1 Introduction

Throughout this paper, we use following notation: \(\mathbb{N}\) is the set of all natural numbers, \(\mathbb{R}\) is the set of all real numbers, and \(\mathbb{R}^{+}\) is the set of all nonnegative real numbers.

Definition 1.1

([2])

A partial metric on a nonempty set X is a function \(p:X\times X\to \mathbb{R}^{+}\) such that, for all \(x,y,z\in X\):

(P1):

\(x=y\) if and only if \(p(x,x)=p(x,y)=p(y,y)\);

(P2):

\(p(x,x)\leq p(x,y)\);

(P3):

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

(P4):

\(p(x,y)\leq p(x,z)+p(z,y)-p(z,z)\).

The pair \((X,p)\) is called a partial metric space. Many fixed point results in partial metric spaces have been proved; see [317]. Recently, Beg and Pathak [1] introduced a weaker form of partial metrics called a weak partial metric.

Definition 1.2

([1])

Let X be a nonempty set. A function \(q:X\times X\rightarrow \mathbb{R}^{+}\) is called a weak partial metric on X if for all \(x,y,z\in X\), the following conditions hold:

\((WP1)\) :

\(q(x,x)=q(x,y)\) if and only if \(x=y\);

\((WP2)\) :

\(q(x, x) \le q(x, y)\);

\((WP3)\) :

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

\((WP4)\) :

\(q(x,y)\leq q(x,z)+q(z,y)\).

The pair \((X, q)\) is called a weak partial metric space.

Examples of weak partial metric spaces [1] are:

  1. (1)

    \((\mathbb{R}^{+},q)\), where \(q:\mathbb{R}^{+}\times \mathbb{R} ^{+}\rightarrow \mathbb{R}^{+}\) is defined as \(q(x,y)=\vert x-y \vert +1\) for \(x,y\in \mathbb{R}^{+}\).

  2. (2)

    \((\mathbb{R}^{+},q)\), where \(q:\mathbb{R}^{+}\times \mathbb{R}^{+}\rightarrow \mathbb{R}^{+}\) is defined as \(q(x,y)=\frac{1}{4}\vert x-y \vert +\max \{x,y\}\) for \(x,y\in \mathbb{R}^{+}\).

  3. (3)

    \((\mathbb{R}^{+},q)\), where \(q:\mathbb{R}^{+}\times \mathbb{R}^{+}\rightarrow \mathbb{R}^{+}\) is defined as \(q(x,y)=\max \{x,y\}+e^{\vert x-y \vert }+1\) for \(x,y\in \mathbb{R} ^{+}\).

Notice that

  • If \(q(x,y)=0\), then \((WP1)\) and \((WP2)\) imply that \(x=y\), but the converse need not be true.

  • \((P1)\) implies \((WP1)\), but the converse need not be true.

  • \((P4)\) implies \((WP4)\), but the converse need not be true.

Example 1.1

([1])

If \(X=\{[a,b]:a,b\in \mathbb{R},a\leq b\}\), then \(q([a,b],[c,d])=\max \{b,d\}-\min \{a,c\}\) is a weak partial metric.

Each weak partial metric q on X generates a \(T_{0}\) topology \(\tau _{q}\) on X. Topology \(\tau _{q}\) has as a base the family of open q-balls \(\{B_{q}(x,\epsilon ):x\in X,\epsilon >0\}\), where \(B_{q}(x,\epsilon )=\{y\in X:q(x,y)< q(x,x)+\epsilon \}\) for all \(x\in X\) and \(\epsilon >0\).

If q is a weak partial metric on X, then the function \(q^{s}:X\times X\to [0,\infty )\) given by \(q^{s}(x,y)=q(x,y)-\frac{1}{2}[q(x,x)+q(y,y)]\) defines a metric on X.

Definition 1.3

Let \((X,q)\) be a weak partial metric space.

  1. (i)

    A sequence \(\{x_{n}\}\) in \((X, q)\) converges to a point \(x\in X\), with respect to \(\tau _{q}\) if \(q(x,x)=\lim_{n\rightarrow \infty }q(x,x_{n})\);

  2. (ii)

    A sequence \(\{x_{n}\}\) in X is said to be a Cauchy sequence if \(\lim_{n, m\rightarrow \infty } q(x_{n}, x_{m})\) exists and is finite;

  3. (iii)

    \((X, q)\) is called complete if every Cauchy sequence \(\{x_{n}\}\) in X converges to \(x\in X\) with respect to topology \(\tau _{q}\).

Clearly, we also have the following:

Lemma 1.1

Let \((X,q)\) be a weak partial metric space. Then

  1. (a)

    A sequence \(\{x_{n}\}\) in X is Cauchy sequence in \((X,q)\) if and only if it is a Cauchy sequence in the metric space \((X, q^{s})\);

  2. (b)

    \((X,q)\) is complete if and only if the metric space \((X,q^{s})\) is complete. Furthermore, a sequence \(\{x_{n}\} \) converges in \((X,q^{s})\) to a point \(x\in X\) if and only if

    $$ \lim_{ n, m \to \infty } q(x_{n}, x_{m}) = \lim_{n \to \infty } q(x_{n}, x) = q(x, x). $$
    (1.1)

Let \((X,q)\) be a weak partial metric space. Let \(CB^{q}(X)\) be the family of all nonempty closed bounded subsets of \((X,q)\). Here, the boundedness is given as follows: E is a bounded subset in \((X,q)\) if there exist \(x_{0}\in X\) and \(M\geq 0\) such that, for all \(a\in E\), we have \(a\in B_{q}(x_{0},M)\), that is, \(q(x_{0},a)< q(a,a)+M\).

For \(E,F\in \mathit{CB}^{q}(X)\) and \(x\in X\), define

$$ q(x,E)=\inf \bigl\{ q(x,a),a\in E \bigr\} ,\qquad \delta _{q}(E,F)=\sup \bigl\{ q(a,F):a\in E \bigr\} $$

and

$$ \delta _{q}(F,E)=\sup \bigl\{ q(b,E):b\in F \bigr\} . $$

Now, \(q(x,E)=0\) implies \(q^{s}(x,E)=0\), where \(q^{s}(x,E)=\inf \{q^{s}(x,a),a\in E\}\).

Remark 1.1

([1])

Let \((X,q)\) be a weak partial metric space, and let E be a nonempty set in (\(X,q\)). Then

$$ a\in \overline{E} \quad \mbox{if and only if}\quad q(a,E)=q(a,a), $$
(1.2)

where denotes the closure of E with respect to the weak partial metric q.

Note that E is closed in \((X,q)\) if and only if \(E=\overline{E}\).

First, we study properties of the mapping \(\delta _{q}:\mathit{CB}^{q}(X)\times \mathit{CB}^{q}(X)\rightarrow[ 0,\infty )\).

Proposition 1.1

([1])

Let (\(X, q\)) be a weak partial metric space,We have the following:

  1. (i)

    \(\delta _{q}(E,E)=\sup \{q(a,a):a\in E\}\);

  2. (ii)

    \(\delta _{q}(E,E)\leq \delta _{q}(E,F)\);

  3. (iii)

    \(\delta _{q}(E,F)=0\) implies \(E\subseteq F\);

  4. (iv)

    \(\delta _{q}(E,F)\leq \delta _{q}(E,H)+\delta _{q}(H,F)\) for all \(E ,F, H\in \mathit{CB}^{q}(X)\).

Definition 1.4

([1])

Let \((X,q)\) be a weak partial metric space. For \(E,F\in \mathit{CB}^{q}(X)\), define

$$ \mathcal{H}_{q}^{+}(E,F)=\frac{1}{2} \bigl\{ \delta _{q}(E,F)+\delta _{q}(F,E) \bigr\} . $$
(1.3)

The following proposition is a consequence of Proposition 1.1.

Proposition 1.2

([1])

Let \((X, q)\) be a weak partial metric space. Then, for all \(E, F, H\in \mathit{CB}^{q}(X)\), we have

  1. (wh1)

    \(\mathcal{H}_{q}^{+}(E,E)\leq \mathcal{H}_{q}^{+}(E,F)\);

  2. (wh2)

    \(\mathcal{H}_{q}^{+}(E,F)=\mathcal{H}_{q}^{+}(F,E)\);

  3. (wh3)

    \(\mathcal{H}_{q}^{+}(E,F)\leq \mathcal{H}_{q}^{+}(E,H)+\mathcal{H}_{q}^{+}(H,F)\).

The mapping \(\mathcal{H}_{q}^{+}:\mathit{CB}^{q}(X)\times \mathit{CB}^{q}(X)\rightarrow[ 0,+\infty )\), is called the \(\mathcal{H}^{+}\)-type Pompeiu–Hausdorff metric induced by q.

Definition 1.5

([1])

Let \((X,q)\) be a complete weak partial metric space. A multivalued map \(T:X\rightarrow \mathit{CB}^{q}(X)\) is called an \(\mathcal{H}_{q}^{+}\)-contraction if

\((1^{\circ })\) :

there exists k in \((0, 1)\) such that

$$ \mathcal{H}_{q}^{+} \bigl(Tx\setminus \{x\}, Ty\setminus \{y \} \bigr) \le k q(x, y) \quad \mbox{for every } x, y \in X, $$
(1.4)
\((2^{\circ })\) :

for all x in \(X, y\) in Tx, and \(\epsilon > 0\), there exists z in Ty such that

$$ q(y, z) \leq \mathcal{H}_{q}^{+}(Ty, Tx) + \epsilon . $$
(1.5)

Beg and Pathak [1] proved the following fixed point theorem.

Theorem 1.1

([1])

Let \((X,q)\) be a complete weak partial metric space. Every \(\mathcal{H}_{q}^{+}\)-type multivalued contraction mapping \(T:X\rightarrow \mathit{CB}^{q}(X)\) with Lipschitz constant \(k<1\) has a fixed point.

In this paper, we generalize the concept of \(\mathcal{H}_{q}^{+}\)-type multivalued contractions by introducing \(\mathcal{H}_{q}^{+}\)-type Suzuki mult-valued contraction mappings.

2 Fixed point results

First, let \(\psi : [0, 1)\rightarrow (0, 1]\) be the nonincreasing function

$$ \psi (r)= \textstyle\begin{cases} 1& \mbox{if } 0\leq r < \frac{1}{2}, \\ 1-r& \mbox{if } \frac{1}{2}\leq r < 1. \end{cases} $$
(2.1)

Now, we state a fixed point result for \(\mathcal{H}_{q}^{+}\)-type Suzuki multivalued contraction mappings.

Theorem 2.1

Let \((X,q)\) be a complete weak partial metric space, and let \(F : X \to \mathcal{CB}^{q}(X)\) be a multivalued mapping. Let \(\psi : [0, 1)\rightarrow (0, 1]\) be the nonincreasing function defined by (2.1). Suppose that there exists \(0 \leq s < 1\) such that T satisfies the condition

$$ \psi (s)q(x,Fx)\leq q(x,y) \quad \textit{implies } \mathcal{H}_{q}^{+} \bigl(Fx\setminus \{x\},Fy\setminus \{y\} \bigr)\leq s q(x,y) $$
(2.2)

for all \(x,y \in X\). Suppose also that, for all x in \(X, y\) in Fx, and \(t > 1\), there exists z in Fy such that

$$ q(y, z) \le t \mathcal{H}_{q}^{+}(Fy, Fx). $$
(2.3)

Then F has a fixed point.

Proof

Let \(s_{1}\in (0, 1)\) be such that \(0\leq s \leq s_{1} < 1\) and \(w_{0}\in X\). Since \(Fw_{0}\) is nonempty, it follows that if \(w_{0} \in Fw_{0}\), then the proof is completed. Let \(w_{0}\notin Fw_{0}\). Then there exists \(w_{1}\in Fw_{0}\) such that \(w_{1}\neq w_{0}\).

Similarly, there exists \(w_{2}\in Fw_{1}\) such that \(w_{1}\neq w_{2}\), and from (2.3) we have

$$ q(w_{1},w_{2})\leq \frac{1}{\sqrt{s_{1}}}H^{+}_{q}(Fw_{0},Fw_{1}). $$
(2.4)

Since

$$ \psi (s)q(w_{1},Fw_{1})\leq q(w_{1},Fw_{1}) \leq q(w_{1},w_{2}), $$

from (2.2) and (2.4) we get

$$\begin{aligned} q(w_{1},w_{2})&\leq \frac{1}{\sqrt{s_{1}}}H^{+}_{q}(Fw_{0},Fw_{1})\leq \frac{1}{\sqrt{s_{1}}}H^{+}_{q} \bigl(Fw_{0} \setminus \{w_{0}\},Fw_{1}\setminus \{w_{1}\} \bigr) \\ & \leq \frac{1}{\sqrt{s_{1}}}. s .q(w_{0},w_{1})< \sqrt{s_{1}}. q(w_{0},w_{1}). \end{aligned}$$

By repeating this process n times we obtain

$$ q(w_{n},w_{n+1})\leq (\sqrt{s_{1}})^{n} \cdot q(w_{0},w_{1}). $$
(2.5)

Hence

$$ \lim_{n\to \infty }q(w_{n},w_{n+1})=0. $$
(2.6)

Now we prove that \(\{w_{n}\}\) is a Cauchy sequence in \((X,q^{s})\). For all \(m\in N\), we have

$$\begin{aligned} q^{s}(w_{n},w_{n+m})&= q(w_{n},w_{n+m})- \frac{1}{2} \bigl[q(w_{n},w_{n})+q(w_{n+m},w_{n+m}) \bigr] \\ &\leq q(w_{n},w_{n+m}) \\ &\leq q(w_{n},w_{n+1})+q(w_{n+1},w_{n+2})+ \cdots+q(w_{n+m-1},w_{n+m}) \\ & \leq \bigl[(\sqrt{s_{1}})^{n}+(\sqrt{s_{1}})^{n+1}+ \cdots+(\sqrt{s_{1}})^{n+m-1} \bigr]q(w_{0},w_{1}) \\ &\leq (\sqrt{s_{1}})^{n}\frac{1}{1-\sqrt{s_{1}}}q(w_{0},w_{1}). \end{aligned}$$

Hence

$$ \lim_{n\to \infty }q^{s}(w_{n},w_{n+m})=0. $$
(2.7)

This implies that \(\{w_{n}\}\) is a Cauchy sequence in the complete metric space \((X,q^{s})\). It follows that there exists \(u\in X \) such that

$$ \lim_{n\to \infty }q(w_{n},u)=\lim _{n,m\to \infty }q(w_{n},w_{m})=q(u,u). $$
(2.8)

From \((WP2)\) we obtain

$$ \frac{1}{2} \bigl[q(w_{n},w_{n})+q(w_{n+1},w_{n+1}) \bigr]\leq q(w_{n},w_{n+1}). $$
(2.9)

By taking the limit as \(n\to \infty \) from (2.6) we get

$$ \lim_{n\to \infty }q(w_{n},w_{n})= \lim_{n\to \infty }q(w_{n+1},w_{n+1})=\lim _{n\to \infty }q(w_{n},w_{n+1})=0. $$
(2.10)

Also, from (2.7) and (2.10) we find

$$ \lim_{n\to \infty }q^{s}(w_{n},w_{n+m})=0= \lim_{n\to \infty }q(w_{n},w_{n+m})-\frac{1}{2} \lim_{n\to \infty } \bigl[q(w_{n},w_{n})+q(w_{n+m},w_{n+m}) \bigr]. $$
(2.11)

Therefore

$$ \lim_{n\to \infty }q(w_{n},w_{n+m})=0= \lim_{n\to\infty }q(w_{n},u)=q(u,u). $$
(2.12)

Now, we prove that

$$ q(u,Fx)\leq 2s q(u,x)\quad \mbox{for all } x\in X\setminus \{u\}. $$
(2.13)

Since \(\lim_{n\to \infty } q(w_{n},u)=0\), there exists \(n_{0}\in \mathbb{N}\) such that

$$ q(w_{n},u)\leq \frac{1}{3}q(x,u)\quad \mbox{for all } n\geq n_{0}. $$

Then

$$\begin{aligned} \psi (s)q(w_{n},Fw_{n}) &\leq q(w_{n},Fw_{n}) \\ &\leq q(w_{n},w_{n+1}) \\ &\leq q(w_{n},u)+q(u,w_{n+1}) \\ &\leq \frac{1}{3}q(u,x)+ \frac{1}{3}q(u,x) \\ &\leq q(u,x)-\frac{1}{3}q(u,x) \\ &\leq q(u,x)-q(u,w_{n})\leq q(x,w_{n}). \end{aligned}$$

This implies that

$$ H^{+}_{q}(Fw_{n},Fx)\leq s q(w_{n},x). $$

Since \(w_{n+1}\in Fw_{n}\), we have

$$\begin{aligned} q(w_{n+1},Fx)&\leq \delta _{q}(Fw_{n},Fx) \\ &\leq 2H^{+}_{q}(Fw_{n},Fx) \\ &\leq 2 s q(w_{n},x) \\ &\leq 2s \bigl[q(w_{n},u)+q(u,x) \bigr]. \end{aligned}$$

By taking the limit as \(n\to \infty \) we get

$$ \lim_{n\to \infty }q(w_{n+1},Fx)\leq 2s q(u,x). $$
(2.14)

Also, since

$$ q(u,Fx)\leq q(u,w_{n+1})+q(w_{n+1},Fx) $$

and

$$ q(w_{n+1},Fx)\leq q(w_{n+1},w_{n})+q(w_{n},u)+q(u,Fx), $$

we have

$$ \lim_{n\to \infty }q(w_{n+1},Fx)= q(u,Fx). $$
(2.15)

From (2.14) and (2.15) we find that

$$ q(u,Fx)\leq 2s q(u,x)\quad \mbox{for all } x\in X\setminus \{u \}. $$
(2.16)

We claim that

$$ H^{+}_{q}(Fx,Fu)\leq s q(u,u) \quad \mbox{for all } x\in X. $$

If \(x=u\), then at that point, this clearly holds. So, let \(x\neq u\). Then for every positive integer \(n\in \mathbb{N}\), there exists \(y_{n} \in Fx \) such that

$$ q(u,y_{n})\leq q(u,Fx)+\frac{1}{n} q(u,x). $$

Therefore

$$\begin{aligned} q(x,Fx)&\leq q(x,y_{n}) \\ &\leq q(x,u)+q(u,y_{n}) \\ &\leq q(x,u)+q(u,Fx)+\frac{1}{n} q(x,u). \end{aligned}$$
(2.17)

From (2.16) and (2.17) we get

$$\begin{aligned} q(x,Fx)&\leq q(u,x)+2s q(u,x)+\frac{1}{n} q(x,u) \end{aligned}$$
(2.18)
$$\begin{aligned} &= \biggl[1+2s+\frac{1}{n} \biggr]q(x,u). \end{aligned}$$
(2.19)

Hence

$$\begin{aligned} \frac{1}{1+2s+\frac{1}{n}} q(x,Fx)\leq q(u,x). \end{aligned}$$

This implies that

$$\begin{aligned} H^{+}_{q}(Fu,Fx)\leq s q(u,x). \end{aligned}$$

Finally, we show that \(u\in Fu\). For this,

$$\begin{aligned} q(u,Fu)&=\lim_{n\to \infty }q(w_{n+1},Fu) \\ &\leq \lim _{n\to \infty }\delta _{q}(Fw_{n},Fu) \\ &\leq 2\lim_{n\to \infty }H^{+}_{q}(Fw_{n},Fu) \\ &\leq 2 s\lim_{n\to \infty } q(w_{n},u)=0. \end{aligned}$$

We deduce that \(q(u,u)=q(u,Fu)=0\). Since Fu is closed, \(u\in \overline{Fu}=Fu\). □

We provide the following example.

Example 2.1

Let \(X=\{0,\frac{1}{2},1\}\) and define a weak partial metric \(q:X\times X \to [0,\infty ) \) as follows: \(q(0,0)=0\), \(q(\frac{1}{2}, \frac{1}{2})=\frac{1}{3}\), \(q(1,1)=\frac{1}{4}\), \(q(0,\frac{1}{2})=q(\frac{1}{2},0)=\frac{1}{2}\), \(q(\frac{1}{2},1)=q(1,\frac{1}{2})=\frac{3}{4}\), and \(q(1,0)=q(0,1)=1\). It is clear that \((X,q)\) is a weak partial metric space. Note that

$$ q(1,0)=1\nleq q \biggl(1,\frac{1}{2} \biggr)+q \biggl(\frac{1}{2},0 \biggr)-q \biggl(\frac{1}{2},\frac{1}{2} \biggr)=\frac{3}{4}+ \frac{1}{2}-\frac{1}{3}. $$

Then \((X,q)\) is not a partial metric space. Define the mapping \(F:X\to \mathit{CB}^{q}(X)\) by \(F(0)=F(\frac{1}{2})=\{0\}\) and \(F(1)=\{0,\frac{1}{3}\}\). Choose \(s=0.5\). From the definition of ψ we have \(\psi (s)=1\).

To prove the contraction condition (2.2), we need the following cases:

Case 1. At \(x=0\), we have

$$ \psi (s)q \bigl(0,F(0) \bigr)=q(0,0)=0\leq q(0,y) \quad \mbox{for all } x\in X. $$

For \(y=0\), we have

$$ H^{+}_{q} \bigl(F(0)\setminus \{0\},F(0)\setminus \{0\}\bigr)=H^{+}_{q}(\phi ,\phi )=0\leq s q(0,0). $$

For \(y=\frac{1}{2}\), we get

$$ H^{+}_{q} \biggl(F(0)\setminus \{0\},F \biggl(\frac{1}{2} \Bigm\backslash \biggl\{ \frac{1}{2} \biggr\} \biggr)=H^{+}_{q} \bigl(\phi ,\{0\} \bigr)=0\leq s q \biggl(0, \frac{1}{2} \biggr). $$

If f \(y=1\), then

$$ H^{+}_{q} \bigl(F(0)\setminus \{0\},F(1)\setminus \{1\} \bigr)=H^{+}_{q} \biggl(\phi , \biggl\{ 0,\frac{1}{2} \biggr\} \biggr)=0\leq s q(0,1). $$

Case 2. At \(x=\frac{1}{2}\), we have

$$ \psi (s)q \biggl(\frac{1}{2},F \biggl(\frac{1}{2} \biggr) \biggr)=q \biggl(\frac{1}{2},0 \biggr)=\frac{1}{2}\leq q \biggl( \frac{1}{2},y \biggr)\quad\mbox{for all } y\in X\Bigm\backslash \biggl\{ \frac{1}{2} \biggr\} . $$

Similarly, if \(y=0, then\)

$$ H^{+}_{q} \biggl(F \biggl(\frac{1}{2} \Bigm\backslash \biggl\{ \frac{1}{2} \biggr\} ,F(0)\setminus \{0\} \biggr)=H^{+}_{q} \bigl(\{0\},\phi \bigr)=0\leq s q \biggl(\frac{1}{2},0 \biggr), $$

If \(y=1\), then

$$ H^{+}_{q} \biggl(F \biggl(\frac{1}{2} \biggr)\Bigm\backslash \biggl\{ \frac{1}{2} \biggr\} ,F(1)\setminus \{1\} \biggr)=H^{+}_{q}\biggl(\{0\}, \biggl\{ 0,\frac{1}{2} \biggr\} \biggr)=\frac{1}{4}< s q \biggl(\frac{1}{2},1 \biggr)=\frac{3}{8}. $$

Case 3. At \(x=1\), we have

$$ \psi (s)q \bigl(1,F(1) \bigr)=q \biggl(1,\frac{1}{2} \biggr)= \frac{3}{4} \leq q(1,y)\quad\mbox{for all } y\in X\setminus \{1\}. $$

Again, if \(y=0\), then

$$ H^{+}_{q} \bigl(F(1)\setminus \{1\},F(0)\setminus \{0\} \bigr)=H^{+}_{q} \biggl( \biggl\{ 0,\frac{1}{2} \biggr\} , \phi \biggr)=0\leq s q(1,0). $$

If \(y=\frac{1}{2}\), then

$$ H^{+}_{q} \biggl(F(1)\setminus \{1\},F \biggl( \frac{1}{2} \Bigm\backslash \biggl\{ \frac{1}{2} \biggr\} \biggr)=H^{+}_{q} \biggl( \biggl\{ 0,\frac{1}{2} \biggr\} ,\{0\} \biggr)=\frac{1}{4}< s q \biggl(1,\frac{1}{2} \biggr)= \frac{3}{8}. $$

Finally, we will enquire the condition (2.3) with \(t=2\). For this, we discuss the following situations:

  1. (i)

    If \(x=0\) or \(x=\frac{1}{2}\), then \(y\in F(0)=F(\frac{1}{2})=\{0\}\). This yields that \(y=0\), so there exists \(z\in F(y)\) such that

    $$ 0=q(y,z)\leq 2 H^{+}_{q} \bigl(F(x),F(y) \bigr). $$
  2. (ii)

    If \(x=1\), then \(y\in F(1)=\{0,\frac{1}{2}\}\). If \(y=0\), then \(z=0\), and condition (2.3) is satisfied.

    Also, If \(y=\frac{1}{2}\), then \(z=0\), so that

    $$ \frac{1}{2}=q(y,z)= 2 H^{+}_{q} \biggl(F(1),F \biggl( \frac{1}{2} \biggr) \biggr)=\frac{1}{2}. $$

    Therefore all conditions of Theorem 2.1 are satisfied, and the function F has a fixed point \(u=0\).

On the other hand, the result of Beg and Pathak [1] is not applicable. Indeed,

$$ H^{+}_{q} \bigl(F(1)\setminus \{1\},F(1)\setminus \{1\} \bigr)= \frac{1}{3}>\frac{1}{2} q(1,1)=\frac{1}{8}. $$

3 Applications

First, we present an application concerning a homotopy result for complete weak partial metric spaces.

Theorem 3.1

Let \((X, q)\) be a complete weak partial metric space, let D be an open subset of X, and let W be a closed subset of X with \(D\subset W\). Let \(F : W \times [0, 1] \to \mathit{CB}^{q} (X)\) be an operator satisfying:

  1. (i)

    \(x\notin F (x, t)\) for each \(x \in W \setminus D\) and each \(t\in [0, 1]\);

  2. (ii)

    there exists \(s \in (0, \frac{1}{2})\) such that, for each \(t \in [0, 1]\) and each \(x, y \in W\), we have

    $$ \psi (s) q \bigl(x,F(x,t) \bigr)\leq q(x,y)\Rightarrow H^{+}_{q} \bigl( F (x, t) \setminus \{x\}, F (y, t) \setminus \{y\} \bigr) \leq s q(x, y); $$
  3. (iii)

    for all \(x \in W\), \(y \in F (x, t)\), and \(h > 1\), there exists \(z \in F (y, t)\) such that

    $$ q(y, z) \leq h H^{+}_{q} \bigl(F (y, t), F (x, t) \bigr); $$
  4. (iv)

    there exists a continuous function \(\eta : [0, 1] \to \mathbb{R}\) such that

    $$ H^{+}_{q} \bigl( F (x, t_{1}) \setminus \{x\}, F (x, t_{2}) \setminus \{x\} \bigr) \leq s\bigl\vert \eta (t_{1}) -\eta (t_{2}) \bigr\vert $$

    for all \(t_{1}, t_{2} \in [0, 1]\) and \(x \in W\);

  5. (v)

    if \(x \in F (x, t)\), then \(F (x, t) = \{x\}\). Then \(F (\cdot , 0)\) has a fixed point if and only if \(F (\cdot , 1)\) has a fixed point.

Proof

Define the set

$$ \Delta := \bigl\{ t\in [0,1] ;x\in F(x, t) \mbox{ for some } x\in D \bigr\} . $$

Since \(F (\cdot , 0)\) has a fixed point, from condition (i), we get \(0\in \Delta \), so \(\Delta \neq\phi \). First, we want to show that Δ is an open set. Let \(t_{1}\in \Delta \) and \(x_{1}\in D\) be such that \(x_{1}\in F(x_{1},t_{1} )\). Since D is open in \((X,q)\), there exists \(r>0\) such that \(B(x_{1},r)\subset D\). Consider \(\epsilon =(\frac{1-2s}{2})(q(x_{1},x_{1})+r)>0\). Since η is continuous at \(t_{1}\), there exists \(\delta (\epsilon )>0\) such that \(\vert \eta (t)-\eta (t_{1}) \vert < \epsilon \) for all \(t\in (t_{1}-\delta (\epsilon ),t_{1}+\delta (\epsilon ))\).

Let \(t\in (t_{1}-\delta (\epsilon ),t_{1}+\delta (\epsilon ))\) and \(x\in B(x_{1},r)=\{x\in X ;q(x_{1},x)\leq q(x_{1},x_{1})+r\}\). Since \(x_{1}\in F(x_{1},t_{1} )\), from \((WP2)\) we have

$$ \psi (s)q \bigl(x_{1},F(x_{1},t_{1} ) \bigr) \leq q(x_{1},x_{1})\leq q(x_{1},x)\quad\mbox{for all } x\in X. $$

Thus

$$\begin{aligned} q \bigl(x_{1},F(x,t) \bigr)&\leq 2H^{+}_{q} \bigl(F(x,t),F(x_{1},t_{1}) \bigr) \\ &\leq 2 \bigl[H^{+}_{q} \bigl(F(x,t),F(x,t_{1}) \bigr)+H^{+}_{q} \bigl(F(x,t_{1}),F(x_{1},t_{1}) \bigr) \bigr] \\ &=2 \bigl[H^{+}_{q} \bigl(F(x,t)\setminus \{x \},F(x,t_{1})\setminus \{x\} \bigr) +H^{+}_{q} \bigl(F(x,t_{1})\setminus \{x \},F(x_{1},t_{1})\setminus \{x_{1}\} \bigr) \bigr] \\ &\leq 2 \bigl[\bigl\vert \eta (t) - \eta (t_{1}) \bigr\vert +s q(x,x_{1}) \bigr] \\ &\leq 2 \bigl[\epsilon +s \bigl(q(x_{1},x_{1})+r \bigr) \bigr] \\ &\leq 2 \biggl[ \biggl(\frac{1-2s}{2} \biggr) \bigl(q(x_{1},x_{1})+r \bigr)+s \bigl(q(x_{1},x_{1})+r \bigr) \biggr] \\ &\leq q(x_{1},x_{1})+r. \end{aligned}$$

Therefore \(F(x,t)\subset B(x_{1},r)\). Since \(F(\cdot ,t):B(x_{1},r)\to \mathit{CB}^{q}(X)\) for each fixed \(t\in (t_{1}-\delta (\epsilon ),t-1+\delta (\epsilon ))\) and (ii) holds, all the hypotheses of Theorem 2.1 are satisfied. We conclude that \(F(\cdot ,t)\) has a fixed point in \(B(x_{1},r)\subset W\). This fixed point must be in D due to (i). Hence \((t_{1}-\delta (\epsilon ),t-1+\delta (\epsilon ))\subset \Delta \), and therefore Δ is open in \([0,1]\).

Second, we prove that Δ is closed in \([0,1]\). To show this, choose a sequence \(\{t_{n}\}\) in Δ such that \(t_{n}\to t^{*}\in [0,1]\) as \(n\to \infty \). We must show that \(t^{*} \in \Delta \). By the definition of Δ there exists \(x_{n} \in D\) with \(x_{n} \in F(x_{n},t_{n})\). Then

$$ \psi (s)q \bigl(x_{n},F(x_{n},t_{n} ) \bigr)\leq q(x_{n},x_{n})\leq q(x_{n},x)\quad\mbox{for all } x \in X. $$

This implies that, for all positive integers \(m, n\in \mathbb{N}\), using (v) and \((Wh3)\), we have

$$\begin{aligned} q(x_{n},x_{m}) &\leq 2H^{+}_{q} \bigl(F(x_{n},t_{n}), F (x_{m},t_{m}) \bigr) \\ & \leq 2H^{+}_{q} \bigl(F(x_{n},t_{n}), F (x_{n},t_{m}) \bigr)+2H^{+}_{q} \bigl(F (x_{n},t_{m}), F (x_{m},t_{m}) \bigr) \\ & =2H^{+}_{q} \bigl(F(x_{n},t_{n}) \setminus \{x_{n}\},F(x_{n},t_{m})\setminus \{x_{n}\} \bigr) \\ &\quad {} +2H^{+}_{q} \bigl(F(x_{n},t_{m}) \setminus \{x_{n}\},F(x_{m},t_{m})\setminus \{x_{m}\} \bigr) \\ & \leq 2s\bigl\vert \eta (t_{n})-\eta (t_{m}) \bigr\vert +2s q(x_{n},x_{m}). \end{aligned}$$

This implies that

$$ q(x_{n},x_{m})\leq \frac{2s}{1-2s} \bigl(\bigl\vert \eta(t_{n})-\eta (t_{m}) \bigr\vert \bigr). $$

Hence \(\lim_{n,m\to \infty }q(x_{n},x_{m})=0\). Therefore \(\{x_{n}\}\) is a Cauchy sequence in \((X, q)\). Since \((X, q)\) is complete, there exists \(x^{*}\in W\) such that

$$ q \bigl(x^{*} ,x^{*} \bigr) = \lim_{n\to \infty }q \bigl(x^{*},x_{n} \bigr)=\lim_{n,m\to \infty } q(x_{n},x_{m})=0. $$

On the other hand, we have

$$\begin{aligned} q \bigl(x_{n},F \bigl(x^{*} ,t^{*} \bigr) \bigr)& \leq 2H^{+}_{q} \bigl(F (x_{n},t_{n}), F \bigl(x^{*} ,t^{*} \bigr) \bigr) \\ & \leq 2H^{+}_{q} \bigl(F (x_{n},t_{n}), F \bigl(x_{n},t^{*} \bigr)+2H^{+}_{q} \bigl(F \bigl(x_{n},t^{*} \bigr) \bigr), F \bigl(x^{*} ,t^{*} \bigr) \bigr) \\ &=2 H^{+}_{q} \bigl( F(x_{n},t_{n}) \setminus \{x_{n}\},F \bigl(x_{n},t^{*} \bigr) \setminus \{x_{n}\} \bigr) \\ &\quad {} +2H^{+}_{q} \bigl( F \bigl(x_{n},t^{*} \bigr)\setminus \{x_{n}\},F \bigl(x^{*} ,t^{*} \bigr)\setminus \bigl\{ x^{*} \bigr\} \bigr) \\ & \leq 2s\bigl\vert \eta (t_{n})-\eta \bigl(t^{*} \bigr) \bigr\vert +2sq \bigl(x_{n},x^{*} \bigr). \end{aligned}$$

Taking the limit as \(n\to \infty \) in the above inequality, we get

$$ q \bigl(x^{*} ,F \bigl(x^{*} ,t^{*} \bigr) \bigr)= \lim_{n\to \infty } q \bigl(x_{n},F \bigl(x^{*} ,t^{*} \bigr) \bigr)=0. $$

It follows that \(x^{*}\in F(x^{*} ,t^{*} )\). Thus \(t^{*}\in \Delta \), and hence Δ is closed in \([0,1]\). By the connectedness of \([0,1]\) we have \(\Delta =[0,1]\).

The reverse implication easily follows by applying the same strategy. This completes the proof. □

Now, we give another application to the solvability of integral inclusions of Fredholm type. Let \(I=[0,1]\), and let \(C(I,\mathbb{R})\) be the space of all continuous functions \(f :I\to R\). Consider the weak partial metric on X given by

$$ q(x,y)=\sup_{t\in I} \bigl\vert x(t)-y(t) \bigr\vert +\alpha $$

for all \(x,y\in C(I,R)\) and \(\alpha >0\). We have \(q^{s}(x,y)=\sup_{t\in I} \vert x(t)-y(t) \vert \), so by Lemma 1.1 \((C(I,\mathbb{R}),q)\) is a complete weak partial metric space. Denote by \(P_{cv}(\mathbb{R})\) the family of all nonempty compact and convex subsets of \(\mathbb{R}\) and by \(P_{cl}(\mathbb{R})\) the family of all nonempty closed subsets of \(\mathbb{R}\).

Theorem 3.2

Consider the integral inclusion of Fredholm type

$$ h(t)\in f(t)+ \int ^{1}_{0} K \bigl(t,u,h(u) \bigr)\,du, \quad t\in [0,1]. $$
(3.1)

Suppose that:

  1. (i)

    \(K:I\times I\times R\to P_{cv}(\mathbb{R})\) is such that \(K_{h}(t,u):=K(t,u,h(u))\) is a lower semicontinuous for all \((t,u)\in I\times I\) and \(h\in C(I,\mathbb{R})\),

  2. (ii)

    \(f \in C(I,R)\);

  3. (iii)

    for each \(t\in I\), there exists \(l(t,\cdot )\in L^{1}(I)\) such that \(\sup_{ t\in I} \int ^{1}_{0} l(t,u)\,du = \frac{s}{2}\) with \(s\in [0,1)\) and

    $$ H^{+}_{q} \bigl( K \bigl(t,u,h(u) \bigr),K \bigl(t,u,r(u) \bigr) \bigr) \leq l(t,u) \Bigl(\sup_{u\in I}\bigl\vert h(u)-r(u) \bigr\vert +\alpha \Bigr) $$

    for all \(t,u\in I\) and all \(h,r\in C(I,\mathbb{R})\).

Then the integral inclusion (3.1) has at least one solution in \(C(I,\mathbb{R})\).

Proof

Consider the multivalued operator \(T:C(I,R)\to P_{CL}(C(I,R))\) defined by

$$ Tx(t)= \biggl\{ h\in C(I,\mathbb{R}) \mbox{ such that } h(t)\in f(t)+ \int ^{1}_{0} K \bigl(t,u,x(u) \bigr)\,du,t \in I \biggr\} $$

for \(x\in C(I,\mathbb{R})\). For each \(K_{x}(t,u):I\times I \to P_{cv}(\mathbb{R})\), by the Michael selection theorem there exists a continuous operator \(k_{x}: I\times I\to \mathbb{R}\) such that \(k_{x}(t,u)\in K_{x}(t,u)\) for all \(t,u\in I\). This implies that \(f(t)+ \int ^{1}_{0} k_{x}(t,u)\,du\in Tx \), and so \(Tx\neq \emptyset \). It is easy to prove that Tx is closed, and so we omit the details (see also [18]). This implies that Tx is closed in \((C(I,\mathbb{R}),q)\).

Now, we will show that T is \(H^{+}_{q}\)-type Suzuki multivalued contraction mapping. Let \(x_{1}, x_{2}\in C(I,\mathbb{R})\) and \(h\in Tx\). Then there exists \(k_{x_{1}}(t,u)\in K_{x_{1}}(t,u)\) with \(t,u\in I\) such that \(h(t)=f(t)+ \int ^{1}_{0} k_{x}(t,u)\,du,t\in I\). Also, by hypothesis (iii),

$$ H^{+}_{q} \bigl( K \bigl(t,u,x_{1}(u) \bigr),K \bigl(t,u,x_{2}(u) \bigr) \bigr) \leq l(t,u) \Bigl(\sup _{u\in I}\bigl\vert x_{1}(u)-x_{2}(u) \bigr\vert +\alpha \Bigr)\quad \forall t,u\in I. $$

Then there exists \(z(t,u)\in K_{x_{2}}(t,u)\) such that

$$ \bigl\vert k_{x_{1}}(t,u)-z(t,u) \bigr\vert +n \leq l(t,u) \bigl[ \bigl\vert x_{1}(u)-x_{2}(u) \bigr\vert + \alpha \bigr] $$

for all \(t,u\in I\). Now, we define the multivalued operator \(M(t,u)\) by

$$ M(t,u)=K_{x_{2}}(t,u)\cap \bigl\{ m\in \mathbb{R}, \bigl\vert k_{x_{1}}(t,u)-m \bigr\vert + \alpha \leq l(t,u) \bigl(\bigl\vert x_{1}(u)-x_{2}(u) \bigr\vert +\alpha \bigr) \bigr\} $$

for \(t,u\in I\). Since M is a lower semicontinuous operator, there exists a continuous operator \(k_{x_{2}} : I\times I \to \mathbb{R}\) such that \(k_{x_{2}}(t,u)\in M(t,u)\) for all \(t,u\in I\) and

$$ w(t)=f(t)+ \int ^{1}_{0} k_{x_{2}}(t,u)\,du\in f(t)+ \int ^{1}_{0} K \bigl(t,u,x_{2}(u) \bigr)\,du. $$

Therefore

$$\begin{aligned} q \bigl(h(t),Tx_{2}(t) \bigr)&\leq q \bigl(h(t),w(t) \bigr) \\ &=\sup _{t\in I}\bigl\vert h(t)-w(t) \bigr\vert +\alpha \\ & =\sup_{t\in I}\biggl\vert \int ^{1}_{0} \bigl[k_{x_{1}}(t,u)-k_{x_{2}}(t,u) \bigr]\,du \biggr\vert +\alpha \\ & \leq \sup_{t\in I}\int ^{1}_{0} \bigl(\bigl\vert k_{x_{1}}(t,u)-k_{x_{2}}(t,u) \bigr\vert +\alpha -\alpha \bigr)\,du+\alpha \\ & \leq \sup_{t\in I}\int ^{1}_{0}l(t,u) \bigl[\bigl\vert x_{1}(u)-x_{2}(u) \bigr\vert + \alpha \bigr]\,du- \int ^{1}_{0}\alpha \,du+\alpha \\ & = \Bigl(\sup_{t\in I}\bigl\vert x_{1}(u)-x_{2}(u)\bigr\vert + \alpha \Bigr)\int ^{1}_{0}l(t,u)\,du \\ & \leq s q \bigl(x_{1}(t),x_{2}(t) \bigr). \end{aligned}$$

Since \(h(t)\in Tx_{1}\) is arbitrary, we have

$$\begin{aligned} \delta _{q}(Tx_{1},Tx_{2})\leq s q(x_{1},x_{2}). \end{aligned}$$
(3.2)

Similarly, we can get

$$\begin{aligned} \delta _{q}(Tx_{2},Tx_{1})\leq s q(x_{1},x_{2}). \end{aligned}$$
(3.3)

From (3.2) and (3.3) we have

$$ H^{+}_{q}(Tx_{1},Tx_{2})= \frac{\delta _{q}(Tx_{1},Tx_{2})+\delta _{q}(Tx_{2},Tx_{1})}{2}\leq s q(x_{1},x_{2}). $$

In particular, the previous inequality holds for any \(t\in I\), so that

$$ \psi (s)q(x_{1},Tx_{1})\leq q(x_{1},x_{2}). $$

Thus all conditions of Theorem 2.1 are satisfied, and hence a solution of (3.1) exists. □

4 Perspectives

In 2010, Romaguera [19] introduced the notions of 0-Cauchy sequences and 0-complete partial metric spaces and proved some characterizations of partial metric spaces in terms of completeness and 0-completeness. Adapting the same concepts, we introduce the concepts of 0-Cauchy sequences and 0-complete weak partial metric spaces.

Definition 4.1

Let \((X,q)\) be a weak partial metric space.

  1. (i)

    A sequence \(\{x_{n}\}\)in X is said to be 0-Cauchy if \(\displaystyle \lim_{n,m\rightarrow \infty } q(x_{n}, x_{m})=0\);

  2. (iii)

    \((X,q)\) is called 0-complete if every 0-Cauchy sequence \(\{x_{n}\}\) in X converges to \(x\in X\) such that \(q(x,x)=0\).

Open problems: Since 0-completeness is more general than completeness, we would like to prove

  1. (i)

    Theorem 1.1 and Theorem 2.1, and

  2. (ii)

    a Hardy–Rogers-type result

in the class of 0-complete weak partial metric spaces.