1 Introduction

Let \((\mathcal{M},\|.\|)\) be a Banach space and \(\mathcal{C}\) a nonempty subset of \(\mathcal{M}\). A map** \(S: \mathcal{C} \to \mathcal{C} \) is said to be nonexpansive if

$$ \bigl\Vert S(u)-S(v) \bigr\Vert \leq \Vert u-v \Vert , \quad\forall u,v \in \mathcal{C}. $$

A point \(u^{\dagger} \in \mathcal{C}\) is said to be a fixed point of S if \(S(u^{\dagger})= u^{\dagger}\). In the context of Banach spaces, a nonexpansive map** may not necessarily possess a fixed point. However, it is possible to obtain fixed points for such map**s by enriching the space with certain geometric properties. In 1965, Browder [5] and Göhde [15] separately established that nonexpansive map**s have fixed points in every uniformly convex Banach space. Kirk [20], on the other hand, extended the fixed-point theorem for nonexpansive map**s to the broader category of reflexive Banach spaces with normal structure. Recall that a Banach space \((\mathcal{M},\|.\|)\) is said to have normal structure, if for each bounded, closed, and convex subset \(\mathcal{C}\) of \(\mathcal{M}\) consisting of more than one point there is a point \(u \in \mathcal{C}\) such that

$$ \sup \bigl\{ \Vert v-u \Vert :~ v \in \mathcal{C} \bigr\} < \operatorname{diam}_{ \Vert . \Vert }(\mathcal{C})= \sup \bigl\{ \Vert v-u \Vert :~ u,~v \in \mathcal{C} \bigr\} . $$

In [3], Baillon and Schöneberg weakened the concept of normal structure and introduced the asymptotic normal structure as follows: A Banach space \((\mathcal{M},\|.\|)\) is said to have asymptotic normal structure, if for each bounded, closed, and convex subset \(\mathcal{C}\) of \(\mathcal{M}\) consisting of more than one point and each sequence \(\{u_{n}\}\) in \(\mathcal{C}\) satisfying \(u_{n} -u_{n+1} \to 0\) as \(n\to \infty \), there is a point \(u \in \mathcal{C}\) such that

$$ \liminf_{n \to \infty} \Vert u_{n}-u \Vert < \operatorname{diam}_{ \Vert . \Vert }(\mathcal{C}). $$

The relationship between fixed-point theory and the geometry of Banach spaces has been highly productive and significant. In the context of metric fixed-point problems, geometric properties are particularly influential. Nonexpansive map**s are a prominent area of study in metric fixed-point theory. Many authors have since derived generalizations and extensions of nonexpansive map**s and their associated results. The literature contains a considerable body of research on classes of map**s that are more general than the nonexpansive ones. Some of the notable extensions and generalizations of nonexpansive map**s can be found in [1, 2, 4, 6, 7, 11, 13, 21, 22, 2426]. Some classes of map**s are not necessarily continuous on their domains, unlike nonexpansive map**s. In 2008, Suzuki [27] introduced a new class of nonexpansive-type map**s, referred to as map**s satisfying condition (C), and derived some significant fixed-point results for them. Suzuki [27] also demonstrated that this class of map**s does not necessarily exhibit continuity, unlike nonexpansive map**s. García-Falset et al. [11] explored a broader version of condition (C), called map**s satisfying condition (E). In 2011, Llorens-Fuster and Moreno-Galvez [21] introduced a general class of map**s called (L)-type map**s (or condition (L)), which is contingent on two conditions. First, the existence of an approximate fixed-point sequence (a.f.p.s.) for S in all nonempty, closed, convex, and S-invariant subsets of C. Secondly, the distances between points and their images in the limiting case from the a.f.p.s. For this class of map**s, the nonexpansiveness condition need not hold for all points but only for certain points in the domain. They obtained several fixed-point results for their new class of nonexpansive-type map**s.

It is noted herein that the normal structure condition depends on the distance between all points of set \(\mathcal{C}\) and point u, while the asymptotic normal structure condition depends on the limiting distance between sequence \(\{u_{n}\}\) and point u. This condition seems similar to the second condition of (L)-type map**s. It looks natural to investigate the fixed-point theorem for (L)-type map**s in the setting of Banach spaces having asymptotic normal structure. In this paper, we present some results concerning the demiclosedness principle of a map** satisfying condition (L-1) in uniform convex spaces. Further, we obtain some fixed-point theorems for (L-1)-type map**s in the setting of Banach spaces having asymptotic normal structure. Moreover, we show that in \(\ell _{1}\) and \(J_{0}\) (James space), (L-1)-type self-map** of a bounded weak closed convex subset has a fixed point. In this way, results in [11, 18, 21, 27] have been extended, generalized, and complemented.

2 Preliminaries

Definition 1

[12]. Let \(\mathcal{C}\) be a nonempty subset of a Banach space \(\mathcal{M}\). A sequence \(\{u_{n}\}\) in \(\mathcal{C}\) is said to be an approximate fixed-point sequence (in short, a.f.p.s.) for a map** \(S: \mathcal{C} \to \mathcal{C}\) if \(\lim_{n \to \infty}\|u_{n}-S(u_{n})\|=0\).

Definition 2

[12]. Let \(\mathcal{C}\) be a subset of a Banach space \(\mathcal{M}\). A map** \(G:\mathcal{C}\rightarrow \mathcal{C}\) is said to be demiclosed if for any sequence \(\{u_{n}\}\) in \(\mathcal{C}\) the following implication holds:

$$ \{u_{n}\} \text{{ converges weakly to }} u \text{{ and }} \lim _{n\rightarrow \infty } \bigl\Vert G(u_{n})-w \bigr\Vert =0 $$

that implies

$$ u \in \mathcal{C} \quad\text{{and}}\quad G(u)=w. $$

Definition 3

[16]. Let \(\mathcal{M}\) be a Banach space and \(u,v \in \mathcal{M}\). A vector u is orthogonal to v if \(\|u\| \leq \|u+\mu v\|\) for all scalars μ. We use to denote \(u \bot v\) if u is orthogonal to v.

In general, the relation ⊥ is not symmetric cf. [18].

Definition 4

[18]. Let \(\mathcal{M}\) be a Banach space. The relation ⊥ is said to be approximately symmetric if for each \(u \in \mathcal{M}\) and \(\varepsilon >0\), there exists a closed, linear subspace \(\mathcal{Y}=\mathcal{Y}(u,\varepsilon )\) such that the following two conditions hold:

  1. (i)

    \(\mathcal{Y}\) has finite codimension;

  2. (ii)

    \(\|z\| \leq \|z+\mu u\|\) for all \(z \in \mathcal{Y}\), \(\|z\|=1\), and each μ with \(\mu \geq \varepsilon \).

Definition 5

[18]. Let \(\mathcal{M}\) be a conjugate space, that is, there exists a normed space \(\mathcal{Z}\) such that \(\mathcal{M}=\mathcal{Z}^{*}\). The relation ⊥ is said to be weak approximately symmetric if conditions (i) and (ii) in Definition 4 hold along with \(\mathcal{Y}\) is weak closed.

Definition 6

[18]. Let \(\mathcal{M}\) be a Banach space. The relation ⊥ is said to be uniformly approximately symmetric (uniformly weak approximately symmetric) if it is approximately symmetric (weak approximately symmetric) and condition (ii) in Definition 4 is replaced by the following stronger condition:

  1. (iii)

    there exists \(\delta =\delta (u,\varepsilon )>0\) such that \(\|z\| \leq \|z+\mu u\|-\delta \), for all \(z \in Y\), \(\|z\|=1\), and each μ with \(\mu \geq \varepsilon \).

In the spaces \(\ell _{p}\), \(p \in (1,\infty )\), the relation ⊥ is uniformly approximately symmetric. In spaces \(\ell _{1}\) and \(J_{0}\) (James space [17]) the relation ⊥ is uniformly weak approximately symmetric. However, in both spaces \(L_{p}\), \(p\neq 2\) and \(c_{0}\), the relation ⊥ fails to be uniformly approximately symmetric.

Lemma 1

(Goebel–Karlovitz) [14]. Let \(\mathcal{C}\) be a subset of a reflexive Banach space \(\mathcal{M}\), and suppose \(\mathcal{C}\) is minimally invariant with respect to being nonempty, bounded, closed, convex, and S-invariant for some nonexpansive map** S. Let \(\{x_{n}\}\) be a sequence in \(\mathcal{C}\) that satisfies \(\lim_{n \to \infty}\|u_{n}-S(u_{n})\|=0\). Then, for each \(u \in \mathcal{C}\), \(\lim_{n \to \infty}\|u_{n}-u\| = \operatorname{diam}(\mathcal{C})\).

Theorem 1

[14] Let \(\mathcal{M}\) be a uniformly convex Banach space. Then, for any \(d>0\), \(\varepsilon >0\) and \(u,v \in X\) with \(\|u\|\leq d, \|v\|\leq d, \|u-v\|\geq \varepsilon \), there exists a \(\delta >0\) such that

$$ \biggl\Vert \frac{1}{2}(u+v) \biggr\Vert \leq \biggl[ 1-\delta \biggl( \frac{\varepsilon}{d} \biggr) \biggr]d. $$

Theorem 2

[23]. Let \(\mathcal{M}\) be a Banach space. The following conditions are equivalent:

  1. (i)

    \(\mathcal{M}\) is strictly convex;

  2. (ii)

    If \(u,v \in \mathcal{M} \) and \(\|u+v\|= \|u\|+\|v\|\), then \(u=0\) or \(v=0\) or \(v= cu\) for some \(c >0\).

Theorem 3

[3]. Let \(\beta \geq 1\) and let \(\mathcal{M}_{\beta}\) be the real space \(\ell _{2}\) renormed according to

$$ \vert u \vert _{\beta}= \max \bigl\{ \Vert u \Vert _{2}, \beta \Vert u \Vert _{\infty} \bigr\} , $$

where \(\|u\|_{\infty}\) denotes the \(\ell _{\infty}\)-norm and \(\|u\|_{2}\) the \(\ell _{2}\) norm. Then,

  1. (1)

    \(\mathcal{M}_{\beta}\) has normal structure if and only if \(\beta < \sqrt{2}\); and

  2. (2)

    \(\mathcal{M}_{\beta}\) has asymptotic normal structure if and only if \(\beta < 2\).

Lemma 2

[3]. Let \(\beta \geq 1\), \(x,y,z \in \mathcal{M}_{\beta}\) and \(\alpha \in [0,1]\). Then,

$$ \bigl\Vert x- \bigl((1-\alpha )y+\alpha z \bigr) \bigr\Vert _{2}^{2} + \alpha (1-\alpha ) \Vert y-z \Vert _{2}^{2}=(1- \alpha ) \Vert x-y \Vert _{2}^{2}+\alpha \Vert x-z \Vert _{2}^{2}. $$

Lemma 3

[3, 9, 10]. Let \(\beta \geq 1\) and \(\mathcal{C}\) be a bounded, closed, and convex subset of \(\mathcal{M}_{\beta}\). Let \(\{u_{n}\}\) be a sequence in \(\mathcal{C}\). Then, there exists a unique point \(w \in \mathcal{C}\) that satisfies the following conditions:

  1. (i)

    \(\limsup_{n \to \infty}\|u_{n}-w\|_{2}^{2} +\|w-u\|_{2}^{2} \leq \limsup_{n \to \infty} \|u_{n}-u\|_{2}^{2}\) for all \(u \in \mathcal{C}\); and

  2. (ii)

    \(2 \limsup_{n \to \infty}\|u_{n}-w\|_{2}^{2} \leq \limsup_{p \to \infty} \{\limsup_{n \to \infty} \|u_{n}-u_{p} \|_{2}^{2}\}\).

Lemma 4

[3]. Let \(1 \leq \beta \leq 2\) and \(\mathcal{C}\) be a bounded, closed, and convex subset of \(\mathcal{M}_{\beta}\) with \(d=\operatorname{diam}_{|.|_{\beta}}(\mathcal{C})\). Let \(\{u_{n}\}\) be a sequence in \(\mathcal{C}\) such that \(u_{n}-u_{n+1} \to 0\) as \(n \to \infty \) and \(\lim_{n \to \infty} |u_{n}-u|_{\beta}=d\) for all \(u \in \mathcal{C}\), let \(w \in \mathcal{C}\) be the \(\|.\|_{2}\)-asymptotic-center of \(\{u_{n}\}\) in \(\mathcal{C}\). Then, \(\limsup_{n \to \infty}\|u_{n}-w\|_{2}^{2} \geq 2 ( \frac{d}{\beta} )^{2}\).

Lemma 5

[3]. Let \(\mathcal{C}\) be a bounded, closed, and convex subset of \(\mathcal{M}_{2}\) and let \(\{v_{n}\}\) be a sequence in \(\mathcal{C}\) such that \(\lim_{n \to \infty} |v_{n}-u|_{2} =d=\operatorname{diam}_{|.|_{\beta}}( \mathcal{C})\) for all \(u \in \mathcal{C}\). Then, \(\lim_{n \to \infty} \|v_{n}-u\|_{\infty} =\frac{d}{2}\) for all \(u \in \mathcal{C}\).

Lemma 6

[3]. Let \(\{x_{n}\}\) and \(\{y_{n}\}\) be two sequences in \(\ell _{2}\). Suppose that \(d>0\), \(y_{n} \rightharpoonup 0\) as \(n \to \infty \), \(\lim_{n \to \infty}\|y_{n}\|_{\infty}=\frac{d}{2}\) and \(\lim_{n \to \infty}\|x_{n}\|_{\infty}=\frac{d}{2}\), \(\|x_{n}-y_{p}\| \leq \frac{d}{2}\) for all \(n,p\) and

$$ \limsup_{n \to \infty} \Bigl\{ \limsup_{p \to \infty} \Vert x_{n}+y_{p} \Vert _{2}^{2} \Bigr\} =d^{2}. $$

Then,

$$ \limsup_{n \to \infty} \Bigl\{ \limsup_{p \to \infty} \Vert x_{n}+y_{p} \Vert _{\infty} \Bigr\} < d. $$

Let \(\{u_{n}\}\) be a bounded sequence in Banach space \(\mathcal{M}\), and \(\mathcal{C}\) be a nonempty subset of \(\mathcal{M}\). The asymptotic radius of \(\{u_{n}\}\) at a point x in \(\mathcal{M}\) is defined by

$$ r \bigl(x,\{u_{n}\} \bigr):=\limsup_{n \to \infty} \Vert u_{n}-x \Vert . $$

The asymptotic radius of \(\{u_{n}\}\) with respect to \(\mathcal{C}\) is defined by

$$ r \bigl(\mathcal{C},\{u_{n}\} \bigr):=\inf \bigl\{ r \bigl(x, \{u_{n}\} \bigr):x \in \mathcal{C} \bigr\} . $$

The asymptotic center of \(\{u_{n}\}\) with respect to \(\mathcal{C}\) is defined as

$$ A \bigl(\mathcal{C},\{u_{n}\} \bigr):= \{ x \in \mathcal{C}: r \bigl(x, \{u_{n}\} \bigr)=r \bigl( \mathcal{C},\{u_{n}\} \bigr). $$

Definition 7

[27]. Let \(\mathcal{M}\) be a Banach space and \(\mathcal{C}\) a nonempty subset of \(\mathcal{M}\). A map** \(S:\mathcal{C} \to \mathcal{C}\) is said to satisfy condition (C) if

$$\begin{aligned} \frac{1}{2} \bigl\Vert u-S(u) \bigr\Vert \leq \Vert u-v \Vert \quad\text{{implies }} \bigl\Vert S(u) - S(v) \bigr\Vert \leq \Vert u-v \Vert \quad \forall u,v \in \mathcal{C}. \end{aligned}$$

Definition 8

[11]. Let \(\mathcal{C}\) be a nonempty subset of a Banach space \(\mathcal{M}\). A map** \(S:\mathcal{C} \to \mathcal{C}\) is said to fulfill condition \((E_{\mu})\) if there exists \(\mu \geq 1\) such that

$$ \bigl\Vert u-S(v) \bigr\Vert \leq \mu \bigl\Vert u-S(u) \bigr\Vert + \Vert u-v \Vert \quad \forall u,v \in \mathcal{C}. $$

We say that S satisfies condition (E) if it satisfies \((E_{\mu})\) for some \(\mu \geq 1\).

3 Class of map**s satisfying condition (L)

Llorens-Fuster and Moreno-Gálvez [21] introduced the following class of nonlinear map**s:

Definition 9

Let \(\mathcal{C}\) be a nonempty subset of a Banach space \((\mathcal{M},\|.\|)\). We say that a map** \(S:\mathcal{C} \to \mathcal{C}\) satisfies condition (L), (or it is an (L)-type map**), if the following two conditions hold:

  1. (1)

    If a set \(\mathcal{D} \subset \mathcal{C}\) is nonempty, closed, convex, and S-invariant, (i.e., \(S(\mathcal{D}) \subset \mathcal{D}\)), then there exists an a.f.p.s. for S in \(\mathcal{D}\).

  2. (2)

    For any a.f.p.s. \(\{u_{n}\}\) of S in \(\mathcal{C}\) and each \(u \in \mathcal{C}\)

    $$ \limsup_{n \to \infty} \bigl\Vert u_{n}-S(u) \bigr\Vert \leq \limsup_{n \to \infty} \Vert u_{n}-u \Vert . $$

In [21] it is shown that the above two conditions in the definition of (L)-type map**s are independent in nature.

It is proved in [21] that the class of (L)-type map**s contains strictly the following classes:

  1. (A)

    nonexpansive map**s;

  2. (B)

    Suzuki generalized nonexpansive map**s (cf. [27]);

  3. (C)

    generalized nonexpansve in many cases, see [21];

  4. (D)

    The class of map**s satisfying condition (E) that in turn satisfy condition (1) in the Definition 9 (cf. [11]).

Now, we consider a subclass of class of (L)-type map**s.

Definition 10

Let \(\mathcal{C}\) be a nonempty subset of a Banach space \((\mathcal{M},\|.\|)\) and a map** \(S:\mathcal{C} \to \mathcal{C}\) satisfies condition (L-1), (or it is an (L-1)-type map**), if the following two conditions hold:

  1. (1)

    If a set \(\mathcal{D} \subset \mathcal{C}\) is nonempty, closed, convex, and S-invariant, (i.e., \(S(\mathcal{D}) \subset \mathcal{D}\)), then there exists an a.f.p.s. for S in \(\mathcal{D}\).

  2. (2)

    For any a.f.p.s. \(\{u_{n}\}\) of S in \(\mathcal{C}\), there exists a sequence \(\{c_{n}\}\) in \([0,\infty )\) such that \(c_{n} \to 0\) as \(n \to \infty \) and each \(u \in \mathcal{C}\), we have

    $$ \bigl\Vert u_{n}-S(u) \bigr\Vert \leq \Vert u_{n}-u \Vert +c_{n}. $$
    (3.1)

Example 1

Let \((\ell ^{2},\|\cdot \|_{2} )\) be the Banach space of square-summable sequences endowed with its standard norm. Assume that \(B [0_{\mathcal{M}}, 1 ]\) is a unit ball centered at \(0_{\mathcal{M}}\) (zero element). Suppose that \(S: B [0_{\mathcal{M}}, 1 ] \rightarrow B [0_{ \mathcal{M}}, 1 ]\) is the map** given by the following definition:

$$ S(u)= \textstyle\begin{cases} \frac{1}{2} \frac{u}{ \Vert u \Vert } & u \in B [0_{\mathcal{M}}, 1 ] \backslash B [0_{\mathcal{M}}, \frac{1}{2} ], \\ 0_{\mathcal{M}} & u \in B [0_{\mathcal{M}}, \frac{1}{2} ].\end{cases} $$

In fact, the unique fixed point of S is \(0_{\mathcal{M}}\). We can have a.f.p.s. \(\{u_{n}\}\) given by \(u_{n} \equiv 0\). Suppose \(\{c_{n}\}= \{ \frac{1}{n} \}\) in \([0,\infty )\), then \(c_{n} \to 0\) as \(n \to \infty \). Then, if \(u \in B [0_{\mathcal{M}}, \frac{1}{2} ]\)

$$ \bigl\Vert u_{n}-S(u) \bigr\Vert = \bigl\Vert 0_{\mathcal{M}}-S(u) \bigr\Vert \leq \Vert 0_{\mathcal{M}}-u \Vert \leq \Vert u_{n}-u \Vert +c_{n} . $$

Again, if \(u \in B [0_{\mathcal{M}}, 1 ] \backslash B [0_{ \mathcal{M}}, \frac{1}{2} ]\)

$$ \bigl\Vert 0_{\mathcal{M}}-S(u) \bigr\Vert = \biggl\Vert \frac{1}{2} \frac{u}{ \Vert u \Vert } \biggr\Vert = \frac{1}{2} \leq \Vert u \Vert = \Vert 0_{ \mathcal{M}}-u \Vert \leq \Vert u_{n}-u \Vert +c_{n}. $$

On the other hand, \(u \in B [0_{\mathcal{M}}, 1 ]\) with \(\|u\|=\frac{1}{2}\) and \(v:=\frac{3}{2} u\), The map** S is not nonexpansive.

Proposition 1

Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1), then S is a map** satisfying condition (L).

Proof

The first conditions in both map**s are the same. Hence, we only compare the second conditions. Since map** S is a map** satisfying condition (L-1), then for any a.f.p.s. \(\{u_{n}\}\) of S in \(\mathcal{C}\), there exists a sequence \(\{c_{n}\}\) in \([0,\infty )\) such that \(c_{n} \to 0\) as \(n \to \infty \) and each \(u \in \mathcal{C}\)

$$ \bigl\Vert u_{n}-S(u) \bigr\Vert \leq \Vert u_{n}-u \Vert +c_{n}. $$

Taking lim sup on both sides, we obtain the desired result. □

In the next theorem, we present the structure of the fixed-point set of class of (L-1)-type map**s.

Theorem 4

Let \(\mathcal{C}\) be a nonempty, closed subset of a Banach space \(\mathcal{M}\) and \(S:\mathcal{C} \to \mathcal{C}\) a map** satisfying condition (L-1) with \(F(S) \neq \emptyset \). Then, the following implications hold:

  1. (i)

    \(F(S)\) is closed in \(\mathcal{C}\);

  2. (ii)

    If \(\mathcal{C}\) is convex and \(\mathcal{M}\) is strictly convex then \(F(S)\) is convex.

Proof

  1. (i)

    Let \(\{w_{n}\}\subseteq F(S)\) such that \(w_{n} \rightarrow w \in \mathcal{C}\) as \(n \rightarrow \infty \). Thus, \(S(w_{n}) = w_{n}\) and \(\{w_{n}\}\) is an a.f.p.s. for S in \(\mathcal{C}\). Since S is a (L-1)-type map**, we have

    $$ \bigl\Vert w_{n}-S(w) \bigr\Vert \leq \Vert w_{n}-w \Vert +c_{n}, $$

    making \(n \to \infty \), which implies that \(S(w)=w\) and \(F(S)\) is closed.

  2. (ii)

    See [8, Theorem 1].

 □

4 Demiclosedness principle in uniformly convex spaces

In this section, we present some results concerning the demiclosedness principle of a map** satisfying condition (L-1).

Lemma 7

Suppose \(\mathcal{C}\) is a bounded convex subset of a uniformly convex Banach space \(\mathcal{M}\) and \(S:\mathcal{C}\rightarrow \mathcal{M}\) is a map** satisfying condition (L-1). If \(\{u_{n}\}\) and \(\{v_{n}\}\) are approximate fixed-point sequences, then \(\{w_{n}\}=\{\frac{1}{2}(u_{n}+v_{n})\}\) is an approximate fixed-point sequence too.

Proof

Suppose the assertion of the lemma is false. Then, there exist sequences \(\{u_{n}\}\) and \(\{v_{n}\}\) satisfying \(\lim_{n\rightarrow \infty }\Vert u_{n}-S(u_{n})\Vert =0\) and \(\lim_{n\rightarrow \infty }\Vert v_{n}-S(v_{n})\Vert =0\) such that \(\Vert w_{n}-S(w_{n})\Vert \geq \varepsilon \) for some \(\varepsilon >0\) and every \(n \in \mathbb{N}\). We can assume by passing to a subsequence that

$$ \lim_{n\rightarrow \infty } \Vert u_{n}-v_{n} \Vert =2r>0. $$

It follows that

$$ \lim_{n\rightarrow \infty } \Vert u_{n}-w_{n} \Vert = \lim_{n\rightarrow \infty } \Vert v_{n}-w_{n} \Vert =r. $$

By the definition of map** S, for a.f.p.s. \(\{u_{n}\}\) of S in \(\mathcal{C}\), there exists a sequence \(\{c_{n,1}\}\) in \([0,\infty )\) such that \(c_{n,1} \to 0\) as \(n \to \infty \), we have

$$\begin{aligned} \bigl\Vert u_{n}-S(w_{n}) \bigr\Vert \leq & \Vert u_{n}-w_{n} \Vert +c_{n,1}. \end{aligned}$$
(4.1)

Similarly,

$$ \bigl\Vert v_{n}-S(w_{n}) \bigr\Vert \leq \Vert v_{n}-w_{n} \Vert +c_{n,2}, $$

where \(c_{n,2} \to 0\) as \(n \to \infty \). Choose \(s>0\) such that \(s<\frac{\varepsilon }{r}\). Hence, for sufficiently large n, we have

$$ s< \frac{\varepsilon }{c_{n,1} + \Vert u_{n}-w_{n} \Vert } $$
(4.2)

and

$$ s< \frac{\varepsilon }{c_{n,2} + \Vert v_{n}-w_{n} \Vert }. $$

Now,

$$ \biggl\Vert u_{n}-\frac{1}{2} \bigl(w_{n}+S(w_{n}) \bigr) \biggr\Vert = \biggl\Vert \frac{u_{n}-S(w_{n})+\frac{(u_{n}-v_{n})}{2}}{2} \biggr\Vert $$

and it can be seen that

$$ \bigl\Vert u_{n}-S(w_{n}) \bigr\Vert \leq \Vert u_{n}-w_{n} \Vert +c_{n,1}. $$

Now,

$$ \Vert u_{n}-w_{n} \Vert = \biggl\Vert u_{n} -\frac{1}{2}(u_{n}+v_{n}) \biggr\Vert = \frac{1}{2} \Vert u_{n}-v_{n} \Vert . $$

Thus,

$$ \biggl\Vert \frac{(u_{n}-v_{n})}{2} \biggr\Vert \leq \Vert u_{n}-w_{n} \Vert +c_{n,1} $$

and \(\Vert w_{n}-S(w_{n})\Vert \geq \varepsilon \). By the uniform convexity of \(\mathcal{M}\) (see Theorem 1), we have

$$\begin{aligned} \biggl\Vert u_{n}-\frac{1}{2} \bigl(w_{n}+S(w_{n}) \bigr) \biggr\Vert \leq & \biggl(1- \delta \biggl( \frac{\varepsilon }{c_{n,1} + \Vert u_{n}-w_{n} \Vert } \biggr) \biggr) \bigl(c_{n,1} + \Vert u_{n}-w_{n} \Vert \bigr). \end{aligned}$$

It is noted that the modulus of convexity, \(\delta (\varepsilon )\), is a nondecreasing function of ε, it follows that

$$ \biggl\Vert u_{n}-\frac{1}{2} \bigl(w_{n}+S(w_{n}) \bigr) \biggr\Vert \leq \bigl(1- \delta (s) \bigr) \bigl(c_{n,1}+ \Vert u_{n}-w_{n} \Vert \bigr). $$
(4.3)

Similarly,

$$\begin{aligned} \biggl\Vert v_{n}-\frac{1}{2} \bigl(w_{n}+S(w_{n}) \bigr) \biggr\Vert \leq & \biggl( 1- \delta \biggl( \frac{\varepsilon }{c_{n,2} + \Vert v_{n}-w_{n} \Vert } \biggr) \biggr) \\ &{}\times \bigl(c_{n,2} + \Vert v_{n}-w_{n} \Vert \bigr) \\ \leq & \bigl(1-\delta (s) \bigr) \bigl(c_{n,2} + \Vert v_{n}-w_{n} \Vert \bigr). \end{aligned}$$
(4.4)

By the triangle inequality, (4.3), and (4.4), we obtain

$$\begin{aligned} \Vert u_{n}-v_{n} \Vert \leq &\biggl\Vert u_{n}- \frac{1}{2}(w_{n}+S(w_{n}) \biggr\Vert + \biggl\Vert v_{n}-\frac{1}{2}(w_{n}+S(w_{n})\biggr\Vert \\ \leq & \bigl(1-\delta (s) \bigr) \bigl\{ \bigl(c_{n,1} + \Vert u_{n}-w_{n} \Vert \bigr)+ \bigl(c_{n,2}+ \Vert v_{n}-w_{n} \Vert \bigr) \bigr\} . \end{aligned}$$

Letting \(n\rightarrow \infty \), we obtain \(2r\leq 2r(1-\delta (s))\), a contradiction and this completes the proof. □

Proposition 2

Suppose \(\mathcal{C}\) is a bounded, closed, and convex subset of a uniformly convex space. Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1). Then, S has a fixed point.

Proof

See [21, Theorem 4.4]. □

Theorem 5

(Demiclosedness principle). Suppose \(\mathcal{C}\) is a closed, convex subset of a uniformly convex space. Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1). Then, the map** \(G=I-S\) is demiclosed on \(\mathcal{C}\).

Proof

Let \(\{u_{n}\}\) be a sequence in \(\mathcal{C}\) such that \(\{u_{n}\}\) converges weakly to \(u^{\dagger}\) and \(\lim_{n\rightarrow \infty }\|u_{n}-S(u_{n})-w\|=0\). Without loss of generality, we assume \(w=0\), as limits are preserved under the translation. Define \(\mathcal{C}_{n}= \overline{\operatorname{conv}}\{u_{n},u_{n+1},\dots \}\), using Proposition 2 on set \(\mathcal{C}_{n}\), there exists \(y_{n} \in \mathcal{C}_{n}\) such that \(S(y_{n})=y_{n}\). Since any weak subsequential limit of \(y_{n}\) lies in \(\bigcap_{n=1}^{\infty}\mathcal{C}_{n}=\{u^{\dagger}\}\), it implies that \(y_{n}\) converges weakly to \(u^{\dagger}\). Therefore, \(u^{\dagger}\) is in the weak closure of the fixed-point set \(F(S)\). Since \(\mathcal{M}\) is uniformly convex, \(\mathcal{M}\) is both reflexive and strictly convex. From Theorem 4, fixed-point set \(F(S)\) is closed and convex, so weakly closed and \(u^{\dagger} \in F(S)\). This completes the proof. □

5 Some fixed-point theorems

In this section, we present some fixed-point results for the class of map**s satisfying condition (L-1).

Theorem 6

Suppose \(\mathcal{C}\) is a closed, convex subset of a uniformly convex space. Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1). If \(\{u_{n}\}\) is an a.f.p.s. for S such that it converges weakly to \(u^{\dagger} \in \mathcal{C}\), then \(u^{\dagger}\) is a fixed point of S.

Proof

It can be easily seen from Theorem 5 that map** \(I-S\) is demiclosed at 0. From the demiclosedness principle it follows that \(u^{\dagger}\) is a fixed point of S. □

Remark 1

The above theorem should be compared with [21, Theorem 4.6] that asserts the same conclusion in view of the Opial property.

Theorem 7

Let \(\mathcal{C}\) be a nonempty bounded, closed, and convex subset of \(\mathcal{M}_{2}\) and \(S:\mathcal{C} \to \mathcal{C}\) a map** satisfying condition (L-1). Assume the following conditions hold:

  1. (1)

    If \(\mathcal{D}\) is minimal with respect to S, and there is an a.f.p.s. \(\{u_{n}\}\) in \(\mathcal{D}\), then \(u_{n}-u_{n+1} \to 0\) as \(n \to \infty \);

  2. (2)

    If \(\mathcal{D}\) is minimal with respect to S, and \(\{u_{n}\}\) is an a.f.p.s. in \(\mathcal{D}\), then \(|u_{n}-u|_{2} \to d=\operatorname{diam}_{|.|_{\beta}}(\mathcal{C})\) for all \(u \in \mathcal{D}\).

Then, S has a fixed point.

Proof

By the application of Zorn’s lemma there is a nonempty, bounded, closed, convex, and S-invariant subset \(\mathcal{D}\) of \(\mathcal{C}\) with no proper subsets, so \(\mathcal{D}\) is minimal with respect to S. Let \(d=\operatorname{diam}_{|.|_{\beta}}(\mathcal{C})\) and assume, for a contradiction, that \(d>0\). Let \(\{u_{n}\}\) be an a.f.p.s. in \(\mathcal{D}\) such that \(u_{n}-u_{n+1} \to 0\) as \(n \to \infty \). Let \(w \in \mathcal{D}\) denote the \(\|.\|_{2}\)-asymptotic center of \(\{u_{n}\}\) in \(\mathcal{D}\). By Lemma 4, we have

$$ \limsup_{n\rightarrow \infty } \Vert u_{n}-w \Vert _{2}^{2} \geq \frac{d^{2}}{2}. $$
(5.1)

Without loss of generality, we may assume there is a subsequence \(\{u_{n_{k}}\}\) of \(\{u_{n}\}\) with \(u_{n_{k}} \rightharpoonup u \in \mathcal{D}\) and

$$ \lim_{k\rightarrow \infty } \Vert u_{n_{k}}-w \Vert _{2}^{2}= \limsup_{n\rightarrow \infty } \Vert u_{n}-w \Vert _{2}^{2}. $$

Again, take a subsequence \(\{u_{m_{k}}\}\) of \(\{u_{n}\}\) with \(u_{n_{k}} \rightharpoonup v \in \mathcal{D}\) and

$$ \lim_{k\rightarrow \infty } \Vert u_{m_{k}}-u \Vert _{2}^{2}= \limsup_{n\rightarrow \infty } \Vert u_{n}-u \Vert _{2}^{2}. $$

By (5.1) and Lemma 3(i), we obtain

$$\begin{aligned} d^{2} \geq & \lim_{k\rightarrow \infty } \Bigl\{ \lim _{p \rightarrow \infty } \Vert u_{n_{k}}-u_{m_{p}} \Vert _{2}^{2} \Bigr\} \\ =& \limsup_{n\rightarrow \infty } \Vert u_{n}-u \Vert _{2}^{2} + \limsup_{n\rightarrow \infty } \Vert u_{n}-w \Vert _{2}^{2}- \Vert w-u \Vert _{2}^{2} \\ \geq & 2 \limsup_{n\rightarrow \infty } \Vert u_{n}-w \Vert _{2}^{2} \geq d^{2}. \end{aligned}$$

From the above inequalities, we have the following:

$$ \limsup_{n\rightarrow \infty } \Vert u_{n}-w \Vert _{2}^{2} = \frac{d^{2}}{2} $$
(5.2)

and

$$ \lim_{k\rightarrow \infty } \Bigl\{ \lim_{p\rightarrow \infty } \Vert u_{n_{k}}-u_{m_{p}} \Vert _{2}^{2} \Bigr\} = d^{2}. $$
(5.3)

Now, we show that

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup_{p \rightarrow \infty } \biggl\Vert \frac{1}{2}(u_{n_{k}}+u_{m_{p}})-u \biggr\Vert _{\infty} \biggr\} =\frac{d}{2} \quad\text{{for all }}u \in \mathcal{C}. $$

Take \(\Gamma _{k} =u_{n_{k}}\) and \(\Delta _{k}=u_{m_{k}}\). From Lemma 2, for \(k,p \in \mathbb{N}\), we have

$$\begin{aligned} & \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)- \frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr\Vert _{2}^{2} + \frac{1}{4} \Vert \Gamma _{k}-\Delta _{p} \Vert _{2}^{2} \\ &\quad= \frac{1}{2} \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\Gamma _{k} \biggr\Vert _{2}^{2}\frac{1}{2} \biggl\Vert S \biggl( \frac{1}{2}(\Gamma _{k} +\Delta _{p}) \biggr)-\Delta _{p} \biggr\Vert _{2}^{2}. \end{aligned}$$
(5.4)

Now,

$$ \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\Gamma _{k} \biggr\Vert _{2} \leq \biggl\vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\Gamma _{k} \biggr\vert _{2}. $$

Since \(\{\Gamma _{k}\}\) is a.f.p.s for S, from the definition of condition (L-1), we have

$$ \biggl\vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\Gamma _{k} \biggr\vert _{2} \leq \biggl\vert \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\Gamma _{k} \biggr\vert _{2}+c_{n,1}, $$

where \(c_{n,1} \to 0\) as \(n \to \infty \). From the above inequality, we obtain

$$\begin{aligned} \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+ \Delta _{p}) \biggr)-\Gamma _{k} \biggr\Vert _{2} \leq & \frac{1}{2} \vert \Gamma _{k}-\Delta _{p} \vert _{2} +c_{n,1} \\ \leq & \frac{d}{2} +c_{n,1} \end{aligned}$$
(5.5)

and, similarly,

$$ \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+ \Delta _{p}) \biggr)-\Delta _{p} \biggr\Vert _{2}^{2} \leq \frac{d}{2} +c_{n,2}, $$
(5.6)

where \(c_{n,2} \to 0\) as \(n \to \infty \). Using (5.5) and (5.6) in (5.4), we have

$$\begin{aligned} \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)- \frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr\Vert _{2}^{2} + \frac{1}{4} \Vert u_{k}-v_{p} \Vert _{2}^{2} \leq \frac{1}{2} \biggl(\frac{d}{2} +c_{n,1} \biggr)^{2}+ \frac{1}{2} \biggl(\frac{d}{2} +c_{n,2} \biggr)^{2}. \end{aligned}$$

Since \(\lim_{k\rightarrow \infty } \{\lim_{p\rightarrow \infty } \|\Gamma _{k}-\Delta _{p}\|_{2}^{2}\}= d^{2}\), from the above inequality, we obtain

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup_{p \rightarrow \infty } \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr\Vert _{2}^{2} \biggr\} =0. $$

Thus,

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup _{p \rightarrow \infty } \biggl\Vert S \biggl(\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr)-\frac{1}{2}(\Gamma _{k}+\Delta _{p}) \biggr\Vert _{\infty} \biggr\} =0. $$
(5.7)

Assume there exists \(u \in \mathcal{C}\) such that

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup _{p \rightarrow \infty } \biggl\Vert \frac{1}{2}(\Gamma _{k}+ \Delta _{p})-u \biggr\Vert _{\infty} \biggr\} < \Omega < \frac{d}{2}. $$
(5.8)

From (5.7) and (5.8), we can choose subsequences \(\{\Gamma _{k_{q}}\}\) and \(\{\Gamma _{p_{q}}\}\) such that for \(q \in \mathbb{N}\):

$$ \bigl\vert S(z_{q})-z_{q} \bigr\vert _{2} \leq \frac{2}{q} \quad\text{{and}}\quad \Vert z_{q}-u \Vert _{ \infty} \leq \Omega, \quad\text{{where }} z_{q}= \frac{1}{2}(\Gamma _{k_{q}}+ \Delta _{p_{q}}). $$

Therefore, \(\lim_{q \rightarrow \infty }|S(z_{q})-z_{q}|_{2}=0\) and \(\limsup_{q \rightarrow \infty } \|z_{q}-u\|_{\infty} \leq \Omega <\frac{d}{2}\), which contradicts Lemma 5 by assumption (2). Hence,

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup _{p \rightarrow \infty } \biggl\Vert \frac{1}{2}(u_{n_{k}}+u_{m_{p}})-u \biggr\Vert _{\infty} \biggr\} =\frac{d}{2} \quad\text{{for all }}u \in \mathcal{C}. $$
(5.9)

In particular, it yields,

$$ \limsup_{k\rightarrow \infty } \biggl\{ \limsup_{p \rightarrow \infty } \biggl\Vert \frac{1}{2}(u_{n_{k}}+u_{m_{p}})- \frac{1}{2}(u+v) \biggr\Vert _{2}^{2} \biggr\} \geq \frac{d^{2}}{4}. $$

From Lemma 2, it follows that

$$\begin{aligned} \frac{d^{2}}{4} \leq & \limsup_{k\rightarrow \infty } \biggl\{ \limsup_{p\rightarrow \infty } \biggl\Vert \frac{1}{2}(u_{n_{k}}+u_{m_{p}})- \frac{1}{2}(u+v) \biggr\Vert _{2}^{2} \biggr\} \\ =& \limsup_{k\rightarrow \infty } \biggl\{ \limsup_{p \rightarrow \infty } \biggl\Vert \frac{1}{2}(u_{n_{k}}-u)+\frac{1}{2}(u_{m_{p}}-v) \biggr\Vert _{2}^{2} \biggr\} \\ =& \frac{1}{4} \lim_{k\rightarrow \infty } \Vert u_{n_{k}}-u \Vert _{2}^{2}+ \frac{1}{4} \lim_{p\rightarrow \infty } \Vert u_{m_{p}}-v \Vert _{2}^{2}. \end{aligned}$$
(5.10)

Since \(u_{n_{k}} \rightharpoonup u \in \mathcal{D}\) as \(k \to \infty \), then for each \(k \in \mathbb{N}\),

$$\begin{aligned} \Vert u_{n_{k}}-w \Vert _{2}^{2} =& \Vert u_{n_{k}}-u+u-w \Vert _{2}^{2} \\ =& \Vert u_{n_{k}}-u \Vert _{2}^{2} +2 \langle u_{n_{k}}-u,u-w \rangle + \Vert u-w \Vert _{2}^{2}. \end{aligned}$$

From (5.2), we have

$$ \lim_{k\rightarrow \infty } \Vert u_{n_{k}}-u \Vert _{2}^{2}= \frac{d^{2}}{2}- \Vert w-u \Vert _{2}^{2}. $$
(5.11)

Similarly,

$$ \lim_{p \rightarrow \infty } \Vert u_{m_{p}}-v \Vert _{2}^{2}= \frac{d^{2}}{2}- \Vert w-v \Vert _{2}^{2}. $$
(5.12)

Using (5.11) and (5.12) in (5.10) it follows that

$$\begin{aligned} \frac{d^{2}}{4} \leq & \frac{1}{4} \biggl(\frac{d^{2}}{2}- \Vert w-u \Vert _{2}^{2} \biggr)+\frac{1}{4} \biggl( \frac{d^{2}}{2}- \Vert w-v \Vert _{2}^{2} \biggr) \\ \leq & \frac{d^{2}}{4}- \frac{1}{4} \bigl( \Vert w-u \Vert _{2}^{2}+ \Vert w-v \Vert _{2}^{2} \bigr) \end{aligned}$$

and it proves that \(u=v=w\). Take \(\sigma _{k}=u_{n_{k}}-w\) and \(\varrho _{k}=u_{m_{k}}-w\). Since \(\{u_{m_{k}}\}\) converges weakly to \(v=w\),

$$ \varrho _{k} \rightharpoonup 0 \quad\text{{as }} k \to \infty. $$
(5.13)

Since \(|u_{m_{k}}-w|_{2} \to d\) and \(|u_{n_{k}}-w|_{2} \to d\) as \(k \to \infty \), from Lemma 5, the following hold:

$$ \Vert \varrho _{k} \Vert _{\infty} \to \frac{d}{2} \quad\text{{and}}\quad \Vert \sigma _{k} \Vert _{\infty} \to \frac{d}{2}\quad \text{{as }} k \to \infty. $$
(5.14)

By the definition of d, the following condition is satisfied:

$$ \text{{for each }} k, p \in \mathbb{N},\quad \Vert \sigma _{k}-\varrho _{k} \Vert _{\infty} \leq \frac{d}{2}. $$
(5.15)

From (5.3), we have

$$\begin{aligned} \lim_{k\rightarrow \infty } \Bigl\{ \lim_{p\rightarrow \infty } \Vert u_{n_{k}}-u_{m_{p}} \Vert _{2}^{2} \Bigr\} =& \lim_{k \rightarrow \infty } \Bigl\{ \lim_{p\rightarrow \infty } \bigl\Vert (u_{n_{k}}-w)-(u_{m_{p}}-w) \bigr\Vert _{2}^{2} \Bigr\} \\ =&\lim_{k\rightarrow \infty } \Bigl\{ \lim_{p \rightarrow \infty } \Vert \sigma _{k}-\varrho _{p} \Vert _{2}^{2} \Bigr\} = d^{2}. \end{aligned}$$
(5.16)

From (5.9), we obtain \(\frac{1}{2} \limsup_{k\rightarrow \infty } \{\limsup_{p\rightarrow \infty } \Vert (u_{n_{k}}-w)+(u_{m_{p}}-w) \Vert _{\infty} \}=\frac{d}{2} \) and it follows that

$$ \limsup_{k\rightarrow \infty } \Bigl\{ \limsup _{p \rightarrow \infty } \Vert \sigma _{k}+\varrho _{p} \Vert _{ \infty} \Bigr\} =d. $$
(5.17)

From Lemma 2, for all \(k,p \in \mathbb{N}\), we have

$$ \Vert \sigma _{k}+\varrho _{p} \Vert _{2}^{2} =2 \Vert \sigma _{k} \Vert _{2}^{2} + \Vert \varrho _{p} \Vert _{2}^{2} - \Vert \sigma _{k} -\varrho _{p} \Vert _{2}^{2}. $$

In view of (5.13), (5.14), (5.15), (5.16), (5.17), and Lemma 6, this implies

$$ \lim_{k\rightarrow \infty } \Bigl\{ \lim_{p\rightarrow \infty } \Vert \sigma _{k} + \varrho _{p} \Vert _{2}^{2} \Bigr\} = d^{2}, $$

which is impossible. This completes the proof. □

Theorem 8

Let \(\mathcal{C}\) be a nonempty, bounded, closed (resp., weak closed), and convex subset of a reflexive Banach space (resp., the conjugate of a separable Banach space) \(\mathcal{M}\). Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1). Suppose that the relationis uniformly approximately symmetric (resp., uniformly weak approximately symmetric) in \(\mathcal{M}\), then \(F(S) \neq \emptyset \).

Proof

By the application of Zorn’s lemma there exists a nonempty, bounded, closed, convex, and S-invariant subset \(\mathcal{D}\) of \(\mathcal{C}\) with no proper subsets, so \(\mathcal{D}\) is minimal with respect to S. Since S satisfies condition (L-1), there exists an a.f.p.s. \(\{u_{n}\}\) for S in \(\mathcal{D}\). By the reflexiveness of \(\mathcal{M}\), there exists a subsequence \(\{u_{n_{k}}\}\) of \(\{u_{n}\}\) such that \(u_{n_{k}}\) converges weakly to \(u^{\dagger}\). After possible extraction of a subsequence, if necessary, we assume that \(\lim_{k \to \infty}\|x_{n_{k}}-x^{\dagger}\|=\Theta \). Take \(v=u^{\dagger}-S(u^{\dagger})\). If \(\Theta =0\) or \(v=0\), then \(S(u^{\dagger})=u^{\dagger}\) and the proof is completed. Therefore, we assume that \(\Theta >0\) and \(v \neq 0\). Following largely the same argument in [18, Theorem 1] let \(\varepsilon =\frac{1}{2\Theta}\). By the assumptions, there exists a closed (resp., weak closed) linear subspace \(\mathcal{Y}\) such that conditions (i) in Definition 4 and (iii) in Definition 6 are satisfied. This implies that there exists a \(\delta >0\) such that

$$ \vert \mu \vert \leq \Vert v+\mu u \Vert - \vert \mu \vert \delta $$
(5.18)

for every \(u \in \mathcal{Y}\), \(\|u\|=1\) and each μ with \(|\mu | \leq 2\Theta \). Further, the subspace spanned by \(\mathcal{Y}\) and v has a finite-dimensional complement \(\mathcal{Z}\). Therefore, for each \(k \in \mathbb{N}\), \(\sigma _{n_{k}} \in \mathcal{Y}\) and \(\varrho _{n_{k}} \in \mathcal{Z}\), we have

$$ u_{n_{k}}-u^{\dagger}=\mu _{n_{k}} v+\sigma _{n_{k}}+\varrho _{n_{k}}. $$
(5.19)

Since \(\mathcal{Z}\) is a finite-dimensional space and noting the convergence of \(u_{n_{k}}-u^{\dagger}\), it follows that \(\mu _{n_{k}} \to 0\) and \(\|\varrho _{n_{k}}\| \to 0\) as \(k \to \infty \). Thus, \(\|\sigma _{n_{k}}\| \to \Theta \) and for sufficiently large k, \(\frac{\|\sigma _{n_{k}}\|}{(1+\mu _{n_{k}})} \leq 2 \Theta \). From (5.18) and (5.19), we have

$$\begin{aligned} \bigl\Vert u_{n_{k}}-S \bigl(u^{\dagger} \bigr) \bigr\Vert =& \bigl\Vert u_{n_{k}}-u^{\dagger}+u^{\dagger}-S \bigl(u^{ \dagger} \bigr) \bigr\Vert = \bigl\Vert (1+\mu _{n_{k}}) y+ \sigma _{n_{k}}+\varrho _{n_{k}} \bigr\Vert \\ \geq & \bigl\Vert (1+\mu _{n_{k}}) v+\sigma _{n_{k}} \bigr\Vert - \Vert \varrho _{n_{k}} \Vert \\ \geq & \vert 1+\mu _{n_{k}} \vert \biggl\Vert v+ \biggl( \frac{ \Vert \sigma _{n_{k}} \Vert }{(1+\mu _{n_{k}})} \biggr) \frac{\sigma _{n_{k}}}{ \Vert \sigma _{n_{k}} \Vert } \biggr\Vert - \Vert \varrho _{n_{k}} \Vert \\ \geq & \Vert \sigma _{n_{k}} \Vert (1+\delta )- \Vert \varrho _{n_{k}} \Vert . \end{aligned}$$
(5.20)

Since the map** S satisfies condition (L-1), we have

$$ \bigl\Vert u_{n_{k}}-S \bigl(u^{\dagger} \bigr) \bigr\Vert \leq \bigl\Vert u_{n_{k}}-u^{\dagger} \bigr\Vert + c_{k}, $$
(5.21)

where \(c_{k} \to 0\) as \(k \to \infty \). Making \(k \to \infty \), \(\|u_{n_{k}}-S(u^{\dagger})\| \to \Theta \). From (5.20), noting that \(\|u_{n_{k}}\| \to \Theta \) and \(\|v_{n_{k}}\| \to 0\) as \(k \to \infty \) we obtain the following inequality

$$ \Theta \geq (1+\delta ) \Theta, $$

which is a contradiction. Therefore, \(\Theta =0\), and this completes the proof. □

Corollary 1

Let \(\mathcal{C}\) be a convex, bounded, and weak closed subset of \(\ell _{1}\) or the James space \(J_{0}\). Let \(S:\mathcal{C} \to \mathcal{C}\) be a map** satisfying condition (L-1). Then, S has a fixed point in \(\mathcal{C}\).

We conclude the paper by posing the following interesting problem.

Kassay [19] showed that the converse of the above theorem is also true. More precisely, a reflexive Banach space having normal structure can be characterized by the fixed-point property for Jaggi-nonexpansive map**s.

5.1 Problem

Can a reflexive Banach space having asymptotic normal structure be characterized by the fixed-point property for map** satisfying condition (L-1)?