^{*}||^2+\sigma _n, \end{aligned}$$
(3.4)

where \(\sigma _n= \alpha _n||x^{*}||^2+\beta _n\mu _n M_0||x^{*}||^2+\beta _n \xi _n+\beta _n\mu _n M_1,~~\forall n\ge 1.\) From (3.4), we have

$$\begin{aligned} || x_{n+1}-x^{*}||^2\le & {} ||x_n-x^{*}||^2-\Big [\alpha _n-\mu _n M_0\Big ]||x_n-x^{*}||^2+\sigma _n. \end{aligned}$$
(3.5)

Since \(\mu _n=o(\alpha _n),~~\lambda _n=o(\alpha _n)\) and \(\xi _n=o(\alpha _n)\), we may assume without loss of generality that there exist constants \(k_0\in (0, 1)\) and \(M_2 > 0\) such that for all \(n\ge 1\),

$$\begin{aligned} \frac{\mu _n}{\alpha _n}M_0\le 1-k_0,~~\mathrm{and}~~\frac{\sigma _n}{\alpha _n}\le M_2. \end{aligned}$$

Thus, we obtain from (3.5) that

$$\begin{aligned} ||x_{n+1}-x^{*}||^2 \le ||x_n-x^{*}||^2-\alpha _nk_0 ||x_n-x^{*}||^2 + \sigma _n. \end{aligned}$$

By Lemma 2.5, we have that

$$\begin{aligned} ||x_n-x^{*}||^2 \le \max \{|| x_1-x^{*}||^2,(1+k_0)M_2\}. \end{aligned}$$

Therefore, \(\{x_n\}\) is bounded. Furthermore, the sequences \(\{y_n\}\) and \(\{u_n\}\) are bounded.

Observe that since T is a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive, then

$$\begin{aligned}&\!\!\!|| T^nx-x^{*}||^{2} \le || x-x^{*}||^{2}+k ||x-T^nx||^{2}+\mu _n\phi (||x-x^{*}||)+\xi _n\nonumber \\&\!\!\!\quad \Rightarrow \langle T^nx-x^{*},T^nx-x^{*}\rangle \le \langle x-x^{*},x-T^nx\rangle +\langle x-x^{*},T^nx-x^{*}\rangle + k || x-T^nx||^{2}\nonumber \\&\!\!\!\qquad +\mu _n\phi (||x-x^{*}||)+\xi _n \nonumber \\&\!\!\!\quad \Rightarrow \langle T^nx-x^{*},T^nx-x\rangle \le \langle x-x^{*},x-T^nx\rangle + k || x-T^nx||^{2} +\mu _n\phi (||x-x^{*}||)+\xi _n \nonumber \\&\!\!\!\quad \Rightarrow \langle T^nx-x,T^nx-x\rangle +\langle x-x^{*},T^nx-x\rangle \le \langle x-x^{*},x-T^nx\rangle + k || x-T^nx||^{2}\nonumber \\&\!\!\!\qquad +\mu _n\phi (||x-x^{*}||)+\xi _n \nonumber \\&\!\!\!\quad \Rightarrow (1-k)|| x-T^nx||^{2}\le 2\langle x-x^{*},x-T^nx\rangle +\mu _n\phi (||x-x^{*}||)+\xi _n. \end{aligned}$$
(3.6)

It follows from (2.2) and (3.6) that

$$\begin{aligned} || x_{n+1}-x^{*}||^{2}= & {} || (1-\beta _n)y_n+\beta _nT^ny_n-x^{*}||^{2}\nonumber \\= & {} || (y_n-x^{*})+\beta _n(Ty_n-y_n)||^{2}\nonumber \\= & {} || y_n-x^{*}||^{2}+\beta _n^{2}|| y_n-T^ny_n||^{2}-2\beta _n\langle y_n-x^{*},y_n-Ty_n\rangle \nonumber \\\le & {} || y_n\,{-}\,x^{*}||^{2}+\beta _n\Big [\beta _n\,{-}\,(1\,{-}\,k))|| y_n\,{-}\,T^ny_n||^{2}+\mu _n\phi (||y_n\,{-}\,x^{*}||)+\xi _n\Big ]\nonumber \\\le & {} || u_n\!-\!x^{*}||^{2}\,{+}\,\beta _n(\beta _n\!-\!(1\!-\!k))|| y_n\!-\!T^ny_n||^{2}\,{+}\,\beta _n\Big [\mu _n\phi (||y_n\!-\!x^{*}||)\,{+}\,\xi _n\Big ]\nonumber \\= & {} || (1-\alpha _n)x_n-x^{*}||^{2}+\beta _n(\beta _n-(1-k))|| y_n-T^ny_n||^{2}\nonumber \\&+\,\beta _n\Big [\mu _n\phi (||y_n-x^{*}||)+\xi _n\Big ]\nonumber \\= & {} (1-\alpha _n)^{2}|| x_n-x^{*}||^{2}+\alpha _n^{2}|| x^{*}||^{2}-\alpha _n(1-\alpha _n)\langle x_n-x^{*},x^{*}\rangle \nonumber \\&+\,\beta _n(\beta _n-(1-k))|| y_n-T^ny_n||^{2}+\beta _n\Big [\mu _n\phi (||y_n-x^{*}||)+\xi _n\Big ]\nonumber \\\le & {} || x_n-x^{*}||^{2}+\alpha _n^{2}|| x^{*}||^{2}-\alpha _n(1-\alpha _n)\langle x_n-x^{*},x^{*}\rangle \nonumber \\&+\,\beta _n(\beta _n-(1-k))|| y_n-T^ny_n||^{2}+\beta _n\Big [\mu _n\phi (||y_n-x^{*}||)+\xi _n\Big ]\nonumber \\\le & {} || x_n-x^{*}||^{2}+\alpha _n^{2}|| x^{*}||^{2}-\alpha _n(1-\alpha _n)\langle x_n-x^{*},x^{*}\rangle \nonumber \\&+\,\beta _n(\beta _n\!-\!(1-k))|| y_n-T^ny_n||^{2}+\beta _n\Big [\mu _n(M_1+M_0||y_n\!-\!x^{*}||^2)+\xi _n\Big ].\nonumber \\ \end{aligned}$$
(3.7)

Since \(\{x_n\}\) and \(\{y_n\}\) are bounded, \(\exists M>0\) such that

$$\begin{aligned} -(1-\alpha _n)\langle x_n-x^{*},x^{*}\rangle \le M,~~ \forall n\ge 1. \end{aligned}$$

Therefore,

$$\begin{aligned} \beta _n((1-k)-\beta _n)|| y_n-T^ny_n||^{2}\le & {} \alpha _n^{2}|| x^{*}||^{2}+\alpha _nM-|| x_{n+1}-x^{*}||^{2}+|| x_n-x^{*}||^{2}\nonumber \\&+\,\mu _n(M_1+M_0||y_n-x^{*}||^2)+\xi _n. \end{aligned}$$
(3.8)

Now we divide the rest of the proof into two cases.

\(\mathbf{Case1}\)

Assume that \(\{||x_n-x^{*}||\}\) is monotonically decreasing sequence. Then \(\{||x_n-x^{*}||\}\) is convergent, obviously

$$\begin{aligned} || x_{n+1}-x^{*}||^2-|| x_n-x^{*}||^2\rightarrow 0, n\rightarrow \infty . \end{aligned}$$
(3.9)

This together with (3.8) and the conditions that \(\alpha _n\rightarrow 0, \mu _n\rightarrow 0\) and \(\xi _n\rightarrow 0\) imply that

$$\begin{aligned} || y_n-T^ny_n||\rightarrow 0, n\rightarrow \infty . \end{aligned}$$

From (3.1) and (2.2), we have that

$$\begin{aligned}&\rho _n\Big (\frac{4h(u_n)}{h(u_n)+l(u_n)}-\rho _n\Big )\frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)} \le (1+M^*\lambda _n)||u_n-x^{*}||^2-||y_n-x^{*}||^2\\&\quad \le (1-\alpha _n)||x_n-x^{*}||^2+\alpha _n||x^{*}||^2-||x_{n+1}-x^{*}||^2+M^*\lambda _n||u_n-x^{*}||^2\\&\quad \le ||x_n-x^{*}||^2-||x_{n+1}-x^{*}||^2 +\alpha _n||x^{*}||^2+M^*\lambda _n||u_n-x^{*}||^2. \end{aligned}$$

Using conditions that \(\alpha _n\rightarrow 0\) and \(\lambda _n\rightarrow 0\), we have that

$$\begin{aligned} \rho _n\Big (\frac{4h(u_n)}{h(u_n)+l(u_n)}-\rho _n\Big )\frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)} \rightarrow 0,~~n\rightarrow \infty . \end{aligned}$$

Hence, we obtain

$$\begin{aligned} \frac{(h(u_n)+l(u_n))^2}{\theta ^2(u_n)} \rightarrow 0,~~n\rightarrow \infty . \end{aligned}$$
(3.10)

Consequently, we have

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }(h(u_n)+l(u_n))=0\Leftrightarrow \underset{n\rightarrow \infty }{\lim }h(u_n)=0~~\mathrm{and}~~\underset{n\rightarrow \infty }{\lim }l(u_n)=0, \end{aligned}$$

because \(\theta ^2(u_n)=||\nabla h(u_n)||^2+||\nabla l(u_n)||^2\) is bounded. This follows from the fact that \(\nabla h\) is Lipschitz continuous with constant \(||A||^2\), \(\nabla l\) is nonexpansive and \(\{u_n\}\) is bounded. More precisely, for any \(x^{*}\) which solves (1.9), we have

$$\begin{aligned} ||\nabla h(u_n)||=||\nabla h(u_n)-\nabla x^{*}|| \le ||A||^2||u_n-x^{*}||~~\mathrm{and} \end{aligned}$$
$$\begin{aligned} ||\nabla l(u_n)||=||\nabla l(u_n)-\nabla x^{*}||\le ||u_n-x^{*}||. \end{aligned}$$

Now, let z be a weak cluster point of \(\{u_n\}\), there exists a subsequence \(\{u_{n_j}\}\) which weakly converges to z. The lower-semicontinuity of h then implies that

$$\begin{aligned} 0\le h(z)\le \underset{j\rightarrow \infty }{\liminf }h(u_{n_j})=\underset{n\rightarrow \infty }{\lim }h(u_n)=0. \end{aligned}$$

That is, \(h(z)=\frac{1}{2}||(I- \mathrm{prox}_{\lambda g})Az||= 0\), i.e., Az is a fixed point of the proximal mapping of g or equivalently, \( 0\in \partial _{pg}(Az)\). In other words, Az is a minimizer of g.

Likewise, the lower-semicontinuity of l implies that

$$\begin{aligned} 0\le l(z)\le \underset{j\rightarrow \infty }{\liminf }l(u_{n_j})=\underset{n\rightarrow \infty }{\lim }l(u_n)=0. \end{aligned}$$

That is, \(l(z)=\frac{1}{2}||(I- \mathrm{prox}_{\tau \lambda f})z||= 0\), i.e., z is a fixed point of the proximal mapping of f or equivalently, \(0\in \partial f(z)\). In other words, z is a minimizer of f. Hence, \(z \in \Gamma \).

Next, we show that \(z \in F(T)\). Since \(x^{*}=\mathrm{prox}_{\lambda \tau _n f}( x^{*}-\mu _n A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\) and \(A^*(I-\mathrm{prox}_{\lambda g})A\) is Lipschitz with constant \(||A||^2\), we have from (2.2) that

$$\begin{aligned}&\!\!\!|| y_n-x^{*}||^2\\&\!\!\!\quad = ||\mathrm{prox}_{\lambda \tau _n f}( u_n-\mu _n A^*(I-\mathrm{prox}_{\lambda g})Au_n)-\mathrm{prox}_{\lambda \tau _n f}\left( x^{*}-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Ax^{*}\right) ||^2\nonumber \\&\!\!\!\quad \le \langle ( u_n-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Au_n)-\left( x^{*}-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Ax^{*}\right) , y_n-x^{*}\rangle \nonumber \\&\!\!\!\quad =\frac{1}{2}\Big [ ||( u_n-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Au_n)-\left( x^{*}-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Ax^{*}\right) ||^2 +||y_n-x^{*}||^2\\&\!\!\!\qquad -\,||( u_n-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Au_n)-\left( x^{*}-\tau _n A^*(I-\mathrm{prox}_{\lambda g})Ax^{*}\right) -(y_n-x^{*})||^2\Big ]\\&\!\!\!\quad \le \frac{1}{2}\Big [ (1+\tau _n||A||^2)^2||u_n-x^{*}||^2 +||y_n-x^{*}||^2-||u_n-y_n -\tau _n(A^*(I-\mathrm{prox}_{\lambda g})Au_n\\&\!\!\!\qquad -\, A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})||^2\Big ]\\&\!\!\!\quad =\frac{1}{2}\Big [ (1+\tau _n||A||^2)^2||u_n-x^{*}||^2 +||y_n-x^{*}||^2\\&\!\!\!\qquad -\,||u_n-y_n||^2+2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle \\&\!\!\!\qquad -\,\tau _n^2||A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})||^2\Big ]. \end{aligned}$$

Thus,

$$\begin{aligned} || y_n-x^{*}||^2\le & {} (1+\tau _n||A||^2)^2||x_n-x^{*}||^2-||u_n-y_n||^2\nonumber \\&+\,2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle \nonumber \\&-\,\tau _n^2||A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})||^2. \end{aligned}$$
(3.11)

We observe that

$$\begin{aligned} 0<\tau _n<4\frac{h(u_n)+l(u_n)}{\theta ^2(u_n)} \rightarrow 0,~~n\rightarrow \infty \end{aligned}$$

implies that \(\tau _n \rightarrow 0,~~n\rightarrow \infty \). Furthermore, we obtain from (3.11) and (2.2) that

$$\begin{aligned} ||u_n-y_n||^2\le & {} (1+\tau _n||A||^2)^2||u_n-x^{*}||^2-|| y_n-x^{*}||^2\nonumber \\&+\,2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle \nonumber \\= & {} ||u_n-x^{*}||^2+\tau _n||A||^2(2+\tau _n||A||^2)||u_n-x^{*}||^2-|| y_n-x^{*}||^2\nonumber \\&+\,2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle \nonumber \\\le & {} (1-\alpha _n)||x_n-x^{*}||^2+\alpha _n||x^{*}||^2-||x_{n+1}-x^{*}||^2\nonumber \\&+\,\tau _n||A||^2(2+\tau _n||A||^2)||u_n-x^{*}||^2+2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n\nonumber \\&-\, A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle \nonumber \\\le & {} ||x_n\!-\!x^{*}||^2\!-\!||x_{n+1}-x^{*}||^2+\alpha _n||x^{*}||^2\!+\!\tau _n||A||^2(2+\tau _n||A||^2)||u_n\!-\!x^{*}||^2\nonumber \\&+\,2\tau _n \langle u_n-y_n, A^*(I-\mathrm{prox}_{\lambda g})Au_n- A^*(I-\mathrm{prox}_{\lambda g})Ax^{*})\rangle . \end{aligned}$$
(3.12)

Since \(\tau _n \rightarrow 0,~~n\rightarrow \infty \) and \(\alpha _n \rightarrow 0,~~n\rightarrow \infty \), we obtain that

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }||u_n-y_n||=0. \end{aligned}$$

We observe that \(||u_n-x_n||\le \alpha _n||x_n||\rightarrow 0,~~n\rightarrow \infty \) and

$$\begin{aligned} ||x_n-y_n||\le ||u_n-x_n||+||u_n-y_n||\rightarrow 0,~~n\rightarrow \infty . \end{aligned}$$

Using Lemma 2.3 (i), we have that

$$\begin{aligned} ||x_{n+1}-x_n||^2= & {} ||x_n-y_n+y_n-x_{n+1}||^2 \nonumber \\\le & {} ||y_n-x_{n+1}||^2+2\langle x_n-y_n,x_n-x_{n+1}\rangle \nonumber \\\le & {} ||y_n-x_{n+1}||^2+2||x_n-y_n||||x_n-x_{n+1}||. \end{aligned}$$
(3.13)

Using (2.2) and (3.13), we have

$$\begin{aligned} ||y_n-T^ny_n||^2= & {} \frac{1}{\beta _n^2}||y_n-x_{n+1}||^2 \nonumber \\\ge & {} \frac{1}{\beta _n^2}\Big [||y_n-x_{n+1}||^2+2||x_n-y_n||||x_n-x_{n+1}||\Big ]. \end{aligned}$$
(3.14)

Since \(\lim \nolimits _{n\rightarrow \infty } ||y_n-T^ny_n||=0\) and \(0<\liminf \nolimits _{n\rightarrow \infty } \beta _n \le \limsup \nolimits _{n\rightarrow \infty }\beta _n<1-k\), then we have that

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }\Big [||y_n-x_{n+1}||^2+2||x_n-y_n||||x_n-x_{n+1}||\Big ]=0, \end{aligned}$$

from which we have

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }||x_n-x_{n+1}||=0. \end{aligned}$$

Consequently,

$$\begin{aligned} ||u_{n+1}-u_n||= & {} ||(1-\alpha _{n+1})x_{n+1}-(1-\alpha _n)x_n||\nonumber \\\le & {} |\alpha _{n+1}-\alpha _n|||x_{n+1}||+(1-\alpha _n)||x_{n+1}-x_n||\rightarrow 0,~~n\rightarrow \infty .\qquad \quad \end{aligned}$$
(3.15)

Now,

$$\begin{aligned} ||y_{n+1}-y_n||\le & {} ||y_{n+1}-u_{n+1}||+||u_n-y_n||+||u_{n+1}-u_n||. \end{aligned}$$
(3.16)

Using (3.15) and the fact that \(\lim \nolimits _{n\rightarrow \infty } ||u_n-y_n||=0\) in (3.16), we obtain

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }||y_{n+1}-y_n||=0. \end{aligned}$$

Using the fact that T is uniformly L-Lipschitzian, we have

$$\begin{aligned} ||Ty_n-y_n||\le & {} ||Ty_n-T^{n+1}y_n||+||T^{n+1}y_n-T^{n+1}y_{n+1}||\nonumber \\&+\,||T^{n+1}y_{n+1}-y_{n+1}||+||y_{n+1}-y_n||\nonumber \\\le & {} L||y_n-T^ny_n||+(L+1)||y_{n+1}-y_n||+||T^{n+1}y_{n+1}-y_{n+1}||\nonumber \\\le & {} L||y_n-T^ny_n||+(L+1)||A|| ||y_{n+1}-y_n||+||T^{n+1}y_{n+1}-y_{n+1}||.\nonumber \\ \end{aligned}$$
(3.17)

By using \(\lim \nolimits _{n\rightarrow \infty } ||y_{n+1}-y_n||=0\) and \(\lim \nolimits _{n\rightarrow \infty } ||y_n-T^ny_n||=0\) in (3.17), we obtain

$$\begin{aligned} || y_n-Ty_n||\rightarrow 0, n\rightarrow \infty . \end{aligned}$$
(3.18)

Using the fact that \(u_{n_j}\rightharpoonup z \in H_1\) and \(||u_n-y_n|| \rightarrow 0,~~n\rightarrow \infty \), we have that \(y_{n_j}\rightharpoonup z \in H_1\). Similarly, \(u_{n_j}\rightharpoonup z \in H_1\) since \(||u_n-x_n||\rightarrow 0,~~n\rightarrow \infty \). Using Lemma 2.4 and (3.18), we have that \(z \in F(T)\). Therefore, \(z \in \Gamma \cap F(T).\)

From (2.2), we have

$$\begin{aligned} || x_{n+1}-x^{*}||^2= & {} ||(1-\beta _n)y_n+\beta _n T^ny_n-x^{*}||^2 \nonumber \\= & {} || (1-\beta _n)(y_n-x^{*})+\beta _n(T^ny_n-x^{*})||^2 \nonumber \\= & {} (1-\beta _n)^2|| y_n-x^{*}||^2 +\beta _n^2|| T^ny_n-x^{*}||^2\nonumber \\&+\, 2\beta _n(1-\beta _n)\langle y_n-x^{*},T^ny_n-x^{*}\rangle \nonumber \\\le & {} (1-\beta _n)^2|| y_n-x^{*}||^2 +\beta _n^2[|| y_n-x^{*}||^2 +k|| y_n-T^ny_n||^2] \nonumber \\&+\,\beta _n^2\mu _n\phi (||y_n-x^{*}||)+\beta _n^2\xi _n+2\beta _n(1-\beta _n)[|| y_n-x^{*}||^2\nonumber \\&-\,\frac{1-k}{2}|| y_n-T^ny_n||^2]\nonumber \\= & {} (1-2\beta _n+\beta _n^2)|| y_n-x^{*}||^2+ \beta _n^2[|| y_n-x^{*}||^2 +k|| y_n-T^ny_n||^2] \nonumber \\&+\,2\beta _n|| y_n-x^{*}||^2-2\beta _n^2|| y_n-x^{*}||^2 -\beta _n(1-\beta _n)(1-k)|| y_n-T^ny_n||^2 \nonumber \\&+\,\beta _n^2\mu _n(M+M_0||y_n-x^{*}||^2)+\beta _n^2\xi _n \nonumber \\= & {} || y_n-x^{*}||^2+[\beta _n^2k -\beta _n(1-\beta _n)(1-k)]|| y_n-T^ny_n||^2\nonumber \\&+\,\beta _n^2\mu _n(M+M_0||y_n-x^{*}||^2)+\beta _n^2\xi _n \nonumber \\= & {} || y_n-x^{*}||^2 + \beta _n[k + \beta _n -1]|| y_n-T^ny_n||^2 \nonumber \\\le & {} || y_n-x^{*}||^2+\beta _n^2\mu _n(M+M_0||y_n-x^{*}||^2)+\beta _n^2\xi _n \nonumber \\= & {} || y_n-x^{*}||^2+a_n, \end{aligned}$$
(3.19)

where \(a_n=\beta _n^2\mu _n(M+M_0||y_n-x^{*}||^2)+\beta _n^2\xi _n.\)

Now, from (3.1), (3.19) and Lemma 2.3 (ii), we have

$$\begin{aligned} ||x_{n+1}-x^{*}||^2\le & {} ||y_n-x^{*}||^2 +a_n\le ||u_n-x^{*}||^2+a_n\nonumber \\= & {} ||(1-\alpha _n)(x_n-x^{*})-\alpha _nx^{*}||^2+a_n \nonumber \\\le & {} (1-\alpha _n)||x_n-x^{*}||^2-2\alpha _n \langle u_n-x^{*},x^{*}\rangle +a_n. \end{aligned}$$
(3.20)

It is clear that

$$\begin{aligned} \underset{n\rightarrow \infty }{\limsup }-2 \langle u_n-x^{*},x^{*}\rangle= & {} \underset{j\rightarrow \infty }{\lim }-2 \langle u_{n_j}-x^{*},x^{*}\rangle -2 \langle z-x^{*},x^{*}\rangle \le 0, \end{aligned}$$

since \(\{u_n\}\) converges weakly to z and \(x^{*}\) is the minimum-norm solution (i.e., \(x^*=P_{\Gamma \cap F(T)}0\)). Also, we observe that \(\sum \nolimits _{n=1}^{\infty } a_n <\infty .\) Now, using Lemma 2.6 in (3.20), we have \(||x_n-x^{*}||\rightarrow 0\). That is, \( x_n\rightarrow x^{*}, n\rightarrow \infty \).

Case 2

Assume that \(\{|| x_n-x^{*}||\}\) is not monotonically decreasing sequence. Set \(\Gamma _{n}=||x_{n}-x^{*}||^{2}\) and let \(\tau :\mathbb {N}\rightarrow \mathbb {N}\) be a mapping for all \(n\ge n_{0}\) (for some \(n_{0}\) large enough)by

$$\begin{aligned} \tau (n):=\max \{k\in \mathbb {N}:k\le n,\Gamma _{k}< \Gamma _{k+1}\} \end{aligned}$$

Clearly, \(\tau \) is a non decreasing sequence such that \(\tau (n)\rightarrow \infty \) as \(n \rightarrow \infty \) and

$$\begin{aligned} \Gamma _{\tau (n)+1}-\Gamma _{\tau (n)}\ge 0, ~~\forall n\ge n_{0}. \end{aligned}$$

From (3.18), it is easy to see that

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }||y_{\tau (n)}-Ty_{\tau (n)}||=0. \end{aligned}$$

Furthermore, we can show that

$$\begin{aligned} \underset{n\rightarrow \infty }{\lim }h(x_{\tau (n)})=0~~\mathrm{and}~~\underset{n\rightarrow \infty }{\lim }l(x_{\tau (n)})=0. \end{aligned}$$

By similar argument as above in Case 1, we conclude immediately that \(x_{\tau (n)}, y_{\tau (n)}\) and \(u_{\tau (n)}\) weakly converge to z as \(\tau (n)\rightarrow \infty .\) At the same time, from (3.20), we note that, for all \(n\ge n_{0},\)

$$\begin{aligned} 0\le & {} ||x_{\tau (n)+1}-x^{*}||^{2}-||x_{\tau (n)}-x^{*}||^{2}\\\le & {} \alpha _{\tau (n)}[-2\langle u_{\tau (n)}-x^{*},x^{*}\rangle - ||x_{\tau (n)}-x^{*}||^{2}]+a_{\tau (n)}, \end{aligned}$$

which implies

$$\begin{aligned} ||x_{\tau (n)}-x^{*}||^{2}\le -2\langle u_{\tau (n)}-x^{*},x^{*}\rangle +a_{\tau (n)}. \end{aligned}$$

Hence, we deduce that

$$\begin{aligned} \lim _{n\rightarrow \infty }||x_{\tau (n)}-x^{*}||=0. \end{aligned}$$

Therefore,

$$\begin{aligned} \lim _{n\rightarrow \infty }\Gamma _{\tau (n)}=\lim _{n\rightarrow \infty }\Gamma _{\tau (n)+1}= 0. \end{aligned}$$

Furthermore, for \(n\ge n_{0}\), it is easy to see that \(\Gamma _{\tau (n)}\le \Gamma _{\tau (n)+1}\) if \(n\ne \tau (n)\) (that is \(\tau (n)<n\)), because \(\Gamma _{j}\ge \Gamma _{j+1}\) for \(\tau (n)+1\le j\le n.\) As a consequence, we obtain for all \(n\ge n_{0}\),

$$\begin{aligned} 0\le \Gamma _{n}\le \max \{\Gamma _{\tau (n)},\Gamma _{\tau (n)+1}\}=\Gamma _{\tau (n)+1}. \end{aligned}$$

Hence, \(\lim \Gamma _{n}=0\), that is, \(\{x_n\}\) converges strongly to \(x^{*}\). This completes the proof. \(\square \)

4 Applications

Applying Theorem 3.1 to the case where \(f =\delta _C\), \(g =\delta _Q\) the indicator functions of two nonempty, closed and convex sets CQ of \(H_1\) and \(H_2\) respectively, we have the following result.

Theorem 4.1

Let C, Q be nonempty, closed and convex subset of real Hilbert spaces \(H_1\) and \(H_2\) respectively. Let \(T:C\rightarrow C\) is a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping such that \(F(T)\ne \emptyset \). Assume that problem (1.2) is consistent (i.e., \(\Gamma \ne \emptyset \)) and \(\Gamma \cap F(T)\ne \emptyset \). Let \(\{\alpha _n\}\) and \(\{\beta _n\}\) be sequences in (0, 1). If the parameters satisfy the following conditions

  1. (a)

    \(\sum \limits _{n=1}^{\infty }\mu _n<\infty ; \sum \limits _{n=1}^{\infty } \xi _n<\infty ;\)

  2. (b)

    \(\mu _n=o(\alpha _n); \xi _n=o(\alpha _n); \underset{n\rightarrow \infty }{\lim }\alpha _n =0; \sum \limits _{n=1}^{\infty } \alpha _n =\infty ;\)

  3. (c)

    \(\epsilon \le \rho _n \le \frac{4 h(u_n)}{h(u_n)+l(u_n)}-\epsilon \) for some \(\epsilon >0\);

  4. (d)

    \(0<\underset{n\rightarrow \infty }{\liminf }\beta _n \le \underset{n\rightarrow \infty }{\limsup }\beta _n<1-k\);

  5. (e)

    there exist constants \(M_0>0, K_1 > 0\) such that \(\phi (t)\le M_0 t^2, \forall t >K_1\);

then sequence \(\{x_n\}\) generated by

$$\begin{aligned} \left\{ \begin{array}{l} u_n=(1-\alpha _n)x_n,\\ y_n =P_C(u_n-\tau _n A^*(I-P_Q)Au_n),\\ x_{n+1} =(1-\beta _n)y_n+ \beta _nT^ny_n,~~n \ge 1, \end{array} \right. \end{aligned}$$
(4.1)

converges strongly to \(x^*\in \Gamma \cap F(T)\) which is also the minimum-norm solution (i.e., \(x^*\in \Gamma \cap F(T)\) and \(||x^*|| =\min \{ ||x||: x \in \Gamma \cap F(T)\}\)).

Proof

Take \(f =\delta _C\) and \(g =\delta _Q\) in Theorem 3.1. Then we have \(\mathrm{prox}_{\lambda f}=P_C\) and \(\mathrm{prox}_{\lambda g}=P_Q\). Furthermore, we see that algorithm reduces to (2.2) reduces to (4.1) and the conclusion of Theorem 4.1. \(\square \)

We next apply our results to split equilibrium problem and fixed point problem. Let CQ be nonempty, closed and convex subsets of \(H_1\) and \(H_2\) respectively. Let f be a bifunction of \(C \times C\) into \(\mathbb {R}\) and g a bifunction of \(Q \times Q\) into \(\mathbb {R}\). Suppose \(A:H_1\rightarrow H_2\) is a bounded linear operator. Let us consider the following split equilibrium problem. The split equilibrium problem is to find \(x^{*} \in C\) such that

$$\begin{aligned} f(x^{*}, y)\ge 0,~~\forall y \in C \end{aligned}$$
(4.2)

and

$$\begin{aligned} g(Ax^{*}, y)\ge 0,~~\forall y \in Q. \end{aligned}$$
(4.3)

We shall denote the solutions set of (4.2)-(4.3) by \(\Gamma \). In order to solve the split equilibrium problem for a bifunctions \(f:C \times C\rightarrow \mathbb {R}\) and \(g:Q \times Q\rightarrow \mathbb {R}\), let us assume that f and g satisfy the following conditions:

  1. (A1)

    \(f(x,x)=0\) for all \(x \in C\) and \(g(x,x)=0\) for all \(x \in Q\);

  2. (A2)

    f and g are monotone, i.e., \(f(x,y)+f(y,x) \le 0\) for all \(x, y, \in C\) and \(g(x,y)+g(y,x) \le 0\) for all \(x, y, \in Q\);

  3. (A3)

    for each \(x, y \in C,~~\lim \nolimits _{t\rightarrow 0} f(tz+(1-t)x, y) \le f(x, y)\) and for each \(x, y \in Q,~~\lim \nolimits _{t\rightarrow 0} g(tz+(1-t)x, y) \le g(x, y)\);

  4. (A4)

    for each \(x \in C,~~y\mapsto f(x,y)\) is convex and lower semicontinuous and for each \(x \in Q,~~y\mapsto g(x,y)\) is convex and lower semicontinuous.

The following lemma follows from [8].

Lemma 4.2

(Combettes and Hirstoaga, [8]) Let CQ be nonempty, closed and convex subsets of \(H_1\) and \(H_2\) respectively. Assume that fg satisfy (A1)-(A4). For \(r>0\), define mappings \(T^f_r:H_1\rightarrow C\) and \(T^g_r:H_2\rightarrow Q\) as follows:

$$\begin{aligned} T^f_r (x)=\{z \in C:f(z,y)+\frac{1}{r} \langle y-z, z-x \rangle \ge 0, \forall y \in C\} \end{aligned}$$

and

$$\begin{aligned} T^g_r (x)=\{z \in Q:g(z,y)+\frac{1}{r} \langle y-z, z-x \rangle \ge 0, \forall y \in C\}. \end{aligned}$$

Then, the following assertions:

  1. 1.

    \(T^f_r\) and \(T^g_r\) are single-valued;

  2. 2.

    \(T^f_r\) and \(T^g_r\) are firmly nonexpansive-type mapping, i.e., for any \(x, y \in H_1\),

    $$\begin{aligned} ||T^f_r x-T^f_r y||^2 \le \langle T^f_r x- T^f_r y, x-y \rangle ; \end{aligned}$$
  3. 3.

    \(F(T^f_r)=EP(f)\) and \(F(T^g_r)=EP(g)\);

  4. 4.

    EP(f) and EP(g) are closed and convex.

Now, we prove the following theorem.

Theorem 4.3

Let \(T:C\rightarrow C\) is a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping such that \(F(T)\ne \emptyset \). Let f be a bifunction from \(C \times C\) and g a bifunction from \(Q \times Q\) both satisfying (\(A1) - (A4\)) such that \(\Gamma \ne \emptyset \)) and \(\Gamma \cap F(T)\ne \emptyset \). Let \(\{\alpha _n\}\) and \(\{\beta _n\}\) be sequences in (0, 1). If the parameters satisfy the following conditions

  1. (a)

    \(\sum \limits _{n=1}^{\infty }\mu _n<\infty ; \sum \limits _{n=1}^{\infty } \xi _n<\infty ;\)

  2. (b)

    \(\mu _n=o(\alpha _n); \xi _n=o(\alpha _n); \underset{n\rightarrow \infty }{\lim }\alpha _n =0; \sum \limits _{n=1}^{\infty } \alpha _n =\infty ;\)

  3. (c)

    \(\epsilon \le \rho _n \le \frac{4 h(u_n)}{h(u_n)+l(u_n)}-\epsilon \) for some \(\epsilon >0\);

  4. (d)

    \(0<\underset{n\rightarrow \infty }{\liminf }\beta _n \le \underset{n\rightarrow \infty }{\limsup }\beta _n<1-k\);

  5. (e)

    there exist constants \(M_0>0, K_1 > 0\) such that \(\phi (t)\le M_0 t^2, \forall t >K_1\);

then sequence \(\{x_n\}\) generated by \(x_1 \in C\);

$$\begin{aligned} \left\{ \begin{array}{l} u_n=(1-\alpha _n)x_n,\\ y_n =T^f_r(u_n-\tau _n A^*(I-T^g_r)Au_n),\\ x_{n+1} =(1-\beta _n)y_n+ \beta _nT^ny_n,~~n \ge 1, \end{array} \right. \end{aligned}$$
(4.4)

converges strongly to \(x^*\in \Gamma \cap F(T)\) which is also the minimum-norm solution (i.e., \(x^*\in \Gamma \cap F(T)\) and \(||x^*|| =\min \{ ||x||: x \in \Gamma \cap F(T)\}\)).

Proof

Replace the proximal mappings of the convex functions f and g in Theorem 3.1 by the resolvent operators associated to the two monotone equilibrium bifunctions \(T^f_r\) and \(T^g_r\). Hence, we have the desired result. \(\square \)

Remark 1

The following are our contributions in this paper.

  1. 1.

    We obtain strong convergence result concerning convex split feasibility problem and fixed point problem in real Hilbert spaces. We recall that Moudafi and Thakur [13] obtained weak convergence result for split feasibility problem alone and thus our result improves on and extends the results of Moudafi and Thakur [13].

  2. 2.

    It is worth mentioning here that our result in this paper is more applicable than the result of Moudafi and Thakur [13] in the sense that our result can be applied to finding an approximate common solution to proximal split feasibility problem and fixed point problem for a uniformly L-Lipschitzian continuous and \((k, \{\mu _n\}, \{\xi _n\}, \phi )\)-total asymptotically strict pseudocontractive mapping.

  3. 3.

    In all our results in this paper, our iterative scheme is proposed with a way of selecting the step-size such that the implementation of our algorithm does not need any prior information about the operator norm ||A|| because the calculation or at least an estimate of the operator norm ||A|| is very difficult, if not an impossible task. Therefore, improved on the results of Chang et al. [6], Yang et al. [22], Cholamjiak and Shehu [7] and other related works.

  4. 4.

    Our iterative algorithm in this paper appears more efficient and implementable. Our algorithm appears simpler than the “CQ” algorithm used [10] and other related papers for similar problems. Furthermore, our iterative scheme gives strong convergence without imposing any extra compactness type condition ( like semi-compactness) on the mapping T. This compactness condition appears strong as only few mappings are semi-compact. Therefore, we improved on the results of Chang et al. [6], Qin et al. [14], Ding and Quan [9] and other related results.