Skip to main content

Fixed point theorems for generalized \((\alpha , \phi )\)-Meir–Keeler type hybrid contractive mappings via simulation function in b-metric spaces

Abstract

In this paper, we introduce the notion of generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mappings of type I and II via simulation function and establish fixed point theorems for such mappings in the setting of complete b-metric spaces. Our results extend and generalize many related fixed point results in the existing literature. Finally, we provide an example in support of our main finding.

1 Introduction

Fixed point theory is one of the most important topics in development of nonlinear and mathematical analysis in general. Also, fixed point theory has been effectively used in many other branches of science such as chemistry, physics, biology, economics, computer science, all engineering fields, and so on. In 1922, Banach [1] introduced a well-known fixed point result, now called Banach contraction principle, which is one of the pivotal results in nonlinear analysis. Due to its importance and fruitful applications, several authors have obtained many interesting extensions and generalizations of the Banach contraction principle in several direction (see, e.g., [2, 3]). These generalizations are achieved either by using contractive conditions or by imposing some additional conditions on the ambient spaces. For example, one of the important and peculiar generalizations is due to Meir and Keeler [4]. The class of Meir–Keeler contractions consists of the class of Banach contractions and many other classes of nonlinear contractions (see, for example, [5]). Meir and Keeler’s theorem was the originator of further exploration in metric fixed point theory. Later on, Meir–Keeler contraction mapping has been generalized by several authors in several ways. For more works in this line of research, we refer to [68], as well as [914]. On the other hand, the notion of a b-metric space was introduced by Bakhtin [15] and Czerwik [16] as a generalization of metric spaces. Since then, several papers have been published on the fixed point theory in such spaces which are interesting extensions and generalizations of the Banach contraction principle. For further works in the setting of b-metric spaces and their generalization, we refer the readers to [1742]. In 2020, Karapinar et al. [43] studied fixed point results for the Meir–Keeler contraction via simulation function in the setting of metric spaces. Inspired and motivated by the work of Karapinar et al. [43], the main objectives of this research is to introduce the notion of generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mappings of type I and II via simulation function and establish fixed point theorems for the introduced mappings in the setting of b-metric spaces. The present results extend and generalize the results of Karapinar et al. [43] and many other related results in the existing literature.

2 Preliminaries

In what follows we recall basic definitions and results on the topics which we use in the sequel.

Notations 1

Throughout this paper, we denote \(\mathbb{R}^{+}\), \(\mathbb{R}\) and \(\mathbb{N}\) respectively by

  • \(\mathbb{R}^{+} = [0,\infty ) \) – the set of all non-negative real numbers;

  • \(\mathbb{R}\) – the set of all real numbers;

  • \(\mathbb{N}\) – the set of all natural numbers.

Khojasteh et al. [44] introduced the notion of a simulation function as follows.

Definition 1

([44])

A weak simulation function is a mapping \(\zeta : \mathbb{R}^{+} \times \mathbb{R}^{+} \rightarrow \mathbb{R}\) satisfying the following conditions:

\((\zeta _{1})\):

\(\zeta (0, 0) = 0\);

\((\zeta _{2})\):

\(\zeta (t, s) < s-t\) for all \(t, s > 0\).

Note

Throughout this paper we denote by \(Z_{w}\) the family of all simulation functions \(\zeta : \mathbb{R}^{+} \times \mathbb{R}^{+} \rightarrow \mathbb{R}\). Due to the axiom \((\zeta _{2})\), we have \(\zeta (t, t) < 0\) for all \(t > 0\).

Recently, Suzuki [45] introduced the following class of mappings and proved the following interesting fixed point result to extend the coverage of Meir–Keeler theorem in the setting of metric spaces. Let \((X, d)\) be a metric space and \(T : X \rightarrow X\) be a self-mapping. Define a mapping \(M : X \times X \rightarrow \mathbb{R}^{+}\) as follows:

$$ M(x, y) = \max \biggl\{ d(x, y), d(x, Tx), d(y, Ty), \frac{d(x, Ty) + d(y, Tx)}{2} \biggr\} . $$

And let \(p : X \times X \rightarrow \mathbb{R}^{+}\) be a mapping satisfies the following conditions:

\((P^{1}_{p}: M)\):

\(x\neq y\) and \(d(x, Tx) \leq d(x, y)\) imply \(p(x, y) \leq M(x, y)\);

\((P^{2}_{p}: c)\):

\(x_{n} \neq y\), \(\lim_{n\rightarrow \infty}d(x_{n}, y) = 0\), and \(\lim_{n\rightarrow \infty}d(x_{n}, Tx_{n}) = 0\) imply

$$ \limsup_{n\rightarrow \infty}d(x_{n}, y)\leq cd(y, Ty),\quad \text{where } c \in [0, 1). $$

Theorem 1

([45])

Let T be a self-mapping on a complete metric space \((X, d)\). Let \(p : X \times X \rightarrow \mathbb{R}^{+} \) be mapping that satisfies the conditions \((P^{1}_{p} : M)\) and \((P^{2}_{p} : c)\) defined above. Suppose also that the following are satisfied:

  1. (i)

    For any \(\epsilon > 0\), there exists \(\delta (\epsilon ) > 0\) such that \(x \neq y\) and \(p(x, y) < \epsilon +\delta (\epsilon )\) imply \(d(Tx, Ty)\leq \epsilon \);

  2. (ii)

    \(x \neq y\) and \(p(x, y) > 0\) imply \(d(Tx, Ty) < p(x, y)\).

Then T has a unique fixed point z. Moreover, the sequence \(\{T^{n}x\}\) converges to z for all \(x \in X\).

Bakhtin [15] and Czerwik [16] defined a b-metric space as follows.

Definition 2

([15, 16])

Let X be a nonempty set and \(s \geq 1 \) be a given real number. A function \(d:X \times X \rightarrow \mathbb{R}^{+} \) is said to be a b-metric if and only if for all \(x, y, z \in X \), the following conditions are satisfied:

  1. (a)

    \(d(x,y) = 0 \) if and only if \(x=y \);

  2. (b)

    \(d(x,y)=d(y,x) \);

  3. (c)

    \(d(x,z) \leq s[d(x,y)+d(y,z)] \).

The pair \((X,d) \) is called a b-metric space.

It should be noted that the class of b-metric spaces is effectively larger than that of metric spaces. A metric space is a b-metric with \(s = 1 \). But, in general, the converse is not true.

Example 1

([32])

Let \(X =\mathbb{R} \) and \(d:X \times X\rightarrow \mathbb{R}^{+} \) be given by \(d(x,y)=|x - y|^{2} \) for \(x ,y \in X \), then d is a b-metric on X with \(s = 2 \) but it is not a metric on X since for \(x=2\), \(y=4 \), and \(z=6 \), we have

$$ d(2,6) > d(2,4)+d(4,6). $$

Hence, the triangle inequality for a metric does not hold.

Definition 3

([46])

Let X be a b-metric space and \(\{x_{n}\} \) a sequence in X. We say that

  1. 1.

    \(\{x_{n}\}\) is b-convergent to \(x \in X \) if \(d(x_{n},x )\rightarrow 0 \) as \(n \rightarrow \infty \).

  2. 2.

    \(\{x_{n} \} \) is a b-Cauchy sequence if \(d(x_{n} ,x_{m} ) \rightarrow 0 \) as \(n,m \rightarrow \infty \).

  3. 3.

    \((X,d )\) is b-complete if every b-Cauchy sequence in X is b-convergent.

Definition 4

([47])

Let \((X,d)\) be a b-metric space with the coefficient \(s \geq 1 \) and let \(T\colon X\rightarrow X \) be a given mapping. We say that T is b-continuous at \(x_{0} \in X \) if and only if for every sequence \({x_{n}} \in X\) such that \({x_{n}}\rightarrow x_{0} \) as \(n\rightarrow \infty \), we have \(Tx_{n} \rightarrow Tx_{0} \) as \(n\rightarrow \infty \). If T is b-continuous at each point \(x \in X \), then we say that T is b-continuous on X.

In general, a b-metric is not necessarily continuous.

Example 2

([48])

Let \(X= \mathbb{N} \cup \{\infty \} \).

Define a mapping \(d : X \times X \rightarrow \mathbb{R}^{+}\) as follows:

$$ d(m,n)= \textstyle\begin{cases} 0&\text{if $ m=n $,} \\ |\frac{1}{m}-\frac{1}{n}|& \text{if one of $ m$ and $n$ is even and the other even or $\infty $}, \\ 5& \text{if one of $ m$ and $n$ is odd and the other odd or $\infty $}, \\ 2&\text{otherwise.} \end{cases} $$

Observe that \(d(m,p)\leq \frac{3}{2}[d(m,n)+d(n,p)]\) for all \(m,n\), \(p\in X \).

Then \((X,d)\) is a b-metric space with \(s = \frac{3}{2} \).

If we choose \(x_{n} = 2n\) for each \(n\in \mathbb{N}\), then

$$ d(x_{n},\infty ) =d(2n,\infty ) = \frac{1 }{2n}\rightarrow 0 \quad \text{as } n\rightarrow \infty , $$

that is, \(x_{n} \rightarrow \infty\) as \(n\rightarrow \infty \).

But \(\lim_{n\rightarrow \infty} d(x_{n},1) = 2\neq 5= d(\infty ,1)\). Hence, d is not continuous.

The following are definitions of α-orbital admissible and triangular α-orbital admissible mappings.

Definition 5

([49])

Let X be a nonempty set and \(\alpha : X \times X \rightarrow \mathbb{R}^{+}\) a function. A mapping \(T: X \rightarrow X\) is said to be α-orbital admissible if, for all \(x\in X\), \(\alpha (x,Tx) \geq 1\) implies \(\alpha (Tx,T^{2} x) \geq 1\).

Definition 6

([49])

Let X be a nonempty set, \(T : X \rightarrow X\), and \(\alpha : X \times X \rightarrow \mathbb{R}^{+}\). We say that T is triangular α-orbital admissible if:

  1. (i)

    T is α-orbital admissible;

  2. (ii)

    for all \(x, y \in X\), \(\alpha (x, y) \geq 1\) and \(\alpha (y, T y) \geq 1\) imply that \(\alpha (x, T y) \geq 1\).

In 2020, Karapinar et al. [43] introduced the class of hybrid contraction mappings of type I and II and studied fixed point results for such mappings.

Definition 7

([43])

Let T be a self-mapping on a metric space \((X, d)\) and \(\zeta \in Z_{w}\). Suppose that \(p : X \times X \rightarrow \mathbb{R}^{+}\) is a function that satisfies only \((P^{1} p : M)\). Then T is called a hybrid contraction of type I if the following conditions are fulfilled:

  1. (a)

    For any \(\epsilon > 0\), there exists \(\delta (\epsilon ) > 0\) such that \(x \neq y\) and \(p(x, y) < \epsilon + \delta (\epsilon )\) imply \(d(Tx, Ty) \leq \epsilon \);

  2. (b)

    \(x \neq y\) and \(p(x, y) > 0\) imply \(\zeta (\alpha (x, y)d(Tx, Ty), p(x, y)) \geq 0\).

Let a mapping \(N : X \times X \rightarrow \mathbb{R}^{+}\) be defined as follows:

$$ N(x, y) = \max \biggl\{ d(y, Ty)\frac{1+d(x,Tx)}{1+d(x,y)}, d(x, y) \biggr\} , $$

where T is a self-mapping defined on a metric space \((X, d)\). We notice that, for any \(x, y \in X\) with \(x = y\), we have \(0 = d(Tx, Ty) \leq N(x, y)\). Moreover, if \(x \neq y\), then \(N(x, y) > 0\).

Definition 8

([43])

Let T be a self-mapping on a metric space \((X, d)\) and \(\zeta \in Z_{w}\). Suppose that \(p : X \times X \rightarrow \mathbb{R}^{+}\) is a function that satisfies \((P^{1} p : N)\) and \((P^{2} p : c)\), for all \(c \in [0, 1)\). Then T is called a hybrid contraction of type II if the following conditions are satisfied:

  1. (a)

    For any \(\epsilon > 0\) there exists \(\delta (\epsilon ) > 0\) such that \(x \neq y\) and \(p(x, y) < \epsilon + \delta (\epsilon )\) imply \(d(Tx, Ty) \leq \epsilon \);

  2. (b)

    \(x \neq y\) and \(p(x, y) > 0\) imply \(\zeta (\alpha (x, y)d(Tx, Ty), p(x, y)) \geq 0\).

Theorem 2

([43])

Let \((X, d)\) be a complete metric space and \(T : X \rightarrow X\) be a hybrid contraction of type I. Assume that the following conditions are satisfied:

  1. (i)

    T is triangular α-orbital admissible;

  2. (ii)

    there exists \(x_{0} \in X\) such that \(\alpha (x_{0}, Tx_{0}) \geq 1\);

  3. (iii)

    T is continuous.

Then T has a fixed point u. Moreover, \(\{T^{n}x\}\) converges to u for all \(x \in X\).

Theorem 3

([43])

Let \((X, d)\) be a complete metric space and \(T : X \times X\) be a hybrid contraction of type II. Assume that the following conditions are fulfilled:

  1. (i)

    T is triangular α-orbital admissible;

  2. (ii)

    there exists \(x_{0} \in X\) such that \(\alpha (x_{0}, Tx_{0}) \geq 1\);

  3. (iii)

    either T is continuous;

  4. (iv)

    or \(T^{2}\) is continuous and \(\alpha (u,Tu)\geq 1\);

  5. (v)

    or \((X, d)\) is regular.

Then T has a fixed point u. Moreover, \(\{T^{n}x\}\) converges to u for all \(x \in X\).

3 Results

In this section, first we introduce generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I in the setting of b-metric spaces and prove fixed point results for such mappings.

Note

In this section, we denote the class of mappings Ψ by

$$ \varPsi =\bigl\{ \phi :\mathbb{R}^{+}\rightarrow \mathbb{R}^{+}:\phi \text{ is continuous, monotone nondecreasing}, \phi (t)=0 \text{ iff } t=0 \bigr\} . $$

Let \((X, d)\) be a b-metric space with \(s\geq 1\) and \(T : X \rightarrow X\) be a self-mapping. We define a mapping \(M_{s} : X \times X \rightarrow \mathbb{R}^{+}\) by

$$ M_{s}(x, y) = \max \biggl\{ d(x, y), d(x, Tx), d(y, Ty), \frac{d(x, Ty) + d(y, Tx)}{2s} \biggr\} . $$

Let also \(p : X \times X \rightarrow \mathbb{R}^{+}\) be a mapping. The following conditions are used in this section:

\((P^{1}_{p} : M_{s})\):

\(x\neq y\) and \(d(x, Tx) \leq d(x, y)\) imply \(p(x, y) \leq M_{s}(x, y)\);

\((P^{2}_{p}: sc)\):

\(x_{n} \neq y\), \(\lim_{n\rightarrow \infty}d(x_{n}, y) = 0\) and \(\lim_{n\rightarrow \infty}d(x_{n}, Tx_{n}) = 0\) imply \(\limsup_{n\rightarrow \infty}(sd(x_{n}, y))\leq cd(y, Ty)\), where \(c \in [0, 1)\).

Definition 9

Let \((X,d)\) be a b-metric space with \(s\geq 1\), \(T:X\rightarrow X\), \(\alpha :X\times X\rightarrow \mathbb{R}^{+}\), \(p : X \times X \rightarrow \mathbb{R}^{+}\) satisfy \((P^{1}_{p} : M_{s})\), and \(\phi \in \varPsi \). Then the mapping T is said to be a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I if it satisfies, for all \(x,y\in X\), the following conditions:

  1. (i)

    For any \(\epsilon > 0\), there exists \(\delta (\epsilon ) > 0\) such that \(x \neq y\) and \(p(x, y) < \epsilon +\delta (\epsilon )\) imply \(d(Tx, Ty)\leq \frac {\epsilon}{s}\);

  2. (ii)

    \(x \neq y\) and \(p(x, y) > 0\) imply \(\zeta (\alpha (x,y)\phi (d(Tx, Ty)), \phi (p(x, y)))\geq 0 \).

Remark 1

If T is a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I, then

$$ \alpha (x,y)\phi \bigl(d(Tx, Ty)\bigr)< \phi \bigl(p(x, y)\bigr)\leq \phi \bigl(M_{s}(x, y)\bigr). $$
(1)

Indeed, we have \(d(x,y)>0\) since \(x \neq y\). If \(p(x,y)=0\), from (ii), we have \(\phi (d(Tx, Ty))<\epsilon \) for any \(\epsilon >0\). But \(\epsilon >0\) is arbitrary, thus we obtain \(Tx=Ty\). In this case, \(\alpha (x,y)\phi (d(Tx, Ty))=0\leq \phi (p(x, y))\). Otherwise, \(p(x, y)>0\), and if \(Tx\neq Ty\), then \(d(Tx, Ty)>0\). If \(\alpha (x,y)=0\), then (1) is satisfied. On the other hand, from \((\zeta _{2})\) and Definition 9(ii), we get

$$ 0\leq \zeta \bigl(\alpha (x,y)\phi \bigl(d(Tx, Ty)\bigr), \phi \bigl(p(x, y) \bigr)\bigr)< \phi \bigl(p(x, y)\bigr)\bigr)- \alpha (x,y)\phi \bigl(d(Tx, Ty)\bigr), $$

so (1) holds.

Now, we give our first main result as follows:

Theorem 4

Let \((X,d)\) be a complete b-metric space with \(s\geq 1\), \(T:X\rightarrow X\), \(\alpha :X\times X\rightarrow \mathbb{R}^{+}\) be mappings, and \(\phi \in \varPsi \). Suppose the following conditions hold:

  1. (i)

    T is generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I;

  2. (ii)

    T is a triangular α-orbital admissible mapping;

  3. (iii)

    There exists \(x_{0}\in X\) such that \(\alpha (x_{0},Tx_{0})\geq 1\);

  4. (iv)

    T is b-continuous.

Then T has a fixed point z. Moreover, \(\{T^{n}x\}\) converges to z for all \(x\in X\).

Proof

By (iii) above, there exists \(x_{0}\in X \) such that \(\alpha (x_{0},Tx_{0})\geq 1\). We construct an iterative sequence \(\{x_{n}\}\) in X by \(x_{n}=Tx_{n-1}\) for \(n\in \mathbb{N}\). Suppose first that \(x_{n_{0}}=x_{n_{0}+1}\) for some \(n_{0}\in \mathbb{N}\). Since \(Tx_{n_{0}} =x_{n_{0}+1} =x_{n_{0}}\), the point \(x_{n_{0}}\) is a fixed point of T and this completes the proof. So from now on, we suppose that \(x_{n}\neq x_{n+1}\) for all \(n\in \mathbb{N}\cup \{0\} \). Since T is triangular α-orbital admissible, \(\alpha (x_{0},Tx_{0} )=\alpha (x_{0},x_{1} )\geq 1\Rightarrow \alpha (Tx_{0},Tx_{1} )=\alpha (x_{1},x_{2} )\geq 1\Rightarrow \alpha (Tx_{1},Tx_{2} )=\alpha (x_{2},x_{3} )\geq 1\). Continuing in this manner, we get

$$ \alpha (x_{n},x_{n+1} )\geq 1\quad \text{for all } n\geq 0. $$
(2)

Again, by using the assumption that T is triangular α-orbital admissible, for all \(n\in \mathbb{N}\cup \{0\} \), (2) yields that \(\alpha (x_{n},x_{n+1})\geq 1\) and \(\alpha (x_{n+1},x_{n+2})\geq 1 \Rightarrow \alpha (x_{n},x_{n+1}) \geq 1\). Recursively, we conclude that \(\alpha (x_{n},x_{n+j})\geq 1\) for all \(n,j\in \mathbb{N}\). In what follows we prove that the sequence \(\{d(x_{n},x_{n+1})\}\) is monotone decreasing. Taking \(x=x_{n}\) and \(y=x_{n+1}\) in \((P^{1}_{p} : M_{s})\), we get

$$ 0< d(x_{n},x_{n+1})=d(x_{n},Tx_{n}) \leq d(x_{n},x_{n+1}), $$

which implies

$$ p(x_{n},x_{n+1})\leq M_{s}(x_{n},x_{n+1}), $$

where

$$ \begin{aligned} M_{s}(x_{n},x_{n+1}) &= \max \biggl\{ d(x_{n}, x_{n+1}), d(x_{n}, Tx_{n}), d(x_{n+1}, Tx_{n+1}), \\ &\quad \frac{d(x_{n}, Tx_{n+1}) + d(x_{n+1}, Tx_{n})}{2s} \biggr\} \\ &= \max \biggl\{ d(x_{n}, x_{n+1}), d(x_{n+1},x_{n+2}), \frac{d(x_{n}, x_{n+2}) + d(x_{n+1}, x_{n+1})}{2s} \biggr\} \\ &= \max \biggl\{ d(x_{n}, x_{n+1}), d(x_{n+1},x_{n+2}), \frac{d(x_{n}, x_{n+2})}{2s} \biggr\} , \end{aligned} $$

and, taking the b-triangle inequality into account, we observe that

$$ \begin{aligned} \frac{d(x_{n}, x_{n+2})}{2s}&\leq \frac{sd(x_{n}, x_{n+1}) + sd(x_{n+1}, x_{n+2})}{2s} \\ &=\frac{d(x_{n}, x_{n+1}) + d(x_{n+1}, x_{n+2})}{2} \\ &\leq \max \bigl\{ d(x_{n}, x_{n+1}), d(x_{n+1}, x_{n+2})\bigr\} , \end{aligned} $$

which gives

$$ M_{s}(x_{n},x_{n+1}) = \max \bigl\{ d(x_{n}, x_{n+1}), d(x_{n+1}, x_{n+2}) \bigr\} . $$

By Definition 9(ii), we get that

$$ \begin{aligned} 0&\leq \zeta \bigl(\alpha (x_{n}, x_{n+1})\phi \bigl(d(Tx_{n}, Tx_{n+1})\bigr), \phi \bigl(P(x_{n}, x_{n+1})\bigr)\bigr) \\ &< \phi \bigl(P(x_{n}, x_{n+1})\bigr)-\alpha (x_{n}, x_{n+1})\phi \bigl(d(Tx_{n}, Tx_{n+1})\bigr), \end{aligned} $$

which is equivalent to

$$ \begin{aligned} \phi \bigl(d(x_{n+1}, x_{n+2}) \bigr)&= \phi \bigl(d(Tx_{n},T x_{n+1})\bigr) \\ &\leq \alpha (x_{n}, x_{n+1})\phi \bigl(d(Tx_{n}, Tx_{n+1})\bigr) \\ &< \phi \bigl(p(x_{n}, x_{n+1})\bigr) \\ &\leq \phi \bigl(M_{s}(x_{n}, x_{n+1}) \bigr). \end{aligned} $$
(3)

If \(M_{s}(x_{n}, x_{n+1})=d(x_{n+1}, x_{n+2})\), then (3) yields a contradiction. Thus, we have

$$\begin{aligned} M_{s}(x_{n}, x_{n+1})=d(x_{n}, x_{n+1}). \end{aligned}$$
(4)

Moreover, from (3), we get

$$ \phi \bigl(d(x_{n+1}, x_{n+2})\bigr)< \phi \bigl(d(x_{n}, x_{n+1})\bigr), $$

which implies, using the monotonicity of ϕ,

$$ d(x_{n+1}, x_{n+2})< d(x_{n}, x_{n+1})\quad \text{for all } n\in \mathbb{N}\cup \{0\}, $$

that is, \(\{d(x_{n},x_{n+1})\}\) is a monotone decreasing sequence of nonnegative real numbers. Thus, there is some \(l\geq 0\) such that \(\lim_{n\rightarrow \infty} d(x_{n}, x_{n+1})=l\). We need to show \(l=0\). Suppose, on the contrary, that \(l>0\) and set \(0<\epsilon =l\). We also note that

$$\begin{aligned} \epsilon =l< d(x_{n}, x_{n+1})\quad \text{for all } n\in \mathbb{N} \cup \{0\}. \end{aligned}$$
(5)

On the other hand, from (3) and (4), we have

$$ p(x_{n}, x_{n+1})\leq d(x_{n}, x_{n+1})< \epsilon +\delta (\epsilon ) $$

for n sufficiently large. So, applying Definition 9(i), we have

$$\begin{aligned} d(Tx_{n}, Tx_{n+1})\leq \frac{\epsilon}{s}. \end{aligned}$$
(6)

Combining (5) together with (6), we obtain

$$ \epsilon < d(x_{n+1}, x_{n+2})=d(Tx_{n}, Tx_{n+1})\leq \frac{\epsilon}{s}, $$

which is a contradiction. We conclude that \(\epsilon =0\), that is,

$$\begin{aligned} \lim_{n\rightarrow \infty}d(x_{n}, x_{n+1})=0. \end{aligned}$$
(7)

Now, we show that \(\{x_{n}\}\) is a b-Cauchy sequence. Let \(\epsilon _{1}>0\) be fixed. From (7), we can choose \(k\in \mathbb{N}\) large enough such that

$$\begin{aligned} d(x_{k}, x_{k+1})< \frac{\delta _{1}}{2s}, \end{aligned}$$
(8)

for some \(\delta _{1}>0\). Without loss of generality, we assume that \(\delta _{1}=\delta _{1}(\epsilon _{1})<\epsilon _{1}\). By induction, we prove that

$$\begin{aligned} d(x_{k},x_{k+m})< \epsilon _{1}+ \frac{\delta _{1}}{2} \quad \text{for all } k,m \in \mathbb{N}\cup \{0\}. \end{aligned}$$
(9)

We already have (9) from (8), for \(m=1\). Suppose that (9) is satisfied for some \(m=j\). Now, we show that (9) holds for \(m=j+1\). On account of (8) and (9), we first observe that

$$ \begin{aligned} &\frac{d(x_{k},x_{k+j+1})+d(x_{k+j},x_{k+1})}{2s} \\ &\quad \leq \frac{sd(x_{k},x_{k+j})+sd(x_{k+j},x_{k+j+1})+sd(x_{k+j},x_{k})+sd(x_{k},x_{k+1})}{2s} \\ &\quad = \frac{d(x_{k},x_{k+j})+d(x_{k+j},x_{k+j+1})+d(x_{k+j},x_{k})+d(x_{k},x_{k+1})}{2} \\ &\quad < \frac{1}{2}\biggl[2\epsilon _{1}+\delta _{1}+\frac{\delta _{1}}{s} \biggr] \\ &\quad \leq \frac{1}{2}[2\epsilon _{1}+2\delta _{1}]= \epsilon _{1}+\delta _{1}. \end{aligned} $$

Thus, we have

$$ \begin{aligned} &M_{s}(x_{k},x_{k+j}) \\ &\quad =\max \biggl\{ d(x_{k},x_{k+j}), d(x_{k},Tx_{k}), d(x_{k+j},Tx_{k+j}), \frac{d(x_{k},Tx_{k+j})+d(x_{k+j},Tx_{k})}{2s} \biggr\} \\ &\quad =\max \biggl\{ d(x_{k},x_{k+j}), d(x_{k},x_{k+1}), d(x_{k+j},x_{k+j+1}), \frac{d(x_{k},x_{k+j+1})+d(x_{k+j},x_{k+1})}{2s} \biggr\} \\ &\quad < \max \biggl\{ \epsilon _{1} +\frac{\delta _{1}}{2}, \frac{\delta _{1}}{2s}, \epsilon _{1} +\delta _{1}\biggr\} \\ &\quad = \epsilon _{1} +\delta _{1}. \end{aligned} $$

From the above inequality, we have

$$ p(x_{k},x_{k+j})\leq M_{s}(x_{k},x_{k+j})=d(x_{k},x_{k+j})< \epsilon _{1}+ \delta _{1}, $$

and, by Definition 9(i), we get

$$\begin{aligned} d(x_{k+1},x_{k+j+1})=d(Tx_{k},Tx_{k+j}) \leq \frac{\epsilon _{1}}{s}. \end{aligned}$$
(10)

Now, using the b-triangle inequality, as well as (8) and (10), we have

$$ \begin{aligned} d(x_{k},x_{k+j+1})&\leq sd(x_{k},x_{k+1})+sd(x_{k+1},x_{k+j+1}) \\ &=sd(x_{k},x_{k+1})+sd(Tx_{k},Tx_{k+j}) \\ &< \frac{\delta _{1}}{2}+\epsilon _{1}. \end{aligned} $$

So, (9) holds for \(m=j+1\). Therefore,

$$ d(x_{k},x_{k+m})< \epsilon _{1}\quad \text{for all } k,m \in \mathbb{N} \cup \{0\}. $$

In other words, for \(m>n\), we have \(\lim_{n,m \rightarrow \infty} d(x_{n},x_{m})=0\) and hence the sequence \(\{x_{n}\}\) is a b-Cauchy sequence. Since, \((X,d)\) is a complete b-metric space, there exists \(u\in X\) such that \(x_{n}\rightarrow u\) as \(n\rightarrow \infty \). By b-continuity of T, we have

$$ u=\lim_{n\rightarrow \infty}x_{n+1}=\lim_{n \rightarrow \infty}Tx_{n}=Tu, $$

that is, u is a fixed point of T. □

Now, replacing continuity of T by continuity of \(T^{2}\) in Theorem 4, we prove the following fixed point result.

Theorem 5

Let \((X,d)\) be a complete b-metric space with \(s\geq 1\) and let \(T:X\rightarrow X\) be a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I satisfying the following conditions:

  1. (i)

    T is a triangular α-orbital admissible mapping;

  2. (ii)

    There exists \(x_{0}\in X\) such that \(\alpha (x_{0},Tx_{0})\geq 1\);

  3. (iii)

    \(T^{2}\) is continuous.

Then \(\{T^{n}x\}\) is converges to z for all \(x\in X\). Moreover, for \(\alpha (z,Tz)\geq 1\), z is a fixed point of T, and T is discontinuous at z if and only if \(\lim_{x\rightarrow z}M_{s}(x,z)\neq 0\).

Proof

Following the proof of Theorem 4, we see that the sequence \(\{x_{n}\}\) in X defined by \(x_{n}=Tx_{n-1}\) for all \(n\in \mathbb{N}\) is convergent to \(z\in X\) and \(\alpha (x_{n}, x_{n+1})\geq 1\) for all \(n\in \mathbb{N}\cup \{0\}\). Regarding the fact that any subsequence of \(\{x_{n}\}\) converges to z, we get \(\lim_{n\rightarrow \infty}x_{n+1 }=\lim_{n \rightarrow \infty}Tx_{n}=z\) and \(\lim_{n\rightarrow \infty}x_{n+2 }=\lim_{n \rightarrow \infty}T^{2}x_{n}=z\). On the other hand, due to the continuity of \(T^{2}\),

$$ T^{2}z=\lim_{n\rightarrow \infty}T^{2}x_{n}=z. $$

We claim that \(Tz=z\). Suppose, on the contrary, that \(Tz\neq z\) and \(p(z,Tz)>0\). Then we have

$$ \begin{aligned} p(z,Tz)&\leq M_{s}(z,Tz) \\ &= \max \biggl\{ d(z,Tz),d(z,Tz),d\bigl(Tz,T^{2}z\bigr), \frac{d(z,T^{2}z)+d(Tz,Tz)}{2s} \biggr\} \\ &=d(z,Tz). \end{aligned} $$

Thus, using (1) together with the hypothesis \(\alpha (z,Tz)\geq 1\), we obtain

$$\begin{aligned} 0\leq \zeta \bigl(\alpha (z,Tz)\phi \bigl(d\bigl(Tz,T^{2}z\bigr) \bigr),\phi \bigl(P(z,Tz)\bigr)\bigr), \end{aligned}$$

and also

$$ \begin{aligned} 0&< \phi \bigl(d(Tz,z)\bigr)=\phi \bigl(d \bigl(Tz,T^{2}z\bigr)\bigr) \\ &\leq \alpha (z,Tz)\phi \bigl(d\bigl(Tz,T^{2}z\bigr)\bigr) \\ &< \phi \bigl(P(z,Tz)\bigr) \\ &\leq \phi \bigl(M_{s}(z,Tz)\bigr) \\ &=\phi \bigl(d(z,Tz)\bigr), \end{aligned} $$

which is a contradiction. So, \(z=Tz\), that is, z is a fixed point of T. □

Definition 10

A b-metric space \((X,d)\) is called regular if for any sequence \(\{x_{n}\}\) in X with \(\lim_{n\rightarrow \infty}d(x_{n},z)=0\) and \(\alpha (x_{n}, x_{n+1})\geq 1\) for all \(n\in \mathbb{N}\cup \{0\}\), one has \(\alpha (x_{n}, z)\geq 1\) for all \(n\in \mathbb{N}\cup \{0\}\).

In the following, we prove the following fixed point theorem, without continuity assumption of T and \(T^{2}\).

Theorem 6

Let \((X,d)\) be a complete b-metric space with \(s\geq 1\) and \(T:X\rightarrow X\) be a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type I. Suppose that \((P_{p}^{2}: sc)\) and the following conditions hold:

  1. (i)

    T is a triangular α-orbital admissible mapping;

  2. (ii)

    There exists \(x_{0}\in X\) such that \(\alpha (x_{0},Tx_{0})\geq 1\);

  3. (iii)

    \((X,d)\) is regular.

Then \(\{T^{n}x\}\) is converges to z for all \(x\in X\). Moreover, z is a fixed point of T.

Proof

Following the proof of Theorem 4, we see that the sequence \(\{x_{n}\}\) in X defined by \(x_{n}=Tx_{n-1}\) for all \(n\in \mathbb{N}\) is convergent to \(z\in X\) and \(\alpha (x_{n}, x_{n+1})\geq 1\) for all \(n\in \mathbb{N}\cup \{0\}\). We notice also that all adjacent terms in \(\{x_{n}\}\) are distinct. Moreover, we note \(T^{n}x\neq z\) for all \(n\in \mathbb{N} \cup \{0\}\). Regarding the limits \(\lim_{n\rightarrow \infty}d(x_{n},z)=0\) and \(\lim_{n\rightarrow \infty}d(x_{n},x_{n+1})=0\), we drive from \((P^{2}_{p}:sc)\) that

$$\begin{aligned} s\limsup_{n\rightarrow \infty}p(x_{n},z)\leq cd(z,Tz) \quad \text{for any } c\in [0,1). \end{aligned}$$
(11)

So, by assumption (iii), we get \(\alpha (x_{n},z)\geq 1\). Now, we prove that z is a fixed point of T. Suppose, on the contrary, that \(Tz\neq z\). Taking \(x=x_{n}\) and \(y=z\) in Definition 9(ii), we obtain that

$$ \begin{aligned} 0&\leq \zeta \bigl(\alpha (x_{n},z)\phi \bigl(d(Tx_{n},Tz)\bigr),\phi \bigl(p(x_{n},z)\bigr)\bigr) \\ &< \phi \bigl(p(x_{n},z)\bigr)-\alpha (x_{n},z)\phi \bigl(d(Tx_{n},Tz)\bigr), \end{aligned} $$

which is equivalent to

$$ \begin{aligned} \phi \bigl(d(x_{n+1},Tz)\bigr)&=\phi \bigl(d(Tx_{n},Tz)\bigr) \\ &\leq \alpha (x_{n},z)\phi \bigl(d(Tx_{n},Tz)\bigr) \\ &< \phi \bigl(p(x_{n},z)\bigr). \end{aligned} $$
(12)

Since ϕ is monotone, (12) yields

$$\begin{aligned} d(x_{n+1},Tz)< p(x_{n},z). \end{aligned}$$
(13)

Applying the b-triangle inequality and using (13), we have

$$ \begin{aligned} d(z,Tz)&\leq sd(z,x_{n+1})+sd(x_{n+1},Tz) \\ &< sd(z,x_{n+1})+sp(x_{n},z). \end{aligned} $$
(14)

Taking the limit as \(n\rightarrow \infty \) in (14) and using \((P^{2}_{p} :sc)\), we obtain that

$$ d(z,Tz)< s\limsup_{n\rightarrow \infty}p(x_{n},z)\leq cd(z,Tz) \quad \text{for any } c\in [0,1), $$

which is a contradiction. Therefore, z is a fixed point of T. □

For the uniqueness of fixed point, we need the following additional condition.

Condition (U)

For all \(x,y \in \mathrm{Fix}(T)\), we have \(\alpha (x,y)\geq 1\), where \(\mathrm{Fix}(T)\) denotes the set of all fixed points of T.

Theorem 7

Adding Condition (U)to the hypotheses of Theorem 4 (resp. Theorems 5and 6), we prove the uniqueness of fixed point of T.

Proof

We argue by contradiction, that is, suppose there exist \(z,w \in X\) such that \(z=Tz\) and \(w=Tw\) with \(z\neq w\). By Condition (U), we have \(\alpha (z,w)\geq 1\). We notice first that the case \(p(z,w)=0\) is impossible since we have \(Tz=Tw\) and

$$ 0< d(z,w)=d(Tz,Tw)=0, $$

which is a contradiction. Thus, we get that \(p(z,w)>0\). Since

$$ 0=d(z,Tz)\leq d(z,w), $$

by \((P^{1}_{p}:M_{s})\), we have

$$ p(z,w)\leq M_{s}(z,w), $$

where

$$ M_{s}(z,w)=\max \biggl\{ d(z,w), d(z,Tz),d(w,Tw), \frac{ d(z,Tw)+d(Tz,w)}{2s} \biggr\} =d(z,w). $$

Using Definition 9(ii), we get

$$ \begin{aligned}0&\leq \zeta \bigl(\alpha (z,w)\phi \bigl(d(Tz,Tw) \bigr),\phi \bigl(p(z,w)\bigr)\bigr) \\ &< \phi \bigl(p(z,w)\bigr)-\alpha (z,w)\phi \bigl(d(Tz,Tw)\bigr), \end{aligned} $$

which imply

$$ \begin{aligned} 0&< \phi \bigl(d(z,w)\bigr)=\phi \bigl(d(Tz,Tw)\bigr) \\ &\leq \alpha (z,w)\phi \bigl(d(Tz,Tw)\bigr) \\ &< \phi \bigl(p(z,w)\bigr) \\ &\leq \phi \bigl(d(z,w)\bigr), \end{aligned} $$

which is a contradiction. Hence, \(d(z,w)=0\), that is, the fixed point of T is unique. □

In the following, we introduce generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type II and study fixed point results for such mappings.

Definition 11

Let \((X,d)\) be a b-metric space with \(s\geq 1\), \(T:X\rightarrow X\), \(\alpha :X\times X\rightarrow \mathbb{R}^{+}\), \(\zeta \in Z_{w}\), \(\phi \in \varPsi \), and suppose \(p :X\times X\rightarrow \mathbb{R}^{+}\) is a function that satisfies \((P^{1}_{p}:N_{s})\) and \((P^{2}_{p}: sc)\). The mapping T is said to be a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type II if it satisfies for all \(x,y\in X\) the following conditions:

  1. (a)

    For any \(\epsilon > 0\), there exists \(\delta (\epsilon ) > 0\) such that \(x \neq y\) and \(p(x, y) < \epsilon +\delta (\epsilon )\) imply \(d(Tx, Ty)\leq \frac{\epsilon}{s}\);

  2. (b)

    \(x \neq y\) and \(p(x, y) > 0 \) imply

    $$ \zeta \bigl(\alpha (x,y)\phi \bigl(d(Tx, Ty)\bigr), \phi \bigl(p(x, y)\bigr) \bigr)\geq 0. $$
    (15)

We define a mapping \(N_{s}:X\times X \rightarrow \mathbb{R^{+}}\) by

$$ \begin{aligned}N_{s}(x,y)&=\max \biggl\{ d(x,y), d(x,Tx), d(y,Ty), \\ &\quad \frac{d(y,Ty)[1+d(x,Tx)]}{1+d(x,y)}, \frac{d(x,Tx)[1+d(y,Ty)]}{1+d(Tx,Ty)} \biggr\} . \end{aligned} $$

We note that, for any \(x,y\in X\) with \(x=y\), we have \(0=d(Tx,Ty)\leq N_{s}(x,y)\). Moreover, if \(x\neq y\), then \(N_{s}(x,y)>0\).

Now, we state and prove the following fixed point theorem.

Theorem 8

Let \((X,d)\) be a complete b-metric space with \(s\geq 1\) and \(T:X\rightarrow X\) be a generalized \((\alpha , \phi )\)-Meir–Keeler hybrid contractive mapping of type II satisfying the following conditions:

  1. (i)

    T is a triangular α-orbital admissible mapping;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha (x_{0},Tx_{0})\geq 1\);

  3. (iii)

    either T is continuous;

  4. (iv)

    or \(T^{2}\) is continuous and \(\alpha (z,Tz)\geq 1\);

  5. (v)

    or \((X,d)\) is regular.

Then T has a fixed point z. Moreover, \(\{T^{n}x\}\) is convergent to z for all \(x\in X\).

Proof

As in the proof of Theorem 4, we construct a recursive sequence \(\{x_{n}\}\) as follows:

$$ x_{n}=Tx_{n-1}\quad \text{for all } n\in \mathbb{N} . $$

One can conclude that \(\alpha (x_{n},x_{n+1})\geq 1\) for all \(n\in \mathbb{N} \cup \{0\}\), due to conditions (i) and (ii). Throughout the proof, we assume \(x_{n}\neq x_{n+1}\) for all \(n\in \mathbb{N} \cup \{0\}\). Indeed, as it was discussed in the proof of Theorem 4, the other case is trivial and is excluded. Now, by letting \(x=x_{n}\) and \(y=x_{n+1}\) in \((P^{1}_{p} : N_{s})\), we have

$$ d(x_{n},Tx_{n})\leq d(x_{n},x_{n+1}), $$

which implies

$$ p(x_{n},x_{n+1})\leq N_{s}(x_{n},x_{n+1}), $$

where

$$ \begin{aligned} N_{s}(x_{n},x_{n+1})&= \max \biggl\{ d(x_{n},x_{n+1}), d(x_{n},Tx_{n}), d(x_{n+1},Tx_{n+1}), \\ &\quad \frac{d(x_{n+1},Tx_{n+1})[1+d(x_{n},Tx_{n})]}{1+d(x_{n},x_{n+1})}, \frac{d(x_{n},Tx_{n})[1+d(x_{n+1},Tx_{n+1})]}{1+d(Tx_{n},Tx_{n+1})} \biggr\} \\ &=\max \biggl\{ d(x_{n},x_{n+1}), d(x_{n},x_{n+1}), d(x_{n+1},x_{n+2}), \\ &\quad \frac{d(x_{n+1},x_{n+2})[1+d(x_{n},x_{n+1})]}{1+d(x_{n},x_{n+1})}, \frac{d(x_{n},x_{n+1})[1+d(x_{n+1},x_{n+2})]}{1+d(x_{n+1},x_{n+2})} \biggr\} \\ &=\max \bigl\{ d(x_{n},x_{n+1}), d(x_{n+1},x_{n+2}) \bigr\} . \end{aligned} $$

By Definition 11(b), we have

$$\begin{aligned} 0\leq \zeta \bigl(\alpha (x_{n},x_{n+1})\phi \bigl(d(Tx_{n},Tx_{n+1})\bigr),\phi \bigl(p(x_{n},x_{n+1}) \bigr)\bigr). \end{aligned}$$

Consequently, the above inequality yields

$$ \begin{aligned} \phi (d(x_{n+1},x_{n+2})&= \phi \bigl(d(Tx_{n},Tx_{n+1})\bigr) \\ &\leq \alpha (x_{n},x_{n+1})\phi \bigl(d(Tx_{n},Tx_{n+1}) \bigr) \\ &< \phi \bigl(p(x_{n},x_{n+1})\bigr) \\ &\leq \phi \bigl(N_{s}(x_{n},x_{n+1})\bigr), \end{aligned} $$
(16)

where

$$\begin{aligned} N_{s}(x_{n},x_{n+1})=\max \bigl\{ d(x_{n},x_{n+1}), d(x_{n+1},x_{n+2})\bigr\} = d(x_{n},x_{n+1}). \end{aligned}$$
(17)

Thus, from (16), (17) and the monotonicity of ϕ, for all \(n\in \mathbb{N} \cup \{0\}\), we have

$$ d(x_{n+1},x_{n+2})< d(x_{n},x_{n+1}), $$

that is, \(\{d(x_{n},x_{n+1})\}\) is nonincreasing sequence of nonnegative real numbers. Consequently, there exists a real number \(r\geq 0\) such that \(d(x_{n},x_{n+1})\rightarrow r\) as \(n\rightarrow \infty \). Suppose that \(r=\epsilon >0\). First, we note that \(r=\epsilon < d(x_{n},x_{n+1})\) for all \(n\in \mathbb{N} \cup \{0\}\). On the other hand, from (16), there exists \(\delta >0\) such that

$$ \begin{aligned} p(x_{n},x_{n+1})&\leq N_{s}(x_{n},x_{n+1}) \\ &= d(x_{n},x_{n+1}) \\ &< \epsilon +\delta (\epsilon ), \end{aligned} $$

for n sufficiently large. Keeping the observations above, Definition 11(a) yields that

$$ d(Tx_{n},Tx_{n+1})\leq \frac{\epsilon}{s}. $$

Thus, we have

$$ \epsilon < d(x_{n+1},x_{n+2})=d(Tx_{n},Tx_{n+1}) \leq \frac{\epsilon}{s}, $$

which is a contradiction. So, we derive that \(\epsilon =0\), that is, \(\lim_{n\rightarrow \infty}d(x_{n},x_{n+1})=0\). In what follows, we show that the sequence \(\{x_{n}\}\) is b-Cauchy. For this aim, let \(m\in \mathbb{N}\) be large enough to satisfy

$$ d(x_{m},x_{m+1})< \frac{\delta _{1}}{s}. $$

Now, we show by induction that

$$\begin{aligned} d(x_{m},x_{m+k})< \epsilon _{1} + \delta _{1} \quad \text{for all } k \in \mathbb{N}. \end{aligned}$$
(18)

Without loss of generality, we assume that \(\delta _{1}=\delta _{1}(\epsilon )<\epsilon \). We have already proved the claim for \(k=1\). Now, we consider the following two cases:

Case (i). If \(d(x_{m+k},x_{m+k+1})\leq d(x_{m},x_{m+k})\), then we get

$$ \frac{d(x_{m+k},x_{m+k+1})}{1+d(x_{m},x_{m+k})}\leq d(x_{m+k},x_{m+k+1}) $$

and

$$ \frac{d(x_{m+k},x_{m+k+1})d(x_{m},x_{m+1})}{1+d(x_{m},x_{m+k})}< d(x_{m},x_{m+1}). $$

Hence, we have

$$ \begin{aligned} p(x_{m},x_{m+k})&\leq N_{s}(x_{m},x_{m+k}) \\ &=\max \biggl\{ d(x_{m},x_{m+k}), d(x_{m},Tx_{m}), d(x_{m+k},Tx_{m+k}), \\ &\quad \frac{d(x_{m+k},Tx_{m+k})[1+d(x_{m},Tx_{m})]}{1+d(x_{m},x_{m+k})}, \frac{d(x_{m},Tx_{m})[1+d(x_{m+k},Tx_{m+k})]}{1+d(Tx_{m},Tx_{m+k})} \biggr\} \\ &=\max \biggl\{ d(x_{m},x_{m+k}), d(x_{m},x_{m+1}), d(x_{m+k},x_{m+k+1}), \\ &\quad \frac{d(x_{m+k},x_{m+k+1})[1+d(x_{m},x_{m+1})]}{1+d(x_{m},x_{m+k})}, \frac{d(x_{m},x_{m+1})[1+d(x_{m+k},x_{m+k+1})]}{1+d(x_{m+1},x_{m+k+1})} \biggr\} \\ &< \max \{\epsilon _{1} +\delta _{1}, 2\delta _{1}, \delta _{1}\}= \epsilon _{1} +\delta _{1}, \end{aligned} $$

and so it follows from Definition 11(a) that

$$ d(Tx_{m},Tx_{m+k})\leq \frac{\epsilon _{1}}{s}. $$

Thus, by the b-triangle inequality, we have

$$ \begin{aligned} d(x_{m},x_{m+k+1})&\leq sd(x_{m},x_{m+1})+sd(x_{m+1},x_{m+k+1}) \\ &=sd(x_{m},x_{m+1})+sd(Tx_{m},Tx_{m+k}) \\ &< \epsilon _{1} +\delta _{1}. \end{aligned} $$

Case (ii). If \(d(x_{m+k},x_{m+k+1})> d(x_{m},x_{m+k})\), then we get

$$ \begin{aligned} d(x_{m},x_{m+k+1})&\leq sd(x_{m},x_{m+k})+sd(x_{m+k},x_{m+k+1}) \\ &< 2sd(x_{m+k},x_{m+k+1}) \\ &< 2s\frac{\delta _{1}}{s} \\ &= 2\delta _{1}< \epsilon _{1} +\delta _{1}. \end{aligned} $$

Thus, by induction, (18) holds for every \(k\in \mathbb{N}\). Since \(\epsilon _{1}>0\) is arbitrary, we get

$$ \limsup_{k\rightarrow \infty}d(x_{m},x_{m+k})=0, $$

which implies that \(\{x_{n}\}\) is a b-Cauchy sequence in a complete b-metric space \((X,d)\). Hence, \(\{x_{n}\}\) b-converges to some \(z\in X\).

Next, we show that z is a fixed point of T. If T is continuous, then we have

$$ z=\lim_{n\rightarrow \infty}x_{n+1}=\lim_{n \rightarrow \infty}Tx_{n}=Tz, $$

that is, z is a fixed point of T.

If \(T^{2}\) is continuous, since \(x_{n} \rightarrow z\), we get that any subsequence of \(\{x_{n}\}\) converges to the same limit point z, so

$$ \lim_{n\rightarrow \infty}x_{n+1}=\lim_{n \rightarrow \infty}Tx_{n}=z \quad \text{and}\quad \lim_{n \rightarrow \infty}x_{n+2}=\lim _{n\rightarrow \infty}T^{2}x_{n}=z. $$

On the other hand, due to the continuity of \(T^{2}\),

$$ T^{2}z=\lim_{n\rightarrow \infty}T^{2}x_{n}=z. $$

We claim that \(Tz=z\). To the contrary, if \(Tz\neq z\), then we have \(p(z,Tz)>0\) and

$$ \begin{aligned} p(z,Tz)&\leq N_{s}(z,Tz) \\ &=\max \biggl\{ d(z,Tz), d(z,Tz), d\bigl(Tz,T^{2}z\bigr), \\ &\quad \frac{d(Tz,T^{z})[1+d(z,Tz)]}{1+d(z,Tz)}, \frac{d(z,Tz)[1+d(Tz,T^{2}z)]}{1+d(Tz,T^{2}z)} \biggr\} \\ &=\max \biggl\{ d(z,Tz), d(z,Tz), d\bigl(Tz,T^{2}z\bigr), \\ &\quad \frac{d(Tz,T^{2}z)[1+d(z,Tz)]}{1+d(z,Tz)}, \frac{d(z,Tz)[1+d(Tz,T^{2}z)]}{1+d(Tz,T^{2}z)} \biggr\} \\ &= d(z,Tz). \end{aligned} $$

Therefore, together with the supplementary hypothesis \(\alpha (z,Tz)\geq 1\), we have

$$ \begin{aligned} 0&\leq \zeta \bigl(\alpha (z,Tz)\phi \bigl(d \bigl(Tz,T^{2}z\bigr)\bigr),\phi \bigl(p(z,Tz)\bigr)\bigr) \\ &< \phi \bigl(p(z,Tz)\bigr)-\alpha (z,Tz)\phi \bigl(d\bigl(Tz,T^{2}z \bigr)\bigr). \end{aligned} $$

From the above inequality, we obtain

$$ \begin{aligned} 0&< \phi \bigl(d(Tz,z)\bigr) \\ &=\phi \bigl(d\bigl(Tz,T^{2}z\bigr)\bigr) \\ &\leq \alpha (z,Tz)\phi \bigl(d\bigl(Tz,T^{2}z\bigr)\bigr) \\ &< \phi \bigl(p(z,Tz)\bigr) \\ &\leq \phi \bigl(N_{s}(z,Tz)\bigr) \\ &=\phi \bigl(d(z,Tz)\bigr), \end{aligned} $$

which is a contradiction. Hence, z is a fixed point of T.

If X is regular, we deduce that \(d(z,Tz)=0\), using the same arguments as in the proof of Theorem 6. That is, z is a fixed point of T. □

The uniqueness of fixed point of T can be deduced as in Theorem 7.

Now, we give an example to illustrate Theorem 8.

Example 3

Let \(X = [0, 4]\) and \(d : X \times X \rightarrow \mathbb{R}^{+}\) be defined by \(d(x, y) = |x- y|^{2}\) for all \(x,y \in X\). Then \((X,d)\) is a complete b-metric space with \(s=2\) which is not a metric space. Let \(T : X \rightarrow X\) be defined by

$$ T(x)= \textstyle\begin{cases} 1& \text{if $x\in [0, 2)$}, \\ \frac{x}{2} & \text{if $x\in [2,4]$.} \end{cases} $$

Also, we define \(\alpha : X \times X \rightarrow \mathbb{R}^{+}\), \(q:X\times X \rightarrow \mathbb{R}^{+}\) and \(\phi :\mathbb{R}^{+}\rightarrow \mathbb{R}^{+}\) as follows:

$$ \alpha (x,y)= \textstyle\begin{cases} 2&\text{if $x,y\in [0,2)$,} \\ 1&\text{if $x,y\in [2,4]$,} \\ 0 & \text{otherwise}, \end{cases} $$

\(q(x,y)=\max \{d(x,y),\frac {d(x,Tx)d(y,Ty)}{1+d(x,y)}, \frac {d(x,Tx)d(y,Ty)}{1+d(Tx,Ty)} \}\) and \(\phi (t)=\frac {t^{2}}{2}\). First, we note that q satisfies condition \((P^{1}_{q}: N_{s})\) and \(q(x, y) > 0\) for all \(x \neq y\). Since, for \(x = 0\) we have \(T0 = 1\) and \(\alpha (0, T0) = \alpha (0, 1) = 2 > 1\), assumption (ii) of Theorem 8 is satisfied. Also, it is easy to see that T is triangular α-orbital admissible. Let \(\zeta \in Z_{w}\) be is given by \(\zeta (t,s) = \frac{2}{3}s- t\). Now, we consider the following cases:

  1. Case 1.

    For \(x,y \in [0, 2)\), \(x \neq y\), we have \(d(Tx, Ty) = 0\), so

    $$ \zeta \bigl( \alpha (x, y)\phi \bigl(d(Tx, Ty)\bigr), \phi \bigl(q(x, y)\bigr) \bigr) = \frac{2\phi (q(x, y))}{3}= \frac{ (q(x, y))^{2}}{3}>0. $$
  2. Case 2.

    For \(x, y \in [2, 4]\), \(x \neq y\), we have

    $$ d(Tx, Ty) = \frac{|x-y|}{2}, \qquad q(x, y) = \max \biggl\{ |x-y|, \frac{\frac{x}{2}.\frac{y}{2}}{1+|x-y|}, \frac{\frac{x}{2}.\frac{y}{2}}{1+|\frac{x-y}{2}|} \biggr\} , $$

    so

    $$\begin{aligned} \zeta \bigl(\alpha (x, y)\phi \bigl(d(Tx, Ty)\bigr), \phi \bigl(q(x, y)\bigr) \bigr) =& \frac{2\phi (q(x, y))}{3}-\phi \biggl(\frac{|x-y|}{2} \biggr) \\ =&\frac{(q(x, y))^{2}}{3}-\frac{(|x-y|)^{2}}{8}\geq 0. \end{aligned}$$
  3. Case 3.

    For \(x \in [0, 2)\) and \(y \in [2, 4]\), we have \(\alpha (x, y) = 0\) and

    $$ \zeta \bigl(\alpha (x, y)\phi \bigl(d(Tx, Ty)\bigr), \phi \bigl(q(x, y)\bigr) \bigr) = \frac{2\phi (q(x, y))}{3}=\frac{(q(x, y))^{2}}{3} > 0. $$

Thus, due to the cases considered above, T satisfies all the conditions of Theorem 8 and has a unique fixed point \(x = 1\).

Now, we give some corollaries to our main findings.

Corollary 1

Let \((X,d)\) be a complete b-metric space with \(s\geq 1\) and let \(T:X\rightarrow X\) be a \((\alpha ,\phi )\)-Meir–Keeler hybrid contractive mapping of type I with \(p(x,y)=d(x,y)\). Assume that the following conditions are satisfied:

  1. (i)

    T is triangular α-orbital admissible;

  2. (ii)

    there exists \(x_{0} \in X\) such that \(\alpha (x_{0}, Tx_{0}) \geq 1\);

  3. (iii)

    either T is continuous or \(T^{2}\) is continuous and \(\alpha (u, Tu) \geq 1\) or \((X, d)\) is regular.

Then T has a fixed point u. Moreover, \(\{T^{n}x\}\) converges to u for all \(x \in X\).

Remark 2

Under the conditions of Corollary 1, since \(x \neq y\) implies \(d(x, y) > 0\), it is obvious that (b) from Definition 9 is equivalent to the following:

(b′):

\(d(x,y)>0\) implies \(\zeta (\alpha (x,y)\phi (d(Tx,Ty)),\phi (d(x,y)))\geq 0\).

Proof

It is clear that d satisfies the conditions \((P^{1}_{d}: M_{s}) \), respectively \((P^{2}_{d} : 0)\), and so all the assumptions of Theorems 4, 5, and 6 are also satisfied. Thus, T has a fixed point. □

Corollary 2

Let \((X, d)\) be a complete b-metric space with \(s\geq 1\), and let \(T : X \rightarrow X\) be a \((\alpha ,\phi )\)-Meir–Keeler hybrid contractive mapping of type I. Let \(\rho : X \times X \rightarrow \mathbb{R}^{+}\) be defined by

$$ \rho (x, y) = a_{1}d(x, y) + a_{2}d(x, Tx) + a_{3}d(y, Ty), $$

where \(a_{1}, a_{2}, a_{3} \in [0, \frac{1}{s})\), \(a_{1} + a_{2} \leq \frac{1}{2s}\) and \(a_{3} \leq \frac{1}{2s}\). Assume also that:

  1. (i)

    T is triangular α-orbital admissible;

  2. (ii)

    there exists \(x_{0} \in X\) such that \(\alpha (x_{0}, Tx_{0}) \geq 1\);

  3. (iii)

    either T is continuous or \(T^{2}\) is continuous and \(\alpha (u, Tu) \geq 1\) or \((X, d)\) is regular.

Then T has a fixed point u. Moreover, \(\{T^{n}x\}\) converges to u for all \(x \in X\).

Proof

Let \(x, y \in X\) be such that \(x\neq y\) and \(d(x, Tx) \leq d(x, y)\). Then,

$$ \begin{aligned} \rho (x, y)&= a_{1}d(x, y) + a_{2}d(x, Tx) + a_{3}d(y, Ty) \\ &\leq (a_{1} + a_{2})d(x, y) + a_{3}d(y, Ty) \\ &\leq \frac{d(x, y) + d(y, Ty)}{2s} \\ & \leq M_{s}(x, y), \end{aligned} $$

which shows that \((P^{1}_{\rho}: M_{s})\) holds. On the other hand, if \(x_{n} \neq y\), then

$$ \lim_{n\rightarrow \infty} d(x_{n}, y) = 0\quad \text{and} \quad \lim_{n\rightarrow \infty}d(x_{n}, x_{n+1}) = 0 $$

hold, and then we have

$$ \limsup_{n\rightarrow \infty} \rho (x_{n}, y) = \limsup _{n\rightarrow \infty} \bigl[a_{1}d(x_{n}, y) + a_{2}d(x_{n}, x_{n+1}) + a_{3}d(y, Ty)\bigr] = a_{3}d(y, Ty). $$

Thus, \((P^{2}_{\rho} : a_{3})\) holds. Hence, T has a fixed point. □

4 Conclusion

In 2020, Karapinar et al. [43] introduced a generalized Meir–Keeler contraction via a simulation function and studied fixed point results for the mappings introduced in the setting of metric spaces. In this work, we introduced generalized \((\alpha ,\phi )\)-Meir–Keeler hybrid contractive mappings of type I and II in the setting of b-metric spaces and proved the existence and uniqueness of fixed points for such mappings. Our results extend and generalize the results of Karapinar et al. [43] and many other related fixed point results in the existing literature. Finally, we have also supported the main result of this work by an illustrative example.

Data availability

Not applicable.

References

  1. Banach, S.: Sur les opérations dans les ensembles abstraits et leur application aux équations intégrales. Fundam. Math. 3(1), 133–181 (1922)

    Article  Google Scholar 

  2. Taş, N.: Interpolative contractions and discontinuity at fixed point. Appl. Gen. Topol. 24(1), 145–156 (2023)

    Article  MathSciNet  Google Scholar 

  3. Edraoui, M., Semami, S.: Fixed points results for various types of interpolative cyclic contraction. Appl. Gen. Topol. 24(2), 247–252 (2023)

    Article  MathSciNet  Google Scholar 

  4. Meir, A., Keeler, E.: A theorem on contraction mappings. J. Math. Anal. Appl. 28, 326–329 (1969)

    Article  MathSciNet  Google Scholar 

  5. Kirk, W.A.: Contraction mappings and extensions. In: Handbook of Metric Fixed Point Theory, pp. 1–34 (2001)

    Chapter  Google Scholar 

  6. Bisht, R.K.: An overview of the emergence of weaker continuity notions, various classes of contractive mappings and related fixed point theorems. J. Fixed Point Theory Appl. 25(1), 11 (2023)

    Article  MathSciNet  Google Scholar 

  7. Zhou, Z., Tan, B., Li, S.: Inertial algorithms with adaptive step sizes for split variational inclusion problems and their applications to signal recovery problem. Math. Methods Appl. Sci. (2023). https://doi.org/10.1002/mma.9436

    Article  Google Scholar 

  8. Lu, N., He, F., Van Dung, N.: On a question concerning Meir–Keeler contractions in complete b-metric spaces. J. Math. Anal. Appl. 527, 127470 (2023)

    Article  MathSciNet  Google Scholar 

  9. Aydi, H., Banković, R., Mitrović, I., Nazam, M.: Nemytzki–Edelstein–Meir–Keeler type results in b-metric spaces. Discrete Dyn. Nat. Soc. 2018, 4745764 (2018)

    Article  MathSciNet  Google Scholar 

  10. Gholamian, N., Khanehgir, M.: Fixed points of generalized Meir–Keeler contraction mappings in b-metric-like spaces. Fixed Point Theory Appl. 2016, 34 (2016)

    Article  MathSciNet  Google Scholar 

  11. Gülyaz, S., Karapinar, E., Erhan, I.M.: Generalized α-Meir–Keeler contraction mappings on Branciari b-metric spaces. Filomat 17(31), 445–5456 (2017)

    MathSciNet  Google Scholar 

  12. Karapinar, E., Czerwik, S., Aydi, H.: \((\alpha , \varphi )\)-Meir–Keeler contraction mappings in generalized b-metric spaces. J. Funct. Spaces 2018(4), 326–4620 (2018)

    MathSciNet  Google Scholar 

  13. Mitrovi, Z.D., Radenovi, S.: On Meir–Keeler contraction in Branciari b-metric spaces. Trans. A. Razmadze Math. Inst. 173(1), 83–90 (2019)

    MathSciNet  Google Scholar 

  14. Mlaiki, N., Souayah, N., Abodayeh, K., Abdeljawad, T.: Meir–Keeler contraction mappings in \(M_{b}\)-metric spaces. J. Comput. Anal. Appl. 4(27) (2019). https://doi.org/10.1155/2014/756298

  15. Bakhtin, I.: The contraction mapping principle in quasi-metric spaces. Func. An., Gos. Ped. Inst. Unianowsk 30, 26–37 (1989)

    Google Scholar 

  16. Czerwik, S.: Contraction mappings in b-metric spaces. Acta Math. Inform. Univ. Ostrav. 1(1), 5–11 (1993)

    MathSciNet  Google Scholar 

  17. Aghajani, A., Abbas, M., Roshan, J.R.: Common fixed point of generalized weak contractive mappings in partially ordered b-metric spaces. Math. Slovaca 64, 941–960 (2014)

    Article  MathSciNet  Google Scholar 

  18. Deshmukh, A., Gopal, D.: Topology of non-triangular metric spaces and related fixed point results. Filomat 35(11), 3557–3570 (2021)

    Article  MathSciNet  Google Scholar 

  19. Ding, H.S., Imdad, M., Radenović, S., Vujaković, J.: On some fixed point results in b-metric. Math. Sci. 22(2), 151–164 (2016)

    MathSciNet  Google Scholar 

  20. Dosenović, T., Pavlović, M., Radenović, S.: Contractive conditions in b-metric spaces. Vojnotehnicki Glasnik/Mil. Tech. Cour. 65, 851–865 (2017)

    Article  Google Scholar 

  21. Faraji, H., Nourouzi, K.: A generalization of Kannan and Chatterjea fixed point theorems on complete b-metric spaces. Sahand Commun. Math. Anal. 6, 77–86 (2017)

    Google Scholar 

  22. Faraji, H., Nourouzi, K.: Fixed and common fixed points for \((y, j)\)-weakly contractive mappings in b-metric spaces. Sahand Commun. Math. Anal. 7, 49–62 (2017)

    Google Scholar 

  23. Faraji, H., Nourouzi, K., O’Regan, D.: A fixed point theorem in uniform spaces generated by a family of b-pseudo metrics. Fixed Point Theory 20, 177–183 (2019)

    Article  MathSciNet  Google Scholar 

  24. Hussain, N., Mitrović, Z.D., Radenović, S.: A common fixed point theorem of Fisher in b-metric spaces. Rev. R. Acad. Cienc. Exactas Fís. Nat., Ser. A Mat. 113, 949–956 (2019)

    Article  MathSciNet  Google Scholar 

  25. Ilchev, A., Zlatanov, B.: On fixed points for Reich maps in b-metric spaces. In: Annual of Konstantin Preslavski University of Shumen, Faculty of Mathematics and Computer Science VI: Shumen, Bulgaria, pp. 77–88 (2016)

    Google Scholar 

  26. Jovanović, M., Kadelburg, Z., Radenović, S.: Common fixed point results in metric-type spaces. Fixed Point Theory Appl. 2010, 978121 (2010)

    Article  MathSciNet  Google Scholar 

  27. Khamsi, M.A., Hussain, N.: KKM mappings in metric type spaces. Nonlinear Anal. 73, 3123–3129 (2010)

    Article  MathSciNet  Google Scholar 

  28. Kirk, W., Shahzad, N.: Fixed Point Theory in Distance Spaces. Springer, Berlin (2014)

    Book  Google Scholar 

  29. Koleva, R., Zlatanov, B.: On fixed points for Chatterjea’s maps in b-metric spaces. Turk. J. Anal. Number Theory 4, 31–34 (2016)

    Google Scholar 

  30. Miculescu, R., Mihail, A.: A generalization of Matkowski’s fixed point theorem and Istratescu’s fixed point theorem concerning convex contractions. J. Fixed Point Theory Appl. 19(2), 1525–1533 (2017)

    Article  MathSciNet  Google Scholar 

  31. Mitrović, Z.D., Hussain, N.: On weak quasicontractions in b-metric spaces. Publ. Math. (Debr.) 94(3–4), 289–298 (2019)

    Article  MathSciNet  Google Scholar 

  32. Roshan, J.R., Parvaneh, V., Altun, I.: Some coincidence point results in ordered b-metric spaces and applications in a system of integral equations. Appl. Math. Comput. 226, 725–737 (2014)

    MathSciNet  Google Scholar 

  33. Sing, S.L., Czerwik, S., Krol, K., Singh, A.: Coincidences and fixed points of hybrid contractions. Tamsui Oxf. J. Math. Sci. 24, 401–416 (2008)

    MathSciNet  Google Scholar 

  34. Suzuki, T.: Basic inequality on a b-metric space and its applications. J. Inequal. Appl. 2017, 256 (2017)

    Article  MathSciNet  PubMed  PubMed Central  Google Scholar 

  35. Zoto, K., Rhoades, B., Radenović, S.: Common fixed point theorems for a class of \((s, q)\)-contractive mappings in b-metric-like spaces and applications to integral equations. Math. Slovaca 69, 233–247 (2019)

    Article  MathSciNet  Google Scholar 

  36. Ma, Z., Nazam, M., Khan, S.U., Li, X.: Fixed point theorems for generalized-contractions with applications. J. Funct. Spaces 2018, 8368546 (2018)

    MathSciNet  Google Scholar 

  37. Nazam, M., Arshad, M., Postolache, M.: Coincidence and common fixed point theorems for four mappings satisfying \((\alpha (s), F)\)-contraction. Nonlinear Anal., Model. Control 23(5), 664–690 (2018)

    Article  MathSciNet  Google Scholar 

  38. Nazam, M., Hussain, N., Hussain, A., Arshad, M.: Fixed point theorems for weakly beta-admissible pair of F-contractions with application. Nonlinear Anal., Model. Control 24(6), 898–918 (2019)

    MathSciNet  Google Scholar 

  39. Nazam, M., Acar, Ö.: Fixed points of \((\alpha , \psi )\)-contractions in Hausdorff partial metric spaces. Math. Methods Appl. Sci. 42(16), 5159–5173 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  40. Kir, M., Kiziltunc, H.: On some well known fixed point theorems in b-metric spaces. Turk. J. Anal. Number Theory 1(1), 13–16 (2013)

    Article  Google Scholar 

  41. Huang, H., Vujaković, J., Radenović, S.: A note on common fixed point theorems for isotone increasing mappings in ordered b-metric spaces. J. Nonlinear Sci. Appl. 8(5), 808–815 (2015)

    Article  MathSciNet  Google Scholar 

  42. Chandok, S.C., Jovanović, M.S., Radenović, S.N.: Ordered b-metric spaces and Geraghty type contractive mappings. Vojnotehnicki Glasnik/Mil. Tech. Cour. 65(2), 331–345 (2017)

    Article  Google Scholar 

  43. Karapinar, E., Fulga, A., Kumam, P.: Revisiting the Meir–Keeler contraction via simulation function. Filomat 34(5), 1645–1657 (2020)

    Article  MathSciNet  Google Scholar 

  44. Khojasteh, F., Shukla, S., Radenovic, S.: A new approach to the study of fixed point theorems for simulation functions. Filomat 29(6), 1189–1194 (2015)

    Article  MathSciNet  Google Scholar 

  45. Suzuki, T.: Fixed point theorems for contractions of rational type in complete metric spaces. J. Nonlinear Sci. Appl. 11, 98–107 (2018)

    Article  MathSciNet  Google Scholar 

  46. Boriceanu, M., Bota, M., Petrusel, A.: Multivalued fractals in b-metric spaces. Cent. Eur. J. Math. 8(2), 367–377 (2010)

    Article  MathSciNet  Google Scholar 

  47. Boriceanu, M.: Strict fixed point theorems for multivalued operators in b-metric spaces. Int. J. Mod. Math. 4, 285–301 (2009)

    MathSciNet  Google Scholar 

  48. Hussain, N., Doric, D., Kadelburg, Z., Radenović, S.: Suzuki-type fixed point results in metric type spaces. Fixed Point Theory Appl. 2012, 126 (2012)

    Article  MathSciNet  Google Scholar 

  49. Popescu, O.: Some new fixed point theorems for α-Geraghty contraction type maps in metric spaces. Fixed Point Theory Appl. 2014(1), 190 (2014)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the College of Natural Sciences, Jimma University for funding this research work.

Funding

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

MAM contributed to the conceptualization, formal analysis, methodology, writing, editing, and approving the manuscript. KKT involved in formal analysis, methodology and writing the original draft. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Kidane Koyas Tola.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abduletif Mamud, M., Koyas Tola, K. Fixed point theorems for generalized \((\alpha , \phi )\)-Meir–Keeler type hybrid contractive mappings via simulation function in b-metric spaces. Fixed Point Theory Algorithms Sci Eng 2024, 4 (2024). https://doi.org/10.1186/s13663-023-00758-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13663-023-00758-7

Keywords