1 Introduction and preliminaries

Development of fixed point theory on a metric space endowed with graph has a lot of activities in last few year. Jachymski [1] introduced the notion of Banach G-contraction. Later on various authors proved many fixed point theorems for single-valued and multi-valued mappings on a metric space endowed with a graph, see, for example [210]. Recently, Asl et al. [3] defined a graph-metric space and proved fixed point theorems on it. In this paper we introduce the notion of \(\alpha \)-integral type G-contraction to generalize the notions of Banach G-contraction and integral G-contraction. We also prove the fixed point theorems for such mappings and state some illustrative examples to claim that our results properly generalizes some results in literature.

Let (Xd) be a metric space and G be an undirected graph such that the set V(G) of its vertices coincides with X and the set E(G) of its edges contains all loops in V(G). Throughout this paper, we assume that G has no parallel edges. We also denote this space by \(G_d\) and call it a graph-metric space [3]. A mapping \(T:G_d\rightarrow G_d\) is said to be G-continuous if for given sequence \(\{x_{n}\}\) such that \(x_{n}\rightarrow x\) as \(n\rightarrow \infty \), where \(x\in G_d\) and \((x_{n},x_{n+1})\in E(G)\) for all \(n \in {\mathbb {N}}\), we have \(Tx_{n}\rightarrow Tx\) as \(n\rightarrow \infty \).

In 2008, Jachymski [1] introduced the notion of Banach G-contraction mappings as follows:

Definition 1.1

[1] Let (Xd) be a metric space endowed with graph G. A mapping \(T:X\rightarrow X\) is called a Banach G-contraction if T preserves the edges of G, i.e.,

$$\begin{aligned} \forall x,y\in X\quad ((x,y)\in E(G) \Rightarrow (Tx,Ty)\in E(G)) \end{aligned}$$
(1)

and

$$\begin{aligned} \exists c\in (0,1)\ \forall x,y\in X\quad ((x,y)\in E(G) \Rightarrow d(Tx,Ty)\le cd(x,y)). \end{aligned}$$

We denote by \(\Phi \) the set of all Lebesgue integrable mappings \(\phi :[0,\infty )\rightarrow [0,\infty )\) which are summable on each compact subset of \([0,\infty )\) and for each \(\epsilon >0\), we have

$$\begin{aligned} \int _{0}^{\epsilon }\phi (t)~dt>0. \end{aligned}$$
(2)

In 2013, Samreen and Kamran [9] extended the notion of Banach G-contraction in the following way:

Definition 1.2

[9] Let (Xd) be a metric space endowed with graph G. A mapping \(T:X\rightarrow X\) is called an integral G-contraction if T preserves the edges of G, i.e.,

$$\begin{aligned} \forall x,y\in X\quad ((x,y)\in E(G) \Rightarrow (Tx,Ty)\in E(G)) \end{aligned}$$
(3)

and

$$\begin{aligned} \forall x,y\in X\quad \left( (x,y)\in E(G) \Rightarrow \int _{0}^{d(Tx,Ty)}\phi (t)~dt\le c\int _{0}^{d(x,y)}\phi (t)~dt\right) , \end{aligned}$$

where \(c\in (0,1)\) and \(\phi \in \Phi \).

Recently, Kamran and Ali [11] introduced the notion of \(\alpha \)-subadmissible mappings in the following way:

Definition 1.3

Let \(\alpha :G_d\times G_d\rightarrow [0,\infty )\) be a mapping. A mapping \(T:G_d\rightarrow G_d\) is said to be \(\alpha \)-subadmissible if

  1. (i)

    for \(x,y\in G_d\), \(\alpha (x,y)\ge 1\Rightarrow \alpha (Tx,Ty)\ge 1\);

  2. (ii)

    for \(x\in G_d\), \(\alpha (T^{n}x,T^{n+1}x)\ge 1\) for all integers \(n\ge 0\) implies \(\alpha (T^{m}x,T^{n}x)\ge 1\) for all integers \(m > n\ge 0\).

By \(\Xi \), we mean the family of functions \(\xi :[0,\infty )\rightarrow [0,\infty )\) such that \(\xi \) is nondecreasing, upper semicontinuous and \(\lim _{n\rightarrow \infty }\xi ^{n}(t)=0\) for each \(t\ge 0\). Note that if \(\xi \in \Xi \), then \(\xi (t)<t\) for each \(t>0\), and \(\xi (0)=0\).

2 Main results

We begin this section with the following definition.

Definition 2.1

A mapping \(T:G_d\rightarrow G_d\) is said to be an \(\alpha \)-integral type G-contraction mapping if there exist three functions \(\alpha :G_d\times G_d\rightarrow [0,\infty )\), \(\phi \in \Phi \) and \(\xi \in \Xi \) such that

$$\begin{aligned} \alpha (x,y)\int _{0}^{d(Tx,Ty)}\phi (t)~dt\le \xi \left( \int _{0}^{d(x,y)}\phi (t)~dt \right) , \end{aligned}$$
(4)

for each \((x,y)\in E(G)\).

Remark 2.2

Let (Xd) be a metric space. Define graph G by \(V(G)=X\) and \(E(G)=X\times X\), \(\alpha (x,y)=1\) for each \(x,y \in X\) and \(\xi (t)=ct\) for all \(t\ge 0\), where \(c\in [0,1)\). Then (4) reduces to

$$\begin{aligned} \int _{0}^{d(Tx,Ty)}\phi (t)~dt\le c\int _{0}^{d(x,y)}\phi (t)~dt, \end{aligned}$$

for each \(x,y\in X\), where \(\phi \in \Phi \), which is the contractive condition considered by Branciari [12].

Note that every integral G-contraction mapping is an \(\alpha \)-integral type G-contraction mapping. The following example shows that the converse is not true in general.

Example 2.3

Let \(X=[0,\infty )\) with the usual metric d and X endowed with graph G is defined by \(V(G)=X\) and \(E(G)=\{(x,y):x\ge y\}\). Define \(T:G_d\rightarrow G_d\) and \(\alpha :G_d\times G_d\rightarrow [0,\infty )\) by

$$\begin{aligned} Tx= {\left\{ \begin{array}{ll} \frac{x}{4}&{}\text{ if } x\ge 4, \\ x^{2}&{}\text{ if } 0\le x<4, \end{array}\right. } \end{aligned}$$
(5)

and

$$\begin{aligned} \alpha (x,y)={\left\{ \begin{array}{ll} 1&{}\text{ if } x,y\ge 4,\\ 0&{}\text{ otherwise }. \end{array}\right. } \end{aligned}$$
(6)

Take \(\xi (t)=\frac{t}{2}\) for each \(t\ge 0\) and \(\phi (t)={\left\{ \begin{array}{ll} 0&{}\text{ if } t=0,\\ \frac{t^{-1/2}}{2}&{}\text{ if } t>0.\end{array}\right. }\) If \(x,y\ge 4\), then we have

$$\begin{aligned} \alpha (x,y)\int _{0}^{d(Tx,Ty)}\phi (t)~dt=\sqrt{\frac{|x-y|}{4}}= \frac{1}{2}\int _{0}^{d(x,y)}\phi (t)~dt, \end{aligned}$$

and for otherwise, we have

$$\begin{aligned} \alpha (x,y)\int _{0}^{d(Tx,Ty)}\phi (t)~dt=0\le \frac{1}{2} \int _{0}^{d(x,y)}\phi (t)~dt. \end{aligned}$$

Thus, (4) holds for each \((x,y)\in E(G)\). Therefore, T is an \(\alpha \)-integral type G-contraction mapping. But T is not an integral G-contraction, since \((4,2)\in E(G)\nRightarrow (T4,T2)\in E(G)\).

Theorem 2.4

Let \(G_d\) be a complete graph-metric space and \(T:G_d\rightarrow G_d\) be an \(\alpha \)-integral type G-contraction mapping satisfying the following assumptions : 

  1. (i)

    T is \(\alpha \)-subadmissible; 

  2. (ii)

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

  3. (iii)

    if \(x,y\in G_d\) and \(\alpha (x,y)\ge 1,\) then \((x,y)\in E(G);\)

  4. (iv)

    T is G-continuous; 

Then T has a fixed point.

Proof

Starting from \(x_{0}\in G_d\) in (ii). Define a sequence \(\{x_{n}\}\) in \(G_d\) such that \(x_{n+1}=Tx_{n}\) for each \(n\in {\mathbb {N}}\cup \{0\}\). Since T is \(\alpha \)-subadmissible, by induction we have

$$\begin{aligned} \alpha (x_{n},x_{n+1})\ge 1\quad \text {for each }n\in {\mathbb {N}}\cup \{0\}. \end{aligned}$$
(7)

From (4), we have

$$\begin{aligned} \int _{0}^{d(x_{n},x_{n+1})}\phi (t)~dt= & {} \int _{0}^{d(Tx_{n-1},Tx_{n})}\phi (t)~dt\nonumber \\\le & {} \alpha (x_{n-1},x_{n})\int _{0}^{d(Tx_{n-1},Tx_{n})}\phi (t)~dt\nonumber \\\le & {} \xi \left( \int _{0}^{d(x_{n-1},x_{n})}\phi (t)~dt\right) , \end{aligned}$$

for all \(n\in {\mathbb {N}}\). By induction, we have

$$\begin{aligned} \int _{0}^{d(x_{n},x_{n+1})}\phi (t)~dt\le \xi ^{n} \left( \int _{0}^{d(x_{0},x_{1})}\phi (t)~dt\right) , \end{aligned}$$
(8)

for all \(n\in {\mathbb {N}}\). Letting \(n\rightarrow \infty \) in (8), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\int _{0}^{d(x_{n},x_{n+1})}\phi (t)~dt=0. \end{aligned}$$
(9)

From (2) and (9), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }d(x_{n},x_{n+1})=0. \end{aligned}$$
(10)

Now, using a similar argument as given by authors in [13, 14], we will show that \(\{x_{n}\}\) is a Cauchy sequence in \(G_d\). Assume on contrary that \(\{x_{n}\}\) is not a Cauchy sequence. Then there exists \(\epsilon >0\) for which we can find two sequences of positive integers \(\{n_{k}\}\) and \(\{m_{k}\}\) such that for each \(k\in {\mathbb {N}}\), we have

$$\begin{aligned} n_{k}>m_{k}>k,\quad d(x_{n_{k}},x_{m_{k}})\ge \epsilon ,\quad d(x_{n_{k}-1},x_{m_{k}})< \epsilon . \end{aligned}$$
(11)

By using triangular inequality and (11), we have

$$\begin{aligned} \epsilon \le d(x_{n_{k}},x_{m_{k}})\le & {} d(x_{n_{k}},x_{n_{k}-1})+ d(x_{n_{k}-1},x_{m_{k}}) \nonumber \\< & {} d(x_{n_{k}},x_{n_{k}-1})+\epsilon . \end{aligned}$$

Letting \(k\rightarrow \infty \) in above inequality and using (10), we get

$$\begin{aligned} \lim _{k\rightarrow \infty }d(x_{n_{k}},x_{m_{k}})=\epsilon . \end{aligned}$$
(12)

Again by using triangular inequality, we have

$$\begin{aligned} d(x_{n_{k}-1},x_{m_{k}-1}) \le d(x_{n_{k}-1},x_{n_{k}})+ d(x_{n_{k}},x_{m_{k}})+d(x_{m_{k}},x_{m_{k}-1}) \end{aligned}$$

and

$$\begin{aligned} d(x_{n_{k}},x_{m_{k}}) \le d(x_{n_{k}},x_{n_{k}-1})+ d(x_{n_{k}-1},x_{m_{k}-1})+d(x_{m_{k}-1},x_{m_{k}}) \end{aligned}$$

for all \(k\in {\mathbb {N}}\). Using (10) and (12) in above two inequalities, we get

$$\begin{aligned} \lim _{k\rightarrow \infty }d(x_{n_{k}-1},x_{m_{k}-1})=\epsilon . \end{aligned}$$
(13)

As T is \(\alpha \)-subadmissible, from (7), we have \(\alpha (x_{n_{k}-1},x_{m_{k}-1})\ge 1\) for all \(k\in {\mathbb {N}}\). Then by Condition (iii) we get \((x_{n_{k}-1},x_{m_{k}-1})\in E(G)\) for all \(k\in {\mathbb {N}}\). From (4), we have

$$\begin{aligned} \int _{0}^{d(x_{n_{k}},x_{m_{k}})}\phi (t)~dt= & {} \int _{0}^{d(Tx_{n_{k}-1},Tx_{m_{k}-1})}\phi (t)~dt\nonumber \\\le & {} \alpha (x_{n_{k}-1},x_{m_{k}-1})\int _{0}^{d(Tx_{n_{k}-1}, Tx_{m_{k}-1})}\phi (t)~dt\nonumber \\\le & {} \xi \left( \int _{0}^{d(x_{n_{k}-1},x_{m_{k}-1})} \phi (t)~dt\right) \end{aligned}$$
(14)

for all \(k\in {\mathbb {N}}\). Using properties of \(\xi \) and let \(k\rightarrow \infty \) in (14), we have

$$\begin{aligned} \int _{0}^{\epsilon }\phi (t)~dt\le \xi \left( \int _{0}^{\epsilon }\phi (t)~dt\right) < \int _{0}^{\epsilon }\phi (t)~dt. \end{aligned}$$

A contradiction to our assumption. Hence \(\{x_{n}\}\) is a Cauchy sequence in \(G_d\). Since \(G_d\) is complete, there exists \(x^{*}\in G_d\) such that \(x_{n}\rightarrow x^{*}\) as \(n\rightarrow \infty \). G-continuity of T implies \(x_{n+1}=Tx_{n}\rightarrow Tx^{*}\) as \(n\rightarrow \infty \). By uniqueness of limit, we have \(Tx^{*}=x^{*}\). \(\square \)

Theorem 2.5

Let \(G_d\) be a complete graph-metric space and \(T:G_d\rightarrow G_d\) be an \(\alpha \)-integral type G-contraction mapping satisfying the following assumptions : 

  1. (i)

    T is \(\alpha \)-subadmissible; 

  2. (ii)

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

  3. (iii)

    if \(x,y\in G_d\) and \(\alpha (x,y)\ge 1,\) then \((x,y)\in E(G);\)

  4. (iv)

    if \(\{x_{n}\}\) is a sequence in \(G_d\) such that \(x_{n}\rightarrow x\) as \(n\rightarrow \infty \) and \(\alpha (x_{n},x_{n+1})\ge 1\) for each \(n\in \mathbb {N}\cup \{0\}\), then \(\alpha (x_{n},x)\ge 1\) for each \(n\in \mathbb {N}\cup \{0\}\).

Then T has a fixed point.

Proof

According to the proof of Theorem 2.4, we know that \(\{x_{n}\}\) be a Cauchy sequence in \(G_d\). Since \(G_d\) is complete, there exists \(x^{*}\in G_d\) such that \(x_{n}\rightarrow x^{*}\) as \(n\rightarrow \infty \). From (7) and Conditions (iv) and (iii), we have

$$\begin{aligned} \alpha (x_{n},x^{*})\ge 1 \text { and }(x_{n},x^{*})\in E(G) \quad \text {for each }n\in {\mathbb {N}}\cup \{0\}. \end{aligned}$$
(15)

By the triangular inequality, we have

$$\begin{aligned} |d(Tx^{*},x^{*})-d(Tx_{n},x^{*})|\le d(Tx^{*},Tx_{n}) \end{aligned}$$
(16)

for all \(n\in {\mathbb {N}} \cup \{0\}\). Thus, we have

$$\begin{aligned} \int _{0}^{|d(Tx^{*},x^{*})-d(Tx_{n},x^{*})|}\phi (t)~dt\le & {} \int _{0}^{d(Tx^{*},Tx_{n})}\phi (t)~dt\nonumber \\\le & {} \alpha (x_{n},x^{*})\int _{0}^{d(Tx^{*},Tx_{n})}\phi (t)~dt\nonumber \\\le & {} \xi \left( \int _{0}^{d(x^{*},x_{n})}\phi (t)~dt\right) \end{aligned}$$
(17)

for all \(n\in {\mathbb {N}} \cup \{0\}\). Letting \(n\rightarrow \infty \) in (17), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\int _{0}^{|d(Tx^{*},x^{*})-d(Tx_{n},x^{*})|}\phi (t)~dt=0. \end{aligned}$$
(18)

From (18), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }|d(Tx^{*},x^{*})-d(Tx_{n},x^{*})|=0. \end{aligned}$$
(19)

This implies that

$$\begin{aligned} d(Tx^{*},x^{*})=\lim _{n\rightarrow \infty }d(Tx_{n},x^{*})= \lim _{n\rightarrow \infty }d(x_{n+1},x^{*})=0. \end{aligned}$$

Hence \(Tx^{*} = x^{*}\). \(\square \)

We use the following condition to obtain the uniqueness of the fixed point.

(A) For each fixed points x and y of T, we have \(\alpha (x,y)\ge 1\).

Theorem 2.6

Adding the Condition \((\mathrm{A})\) to the hypotheses of Theorem 2.4 (resp. Theorem 2.5), we get the uniqueness of the fixed point of T.

Proof

Suppose that \(x^{*}\) and \(y^{*}\) are two distinct fixed points of T. From Condition (A), we get \(\alpha (x^*,y^*)\ge 1\) and hence

$$\begin{aligned} \int _{0}^{d(x^{*},y^{*})}\phi (t)~dt\le & {} \alpha (x^{*},y^{*})\int _{0}^{d(Tx^{*},Ty^{*})}\phi (t)~dt\nonumber \\\le & {} \xi \left( \int _{0}^{d(x^{*},y^{*})}\phi (t)~dt\right) \nonumber \\< & {} \int _{0}^{d(x^{*},y^{*})}\phi (t)~dt. \end{aligned}$$
(20)

A contradiction to our assumption. Hence T has unique fixed point. \(\square \)

Example 2.7

Let \(X={\mathbb {R}}\) with the usual metric d and X endowed with the graph G such that \(V(G)=X\) and \(E(G)=\{(x,y): x,y\ge -1\}\). Define \(T:G_d\rightarrow G_d\) and \(\alpha :G_d\times G_d\rightarrow [0,\infty )\) by

$$\begin{aligned} Tx= {\left\{ \begin{array}{ll} \frac{x}{2}&{}\text{ if } x\ge 0,\\ x&{}\text{ if } x<0, \end{array}\right. } \end{aligned}$$
(21)

and

$$\begin{aligned} \alpha (x,y)={\left\{ \begin{array}{ll} 1&{}\text{ if } x,y\ge 0,\\ 0&{}\text{ otherwise }.\end{array}\right. } \end{aligned}$$
(22)

Take \(\xi (t)=\frac{t}{4}\) and \(\phi (t)=t\) for each \(t\ge 0\). If \(x,y\ge 0\), then we have

$$\begin{aligned} \alpha (x,y)\int _{0}^{d(Tx,Ty)}\phi (t)~dt=\frac{1}{2} \left( \frac{|x-y|}{2}\right) ^{2}= \frac{1}{4}\int _{0}^{d(x,y)}\phi (t)~dt=\xi \left( \int _{0}^{d(x,y)}\phi (t)~dt\right) , \end{aligned}$$

and for otherwise, we have

$$\begin{aligned} \alpha (x,y)\int _{0}^{d(Tx,Ty)}\phi (t)~dt=0\le \frac{1}{4} \int _{0}^{d(x,y)}\phi (t)~dt=\xi \left( \int _{0}^{d(x,y)}\phi (t)~dt\right) . \end{aligned}$$

This shows that (4) holds for each \((x,y)\in E(G)\). Therefore, T is an \(\alpha \)-integral type G-contraction mapping. If \(x,y\in G_d\) and \(\alpha (x,y)=1\), then \(x,y\ge 0\). By definition of T and \(\alpha \), we have \(\alpha (Tx,Ty)=1\). For each \(x\in G_d\) such that \(\alpha (T^{n}x,T^{n+1}x) \ge 1\) for all integers \(n\ge 0\), we have \(\alpha (T^{m}x,T^{n}x)=1\) for all integers \(m > n\ge 0\). Hence T is \(\alpha \)-subadmissible. Also, we have \(x_{0}=1\in G_d\) such that \(\alpha (1,T1)=\alpha (1,1/2)=1\). Moreover, if \(x,y\in G_d\) such that \(\alpha (x,y) \ge 1\), then \(x,y\ge 0\) which implies that \((x,y)\in E(G)\). Finally, it is easy to see that the continuity of T implies that T is G-continuous. Therefore, all the hypotheses of Theorem 2.4 hold. Hence T has a fixed point, that is, a point 0.

Note that T is neither a Banach G-contraction mapping nor an integral G-contraction mapping. Indeed, if we take \((x,y) = (-1,-1/2)\in E(G)\), we get

$$\begin{aligned} d(Tx,Ty)=d(T(-1),T(-1/2)) > cd(-1,-1/2)=cd(x,y) \end{aligned}$$

and

$$\begin{aligned} \int _{0}^{d(Tx,Ty)}\phi (t)~dt = \int _{0}^{d(T(-1),T(-1/2))}\phi (t)~dt > c\int _{0}^{d(-1,-1/2)}\phi (t)~dt = c\int _{0}^{d(x,y)}\!\phi (t)~dt \end{aligned}$$

for all \(c\in (0,1)\) and \(\phi \in \Phi \).