1 Introduction and preliminaries

Throughout this paper, we denote \(\mathbb{R}^{+}_{0} = [0, +\infty)\) and \(\mathbb{N}_{0} = \mathbb{N} \cup\{0\}\), where \(\mathbb{N}\) is the set of all positive integers. First, we recall some basic concepts and notation.

The concept of b-metric was introduced by Czerwik [1] as a generalization of metric (see also Bakhtin [2, 3]) to extend the celebrated Banach contraction mapping principle. Following the initial paper of Czerwik [1], a number of researchers in nonlinear analysis investigated the topology of the paper and proved several fixed point theorems in the context of complete b-metric spaces (see [48] and references therein).

Definition 1.1

[1]

Let X be a nonempty set, and \(s\geq1\) be a given real number. A mapping \(d : X \times X\to[0, +\infty)\) is said to be a b-metric if for all \(x, y, z \in X\), the following conditions are satisfied:

(b1):

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

(b2):

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

(b3):

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

In this case, the pair \((X, d)\) is called a b-metric space (with constant s).

Definition 1.2

[9]

Let X be a nonempty set, and \(s\geq1\) be a given real number. A mapping \(d : X \times X\to[0, +\infty)\) is said to be a quasi-b-metric if for all \(x, y, z \in X\), the following conditions are satisfied:

(bm1):

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

(bm2):

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

In this case, the pair \((X,d)\) is called a quasi-b-metric space (with constant s).

Definition 1.3

[10]

Let X be a nonempty set, and \(s\geq1\) be a given real number. A mapping \(d : X \times X\to[0, +\infty)\) is said to be a quasi-b-metric-like if for all \(x, y, z \in X\), the following conditions are satisfied:

(bM1):

\(d(x, y) =0\) implies \(x = y\);

(bM2):

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

In this case, the pair \((X, d)\) is called a quasi-b-metric-like space (with constant s).

Example 1.4

Let \(X=\{0,\frac{1}{2},\frac{1}{3}\} \cup [1, \infty)\), and let \(d : X \times X\to[0, +\infty)\) be defined as

$$d(x,y)=\left \{ \textstyle\begin{array}{l@{\quad}l} 6& \text{if } x=y=0, \\ 3 &\text{if } x=y=\frac{1}{3}, \\ 2 &\text{if } x=0, y=\frac{1}{2}, \\ \frac{1}{2} &\text{if } x=0, y=\frac{1}{3}, \\ \frac{3}{2} &\text{if } x=\frac{1}{3}, y=0, \\ |x-y| &\text{otherwise}. \end{array}\displaystyle \right . $$

It is clear that \((X,d)\) is a quasi-b-metric-like space with constant \(s=9\).

Definition 1.5

(see e.g. [10])

Let \((X, d)\) be a quasi-b-metric-like space. Then:

(i)a :

a sequence \(\{x_{n}\}\) in X is called a left-Cauchy sequence if and only if for every \(\varepsilon>0\), there exists a positive integer \(N=N(\varepsilon)\) such that \(d(x_{n}, x_{m})<\varepsilon\) for all \(n>m>N\);

(ii)b :

a sequence \(\{x_{n}\}\) in X is called a right-Cauchy sequence if and only if for every \(\varepsilon>0\), there exists a positive integer \(N=N(\varepsilon)\) such that \(d(x_{n}, x_{m})<\varepsilon\) for all \(m>n>N\);

(iii)a :

a quasi-partial metric space is said to be left-complete if every left-Cauchy sequence \(\{x_{n}\}\) in X converges with respect to d to a point \(u \in X\) such that

$$\lim_{n \to\infty}d( x_{n},u) = d(u,u) =\lim _{n,m\to\infty}d(x_{m}, x_{n})=0,\quad \text{where } m\geq n; $$
(iii)b :

a quasi-partial metric space is said to be right-complete if every left-Cauchy sequence \(\{x_{n}\}\) in X converges with respect to d to a point \(u \in X\) such that

$$\lim_{n \to\infty}d(u, x_{n}) = d(u,u) =\lim _{n,m\to\infty}d(x_{n}, x_{m})=0,\quad \text{where } m\geq n. $$

Let \((X, d)\) and \((Y, \alpha)\) be quasi-b-metric-like spaces, and let \(f : X \to Y\) be a continuous mapping. Then

$$\lim_{n \to\infty} x_{n}=u \quad \Rightarrow\quad \lim _{n \to\infty} fx_{n}=fu. $$

In 2012, Samet et al. [11] introduced the concept of α-admissible mappings, and in 2013, Karapınar et al. [12] improved this notion as triangular α-admissible mappings.

Definition 1.6

[11, 12]

Let \(\alpha: X \times X \rightarrow[0, +\infty)\) be a function. A self-mapping f is called an α-admissible mapping if

$$\alpha(x,y)\geq1 \quad \Rightarrow \quad \alpha(fx,fy)\geq1 $$

for all \(x,y \in X\). If, further, f satisfies the condition

$$\alpha(x,z)\geq1 \quad \text{and}\quad \alpha(z,y)\geq1 \quad \Rightarrow \quad \alpha(x,y)\geq1 $$

for all \(x,y,z \in X\), then it is called triangular α-admissible mapping.

Very recently, Popescu [13] improved these notions as follows.

Definition 1.7

[13]

Let \(\alpha:X\times X\rightarrow[0,\infty)\) be a function. If \(f: X\to X\) satisfies the condition

$$ (\mathrm{T}1)'\quad \alpha(x,fx)\geq1\quad \Rightarrow\quad \alpha \bigl(fx,f^{2}x\bigr) \geq1 $$

for all \(x\in X\), then it is called a right-α-orbital admissible mapping. If f satisfies the condition

$$ (\mathrm{T}1)'' \quad \alpha(fx,x)\geq1 \quad \Rightarrow\quad \alpha\bigl(f^{2}x,fx\bigr) \geq1 $$

for all \(x\in X\), then it is called a left-α-orbital admissible mapping. Furthermore, if f is both right-α-orbital admissible and left-α-orbital admissible, then f is called an α-orbital admissible mapping.

Triangular α-admissible mappings defined by Popescu [13] impose the following definitions.

Definition 1.8

[13]

Let \(f:X\rightarrow X\) be a self-mapping, and \(\alpha:X\times X\rightarrow[0,\infty)\) be a function. Then f is said to be triangular right-α-orbital admissible if f is right-α-orbital admissible and

$$ (\mathrm{T}2)'\quad \alpha(x,y)\geq1\quad \text{and}\quad \alpha(y,fy)\geq 1 \quad \Rightarrow\quad \alpha(x,fy) \geq1 $$

and is said to be triangular left-α-orbital admissible if f is α-orbital admissible and

$$ (\mathrm{T}2)'' \quad \alpha(fx,x)\geq1 \quad \text{and} \quad \alpha(x,y)\geq 1\quad \Rightarrow\quad \alpha(fx,y) \geq1. $$

If T satisfies both (T2)′ and (T2)″, then it is called triangular α-orbital admissible.

It is easy to conclude that each α-admissible mapping is an α-orbital admissible mapping and each triangular α-admissible mapping is a triangular α-orbital admissible mapping. However, the converses of the statements are false. In the following example, we see that a mapping that is triangular α-orbital admissible need not be triangular α-admissible.

Example 1.9

Let \(X = \{x_{i}: i=1,\ldots, n\}\) for some \(n\geq4\), and \(d : X \times X \to\mathbb{R}_{0}^{+}\) with \(d(x, y) = |x-y|\). We define a self-mapping \(f : X \to X \) such that \(fx_{i}=x_{i}\) for \(i=1,2\), \(fx_{i}=x_{j}\) for \(i,j \in\{ 3,4\}\), \(i\neq j\), \(fx_{i}=x_{i+1}\) for \(i \in\{5,\ldots, n-1\}\), and \(fx_{n}=fx_{5}\). Moreover, let \(\alpha: X \times X \to\mathbb{R}_{0}^{+}\) be such that

$$\alpha(x, y) = \left \{ \textstyle\begin{array}{l@{\quad}l} 1 & \text{if } (x, y) \in \{(x_{1},x_{3}), (x_{1},x_{4}), (x_{3},x_{3}), (x_{4},x_{4}), \\ &\hphantom{\text{if } (x, y) \in{}}{}(x_{3},x_{4}), (x_{4},x_{3}), (x_{3},x_{2}), (x_{4},x_{2}) \}, \\ 0 & \text{otherwise}. \end{array}\displaystyle \right . $$

Note that f is α-orbital admissible since \(\alpha(x_{3},fx_{3}) = \alpha(x_{3},x_{4}) = 1 \) and \(\alpha(x_{4},fx_{4}) = \alpha(x_{4},x_{3}) =1 \). On the other hand, we have \(\alpha(x_{1},x_{3}) = \alpha(x_{3},x_{2}) = 1 \), but \(\alpha(x_{1},x_{2})=0\). Hence, T is not triangular α-admissible.

Definition 1.10

[13]

Let \((X,d)\) be a quasi-b-metric-like space. Then X is said to be α-regular if for every sequence \(\{x_{n}\}\) in X such that \(\alpha(x_{n},x_{n+1})\geq1\) for all n and \(x_{n}\rightarrow x\in X\) as \(n\rightarrow\infty\), there exists a subsequence \(\{x_{n(k)}\}\) of \(\{x_{n}\}\) such that \(\alpha(x_{n(k)},x)\geq1\) for all k.

2 Main result

The notion of \((b)\)-comparison was introduced by Berinde [14] in order to extend the notion of \((c)\)-comparison.

Definition 2.1

[14]

Let \(s\geq1\) be a real number. A mapping \(\psi:\mathbb{R}_{0}^{+}\to\mathbb{R}_{0}^{+}\) is called a \((b)\)-comparison function if the following conditions are fulfilled:

  1. (1)

    ψ is monotone increasing;

  2. (2)

    there exist \(k_{0}\in\mathbb{N}\), \(a\in(0,1)\), and a convergent series of nonnegative terms \(\sum_{k=1}^{\infty}v_{k}\) such that \(s^{k+1}\psi ^{k+1}(t)\leq a s^{k}\psi^{k}(t)+v_{k}\) for all \(k\geq k_{0}\) and \(t\in[0,\infty)\).

The class of \((b)\)-comparison functions will be denoted by \(\Psi_{b}\). Notice that the notion of a \((b)\)-comparison function reduces to the concept of a \((c)\)-comparison function if \(s=1\).

The following lemma will be used in the proof of our main result.

Lemma 2.2

[15, 16]

Let \(s\geq1\) be a real number. If \(\psi :\mathbb{R}_{0}^{+}\to\mathbb{R}_{0}^{+}\) is a \((b)\)-comparison function, then:

  1. (1)

    the series \(\sum_{k=0}^{\infty}s^{k}\psi^{k}(t)\) converges for any \(t\in\mathbb{R}_{0}^{+}\);

  2. (2)

    the function \(p_{s}:[0,\infty)\to[0,\infty)\) defined by

    $$p_{s}(t)= \sum_{k=0}^{\infty}s^{k}\psi^{k}(t)\quad \textit{for all } t\in[0,\infty) $$

    is increasing and continuous at 0.

Remark 2.3

It is easy to see that if \(\psi(t) \in\Psi _{b}\), then \(\psi(t)< t\) for all \(t>0\). In fact, if there is a \(t^{*}>0\) such that \(\psi(t^{*})\geq t^{*}\), then we have \(\psi^{2}(t^{*})\geq\psi(t^{*})\geq t^{*}\) (since ψ is increasing). Continuing in the same manner, we get \(\psi^{n}(t^{*})\geq t^{*}>0\), \(n\in \mathbb{N}\). This contradicts Lemma 2.2.

Definition 2.4

Let \((X,d)\) be a complete quasi-b-metric-like space with a constant \(s \geq1\). A self-mapping \(f: X \to X\) is called \((\alpha ,\psi)\)-contractive mapping if there exist two functions \(\psi\in \Psi_{b}\) and \(\alpha: X \times X \to[0,\infty)\) satisfying the following condition:

$$ \alpha(x,y)d(fx,fy) \leq\psi\bigl(d(x,y)\bigr) $$
(2.1)

for all \(x,y \in X\).

Theorem 2.5

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be an \((\alpha,\psi)\)-contractive mapping. Suppose also that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    f is continuous.

Then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

By (ii) there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0})\geq1\) and \(\alpha(fx_{0},x_{0})\geq1\). Define the iterative sequence \(\{x_{n}\}\) in X by \(x_{n+1}=fx_{n}\) for all \(n\in\mathbb{N}_{0}\). Note that if there exists \(n_{0} \in\mathbb{N}_{0}\) such that \(x_{n_{0}}= x_{n_{0}+1}\), then \(x_{n_{0}}\) becomes a fixed point, which completes the proof. Hence, throughout the proof, we suppose that \(x_{n}\neq x_{n+1}\) for all \(n\in\mathbb{N}_{0}\). Regarding the fact that f is α-orbital admissible, from (ii) we derive that

$$\alpha(x_{0},x_{1})= \alpha(x_{0}, fx_{0}) \geq1 \quad \Rightarrow \quad \alpha(fx_{0},fx_{1})= \alpha(x_{1},x_{2}) \geq1. $$

Inductively, we get that

$$ \alpha(x_{n},x_{n+1})\geq1\quad \text{for all } n\in\mathbb{N}_{0}. $$
(2.2)

Analogously, again by (ii) and the fact that f is α-orbital admissible we find that

$$\alpha(x_{1},x_{0})= \alpha(fx_{0}, x_{0}) \geq1 \quad \Rightarrow \quad \alpha(fx_{1},fx_{0})= \alpha(x_{2},x_{1}) \geq1. $$

Consequently, we observe that

$$ \alpha(x_{n+1},x_{n})\geq1 \quad \text{for all } n\in\mathbb{N}_{0}. $$
(2.3)

From (2.1), by taking \(x=x_{n}\) and \(y=x_{n-1}\), we find that

$$\begin{aligned} d(x_{n+1},x_{n})&= d(fx_{n},fx_{n-1}) \\ &\leq\alpha(x_{n},x_{n-1})d(fx_{n},fx_{n-1}) \\ & \leq\psi\bigl(d(x_{n},x_{n-1})\bigr). \end{aligned}$$

In view of Remark 2.3, we get that

$$ d(x_{n+1},x_{n}) \leq\psi\bigl(d(x_{n},x_{n-1}) \bigr) < d(x_{n},x_{n-1}) \quad \text{for all } n\in\mathbb {N}. $$
(2.4)

By analogy, again by (2.1) and by substituting \(x=x_{n-1}\) and \(y=x_{n}\), we have

$$\begin{aligned} d(x_{n},x_{n+1})&= d(fx_{n-1},fx_{n}) \\ &\leq\alpha(x_{n-1},x_{n })d(fx_{n-1},fx_{n }) \\ & \leq\psi\bigl(d(x_{n-1},x_{n})\bigr). \end{aligned}$$

Consequently,

$$ d(x_{n},x_{n+1})\leq\psi\bigl(d(x_{n-1},x_{n}) \bigr) < d(x_{n-1},x_{n}) \quad \text{for all } n\in\mathbb {N}. $$
(2.5)

From (2.4) and (2.5) we derive that

$$ d(x_{n},x_{n+1})\leq\psi^{n} \bigl(d(x_{0},x_{1})\bigr)\quad \text{and}\quad d(x_{n+1},x_{n})\leq\psi^{n}\bigl(d(x_{1},x_{0}) \bigr) \quad \text{for all } n\in\mathbb {N}. $$
(2.6)

By Lemma 2.2(1) and letting \(n \to\infty\) in (2.6), we have \(\lim_{n\rightarrow\infty}d(x_{n},x_{n+1})=\lim_{n\rightarrow \infty}d(x_{n+1}, x_{n})=0\).

We further prove that the sequence \(\{x_{n}\}\) is right-Cauchy and left-Cauchy. For all \(n,p\in\mathbb{N}\), we have

$$\begin{aligned} d(x_{n},x_{n+p})&\leq \sum_{i=1}^{p-1} s^{i}d(x_{n+i-1},x_{n+i})+s^{p-1}d(x_{n+p-1},x_{n+p}) \\ &< \sum_{i=1}^{p} s^{i} \psi^{n+i-1}\bigl(d(x_{0},x_{1})\bigr) \\ &= \frac{1}{s^{n-1} } \sum^{n+p-1}_{k=n} s^{k} \psi^{k}\bigl(d(x_{0},x_{1}) \bigr). \end{aligned}$$

By letting \(n,p \to\infty\) we get that

$$ \lim_{n,p\to\infty} d(x_{n},x_{n+p})=0, $$

that is, the sequence \(\{x_{n}\}\) is right-Cauchy.

Analogously,

$$ \lim_{n,p\to\infty} d(x_{n+p},x_{n})=0, $$

that is, the sequence \(\{x_{n}\}\) is left-Cauchy. As a result, the sequence \(\{x_{n}\}\) is a Cauchy sequence. Since \((X,d)\) is complete, there exists a point \(u \in X\) such that

$$ \lim_{n \to\infty}d(u, x_{n})=\lim _{n \to\infty}d( x_{n},u) = d(u,u) =\lim_{n,m\to\infty}d(x_{n}, x_{m})= \lim_{n,m\to\infty}d(x_{m}, x_{n})=0. $$
(2.7)

Since f is continuous, we have

$$u=\lim_{n \to\infty} x_{n+1}= \lim_{n \to\infty} fx_{n}=fu. $$

 □

Example 2.6

Let \((X,d)\) be a quasi b-metric like space defined in Example 1.4, and let the mapping \(f:X\mapsto X\) be defined as

$$fx=\left \{ \textstyle\begin{array}{l@{\quad}l} \frac{1}{3} &\text{if } x=0, \\ \frac{1}{2} &\text{if } x=\frac{1}{3}, \\ \frac{1}{2} &\text{if } x=\frac{1}{2}, \\ x+1 &\text{if } x\geq1. \end{array}\displaystyle \right . $$

Let \(\psi(t)=\frac{t}{2}\), \(t\geq0\), and let \(\alpha:X\times X\to [0,\infty)\) be defined as

$$\alpha(x,y)=\left \{ \textstyle\begin{array}{l@{\quad}l} 1 &\text{if } x,y\in\{0,\frac{1}{3},\frac{1}{2}\}, \\ 0 &\text{otherwise}. \end{array}\displaystyle \right . $$

Then \(\psi\in\Psi_{b}\), and f is an \((\alpha,\psi)\)-contractive mapping. Since the conditions of Theorem 2.5 are satisfied, it follows that f has a fixed point in X.

It is possible to remove the heavy condition of continuity of the self-mapping f in Theorem 2.5. For this purpose, we need the following result, which is inspired from the results in [17].

Lemma 2.7

Let \((X,d)\) be a quasi-b-metric-like space with constant s and assume that \(\{x_{n}\}\) and \(\{y_{n}\}\) are sequences in X converging to x and y, respectively. Then

$$\begin{aligned} \frac{1}{s^{2}} d(x,y)-\frac{1}{s} d(x,x)-d(y,y) &\leq\liminf _{n\rightarrow\infty} d(x_{n},y_{n}) \leq \limsup _{n\rightarrow\infty} d(x_{n},y_{n}) \\ &\leq s d(x,x) + s^{2} d(y,y) + s^{2} d(x,y) . \end{aligned}$$

In particular, if \(d(x,y)=0 \), then \(\lim_{n\rightarrow \infty} d(x_{n},y_{n})=0 \).

Moreover, for each \(z\in X \), we have

$$ \frac{1}{s} d(x,z)- d(x,x) \leq\liminf_{n\rightarrow\infty} d(x_{n},z) \leq\limsup_{n\rightarrow\infty} d(x_{n},z) \leq s d(x,z) + s d(x,x). $$
(2.8)

If \(d(x,x)=0\), then

$$\frac{1}{s} d(x,z) \leq\liminf_{n\rightarrow\infty} d(x_{n},z) \leq\limsup_{n\rightarrow\infty} d(x_{n},z) \leq s d(x,z). $$

Theorem 2.8

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be an \((\alpha,\psi)\)-contractive mapping. Suppose also that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    X is α-regular.

Then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

By verbatim of the proof of Theorem 2.5 we find an iterative sequence \(\{x_{n}\}\) that converges to a point \(u \in X\) such that (2.7) holds.

Since \(d(u,u)=0\), by Lemma 2.7 we have

$$\begin{aligned} \frac{1}{s} d(u,fu) &\leq\liminf_{n\rightarrow\infty} d(x_{n+1},fu) \\ &\leq\limsup_{n\rightarrow\infty} d(x_{n+1},fu) \\ &=\limsup_{n\rightarrow\infty} d(fx_{n},fu) \\ &\leq\limsup_{n\rightarrow\infty}\alpha (x_{n},u)d(fx_{n},fu) \\ & \leq\limsup_{n\rightarrow\infty}\psi\bigl(d(x_{n},u)\bigr). \end{aligned}$$

By letting \(n \to\infty\) in these inequalities we derive that \(\frac {1}{s} d(u,fu) =0\) and hence \(fu=u\). □

It is natural to consider the uniqueness of a fixed point of an \((\alpha,\psi)\)-contractive mapping. We notice that we need to add an additional condition to guarantee the uniqueness.

  1. (U)

    For all \(x,y\in\operatorname{Fix}(f)\), either \(\alpha(x,y)\geq1\) or \(\alpha (y,x)\geq1\).

Here, \(\operatorname{Fix}(f)\) denotes the set of all fixed points of f.

Theorem 2.9

Adding condition (U) to hypotheses of Theorem  2.5 (or Theorem  2.8), we obtain the uniqueness of a fixed point of f.

Proof

Suppose that \(x^{*}\) and \(y^{*}\) are two distinct fixed points of f, so that \(d(x^{*},y^{*})>0\).

If, for example, \(\alpha(x^{*},y^{*})\geq1\), then

$$\begin{aligned} d\bigl(x^{*},y^{*}\bigr)&= d\bigl(fx^{*},fy^{*}\bigr) \\ &\leq\alpha\bigl(x^{*},y^{*}\bigr)d\bigl(fx^{*},fy^{*}\bigr) \\ &\leq\psi\bigl(d\bigl(x^{*},y^{*}\bigr)\bigr) \\ &< d\bigl(x^{*},y^{*}\bigr), \end{aligned}$$

which is a contradiction. □

Definition 2.10

Let \((X,d)\) be a complete quasi-b-metric-like space with a constant \(s \geq1\). A self-mapping \(f: X \to X\) is called a generalized \((\alpha,\psi)\)-contractive mapping of type \((A)\) if there exist two functions \(\psi\in\Psi_{b}\) and \(\alpha: X \times X \to[0,\infty)\) satisfying the following condition:

$$ \alpha(x,y) d(fx,fy) \leq\psi\bigl(M(x,y)\bigr) $$
(2.9)

for all \(x,y \in X\), where

$$ M(x,y)=\max \bigl\{ d(x,y),d(x,fx),d(y,fy) \bigr\} . $$
(2.10)

Theorem 2.11

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a generalized \((\alpha,\psi)\)-contractive mapping of type \((A)\). Assume that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    f is continuous.

Then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

As in the proof of Theorem 2.5, we construct an iterative sequence \(x_{n+1}=fx_{n}\), \(n\in\mathbb{N}_{0}\), where the existence of \(x_{0}\in X\) is guaranteed by (ii). By the same reason as in the proof of Theorem 2.5, we may assume that \(x_{n}\neq x_{n+1}\) for all \(n\in\mathbb{N}_{0}\), and we can conclude that

$$ \alpha(x_{n},x_{n+1})\geq1 \quad \text{and} \quad \alpha (x_{n+1},x_{n})\geq1\quad \text{for all } n\in \mathbb{N}_{0}. $$
(2.11)

From (2.9) we have

$$\begin{aligned} \begin{aligned} d(x_{n},x_{n+1})&= d(fx_{n-1},fx_{n}) \\ &\leq\alpha(x_{n-1},x_{n})d(fx_{n-1},fx_{n}) \\ &\leq\psi\bigl(M(x_{n-1},x_{n})\bigr) \end{aligned} \end{aligned}$$

for all \(n \in\mathbb{N} \), where

$$\begin{aligned} M(x_{n-1},x_{n})&=\max \bigl\{ d(x_{n-1},x_{n }),d(x_{n},fx_{n}),d(x_{n-1},fx_{n-1}) \bigr\} \\ &=\max \bigl\{ d(x_{n-1},x_{n}),d(x_{n},x_{n+1}) \bigr\} . \end{aligned}$$

If \(M(x_{n-1},x_{n})=d(x_{n},x_{n+1})\), then since we assumed that \(x_{n}\neq x_{n+1}\),

$$d(x_{n},x_{n+1})\leq\psi\bigl(d(x_{n},x_{n+1}) \bigr)< d(x_{n},x_{n+1}), $$

which is a contradiction. It allows us to conclude that \(M(x_{n-1},x_{n})=d(x_{n-1},x_{n})\), \(n\in\mathbb{N}\).

Thus,

$$d(x_{n},x_{n+1})\leq\psi\bigl(d(x_{n-1},x_{n}) \bigr)< d(x_{n-1},x_{n})\quad \text{for all } n\in\mathbb{N} $$

and

$$ d(x_{n},x_{n+1})\leq\psi^{n} \bigl(d(x_{0},x_{1})\bigr)\quad \text{for all } n\in \mathbb{N}. $$
(2.12)

Analogously, letting \(x=x_{n}\) and \(y=x_{n-1}\) in (2.9), we get

$$\begin{aligned} d(x_{n+1},x_{n}) =& d(fx_{n},fx_{n-1}) \\ \leq& \alpha(x_{n},x_{n-1})d(fx_{n},fx_{n-1}) \\ \leq& \psi\bigl(M(x_{n},x_{n-1})\bigr) \end{aligned}$$
(2.13)

for all \(n \in\mathbb{N} \), where

$$\begin{aligned} M(x_{n},x_{n-1}) =&\max \bigl\{ d(x_{n},x_{n-1}),d(x_{n},fx_{n}),d(x_{n-1},fx_{n-1}) \bigr\} \\ =&\max \bigl\{ d(x_{n},x_{n-1}),d(x_{n},x_{n+1}),d(x_{n-1},x_{n}) \bigr\} . \end{aligned}$$

For the estimation of \(d(x_{n+1},x_{n})\), we will consider three different cases.

Case 1. If \(M(x_{n},x_{n-1})=d(x_{n-1},x_{n})\), then, by (2.13),

$$ d(x_{n+1},x_{n})\leq\psi\bigl(d(x_{n-1},x_{n}) \bigr). $$
(2.14)

Case 2. If \(M(x_{n},x_{n-1})=d(x_{n},x_{n+1})\), then

$$ d(x_{n+1},x_{n})\leq\psi\bigl(d(x_{n},x_{n+1}) \bigr). $$

By Remark 2.3 we find that

$$d(x_{n+1},x_{n})\leq\psi\bigl(d(x_{n},x_{n+1}) \bigr)< \psi^{n+1} \bigl(d(x_{0},x_{1}) \bigr). $$

Case 3. Otherwise, \(M(x_{n},x_{n-1})=d(x_{n},x_{n-1})\) and

$$ d(x_{n+1},x_{n})\leq\psi\bigl(d(x_{n},x_{n-1}) \bigr). $$
(2.15)

Observing (2.14) and (2.15), it follows that, for any \(n\in \mathbb{N}\),

$$ d(x_{n+1},x_{n})\leq\max \bigl\{ \psi^{n} \bigl(d(x_{0},x_{1})\bigr), \psi ^{n} \bigl(d(x_{1},x_{0})\bigr) \bigr\} . $$
(2.16)

Obviously, in all considered cases, we deduce that

$$\lim_{n\rightarrow\infty}d(x_{n+1},x_{n})=\lim _{n\rightarrow\infty }d(x_{n},x_{n+1}) =0. $$

Since ψ is an increasing function, let

$$v=\max \bigl\{ d(x_{0},x_{1}), d(x_{1},x_{0}) \bigr\} . $$

Consequently, we have that \(d(x_{n+1},x_{n})\leq\psi^{n}(v)\) and \(d(x_{n},x_{n+1})\leq\psi^{n}(v)\). By applying (bM2) for any \(n,p\in\mathbb{N}\) it follows that

$$\begin{aligned} d(x_{n},x_{n+p}) \leq&\sum_{i=1}^{p-1} s^{i}d(x_{n+i-1},x_{n+i})+s^{p-1}d(x_{n+p-1},x_{n+p}) \\ \leq&\sum_{i=1}^{p} s^{i}d(x_{n+i-1},x_{n+i}) \\ \leq&\sum_{i=1}^{p} s^{i} \psi^{n+i-1}(v) \\ =&\frac{1}{s^{n-1}}\sum_{i=1}^{p} s^{n+i-1}\psi^{n+i-1}(v). \end{aligned}$$

Therefore, \(\lim_{n,p\to\infty} d(x_{n},x_{n+p})=0\) and, likewise, \(\lim_{n,p\to\infty} d(x_{n+p},x_{n})=0\). Since, X is complete, there exists \(u\in X\) such that \(\lim_{n\rightarrow \infty}x_{n}=u\) and

$$ \lim_{n\rightarrow\infty}d(u,x_{n})=\lim _{n\rightarrow\infty }d(x_{n},u)=d(u,u)=0. $$
(2.17)

Furthermore, f is a continuous mapping, and hence \(u=\lim_{n\rightarrow\infty} x_{n}=\lim_{n\rightarrow\infty}fx_{n-1}=fu\). □

Theorem 2.12

Adding condition (U) to hypotheses of Theorem  2.11, we obtain the uniqueness of a fixed point of T.

Proof

Suppose that \(fx^{*}=x^{*}\) and \(fy^{*}=y^{*}\). Then

$$\begin{aligned} d\bigl(x^{*},y^{*}\bigr)&= d\bigl(fx^{*},fy^{*}\bigr) \\ &\leq\alpha\bigl(x^{*},y^{*}\bigr)\psi\bigl( d\bigl(fx^{*},fy^{*}\bigr)\bigr) \\ &\leq\psi\bigl(M\bigl(x^{*},y^{*}\bigr)\bigr) \\ &=\psi\bigl(d\bigl(x^{*},y^{*}\bigr)\bigr), \end{aligned}$$

so that \(d(x^{*},y^{*})=0 \Rightarrow x^{*}=y^{*}\). □

In the following example, we show the existence of a function satisfying conditions of Theorem 2.11 but not satisfying conditions of Theorem 2.5.

Example 2.13

Let \((X,d)\) be a quasi-b-metric-like space described in Example 1.4, and \(f:X\mapsto X\) the mapping

$$fx=\left \{ \textstyle\begin{array}{l@{\quad}l} \frac{1}{2} &\text{if } x=0, \\ 0, &\text{if } x=\frac{1}{3}, \\ \frac{1}{2} &\text{if } x=\frac{1}{2}, \\ x+1 &\text{if } x\geq1. \end{array}\displaystyle \right . $$

Let \(\psi(t)=\frac{t}{2}\), \(t\geq0\), and let \(\alpha:X\times X\to [0,\infty)\) be defined as

$$\alpha(x,y)=\left \{ \textstyle\begin{array}{l@{\quad}l} 1 &\text{if } (x,y)\in \{(0,\frac{1}{3}),(0,\frac {1}{2}),(\frac{1}{2},0),(\frac{1}{2},\frac{1}{3}),(\frac {1}{2},\frac{1}{2}) \}, \\ 0 &\text{otherwise}. \end{array}\displaystyle \right . $$

Then (2.1) does not hold, for example, for \(x=0\) and \(y=\frac {1}{3}\), but (2.9) holds, f has a unique fixed point \(u=\frac {1}{2}\), and \(d(u,u)=0\).

Definition 2.14

Let \((X,d)\) be a complete quasi-b-metric-like space with a constant \(s \geq1\). A self-mapping \(f: X \to X\) is called a generalized \((\alpha,\psi)\)-contractive mapping of type \((B)\) if there exist two functions \(\psi\in\Psi_{b}\) and \(\alpha: X \times X \to[0,\infty)\) satisfying the following condition:

$$ \alpha(x,y) d(fx,fy) \leq\psi\bigl(N(x,y)\bigr) $$
(2.18)

for all \(x,y \in X\), where

$$ N(x,y)=\max \biggl\{ d(x,y), \frac{d(x,fx)+d(y,fy)}{2} \biggr\} . $$
(2.19)

The following theorem can be deduced from the inequality \(N(x,y) \leq M(x,y)\) for all x, y, together with the monotonicity of ψ.

Theorem 2.15

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a generalized \((\alpha,\psi)\)-contractive mapping of type \((B)\). Assume that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    f is continuous.

Then f has a fixed point u in X, and \(d(u,u)=0\).

Definition 2.16

Let \((X,d)\) be a complete quasi-b-metric-like space with a constant \(s \geq1\). A self-mapping \(f: X \to X\) is called a generalized \((\alpha,\psi)\)-contractive mapping of type \((C)\) if there exist two functions \(\psi\in\Psi_{b}\) and \(\alpha: X \times X \to[0,\infty)\) satisfying the following condition:

$$ s\alpha(x,y) d(fx,fy) \leq\psi\bigl(M(x,y)\bigr) $$
(2.20)

for all \(x,y \in X\), where

$$ M(x,y)=\max \bigl\{ d(x,y),d(x,fx),d(y,fy) \bigr\} . $$
(2.21)

The following theorem is easily observed from Theorem 2.11 since inequality (2.9) can be easily derived from inequality (2.20).

Theorem 2.17

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a generalized \((\alpha,\psi)\)-contractive mapping of type \((C)\). Assume that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    f is continuous.

Then f has a fixed point u in X, and \(d(u,u)=0\).

In the next theorems, we establish a fixed point result for a generalized \((\alpha,\psi)\)-contractive mapping of type \((C)\) without any continuity assumption on the mapping f.

Theorem 2.18

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a generalized \((\alpha,\psi)\)-contractive mapping of type \((C)\). Suppose that

  1. (i)

    f is α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    X is α-regular.

Then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

As in the proof of Theorem 2.5, we consider an iterative sequence \(\{x_{n}\}\), and we obtain the existence of \(u\in X\) such that (2.17) holds. By Lemma 2.7 we get

$$\begin{aligned} d(u,fu) \leq& s \liminf_{n\to\infty} d(x_{n+1},fu) \\ \leq& s\limsup_{n\to\infty} d(x_{n+1},fu) \\ \leq& s\limsup_{n\to\infty} \alpha(x_{n},u) d(fx_{n},fu) \\ \leq& \limsup_{n\to\infty} \psi\bigl(M(x_{n},u)\bigr), \end{aligned}$$

where

$$M(x_{n},u)=\max \bigl\{ d(x_{n-1},u), d(x_{n-1},x_{n}),d(u,fu) \bigr\} . $$

According to (2.17) and the fact that \(\lim_{n\rightarrow\infty }d(x_{n-1},x_{n})=0\), it remains to discuss only the case \(M(x_{n},u)=d(u,fu)\) because otherwise it follows \(d(u,fu)=0 \Rightarrow u=fu\).

Notice that, under this assumption, \(d(u,fu)\leq\psi(d(u,fu))\) also implies \(d(u,fu)=0\) since \(\psi(t)< t\) for any \(t>0\). Hence, u is a fixed point of the mapping f. □

Theorem 2.19

Adding condition (U) to hypotheses of Theorem  2.17 (or Theorem  2.18), we obtain the uniqueness of a fixed point of T.

Example 2.20

Let \((X,d)\) be a quasi b-metric like space defined in Example 1.4, and let the mapping \(f:X\mapsto X\) be defined as

$$fx=\left \{ \textstyle\begin{array}{l@{\quad}l} 1 &\text{if } x=0, \\ \frac{1}{2} &\text{if } x=\frac{1}{3}, \\ \frac{1}{2} &\text{if } x=\frac{1}{2}, \\ x+1 &\text{if } x\geq1. \end{array}\displaystyle \right . $$

Let \(\psi(t)=\frac{t}{2}\), \(t\geq0\), and \(\alpha:X\times X\to [0,\infty)\) be defined as

$$\alpha(x,y)=\left \{ \textstyle\begin{array}{l@{\quad}l} 1 &\text{if } (x,y)\in \{(\frac{1}{3},\frac {1}{2}),(\frac{1}{2},\frac{1}{2}) \}, \\ \frac{1}{9} &\text{if } (x,y)\in \{0,\frac{1}{3},\frac {1}{2} \}\times \{0,\frac{1}{3},\frac{1}{2} \} \setminus \{(\frac{1}{3},\frac{1}{2}),(\frac{1}{2},\frac {1}{2}) \}, \\ 0 &\text{otherwise}. \end{array}\displaystyle \right . $$

Then \(\psi\in\Psi_{b}\), and f is a generalized \((\alpha,\psi )\)-contractive mapping of type \((C)\). Since the conditions of Theorem 2.18 are satisfied, it follows that f has a fixed point in X.

Definition 2.21

Let \((X,d)\) be a complete quasi-b-metric-like space with a constant \(s \geq1\). A self-mapping \(f: X \to X\) is called a generalized \((\alpha,\psi)\)-contractive mapping of type \((D)\) if there exist two functions \(\psi\in\Psi_{b}\) and \(\alpha: X \times X \to[0,\infty)\) satisfying the following condition:

$$ \alpha(x,y) d(fx,fy) \leq\psi\bigl(L(x,y)\bigr) $$
(2.22)

for all \(x,y \in X\), where

$$ L(x,y)=\max \biggl\{ d(x,y), \frac{d(x,fx)+d(y,fy)}{2s} \biggr\} . $$
(2.23)

Theorem 2.22

Let \((X,d)\) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a generalized \((\alpha,\psi)\)-contractive mapping of type \((D)\). Suppose that

  1. (i)

    f is triangular α-orbital admissible;

  2. (ii)

    there exists \(x_{0}\in X\) such that \(\alpha(x_{0},fx_{0}) \geq1\) and \(\alpha(fx_{0},x_{0})\geq1\);

  3. (iii)

    X is α-regular.

Then f has a fixed point in X, that is, there exists \(u \in X\) such that \(fu=u\) and \(d(u,u)=0\).

Proof

As in the proof of Theorem 2.11, we consider an iterative sequence \(\{x_{n}\}\) and obtain the existence of \(u\in X\) such that (2.17) holds. By Lemma 2.7 we get

$$\begin{aligned} \frac{ 1}{s} d(u,fu) \leq& \liminf_{n\to\infty} d(x_{n+1},fu) \\ \leq& \limsup_{n\to\infty} d(x_{n+1},fu) \\ \leq& \limsup_{n\to\infty} \alpha(x_{n},u) d(fx_{n},fu) \\ \leq& \limsup_{n\to\infty} \psi\bigl(N(x_{n},u)\bigr), \end{aligned}$$

where

$$N(x_{n},u)=\max \biggl\{ d(x_{n-1},u), \frac {d(x_{n-1},x_{n})+d(u,fu)}{2s} \biggr\} . $$

If \(N(x_{n},u)=d(x_{n-1},u)\), then we conclude the result due to (2.17). Taking \(\lim_{n\rightarrow\infty}d(x_{n-1}, x_{n})=0\) into account, we deduce that \(\lim_{n\rightarrow\infty} N(x_{n},u)=\frac{ d(u,fu)}{2s}\). Notice that, under this assumption, \(\frac{ 1}{s}d(u,fu)\leq\psi (\frac{ d(u,fu)}{2s})\) also implies \(d(u,fu)=0\) since \(\psi(t)< t\) for any \(t>0\). Hence, u is a fixed point of the mapping f. □

Theorem 2.23

Adding condition (U) to hypotheses of Theorem  2.15 (and respectively, Theorem  2.22), we obtain the uniqueness of a fixed point of T.

3 Consequences

In this section, we will list some consequences of our main results.

3.1 For standard quasi-b-metric-like

Corollary 3.1

Let \((X,d)\) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a mapping such that

$$ d(fx,fy) \leq\psi\bigl(\max \bigl\{ d(x,y),d(x,fx),d(y,fy) \bigr\} \bigr) $$
(3.1)

for all \(x,y \in X\), where \(\psi\in\Psi_{b}\). If f is continuous, then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

The proof of Corollary 3.1 follows from Theorem 2.12 by taking \(\alpha(x,y)=1\) for all \(x,y \in X\), so (ii) is satisfied for any \(x_{0}\in X\), f is obviously an α-orbital admissible, and (U) holds. Inequality (3.1) allows us to conclude that f is a generalized \((\alpha,\psi)\)-contractive mapping of type \((A)\). □

Corollary 3.2

Let \((X,d)\) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a continuous mapping such that

$$ d(fx,fy) \leq\psi \biggl(\max \biggl\{ d(x,y),\frac {d(x,fx)+d(y,fy)}{2} \biggr\} \biggr) $$
(3.2)

for all \(x,y \in X\), where \(\psi\in\Psi_{b}\). Then f has a fixed point u in X, and \(d(u,u)=0\).

Proof

The proof of Corollary 3.2 follows from Theorem 2.15 by taking \(\alpha(x,y)=1\) for all \(x,y \in X\) since then (2.18) follows from (3.2). □

Notice that the continuity condition of f in Corollary 3.1 can be removed by adding an extra term s.

Corollary 3.3

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a mapping such that

$$ d(fx,fy) \leq s \psi\bigl(\max \bigl\{ d(x,y),d(x,fx),d(y,fy) \bigr\} \bigr) $$
(3.3)

for all \(x,y \in X\), where \(\psi\in\Psi_{b}\). Then f has a fixed point u in X such that \(d(u,u)=0\).

Proof

The proof of Corollary 3.3 follows from Theorem 2.18 by taking \(\alpha(x,y)=1\) for all \(x,y \in X\). Then f is an α-orbital admissible mapping, and both inequalities in (ii) hold for any \(x_{0}\in X\). Notice that since \(\alpha(x,y)=1\), any constructive sequence turns to be regular, and thus X is α-regular. □

Corollary 3.4

Let \((X,d) \) be a complete quasi-b-metric-like space, and let \(f: X \to X\) be a mapping such that

$$ d(fx,fy) \leq\psi\bigl(d(x,y)\bigr) $$
(3.4)

for all \(x,y \in X\), where \(\psi\in\Psi_{b}\). Then f has a fixed point u in X such that \(d(u,u)=0\).

Proof

The proof of Corollary 3.4 follows from Theorem 2.8 by taking \(\alpha(x,y)=1\) for all \(x,y \in X\) and observing that X is α-regular and that (i) and (ii) hold. □

3.2 For standard quasi-b-metric-like spaces with a partial order

In this section, we deduce various fixed point results on a quasi-b-metric-like space endowed with a partial order. We, first, recollect some basic notions and notation.

Definition 3.5

Let \((X,\preceq)\) be a partially ordered set, and \(f: X\to X\) be a given mapping. We say that f is nondecreasing with respect to ⪯ if for all \(x,y\in X\),

$$x\preceq y\quad \Rightarrow \quad fx\preceq fy. $$

Definition 3.6

Let \((X,\preceq)\) be a partially ordered set. A sequence \(\{x_{n}\} \subseteq X\) is said to be nondecreasing (respectively, nonincreasing) with respect to ⪯ if \(x_{n}\preceq x_{n+1}\), \(n\in\mathbb{N}\) (respectively, \(x_{n+1}\preceq x_{n}\), \(n\in\mathbb{N}\)).

Definition 3.7

Let \((X,\preceq)\) be a partially ordered set, and d be a b-metric-like on X. We say that \((X,\preceq,d)\) is regular if for every nondecreasing (respectively, nonincreasing) sequence \(\{x_{n}\} \subseteq X\) such that \(x_{n}\to x\in X\) as \(n\to\infty\), there exists a subsequence \(\{ x_{n_{k}}\}\) of \(\{x_{n}\}\) such that \(x_{n_{k}}\preceq x\) (respectively, \(x\preceq x_{n_{k}}\)) for all k.

We have the following result.

Corollary 3.8

Let \((X,\preceq)\) be a partially ordered set (which does not contain an infinite totally unordered subset), and d be a b-metric-like on X with constant \(s\geq1\) such that \((X,d)\) is complete. Let \(f: X\to X\) be a nondecreasing mapping with respect to ⪯. Suppose that there exists \(\psi\in\Psi_{b}\) such that

$$ d(fx,fy) \leq\psi\bigl(M(x,y)\bigr) $$
(3.5)

for all \(x,y \in X\) with \(x\preceq y\) or \(y\preceq x\), where \(M(x,y)\) is defined as in (2.10). Suppose also that the following conditions hold:

(i):

there exists \(x_{0}\in X\) such that \(x_{0}\preceq fx_{0}\) and \(fx_{0}\preceq x_{0}\);

(ii):

f is continuous or

(ii)′:

\((X,\preceq,d)\) is regular, and d is continuous.

Then f has a fixed point \(u \in X\) with \(d(u,u)=0\). Moreover, if for all \(x,y\in X\), there exists \(z\in X\) such that \(x\preceq z\) and \(y\preceq z\), then f has a unique fixed point.

Proof

Define the mapping \(\alpha: X\times X\to[0,\infty)\) by

$$ \alpha(x,y)=\left \{ \textstyle\begin{array}{l@{\quad}l} 1 &\mbox{if } x\preceq y \mbox{ or } x\succeq y, \\ 0 &\mbox{otherwise}. \end{array}\displaystyle \right . $$

Clearly, f satisfies (2.20), that is,

$$\alpha(x,y)s d(fx, fy)\leq\psi\bigl(M(x,y)\bigr) $$

for all \(x,y\in X\). From condition (i) we have \(x_{0}\preceq fx_{0}\) and \(fx_{0}\preceq x_{0}\). Moreover, for all \(x,y\in X\), from the monotone property of f we have

$$\begin{aligned}& \alpha(x,y)\geq1\quad \Rightarrow \quad x\succeq y\quad \mbox{or} \\& x \preceq y \quad \Rightarrow \quad fx\succeq fy \quad \mbox{or} \\& fx\preceq fy \quad \Rightarrow \quad \alpha(fx,fy)\geq1. \end{aligned}$$

Hence, the self-mapping f is α-admissible. Similarly, we can prove that f is triangular α-admissible and so triangular α-orbital admissible. Now, if f is continuous, then the existence of a fixed point follows from Theorem 2.17.

Suppose that \((X,\preceq,d)\) is regular. Let \(\{x_{n}\}\) be a sequence in X such that \(\alpha(x_{n},x_{n+1})\geq1\) for all n and \(x_{n} \rightarrow x\in X\) as \(n\rightarrow\infty\). By the regularity hypothesis, since X does not contain an infinite totally unordered subset, there exists a subsequence \(\{x_{n_{k}}\}\) of \(\{x_{n}\}\) such that \(x_{n_{k}}\preceq x\) or \(x\preceq x_{n_{k}}\) for all k.

This implies from the definition of α that \(\alpha (x_{n_{k}},x)\geq1\) for all k. In this case, the existence of a fixed point follows again from Theorem 2.18. □

Corollary 3.9

Let \((X,\preceq)\) be a partially ordered set (which does not contain an infinite totally unordered subset), and d be a b-metric-like on X with constant \(s\geq1\) such that \((X,d)\) is complete. Let \(f: X\to X\) be a nondecreasing mapping with respect to ⪯. Suppose that there exists \(\psi\in\Psi_{b}\) such that

$$ d(fx,fy) \leq\psi\bigl(d(x,y)\bigr) $$
(3.6)

for all \(x,y \in X\) with \(x\succeq y\) or \(y\succeq x\). Suppose also that the following conditions hold:

  1. (i)

    there exists \(x_{0}\in X\) such that \(x_{0}\preceq fx_{0}\) and \(fx_{0}\preceq x_{0}\);

  2. (ii)

    f is continuous.

Then T has a fixed point \(u \in X\) with \(d(u,u)=0\). Moreover, if for all \(x,y\in X\), there exists \(z\in X\) such that \(x\preceq z\) and \(y\preceq z\), we have the uniqueness of a fixed point.