• No results found

Coupled coincidence point and common coupled fixed point theorems in complex valued metric spaces

N/A
N/A
Protected

Academic year: 2020

Share "Coupled coincidence point and common coupled fixed point theorems in complex valued metric spaces"

Copied!
14
0
0

Loading.... (view fulltext now)

Full text

(1)

ISSN: 2008-6822 (electronic)

http://dx.doi.org/10.22075/ijnaa.2017.521

Coupled coincidence point and common coupled

fixed point theorems in complex valued metric

spaces

Fayyaz Rouzkarda,∗, Mohammad Imdadb

aFarhangian University, Shariati Pardis, Sari, Mazandaran Iran

bDepartment of Mathematics, Aligarh Muslim University, Aligarh, 202002, India (Communicated by A. Ebadian)

Abstract

In this paper, we introduce the concept of a w-compatible mappings and utilize the same to discuss the ideas of coupled coincidence point and coupled point of coincidence for nonlinear contractive mappings in the context of complex valued metric spaces besides proving existence theorems which are following by corresponding unique coupled common fixed point theorems for such mappings. Some illustrative examples are also given to substantiate our newly proved results.

Keywords: Common fixed point; Contractive type mapping; coupled coincidence point; coupled point of coincidence; Complex valued metric space.

2000 MSC: Primary 47H10; Secondary 54H25.

1. Introduction and preliminaries

The axiomatic development of a metric space was essentially carried out by French mathematician Frechet in the year 1906. The utility of metric spaces in the natural growth of Functional Analysis is enormous. Inspired from the impact of this natural idea to mathematics in general and to Functional Analysis in particular, several researchers attempted various generalizations of this notion in the recent past such as: rectangular metric spaces, semi metric spaces, quasi metric spaces, quasi semi metric spaces, pseudo metric spaces, probabilistic metric spaces, 2-metric spaces, D-metric spaces,

Corresponding author

Email addresses: fayyazrouzkard@gmail.com, fayyaz_rouzkard@yahoo.com(Fayyaz Rouzkard),

mhimdad@yahoo.co.in(Mohammad Imdad)

(2)

G-metric spaces, , K-metric spaces, Cone metric spaces etc and by now there exists considerable literature on all these generalizations of metric spaces (see [8]-[5]).

Most recently, Azam et al. [1] and Fayyaz et al. [12] studied complex valued metric spaces wherein some fixed point theorems for mappings satisfying a rational inequality were established. Naturally, this new idea can be utilized to define complex valued normed spaces and complex valued inner product spaces which, in turn, offer a lot of scopes for further investigation. Though complex valued metric spaces form a special class of cone metric space, yet this idea is intended to define rational expressions which are not meaningful in cone metric spaces and thus many results of analysis cannot be generalized to cone metric spaces. Indeed the definition of a cone metric space banks on the underlying Banach space which is not a division Ring . However, in complex valued metric spaces, we can study improvements of a host of results of analysis involving divisions.

In this paper we prove common fixed point theorems involving two pairs of weakly compatible mappings satisfying certain rational expressions in complex valued metric space.

In what follows, we recall some notations and definitions that will be utilized in our subsequent discussion. Let C be the set of complex numbers and z1, z2 ∈ C. Define a partial order - on C as

follows:

z1 -z2 if and only if Re(z1)≤Re(z2), Im(z1)≤Im(z2).

Consequently, one can infer that z1 -z2 if one of the following conditions is satisfied:

(i) Re(z1) = Re(z2), Im(z1)< Im(z2),

(ii) Re(z1)< Re(z2), Im(z1) =Im(z2),

(iii) Re(z1)< Re(z2), Im(z1)< Im(z2),

(iv) Re(z1) =Re(z2), Im(z1) =Im(z2).

In particular, we write z1 z2 if z1 6= z2 and one of (i), (ii), and (iii) is satisfied and we write

z1 ≺z2 if only (iii) is satisfied. Notice that 0-z1 z2 ⇒ |z1|<|z2|,andz1 -z2, z2 ≺z3 ⇒z1 ≺z3.

Definition 1.1. [1] Let X be a nonempty set whereas C be the set of complex numbers. Suppose that the mappingd:X×X →C, satisfies the following conditions:

(d1). 0-d(x, y), for all x, y ∈X and d(x, y) = 0 if and only if x=y;

(d2). d(x, y) =d(y, x) for all x, y ∈X;

(d3). d(x, y)-d(x, z) +d(z, y), for all x, y, z ∈X.

Then d is called a complex valued metric on X, and (X, d) is called a complex valued metric space.

Example 1.2. [12]. Let X =C be a set of complex number. Define d:C×CC, by

d(z1, z2) =|x1−x2|+i|y1−y2|

where z1 =x1+iy1 and z2 =x2+iy2. Then (X, d) is a complex valued metric space.

Example 1.3. [14] Let X =C be a set of complex number. Define d:C×CC,by

d(z1, z2) = eik|z1−z2|

where 0≤k ≤ π

2, z1 =x1+iy1 and z2 =x2+iy2. Then (X, d) is a complex valued metric space.

(3)

(i) b∈B is called an interior point of a set B whenever there is 0≺r∈C such that

N(b, r)⊆B

where N(b, r) ={y ∈X :d(b, y)≺r}.

(ii) A point x∈X is called a limit point of B whenever for every 0≺r∈C,

N(x, r)∩(B\X)6=∅.

(iii) A subset A⊆X is called open whenever each element of A is an interior point of A.

(iv) A subset B ⊆X is called closed whenever each limit point of B belongs to B.

(v)The family

F ={N(x, r) :x∈X,0≺r}

is a sub-basis for a topology on X. We denote this complex topology by τc. Indeed, the topology τc

is Hausdorff.

Definition 1.5. [1] Let (X, d) be a complex valued metric space and {xn}n≥1 be a sequence in X

and x∈X. We say that

(i) the sequence{xn}n≥1 converges to x if for every c ∈ C, with 0 ≺ c there is n0 ∈ N such that

for all n > n0, d(xn, x)≺c, . We denote this by limnxn=x,or xn→x, asn → ∞,

(ii) the sequence{xn}n≥1 is Cauchy sequence if for every c ∈C with 0 ≺ c there is n0 ∈ N such

that for all n > n0, d(xn, xn+m)≺c,

(iii) the metric space (X, d) is a complete complex valued metric space If every Cauchy sequence is convergent.

In [1], Azam et al. established the following two lemmas.

Lemma 1.6. Let (X, d) be a complex valued metric space and let {xn} be a sequence in X. Then {xn} converges to x if and only if |d(xn, x)| →0 as n→ ∞.

Lemma 1.7. Let (X, d) be a complex valued metric space and let {xn} be a sequence in X. Then

{xn} is a Cauchy sequence if and only if |d(xn, xn+m)| →0 as n→ ∞.

2. Main results

Bhashkar et al. [2] introduced the concept of coupled fixed point of a mapping F : X ×X → X

(4)

Definition 2.1. [2] An element (x, y) ∈ X ×X is called a coupled fixed point of mapping F :

X×X →X if x=F(x, y) and y=F(y, x).

Inspired with Definition 2.1 in the following we introduce the concept of a coupled fixed point of a mapping F :X×X→X.

Definition 2.2. An element (x, y)∈X×X is called

(g1) a coupled coincidence point of mappingF :X×X →X and g :X →X if g(x) =F(x, y) and

g(y) =F(y, x), and (gx, gy) is called coupled point of coincidence;

(g2) a common coupled fixed point of mappingsF :X×X →X andg :X →Xifx=g(x) = F(x, y) and y=g(y) =F(y, x).

Note that ifgis the identity mapping, then Definition 2.2 reduces to Definition 2.1. We introduce the following definition.

Definition 2.3. The mappings F : X ×X → X and g : X → X are called w−compatible if

g(F(x, y)) = F(gx, gy) whenever g(x) =F(x, y) and g(y) =F(y, x).

Definition 2.4. [10] Let (X,) be a partially ordered set and letF :X×X →X andg :X →X.

The mapping F is said to have a mixed g-monotone property if F is monotone g-nondecreasing in its first argument and monotone g-nonincreasing in its second argument, that is, for any x, y ∈X,

x1, x2 ∈X, gx1 gx2 ⇒F(x1, y)F(x2, y)

and, y1, y2 ∈X, gy1 gy2 ⇒F(x, y2)F(x, y1).

Now we prove our main result.

Theorem 2.5. Let (X, d) be a complex valued metric space, F : X ×X → X and g : X → X be mappings satisfying

d(F(x, y), F(u, v))-a1d(gx, gu) +a2d(gy, gv)

+a3d(F(x, y), gx) +a4d(F(x, y), gu) 1 +d(u, v)

+a5d(F(u, v), gu) +a6d(F(u, v), gx) 1 +d(x, y) ,

(2.1)

for all x, y, u, v ∈X, where ai, i= 1,2, . . . ,6 are nonnegative real numbers such that

P6

i=1ai <1. If

F(X×X)⊆g(X)and g(X)is complete subset of X,then F and g have a coupled coincidence point in X.

Proof . Letx0, y0 be any two arbitrary point in X.Setg(x1) = F(x0, y0) and g(y1) = F(y0, x0),this

can be done because F(X ×X)⊆g(X). Continuing this process we obtain two sequence {xn} and {yn} inX such that

(5)

From 2.1, we have

d(gxn, gxn+1) =d(F(xn−1, yn−1), F(xn, yn))

-a1d(gxn−1, gxn) +a2d(gyn−1, gyn)

+ a3d(F(xn−1, yn−1), gxn−1) +a4d(F(xn−1, yn−1), gxn) 1 +d(xn, yn)

+ a5d(F(xn, yn), gxn) +a6d(F(xn, yn), gxn−1) 1 +d(xn−1, yn−1)

=a1d(gxn−1, gxn) +a2d(gyn−1, gyn)

+ a3d(gxn, gxn−1) +a4d(gxn, gxn) 1 +d(xn, yn)

+ a5d(gxn+1, gxn) +a6d(gxn+1, gxn−1) 1 +d(xn−1, yn−1)

-a1d(gxn−1, gxn) +a2d(gyn−1, gyn) +

a3d(gxn, gxn−1)

1 +d(xn, yn)

+ a5d(gxn+1, gxn) +a6d(gxn+1, gxn) +a6d(gxn, gxn−1) 1 +d(xn−1, yn−1)

.

Therefore

d(gxn, gxn+1)-a1d(gxn−1, gxn) +a2d(gyn−1, gyn) +

a3d(gxn, gxn−1)

1 +d(xn, yn)

+a5d(gxn+1, gxn) +a6d(gxn+1, gxn) +a6d(gxn, gxn−1) 1 +d(xn−1, yn−1)

so that

|d(gxn, gxn+1)| ≤a1|d(gxn−1, gxn)|+a2|d(gyn−1, gyn)|+

a3|d(gxn, gxn−1)| |1 +d(xn, yn)|

+a5|d(gxn+1, gxn)|+a6|d(gxn+1, gxn)|+a6|d(gxn, gxn−1)|

|1 +d(xn−1, yn−1)|

since |1 +d(xn, yn)| ≥1 and |1 +d(xn−1, yn−1)| ≥1,therefore

|d(gxn, gxn+1)| ≤a1|d(gxn−1, gxn)|+a2|d(gyn−1, gyn)|+a3|d(gxn, gxn−1)|

+a5|d(gxn+1, gxn)|+a6|d(gxn+1, gxn)|+a6|d(gxn, gxn−1)|

= (a1+a3 +a6)|d(gxn, gxn−1)|+ (a5+a6)|d(gxn+1, gxn)|

+a2|d(gyn−1, gyn)|

from which it follows

(1−a5 −a6)|d(gxn, gxn+1)| ≤(a1+a3+a6)|d(gxn, gxn−1)|+a2|d(gyn−1, gyn)|. (2.2)

Similarly, one can prove that

(6)

Because of the symmetry in 2.1,

d(gxn+1, gxn) = d(F(xn, yn), F(xn−1, yn−1)) -a1d(gxn, gxn−1) +a2d(gyn, gyn−1)

+a3d(F(xn, yn), gxn) +a4d(F(xn, yn), gxn−1) 1 +d(xn−1, yn−1)

+a5d(F(xn−1, yn−1), gxn−1) +a6d(F(xn−1, yn−1), gxn) 1 +d(xn, yn)

=a1d(gxn, gxn−1) +a2d(gyn, gyn−1)

+a3d(gxn+1, gxn) +a4d(gxn+1, gxn−1) 1 +d(xn−1, yn−1)

+a5d(gxn, gxn−1) +a6d(gxn, gxn) 1 +d(xn, yn)

-a1d(gxn, gxn−1) +a2d(gyn, gyn−1)

+a3d(gxn+1, gxn) +a4d(gxn+1, gxn) +a4d(gxn, gxn−1) 1 +d(xn, yn)

+a5d(gxn, gxn−1) 1 +d(xn, yn)

.

So

|d(gxn+1, gxn)| ≤a1|d(gxn, gxn−1)|+a2|d(gyn, gyn−1)|+a3|d(gxn+1, gxn)|

+a4|d(gxn+1, gxn)|+a4|d(gxn, gxn−1)|+a5|d(gxn, gxn−1)|.

From which it follows

(1−a3 −a4)|d(gxn+1, gxn)| ≤(a1+a4+a5)|d(gxn−1, gxn)|+a2|d(gyn, gyn−1)|. (2.4)

Similarly, we can prove that

(1−a3 −a4)|d(gyn+1, gyn)| ≤(a1+a4+a5)|d(gyn−1, gyn)|+a2|d(gxn, gxn−1)|. (2.5)

Letδn=|d(gxn, gxn+1)|+|d(gyn, gyn+1)|.Now, from 2.2 and 2.3 respectively 2.4 and 2.5 we obtain:

(1−a5 −a6)δn ≤(a1+a2+a3+a6)δn−1. (2.6)

(1−a3 −a4)δn ≤(a1+a2+a4+a5)δn−1. (2.7)

Finally, from 2.6 and 2.7 we have

(2−a3 −a4−a5−a6)δn≤(2a1+ 2a2+a3+a4+a5+a6)δn−1,

that is,

δn ≤ηδn−1, η=

2a1+ 2a2+a3+a4 +a5+a6

2−a3−a4−a5−a6

<1. (2.8)

Consequently, we have

(7)

Ifδ0 = 0 then (x0, y0) is a coupled coincidence point of F and g.So let 0< δ0. If m > n, we have

d(gxn+m, gxn)-d(gxn+m, gxn+m−1) +d(gxn+m−1, gxn+m−2) +· · ·+d(gxn−1, gxn),

and

d(gyn+m, gyn)-d(gyn+m, gyn+m−1) +d(gyn+m−1, gyn+m−2) +· · ·+d(gyn−1, gyn),

so that

|d(gxn+m, gxn)| ≤ |d(gxn+m, gxn+m−1)|+|d(gxn+m−1, gxn+m−2)|+· · ·+|d(gxn−1, gxn)|,

and

|d(gyn+m, gyn)| ≤ |d(gyn+m, gyn+m−1)|+|d(gyn+m−1, gyn+m−2)|+· · ·+|d(gyn−1, gyn)|.

Therefore,

|d(gxn+m, gxn)|+|d(gyn+m, gyn)| ≤(|d(gxn+m, gxn+m−1)|+|d(gyn+m, gyn+m−1)|)

+ (|d(gxn+m−1, gxn+m−2)|+|d(gyn+m−1, gyn+m−2)|)

+· · ·+ (|d(gxn−1, gxn)|+|d(gyn−1, gyn)|),

that is,

|d(gxn+m, gxn)|+|d(gyn+m, gyn)| ≤δn+m−1+δn+m−2+· · ·+δn ≤(ηn+m−1+ηn+m−2+· · ·+ηn)δ0

= η

n

1−ηδ0 →0, as n→ ∞,

hence |d(gxn+m, gxn)| → 0 and |d(gyn+m, gyn)| → 0 as n → ∞. Thus by Lemma 1.7, {gxn} and {gyn} are Cauchy sequence in g(X). Since g(X) is complete subset of X, so there exists x and y in

X such that gxn →gx and gyn →gy. Now, we prove thatF(x, y) =gx and F(y, x) =gy. For that,

we have

d(F(x, y), gx)-d(F(x, y), gxn+1) +d(gxn+1, gx)

=d(F(x, y), F(xn, yn)) +d(gxn+1, gx)

-a1d(gx, gxn) +a2d(gy, gyn) +

a3d(F(x, y), gx) +a4d(F(x, y), gxn)

1 +d(xn, yn)

+a5d(F(xn, yn), gxn) +a6d(F(xn, yn), gx)

1 +d(x, y) +d(gxn+1, gx)

=a1d(gx, gxn) +a2d(gy, gyn) +

a3d(F(x, y), gx) +a4d(F(x, y), gxn)

1 +d(xn, yn)

+a5d(gxn+1, gxn) +a6d(gxn+1, gx)

1 +d(x, y) +d(gxn+1, gx)

-a1d(gx, gxn) +a2d(gy, gyn) +

a3d(F(x, y), gx) +a4d(F(x, y), gx) +a4d(gx, gxn)

1 +d(xn, yn)

+a5d(gxn+1, gx) +a5d(gx, gxn) +a6d(gxn+1, gx)

(8)

so that

|d(F(x, y), gx)| ≤a1|d(gx, gxn)|+a2|d(gy, gyn)|+a3|d(F(x, y), gx)|

+a4|d(F(x, y), gx)|+a4|d(gx, gxn)|+a5|d(gxn+1, gx)|+a5|d(gx, gxn)|

+a6|d(gxn+1, gx)|+|d(gxn+1, gx)|,

which further implies that,

|d(F(x, y), gx)| ≤ a1+a4+a5

1−a3−a4

|d(gxn, gx)|+

a5+a6

1−a3−a4

|d(gxn+1, gx)|

+ a2

1−a3−a4

|d(gyn, gy)|.

(2.10)

Sincegxn→gxandgyn→gyasn → ∞,then,|d(F(x, y), gx)|= 0,it follows thatd(F(x, y), gx) = 0,

and hence F(x, y) = gx. Similarly, we can prove F(y, x) = gy. Hence (x, y) is coupled coincidence point of the mappings F and g.

By setting a1 = a2 =α, a3 =a4 = β and a5 = a6 =γ in Theorem 2.5, we deduce the following

corollary.

Corollary 2.6. Let (X, d) be a complex valued metric space, F : X ×X → X and g : X → X be mappings satisfying

d(F(x, y), F(u, v))-α[d(gx, gu) +d(gy, gv)] +βd(F(x, y), gx) +d(F(x, y), gu)

1 +d(u, v)

+γd(F(u, v), gu) +d(F(u, v), gx)

1 +d(x, y) ,

(2.11)

for all x, y, u, v ∈ X, where α, β and γ are nonnegative real numbers such that α+β +γ < 12. If

F(X×X)⊆g(X)and g(X)is complete subset of X,then F and g have a coupled coincidence point in X.

Now, we present two examples showing that Theorem 2.5 is a proper extension of known results.

Example 2.7. Let X = [0,∞).The mapping d:X×X →C defined as:

d(x, y) =|x−y|i.

Clearly, (X, d) is a complete complex valued metric space. We define the functionsF :X×X →X

and g :X →X by

g(x) = 3x and F(x, y) = x 5 +

y

5.

It is easy to verify that F and g satisfy all the conditions of Theorem 2.5, takinga1 =a2 = 101, a3 =

a4 = 121 and a5 =a6 = 141. Moreover (0,0), is common coupled coincidence point ofF and g.

Example 2.8. LetX = [0,∞) andd:X×X →Cis a mapping defined by: d(x, y) =|x−y|+|x−y|i.

Clearly, (X, d) is a complete complex valued metric space. We define the functionsF :X×X →X

and g :X →X by

g(x) = 3x and F(x, y) =x+ |siny| 5 .

It is easy to verify that F and g satisfy all the conditions of Theorem 2.5, taking a1 = 23, a2 = 1

9 and a3 =a4 =a5 =a6 = 1

(9)

Theorem 2.9. Let (X, d)be a complex valued metric space, F :X×X →X andg :X →X be two mappings which satisfy all the conditions of Theorem 2.5. If F and g arew−compatible, then F and

g have unique common coupled fixed point. Moreover, common fixed point of F and g has the form

(u, u) for someu∈X.

Proof . First we claim that coupled point of coincidence is unique. suppose that (x, y),(x∗, y∗) ∈

X×X with g(x) =F(x, y), g(y) =F(y, x), and g(x∗) =F(x∗, y∗), g(y∗) =F(y∗, x∗).Using 2.1, we get

d(gx, gx∗) =d(F(x, y), F(x∗, y∗))

-a1d(gx, gx∗) +a2d(gy, gy∗)

+ a3d(F(x, y), gx) +a4d(F(x, y), gx

)

1 +d(x∗, y)

+ a5d(F(x

, y), gx) +a

6d(F(x∗, y∗), gx)

1 +d(x, y)

=a1d(gx, gx∗) +a2d(gy, gy∗) +

a4d(gx, gx∗)

1 +d(x∗, y) +

a6d(gx∗, gx)

1 +d(x, y) ,

so that

|d(gx, gx∗)| ≤a1|d(gx, gx∗)|+a2|d(gy, gy∗)|+|

a4d(gx, gx∗)

1 +d(x∗, y)|+|

a6d(gx∗, gx)

1 +d(x, y) |,

and

|d(gx, gx∗)| ≤a1|d(gx, gx∗)|+a2|d(gy, gy∗)|+a4|d(gx, gx∗)|+a6|d(gx∗, gx).

Therefore

|d(gx, gx∗)| ≤(a1+a4+a6)|d(gx, gx∗)|+a2|d(gy, gy∗)|. (2.12)

Similarly, one gets

|d(gy, gy∗)| ≤(a1+a4+a6)|d(gy, gy∗)|+a2|d(gx, gx∗)|. (2.13)

Thus

|d(gx, gx∗)|+|d(gy, gy∗)| ≤(a1+a2+a4+a6)(|d(gx, gx∗)|+|d(gy, gy∗)|).

Since a1 +a2 +a4 +a6 < 1, therefore we have |d(gx, gx∗)|+|d(gy, gy∗)| = 0, which implies that

gx = gx∗ and gy =gy∗. Similarly we can prove that gx =gy∗ and gy =gx∗. Therefore (gx, gy) is unique coupled point of coincidence ofF andg.Now, letg(x) = u.Then we haveu=g(x) =F(x, x).

Byw−compatibility of F and g, we have

g(u) = g(g(x)) = g(F(x, x)) =F(gx, gx) =F(u, u).

Then (gu, gu) is coupled point of coincidence ofF and g.Consequentlygu=gx.Thereforeu=gu=

(10)

Theorem 2.10. Let (X, d) be a complex valued metric space, F : X×X → X and g : X → X be

w−compatible mappings such that

d(F(x, y), F(u, v))kU +mV (2.14)

for all x, y, u, v ∈X, where,

U, V ∈Su,vx,y ={d(gx, gu), d(gy, gv),d(F(x, y), gx)

1 +d(u, v) ,

d(F(x, y), gu) 1 +d(u, v) ,

d(F(u, v), gu) 1 +d(x, y) },

andk, m are nonnegative real numbers such thatk+m <1. If F(X×X)⊆g(X)than F andg have unique common coupled fixed point having the form (u, u) for some u∈X.

Proof . Following similar arguments to those given in Theorem 2.5, we construct two sequences

{xn} and {yn} inX such that

g(xn+1) =F(xn, yn) and g(yn+1) =F(yn, xn).

Now, from 2.14, we have

d(gxn, gxn+1) =d(F(xn−1, yn−1), F(xn, yn))kU+mV, (2.15)

where U, V ∈Sxn−1,yn−1 xn,yn , and

d(gyn, gyn+1) =d(F(yn−1, xn−1), F(yn, xn))kU+mV, (2.16)

where U, V ∈Syn−1,xn−1 yn,xn .

We have the following 15 cases:

(i) :U =d(gxn−1, gxn) and V =d(gxn−1, gxn).

(ii) :U =d(gxn−1, gxn) and V =d(gyn−1, gyn).

(iii) :U =d(gxn−1, gxn) and V = d(F(x1+n−d1,y(gxnn,gy1),gxn)n−1).

(iv) :U =d(gxn−1, gxn) and V =

d(F(xn−1,yn−1),gxn)

1+d(gxn,gyn) .

(v) :U =d(gxn−1, gxn) and V = 1+d(dF((gxxnn,yn1,gy),gxnn)1).

(vi) :U =d(gyn−1, gyn) and V =d(gyn−1, gyn).

(vii) :U =d(gyn−1, gyn) and V =

d(F(xn−1,yn−1),gxn−1) 1+d(gxn,gyn) .

(viii) :U =d(gyn−1, gyn) and V = d(F1+(xnd(1,ygxnn,gy−1)n,gx) n).

(ix) :U =d(gyn−1, gyn) and V = 1+d(dF((gxxn,yn),gxn)

n−1,gyn−1).

(x) :U = d(F(xn−1,yn−1),gxn−1)

1+d(gxn,gyn) and V =

d(F(xn−1,yn−1),gxn−1) 1+d(gxn,gyn) .

(xi) :U = d(F(xn−1,yn−1),gxn−1)

1+d(gxn,gyn) and V =

d(F(xn−1,yn−1),gxn)

(11)

(xii) :U = d(F(xn−1,yn−1),gxn−1)

1+d(gxn,gyn) and V =

d(F(xn,yn),gxn)

1+d(gxn−1,gyn−1).

(xiii) :U = d(F(xn−1,yn−1),gxn)

1+d(gxn,gyn) and V =

d(F(xn−1,yn−1),gxn)

1+d(gxn,gyn) .

(xiv) :U = d(F(xn−1,yn−1),gxn)

1+d(gxn,gyn) and V =

d(F(xn,yn),gxn)

1+d(gxn−1,gyn−1).

(xv) :U = d(F(xn,yn),gxn)

1+d(gxn−1,gyn−1) and V =

d(F(xn,yn),gxn)

1+d(gxn−1,gyn−1).

In case (i),according to 2.15 and 2.16 we get

d(gxn, gxn+1) =d(F(xn−1, yn−1), F(xn, yn))(k+m)d(gxn−1, gxn),

and

d(gyn, gyn+1) =d(F(yn−1, xn−1), F(yn, xn))(k+m)d(gyn−1, gyn).

Therefore,

|d(gxn, gxn+1)| ≤(k+m)|d(gxn−1, gxn)|,

and

|d(gyn, gyn+1)| ≤(k+m)|d(gyn−1, gyn)|.

Hence,

(|d(gxn, gxn+1)|+|d(gyn, gyn+1)|)≤(k+m)(|d(gxn−1, gxn)|+|d(gyn−1, gyn)|),

and also, in the cases (ii),(iii),(vi),(vii),(x),(xiii) and (xv) according to 2.15 and 2.16 we obtain that

(|d(gxn, gxn+1)|+|d(gyn, gyn+1)|)≤(k+m)(|d(gxn−1, gxn)|+|d(gyn−1, gyn)|).

Similarly, in the cases (ix),(xii) and (xiv), according to 2.15 and 2.16 we again obtain that

(|d(gxn, gxn+1)|+|d(gyn, gyn+1)|)≤

k

1−m(|d(gxn−1, gxn)|+|d(gyn−1, gyn)|),

and, in the cases (iv),(v),(viii) and (xi) we have

(|d(gxn, gxn+1)|+|d(gyn, gyn+1)|)≤k(|d(gxn−1, gxn)|+|d(gyn−1, gyn)|).

Thus we conclude that

(|d(gxn, gxn+1)|+|d(gyn, gyn+1)|)≤β(|d(gxn−1, gxn)|+|d(gyn−1, gyn)|),

for some β∈ {k,(k+m),1km}(β <1) and for alln ≥1.

Following similar arguments to those given in Theorem 2.5, (x, y) is common coupled coincidence point of F and g, where gxn →gx and gyn → gy. Now we prove that coupled point of coincidence

(12)

Suppose that (x, y),(x∗, y∗) ∈ X ×X such that g(x) = F(x, y), g(y) = F(y, x) and g(x∗) =

F(x∗, y∗), g(y∗) =F(y∗, x∗).

By using 2.14 we get

d(gx, gx∗) =d(F(x, y), F(x∗, y∗))kU+mV (2.17)

where U, V ∈Sxx,y∗,y∗, and

d(gy, gy∗) = d(F(y, x), F(y∗, x∗))kU +mV, (2.18)

where U, V ∈Syy,x∗,x∗.

Again we have the following 15 cases: (i) :U =d(gx, gx∗) and V =d(gx, gx∗).

(ii) :U =d(gx, gx∗) and V =d(gy, gy∗).

(iii) :U =d(gx, gx∗) and V = 1+d(Fd((gxx,y∗),gy,gx∗)).

(iv) :U =d(gx, gx∗) and V = d1+(Fd((x,ygx∗),gx,gy∗∗)).

(v) :U =d(gx, gx∗) and V = d(F1+(xd(,ygx,gy∗),gx)∗).

(vi) :U =d(gy, gy∗) and V =d(gy, gy∗).

(vii) :U =d(gy, gy∗) and V = 1+d(Fd((gxx,y∗),gy,gx∗)).

(viii) :U =d(gy, gy∗) and V = d1+(Fd((x,ygx∗),gx,gy∗∗)).

(ix) :U =d(gy, gy∗) and V = d(F1+(xd(,ygx,gy∗),gx)∗).

(x) :U = 1+d(Fd((gxx,y∗),gy,gx∗)) and V =

d(F(x,y),gx) 1+d(gx∗,gy).

(xi) :U = 1+d(Fd((gxx,y∗),gy,gx∗)) and V =

d(F(x,y),gx∗) 1+d(gx∗,gy).

(xii) :U = 1+d(Fd((gxx,y∗),gy,gx∗)) and V =

d(F(x∗,y),gx)

1+d(gx,gy) .

(xiii) :U = d1+(Fd((x,ygx∗),gx,gy∗∗)) and V =

d(F(x,y),gx∗)

1+d(gx∗,gy).

(xiv) :U = d1+(Fd((x,ygx∗),gx,gy∗∗)) and V =

d(F(x∗,y∗),gx∗) 1+d(gx,gy) .

(xv) :U = d(F1+(xd(,ygx,gy∗),gx)∗) and V = d(F1+(xd(,ygx,gy∗),gx)∗).

In the cases (i),(ii),(iv),(vi),(viii) and (xiii) according to 2.17 and 2.18, we obtain that

|d(gx, gx∗)|+|d(gy, gy∗)| ≤(k+m)(|d(gx, gx∗)|+|d(gy, gy∗)|). (2.19)

Similarly, from 2.17 and 2.18, in the cases (iii),(v),(vii),(ix),(x),(xi),(xii),(xiv) and (xv) we again obtain that

(13)

where α ∈ {0, k, m} (α < 1). Since k+m < 1 and α < 1, therefore from 2.19 and 2.20, we have

|d(gx, gx∗)|+|d(gy, gy∗)| = 0, i.e. gx = gx∗ and gy =gy∗. That is, (gx, gx) is the unique common coupled point of coincidence. SinceF and g are w−compatible maps, we have,

g(g(x)) = F(gx, gx). (2.21)

Letu =g(x). By 2.21, we have g(u) =F(u, u). Therefore (gu, gu) is a coupled point of coincidence of F and g. Consequently, u = g(u) = F(u, u). Hence (u, u) is unique common coupled fixed point of F and g.

Example 2.11. Let X = {(x,0) : x ∈ [0,∞)} ∪ {(0, x) : x ∈ [0,∞)}. Define d : X ×X → C by

d(x, y) = |x1 −y1|+|x2 −y2|i, where x = (x1, x2) and y = (y1, y2). Clearly, (X, d) is a complete

complex valued metric space. Consider mappingsF :X×X →X and g :X →X,given by

g(x) = (

(0,t); x= (t,0), t∈[0,∞)

(t,0); x= (0, t), t∈[0,∞) and F (x1, x2),(y1, y2)

= (x1 2 ,

x2

2).

Note that F and g satisfy all the conditions of Theorem 2.9, if we take a1 = a2 = a3 = a4 = a5 =

a6 = 121. Moreover (0,0) is the unique common coupled fixed point of F and g.

Next, we give a example which supports Theorems 2.5 and 2.9, while a lot of results in partially ordered metric spaces, cone metric spaces, ordered cone metric spaces and complex valued metric spaces for example the results of Nashine et al. [11] do not.

Example 2.12. LetX ={ix:x∈[0,1]}.Defined:X×X →Cbyd(x, y) = i|x−y|, where x, y ∈

X. Clearly, (X, d) is a complete complex valued metric space. Consider mappings F :X×X →X

and g :X →X,given by

F(x, y) =ix

4+y4

16 and g(x) =i

x4

4 .

Ify1 = 34iand y2 = 12i,then g(y2) =i 1 16

4 i 81 44

4 =g(y1), but for x=i,we get

F(x, y2) =F(i,

1 2i) = i

1 + 161 16 i

1 + 8144

16 =F(i, 3

4i) =F(x, y1).

So, the mappings F and g do not satisfy the mixed g-monotone property. Therefore, Theorems 3.1 and 3.2 of Nashine et al. [11] cannot be supported to reach this conclusion.

Now, we show that Theorems 2.5 and 2.9 can be used for this case.

d(F(x, y), F(u, v)) =i|ix

4+y4

16 −i

u4+v4

16 |= 1 4i|i

x4−u4

4 +i

y4−v4

4 |

1

4i|i

x4u4

4 |+ 1 4i|i

y4v4

4 |= 1 4i|

x4u4

4 |+ 1 4i|

y4v4

4 |

= 1

4d(g(x), g(u)) + 1

4d(g(y), g(v)),

where a1 =a2 = 14 and a3 =a4 =a5 =a6 = 0. Note thata1+a2+a3+a4+a5+a6 <1.

F(X×X) ={ix:x∈[0,1

(14)

and g(X) is a complete subset of X. Hence, the conditions of Theorem 2.5 are satisfied. Therefore,

F and g have a coupled coincidence point in X.

Since, g(0) =F(0,0) and

g(F(0,0)) =g(0) =F(0,0) =F(g(0), g(0)),

then F and g are W-compatible. Therefore, conditions of Theorem 2.9 are satisfied. Now, we can apply Theorem 2.9 to conclude the existence of a unique common coupled fixed point of F and g that is a point (0,0).

References

[1] A. Azam, B. Fisher and M. Khan, Common Fixed Point Theorems in Complex Valued Metric Spaces, Numer. Funct. Anal. Optim. 32 (2011) 243–253.

[2] T.G. Bhashkar and V. Lakshmikantham, Fixed point theorems in partially ordered cone metric spaces and appli-cations, Nonlinear Anal.: Theory, Methods Appl. 65 (2006) 825–832.

[3] S.K. Chatterjea, Fixed point theorems, C.R. Bulgar Sci. 25 (1972) 727–730.

[4] O. Hadzic and E. Pap,Fixed Point Theory in PM-Spaces, Kluwer Academic, Dordrecht, The Netherlands, 2001. [5] G.E. Hardy and T.D. Rogers,A generalization of a fixed point theorems of Reich, Canadian Math. Bull. 16 (1973)

201–206.

[6] L.G. Huang and X. Zhang,Cone metric spaces and fixed point theorems of contractive mappings, J. Math. Anal. Appl. 332 (2007) 1468–1476.

[7] M. Imdad, J. Ali and M. Tanveer, Coincidence and common fixed point theorems for nonlinear contractions in Menger PM spaces, Chaos Solitons Fractals 42 (2009) 3121–3129.

[8] G. jungck and B.E. Rhoades,Fixed point for set valued functions without continuity, Indian J. Pure Appl. Math. 29 (1998) 227–238.

[9] R. Kannan, Some results on fixed points, Bull. Calcutta Math. Soc. 60 (1968) 71–76.

[10] V. Lakshmikantham and L. Ciric, Coupled fixed point theorems for nonlinear contractive in partially ordered metric spaces, Nonlinear Anal. 70 (2009) 4341–4349.

[11] H.K. Nashine, Z. Kadelburg and S. Radenovic,Coupled common fixed point theorems for w∗-compatible mappings in ordered cone metric spaces, Appl. Math. Comput. 218 (2012) 5422–5432.

[12] F. Rouzkard and M. Imdad,Some common fixed point theorems on complex valued metric spaces, Comput. Math. Appl. 64 (2012) 1866–1874.

[13] M. Telci and B. Fisher, On a fixed point theorem for fuzzy mappings in quasi-metric spaces, Thai J. Math. 2 (2003) 1–8.

References

Related documents

As a result of numerical analyses on AR(1), the incompletely estimated parameter (IEP) follows a Cauchy distribution, which has a power law of the slope -2 in the

Instead of the personal surveillance of the liberalistic approach to security, which can be regarded as a logical consequence of the methodological individualism –

To examine the nature of the association between atopy and glioma, we implemented two-sample MR [ 17 ] to estimate associations between atopy-associated SNPs and glioma risk

Three goals are defined for the Product Integration process area: (i) prepare for product integration, (ii) ensure interface compatibility and (iii) assemble product components

HC, healthy control individuals, no sign of neurological disease; MCI, patients with a diagnosis of mild cognitive impairment; AD, patients with a diagnosis of probable Alzheimer ’

Clearly, models to study AD initiation and progression in the context of diff erent human neural circuits in vitro, including cortico-cortical, cortico-thalamic

The present data (i) add to previous studies reporting the rare occurrence of white matter PrP amyloid plaques in patients with an otherwise classic sCJDMM1 pheno- type; (ii)

Moni torin g tools are nec essary for lon g-term evaluation – Macd uff (2009 ) UK Full report - Evaluation of Cleanlin ess Cham pions Prog ram using a mix of qua litative and