1 Introduction and preliminaries

Let C be a nonempty subset of a metric space \((X, d)\).

Recall that a map** \(T : C \to X\) is said to be:

(i) nonexpansive if \(d(T x, T y) \le d(x, y)\) for all \(x, y\in C\);

(ii) asymptotically nonexpansive [1] if there exists a sequence \(\{k_{n}\} \subset [1, \infty )\) with \(\lim_{n\to \infty} k_{n} = 1\) such that \(d(T^{n} x, T^{n} y) \le k_{n} d(x, y)\) for all \(x, y \in C\) and \(n \in \mathscr{N}\), where \(\mathscr{N}\) denotes the set of positive integers. The class of asymptotically nonexpansive map**s includes a class of nonexpansive map**s as a proper subclass.

(iii) In 1993, Bruck, Kuczumow, and Reich [2] introduced the concept of asymptotically nonexpansive map** in the intermediate sense. A map** \(T : C \to C\) is said to be asymptotically nonexpansive in the intermediate sense if T is uniformly continuous and the following inequality holds:

$$ \limsup_{n \to \infty} \sup_{x,y\in C}\bigl\{ d \bigl(T^{n} x, T^{n} y\bigr)- d (x, y) \bigr\} \le 0. $$
(1.1)

It is easy to know that the class of asymptotically nonexpansive map**s in the intermediate sense is more general than the class of asymptotically nonexpansive map**s.

Definition 1.1

([3])

A map** \(T : C \to C\) is said to be \((\{\mu _{n}\}, \{\nu _{n}\}, \zeta )\)-total asymptotically nonexpansive if there exist nonnegative sequences \(\{\mu _{n}\}\), \(\{\nu _{n}\}\) with \(\mu _{n} \to 0\), \(\nu _{n} \to 0\) and a strictly increasing continuous function \(\zeta : [0, \infty ) \to [0, \infty )\) with \(\zeta (0) = 0\) such that

$$ d\bigl(T^{n} x, T^{n} y\bigr) \le d(x,y) + \nu _{n} \zeta \bigl(d(x,y)\bigr) + \mu _{n}, \quad \forall n \ge 1, x, y \in C. $$
(1.2)

The concept of total asymptotically nonexpansive map**s is more general than that of asymptotically nonexpansive map**s in the intermediate sense. In fact, if \(T : C \to C\) is an asymptotically nonexpansive map** in the intermediate sense, denote by \(\mu _{n} = \max \{\sup_{x,y \in C} (d(T^{n} x, T^{n} y) - d(x, y)), 0\}\). Then \(\mu _{n} \ge 0\), \(\lim_{n \to \infty}\mu _{n} = 0\), and

$$ d\bigl(T^{n} x, T^{n} y\bigr) \le d(x, y) + \mu _{n},\quad \forall x, y, \in C, n \ge 1. $$
(1.3)

Taking \(\{\nu _{n} = 0\}\), \(\zeta = t\), \(t \ge 0\), then (1.3) can be written as

$$ d\bigl(T^{n} x, T^{n} y\bigr) \le d(x,y) + \nu _{n} \zeta \bigl(d(x,y)\bigr) + \mu _{n}, \quad \forall n \ge 1, x, y \in C, $$

i.e., T is a total asymptotically nonexpansive map**.

Definition 1.2

A map** \(T: C \to C\) is said to be uniformly L-Lipschitzian if there exists a constant \(L > 0\) such that

$$ d\bigl(T^{n} x, T^{n} y\bigr) \le L d(x, y) \quad \forall x, y, \in C \text{ and } \forall n \ge 1. $$

In recent years, CAT(0) spaces (the precise definition of a CAT(0) space is given below) have attracted the attention of many authors because they have played a very important role in different aspects of geometry [4]. Kirk [5, 6] showed that a nonexpansive map** defined on a bounded closed convex subset of a complete CAT(0) space has a fixed point.

In 2012, Chang et al. [7] studied the demiclosedness principle and Δ-convergence theorems for total asymptotically nonexpansive map**s in the setting of CAT(0) spaces. Since then the convergence of several iteration procedures for this type of map**s has been rapidly developed, and many of articles have appeared (see, e.g., [817]). In 2013, under some suitable assumptions, Karapinar et al. [9] obtained the demiclosedness principle, fixed point theorems, and convergence theorems for the following iteration:

Let C be a nonempty closed convex subset of a CAT(0) space X and \(T : C \to C\) be a total asymptotically nonexpansive map**. Given \(x_{1} \in C\), let \(\{x_{n}\} \subset C\) be defined by

$$ x_{n+1} = (1 - \alpha _{n})x_{n} \oplus \alpha _{n} T^{n} \bigl((1 -\beta _{n})x_{ \oplus } \beta _{n} T^{n} (x_{n})\bigr), \quad n \in \mathscr{N}, $$

where \(\{\alpha _{n}\}\) and \(\{\beta _{n}\}\) are sequences in \([0; 1]\).

It is well known that any \(\operatorname{CAT} (\kappa )\) space is a \(\operatorname{CAT} (\kappa _{1})\) space for \(\kappa _{1} \ge \kappa \). Thus, all results for CAT(0) spaces immediately apply to any \(\operatorname{CAT} (\kappa )\) space with \(\kappa \le 0\).

Very recently, Panyanak [10] obtained the demiclosedness principle, fixed point theorems, and convergence theorems for total asymptotically nonexpansive map**s on \(\operatorname{CAT} (\kappa )\) space with \(\kappa > 0\), which generalize the results of Chang et al. [7], Tang et al. [8], Karapinar et al. [9].

Motivated by the work going on in this direction, in this paper we aim to study the strong convergence of a sequence generated by an infinite family of total asymptotically nonexpansive map**s in \(\operatorname{CAT} (\kappa )\) spaces with \(\kappa > 0\). Our results are new, they extend and improve the corresponding results of Chang et al. [7], Tang et al. [8], Karapinar et al. [9], Panyanak [10], Hea et al. [18], and many others.

2 Preliminaries

In this section, we first recall some definitions, notations, and conclusions that will be needed in our paper.

Let \((X, d)\) be a metric space. A geodesic path joining \(x \in X\) to \(y \in X\) is a map** c from a closed interval \([0, l] \subset \mathscr{R}\) to X such that \(c(0) = x\), \(c(l) = y\), and \(d(c(t); c(t^{\prime })) = |t - t^{\prime }|\) for all \(t, t^{\prime } \in [0; l]\). In particular, c is an isometry and \(d(x, y) = l\). The image \(c([0, l])\) of c is called a geodesic segment joining x and y. When it is unique, this geodesic segment is denoted by \([x, y]\). This means that \(z \in [x, y]\) if and only if there exists \(\alpha \in [0; 1]\) such that

$$ d(x, z) = (1 - \alpha ) d(x,y), \quad \text{and} \quad d(y, z) = \alpha d(x; y). $$

In this case, we write \(z = \alpha x \oplus (1-\alpha )y\).

A metric space \((X, d)\) is said to be a geodesic space (D-geodesic space) if every two points of X (every two points of distance smaller than D) are joined by a geodesic, and X is said to be uniquely geodesic (D-uniquely geodesic) if there is exactly one geodesic joining x and y for each \(x, y \in X\) (for \(x, y \in X\) with \(d(x, y) < D\)). A subset C of X is said to be convex if C includes every geodesic segment joining any two of its points.

Now we introduce the concept of model spaces \(M^{n}_{\kappa}\). For more details on these spaces, the reader is referred to [4, 14]. Let \(n \in \mathscr{N}\). We denote by \(E^{n}\) the metric space \(\mathscr{R}^{n}\) endowed with the usual Euclidean distance. We denote by \((\cdot | \cdot )\) the Euclidean scalar product in \(\mathscr{R}^{n}\), that is,

$$ (x|y) = x_{1}y_{1} + \cdots + x_{n} y_{n} \quad \text{where } x = (x_{1}, \ldots x_{n}), y = (y_{1},\ldots y_{n}). $$

Let \(\mathscr{S}^{n}\) denote the n-dimensional sphere defined by

$$ \mathscr{S}^{n} = \bigl\{ x = (x_{1}, \ldots , x_{n+1})\in \mathscr{R}^{n+1} : (x|x) = 1\bigr\} $$

with metric

$$ d_{\mathscr{S}^{n}}(x, y) = \arccos(x|y), \quad x, y\in \mathscr{S}^{n}. $$

Let \(E^{n;1}\) denote the vector space \(\mathscr{R}^{n+1}\) endowed with the symmetric bilinear form that associates to vectors \(u = (u_{1}, \ldots , u_{n+1})\) and \(v = (v_{1}, \ldots , v_{n+1})\), and the real number \(\langle u|v \rangle \) is defined by

$$ \langle u|v \rangle = u- u_{n+1} v_{n+1} + \sum _{i=1}^{n} u_{i} v_{i}. $$

Let \(\mathscr{H}^{n}\) denote the hyperbolic n-space defined by

$$ \mathscr{H}^{n} = \bigl\{ u = (u_{1}, \ldots , u_{n+1)} \in E^{n;1} : \langle u|u \rangle = -1, u_{n+1} > 0\bigr\} $$

with metric \(d_{\mathscr{H}^{n}}\) such that

$$ \cosh d_{\mathscr{H}^{n}}(x,y) = - \langle x| y \rangle , \quad x, y \in \mathscr{H}^{n}. $$

Definition 2.1

Given \(\kappa \in \mathscr{R}\), we denote by \(M^{n}_{\kappa}\) the following metric spaces:

(i) if \(\kappa = 0\), then \(M^{n}_{0}\) is the Euclidean space \(E^{n}\);

(ii) if \(\kappa > 0\), then \(M^{n}_{\kappa}\) is obtained from the spherical space \(\mathscr{S}^{n}\) by multiplying the distance function by the constant \(\frac{1}{\sqrt{\kappa}}\);

(iii) if \(\kappa < 0\), then \(M^{n}_{\kappa}\) is obtained from the hyperbolic space \(\mathscr{H}^{n}\) by multiplying the distance function by the constant \(\frac{1}{\sqrt{ - \kappa}}\).

A geodesic triangle \(\Delta (x, y, z)\) in a geodesic space \((X, d)\) consists of three points x, y, z in X (the vertices of Δ) and three geodesic segments between each pair of vertices (the edges of Δ). A comparison triangle for a geodesic triangle \(\Delta (x; y; z)\) in \((X, d)\) is a triangle \(\bar{\Delta}(\bar{x}, \bar{y},\bar{z})\) in \(M^{2}_{\kappa}\) such that

$$ d(x; y) = d_{M^{2}_{\kappa}}(\bar{x}, \bar{y}), \qquad d(y,z) = d_{M^{2}_{ \kappa}}( \bar{y}, \bar{z}), \qquad d(z,x) = d_{M^{2}_{\kappa}}(\bar{z}, \bar{x}). $$

If \(\kappa < 0\), then such a comparison triangle always exists in \(M^{2}_{\kappa}\). If \(\kappa > 0\), then such a triangle exists whenever \(d(x, y) + d(y, z) + d(z, x) < 2D_{\kappa}\), where \(D_{\kappa}= \frac{\pi}{\sqrt{\kappa}}\).

A point \(\bar{p} \in [\bar{x}, \bar{y}]\) is called a comparison point for \(p \in [x, y]\) if \(d(x, p) = d_{M^{2}_{\kappa}}(\bar{x}, \bar{p})\).

A geodesic triangle \(\Delta (x, y, z)\) in X is said to satisfy the CAT(κ) inequality if for any \(p, q \in \Delta (x, y, z)\) and for their comparison points \(\bar{p}, \bar{q} \in \Delta (\bar{x}, \bar{y}, \bar{z})\), one has

$$ d(p, q) \le d_{M^{2}_{\kappa}} (\bar{p},\bar{q}). $$

Definition 2.2

A metric space \((X, d)\) is called a CAT(0) space if X is a geodesic space such that all of its geodesic triangles satisfy the \(\operatorname{CAT} (\kappa )\) inequality.

If \(\kappa > 0\), then X is called a \(\operatorname{CAT} (\kappa )\) space if X is \(D_{\kappa}\)-geodesic and any geodesic triangle \(\Delta (x, y, z)\) in X with \(d(x, y) + d(y, z) + d(z, x) < 2D_{\kappa}\) satisfies the \(\operatorname{CAT} ( \kappa )\) inequality.

Definition 2.3

A geodesic space \((X, d)\) is said to be R-convex with \(R \in (0, 2]\) (see [16]) if for any three points \(x, y, z \in X\), we have

$$ d^{2}\bigl(x, (1-\alpha )y \oplus \alpha z\bigr)\le (1-\alpha ) d^{2}(x,y) + \alpha d^{2}(x,z) - \frac{R}{2}\alpha (1- \alpha )d^{2}(y, z). $$
(2.1)

Notice that if \((X,d)\) is a geodesic space, then the following statements are equivalent:

(i) \((X,d)\) is a CAT(0) space;

(ii) \((X, d)\) is R-convex with \(R=2\), i.e., it satisfies the following inequality:

$$ d^{2}\bigl(x, (1-\alpha )y \oplus \alpha z\bigr)\le (1-\alpha ) d^{2}(x,y) + \alpha d^{2}(x,z) - \alpha (1-\alpha )d^{2}(y; z) $$
(2.2)

for all \(\alpha \in (0, 1]\) and \(x, y, z \in X\).

The following lemma is a consequence of Proposition 3.1 in [19].

Lemma 2.4

Let \(\kappa > 0\) and \((X, d)\) be a CAT(κ) space with \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). Then \((X,d)\) is R-convex with \(R = (\pi - 2\epsilon ) \tan(\epsilon )\).

Lemma 2.5

([20, page 176])

Let \(\kappa > 0\) and \((X,d)\) be a complete CAT(κ) space with \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). Then

$$ d\bigl((1 -\alpha )x \oplus \alpha y, z\bigr) \le (1 -\alpha ) d(x, z) + \alpha d(y, z) $$
(2.3)

for all \(x, y, z \in X\) and \(\alpha \in [0, 1]\).

We now collect some elementary facts about \(\operatorname{CAT}(\kappa )\) spaces, \(\kappa > 0\).

Let \(\{x_{n}\}\) be a bounded sequence in a CAT(κ) space \((X, d)\). For \(x \in X\), we set

$$ r\bigl(x, \{x_{n}\}\bigr) = \lim \sup_{n \to \infty} d(x, x_{n}). $$

The asymptotic radius \(r(\{x_{n}\})\) of \(\{x_{n}\}\) is given by

$$ r\bigl(\{x_{n}\}\bigr) = \inf \bigl\{ r\bigl(x, \{x_{n}\} \bigr) : x \in X\bigr\} . $$

The asymptotic radius \(r(\{x_{n}\})\) of \(\{x_{n}\}\) is given by

$$ r\bigl(\{x_{n}\}\bigr) = \inf \bigl\{ r\bigl(x, \{x_{n}\} \bigr) : x \in X\bigr\} . $$
(2.4)

The asymptotic center \(A(\{x_{n}\})\) of \(\{x_{n}\}\) is the set

$$ A\bigl(\{x_{n}\}\bigr) = \bigl\{ x \in X: r\bigl(x, \{x_{n}\} \bigr) = r\bigl(\{x_{n}\}\bigr)\bigr\} . $$
(2.5)

It is known from Proposition 4.1 of [21] that in a CAT(κ) space X with \(\operatorname{diam}(X) < \frac{\pi}{2\sqrt{\kappa}}\), \(A(\{x_{n}\})\) consists of exactly one point.

We now give the concept of Δ-convergence and collect some of its basic properties.

Definition 2.6

([22, 23])

A sequence \(\{x_{n}\}\) in X is said to Δ-converge to \(x \in X\) if x is the unique asymptotic center of \(\{u_{n}\}\) for every subsequence \(\{u_{n}\}\) of \(\{x_{n}\}\). In this case we write \(\Delta -\lim_{n \to \infty} x_{n} = x\), and x is called the Δ-limit of \(\{x_{n}\}\).

Lemma 2.7

Let \((X, d)\) be a complete CAT(κ) space with \(\kappa > 0\) and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). Then the following statements hold:

  1. (i)

    [17, Corollary 4.4] Every sequence in X has a Δ-convergence subsequence;

  2. (ii)

    [17, Proposition 4.5] If \(\{x_{n}\} \subset X\) and \(\Delta -\lim x_{n} = x\), then

    $$ x\in \bigcap_{n=1}^{\infty }\overline{ \mathrm{conv}}\{x_{n}, x_{n+1}, \ldots \}, $$

where \(\overline{\mathrm{conv}}(A) = \bigcap \{B : B \supseteq A \textit{and} B \textit{is} \textit{closed} \textit{and} \textit{convex}\}\).

By the uniqueness of asymptotic centers, we can obtain the following lemma (cf. [24, Lemma 2.8]).

Lemma 2.8

Let \((X,d)\) be a complete CAT(κ) space with \(\kappa > 0\) and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). If \(\{x_{n}\}\) is a sequence in X with \(A(\{x_{n}\}) = \{x\}\) and if \(\{u_{n}\}\) is a subsequence of \(\{x_{n}\}\) with \(A(\{u_{n}\}) = \{u\}\) and the sequence \(\{d(x_{n}, u)\}\) converges, then \(x = u\).

In the sequel, we use \(F(T)\) to denote the fixed point set of a map** T.

Definition 2.9

([25])

(1) A triple \((X, d, W)\) is called a hyperbolic space if \((X, d)\) is a metric space and \(W : X \times X \times [0, 1] \to X\) is a map** such that \(\forall x, y, z, w \in X\), \(\alpha , \beta , \in [0, 1]\), the following hold:

(W1) \(d(z, W(x, y, \alpha ) \le \alpha d(z, x) + (1-\alpha )d(z, y)\);

(W2) \(d(W(x, y, \alpha ), W(x, y, \beta ) = |\alpha - \beta | d(x, y)\);

(W3) \(W(x, y, \alpha ) = W(y, x, 1-\alpha )\);

(W4) \(d(W(x, z, \alpha ), W(y, w, \alpha )\le \alpha d(x, y) + (1- \alpha )d(z, w)\).

(2) A hyperbolic space \((X, d, W)\) is called uniformly convex if for any \(r > 0\) and \(\epsilon \in (0, 2]\), there exists \(\delta \in (0, 1]\) such that, for all \(x, y, z\in X\),

$$ \left. \begin{aligned} &d(x, z) \le r \\ &d(y, z) \le r \\ &d(x, y) \ge \epsilon \cdot r \end{aligned} \right\} \Rightarrow d\biggl( \frac{1}{2} x \oplus \frac{1}{2} y, z\biggr) \le (1 - \delta )r. $$
(2.6)

A map** \(\eta : (0, \infty ) \times (0, 2] \to (0, 1]\) providing such \(\delta : = \eta (r, \epsilon )\) for given \(r > 0\) and \(\epsilon \in (0, 2]\) is called a modulus of uniform convexity.

Lemma 2.10

([25])

Let \((X, d, W)\) be a uniformly convex hyperbolic space with modulus of uniform convexity η. For any \(r > 0\), \(\epsilon \in (0, 2]\), \(\lambda \in [0, 1]\), and \(x, y, z \in X\),

$$ \left. \begin{aligned} &d(x, z) \le r \\ &d(y, z) \le r \\ &d(x, y) \ge \epsilon \cdot r \end{aligned} \right\} \Rightarrow d\bigl((1 - \lambda ) x \oplus \lambda y, z\bigr) \le \bigl(1 - 2\lambda (1-\lambda )\eta (r, \epsilon )\bigr)r. $$
(2.7)

Proposition 2.11

Let \((X,d)\) be a complete uniformly convex CAT(κ) space \(\kappa > 0\) with modulus of uniform convexity η and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). Let \(x\in X\) be a given point and \(\{t_{n}\}\) be a sequence in \([b, c]\) with \(b, c \in (0, 1)\) and \(0 < b(1-c) \le \frac{1}{2}\). Let \(\{x_{n}\}\) and \(\{y_{n}\}\) be any sequences in X such that

$$\begin{aligned}& \limsup_{n \to \infty} d(x_{n}, x)\le r, \qquad \limsup_{n \to \infty}d(y_{n}, x)\le r\quad\textit{and} \\& \lim_{n \to \infty}d\bigl((1-t_{n})x_{n} \oplus t_{n} y_{n}\bigr), x) = r \quad \textit{for some } r \ge 0. \end{aligned}$$

Then

$$ \lim_{n \to \infty}d(x_{n}, y_{n})=0. $$
(2.8)

Proof

By the assumption that \((X,d)\) is a complete CAT(κ) space \(\kappa > 0\) and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\), it follows from Lemma 2.5 that for all \(x, y, z \in X\) and \(\alpha \in [0, 1]\)

$$ d\bigl((1 -\alpha )x \oplus \alpha y, z\bigr) \le (1 -\alpha ) d(x, z) + \alpha d(y, z). $$

Letting \(W(x, y, \alpha ): = (1 -\alpha )x \oplus \alpha y\). It is easy to prove that \(W(x, y, \alpha )\) satisfies conditions \((W1)-(W4)\). Hence \((X, d, W)\) is a hyperbolic space. Again, since \((X,d)\) is uniformly convex with modulus of uniform convexity η, this implies that \((X, d, W)\) is a uniformly convex hyperbolic space with modulus of uniform convexity η.

Now we consider two cases.

1. If \(r = 0\), then the conclusion of Proposition 2.11 is obvious.

2. The case of \(r > 0\). If it is not the case that \(d(x_{n}, y_{n}) \to 0\) as \(n \to \infty \), then there are subsequences (denoted by \(\{x_{n}\}\) and \(\{y_{n}\}\) again) such that

$$ \inf_{n} d(x_{n}, y_{n}) > 0. $$
(2.9)

Choose \(\epsilon \in (0, 1]\) such that

$$ d(x_{n}, y_{n}) \ge \epsilon (r + 1) > 0,\quad \forall n \in \mathscr{N}. $$
(2.10)

Since \(0 < b(1-c) < \frac{1}{2}\) and \(0 < \eta (r,\epsilon ) \le 1\), \(0 < 2b(1-c)\eta (r,\epsilon ) \le 1\). This implies \(0 \le 1 - 2b(1-c)\eta (r,\epsilon ) < 1\). Choose \(R \in (r, r +1)\) such that

$$ (1 - 2b(1- c)\eta (r,\epsilon )R < r. $$
(2.11)

Since

$$ \lim \sup_{n} d(x_{n}, x) \le r,\qquad \lim \sup _{n} d(y_{n}, x) \le r,\quad r < R, $$
(2.12)

there are further subsequences again denoted by \(\{x_{n}\}\) and \(\{y_{n}\}\) such that

$$ d(x_{n}, x) \le R,\qquad d(y_{n}, x) \le R,\qquad d(x_{n}, y_{n}) \ge \epsilon R,\quad \forall n \in \mathscr{N}. $$
(2.13)

Then, by Lemma 2.10 and (2.11),

$$ \begin{aligned} d\bigl((1 - t_{n})x_{n}, t_{n}y_{n}, x\bigr) &\le \bigl(1- 2t_{n}(1- t_{n})\eta (R, \epsilon )\bigr)R \\ & \le \bigl(1 - 2b(1 -c)\eta (r,\epsilon )\bigr)R < r \end{aligned} $$
(2.14)

for all \(n \in \mathscr{N}\). Taking \(n\to \infty \), we obtain

$$ \lim_{n \to \infty} d\bigl((1-t_{n}) x_{n} \oplus t_{n} y_{n}, x\bigr) < r, $$
(2.15)

which contradicts the hypothesis.

The conclusion of Proposition 2.11 is proved. □

Lemma 2.12

Let \(\{ a_{n}\}\), \(\{ \lambda _{n}\}\), and \(\{ c_{n}\}\) be the sequences of nonnegative numbers such that

$$ a_{n+1} \le (1+ \lambda _{n})a_{n} + c_{n}, \quad \forall n \ge 1. $$

If \(\sum_{n = 1}^{\infty }\lambda _{n} < \infty \) and \(\sum_{n = 1}^{\infty }c_{n} < \infty \), then \(\lim_{n \to \infty} a_{n} \) exists. In addition, if there exists a subsequence \(\{a_{n_{i}}\} \subset \{a_{n}\}\) such that \(a_{n_{i}} \to 0\), then \(\lim_{n \to \infty} a_{n} = 0\).

3 Strong convergence theorems for total asymptotically nonexpansive map**s in CAT(κ) spaces

Lemma 3.1

([26])

(1) For each positive integer \(n\ge 1\), the unique solutions \(i(n)\) and \(k(n)\) with \(k(n) \ge i(n)\) to the following positive integer equation

$$ n = i(n) + \frac{(k(n) -1)k(n)}{2} $$
(3.1)

are as follows:

$$ \textstyle\begin{cases} i(n) = n - \frac{(k(n)-1)k(n)}{2}, \\ k(n) =[\frac{1}{2} + \sqrt[2]{2n - \frac{7}{4}}], & k(n) \ge i(n), \end{cases} $$

and \(k(n) \to \infty\) (as \(n \to \infty \)), where \([x]\) denotes the maximal integer that is not larger than x.

(2) For each \(i \ge 1\), denote by

$$ \textstyle\begin{cases} \Gamma _{i}: = \{n \in \mathscr{N}: n = i + \frac{(k(n)-1)k(n)}{2}, k(n) \ge i\},\quad \textit{and} \\ K_{i}: = \{k(n): n \in \Gamma _{i}, n = i + \frac{(k(n)-1)k(n)}{2}, k(n) \ge i\}, \end{cases} $$

then \(k(n)+1 = k(n+1)\), \(\forall n \in \Gamma _{i}\).

In this section we prove some strong convergence theorems for the following iterative scheme:

$$ \textstyle\begin{cases} x_{1} \in C, \\ x_{n+1} = (1 - \alpha _{n}) x_{n} \oplus \alpha _{n} T_{i(n)}^{k(n)}y_{n}, \\ y_{n} = (1-\beta _{n})x_{n} \oplus \beta _{n} T_{i(n)}^{k(n)}x_{n}, \end{cases}\displaystyle n \ge 1, $$
(3.2)

where C is a nonempty closed and convex subset of a complete CAT(κ) space X, \(\kappa > 0\), for each \(i \ge 1\), \(T_{i}: C\rightarrow C \) is uniformly \(L_{i}\)-Lipschitzian and \((\{\nu _{n}^{(i)}\}, \{\mu _{n}^{(i)}\}, \zeta ^{(i)})\)-total asymptotically nonexpansive map**s defined by (1.2); and for each positive integer \(n\ge 1\), \(i(n)\) and \(k(n)\) are the unique solutions of the positive integer equation (3.1).

Theorem 3.2

Let \((X,d)\) be a complete uniformly convex CAT(κ) space with \(\kappa > 0\) and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\). Let C be a nonempty closed and convex subset of X and, for each \(i \ge 1\), let \(T_{i}: C\rightarrow C\) be uniformly \(L_{i}\)-Lipschitzian and \((\{\nu _{n}^{(i)}\}, \{\mu _{n}^{(i)}\}, \zeta ^{(i)})\)-total asymptotically nonexpansive map**s defined by (1.2) such that

  1. (i)

    \(\sum_{i=1}^{\infty }\sum_{n=1}^{\infty }\nu _{n}^{(i)} < \infty \), \(\sum_{i=1}^{\infty}\sum_{n=1}^{\infty }\mu _{n}^{(i)} < \infty \),

  2. (ii)

    there exists a constant \(M_{*} > 0\) such that \(\zeta ^{(i)}(r)\leq M_{*} r\), \(\forall r\geq 0\), \(i= 1, 2, \ldots \) ;

  3. (iii)

    there exist constants \(a, b \in (0, 1)\) with \(0 < b(1-a) \le \frac{1}{2}\) such that \(\{\alpha _{n}\}, \{\beta _{n}\} \subset [a, b]\).

If \(\mathscr{F} := \bigcap_{i = 1}^{\infty }F(T_{i})\neq \emptyset \) and there exist a map** \(T_{n_{0}} \in \{T_{i}\}_{i=1}^{\infty}\) and a nondecreasing function \(f: [0, \infty ) \to [0, \infty )\) with f(0)= 0 and \(f(r) > 0\) \(\forall r > 0\) such that

$$ f\bigl(d(x_{n}, \mathscr{F})\bigr) \le d(x_{n}, T_{n_{0}} x_{n}), \quad \forall n \ge 1, $$
(3.3)

then the sequence \(\{x_{n}\}\) defined by (3.2) converges strongly (i.e., in metric topology) to some point \(p^{*} \in \mathscr{F}\).

Proof

First we observe that for each \(i\ge 1\), \(T_{i}: C \to C\) is a \((\{\nu _{n}^{(i)}\}, \{\mu _{n}^{(i)}\}, \zeta ^{(i)})\)-total asymptotically nonexpansive map**. By condition (ii), for each \(n \ge 1\) and any \(x, y \in C\), we have

$$ d\bigl(T_{i}^{n} x, T_{i}^{n} y\bigr) \le d(x,y) + \nu _{n}^{(i)}\zeta ^{i}\bigl( d(x,y) \bigr)+ \mu _{n}^{(i)} \le \bigl(1+ \nu _{n}^{(i)} M_{*}\bigr)d(x,y) + \mu _{n}^{(i)}. $$
(3.4)

(I) First we prove that the following limits exist:

$$ \lim_{n \to \infty}d(x_{n}, \mathscr{F}), \quad \text{and} \quad \lim_{n \to \infty}d(x_{n}, p)\quad\text{for each } p \in \mathscr{F}. $$
(3.5)

In fact, since \(p \in \mathscr{F}\) and \(T_{i}\), \(i \ge 1\) is a total asymptotically nonexpansive map**, it follows from (3.4) and Lemma 2.5 that

$$ \begin{aligned} d(y_{n}, p)& = d\bigl((1-\beta _{n}) x_{n} \oplus \beta _{n} T_{i(n)}^{k(n)}x_{n}, p\bigr) \\ & \le (1-\beta _{n})d( x_{n}, p) + \beta _{n} d \bigl(T_{i(n)}^{k(n)}x_{n}, p\bigr) \\ & = (1-\beta _{n}) d(x_{n}, p) + \beta _{n} \bigl\{ d(x_{n}, p)+ \nu _{k(n)}^{i(n)} \zeta ^{i(n)} \bigl(d(x_{n}, p)\bigr)+ \mu ^{i(n)}_{k(n)}\bigr\} \\ & \le d(x_{n}, p)+ \nu _{k(n)}^{i(n)}\zeta ^{i(n)}\bigl(d(x_{n}, p)\bigr)+ \mu _{k(n)}^{i(n)} \\ & \le \bigl(1+ \nu _{k(n)}^{i(n)} M_{*} \bigr)d(x_{n}, p) + \mu _{k(n)}^{i(n)} \end{aligned} $$
(3.6)

and

$$ \begin{aligned} d(x_{n+1}, p)& = d\bigl((1 - \alpha _{n}) x_{n} \oplus \alpha _{n} T_{i(n)}^{k(n)}y_{n}, p\bigr) \\ & \le (1- \alpha _{n}) d( x_{n}, p) + \alpha _{n} d\bigl(T_{i(n)}^{k(n)}y_{n}, p\bigr) \\ & = (1- \alpha _{n})d(x_{n}, p) + \alpha _{n} \bigl\{ d(y_{n}, p)+ \nu ^{i(n)}_{k(n)} \zeta ^{i(n)}\bigl( d(y_{n}, p)\bigr) + \mu ^{i(n)}_{k(n)} \bigr\} \\ & \le (1- \alpha _{n}) d(x_{n}, p) + \alpha _{n} \bigl\{ \bigl(1+ \nu ^{i(n)}_{k(n)} M_{*} \bigr)d(y_{n}, p) + \mu ^{i(n)}_{k(n)}\bigr\} . \end{aligned} $$
(3.7)

Substituting (3.6) into (3.7) and simplifying it, we have

$$ d(x_{n+1}, p) \le (1 + \sigma _{n})d(x_{n}, p) + \xi _{n}, \quad \forall n \ge 1 \text{ and } p\in \mathscr{F}, $$
(3.8)

and so

$$ d(x_{n+1}, \mathscr{F}) \le (1 + \sigma _{n})d(x_{n}, \mathscr{F}) + \xi _{n}, \quad \forall n \ge 1, $$
(3.9)

where \(\sigma _{n} = b\nu ^{i(n)}_{k(n)} M_{*}(2 + \nu ^{i(n)}_{k(n)} M_{*})\), \(\xi _{n} = b (2 + \nu ^{i(n)}_{k(n)} M_{*})\mu ^{i(n)}_{k(n)}\). By virtue of condition (i),

$$ \sum_{n=1}^{\infty }\sigma _{n} < \infty \quad \text{and}\quad \sum_{n=1}^{ \infty } \xi _{n} < \infty . $$
(3.10)

By Lemma 2.12, the limits \(\lim_{n \to \infty}d(x_{n}, \mathscr{F})\) and \(\lim_{n \to \infty}d(x_{n}, p)\) exist for each \(p \in \mathscr{F}\).

(II) Next we prove that for each \(i \ge 1\) there exists some subsequence \(\{x_{m(\in \Gamma _{i})}\} \subset \{x_{n}\}\) such that

$$ \lim_{m(\in \Gamma _{i})\to \infty}d(x_{m}, T_{i} x_{m}) = 0, $$
(3.11)

where \(\Gamma _{i}\) is the set of positive integers defined by Lemma 3.1(2).

In fact, it follows from (3.5) that for each given \(p \in \mathscr{F}\), the limit \(\lim_{n \to \infty}d(x_{n}, p)\) exists. Without loss of generality, we can assume that

$$ \lim_{n \to \infty}d(x_{n}, p) = r \ge 0. $$
(3.12)

From (3.6) we have

$$ \limsup_{n \to \infty}d(y_{n}, p)\le \lim _{n \to \infty} \bigl\{ \bigl(1+ \nu _{k(n)}^{i(n)} M_{*}\bigr)d(x_{n}, p) + \mu _{k(n)}^{i(n)} \bigr\} = r. $$
(3.13)

Since

$$ \begin{aligned} d\bigl(T_{i(n)}^{k(n)} y_{n}, p\bigr) & \le d(y_{n}, p) + \nu ^{i(n)}_{k(n)} \zeta ^{i(n)}\bigl(d(y_{n}, p)\bigr) + \nu ^{i(n)}_{k(n)} \\ & \le \bigl(1 + \nu ^{i(n)}_{k(n)} M_{*}\bigr) d(y_{n}, p) + \mu ^{i(n)}_{k(n)}, \quad \forall n \ge 1, \end{aligned} $$

from (3.13) we have

$$ \limsup_{n \to \infty}d\bigl(T_{i(n)}^{k(n)} y_{n}, p\bigr)\le r. $$
(3.14)

In addition, it follows from (3.8) that

$$ \begin{aligned} d(x_{n+1}, p)& = d\bigl((1 - \alpha _{n}) x_{n} \oplus \alpha _{n} T_{i(n)}^{k(n)}y_{n}, p\bigr) \\ &\le (1 + \sigma _{n})d(x_{n}, p) + \xi _{n}. \end{aligned} $$

This implies that

$$ \lim_{n \to \infty}d\bigl((1 - \alpha _{n}) x_{n} \oplus \alpha _{n} T_{i(n)}^{k(n)}y_{n}, p \bigr)= r. $$
(3.15)

From (3.12), (3.14), (3.15), and Proposition 2.11, we have

$$ \lim_{n \to \infty}d\bigl( x_{n}, T_{i(n)}^{k(n)}y_{n} \bigr) = 0. $$
(3.16)

Since

$$ \begin{aligned} d(x_{n}, p) &\le d\bigl(x_{n}, T_{i(n)}^{k(n)}y_{n}\bigr) + d\bigl(T_{i(n)}^{k(n)}y_{n}, p\bigr) \\ & \le d\bigl(x_{n}, T_{i(n)}^{k(n)}y_{n} \bigr) + \bigl\{ d(y_{n}, p) +\nu ^{i(n)}_{k(n)} \zeta ^{i(n)}\bigl(d(y_{n}, p)\bigr) + \mu ^{i(n)}_{k(n)} \bigr\} \\ & \le d\bigl(x_{n}, T_{i(n)}^{k(n)}y_{n} \bigr) + \bigl(1 + \nu ^{i(n)}_{k(n) }M_{*} \bigr)d(y_{n}, p) + \mu ^{i(n)}_{k(n)}. \end{aligned} $$

Taking lim inf on both sides of the above inequality, from (3.16) we have

$$ \liminf_{n \to \infty}d(y_{n}, p) \ge r. $$

This together with (3.13) shows that

$$ \lim_{n \to \infty}d(y_{n}, p) = r. $$
(3.17)

Using (3.6) we have

$$ \begin{aligned} r &= \lim_{n \to \infty}d(y_{n}, p) = \lim_{n \to \infty}\bigl\{ d\bigl((1- \beta _{n}) x_{n} \oplus \beta _{n} T_{i(n)}^{k(n)}x_{n}, p\bigr)\bigr\} \\ &\le \lim_{n \to \infty} \bigl\{ \bigl(1+ \nu _{k(n)}^{i(n)} M_{*}\bigr)d(x_{n}, p) + \mu _{k(n)}^{i(n)} \bigr\} = r. \end{aligned} $$
(3.18)

This implies that

$$ \lim_{n \to \infty}\bigl\{ d\bigl((1-\beta _{n}) x_{n} \oplus \beta _{n} T_{i(n)}^{k(n)}x_{n}, p\bigr)\bigr\} =r. $$
(3.19)

Similarly, we can also prove that

$$ \limsup_{n \to \infty} d\bigl(T_{i(n)}^{k(n)}x_{n}, p\bigr) \le \limsup_{n \to \infty}\bigl\{ d(x_{n},p) + \nu _{k(n)}^{i(n)}\zeta ^{i(n)}\bigl(d(x_{n},p) \bigr) + \mu _{k(n)}^{i(n)}\bigr\} \le r. $$

This together with (3.12), (3.19), and Lemma 2.11 gives that

$$ \lim_{n \to \infty}d\bigl(x_{n},T_{i(n)}^{k(n)}x_{n} \bigr) = 0. $$
(3.20)

Therefore we have

$$ \begin{aligned} d(x_{n}, y_{n}) & = d \bigl(x_{n}, (1-\beta _{n})x_{n} \oplus \beta _{n} T_{i(n)}^{k(n)}x_{n}\bigr) \\ & \le \beta _{n} d\bigl(x_{n},T_{i(n)}^{k(n)}x_{n} \bigr) \to 0 \quad (\text{as } n \to \infty ). \end{aligned} $$
(3.21)

Furthermore, it follows from (3.16) that

$$ \begin{aligned} d(x_{n +1}, x_{n}) & = d((1 - \alpha _{n}) x_{n} \oplus \alpha _{n} d \bigl(T_{i(n)}^{k(n)}y_{n}, x_{n}\bigr) \\ & \le \alpha _{n} d\bigl(T_{i(n)}^{k(n)}y_{n}, x_{n}\bigr) \to 0 \quad (\text{as } n \to \infty ). \end{aligned} $$
(3.22)

This together with (3.21) shows that

$$ d(x_{n +1}, y_{n}) \le d(x_{n +1}, x_{n}) + d(x_{n}, y_{n}) \to 0 \quad (\text{as } n \to \infty ). $$
(3.23)

From Lemma 3.1, (3.16), (3.20), (3.22), and (3.23), we have that for each given positive integer \(i \ge 1\), there exist subsequences \(\{x_{m}\}_{m\in \Gamma _{i}}\), \(\{y_{m}\}_{m \in \Gamma _{i}}\), and \(\{k(m)\}_{m \in \Gamma _{i}} \subset K_{i}: = \{k(m): m \in \Gamma _{i}, m = i + \frac{(k(m)-1)k(m)}{2}, k(m) \ge i\}\) such that

$$ \begin{aligned} d(x_{m}, T_{i} x_{m}) \le{}& d\bigl(x_{m}, T_{i}^{k(m)}x_{m} \bigr) + d\bigl(T_{i}^{k(m)}x_{m}, T_{i}^{k(m)}y_{m-1}\bigr) + d\bigl(T_{i}^{k(m)}y_{m-1}, T_{i} x_{m}\bigr) \\ \le{}& d\bigl(x_{m}, T_{i}^{k(m)}x_{m} \bigr) + \bigl\{ d(x_{m}, y_{m-1}) + \nu _{k(m)}^{(i)} \zeta ^{(i)}\bigl(d(x_{m},y_{m-1})\bigr) + \mu _{k(m)}^{(i)}\bigr\} \\ & {} + L_{i} (d\bigl(T_{i}^{k(m)-1}y_{m-1}, x_{m}\bigr) \\ \le{}& d\bigl(x_{m}, T_{i}^{k(m)}x_{m} \bigr) + \bigl\{ d(x_{m},y_{m-1}) + \nu _{k(m)}^{(i)} \zeta ^{(i)}\bigl(d(x_{m},y_{m-1})\bigr) + \mu _{k(m)}^{(i)}\bigr\} \\ & {} + L_{i} (d\bigl(T_{i}^{k(m-1)}y_{m-1}, x_{m-1}\bigr) + L_{i} (d(x_{m-1}, x_{m}) \to 0 (\text{as} m \to \infty ). \end{aligned} $$
(3.24)

The conclusion (3.11) is proved.

(III) Now we prove that \(\{x_{n}\}\) converges strongly (i.e., in the metric topology) to some point \(p^{*} \in \mathscr{F}\).

In fact, it follows from (3.11) and (3.24) that for given map** \(T_{n_{0}}\) there exists some subsequence \(\{x_{m}\}_{m \in \Gamma _{n_{0}}}\) of \(\{x_{n}\}\) such that

$$ \lim_{m(\in \Gamma _{n_{0}}) \to \infty}d(x_{m}, T_{n_{0}} x_{m}) = 0. $$

By (3.3) we have

$$ f\bigl(d(x_{m}, \mathscr{F})\bigr) \le d(x_{m}, T_{n_{0}} x_{m})\quad \forall m \ge 1. $$

Let \(m \to \infty \), and taking limsup on the above inequality, we have \(\lim_{m\to \infty} f(d(x_{m}, \mathscr{F})) = 0\). By the property of f, this implies that

$$ \lim_{m(\in \Gamma _{n_{0}}) \to \infty} d(x_{m}, \mathscr{F}) = 0. $$
(3.25)

Next we prove that \(\{x_{m}\}_{m \in \Gamma _{n_{0}}}\) is a Cauchy sequence in C. In fact, it follows from (3.8) that for any \(p \in \mathscr{F}\)

$$ d(x_{m+1}, p) \le (1 + \sigma _{m})d(x_{m}, p) + \xi _{m}, \quad \forall m ( \in \Gamma _{n_{0}}) \ge 1, $$

where \(\sum_{m = 1}^{\infty }\sigma _{m} < \infty \) and \(\sum_{m = 1}^{\infty }\xi _{m} < \infty \). Hence, for any positive integers \(j, n \in \Gamma _{n_{0}}\), \(n > j\), and \(n = m + j\) for some positive integer m, we have

$$ \begin{aligned} d(x_{n}, x_{j}) &= d(x_{j+m}, x_{j}) \le d(x_{j+m},p) + d(x_{j}, p) \\ & \le (1 + \sigma _{j+m-1})d(x_{j+m-1}, p) + \xi _{j+m-1} + d(x_{j}, p). \end{aligned} $$

Since for each \(x \ge 0\), \(1+x \le e^{x}\), we have

$$ \begin{aligned} d(x_{n}, x_{j}) ={}& d(x_{j+m}, x_{j}) \\ \le{}& e^{\sigma _{j+m-1}} d(x_{j+m-1}, p) + \xi _{j+m-1} + d(x_{j}, p) \\ \le{}& e^{\sigma _{j+m-1} + \sigma _{j+m-2}}d(x_{j+m-2}, p) + e^{ \sigma _{j+m-1}}\xi _{j+m-2} + \xi _{j+m-1} + d(x_{j}, p) \\ \le{}& \cdots \\ \le{}& e^{\sum _{i=j}^{j+m-1}\sigma _{i}} d(x_{j}, p) + e^{\sum _{i=j+1}^{j+m-1} \sigma _{i}}\xi _{j} + e^{\sum _{i=j+2}^{j+m-2}\sigma _{i}}\xi _{j+1} + \cdots \\ & {} + e^{\sigma _{j+m-1}}\xi _{j+m-2} + \xi _{j+m-1} + d(x_{j}, p) \\ \le {}& (1+ M)d(x_{j}, p) + M\sum_{i=j}^{j+m - 1} \xi _{i} \\ ={}& (1+ M)d(x_{j}, p) + M\sum_{i=j}^{n - 1} \xi _{i}, \quad \text{for each } p\in \mathscr{F}. \end{aligned} $$

Therefore we have

$$ d(x_{n}, x_{j}) = d(x_{j+m}, x_{j}) \le (1+ M)d(x_{j}, \mathscr{F}) + M \sum_{i=j}^{n-1} \xi _{i}, $$

where \(M = e^{\sum _{i=1}^{\infty}}\sigma _{i} < \infty \). By (3.25) we have

$$ d(x_{n}, x_{j}) \le (1+ M)d(x_{j}, \mathscr{F}) + M\sum_{i=j}^{n-1} \xi _{i} \to 0 \quad \bigl(\text{as } n, j (\in \Gamma _{n_{0}}) \to \infty \bigr). $$

This shows that the subsequence \(\{x_{m}\}_{m \in \Gamma _{n_{0}}}\) is a Cauchy sequence in C. Since C is a closed subset in a complete CAT(κ) space X, it is complete. Without loss of generality, we can assume that the subsequence \(\{x_{m}\}\) converges strongly (i.e., in metric topology in X) to some point \(p^{*} \in C\). It is easy to know that \(\mathscr{F}\) is a closed subset in C. Since \(\lim_{m \to \infty} d(x_{m}, \mathscr{F}) = 0\), \(p^{*} \in \mathscr{F}\). By using (3.5), it yields that the whole sequence \(\{x_{n}\}\) converges strongly to \(p^{*} \in \mathscr{F}\).

This completes the proof of Theorem 3.2. □

Remark 3.3

It should be pointed out that if \((X, d)\) is a CAT(0) space, then X is uniformly convex, its modulus of uniform convexity \(\eta (r,\epsilon ) = \frac{\epsilon ^{2}}{8}\) [25, 27] and all of its geodesic triangles satisfy the \(\operatorname{CAT} (\kappa )\) inequality. These imply that if \((X, d)\) is a CAT(0) space, then the conditions that appeared in Theorem 3.2\((X,d)\) is uniformly convex and \(\operatorname{diam}(X)\le \frac{\frac{\pi}{2} - \epsilon}{\sqrt{\kappa}}\) for some \(\epsilon \in (0, \frac{\pi}{2})\)” are of no use here. Therefore from Theorem 3.2 we can obtain the following.

Theorem 3.4

Let \((X,d)\) be a complete CAT(0) space. Let C be a nonempty closed and convex subset of X, and for each \(i \ge 1\), let \(T_{i}: C\rightarrow C\) be uniformly \(L_{i}\)-Lipschitzian and \((\{\nu _{n}^{(i)}\}, \{\mu _{n}^{(i)}\}, \zeta ^{(i)})\)-total asymptotically nonexpansive map**s defined by (1.2) such that

  1. (i)

    \(\sum_{i=1}^{\infty }\sum_{n=1}^{\infty }\nu _{n}^{(i)} < \infty \), \(\sum_{i=1}^{\infty}\sum_{n=1}^{\infty }\mu _{n}^{(i)} < \infty \),

  2. (ii)

    there exists a constant \(M_{*} > 0\) such that \(\zeta ^{(i)}(r)\leq M_{*} r\), \(\forall r\geq 0\), \(i= 1, 2, \ldots \) ;

  3. (iii)

    there exist constants \(a, b \in (0, 1)\) with \(0 < b(1-a) \le \frac{1}{2}\) such that \(\{\alpha _{n}\}, \{\beta _{n}\} \subset [a, b]\).

If \(\mathscr{F} := \bigcap_{i = 1}^{\infty }F(T_{i})\neq \emptyset \) and there exist a map** \(T_{n_{0}} \in \{T_{i}\}_{i=1}^{\infty}\) and a nondecreasing function \(f: [0, \infty ) \to [0, \infty )\) with \(f(0)= 0\) and \(f(r) > 0\) \(\forall r > 0\) such that

$$ f\bigl(d(x_{n}, \mathscr{F})\bigr) \le d(x_{n}, T_{n_{0}} x_{n}),\quad \forall n \ge 1, $$

then the sequence \(\{x_{n}\}\) defined by (3.2) converges strongly (i.e., in metric topology) to some point \(p^{*} \in \mathscr{F}\).