You are on page 1of 9

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/316629639

Modified F-contractions via α-admissible


mappings and application to integral equations

Article · January 2017


DOI: 10.2298/FIL1705141A

CITATIONS READS

0 54

3 authors:

Hassen Aydi Erdal Karapinar


University of Dammam Atilim University
143 PUBLICATIONS 2,646 CITATIONS 349 PUBLICATIONS 4,928 CITATIONS

SEE PROFILE SEE PROFILE

Habib Yazidi
Ecole Supérieure des Sciences et Techniques …
20 PUBLICATIONS 121 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Scientific Hypothesis as a source of innovation and discovery View project

Chapter on a Book View project

All content following this page was uploaded by Hassen Aydi on 10 May 2017.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Filomat 31:5 (2017), 1141–1148 Published by Faculty of Sciences and Mathematics,
DOI 10.2298/FIL1705141A University of Niš, Serbia
Available at: http://www.pmf.ni.ac.rs/filomat

Modified F-Contractions via α-Admissible Mappings


and Application to Integral Equations

Hassen Aydia,b , Erdal Karapinarc , Habib Yazidid


a University of Dammam, Department of Mathematics. College of Education of Jubail, P.O: 12020, Industrial Jubail 31961. Saudi Arabia.
b Department of Medical Research, China Medical University Hospital, China Medical University, Taichung, Taiwan.
c Atilim University, Department of Mathematics, 06836, İncek, Ankara, Turkey.
d Jizan University, College of Sciences, Department of Mathematics, Saudi Arabia.

Abstract. In this paper, we introduce the concept of a modified F-contraction via α-admissible mappings
and propose some theorems that guarantee the existence and uniqueness of fixed point for such mappings
in the frame of complete metric spaces. We also provide some illustrative examples. Moreover, we consider
an application solving an integral equation.

1. Introduction and Preliminaries

In 2012, Wardowski [15] defined a new concept of F-contractions as follows.


Definition 1.1. Let (X, d) be a metric space. A self-mapping T : X → X is said to be an F-contraction if there exists
τ > 0 such that

d(Tx, Ty) > 0 ⇒ τ + F(d(Tx, Ty)) ≤ F(d(x, y)), (1)

for all x, y ∈ X where F : R+ → R is a mapping satisfying the following conditions:


(F1 ) F is increasing, i.e., for all α, β ∈ R+ such that α < β, F(α) < F(β),
(F2 ) For any sequence {αn }∞
n=1
of positive real numbers, lim αn = 0 if and only if lim F(αn ) = −∞.
n→∞ n→∞

(F3 ) There exists k ∈ (0, 1) such that lim+ α F(α) = 0.


k
α→0

We set F as the collections of all functions F satisfying (F1 )-(F3 ). Consider F(y) = ln(y), G(y) = ln(y) + y and
H(y) = − √1y for y > 0. It is clear that F, G, H ∈ F. For more examples and details, see e.g. [15].
In what follows we recollect the main result of Wardowski [15] which is a generalization of the Banach
Contraction Mapping Principle [5]:

2010 Mathematics Subject Classification. Primary 47H10; Secondary 54H25.


Keywords. α-admissible mappings; F-contraction; fixed point; integral equation.
Received: 20 January 2015; Accepted: 13 June 2016
Communicated by Ljubomir Ćirić
Corresponding author: Hassen Aydi
Email addresses: hmaydi@uod.edu.sa (Hassen Aydi), erdalkarapinar@yahoo.com, ekarapinar@atilim.edu.tr (Erdal
Karapinar), yazidi2001@yahoo.com (Habib Yazidi)
H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1142

Theorem 1.1. Let (X, d) be a complete metric space and T : X → X be an F-contraction. Then T has a unique fixed
point.
The concept of an F-contraction as well as Theorem 1.1 have been generalized in many directions, see e.g.
[8, 9, 12, 13].
Let Ψ be the family of functions ψ : [0, ∞) → [0, ∞) satisfying the following conditions:
i) ψ is nondecreasing;
+∞
X
(ii) ψn (t) < ∞ for all t > 0.
n=1

If ψ ∈ Ψ, then it is called c-comparison function. It is easy to show that ψ(t) < t for all t > 0 and ψ is
continuous at 0.
In 2012, Samet et al. [11] introduced the class of α-admissible mappings.

Definition 1.2. [11] Let α : X × X → [0, ∞) be given mapping where X , ∅. A selfmapping T is called α-admissible
if for all x, y ∈ X, we have

α(x, y) ≥ 1 =⇒ α(Tx, Ty) ≥ 1. (2)

In what follows we extend the notion of F-contraction.


Definition 1.3. Let (X, d) be a metric space. A self-mapping T : X → X is said to be a modified F-contraction via
α-admissible mappings if there exists τ > 0 such that

d(Tx, Ty) > 0 ⇒ τ + F(α(x, y)d(Tx, Ty)) ≤ F(ψ(d(x, y))), (3)

for all x, y ∈ X, where the mapping F ∈ F and ψ ∈ Ψ.


If we let F(t) = ln(t) for t > 0, the contraction form (3) becomes

α(x, y)d(Tx, Ty) ≤ e−τ ψ(d(x, y)) ≤ ψ(d(x, y)) for all x, y ∈ X, Tx , Ty. (4)

(4) is considered as an α − ψ-contraction which was introduced by Samet et al. [11].


In this paper, we prove some fixed point results of certain contractions whose frames are drawn above.
The obtained theorems shall be supported by a concrete example. An application of the observed result is
considered in the frame of integral equation theory.

2. Main Results

The following theorem is the first main result of this paper:


Theorem 2.1. Let (X, d) be a complete metric space and T : X → X be a modified F-contraction via α-admissible
mappings. Suppose that
(i) T is α-admissible;
(ii) there exists x0 ∈ X such that α(x0 , Tx0 ) ≥ 1;
(iii) T is continuous.
Then T has a fixed point.
Proof. By assumption (ii), there exists a point x0 ∈ X such that α(x0 , Tx0 ) ≥ 1. We define a sequence {xn } in X
by xn+1 = Txn = Tn+1 x0 for all n ≥ 0. Suppose that xn0 = xn0 +1 for some n0 . So the proof is completed. Now,
we assume that

xn , xn+1 for all n. (5)


H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1143

Since α(x0 , x1 ) = α(x0 , Tx0 ) ≥ 1 and T is α-admissible, we get

α(xn , xn+1 ) ≥ 1, for all n = 0, 1, . . . . (6)

From (3) and (5), we have

τ + F(α(xn−1 , xn )d(Txn−1 , Txn )) ≤ F(ψ(d(xn−1 , xn ))).

On account of (F1 ) and (6), we find

τ + F(d(xn , xn+1 )) ≤ F(d(xn−1 , xn )) for all n ≥ 1.

By letting dn = d(xn , xn+1 ), the inequalities above infer that

F(dn ) ≤ F(dn−1 ) − τ ≤ F(d0 ) − nτ for all n ≥ 1.

Consequently, we obtain

lim F(dn ) = −∞.


n→∞

By a property (F2 ), we have

lim dn = 0. (7)
n→∞

Now, due to (F3 ), we have lim dkn F(dn ) = 0, where k ∈ (0, 1). By (7), the following holds for all n ≥ 0
n→∞

0 ≤ dkn F(dn ) − dkn F(d0 ) ≤ dkn (F(d0 ) − nτ) − dkn F(d0 ) = −nτdkn ≤ 0. (8)

Letting n → ∞ in (8), we find that

lim ndkn = 0.
n→∞

So there exists n1 ∈ N such that dn ≤ 1


1 , for all n ≥ n1 . For m, n ∈ N with m > n ≥ n1 , we have
nk


X 1
d(xn , xm ) ≤ dn + dn+1 + ... + dm−1 ≤ 1
.
i=n ik
X 1
Since 1
converges, the sequence {xn } is Cauchy in (X, d). From the completeness of X, there exists u ∈ X
ik
i≥1
such that lim xn = u. Finally, the continuity of T yields Tu = u, which completes the proof.
n→∞

Theorem 2.1 remains true if we replace the continuity hypothesis by the following property:

(H) If {xn } is a sequence in X such that α(xn , xn+1 ) ≥ 1 for all n and xn → x ∈ X as n → ∞, then there exists a
subsequence {xn(k) } of {xn } such that α(xn(k) , x) ≥ 1 for all k.

Theorem 2.2. Let (X, d) be a complete metric space and T : X → X be a modified F-contraction via α-admissible
mappings. Suppose that
(i) T is α-admissible;
(ii) there exists x0 ∈ X such that α(x0 , Tx0 ) ≥ 1;
(iii) (H) holds.
H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1144

Then there exists u ∈ X such that Tu = u.


Proof. Following the lines in the proof of Theorem 2.1, we construct a sequence {xn } in (X, d) which is Cauchy
and converges to some u ∈ X.
Suppose that there exists an increasing sequence {n(k)} ⊂ N such that xn(k) = Tu for all k ∈ N. Letting
k → ∞, by the uniqueness of the limit, we find Tu = u. Hence, the proof is completed. As a result, we
shall assume that there exists k0 ∈ N such that xn(k) , Tu for all k ∈ N with k ≥ k0 . Consequently, we have
Txn(k)−1 , Tu for all k ≥ k0 .
Therefore, by (3), we have
τ + F(α(xn(k)−1 , u)d(Txn(k)−1 , Tu)) ≤ F(ψ(d(xn(k)−1 , u))).
Regarding α(xn(k)−1 , x) ≥ 1 and (F1 ),
d(xn(k) , Tu) = d(Txn(k)−1 , Tu) ≤ ψ(d(xn(k)−1 , u)).
Since ψ is continuous at 0 and d(xn(k)−1 , u) → 0,
lim ψ(d(xn(k)−1 , u)) = 0.
k→∞

Thus, lim d(xn(k)+1 , Tu) = 0. By the uniqueness of limit, Tu = u.


k→∞

We provide the following example.


Example 2.1. Take X = {0, 1, 2} and T : X → X such that
T0 = 0 and T1 = T2 = 1. (9)
Consider α : X × X → [0, ∞) as
α(1, 2) = α(2, 1) = α(1, 1) = 1,
and 0 otherwise. Clearly, if α(x, y) = 1, α(Tx, Ty) = 1. Then T is α-admissible. Notice also that α(1, T1) = α(1, 1) =
1.
Let x, y ∈ X such that Tx , Ty, so (x, y) is equal to (0, 1), (0, 2), (1, 0) or (2, 0). For these four cases, α(x, y) = 0,
so (4) holds. In other words, (3) holds for F(t) = ln(t) and for any ψ ∈ Ψ and any metric d. It is obvious also that the
hypothesis (H) is satisfied. Thus, applying Theorem 2.2, the mapping T has a fixed point. Here, we have two fixed
points which are u = 0 and u = 1.
Here, we underline the fact that the mapping considered in above example has two fixed points, 0 and 1.
Notice also that α(0, 1) = 0 < 1. For the uniqueness, we need an additional condition:
(U) For all x, y ∈ Fix(T), we have α(x, y) ≥ 1, where Fix(T) denotes the set of fixed points of T.
Theorem 2.3. Adding condition (U) to the hypotheses of Theorem 2.1 (resp. Theorem 2.2), we obtain that u is the
unique fixed point of T.
Proof. We shall prove the uniqueness by the method of Reductio and Absurdum.
Suppose, on the contrary, that there exist u, v ∈ X such that u = Tu and v = Tv with u , v. Then Tu , Tv,
so by (3), we get
τ + F(α(u, v)d(Tu, Tv)) ≤ F(ψ(d(u, v))),
that is,
τ + F(α(u, v)d(u, v)) ≤ F(ψ(d(u, v))) < F(d(u, v)).
Again by (F1 ), we have
τ + F(d(u, v)) ≤ F(ψ(d(u, v))) < F(d(u, v)),
which is a contradiction. Thus, u = v which completes the proof.
H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1145

The following corollaries are immediate.

Corollary 2.1. Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose there exists τ > 0
such that

d(Tx, Ty) > 0 ⇒ τ + F(d(Tx, Ty)) ≤ F(ψ(d(x, y))), (10)

for all x, y ∈ X where F satisfies (F1 ) − (F3 ). Then T has a unique fixed point.

Proof. It sufficient to take α(x, y) = 1 in Theorem 2.3.

Corollary 2.2. Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose there exists τ > 0
such that

d(Tx, Ty) > 0 ⇒ τ + F(d(Tx, Ty)) ≤ F(cd(x, y))), (11)

for all x, y ∈ X where F satisfies (F1 ) − (F3 ) and c ∈ (0, 1). Then T has a unique fixed point.

Proof. It follows from Corollary 2.1 with ψ(t) = ct.

The investigation of existence of fixed points on metric spaces endowed with a partial order was initiated
by Turinici [14]. Then, several interesting and valuable results appeared in this direction, see e.g. [1–4, 7, 10].

Definition 2.1. Let (X, ) be a partially ordered set and T : X → X be a given mapping. We say that T is
nondecreasing with respect to  if
x, y ∈ X, x  y =⇒ Tx  Ty.
Furthermore, a sequence {xn } ⊂ X is said to be nondecreasing with respect to  if xn  xn+1 for all n.

Definition 2.2. Let (X, ) be a partially ordered set and d be a metric on X. We say that (X, , d) is regular if for
every nondecreasing sequence {xn } ⊂ X such that xn → x ∈ X as n → ∞, there exists a subsequence {xn(k) } of {xn }
such that xn(k)  x for all k.

Under the set-up of partially ordered metric spaces, we have the following result.

Corollary 2.3. Let (X, ) be a partially ordered set and d be a metric on X such that (X, d) is complete. Let T : X → X
be a nondecreasing mapping with respect to . Suppose that there exist τ > 0, ψ ∈ Ψ and F ∈ F such that

τ + F(d(Tx, Ty)) ≤ F(ψ(d(x, y)),

for x, y ∈ X with x  y and Tx , Ty. Suppose also that the following conditions hold:

(i) there exists x0 ∈ X such that x0  Tx0 ;

(ii) either T is continuous

(ii)’ or (X, , d) is regular.

Then T has a fixed point.

Proof. Define the mapping α : X × X → [0, ∞) by


(
1 if x  y or x  y,
α(x, y) =
0 otherwise.
Clearly, T is a modified F-contractive mapping via α-admissible mappings, that is,

τ + F(α(x, y)d(Tx, Ty)) ≤ F(ψ(d(x, y))),


H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1146

for all x  y with Tx , Ty. From condition (i), we have α(x0 , Tx0 ) ≥ 1. Moreover, for all x, y ∈ X, from the
monotone property of T, we have

α(x, y) ≥ 1 =⇒ x  y or x  y =⇒ Tx  Ty or Tx  Ty =⇒ α(Tx, Ty) ≥ 1.

Thus, T is α-admissible. Now, if T is continuous, then the existence of a fixed point follows from Theorem
2.1. Suppose now that (X, , d) is regular. Let {xn } be a sequence in X such that α(xn , xn+1 ) ≥ 1 for all n and
xn → x ∈ X as n → ∞. From the regularity hypothesis, there exists a subsequence {xn(k) } of {xn } such that
xn(k)  x for all k. This implies from the definition of α that α(xn(k) , x) ≥ 1 for all k. In this case, the existence
of a fixed point follows from Theorem 2.2.

Now, we present an example which guarantees the uniqueness of the fixed point.

Example 2.2. Let X = [0, ∞) and d(x, y) = |x − y| for all x, y ∈ X. Take τ > 0. Consider the mapping T : X → X
given by

3
e−τ 4 x if x ∈ [0, 1],

Tx = 

e−τ 34 if x > 1.

T is continuous in (X, d). Define the mapping α : X × X → [0, ∞) by


(
1 if x, y ∈ [0, 1],
α(x, y) =
0 otherwise.
Consider the function ψ : [0, ∞) → [0, ∞) by
( 3
t if t ∈ [0, 1],
ψ(t) = 42
5 t otherwise.

Let x, y ∈ X such that α(x, y) ≥ 1, so x, y ∈ [0, 1]. Then Tx, Ty ∈ [0, 1], that is, α(Tx, Ty) = 1. Hence, T is
α-admissible. Mention that ψ ∈ Ψ and α(0, T0) = 1. In the case where x, y ∈ [0, 1] such that Tx , Ty, we have
3
α(x, y)d(Tx, Ty) = d(Tx, Ty) = e−τ |x − y| ≤ e−τ ψ(d(x, y).
4
In the other case where x or y is not in [0, 1], α(x, y) = 0, so the above inequality is satisfied for all x, y ∈ X with
Tx , Ty. Thus, (3) is satisfied with F(t) = ln(t) for t > 0. Moreover, it is easy that the hypothesis (U) is true. Thus,
applying Theorem 2.3, the mapping T has a unique fixed point, which is u = 0.

3. Application

In this section, we present an application on existence of a solution of an integral equation. In particular,


inspired from [6] and using Corollary 2.3, we will prove the existence of a solution of the following integral
equation.
Z 1
x(t) = 1(t) + S(t, s) K(t, x(s)) ds, (12)
0

where K : [0, 1] × R → R and 1 : [0, 1] → R are continuous functions, and S : [0, 1] × [0, 1] → [0, ∞) is a
function such that S(t, .) ∈ L1 ([0, 1]) for all t ∈ [0, 1].
Throughout this section, let X = C([0, 1], R) be the set of real continuous functions defined on [0, 1]. Take
the metric d : X × X → [0, ∞) be defined by

d(x, y) = kx − yk∞ = max |x(s) − y(s)|.


s∈[0,1]
H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1147

It is known that (X, d) is a complete metric space.


Now, take the operator T : X → X defined by
Z 1
Tx(t) = 1(t) + S(t, s) K(t, x(s)) ds. (13)
0

Mention that (12) has a solution if and only if the operator T has a fixed point.

Theorem 3.1. Assume that


• (i) there exist λ ∈ (0, 1), τ > 0, ν : X × X → [0, ∞) and α : X × X → [0, ∞) such that if α(x, y) ≥ 1 for all
x, y ∈ X then, for every s ∈ [0, 1], we have

0 ≤ |K(s, x(s)) − K(s, y(s))| ≤ ν(x, y)|x(s) − y(s)|,

and
Z 1
k S(t, s)ν(x, y) dsk∞ ≤ λ e−τ .
0

• (ii) K is non-decreasing with respect to its second variable;


R1
• (iii) there exists x0 ∈ X such that x0 (t)  1(t) + 0 S(t, s) K(t, x(s)) ds.
Then T has a fixed point in X.

Proof. Define the mapping α : X × X → [0, ∞) by


(
1 if x  y,
α(x, y) =
0 otherwise.
Take ψ(t) = λt. We define

x, y ∈ X, x  y if and only if x(t) ≤ y(t) for all t ∈ [0, 1],

where ≤ denotes the usual order of real numbers. If x  y, then α(x, y) ≥ 1. By condition (i)
Z 1
|A(x)(t) − A(y)(t)| ≤ S(t, s)|K(t, s, x(s)) − K(t, s, y(s))| ds
0
Z 1
≤ S(t, s)ν(x, y)|x(s) − y(s)| ds
0
Z 1
≤ kx − yk∞ S(t, s)ν(x, y) ds
0
≤ e−τ λ kx − yk∞
= e−τ ψ(kx − yk∞ ).

We deduce for all x, y ∈ X such that x  y and Tx , Ty

α(x, y)d(Tx, Ty) = d(Tx, Ty) = kTx − Tyk∞ ≤ e−τ ψ(d(x, y)). (14)

It is clear that if Tx = Ty, then (14) holds.


Since K is non-decreasing with respect to its second variable, so for all x, y ∈ X with x  y, we get
Tx(t) ≤ T(y)(t) for all t ∈ [0, 1], that is, if α(x, y) ≥ 1, α(Tx, Ty) ≥ 1. Moreover, the condition (iii) yields that
there exists x0 ∈ X such that α(x0 , Tx0 ) ≥ 1. Therefore, the integral operator A satisfies all the hypotheses of
Corollary 2.3 with F(t) = ln(t) for t > 0. Consequently, T has a fixed point, that is, the integral equation (12)
has a solution x ∈ X.
H. Aydi et al. / Filomat 31:5 (2017), 1141–1148 1148

References
[1] H. Aydi, Coincidence and common fixed point results for contraction type maps in partially ordered metric spaces, Int. Journal
of Math. Analysis, 5 (13) 631-642 (2011).
[2] H. Aydi, E. Karapınar, W. Shatanawi, Coupled coincidence points in partially ordered cone metric spaces with a c-distance,
Journal of Applied Mathematics, Volume 2012, Article ID 312078, 15 pages.
[3] H. Aydi, B. Damjanović, B. Samet, W. Shatanawi, Coupled fixed point theorems for nonlinear contractions in partially ordered
G-metric spaces, Mathematical and Computer Modelling, (54) 2443-2450 (2011).
[4] H. Aydi, W. Shatanawi, M. Postolache, Coupled fixed point results for (ψ, φ)-weakly contractive mappings in ordered G-metric
spaces, Computers and Mathematics with Applications, (63) 298-309 (2012).
[5] B. Banach, Sur les opérations dans les ensembles abstraits et leur application aux équations intégrales. Fundam. Math. 3, 133-181
(1922).
[6] M. Jleli, B. Samet, C. Vetro, F. Vetro, Fixed points for multivalued mappings in b-metric spaces, in press.
[7] J.J. Nieto and R. Rodrı́guez-López, Contractive Mapping Theorems in Partially Ordered Sets and Applications to Ordinary
Differential Equations. Order. 22 223-239, (2005).
[8] D. Paesano, C. Vetro, Multi-valued F-contractions in 0-complete partial metric spaces with application to Volterra type integral
equation, RACSAM, DOI 10.1007/s13398-013-0157-z.
[9] H. Piri, P. Kumam, Some fixed point theorems concerning F-contraction in complete metric spaces, Fixed Point Theory Appl.
2014, 2014:210.
[10] A.C.M. Ran and M.C.B. Reurings, A fixed point theorem in partially ordered sets and some applications to matrix equations,
Proc. Amer. Math. Soc. 132 (2003) 1435–1443.
[11] B. Samet, C. Vetro and P. Vetro, Fixed point theorems for α-ψ-contractive type mappings, Nonlinear Anal. 75 2154-2165 (2012).
[12] S. Shukla, S. Radenović, Some common fixed point theorems for F-contraction type mappings in 0-complete partial metric spaces,
J. Ineq. Appl. 2014.
[13] M. Sgroi, C. Vetro, Multi-valued F-contractions and the solution of certain functional and integral equations. Filomat. (27)
1259-1268 (2013).
[14] M. Turinici, Abstract comparison principles and multivariable Gronwall-Bellman inequalities, J. Math. Anal. Appl. (117) 100–127
(1986).
[15] D. Wardowski, Fixed points of a new type of contractive mappings in complete metric spaces. Fixed Point Theory Appl. 2012, 94
(2012).

View publication stats

You might also like