Skip to main content

Best proximity point results in partially ordered metric spaces via simulation functions

Abstract

We obtain sufficient conditions for the existence and uniqueness of best proximity points for a new class of non-self mappings involving simulation functions in a metric space endowed with a partial order. Some interesting consequences including fixed point results via simulation functions are presented.

1 Introduction

Recently, in [1] the authors introduced the class of simulation functions as follows.

Definition 1.1

We say that \(\xi:[0,\infty)\times[0,\infty)\to\mathbb{R}\) is a simulation function if it satisfies the following conditions:

  1. (i)

    \(\xi(0,0)=0\);

  2. (ii)

    \(\xi(t,s)< s-t\), for every \(t,s>0\);

  3. (iii)

    if \(\{a_{n}\}\) and \(\{b_{n}\}\) are two sequences in \((0,\infty )\), then

    $$\lim_{n\to\infty} a_{n}=\lim_{n\to\infty} b_{n}>0\quad \Longrightarrow\quad \limsup_{n\to\infty} \xi(a_{n},b_{n})< 0. $$

Various examples of simulation functions were presented in [1]. The class of such functions will be denoted by \(\mathcal{Z}\).

Definition 1.2

([1])

Let \(T: X\to X\) be a given operator, where X is a nonempty set equipped with a metric d. We say that T is a \(\mathcal {Z}\)-contraction with respect to \(\xi\in\mathcal{Z}\) if

$$\xi\bigl(d(Tx,Ty),d(x,y)\bigr)\geq0,\quad \mbox{for all } x,y\in X. $$

In [1], the authors established the following fixed point theorem that generalizes many previous results from the literature including the Banach fixed point theorem.

Theorem 1.3

([1])

Let \(T: X\to X\) be a given map, where X is a nonempty set equipped with a metric d such that \((X,d)\) is complete. Suppose that T is a \(\mathcal{Z}\)-contraction with respect to \(\xi\in\mathcal{Z}\). Then T has a unique fixed point. Moreover, for any \(x\in X\), the sequence \(\{T^{n}x\}\) converges to this fixed point.

For other results via simulation functions, we refer to [2–7].

Let \((X,d)\) be a metric space. Consider a mapping \(T: A\to B\), where A and B are nonempty subsets of X. If \(d(x,Tx)>0\) for every \(x\in A\), then the set of fixed points of T is empty. In this case, we are interested in finding a point \(x\in A\) such that \(d(x,Tx)\) is minimum in some sense.

Definition 1.4

We say that \(z\in A\) is a best proximity point of T if

$$d(z,Tz)=d(A,B):=\inf\bigl\{ d(x,y): x\in A, y\in B\bigr\} . $$

Observe that if \(d(A,B)=0\), then a best proximity point of T is a fixed point of T.

The study of the existence of best proximity points is an interesting field of optimization and it attracted recently the attention of several researchers (see [1, 8–23] and the references therein).

In the sequel, we will use the following notations. Set

$$A_{0}=\bigl\{ x \in A : d(x,y)=d(A,B), \mbox{for some }y \in B\bigr\} $$

and

$$B_{0}=\bigl\{ y\in B: d(x,y)=d(A,B), \mbox{for some }x \in A\bigr\} . $$

We refer to [19] for sufficient conditions that guarantee that \(A_{0}\) and \(B_{0}\) are nonempty.

Now, we endow the set X with a partial order ⪯. Let us introduce the following class of mappings. For a given simulation function \(\xi\in\mathcal{Z}\), we denote by \(\mathcal{T}_{\xi}\) the set of mappings \(T: A\to B\) satisfying the following conditions:

  1. (C1)

    for every \(x_{1},x_{2},y_{1},y_{2}\in A\), we have

    $$y_{1}\preceq y_{2},\quad d(x_{1},Ty_{1})=d(x_{2},Ty_{2})=d(A,B) \quad \Longrightarrow\quad x_{1}\preceq x_{2}; $$
  2. (C2)

    for every \(x,y,u_{1},u_{2}\in A\), we have

    $$x\preceq y, x\neq y,\quad d(u_{1},Tx)=d(u_{2},Ty)=d(A,B) \quad \Longrightarrow\quad \xi \bigl(d(u_{1},u_{2}),m(x,y) \bigr)\geq0, $$

    where

    $$m(x,y)=\max \biggl\{ \frac{d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} . $$

Our aim in this paper is to study the existence and uniqueness of best proximity points for non-self mappings \(T: A\to B\) that belong to the class \(\mathcal{T}_{\xi}\), for some simulation function \(\xi\in\mathcal{Z}\).

2 Main results

Our first main result is the following.

Theorem 2.1

Let \(T\in\mathcal{T}_{\xi}\), for some \(\xi\in\mathcal{Z}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    A is closed with respect to the metric d;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B), \quad x_{0}\preceq x_{1}; $$
  5. (5)

    T is continuous.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Proof

By condition (4), we have

$$d(x_{1},Tx_{0})=d(A,B), $$

for some \(x_{0},x_{1}\in A_{0}\) such that \(x_{0}\preceq x_{1}\). Condition (3) implies that \(Tx_{1}\in B_{0}\), which yields

$$d(x_{2},Tx_{1})=d(A,B), $$

for some \(x_{2}\in A_{0}\). Since \(x_{0}\preceq x_{1}\), condition (C1) implies that \(x_{1}\preceq x_{2}\). Continuing this process, by induction, we can construct a sequence \(\{x_{n}\}\subset A_{0}\) such that

$$ d(x_{n+1},Tx_{n})=d(A,B),\quad n=0,1,2,\ldots $$
(2.1)

and

$$x_{0}\preceq x_{1}\preceq x_{2}\preceq\cdots \preceq x_{n}\preceq x_{n+1}\preceq\cdots. $$

Suppose that for some \(p=0,1,2,\ldots \) , we have \(x_{p+1}=x_{p}\). In this case, we get \(d(x_{p},Tx_{p})=d(A,B)\), that is, \(x_{p}\) is a best proximity point of T. So, without restriction of the generality, we may suppose that

$$x_{n}\neq x_{n+1},\quad n=0,1,2,\ldots. $$

Since

$$x_{n}\preceq x_{n+1}, x_{n}\neq x_{n+1},\quad d(x_{n},Tx_{n-1})=d(x_{n+1},Tx_{n})=d(A,B), \quad n=1,2,3,\ldots, $$

it follows from condition (C2) that

$$\xi\bigl(d(x_{n},x_{n+1}),m(x_{n-1},x_{n}) \bigr)\geq0,\quad n=1,2,3,\ldots, $$

where

$$\begin{aligned} m(x_{n-1},x_{n}) =&\max \biggl\{ \frac {d(x_{n-1},x_{n})d(x_{n},x_{n+1})}{d(x_{n-1},x_{n})},d(x_{n-1},x_{n}) \biggr\} \\ =& \max \bigl\{ d(x_{n},x_{n+1}),d(x_{n-1},x_{n}) \bigr\} . \end{aligned}$$

Suppose that for some \(n_{0}=1,2,3,\ldots \) , we have

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

In this case, we obtain

$$0\leq\xi\bigl(d(x_{n_{0}},x_{n_{0}+1}),d(x_{n_{0}},x_{n_{0}+1}) \bigr). $$

On the other hand, since \(d(x_{n_{0}},x_{n_{0}+1})>0\), using the property (ii) of a simulation function, we obtain

$$\xi\bigl(d(x_{n_{0}},x_{n_{0}+1}),d(x_{n_{0}},x_{n_{0}+1}) \bigr)< 0, $$

which is a contradiction. As a consequence,

$$ \max \bigl\{ d(x_{n},x_{n+1}),d(x_{n-1},x_{n}) \bigr\} = d(x_{n-1},x_{n}),\quad n=1,2,3,\ldots. $$
(2.2)

Thus, we obtain

$$ \xi\bigl(d(x_{n},x_{n+1}),d(x_{n-1},x_{n}) \bigr)\geq0,\quad n=1,2,3,\ldots. $$
(2.3)

From (2.2), we deduce that the sequence \(\{r_{n}\}\) defined by

$$r_{n}=d(x_{n},x_{n+1}),\quad n=0,1,2,\ldots $$

is decreasing, which yields

$$\lim_{n\to\infty}r_{n}=r, $$

where \(r\in[0,\infty)\). Suppose that \(r>0\). Using (2.3) and the property (iii) of a simulation function, we deduce that

$$0\leq\limsup_{n\to\infty} \xi\bigl(d(x_{n},x_{n+1}),d(x_{n-1},x_{n}) \bigr)< 0, $$

which is a contradiction. As consequence, we have

$$ \lim_{n\to\infty} d(x_{n},x_{n+1})=0. $$
(2.4)

Let us prove now that \(\{x_{n}\}\) is a Cauchy sequence. We argue by contradiction by supposing that \(\{x_{n}\}\) is not a Cauchy sequence. In this case, there is some \(\varepsilon>0\) for which there are subsequences \(\{x_{m(k)}\}\) and \(\{x_{n(k)}\}\) of \(\{x_{n}\}\) such that

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

Using the triangle inequality, we have

$$ \varepsilon\leq d(x_{m(k)},x_{n(k)}) \leq d(x_{m(k)},x_{n(k)-1})+d(x_{n(k)-1},x_{n(k)}) < \varepsilon+d(x_{n(k)-1},x_{n(k)}). $$

Thus we have

$$\varepsilon\leq d(x_{m(k)},x_{n(k)})< \varepsilon +d(x_{n(k)-1},x_{n(k)}),\quad \mbox{for all } k. $$

Letting \(k\to\infty\) and using (2.4), we obtain

$$ \lim_{n\to\infty} d(x_{m(k)},x_{n(k)})= \varepsilon. $$
(2.5)

Again, the triangle inequality yields

$$\bigl\vert d(x_{n(k)-1},x_{m(k)})-d(x_{m(k)},x_{n(k)}) \bigr\vert \leq d(x_{n(k)-1},x_{n(k)}),\quad \mbox{for all } k. $$

Letting \(k\to\infty\), using (2.4) and (2.5), we obtain

$$ \lim_{n\to\infty} d(x_{n(k)-1},x_{m(k)})= \varepsilon. $$
(2.6)

Similarly, we have

$$\bigl\vert d(x_{n(k)-1},x_{m(k)-1})-d(x_{n(k)-1},x_{m(k)}) \bigr\vert \leq d(x_{m(k)-1},x_{m(k)}),\quad \mbox{for all } k. $$

Letting \(k\to\infty\), using (2.4) and (2.6), we obtain

$$ \lim_{n\to\infty} d(x_{n(k)-1},x_{m(k)-1})= \varepsilon. $$
(2.7)

Observe that for k large enough, we have

$$x_{m(k)-1}\preceq x_{n(k)-1},\qquad x_{m(k)-1}\neq x_{n(k)-1} $$

and

$$d(x_{m(k)},Tx_{m(k)-1})= d(x_{n(k)},Tx_{n(k)-1})=d(A,B). $$

Then condition (C2) yields

$$ \xi\bigl(d(x_{m(k)},x_{n(k)}),m(x_{m(k)-1},x_{n(k)-1}) \bigr)\geq0, \quad \mbox{for all } k. $$
(2.8)

On the other hand, for all k, we have

$$m(x_{m(k)-1},x_{n(k)-1})=\max \biggl\{ \frac {d(x_{m(k)-1},x_{m(k)})d(x_{n(k)-1},x_{n(k)})}{d(x_{m(k)-1},x_{n(k)-1})},d(x_{m(k)-1},x_{n(k)-1}) \biggr\} . $$

Passing \(k\to\infty\) and using (2.4) and (2.7), we get

$$ \lim_{k\to\infty}m(x_{m(k)-1},x_{n(k)-1})= \varepsilon. $$
(2.9)

Using (2.5), (2.9), (2.8) and the condition (iii) of a simulation function, we have

$$0\leq\limsup_{k\to\infty} \xi \bigl(d(x_{m(k)},x_{n(k)}),m(x_{m(k)-1},x_{n(k)-1}) \bigr)< 0, $$

which is a contradiction. As consequence, the sequence \(\{x_{n}\}\) is Cauchy. Since A is a closed subset of the complete metric space \((X,d)\) (from conditions (1) and (2)), there is some \(z\in A\) such that

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

The continuity of T (from condition (5)) yields

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

Since \(d(x_{n+1},Tx_{n})=d(A,B)\) for all \(n=0,1,2,\ldots \) , we obtain

$$d(A,B)=\lim_{n\to\infty}d(x_{n+1},Tx_{n})=d(z,Tz), $$

that is, \(z\in A\) is a best proximity point of T. This ends the proof. □

Next, we obtain a best proximity point result for mappings \(T\in \mathcal{T}_{\xi}\) that are not necessarily continuous.

We say that the set A is \((d,\preceq)\)-regular if it satisfies the following property:

$$\{a_{n}\} \subset A \mbox{ is nondecreasing w.r.t.}\preceq\quad \mbox{and}\quad \lim_{n\to\infty}d(a_{n},a)=0\quad \Longrightarrow \quad a=\sup\{a_{n}\}. $$

Theorem 2.2

Let \(T\in\mathcal{T}_{\xi}\), for some \(\xi\in\mathcal{Z}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    \(A_{0}\) is closed;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B),\quad x_{0}\preceq x_{1}; $$
  5. (5)

    A is \((d,\preceq)\)-regular.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Proof

Let us consider the sequence \(\{x_{n}\}\subset A_{0}\) defined by (2.1). Following the proof of Theorem 2.1, we know that \(\{x_{n}\}\) is a Cauchy sequence. Since \(A_{0}\) is closed, there is some \(z\in A_{0}\) such that

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

From condition (3), we have \(Tz\in B_{0}\), which yields

$$d(y_{1},Tz)=d(A,B), $$

for some \(y_{1}\in A_{0}\). On the other hand, the regularity condition (5) implies that

$$x_{n}\preceq z,\quad \mbox{for all } n. $$

Since for all n,

$$x_{n}\preceq z,\quad d(x_{n+1},Tx_{n})=d(y_{1},Tz)=d(A,B), $$

condition (C1) yields

$$x_{n+1}\preceq y_{1}, \quad \mbox{for all } n. $$

On the other hand, we know that \(z=\sup\{x_{n}\}\), which implies that

$$z\preceq y_{1}. $$

Thus we have

$$d(y_{1},Tz)=d(A,B),\quad z\preceq y_{1}. $$

Again, since \(Ty_{1}\in B_{0}\), there is some \(y_{2}\in A_{0}\) such that \(d(y_{2},Ty_{1})=d(A,B)\). Condition (C1) yields \(y_{1}\preceq y_{2}\). Thus we have

$$d(y_{2},Ty_{1})=d(A,B),\quad y_{1}\preceq y_{2}. $$

Set \(y_{0}=z\) and continuing this process, we can build a sequence \(\{ y_{n}\}\subset A_{0}\) such that

$$d(y_{n+1},Ty_{n})=d(A,B),\quad n=0,1,2,\ldots $$

and

$$y_{0}\preceq y_{1}\preceq y_{2}\preceq\cdots \preceq y_{n}\preceq y_{n+1}\preceq\cdots. $$

Following similar arguments as in the proof of Theorem 2.1, we can prove that \(\{y_{n}\}\) is a Cauchy sequence in the closed subset \(A_{0}\) of the complete metric space \((X,d)\), which yields

$$\lim_{n\to\infty} d(y_{n},y)=0, $$

for some \(y\in A_{0}\). The regularity assumption (5) implies that \(y=\sup \{y_{n}\}\). So, we have

$$x_{n}\preceq z=y_{0}\preceq y_{1}\preceq\cdots \preceq y_{n}\preceq y,\quad \mbox{for all } n. $$

We claim that \(z=y\). In order to prove our claim, suppose that \(d(z,y)>0\). Set

$$I=\{n: x_{n}=z\}. $$

We consider two cases.

Case 1. If \(|I|=\infty\).

In this case, there is a subsequence \(\{x_{n_{k}}\}\) of \(\{x_{n}\}\) such that

$$x_{n_{k}}=z,\quad \mbox{for all } k, $$

which implies that z is a best proximity point. So, this case is trivial.

Case 2. If \(|I|<\infty\).

In this case, for n large enough, we have

$$x_{n}\neq z, \quad x_{n}\preceq z\preceq y_{n}, \quad \mbox{for all } n. $$

From condition (C2), for n large enough, we obtain

$$\xi\bigl(d(x_{n+1},y_{n+1}),m(x_{n},y_{n}) \bigr)\geq0, $$

where

$$m(x_{n},y_{n})=\max \biggl\{ \frac {d(x_{n},x_{n+1})d(y_{n},y_{n+1})}{d(x_{n},y_{n})},d(x_{n},y_{n}) \biggr\} . $$

Observe that

$$\lim_{n\to\infty}d(x_{n+1},y_{n+1})=\lim _{n\to\infty}m(x_{n},y_{n})=d(z,y)>0. $$

From the property (iii) of simulation functions, we obtain

$$0\leq\limsup_{n\to\infty} \xi\bigl(d(x_{n+1},y_{n+1}),m(x_{n},y_{n}) \bigr) < 0, $$

which is a contradiction. As consequence, we have \(z=y\).

Since \(z=y\), we obtain

$$x_{n}\preceq z=y_{0}\preceq y_{1}\preceq\cdots \preceq y_{n}\preceq y=z, \quad \mbox{for all } n, $$

which implies that

$$y_{n}=z,\quad \mbox{for all } n. $$

Since \(d(y_{n+1},Ty_{n})=d(A,B)\), we have \(d(z,Tz)=d(A,b)\), that is, z is a best proximity point of T. This completes the proof. □

Note that the assumptions in Theorems 2.1 and 2.2 do not guarantee the uniqueness of the best proximity point. The next example shows this fact.

Example 2.3

Let X be the subset of \(\mathbb{R}^{3}\) given by

$$X=\bigl\{ (0,0,1),(1,0,0),(0,0,-1),(-1,0,0)\bigr\} . $$

We endow X with the partial order ⪯ defined by

$$(x,y,z)\preceq\bigl(x',y',z'\bigr)\quad \Longleftrightarrow\quad x\leq x',y\leq y', z\leq z'. $$

Let d be the Euclidean metric on \(\mathbb{R}^{3}\). Then \((X,d)\) is a complete metric space. Set

$$A=\bigl\{ (0,0,1),(1,0,0)\bigr\} \quad \textit{and}\quad B=\bigl\{ (0,0,-1),(-1,0,0) \bigr\} . $$

In this case, we have

$$d(A,B)=\sqrt{2},\qquad A_{0}=A,\qquad B_{0}=B. $$

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

$$T(x,y,z)=(-z,-y,-x), \quad (x,y,z)\in A. $$

Then T is continuous and \(T\in\mathcal{T}_{\xi}\) for every \(\xi\in \mathcal{Z}\). Moreover, it can be shown that all the other conditions of Theorems 2.1 and 2.2 are satisfied. However, \(z_{1}=(0,0,1)\) and \(z_{2}=(1,0,0)\) are two best proximity points of T.

In the next theorem, we give a sufficient condition for the uniqueness of the best proximity point.

Theorem 2.4

In addition to the assumptions of Theorem  2.1 (resp. Theorem  2.2), suppose that

$$\textit{for every } (x,y)\in A_{0}\times A_{0}, \textit{there is some } w\in A_{0} \textit{ such that } x\preceq w, y \preceq w. $$

Then T has a unique best proximity point.

Proof

From Theorem 2.1 (resp. Theorem 2.2), the set of best proximity points of T is not empty. Suppose that \(z_{1},z_{2}\in A_{0}\) are two distinct best proximity points of T, that is,

$$d(z_{1},Tz_{1})=d(z_{2},Tz_{2})=d(A,B), \qquad d(z_{1},z_{2})>0. $$

We consider two cases.

Case 1. If \(z_{1}\) and \(z_{2}\) are comparable.

We may assume that \(z_{1}\preceq z_{2}\). From condition (C2), we have

$$\xi\bigl(d(z_{1},z_{2}),m(z_{1},z_{2}) \bigr)\geq0, $$

where

$$m(z_{1},z_{2})=\max \biggl\{ \frac{d(z_{1},z_{1})d(z_{2},z_{2})}{ d(z_{1},z_{2})},d(z_{1},z_{2}) \biggr\} =d(z_{1},z_{2}). $$

Thus we have

$$\xi\bigl(d(z_{1},z_{2}),d(z_{1},z_{2}) \bigr)\geq0, $$

which is a contradiction with the property (ii) of a simulation function.

Case 2. If \(z_{1}\) and \(z_{2}\) are not comparable.

In this case, there is some \(w\in A_{0}\) such that

$$z_{1}\preceq w, \qquad z_{2}\preceq w,\quad w\notin \{z_{1},z_{2}\}. $$

Since \(T(A_{0})\subseteq B_{0}\), we can build a sequence \(\{w_{n}\}\subset A_{0}\) such that

$$d(w_{n+1},Tw_{n})=d(A,B),\quad n=0,1,2,\ldots $$

with \(w_{0}=w\). From condition (C1), we get

$$z_{1}\preceq w_{n}, \quad n=0,1,2,\ldots. $$

If for some k, we have \(z_{1}=w_{k}\), using condition (C1), we have \(w_{k+1}\preceq z_{1}\), which yields \(w_{k+1}=z_{1}\). Arguing similarly, we obtain \(w_{n}=z_{1}\) for every \(n\geq k\). Thus we have

$$\lim_{n\to\infty} d(w_{n},z_{1})=0. $$

If \(w_{n}\neq z_{1}\) for every n, from condition (C2), we have

$$\xi\bigl(d(z_{1},w_{n+1}),m(z_{1},w_{n}) \bigr)\geq0,\quad n=0,1,2,\ldots, $$

where

$$m(z_{1},w_{n})=\max \biggl\{ \frac{d(z_{1},z_{1})d(w_{n},w_{n+1})}{d(z_{1},w_{n})}, d(z_{1},w_{n}) \biggr\} =d(z_{1},w_{n}). $$

Thus we have

$$\xi\bigl(d(z_{1},w_{n+1}),d(z_{1},w_{n}) \bigr)\geq0, \quad n=0,1,2,\ldots. $$

On the other hand, from the property (ii) of a simulation function, we have

$$0\leq\xi\bigl(d(z_{1},w_{n+1}),d(z_{1},w_{n}) \bigr)< d(z_{1},w_{n})-d(z_{1},w_{n+1}), \quad n=0,1,2,\ldots. $$

We deduce that the sequence \(\{s_{n}\}\) defined by

$$s_{n}=d(z_{1},w_{n}),\quad n=0,1,2,\ldots $$

converges to some \(s\geq0\). But the property (ii) of a simulation function gives us that \(s=0\). Thus, in all cases, we have

$$\lim_{n\to\infty}d(w_{n},z_{1})=0. $$

Analogously, we can prove that

$$\lim_{n\to\infty}d(w_{n},z_{2})=0. $$

Finally, the uniqueness of the limit yields the desired result. □

In the following corollaries we deduce some known and some new results in best proximity point theory via various choices of simulation functions.

We denote by \(\mathcal{F}\) the set of mappings \(T: A\to B\) satisfying the following conditions:

  1. (F1)

    for every \(x_{1},x_{2},y_{1},y_{2}\in A\), we have

    $$y_{1}\preceq y_{2},\quad d(x_{1},Ty_{1})=d(x_{2},Ty_{2})=d(A,B) \quad \Longrightarrow\quad x_{1}\preceq x_{2}; $$
  2. (F2)

    for every \(x,y,u_{1},u_{2}\in A\), we have

    $$\begin{aligned}& x\preceq y, x\neq y,\quad d(u_{1},Tx)=d(u_{2},Ty)=d(A,B) \\& \quad \Longrightarrow\quad d(u_{1},u_{2})\leq k \max \biggl\{ \frac {d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} , \end{aligned}$$

    for some constant \(k\in(0,1)\).

Take \(\xi(t,s)=ks-t\), for \(t,s\geq0\), we deduce from Theorems 2.1, 2.2 and 2.4 the following results.

Corollary 2.5

Let \(T\in\mathcal{F}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    A is closed with respect to the metric d;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B), \quad x_{0}\preceq x_{1}; $$
  5. (5)

    T is continuous.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.6

Let \(T\in\mathcal{F}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    \(A_{0}\) is closed;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B), \quad x_{0}\preceq x_{1}; $$
  5. (5)

    A is \((d,\preceq)\)-regular.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.7

In addition to the assumptions of Corollary  2.5 (resp. Corollary  2.6), suppose that

$$\textit{for every } (x,y)\in A_{0}\times A_{0}, \textit{there is some } w\in A_{0} \textit{ such that } x\preceq w, y \preceq w. $$

Then T has a unique best proximity point.

We denote by \(\mathcal{G}\) the set of mappings \(T: A\to B\) satisfying the following conditions:

  1. (G1)

    for every \(x_{1},x_{2},y_{1},y_{2}\in A\), we have

    $$y_{1}\preceq y_{2},\quad d(x_{1},Ty_{1})=d(x_{2},Ty_{2})=d(A,B) \quad \Longrightarrow \quad x_{1}\preceq x_{2}; $$
  2. (G2)

    for every \(x,y,u_{1},u_{2}\in A\), we have

    $$\begin{aligned}& x\preceq y, x\neq y,\quad d(u_{1},Tx)=d(u_{2},Ty)=d(A,B) \\& \quad \Longrightarrow \quad d(u_{1},u_{2})\leq\max \biggl\{ \frac {d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} \\& \hphantom{\quad \Longrightarrow \quad d(u_{1},u_{2})\leq{}}{}-\varphi \biggl(\max \biggl\{ \frac {d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} \biggr), \end{aligned}$$

    where \(\varphi:[0,\infty)\to[0,\infty)\) is lower semi-continuous function and \(\varphi^{-1}(\{0\})=\{0\}\).

Take \(\xi(t,s)=s-\varphi(s)-t\), for \(t,s\geq0\), we deduce from Theorems 2.1, 2.2 and 2.4 the following results obtained in [23].

Corollary 2.8

Let \(T\in\mathcal{G}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    A is closed with respect to the metric d;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B), \quad x_{0}\preceq x_{1}; $$
  5. (5)

    T is continuous.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.9

Let \(T\in\mathcal{G}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    \(A_{0}\) is closed;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B), \quad x_{0}\preceq x_{1}; $$
  5. (5)

    A is \((d,\preceq)\)-regular.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.10

In addition to the assumptions of Corollary  2.8 (resp. Corollary  2.9), suppose that

$$\textit{for every } (x,y)\in A_{0}\times A_{0}, \textit{there is some } w\in A_{0} \textit{ such that } x\preceq w, y \preceq w. $$

Then T has a unique best proximity point.

We denote by \(\mathcal{H}\) the set of mappings \(T: A\to B\) satisfying the following conditions:

  1. (H1)

    for every \(x_{1},x_{2},y_{1},y_{2}\in A\), we have

    $$y_{1}\preceq y_{2}, \quad d(x_{1},Ty_{1})=d(x_{2},Ty_{2})=d(A,B) \quad \Longrightarrow\quad x_{1}\preceq x_{2}; $$
  2. (H2)

    for every \(x,y,u_{1},u_{2}\in A\), we have

    $$\begin{aligned}& x\preceq y, x\neq y,\quad d(u_{1},Tx)=d(u_{2},Ty)=d(A,B) \\& \quad \Longrightarrow \quad d(u_{1},u_{2})\leq\varphi \biggl(\max \biggl\{ \frac {d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} \biggr) \\& \hphantom{\quad \Longrightarrow \quad d(u_{1},u_{2})\leq{}}{}\times\max \biggl\{ \frac {d(x,u_{1})d(y,u_{2})}{d(x,y)},d(x,y) \biggr\} , \end{aligned}$$

    where \(\varphi:[0,\infty)\to[0,1)\) is a function such that \(\limsup_{t\to r^{+}}\varphi(t)<1\), for all \(r>0\).

Take \(\xi(t,s)=s\varphi(s)-t\), for \(t,s\geq0\), we deduce from Theorems 2.1, 2.2 and 2.4 the following results.

Corollary 2.11

Let \(T\in\mathcal{H}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    A is closed with respect to the metric d;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B),\quad x_{0}\preceq x_{1}; $$
  5. (5)

    T is continuous.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.12

Let \(T\in\mathcal{H}\). Suppose that the following conditions hold:

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    \(A_{0}\) is closed;

  3. (3)

    \(T(A_{0})\subseteq B_{0}\);

  4. (4)

    there exist \(x_{0},x_{1}\in A_{0}\) such that

    $$d(x_{1},Tx_{0})=d(A,B),\quad x_{0}\preceq x_{1}; $$
  5. (5)

    A is \((d,\preceq)\)-regular.

Then T has a best proximity point, that is, there is some \(z\in A\) such that \(d(z,Tz)=d(A,B)\).

Corollary 2.13

In addition to the assumptions of Corollary  2.11 (resp. Corollary  2.12), suppose that

$$\textit{for every } (x,y)\in A_{0}\times A_{0}, \textit{there is some } w\in A_{0} \textit{ such that } x\preceq w, y \preceq w. $$

Then T has a unique best proximity point.

Finally, take \(A=B=X\) in Theorems 2.1, 2.2 and 2.4, we obtain the following fixed point theorems.

For a given simulation function \(\xi\in\mathcal{Z}\), we denote by \(\mathcal{C}_{\xi}\) the class of mappings \(T: X\to X\) satisfying the following conditions:

  1. (I)

    for every \(x,y\in X\), we have

    $$x\preceq y \quad \Longrightarrow\quad Tx\preceq Ty; $$
  2. (II)

    for every \(x,y\in X\), we have

    $$x\preceq y, x\neq y \quad \Longrightarrow\quad \xi \biggl(d(Tx,Ty),\max \biggl\{ \frac{d(x,Tx)d(y,Ty)}{d(x,y)},d(x,y) \biggr\} \biggr)\geq0. $$

Corollary 2.14

Let \(T\in\mathcal{C}_{\xi}\), for some \(\xi\in\mathcal{Z}\). Suppose that

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    there exists some \(x_{0}\in X\) such that \(x_{0}\preceq Tx_{0}\);

  3. (3)

    T is continuous.

Then T has a fixed point, that is, there is some \(z\in X\) such that \(z=Tz\).

Corollary 2.15

Let \(T\in\mathcal{C}_{\xi}\), for some \(\xi\in\mathcal{Z}\). Suppose that

  1. (1)

    \((X,d)\) is complete;

  2. (2)

    there exists some \(x_{0}\in X\) such that \(x_{0}\preceq Tx_{0}\);

  3. (3)

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

Then T has a fixed point, that is, there is some \(z\in X\) such that \(z=Tz\).

Corollary 2.16

In addition to the assumptions of Corollary  2.14 (resp. Corollary  2.15), suppose that

$$\textit{for every } (x,y)\in X\times X, \textit{there is some } w\in X \textit{ such that } x\preceq w, y\preceq w. $$

Then T has a unique fixed point.

References

  1. Khojasteh, F, Shukla, S, Radenović, S: A new approach to the study of fixed point theory for simulation functions. Filomat 29(6), 1189-1194 (2015)

    Article  MathSciNet  Google Scholar 

  2. Argoubi, H, Samet, B, Vetro, C: Nonlinear contractions involving simulation functions in a metric space with a partial order. J. Nonlinear Sci. Appl. 8, 1082-1094 (2015)

    MathSciNet  Google Scholar 

  3. Du, WS, Khojasteh, F: New results and generalizations for approximate fixed point property and their applications. Abstr. Appl. Anal. 2014, Article ID 581267 (2014)

    MathSciNet  Google Scholar 

  4. Du, WS, Khojasteh, F, Chiu, YN: Some generalizations of Mizoguchi-Takahashi’s fixed point theorem with new local constraints. Fixed Point Theory Appl. 2014, 31 (2014)

    Article  MathSciNet  Google Scholar 

  5. Khojasteh, F, Karapinar, E, Radenović, S: Metric space: a generalization. Math. Probl. Eng. 2013, Article ID 504609 (2013)

    Article  Google Scholar 

  6. Roldán-López-de-Hierro, AF, Karapinar, E, Roldán-López-de-Hierro, C, Martínez-Moreno, J: Coincidence point theorems on metric spaces via simulation functions. J. Comput. Appl. Math. 275, 345-355 (2015)

    Article  MATH  MathSciNet  Google Scholar 

  7. Roldán-López-de-Hierro, AF, Shahzad, N: New fixed point theorem under R-contractions. Fixed Point Theory Appl. 2015, 98 (2015)

    Article  Google Scholar 

  8. Abkar, A, Gabeleh, M: Best proximity points for cyclic mappings in ordered metric spaces. J. Optim. Theory Appl. 150(1), 188-193 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  9. Basha, SS: Discrete optimization in partially ordered sets. J. Glob. Optim. 54(3), 511-517 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  10. Bilgili, N, Karapinar, E, Sadarangani, K: A generalization for the best proximity point of Geraghty-contractions. J. Inequal. Appl. 2013, 286 (2013)

    Article  MathSciNet  Google Scholar 

  11. de la Sen, M, Agarwal, RP: Some fixed point-type results for a class of extended cyclic self-mappings with a more general contractive condition. Fixed Point Theory Appl. 2011, 59 (2011)

    Article  Google Scholar 

  12. Eldred, AA, Veeramani, P: Existence and convergence of best proximity points. J. Math. Anal. Appl. 323(2), 1001-1006 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  13. Jleli, M, Karapinar, E, Samet, B: A best proximity point result in modular spaces with the Fatou property. Abstr. Appl. Anal. 2013, Article ID 329451 (2013)

    MathSciNet  Google Scholar 

  14. Jleli, M, Karapinar, E, Samet, B: A short note on the equivalence between best proximity points and fixed point results. J. Inequal. Appl. 2014, 246 (2014)

    Article  Google Scholar 

  15. Jleli, M, Samet, B: An optimization problem involving proximal quasi-contraction mappings. Fixed Point Theory Appl. 2014, 141 (2014)

    Article  Google Scholar 

  16. Karapinar, E: Fixed point theory for cyclic weak Ï•-contraction. Appl. Math. Lett. 24(6), 822-825 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  17. Karapinar, E, Pragadeeswarar, V, Marudai, M: Best proximity point for generalized proximal weak contractions in complete metric space. J. Appl. Math. 2014, Article ID 150941 (2014)

    MathSciNet  Google Scholar 

  18. Kim, WK, Lee, KH: Existence of best proximity pairs and equilibrium pairs. J. Math. Anal. Appl. 316(2), 433-446 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  19. Kirk, WA, Reich, S, Veeramani, P: Proximinal retracts and best proximity pair theorems. Numer. Funct. Anal. Optim. 24(7-8), 851-862 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  20. Nashine, HK, Kumam, P, Vetro, C: Best proximity point theorems for rational proximal contractions. Fixed Point Theory Appl. 2013, 95 (2013)

    Article  MathSciNet  Google Scholar 

  21. Raj, VS: A best proximity point theorem for weakly contractive non-self-mappings. Nonlinear Anal. 74(14), 4804-4808 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  22. Srinivasan, PS, Veeramani, P: On existence of equilibrium pair for constrained generalized games. Fixed Point Theory Appl. 2004(1), 21-29 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  23. Pragadeeswarar, V, Marudai, M: Best proximity points for generalized proximal weak contractions satisfying rational expression on ordered metric spaces. Abstr. Appl. Anal. 2015, Article ID 361657 (2015)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author extends his sincere appreciation to the Deanship of Scientific Research at King Saud University for its funding this Prolific Research group (PRG-1436-10).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bessem Samet.

Additional information

Competing interests

The author declares that he has no competing interests.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Samet, B. Best proximity point results in partially ordered metric spaces via simulation functions. Fixed Point Theory Appl 2015, 232 (2015). https://doi.org/10.1186/s13663-015-0484-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13663-015-0484-1

MSC

Keywords