1 Introduction

Let \((X,d)\) be a metric space. For a given function \(\varphi: X\to[0,\infty)\), we define the set

$$Z_{\varphi}=\bigl\{ x\in X: \varphi(x)=0\bigr\} . $$

Let \(T: X\to X\) be a given operator. The set of fixed points of T is denoted by \(F_{T}\), that is,

$$F_{T}=\{x\in X: Tx=x\}. $$

Definition 1.1

We say that the set \(F_{T}\) is φ-admissible if and only if \(F_{T}\neq\emptyset\) and \(F_{T}\subseteq Z_{\varphi}\).

Let \(\mathcal{F}\) be the set of functions \(F:[0,\infty)^{3}\to[0,\infty )\) satisfying the following conditions:

(F1):

\(\max\{a,b\} \leq F(a,b,c)\), for all \(a, b, c \geq0\);

(F2):

\(F(a,0,0)=a\), for all \(a\geq0\);

(F3):

F is continuous.

As examples, the following functions belong to \(\mathcal{F}\):

  1. 1.

    \(F(a,b,c)=a+b+c\),

  2. 2.

    \(F(a,b,c)=\max\{a,b\}+\ln(c+1)\),

  3. 3.

    \(F(a,b,c)= a+b+c(c+1)\),

  4. 4.

    \(F(a,b,c)=(a+b)e^{c}\),

  5. 5.

    \(F(a,b,c)=(a+b)(c+1)^{n}\), \(n\in\mathbb{N}\).

Let Ψ be the set of functions \(\psi:[0,\infty)\to[0,\infty)\) satisfying the following conditions:

(Ψ1):

ψ is upper semi-continuous from the right;

(Ψ2):

\(\psi(t)< t\), for all \(t>0\).

For given functions \(\varphi:X\to[0,\infty)\), \(F\in\mathcal{F}\), and \(\psi\in\Psi\), we denote by \(\mathcal{T}(\varphi,F,\psi)\) the class of operators \(T: X\to X\) satisfying

$$ F \bigl(d(Tx,Ty),\varphi(Tx),\varphi(Ty) \bigr)\leq\psi \bigl(F \bigl(d(x,y),\varphi(x),\varphi(y) \bigr) \bigr),\quad (x,y)\in X\times X. $$
(1.1)

The aim of this paper is to study the φ-admissibility of the set \(F_{T}\), where T belongs to the class of operators \(\mathcal{T}(\varphi,F,\psi)\), \((F,\psi)\in\mathcal{F}\times\Psi\). As applications, we obtain an homotopy result and a partial metric version of the Boyd-Wong fixed point theorem.

2 Main result

Our main result is given in the following theorem.

Theorem 2.1

Let \((X,d)\) be a complete metric space and \(T: X\to X\) be a given operator. Suppose that the following conditions hold:

  1. (i)

    there exist \(\varphi: X\to[0,\infty)\), \(F\in\mathcal{F}\), and \(\psi\in\Psi\) such that \(T\in\mathcal{T}(\varphi,F,\psi)\);

  2. (ii)

    φ is lower semi-continuous.

Then the set \(F_{T}\) is φ-admissible. Moreover, the operator T has a unique fixed point.

Proof

Let ξ be an arbitrary element of the set \(F_{T}\). Take \(x=y=\xi\) in (1.1), and we get

$$ F\bigl(0,\varphi(\xi),\varphi(\xi)\bigr)\leq\psi \bigl(F \bigl(0, \varphi(\xi),\varphi (\xi) \bigr) \bigr). $$
(2.1)

If \(F(0,\varphi(\xi),\varphi(\xi))\neq0\), from (Ψ2), we get

$$\psi \bigl(F \bigl(0,\varphi(\xi),\varphi(\xi) \bigr) \bigr)< F\bigl(0,\varphi( \xi ),\varphi(\xi)\bigr), $$

which is impossible from (2.1). Consequently, we have

$$F\bigl(0,\varphi(\xi),\varphi(\xi)\bigr)=0. $$

Using the above equality and (F1), we obtain

$$\varphi(\xi)\leq F\bigl(0,\varphi(\xi),\varphi(\xi)\bigr)=0, $$

which yields

$$\varphi(\xi)=0. $$

Consequently, we have

$$ F_{T}\subseteq Z_{\varphi}. $$
(2.2)

Now, we have to prove that \(F_{T}\) is a nonempty set. Let \(x_{0}\) be an arbitrary element of X. Consider the Picard sequence \(\{x_{n}\}\subset X\) defined by

$$x_{n}=T^{n}x_{0},\quad n\in\mathbb{N}=\{0,1,2,\ldots\}, $$

where \(T^{n}\) is the nth iterate of T. If for some \(N\in \mathbb{N}\) we have \(x_{N}=x_{N+1}\), then \(x_{N}\) will be an element of \(F_{T}\). As a result we can suppose that

$$ d(x_{n},x_{n+1})>0,\quad n\in\mathbb{N}. $$
(2.3)

Using (1.1), we have

$$\begin{aligned}& F \bigl(d(Tx_{n},Tx_{n-1}),\varphi(Tx_{n}), \varphi(Tx_{n-1}) \bigr) \\& \quad\leq \psi \bigl(F \bigl(d(x_{n},x_{n-1}), \varphi(x_{n}),\varphi(x_{n-1}) \bigr) \bigr),\quad n\in\mathbb{N}^{*}; \end{aligned}$$
(2.4)

here \(\mathbb{N}^{*}=\{1,2,\ldots\}\). If for some \(N\in\mathbb{N}^{*}\), we have

$$F \bigl(d(x_{N},x_{N-1}),\varphi(x_{N}), \varphi(x_{N-1}) \bigr)=0, $$

then property (F1) yields

$$d(x_{N},x_{N-1})\leq F \bigl(d(x_{N},x_{N-1}), \varphi(x_{N}),\varphi (x_{N-1}) \bigr)=0, $$

which is a contradiction with (2.3). Thus

$$ F \bigl(d(x_{n},x_{n-1}),\varphi(x_{n}), \varphi(x_{n-1}) \bigr)>0,\quad n\in \mathbb{N}^{*}. $$
(2.5)

Using (2.4), (2.5), the definition of the sequence \(\{x_{n}\}\), and (Ψ2), we have

$$\begin{aligned} \textstyle\begin{cases} F (d(x_{n+1},x_{n}),\varphi(x_{n+1}),\varphi(x_{n}) )\leq\psi (F (d(x_{n},x_{n-1}),\varphi(x_{n}),\varphi(x_{n-1}) ) ),\\ \psi (F (d(x_{n},x_{n-1}),\varphi(x_{n}),\varphi(x_{n-1}) ) )< F (d(x_{n},x_{n-1}),\varphi(x_{n}),\varphi(x_{n-1}) ), \end{cases}\displaystyle \displaystyle \quad n\in\mathbb{N}^{*}. \end{aligned}$$
(2.6)

It follows immediately from (2.6) that there exists some \(c\geq 0\) such that

$$\begin{aligned}& \lim_{n\to\infty} F \bigl(d(x_{n+1},x_{n}), \varphi(x_{n+1}),\varphi (x_{n}) \bigr) \\& \quad=\lim_{n\to\infty}\psi \bigl(F \bigl(d(x_{n},x_{n-1}), \varphi(x_{n}),\varphi (x_{n-1}) \bigr) \bigr)=c. \end{aligned}$$
(2.7)

Suppose now that \(c>0\). Using the properties (Ψ1)-(Ψ2), we deduce from (2.7) that

$$c=\limsup_{n\to\infty}\psi \bigl(F \bigl(d(x_{n},x_{n-1}), \varphi (x_{n}),\varphi(x_{n-1}) \bigr) \bigr)\leq\psi(c)< c, $$

which is a contradiction. As a consequence, we have \(c=0\), that is,

$$\begin{aligned}& \lim_{n\to\infty} F \bigl(d(x_{n+1},x_{n}), \varphi(x_{n+1}),\varphi (x_{n}) \bigr) \\& \quad=\lim_{n\to\infty}\psi \bigl(F \bigl(d(x_{n},x_{n-1}), \varphi(x_{n}),\varphi (x_{n-1}) \bigr) \bigr)=0. \end{aligned}$$
(2.8)

Using (F1) and (2.8), we get

$$ \lim_{n\to\infty} d(x_{n+1},x_{n})= \lim_{n\to\infty}\varphi(x_{n})=0. $$
(2.9)

Next, we show that \(\{x_{n}\}\) is a Cauchy sequence in the metric space \((X,d)\). Suppose that \(\{x_{n}\}\) is not a Cauchy sequence. Then there exists \(\varepsilon>0\) for which we can find two sequences of positive integers \(\{m(k)\}\) and \(\{n(k)\}\) such that, for all \(k\in\mathbb{N}\),

$$ n(k)>m(k)>k,\qquad d(x_{m(k)}, x_{n(k)}) \geq \varepsilon,\quad\quad d(x_{m(k)},x_{n(k)-1})< \varepsilon. $$
(2.10)

Using (2.10), for all \(k\in\mathbb{N}\) we have

$$\begin{aligned} \varepsilon \leq& d(x_{m(k)}, x_{n(k)}) \\ \leq& d(x_{m(k)},x_{n(k)-1})+d(x_{n(k)-1},x_{n(k)}) \\ < & \varepsilon+ d(x_{n(k)-1},x_{n(k)}), \end{aligned}$$

which yields

$$ \varepsilon\leq d(x_{m(k)}, x_{n(k)})< \varepsilon+ d(x_{n(k)-1},x_{n(k)}), \quad k\in\mathbb{N}. $$
(2.11)

Letting \(k\to\infty\) in the above inequality and using (2.9), we obtain

$$ \begin{aligned}&\lim_{k\to\infty} d(x_{m(k)}, x_{n(k)})=\varepsilon^{+} \quad \textit{i.e.} \\ & \lim_{k\to\infty} d(x_{m(k)}, x_{n(k)})=\varepsilon\quad \hbox{and}\quad d(x_{m(k)}, x_{n(k)}) \geq \epsilon\quad \mbox{for } k\in\mathbb{N}. \end{aligned} $$
(2.12)

Using the properties (F2)-(F3), (2.9), and (2.12), we get

$$\lim_{k\to\infty}F \bigl(d(x_{n(k)},x_{m(k)}), \varphi(x_{n(k)}),\varphi (x_{m(k)}) \bigr)=F(\varepsilon,0,0)= \varepsilon^{+}. $$

Using the above limit and (Ψ1), we obtain

$$ \limsup_{k\to\infty} \psi \bigl(F \bigl(d(x_{n(k)},x_{m(k)}), \varphi (x_{n(k)}),\varphi(x_{m(k)}) \bigr) \bigr)\leq\psi( \varepsilon). $$
(2.13)

On the other hand, using (1.1) and (F1), for all \(k\in\mathbb {N}\) we have

$$\begin{aligned} \varepsilon \leq& d(x_{n(k)},x_{m(k)})\leq d(x_{n(k)},x_{n(k)+1})+d(x_{n(k)+1},x_{m(k)+1})+d(x_{m(k)+1},x_{m(k)}) \\ \leq& d(x_{n(k)},x_{n(k)+1})+F \bigl(d(x_{n(k)+1},x_{m(k)+1}), \varphi (x_{n(k)+1}),\varphi(x_{m(k)+1}) \bigr)+d(x_{m(k)+1},x_{m(k)}) \\ \leq& d(x_{n(k)},x_{n(k)+1})+\psi \bigl(F \bigl(d(x_{n(k)},x_{m(k)}), \varphi(x_{n(k)}),\varphi(x_{m(k)}) \bigr) \bigr)+d(x_{m(k)+1},x_{m(k)}). \end{aligned}$$

Passing to the limit superior as \(k\to\infty\), using (2.9), (2.13), and (Ψ2), we obtain

$$\varepsilon\leq\psi(\varepsilon)< \varepsilon, $$

which is a contradiction. As a consequence, \(\{x_{n}\}\) is a Cauchy sequence. Since \((X,d)\) is a complete metric space, there is a \(z\in X\) such that

$$ \lim_{n\to\infty}d(x_{n},z)=0. $$
(2.14)

Since φ is lower semi-continuous, it follows from (2.14) and (2.9) that

$$0\leq\varphi(z)\leq\liminf_{n\to\infty}\varphi(x_{n})=0, $$

which yields

$$ z\in Z_{\varphi}. $$
(2.15)

Now we show that \(z\in F_{T}\). Using (1.1), (F1), and (2.15), we have

$$ d(x_{n+1},Tz)\leq \psi \bigl(F \bigl(d(x_{n},z), \varphi(x_{n}),0 \bigr) \bigr),\quad n\in\mathbb{N}. $$
(2.16)

Also using the continuity of F, (F2), (2.14), and (2.9), we have

$$\lim_{n\to\infty}F \bigl(d(x_{n},z),\varphi(x_{n}),0 \bigr)=F(0,0,0)=0. $$

Note that from (Ψ2), we have

$$\lim_{t\to0^{+}}\psi(t)=0. $$

Then

$$ \lim_{n\to\infty} \psi \bigl(F \bigl(d(x_{n},z), \varphi(x_{n}),0 \bigr) \bigr)=\lim_{t\to0^{+}}\psi(t)=0. $$
(2.17)

Now, passing \(n\to\infty\) in (2.16) and using (2.17), we get

$$\lim_{n\to\infty}d(x_{n+1},Tz)=0. $$

The uniqueness of the limit yields \(z=Tz\). Thus \(F_{T}\) is a nonempty set, and the φ-admissibility of \(F_{T}\) is proved. Finally, in order to prove the uniqueness of the fixed point, let us assume that \(w\in F_{T}\) with \(d(z,w)>0\). Since \(F_{T}\) is φ-admissible, we know that \(z,w\in Z_{\varphi}\). Now, applying (1.1) with \((x,y)=(z,w)\), we obtain

$$F \bigl(d(z,w),0,0 \bigr)\leq\psi \bigl(F \bigl(d(z,w),0,0 \bigr) \bigr). $$

Using the properties (F2) and (Ψ2), we get

$$d(z,w)\leq\psi\bigl(d(z,w)\bigr)< d(z,w), $$

which is a contradiction. Thus T has a unique fixed point. □

Remark 2.2

Take \(\varphi\equiv0\) and \(F(a,b,c)=a+b+c\) in Theorem 2.1, and we recover the Boyd-Wong fixed point theorem [1].

Now, we give some examples to illustrate our main result given by Theorem 2.1.

Example 2.3

We endow the set \(X=[0,\infty)\) with the standard metric

$$d(x,y)=|x-y|, \quad(x,y)\in X\times X. $$

Let \(T: X\to X\) be the mapping defined by

$$ Tx=\textstyle\begin{cases} 0 &\mbox{if } 0\leq x\leq1,\\ \frac{x}{2} &\mbox{if } 1< x. \end{cases} $$

Observe that T is not continuous in X. So, there is no \(\psi\in\Psi \) such that

$$d(Tx,Ty)\leq\psi\bigl(d(x,y)\bigr),\quad (x,y)\in X\times X. $$

Then the Boyd-Wong fixed point theorem cannot be applied in this case. Let \(\varphi: X\to[0,\infty)\) be the function defined by

$$\varphi(x)=x^{n},\quad x\in X, \mbox{ for some } n\in\mathbb{N}^{*}. $$

Let \(F: [0,\infty)^{3}\to[0,\infty)\) be the function defined by

$$F(a,b,c)=a+b+c, \quad a,b,c\geq0. $$

Let \(\psi:[0,\infty)\to[0,\infty)\) be the function defined by

$$\psi(t)=\frac{t}{2},\quad t\geq0. $$

Observe that F belongs to the set \(\mathcal{F}\) and ψ belongs to the set Ψ. We claim that \(T\in\mathcal{T}(\varphi,F,\psi)\), that is,

$$ d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)\leq\frac{1}{2}\bigl[d(x,y)+ \varphi (x)+\varphi(y)\bigr],\quad (x,y)\in X\times X. $$
(2.18)

In order to prove our claim, we distinguish three cases.

Case 1. \((x,y)\in[0,1]\times[0,1]\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=0\leq\frac{1}{2}\bigl[d(x,y)+\varphi (x)+ \varphi(y)\bigr]. $$

Case 2. \((x,y)\in[0,1]\times(1,\infty)\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=\frac{y}{2}+ \biggl(\frac{y}{2} \biggr)^{n}, $$

while

$$\frac{1}{2}\bigl[d(x,y)+\varphi(x)+\varphi(y)\bigr]= \frac{1}{2} \bigl[y-x+x^{n}+y^{n}\bigr]. $$

Then we have to prove that

$$y^{n} \biggl(\frac{1}{2^{n-1}}-1 \biggr)\leq x^{n}-x. $$

Observe that the function \(h: [0,1]\to\mathbb{R}\) defined by

$$h(x)=x^{n}-x,\quad x\in[0,1], $$

has a global minimum at \(x_{n}= (\frac{1}{n} )^{\frac{1}{n-1}}\) which is equal to \(x_{n} (\frac{1-n}{n} )\geq\frac{1-n}{n}\). So, we have just to check that

$$y^{n} \biggl(\frac{1}{2^{n-1}}-1 \biggr)\leq\frac{1}{n}-1. $$

Since \(y>1\), we have

$$y^{n} \biggl(\frac{1}{2^{n-1}}-1 \biggr)\leq\frac{1}{2^{n-1}}-1\leq \frac{1}{n}-1. $$

Then our claim holds in this case.

Case 3. \(x,y>1\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=\frac{|x-y|}{2}+ \biggl(\frac{x}{2} \biggr)^{n}+ \biggl(\frac{y}{2} \biggr)^{n} $$

and

$$\frac{1}{2}\bigl[d(x,y)+\varphi(x)+\varphi(y)\bigr]=\frac{|x-y|}{2}+ \frac{x^{n}+y^{n}}{2}. $$

Obviously, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)\leq\frac{1}{2}\bigl[d(x,y)+\varphi (x)+ \varphi(y)\bigr]. $$

Finally, in all cases our claim (2.18) holds, which yields \(T\in \mathcal{T}(\varphi,F,\psi)\). By Theorem 2.1, the set \(F_{T}\) is φ-admissible and T has a unique fixed point. In this example, \(F_{T}=\{0\}\) and \(\varphi(0)=0\).

Example 2.4

We endow the set \(X=[\sqrt{2},\infty)\) with the standard metric

$$d(x,y)=|x-y|, \quad(x,y)\in X\times X. $$

Let \(T: X\to X\) be the mapping defined by

$$ Tx=\textstyle\begin{cases} \sqrt{2} &\mbox{if } \sqrt{2}\leq x\leq2\sqrt{2},\\ \frac{x}{2} &\mbox{if } 2\sqrt{2}< x. \end{cases} $$

As in the previous example, the Boyd-Wong fixed point theorem cannot be applied in this case. Let \(\varphi: X\to[0,\infty)\) be the function defined by

$$\varphi(x)=x^{2}-2,\quad x\in X. $$

Let \(F: [0,\infty)^{3}\to[0,\infty)\) be the function defined by

$$F(a,b,c)=a+b+c, \quad a,b,c\geq0. $$

Let \(\psi:[0,\infty)\to[0,\infty)\) be the function defined by

$$\psi(t)=\frac{t}{2},\quad t\geq0. $$

We distinguish three cases.

Case 1. \((x,y)\in[\sqrt{2}, 2\sqrt{2}]\times [\sqrt{2}, 2\sqrt{2}]\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=0\leq\frac{1}{2}\bigl[d(x,y)+\varphi (x)+ \varphi(y)\bigr]. $$

Case 2. \((x,y)\in[\sqrt{2}, 2\sqrt{2}]\times (2\sqrt{2},\infty)\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=\frac{y}{2}-\sqrt{2} +\frac{y^{2}}{4}-2, $$

while

$$\frac{1}{2}\bigl[d(x,y)+\varphi(x)+\varphi(y)\bigr]=\frac{y}{2}- \frac{x}{2}+\frac {x^{2}}{2}+\frac{y^{2}}{2}-2. $$

Clearly, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)\leq\frac{1}{2}\bigl[d(x,y)+\varphi (x)+ \varphi(y)\bigr]. $$

Case 3. \((x,y)\in(2\sqrt{2},\infty)\).

In this case, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)=\frac{|x-y|}{2}+\frac{x^{2}}{4}+ \frac{y^{2}}{4}-4, $$

while

$$\frac{1}{2}\bigl[d(x,y)+\varphi(x)+\varphi(y)\bigr]=\frac{|x-y|}{2}+ \frac {x^{2}}{2}+\frac{y^{2}}{2}-2. $$

Also, we have

$$d(Tx,Ty)+\varphi(Tx)+\varphi(Ty)\leq\frac{1}{2}\bigl[d(x,y)+\varphi (x)+ \varphi(y)\bigr]. $$

As a consequence, the mapping T belongs to \(\mathcal{T}(\varphi,F,\psi )\). By Theorem 2.1, the set \(F_{T}\) is φ-admissible and T has a unique fixed point. In this example, \(F_{T}=\{\sqrt{2}\}\) and \(\varphi(\sqrt{2})=0\).

Example 2.5

Let \((X,d)\) be the metric space considered in Example 2.4. We take the functions φ, F, and ψ defined in Example 2.4. Let \(T: X\to X\) be the mapping defined by

$$\begin{aligned} Tx=\textstyle\begin{cases} \sqrt{2} &\mbox{if } \sqrt{2}\leq x\leq2\sqrt{2},\\ \frac{\sin x}{2} &\mbox{if } 2\sqrt{2}< x. \end{cases}\displaystyle \displaystyle \end{aligned}$$

Similarly, we have \(T\in\mathcal{T}(\varphi,F,\psi)\). By Theorem 2.1, the set \(F_{T}\) is φ-admissible and T has a unique fixed point. In this example, \(F_{T}=\{\sqrt{2}\}\) and \(\varphi(\sqrt{2})=0\).

3 Applications

3.1 An homotopy result

Let us denote by \(\mathcal{F}^{*}\) the set of functions \(F\in \mathcal{F}\) satisfying the following property:

(F4):

for all \(a,b,c,d\geq0\),

$$a\leq c+d \quad\Longrightarrow\quad F(a,b,0)\leq F(c,b,0)+d. $$

As examples, the following functions belong to \(\mathcal{F}^{*}\):

  1. 1.

    \(F(a,b,c)=(a+b)e^{c}\),

  2. 2.

    \(F(a,b,c)=(a+b)(c+1)^{n}\), \(n\in\mathbb{N}\).

Observe that \(\mathcal{F}^{*}\subsetneq\mathcal{F}\). To see this, let us consider the function

$$F(a,b,c)=ae^{c+b}+be^{a+c},\quad a,b,c\geq0. $$

It is not difficult to check that \(F\in\mathcal{F}\) but \(F\notin \mathcal{F}^{*}\).

We have the following homotopy result.

Theorem 3.1

Let \((X,d)\) be a complete metric space, U be an open subset of X, and V be a closed subset of X with \(U\subset V\). Suppose that \(H: V\times[0,1]\to X\) has the following properties:

(C1):

\(x\neq H(x,\lambda)\) for every \(x\in V\backslash U\) and \(\lambda\in[0,1]\);

(C2):

there exist a continuous function \(\varphi: X\to[0,\infty )\), \(L\in(0,1)\), and \(F\in\mathcal{F}^{*}\) such that for all \(x,y\in V\) and \(\lambda\in[0,1]\),

$$F \bigl(d\bigl(H(x,\lambda),H(y,\lambda)\bigr),\varphi\bigl(H(x,\lambda)\bigr), \varphi \bigl(H(y,\lambda) \bigr)\bigr)\leq L F \bigl(d(x,y),\varphi(x), \varphi(y) \bigr); $$
(C3):

there exists a continuous function \(\eta:[0,1]\to\mathbb {R}\) such that for all \(x\in V\) and \(\lambda,\mu\in[0,1]\),

$$F \bigl(d\bigl(H(x,\lambda),H(x,\mu)\bigr),\varphi\bigl(H(x,\lambda)\bigr), \varphi \bigl(H(x,\mu ) \bigr)\bigr)\leq \bigl|\eta(\lambda)-\eta(\mu)\bigr|. $$

Then \(H(\cdot,0)\) has a fixed point if and only if \(H(\cdot,1)\) has a fixed point.

Proof

Suppose that \(H(\cdot,0)\) has a fixed point. Consider the set

$$Q=\bigl\{ t\in[0,1]: x= H(x,t)\mbox{ for some }x \in U\bigr\} . $$

From (C1), clearly 0 is an element of Q, so Q is a nonempty set. We will show that Q is both closed and open in \([0,1]\), and so by the connectedness of \([0,1]\), we are finished since \(Q=[0,1]\). First, let us prove that Q is open in \([0,1]\). Let \(t_{0}\in Q\) and \(x_{0}\in U\) with \(x_{0}= H(x_{0},t_{0})\). Using (C2) with \(x=y=x_{0}\) and \(\lambda=t_{0}\), we obtain

$$F\bigl(0,\varphi(x_{0}), \varphi(x_{0})\bigr)\leq L F \bigl(0,\varphi(x_{0}), \varphi(x_{0})\bigr), $$

which implies since \(L\in(0,1)\) that

$$F\bigl(0,\varphi(x_{0}), \varphi(x_{0})\bigr)=0. $$

Then (F1) yields

$$\varphi(x_{0})=0. $$

Moreover, observe that, for all \(t\in[0,1]\), if \(x\in U\) is a fixed point of \(H(\cdot,t)\), then \(\varphi(x)=0\). On the other hand, since U is open in \((X,d)\), there exists \(r>0\) such that \(B(x_{0},r) \subseteq U\), where

$$B(x_{0},r)=\bigl\{ z\in X: d(x_{0},z)< r\bigr\} . $$

Consider the set

$$\Lambda(x_{0},\varphi)=\bigl\{ z\in X: F\bigl(d(z,x_{0}), \varphi(z),0\bigr) < r\bigr\} . $$

Clearly \(\Lambda(x_{0},\varphi)\) is nonempty (since \(x_{0}\in \Lambda (x_{0},\varphi)\)) and \(\Lambda(x_{0},\varphi)\subseteq B(x_{0},r)\). Let \(\varepsilon= (1-L)r>0\). Since η is continuous on \(t_{0}\), there exists \(\alpha(\varepsilon)>0\) such that

$$t\in\bigl(t_{0}-\alpha(\varepsilon),t_{0}+\alpha( \varepsilon)\bigr)\cap [0,1]\quad\Longrightarrow\quad\bigl|\eta(t)-\eta(t_{0})\bigr|< \varepsilon. $$

Let \(t\in(t_{0}-\alpha(\varepsilon),t_{0}+\alpha(\varepsilon))\cap[0,1]\). For \(x\in\overline{\Lambda(x_{0},\varphi)}\) (the closure of \(\Lambda(x_{0},\varphi)\)), we have

$$F \bigl(d\bigl(H(x,t),x_{0}\bigr),\varphi\bigl(H(x,t)\bigr),0 \bigr) =F \bigl(d\bigl(H(x,t),H(x_{0},t_{0})\bigr),\varphi \bigl(H(x,t)\bigr),0 \bigr). $$

Also since

$$d\bigl(H(x,t),H(x_{0},t_{0})\bigr)\leq d \bigl(H(x,t),H(x,t_{0})\bigr)+d\bigl(H(x,t_{0}),H(x_{0},t_{0}) \bigr), $$

using the properties (F1), (F4) we get

$$\begin{aligned}& F \bigl(d\bigl(H(x,t),H(x_{0},t_{0})\bigr),\varphi \bigl(H(x,t)\bigr),0 \bigr) \\& \quad\leq F \bigl(d\bigl(H(x,t),H(x,t_{0})\bigr),\varphi\bigl(H(x,t) \bigr),0 \bigr)+d\bigl(H(x,t_{0}),H(x_{0},t_{0}) \bigr) \\& \quad\leq F \bigl(d\bigl(H(x,t),H(x,t_{0})\bigr),\varphi\bigl(H(x,t) \bigr),0 \bigr)+F \bigl(d\bigl(H(x,t_{0}),H(x_{0},t_{0}) \bigr),\varphi\bigl(H(x,t_{0})\bigr),0 \bigr) \\& \quad\leq\bigl|\eta(t)-\eta(t_{0})\bigr|+L F \bigl(d(x,x_{0}), \varphi(x),0 \bigr) \\& \quad< \varepsilon+Lr=r. \end{aligned}$$

Thus we proved that, for all \(t\in(t_{0}-\alpha(\varepsilon),t_{0}+\alpha (\varepsilon))\cap[0,1]\), the operator

$$H(\cdot,t): \overline{\Lambda(x_{0},\varphi)}\to\overline{\Lambda (x_{0},\varphi)} $$

is well defined. Now, using (C2) and Theorem 2.1, we deduce that, for all \(t\in (t_{0}-\alpha(\varepsilon), t_{0}+\alpha(\varepsilon))\cap[0,1]\), the operator \(H(\cdot,t)\) has a fixed point in V. However, such a fixed point should be in U from (C1). As a consequence,

$$\bigl(t_{0}-\alpha(\varepsilon),t_{0}+\alpha(\varepsilon) \bigr)\cap[0,1]\subseteq Q, $$

which proves that Q is open in \([0,1]\). Next, we show that Q is closed in \([0,1]\). To see this, let \(\{t_{n}\}\) be a sequence in Q with \(t_{n}\to t\in[0,1]\) as \(n\to\infty\). We have to prove that \(t\in Q\). From the definition of Q, for all \(n\in\mathbb{N}\), there exists \(x_{n}\in U\) with

$$x_{n} =H(x_{n},t_{n}) \quad\mbox{and}\quad \varphi(x_{n})=0. $$

Also for all \(m,n\in\mathbb{N}\), we have

$$\begin{aligned} d(x_{n},x_{m}) =&d\bigl(H(x_{n},t_{n}),H(x_{m},t_{m}) \bigr)\\ \leq& d\bigl(H(x_{n},t_{n}),H(x_{n},t_{m}) \bigr)+d\bigl(H(x_{n},t_{m}),H(x_{m},t_{m}) \bigr) \\ \leq& F \bigl(d\bigl(H(x_{n},t_{n}),H(x_{n},t_{m}) \bigr),\varphi\bigl(H(x_{n},t_{n})\bigr),\varphi \bigl(H(x_{n},t_{m})\bigr) \bigr) \\ &{}+F \bigl(d\bigl(H(x_{n},t_{m}),H(x_{m},t_{m}) \bigr),\varphi\bigl(H(x_{n},t_{m})\bigr),\varphi \bigl(H(x_{m},t_{m})\bigr) \bigr) \\ \leq&\bigl|\eta(t_{n})-\eta(t_{m})\bigr|+LF \bigl(d(x_{n},x_{m}),0,0 \bigr) \\ =&\bigl|\eta(t_{n})-\eta(t_{m})\bigr|+Ld(x_{n},x_{m}), \end{aligned}$$

which yields

$$d(x_{n},x_{m})\leq\frac{|\eta(t_{n})-\eta(t_{m})|}{1-L},\quad m,n\in\mathbb{N}. $$

Letting \(m,n\to\infty\) in the above inequality and using the continuity of η, we get \(d(x_{n},x_{m})\to0\) as \(m,n\to\infty\), which implies that \(\{x_{n}\}\) is a Cauchy sequence in the complete metric space \((X,d)\). Then there is some \(z\in V\) (since V is closed) such that

$$\lim_{n\to\infty} d(x_{n},z)=0 \quad\mbox{and}\quad \varphi(z)=0, $$

since φ is lower semi-continuous. Now, for all \(n\in\mathbb {N}\) we have

$$\begin{aligned} d\bigl(x_{n},H(z,t)\bigr) =& d\bigl(H(x_{n},t_{n}),H(z,t) \bigr) \leq d\bigl(H(x_{n},t_{n}),H(x_{n},t)\bigr)+d \bigl(H(x_{n},t),H(z,t)\bigr) \\ \leq& F \bigl(d\bigl(H(x_{n},t_{n}),H(x_{n},t) \bigr),\varphi\bigl(H(x_{n},t_{n})\bigr),\varphi \bigl(H(x_{n},t)\bigr) \bigr)\\ &{}+ F \bigl(d\bigl(H(x_{n},t),H(z,t) \bigr),\varphi\bigl(H(x_{n},t)\bigr),\varphi\bigl(H(z,t)\bigr) \bigr) \\ \leq&\bigl|\eta(t_{n})-\eta(t)\bigr|+LF \bigl(d(x_{n},z),0,0 \bigr) \\ =&\bigl|\eta(t_{n})-\eta(t)\bigr|+Ld(x_{n},z). \end{aligned}$$

Letting \(n\to\infty\) in the above inequality, we obtain

$$\lim_{n\to\infty}d\bigl(x_{n},H(z,t)\bigr)=0. $$

The uniqueness of the limit yields \(z=H(z,t)\). Using (C1), we deduce that \(z\in U\) and \(t\in Q\). Thus Q is closed in \([0,1]\).

For the reverse implication, we use the same technique. □

3.2 A partial metric version of Boyd-Wong fixed point theorem

In this part, using Theorem 2.1, we obtain a partial metric version of the Boyd-Wong fixed point theorem.

We start by recalling some basic definitions and properties of partial metric spaces. For more details of such spaces, we refer the reader to [220].

A partial metric on a nonempty set X is a function \(p: X\to X \to[0, \infty)\) such that for all \(x,y,z\in X\), we have

  1. (i)

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

  2. (ii)

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

  3. (iii)

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

  4. (iv)

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

A partial metric space is a pair \((X,p)\) such that X is a nonempty set and p is a partial metric on X. It is clear that, if \(p(x,y)=0\), then from (i)-(ii), \(x=y\); but if \(x=y\), \(p(x,y)\) may not be 0. A basic example of a partial metric space is the pair \(([0, \infty),p)\), where \(p(x,y)= \max\{x,y\}\).

Each partial metric p on X generates a \(T_{0}\) topology \(\tau_{p}\) on X which has as a base the family of open p-balls \(\{ B_{p}(x,\varepsilon): x\in X, \varepsilon>0\}\), where

$$B_{p}(x,\varepsilon):=\bigl\{ y\in X: p(x,y)< p(x,x)+\varepsilon\bigr\} . $$

Let \((X,p)\) be a partial metric space. A sequence \(\{x_{n}\}\subset X\) converges to some \(x\in X\) with respect to p if and only if

$$\lim_{n\to\infty} p(x_{n},x)=p(x,x). $$

A sequence \(\{x_{n}\}\subset X\) is said to be a Cauchy sequence if and only if \(\lim_{m,n\to\infty}p(x_{n},x_{m})\) exists and is finite. The partial metric space \((X,p)\) is said to be complete if and only if every Cauchy sequence \(\{x_{n}\}\) in X converges to some \(x\in X\) such that \(\lim_{n,m\to\infty}p(x_{n},x_{m})=p(x,x)\).

If p is a partial metric on X, then the function \(d_{p}: X\to X\to [0,\infty)\) defined by

$$ d_{p}(x,y)=2p(x,y)-p(x,x)-p(y,y),\quad (x,y)\in X^{2}, $$
(3.1)

is a metric on X.

Lemma 3.2

Let \((X,p)\) be a partial metric space. Then:

  1. (i)

    \(\{x_{n}\}\) is a Cauchy sequence in \((X,p)\) if and only if \(\{x_{n}\}\) is a Cauchy sequence in the metric space \((X,d_{p})\);

  2. (ii)

    the partial metric space \((X,p)\) is complete if and only if the metric space \((X,d_{p})\) is complete. Furthermore, \(\lim_{n\to\infty} d_{p}(x_{n}, x)=0\) if and only if

    $$\lim_{n\to\infty}p(x_{n},x)=p(x,x)=\lim _{m,n\to\infty}p(x_{n},x_{m}). $$

We have the following result.

Corollary 3.3

Let \((X,p)\) be a complete partial metric space and let \(T: X\to X\) be an operator such that

$$ p(Tx,Ty)\leq\psi\bigl(p(x,y)\bigr), \quad(x,y)\in X\times X, $$
(3.2)

where \(\psi\in\Psi\). We have the following results:

  1. (i)

    if \(z\in X\) is a fixed point of T then \(p(z,z)=0\);

  2. (ii)

    T has a unique fixed point.

Proof

Let \(d_{p}\) be the metric on X defined by (3.1). We have

$$p(x,y)=d(x,y)+\varphi(x)+\varphi(y), \quad(x,y)\in X\times X, $$

where

$$d(x,y)=\frac{d_{p}(x,y)}{2}, \quad\quad\varphi(x)=\frac{p(x,x)}{2}. $$

Then (3.2) yields

$$F\bigl(d(Tx,Ty),\varphi(Tx),\varphi(Ty)\bigr)\leq\psi \bigl(F \bigl(d(x,y), \varphi (x),\varphi(y) \bigr) \bigr), \quad(x,y)\in X\times X, $$

where

$$F(a,b,c)=a+b+c, \quad a,b,c\geq0. $$

From (ii) Lemma 3.2, the metric space \((X,d)\) is complete and the function φ is continuous with respect to the topology of d. Finally the desired result follows from Theorem 2.1. □

Remark 3.4

Take in Corollary 3.3, \(\psi(t)= kt\) with \(k\in(0,1)\), and we recover Matthews fixed point theorem [9].