Geraghty-type theorems in modular metric spaces with an application to partial differential equation

12
RESEARCH Open Access Geraghty-type theorems in modular metric spaces with an application to partial differential equation Parin Chaipunya 1 , Yeol Je Cho 2* and Poom Kumam 1* * Correspondence: [email protected]; [email protected] 1 Department of Mathematics, Faculty of Science, King Mongkuts University of Technology Thonburi, Bangkok 10140, Thailand 2 Department of Mathematics Education and the RINS, Gyeongsang National University, Chinju 660-701, Korea Full list of author information is available at the end of the article Abstract In this article, we prove some fixed point theorems of Geraghty-type concerning the existence and uniqueness of fixed points under the setting of modular metric spaces. Also, we give an application of our main results to establish the existence and uniqueness of a solution to a nonhomogeneous linear parabolic partial differential equation in the last section. Mathematics Subject Classification (2010): 47H10, 54H25, 35K15. Keywords: modular metric space, ordered set, Geraghtys theorem, initial value pro- blems, parabolic equation. Introduction and preliminaries Throughout this article, let + denote the set of all positive real numbers and let + denote the set of all nonnegative real numbers. Since the year 1922, Banachs contraction principle, due to its simplicity and applic- ability, has became a very popular tool in modern analysis, especially in nonlinear ana- lysis including its applications to differential and integral equations, variational inequality theory, complementarity problems, equilibrium problems, minimization pro- blems and many others. Also, many authors have improved, extended and generalized this contraction principle in several ways (see e.g. [1-10]). In 1973, Geraghty [11] gave an interesting generalization of the contraction principle using the class S of the functions b: + ® [0, 1) satisfying the following condition: β (t n ) 1 implies t n 0. Theorem 1.1. [11]Let (X, d) be a complete metric space and f be a self-mapping on X such that there exists β S satisfying d(fx, fy) β (d(x, y))d(x, y) (1:1) for all x, y Î X. Then the sequence {x n } defined by x n = fx n-1 for each n 1 converges to the unique fixed point of f in X. Later, Amini-Harandini et al. [12] extended Geraghtys fixed point theorem to the setting of partially ordered metric spaces as follows: Chaipunya et al. Advances in Difference Equations 2012, 2012:83 http://www.advancesindifferenceequations.com/content/2012/1/83 © 2012 Chaipunya et al; licensee Springer. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Transcript of Geraghty-type theorems in modular metric spaces with an application to partial differential equation

RESEARCH Open Access

Geraghty-type theorems in modular metricspaces with an application to partial differentialequationParin Chaipunya1, Yeol Je Cho2* and Poom Kumam1*

* Correspondence: [email protected];[email protected] of Mathematics,Faculty of Science, King Mongkut’sUniversity of Technology Thonburi,Bangkok 10140, Thailand2Department of MathematicsEducation and the RINS,Gyeongsang National University,Chinju 660-701, KoreaFull list of author information isavailable at the end of the article

Abstract

In this article, we prove some fixed point theorems of Geraghty-type concerning theexistence and uniqueness of fixed points under the setting of modular metric spaces.Also, we give an application of our main results to establish the existence anduniqueness of a solution to a nonhomogeneous linear parabolic partial differentialequation in the last section.Mathematics Subject Classification (2010): 47H10, 54H25, 35K15.

Keywords: modular metric space, ordered set, Geraghty’s theorem, initial value pro-blems, parabolic equation.

Introduction and preliminariesThroughout this article, let ℝ+ denote the set of all positive real numbers and let ℝ+

denote the set of all nonnegative real numbers.

Since the year 1922, Banach’s contraction principle, due to its simplicity and applic-

ability, has became a very popular tool in modern analysis, especially in nonlinear ana-

lysis including its applications to differential and integral equations, variational

inequality theory, complementarity problems, equilibrium problems, minimization pro-

blems and many others. Also, many authors have improved, extended and generalized

this contraction principle in several ways (see e.g. [1-10]).

In 1973, Geraghty [11] gave an interesting generalization of the contraction principle

using the class S of the functions b: ℝ+ ® [0, 1) satisfying the following condition:

β(tn) → 1 implies tn → 0.

Theorem 1.1. [11]Let (X, d) be a complete metric space and f be a self-mapping on X

such that there exists β ∈ S satisfying

d(fx, fy) ≤ β(d(x, y))d(x, y) (1:1)

for all x, y Î X. Then the sequence {xn} defined by xn = fxn-1 for each n ≥ 1 converges

to the unique fixed point of f in X.

Later, Amini-Harandini et al. [12] extended Geraghty’s fixed point theorem to the

setting of partially ordered metric spaces as follows:

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

© 2012 Chaipunya et al; licensee Springer. This is an Open Access article distributed under the terms of the Creative CommonsAttribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction inany medium, provided the original work is properly cited.

Theorem 1.2. [12]Let (X, ⊑) be a partially ordered metric set and suppose that there

exists a metric d in X such that (X, d) is a complete metric space. Let f be a nonde-

creasing self-mapping on X which satisfies the inequality (1.1) whenever x, y Î X are

comparable. Assume that f is either continuous or

if a nondecreasing sequence {xn} converges to x∗, then xn � x∗for each n ≥ 1. (1:2)

If, additionally, the following condition is satisfied:

for any x, y ∈ X, there exists z ∈ X which is comparable to both x and y, (1:3)

then the sequence {xn} converges to the unique fixed point of f in X.

Let Ψ denote the class of functions ψ: ℝ+ ® ℝ+ satisfying the following conditions:

(a) ψ is nondecreasing.

(b) ψ is continuous.

(c) ψ(t) = 0 if and only if t = 0.

Using this class, Eshaghi Gordji et al. [13] extended the Theorem 1.2 as follows:

Theorem 1.3. [13]Let (X, ⊑) be a partially ordered metric set and suppose that there

exists a metric d in X such that (X, d) is a complete metric space. Let f be a nonde-

creasing self-mapping on X such that there exists x0 Î X with x0 ⊑ fx0. Suppose that

there exist β ∈ S and ψ Î Ψ such that

ψ(d(fx, fy)) ≤ β(ψ(d(x, y)))ψ(d(x, y)),

whenever x, y Î X are comparable. Assume also that the condition (1.2) holds. Then f

has a fixed point.

On the other hand, in 2010, Chistyakov [14] introduced the notion of a modular

metric space which is raised in an attempt to avoid some restrictions of the concept of

a modular space (for the literature of a modular space, see e.g. [15-21] and references

therein). Some of the early investigations on metric fixed point theory in this space

refer to [22-24].

For the rest of this section, we present some notions and basic facts of modular

metric spaces.

Definition 1.4. [14] Let X be a nonempty set. A function ω: ℝ+ × X × X ® ℝ+ ∪ {∞}

is said to be a metric modular on X if, for all x, y, z Î X, the following conditions hold:

(a) ωl(x, y) = 0 for all l >0 if and only if x = y.

(b) ωl(x, y) = ωl(y, x) for all l >0.

(c) ωl+μ(x, y) ≤ ωl(x, z) + ωμ(z, y) for all l, μ >0.

For any xι Î X, the set Xω(xι) = {x Î X: liml®∞ ωl(x, xι) = 0} is called a modular

metric space generated by xι and induced by ω. If its generator xι does not play any

role in the situation (that is, Xω is independent of generators), we write Xω instead of

Xω(xι).

Observe that a metric modular ω on X is nonincreasing with respect to l >0. We

can simply show this assertion using the condition (c). For any x, y Î X and 0 < μ < l,we have

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 2 of 12

ωλ(x, y) ≤ ωλ−μ(x, x) + ωμ(x, y) = ωμ(x, y). (1:4)

For any x, y Î X and l >0, we set

ωλ+ (x, y) := limε↓0

ωλ+ε(x, y), ωλ−(x, y) := limε↓0

ωλ−ε(x, y).

Consequently, from (1.4), it follows that

ωλ+(x, y) ≤ ωλ(x, y) ≤ ωλ−(x, y).

For any x, y Î X, if a metric modular ω on X possesses a finite value and ωl(x, y) =

ωμ(x, y) for all l, μ >0, then d(x, y): = ωl(x, y) is a metric on X.

Later, Chaipunya et al. [23] has altered the notion of convergent and Cauchy

sequences in modular metric spaces under the direction of Mongkolkeha et al. [24].

Definition 1.5. [23,24] Let Xω be a modular metric space and {xn} be a sequence

in Xω.

(1) A point x Î Xω is called a limit of {xn} if, for each l, � >0, there exists n0 Î N

such that ωl(xn, x) < � for all n ≥ n0. A sequence that has a limit is said to be con-

vergent (or converges to x), which is written as limn®∞ xn = x.

(2) A sequence {xn} in Xω is said to be a Cauchy sequence if, for each l, � >0, there

exists n0 Î N such that ωl(xn, xm) < � for all m, n ≥ n0.

(3) If every Cauchy sequences in X converges, X is said to be complete.

In this article, we prove a generalization of Geraghty’s theorem which also improves

the result of Eshagi Gordji et al. [13] under the influence of a modular metric space.

An application to partial differential equation is also provided.

Main resultsBefore stating our main results, we first introduce the following classes for a more con-

venience of usage.

For each n Î N, let Sn denote the class of n-tuples of functions (b1, b2, ..., bn), wherefor each i Î {1, 2, ..., n}, bi: ℝ+ ∪ {∞} ® [0, 1) and the following implication holds:

β(tk) := β1(tk) + β2(tk) + · · · + βn(tk) → 1 implies tk → 0.

Actually, Geraghty’s class S is equivalent to the class S1 when ∞ is not considered.

It follows that, for each m Î {1, 2, ..., n}, if (β1, β2, . . . , βm) ∈ Sm , then

(β1, β2, . . . , βm, θ , θ , . . . , θ︸ ︷︷ ︸n−m entries

) ∈ Sn , where θ denotes the zero function. Also, note that, if

(β , β , . . . , β)︸ ︷︷ ︸n entries

∈ Sn, then we also have the following:

β(tk) → 1n

implies tk → 0.

Besides, if (β1, β2 . . . , βn) ∈ Sn , then π((β1, β2 . . . , βn)) ∈ Sn , where π((b1, b2 ..., bn))is a permutation of (b1, b2 ..., bn). It is also important to know that, if

(β1, β2 . . . , βn) ∈ Sn , then (βn1 , βn2 , . . . , βnm ) ∈ Sm for each m Î {1, 2, ..., n}, where

each βni is selected from {b1, b2, ..., bn} and βni = βnj for all i, j Î {1, 2, ..., m}.

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 3 of 12

Let ̄ denote the class of functions ψ: ℝ+ ∪ {∞} ® ℝ+ ∪ {∞} satisfying the following

conditions:

(a) If 0 < t < ∞, then ψ(t) < ∞.

(b) ψ |R+ ∈ .

Now, we are ready to give our main results in this article.

Theorem 2.1. Let Xω be a complete modular metric space with a partial ordering ⊑and f be a self-mapping on Xω such that, for each l >0, there exists h(l) Î (0, l) suchthat

ψ(ωλ(fx, fy)) ≤α(ψ(ωλ(x, y)))ψ(ωλ+η(λ)(x, y)) + β(ψ(ωλ(x, y)))ψ(ωλ(x, fx))

+ γ (ψ(ωλ(x, y)))ψ(ωλ(y, fy)),(2:1)

where ψ ∈ ̄ and (α, β , γ ) ∈ S3 with a(t) + 2 max{supt≥0 b(t), supt≥0 g(t)} <1.

Assume also that the condition (1.2) holds. If there exists x0 Î Xω such that ωl(x0, fx0)

< ∞ for all l >0, then the following hold:

(1) f has a fixed point x∞ Î Xω.

(2) The sequence {fnx0} converges to x∞.

Proof. It is clear that the sequence {fnx0} is nondecreasing. Assume that, for each n ≥ 1,

there exists ln >0 such that ωλn (f nx0, f n+1x0) = 0 . Otherwise, the proof is complete. For

each n ≥ 1, if 0 < l ≤ ln, then we also have ωl(fnx0, f

n+1x0) ≠ 0. Since fnx0⊑f n+1x0, forany 0 < l ≤ ln, we have

ψ(ωλn(f nx0, f n+1x0)) ≤ ψ(ωλ(f nx0, f n+1x0))

≤ α(ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ+η(λ)(f n−1x0, f nx0))

+ β(ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ(f n−1x0, f nx0))

+ γ (ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ(f nx0, f n+1x0))

≤ α(ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ(f n−1x0, f nx0))

+ β(ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ(f n−1x0, f nx0))

+ γ (ψ(ωλ(f n−1x0, f nx0)))ψ(ωλ(f nx0, f n+1x0)),

which implies that

ψ(ωλ(f nx0, f n+1x0)) ≤ α(ψ(ωλ(f n−1x0,f nx0)))+β(ψ(ωλ(f n−1x0,f nx0)))1−γ (ψ(ωλ(f n−1x0,f nx0))) ψ(ωλ(f n−1x0, f nx0))

≤ ψ(ωλ(f n−1x0, f nx0))

...

≤ ψ(ωλ(x0, f x0)

< ∞.

Therefore, {ψ(ωl(fnx0, f

n+1x0))} is nonincreasing and bounded below. So, the

sequence converges to some number r ≥ 0. Assume r >0. Observe that

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 4 of 12

ψ(ωλ(f nx0, f n+1x0)) ≤ [α(ψ(ωλ(f n−1x0, f nx0))) + β(ψ(ωλ(f n−1x0, f nx0)))

+γ (ψ(ωλ(f n−1x0, f nx0)))]ψ(ωλ(f n−1x0, f nx0)).

Taking n ® ∞, we have

1 ≤ lim infn→∞

[α(ψ(ωλ(f n−1x0, f nx0))) + β(ψ(ωλ(f n−1x0, f nx0)))

+γ (ψ(ωλ(f n−1x0, f nx0)))]

.

So, we have limn®∞ ψ(ωl(fn-1x0, f

nx0)) = 0 and hence

limn→∞ ωλ(f n−1x0, f nx0) = 0,

which is a contradiction of our assumption. Therefore, limn®∞ ψ(ωl(fnx0, f

n+1x0)) = 0 and

so, we have limn®∞ ωl(fnx0, f

n+1x0) = 0. Moreover, we have limn®∞ ωl(fnx0, f

n+1x0) = 0 for

all l >0.

Next, we show that {fnx0} is a Cauchy sequence. Assume the contrary. So, there

exists l0, �0 >0 for which we can define two subsequences {f mkx0} and {f nkx0} of the

sequence {fnx0} such that, for any nk > mk > k, ωλ0(f mkx0, f nk x0) ≥ ε0 , but

ωλ0(f mkx0, f nk−1x0) < ε0 . Now, since f mkx0 � f nkx0 , we observe that

ψ(ε0) ≤ ψ(ωλ0 (f mkx0, f nkx0))

≤ α(ψ(ωλ0 (f mk−1x0, f nk−1x0)))ψ(ωλ0+η(λ0)(f mk−1x0, f nk−1x0))

+ β(ψ(ωλ0(f mk−1x0, f nk−1x0)))ψ(ωλ0(f mk−1x0, f mk x0))

+ γ (ψ(ωλ0(f mk−1x0, f nk−1x0)))ψ(ωλ0(f nk−1x0, f nkx0))

≤ ψ(ωη(λ0)(f mk−1x0, f mkx0) + ωλ0(f mkx0, f nk−1x0))

+ ψ(ωλ0(f mk−1x0, f mk x0)) + ψ(ωλ0(f nk−1x0, f nk x0))

≤ ψ(ωη(λ0)(f mk−1x0, f mkx0) + ε0) + ψ(ωλ0(f mk−1x0, f mk x0))

+ ψ(ωλ0(f nk−1x0, f nk x0)).

Letting k ® ∞, we obtain that limk→∞ψ(ωλ0 (f mkx0, f nk x0)) = ψ(ε0) . So, we have

limk→∞

ωλ0(f mkx0, f nk x0) = ε0.

Observe again that

ψ(ωλ0(f mkx0, f nkx0)) ≤ ψ(ωλ0+η(λ0)(f mk−1x0, f nk−1x0)) + ψ(ωλ0(f mk−1x0, f mk x0))

+ ψ(ωλ0(f nk−1x0, f nkx0))

≤ ψ

(ωη(λ0)

2(f mk−1x0, f mkx0) + ωλ0 (f mkx0, f nkx0)

+ωη(λ0)2

(f nkx0, f nk−1x0))

+ ψ(ωλ0(f mk−1x0, f mkx0))

+ ψ(ωλ0(f nk−1x0, f nkx0)).

Letting k ® ∞, we deduce that limk→∞ψ(ωλ0 (f mk−1x0, f nk−1x0)) = ψ(ε0). Similarly,

we have

limk→∞

ωλ0(f mk−1x0, f nk−1x0) = ε0.

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 5 of 12

Thus, it follows that

1 ≤ lim infk→∞

α(ψ(ωλ0 (f mk−1x0, f nk−1x0))).

Therefore, we conclude that limk→∞ψ(ωλ0 (f mk−1x0, f nk−1x0)) = 0 , which implies

that

limk→∞

ωλ0(f mk−1x0, f nk−1x0) = 0.

This is a contradiction. Therefore, it follows that {fnx0} is a Cauchy sequence. Due to

the completeness of Xω, {fnx0} converges to some point x∞ Î Xω.

Now, we show that x∞ is a fixed point of f. Let l >0 be arbitrary. By virtue of the

condition (1.2), we consider that

ψ(ωλ(f n+1x0, f x∞) ≤ ψ(ωλ(f nx0, x∞)) + ψ(ωλ(f nx0, f n+1x0)).

Letting n ® ∞, we obtain that ψ(ωl(x∞, fx∞) ≤ 0 for all l >0. Therefore, x∞ is a fixed

point of f. ■Theorem 2.2. Additional to the Theorem 2.1, if ψ is subadditive and the following

condition holds:

for any x, y ∈ Xω, there exists w ∈ Xωwith w � fw and ωλ

(w, fw

)< ∞ for all λ > 0

such that w is comparable to both x and y,(2:2)

then the fixed point in Theorem 2.1 is unique.

Proof. By Theorem 2.1, we know that f has a fixed point x∞ Î Xω. Assume that y∞ ÎXω is also another fixed point of f. Thus, we can find w Î Xω with w ⊑ fw and com-

parable to both x∞ and y∞. It follows that fnw is comparable with both x∞ and y∞ for

each n Î N. Observe that, for any l >0,

ψ(ωλ(f n+1w, x∞) = ψ(ωλ(f n+1w, f x∞)

≤ α(ψ(ωλ(f nw, x∞)))ψ(ωλ(f nw, x∞))

+ β(ψ(ωλ(f nw, x∞)))ψ(ωλ(f nw, f n+1w))

≤ α(ψ(ωλ(f nw, x∞)))ψ(ωλ(f nw, x∞))

+ β(ψ(ωλ(f nw, x∞)))ψ(ωλ(f nw, x∞))

+ β(ψ(ωλ(f nw, x∞)))ψ(ωλ(x∞, f n+1w)).

(2:3)

Therefore, without loss of generality, we have

ψ(ωλ(f n+1w, x∞) ≤ α(ψ(ωλ(f nw,x∞)))+β(ψ(ωλ(f nw,x∞)))1−β(ψ(ωλ(f nw,x∞))) ψ(ωλ(f nw, x∞))

≤ ψ(ωλ(f nw, x∞))

...

≤ ψ(ωλ(w, x∞))

< ∞.

Therefore, {ψ(ωl(fnw, x∞))} is nonincreasing and bounded below. So, it converges to

some real number h ≥ 0. Assume that h >0. According to the proof of Theorem 2.1,

we know that limn®∞ ωl(fnw, f n+1w) = 0 for all l >0. Thus, letting n ® ∞ in the

inequality (2.3), we have

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 6 of 12

1 ≤ lim infn→∞ α(ψ(ωλ(f nw, x∞))).

Thus, we have {fnw} converges to x∞. Similarly, we obtain that {fnw } converges also

to y∞. Since the limit is unique, we have x∞ = y∞. This contradicts our assumption.

Therefore, the theorem is proved. ■Corollary 2.3. Additional to Theorem 2.1, if Xω is totally ordered, then the fixed

point in Theorem 2.1 is unique.

Proof. Since Xω is totally ordered, the condition (2.2) is satisfied. Thus, applying The-

orem 2.2, we obtain the result. ■The following two corollaries nicely broaden the results in [24] (see Theorems 3.2

and 3.6 [24]).

Corollary 2.4. Let Xω be a complete modular metric space with a partial ordering ⊑and f be a self-mapping on Xω such that, for any l >0, there exists h(l) Î (0, l) suchthat

ψ(ωλ(fx, fy)) ≤ α(ψ(ωλ(x, y)))ψ(ωλ+η(λ)(x, y)),

where α ∈ S and ψ ∈ ̄ . Assume also that f is continuous or the condition (1.2)

holds. Then f has a fixed point in Xω. Moreover, if the condition (2.2) is satisfied, the

fixed point is unique.

Proof. Since α ∈ S , we have (α, θ , θ) ∈ S3 . Thus, apply Theorems 2.1 and 2.2, we

have the conclusion. ■Corollary 2.5. Let Xω be a complete modular metric space with a partial ordering ⊑

and f be a self-mapping on Xω such that, for any l >0, there exist ζ(l), μ(l) Î (0, l)such that

ψ(ωλ(fx, fy)) ≤ β(ψ(ωλ(x, y)))ψ(ωλ(x, fx)) + γ (ψ(ωλ(x, y)))ψ(ωλ(y, fy)),

where ψ ∈ ̄ and (β , γ ) ∈ S2 with max{supt≥0 b(t), supt≥0 g(t)} <1. Assume also

that f is continuous or that the condition (1.2) holds. Then f has a fixed point in Xω.

Moreover, if the condition (2.2) is satisfied, the fixed point is unique.

Proof. Since (β , γ ) ∈ S2 , we have (θ , β , γ ) ∈ S3 . Thus, apply Theorems 2.1 and 2.2,

we have the conclusion. ■

ApplicationsIn this section, we give an application of our theorems to establish the existence and

uniqueness of a solution to a nonhomogeneous linear parabolic partial differential

equation satisfying a given initial condition.

Consider the following initial value problem{ut(x, t) = uxx(x, t) + F(x, t, u(x, t), ux(x, t)), −∞ < x < ∞, 0 < t ≤ Tu(x, 0) = ϕ(x) ≥ 0, −∞ < x < ∞,

(3:1)

where we assume � to be continuously differentiable such that � and �’ are bounded

and F is continuous.

By a solution of the system (3.1), we meant a function u ≡ u(x, t) defined on ℝ × I,

where I: = [0, T], satisfying the following conditions:

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 7 of 12

(a) u, ut, ux, uxx Î C(ℝ × I).

(b) u and ux are bounded in ℝ × I.

(c) ut(x, t) = uxx(x, t) + F(x, t, u(x, t), ux(x, t)) for all (x, t) Î ℝ × I.

(d) u(x, 0) = �(x) for all x Î ℝ.

Now, we consider the following space:

� := {u(x, t) : u, ux ∈ C(R × I) and||u|| < ∞},

where

‖u‖ := supx∈R,t∈I

|u(x, t)| + supx∈R,t∈I

|ux(x, t)|.

Obviously, the function ω: ℝ+ × Ω × Ω ® ℝ+ given by

ωλ(x, y) :=1

1 + λ||u − v||

is a metric modular on Ω. Clearly, the set Ωω is a complete modular metric space

independent of generators. Define a partial ordering ⊑ on Ωω by

u, v ∈ �ω, u � v ⇔ u(x, t) ≤ v(x, t) and ux(x, t) ≤ vx(x, t) at each (x, t) ∈ R × I.

Taking a nondecreasing sequence {un} in Ωω converging to u Î Ωω. For any (x, t) Îℝ × I,

we have

u1(x, t) ≤ u2(x, t) ≤ · · · ≤ un(x, t) ≤ · · ·

and

(u1)x(x, t) ≤ (u2)x(x, t) ≤ · · · ≤ (un)x(x, t) ≤ · · · .

Since the sequences {un(x, t)} and {(un)x(x, t)} converges to u(x, t) and ux(x, t),

respectively, it follows that, for any (x, t)Îℝ × I,

un(x, t) ≤ u(x, t) and (un)x(x, t) ≤ ux(x, t)

for all n ≥ 1. Therefore, un ⊑ u for all n ≥ 1. So, the space Ωω satisfies the condition (1.2).

Theorem 3.1. Consider the problem (3.1) and assume the following:

(1) For any c >0 with |s| < c and |p| < c, the function F (x, t, s, p) is uniformly

Hölder continuous in X and t for each compact subset of ℝ × I.

(2) There exists a constantcF ≤ (T + 2π

−12 T

12 )−1 such that, for any l > 0, there

exists h(l) Î (0, l) such that

0 ≤ 11+λ

[F(x, t, s2, p2) − F(x, t, s1, p1)]

≤ cF

[1

1+λ+η(λ)ρ

(�

(s2 − s1 + p2 − p1

1 + λ

))+ 1

1+λσ

(�

(s2 − s1 + p2 − p1

1 + λ

))]

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 8 of 12

for all (s1, p1), (s2, p2) Î ℝ × ℝ with s1 ≤ s2 and p1 ≤ p2, where � ∈ ̄ is sublinear

with Ξ(x) ≤ t and r, s are nondecreasing functions on ℝ+ such that r(t) <(1 -k)t

and s(t) <(1 -k)kt for all t >0 and for some fixed k Î (0, 1).

(3) The two functions Γ, ϒ: ℝ+ ® [0, 1) given by

�(t) =

{0 if t = 0,

ρ(t)(1−k)t if t > 0,

ϒ(t) =

{0 if t = 0,

σ (t)(1−k)t if t > 0,

are such that (�, ϒ , ϒ) ∈ S3 , Γ + 2ϒ <1.

(4) F(x, t, s, 0) ≥ s∫ t0

∫ ∞∞ k(x − ξ , t − τ )dξdτ

for all s ≥ 0.

(5) F is bounded for bounded s and p.

Then, the existence and uniqueness of the solution of the system (3.1) is affirmative.

It is essential to note that the problem (3.1) is equivalent (under the assumption of

Theorem 3.1) to the integral equation:

u(x, t) =

∞∫−∞

k(x− ξ , t)ϕ(ξ)dξ +

t∫0

∞∫−∞

k(x− ξ , t − τ )F(ξ , τ , u(ξ , τ ), ux(ξ , τ ))dξdτ (3:2)

for all x Î ℝ and 0 <t ≤ T, where

k(x, t) :=1√4π t

e− x2

4t

for all x Î ℝ and t >0. The system (3.1) possesses a unique solution if and only if the

equation (3.2) possesses a unique solution u such that u and ux are both continuous

and bounded for all x Î ℝ and 0 < t ≤ T.

Define a mapping Λ: Ωω ® Ωω by

(�u)(x, t) :=

∞∫−∞

k(x−ξ , t)ϕ(ξ)dξ+

t∫0

∞∫−∞

k(x−ξ , t−τ )F(ξ , τ , u(ξ , τ ), ux(ξ , τ ))dξdτ

for all (x, t) ℝ × I. Then the problem of finding the solution to the equation (3.2) is

equivalent to the problem of finding the fixed point of Λ.

Proof (Theorem 3.1). It is easy to see that the mapping Λ is nondecreasing by the

definition. Let u, v Î Ωω with u ⊑ v. Suppose that u ≠ v. Besides, we have

11+λ

∣∣(�v)(x, t) − (�u)(x, t)∣∣

≤ 11+λ

t∫0

∞∫−∞

k(x − ξ , t − τ )|F(ξ , τ , v(ξ , τ ), vx(ξ , τ )) − F(ξ , τ , u(ξ , τ ), ux(ξ , τ ))|dξdτ

≤t∫

0

∞∫−∞

k(x − ξ , t − τ )cF

[1

1+λ+η(λ)ρ(�

( 11+λ

(v(ξ , τ ) − u(ξ , τ ) + vx(ξ , τ ) − ux(ξ , τ )

)))+ 1

1+λσ

(�

( 11+λ

(v(ξ , τ ) − u(ξ , τ ) + vx(ξ , τ ) − ux(ξ , τ )

)))]dξdτ

≤ cF

[1

1+λ+η(λ)ρ(�(ωλ(u, v))) + 11+λ

σ (�(ωλ(u, v)))] t∫

0

∞∫−∞

k(x − ξ , t − τ )dξdτ

≤ cF

[1

1+λ+η(λ)ρ(�(ωλ(u, v))) + 11+λ

σ (�(ωλ(u, v)))]

T.

(3:3)

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 9 of 12

Similarly, we have

11+λ

∣∣(�v)x(x, t) − (�u)x(x, t)∣∣

≤ cF

[1

1+λ+η(λ)ρ(�(ωλ(u, v))) + 11+λ

σ (�(ωλ(u, v)))] t∫

0

∞∫−∞

|kx(x − ξ , t − τ )|dξdτ

≤ 2π− 1

2 T− 12 cF

[1

1+λ+η(λ)ρ(�(ωλ(u, v))) + 11+λ

σ (�(ωλ(u, v)))]

.

(3:4)

Note that by (2), we have supt≥0 ϒ(t) ≤ k < 1. Together with (3.3) and (3.4), we

obtain

ωλ(�u, �v) ≤ (T + 2π− 1

2 T12 )cF

[1

1+λ+η(λ)ρ(�(ωλ(u, v))) + 11+λ

σ (�(ωλ(u, v)))]

≤ 11+λ+η(λ)ρ(�(ωλ(u, v))) + 1

1+λσ (�(ωλ(u, v)))

= ||u−v||1+λ+η(λ)

ρ(�(ωλ(u, v)))||u − v|| + ||u−v||

1+λ

σ (�(ωλ(u, v)))||u − v||

≤ ωλ+η(λ)(u, v)ρ(�(ωλ(u, v)))

ωλ(u, v)+ ωλ(u, �u)

σ (�(ωλ(u, v)))

ωλ(u, v))

+ ωλ(v, �v)σ (�(ωλ(u, v)))

ωλ(u, v)+ ωλ(�u, �v)

σ (�(ωλ(u, v)))ωλ(u, v)

≤ ωλ+η(λ)(u, v)ρ(�(ωλ(u, v)))

ωλ(u, v)+ ωλ(u, �u)

σ (�(ωλ(u, v)))ωλ(u, v)

+ ωλ(v, �v)σ (�(ωλ(u, v)))

ωλ(u, v)+ ωλ(�u, �v)ϒ(ωλ(u, v))

≤ ωλ+η(λ)(u, v)ρ(�(ωλ(u, v)))

ωλ(u, v)+ ωλ(u, �u)

σ (�(ωλ(u, v)))ωλ(u, v)

+ ωλ(v, �v)σ (�(ωλ(u, v)))

ωλ(u, v)+ kωλ(�u, �v)

≤ ωλ+η(λ)(u, v)ρ(�(ωλ(u, v)))

�(ωλ(u, v))+ ωλ(u, �u)

σ (�(ωλ(u, v)))�(ωλ(u, v))

+ ωλ(v, �v)σ (�(ωλ(u, v)))

�(ωλ(u, v))+ kωλ(�u, �v).

Further, we obtain

ωλ(�u, �v) ≤ ωλ+η(λ)(u, v)�(�(ωλ(u, v))) + ωλ+ζ (λ)(u, �u)ϒ(�(ωλ(u, v)))

+ ωλ+μ(λ)(v, �v)ϒ(�(ωλ(u, v))).

Moreover, since ξ is a sublinear (real) function in the class Ψ, it follows that

�(ωλ(�u, �v)) ≤ �(ωλ+η(λ)(u, v))�(�(ωλ(u, v))) + �(ωλ+ζ (λ)(u, �u))ϒ(�(ωλ(u, v)))

+ �(ωλ+μ(λ)(v, �v))ϒ(�(ωλ(u, v))).

For the case u = v, it is obvious that the above inequality is satisfied. Thus, we now

have the inequality (2.1) holds for any comparable u, v Î Ωω.

Note that any constant functions are contained in Ω. Now, for any u, v Î Ωω, we

may choose a constant function w Î Ωω for which u, v ⊑ w and θ ⊑ w. Consequently,

we have w(x, t) = ||w|| for all (x, t) Î ℝ × I. Also, observe that this w attains the

following:

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 10 of 12

(�w)(x, t) =

∞∫−∞

k(x − ξ , t)ϕ(ξ)dξ+

t∫0

∞∫−∞

k(x − ξ , t − τ )F(ξ , τ , ||w||, 0)dξdτ

≥∞∫

−∞k(x − ξ , t)ϕ(ξ)dξ + ||w||

≥ ||w||= w(x, t).

Hence, the condition (2.2) is satisfied. Therefore, by applying the Theorems 2.1 and

2.2, the result follows. This completes the proof. ■

AcknowledgementsThe authors were supported by the Higher Education Research Promotion and National Research University Project ofThailand, Office of the Higher Education Commission (NRU-CSEC No.55000613). This research was partially finished atDepartment of Mathematics Education and the RINS, Gyeongsang National University, Chinju, Korea, while the firstand third authors visit here. Also, the second author was supported by the Basic Science Research Program throughthe National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science, and Technology(Grant no. 2011-0021821). Furthermore, the authors are grateful for the reviewers for the careful reading of the articleand for the suggestions which improved the quality of this work.

Author details1Department of Mathematics, Faculty of Science, King Mongkut’s University of Technology Thonburi, Bangkok 10140,Thailand 2Department of Mathematics Education and the RINS, Gyeongsang National University, Chinju 660-701, Korea

Authors’ contributionsAll authors read and approved the final manuscript.

Competing interestsThe authors declare that they have no competing interests.

Received: 3 May 2012 Accepted: 20 June 2012 Published: 20 June 2012

References1. Golubović, Z, Kadelburg, Z, Radenović, S: Coupled coincidence points of mappings in ordered partial metric spaces.

Abstr Appl Anal 2012, 18 (2012). Article ID 1925812. Khamsi, MA, Latif, A, Al-Sulami, H: KKM and Ky Fan theorems in modular function spaces. Fixed Point Theory Appl.

2011, 57 (2011). doi:10.1186/1687-1812-2011-573. Khamsi, MA: Quasicontraction mappings in modular spaces without Δ2-condition. Fixed Point Theory Appl 2008, 6

(2008). Article ID 9161874. Kuaket, K, Kumam, P: Fixed points of asymptotic pointwise contractions in modular spaces. Appl Math Lett. 24,

1795–1798 (2011). doi:10.1016/j.aml.2011.04.0355. Amini-harandini, A: Fixed point theory for generalized quasicontraction maps in vector modular spaces. Comput Math

Appl. 61, 1891–1897 (2011). doi:10.1016/j.camwa.2011.02.0216. Aydi, H, Nashine, HK, Samet, B, Yazidi, H: Coincidence and common fixed point results in partially ordered cone metric

spaces and applications to integral equations. Nonlinear Anal. 74(17), 6814–6825 (2011). doi:10.1016/j.na.2011.07.0067. Aydi, H: Common fixed point results for mappings satisfying (ψ, ϕ)-weak contractions in ordered partial metric spaces.

Int J Math Stat. 12(2), 53–64 (2012)8. Sintunavarat, W, Kumam, P: Weak condition for generalized multi-valued (f, α, β)-weak contraction mappings. Appl Math

Lett. 24, 460–465 (2011). doi:10.1016/j.aml.2010.10.0429. Sintunavarat, W, Cho, YJ, Kumam, P: Common fixed point theorems for c-distance in ordered cone metric spaces.

Comput Math Appl. 62, 1969–1978 (2011). doi:10.1016/j.camwa.2011.06.04010. Mongkolkeha, C, Kumam, P: Fixed point theorems for generalized asymptotic pointwise ρ-contraction mappings

involving orbits in modular function spaces. Appl Math Lett.2012http://dx.doi.org/10.1016/j.aml.2011.11.02711. Geraghty, M: On contractive mappings. Proc Am Math Soc. 40, 604–608 (1973). doi:10.1090/S0002-9939-1973-0334176-512. Amini-Harandi, A, Emani, H: A fixed point theorem for contraction type maps in partially ordered metric spaces and

application to ordinary differential equations. Nonlinear Anal. 72, 2238–2242 (2010). doi:10.1016/j.na.2009.10.02313. Eshaghi Gordji, M, Ramezani, M, Cho, YJ, Pirbavafa, S: A generalization of Geraghty’s theorem in partially ordered metric

spaces and application to ordinary differential equations. Fixed Point Theory Appl. 2012, 74 (2012)14. Chistyakov, VV: Modular metric spaces, I: basic concepts. Nonlinear Anal. 72, 1–14 (2010). doi:10.1016/j.na.2009.04.05715. Nakano, H: Modulared Semi-Ordered Linear Spaces. In Tokyo Mathematics Book Series, vol. 1,Maruzen Co., Tokyo (1950)16. Nakano, H: Topology and Linear Topological Spaces. In Tokyo Mathematics Book Series, vol. 3,Maruzen Co., Tokyo

(1951)17. Kumam, P: Some geometric properties and fixed point theorem in modular spaces. In: Garcia Falset J, Fuster L, Sims B

(eds.) Fixed Point Theorem and its Applications. pp. 173–188. Yokohama Publishers, Yokohama (2004)

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 11 of 12

18. Kumam, P: Fixed point theorems for nonexpansive mappings in, odular spaces. Arch Math (BRONO). 40(4), 345–353(2004)

19. Kumam, P: On uniform opial condition and uniform Kadec-Klee property in modular spaces. J Interdiscip Math. 8(3),377–385 (2005)

20. Kumam, P: On nonsquare and Jordan-Von Neumann constants of modular spaces. Southeast Asian Bull Math. 30(1),67–77 (2006)

21. Mongkolkeha, C, Kumam, P: Common fixed point for generallized weak contraction mapping in modular spaces. SciMath Japon117–127. e-2012 (online)

22. Cho, YJ, Saadati, R, Sadeghi, G: Quasi-contractive mappings in modular metric spaces. J Appl Math 2012, 5 (2012).Article ID 907951

23. Chaipunya, P, Mongkolkeha, C, Sintunavarat, W, Kumam, P: Fixed-point theorems for multivalued mappings in modularmetric spaces. Abstr Appl Anal 2012, 14 (2012). Article ID 503504

24. Mongkolkeha, C, Sintunavarat, W, Kumam, P: Fixed point theorems for contraction mappings in modular metric spaces.Fixed Point Theory Appl. 2011, 93 (2011). doi:10.1186/1687-1812-2011-93

doi:10.1186/1687-1847-2012-83Cite this article as: Chaipunya et al.: Geraghty-type theorems in modular metric spaces with an application topartial differential equation. Advances in Difference Equations 2012 2012:83.

Submit your manuscript to a journal and benefi t from:

7 Convenient online submission

7 Rigorous peer review

7 Immediate publication on acceptance

7 Open access: articles freely available online

7 High visibility within the fi eld

7 Retaining the copyright to your article

Submit your next manuscript at 7 springeropen.com

Chaipunya et al. Advances in Difference Equations 2012, 2012:83http://www.advancesindifferenceequations.com/content/2012/1/83

Page 12 of 12