1 Introduction

The concept of G-metric space was introduced by Mustafa and Sims [1] in order to extend and generalize the notion of metric space. In this paper, the authors characterized the Banach contraction mapping principle [2] in the context of a G-metric space. Following this initial report, a number of authors have characterized many well-known fixed point theorems in the setting of G-metric space (see, e.g., [1, 336]). Since one is adapted from the other, there is a close relation between a usual metric space and a G-metric space (see, e.g., [1, 2327]). In fact, the nature of a G-metric space is to understand the geometry of three points instead of two points via perimeter of a triangle. However, most of the published papers dealing with a G-metric space did not give much importance to these details. Consequently, a great majority of results were obtained by transforming the contraction conditions from the usual metric space context to a G-metric space without carrying enough of the characteristics of the G-metric.

Very recently, Samet et al. [37] and Jleli-Samet [38] observed that some fixed point theorems in the context of a G-metric space in the literature can be concluded by some existing results in the setting of a (quasi-)metric space. In fact, if the contraction condition of the fixed point theorem on a G-metric space can be reduced to two variables instead of three variables, then one can construct an equivalent fixed point theorem in the setup of a usual metric space. More precisely, in [37, 38], the authors noticed that d(x,y)=G(x,y,y) forms a quasi-metric. Hence, if one can transform the contraction condition of existence results in a G-metric space in such terms, G(x,y,y), then the related fixed point results become the known fixed point results in the context of a quasi-metric space.

In this paper, we notice that the techniques used in [37, 38] are valid if the contraction condition in the statement of the theorem can be expressed in two variables. Furthermore, we prove some fixed point theorems in the context of a G-metric space for which the techniques in [37, 38] are inapplicable.

2 Preliminaries

In this section we recollect basic definitions and a detailed overview of the fundamental results. Throughout this paper, ℕ is the set of nonnegative integers, and N is the set of positive integers.

Definition 2.1 (See [1])

Let X be a non-empty set and let G:X×X×X R + be a function satisfying the following properties:

  1. (G1)

    G(x,y,z)=0 if x=y=z,

  2. (G2)

    0<G(x,x,y) for all x,yX with xy,

  3. (G3)

    G(x,x,y)G(x,y,z) for all x,y,zX with yz,

  4. (G4)

    G(x,y,z)=G(x,z,y)=G(y,z,x)= (symmetry in all three variables),

  5. (G5)

    G(x,y,z)G(x,a,a)+G(a,y,z) for all x,y,z,aX (rectangle inequality).

Then the function G is called a generalized metric or, more specifically, a G-metric on X, and the pair (X,G) is called a G-metric space.

Every G-metric on X defines a metric d G on X by

d G (x,y)=G(x,y,y)+G(y,x,x)for all x,yX.
(1)

Example 2.1 Let (X,d) be a metric space. The function G:X×X×X[0,+), defined as

G(x,y,z)=max { d ( x , y ) , d ( y , z ) , d ( z , x ) }

or

G(x,y,z)=d(x,y)+d(y,z)+d(z,x)

for all x,y,zX, is a G-metric on X.

Definition 2.2 (See [1])

Let (X,G) be a G-metric space, and let { x n } be a sequence of points of X. We say that { x n } is G-convergent to xX if

lim n , m + G(x, x n , x m )=0,

that is, for any ε>0, there exists NN such that G(x, x n , x m )<ε for all n,mN. We call x the limit of the sequence and write x n x or lim n + x n =x.

Proposition 2.1 (See [1])

Let (X,G) be a G-metric space. The following are equivalent:

  1. (1)

    { x n } is G-convergent to x,

  2. (2)

    G( x n , x n ,x)0 as n+,

  3. (3)

    G( x n ,x,x)0 as n+,

  4. (4)

    G( x n , x m ,x)0 as n,m+.

Definition 2.3 (See [1])

Let (X,G) be a G-metric space. A sequence { x n } is called a G-Cauchy sequence if, for any ε>0, there is NN such that G( x n , x m , x l )<ε for all m,n,lN, that is, G( x n , x m , x l )0 as n,m,l+.

Proposition 2.2 (See [1])

Let (X,G) be a G-metric space. Then the following are equivalent:

  1. (1)

    the sequence { x n } is G-Cauchy,

  2. (2)

    for any ε>0, there exists NN such that G( x n , x m , x m )<ε for all m,nN.

Definition 2.4 (See [1])

A G-metric space (X,G) is called G-complete if every G-Cauchy sequence is G-convergent in (X,G).

We will use the following result which can be easily derived from the definition of a G-metric space (see, e.g., [1]).

Lemma 2.1 Let (X,G) be a G-metric space. Then

G(x,x,y)2G(x,y,y) for all x,yX.

Definition 2.5 (See [1])

Let (X,G) be a G-metric space. A mapping T:XX is said to be G-continuous if {T( x n )} is G-convergent to T(x) where { x n } is any G-convergent sequence converging to x.

In [22], Mustafa characterized the well-known Banach contraction mapping principle in the context of G-metric spaces in the following ways.

Theorem 2.1 (See [22])

Let (X,G) be a complete G-metric space and let T:XX be a mapping satisfying the following condition for all x,y,zX:

G(Tx,Ty,Tz)kG(x,y,z),
(2)

where k[0,1). Then T has a unique fixed point.

Theorem 2.2 (See [22])

Let (X,G) be a complete G-metric space and let T:XX be a mapping satisfying the following condition for all x,yX:

G(Tx,Ty,Ty)kG(x,y,y),
(3)

where k[0,1). Then T has a unique fixed point.

Remark 2.1 The condition (2) implies the condition (3). The converse is true only if k[0, 1 2 ). For details, see [22].

Theorem 2.3 (See [26])

Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)aG(x,y,z)+bG(x,Tx,Tx)+cG(y,Ty,Ty)+dG(z,Tz,Tz)
(4)

for all x, y, z, where a, b, c, d are positive constants such that k=a+b+c+d<1. Then there is a unique xX such that Tx=x.

Theorem 2.4 (See [27])

Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)k [ G ( x , T x , T x ) + G ( y , T y , T y ) + G ( z , T z , T z ) ]
(5)

for all x, y, z, where k[0, 1 3 ). Then there is a unique xX such that Tx=x.

Theorem 2.5 (See [26])

Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)aG(x,y,z)+b [ G ( x , T x , T x ) + G ( y , T y , T y ) + G ( z , T z , T z ) ]
(6)

for all x, y, z, where a, b are positive constants such that k=a+b<1. Then there is a unique xX such that Tx=x.

Theorem 2.6 (See [26])

Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)aG(x,y,z)+bmax { G ( x , T x , T x ) , G ( y , T y , T y ) , G ( z , T z , T z ) }
(7)

for all x, y, z, where a, b are positive constants such that k=a+b<1. Then there is a unique xX such that Tx=x.

Theorem 2.7 (See [25])

Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G ( T x , T y , T z ) k max { G ( x , y , z ) , G ( x , T x , T x ) , G ( y , T y , T y ) , G ( z , T z , T z ) , G ( z , T x , T x ) , G ( x , T y , T y ) , G ( y , T z , T z ) }
(8)

for all x, y, z, where k[0, 1 2 ). Then there is a unique xX such that Tx=x.

Theorem 2.8 (See, e.g., [38])

Let (X,G) be a complete G-metric space and let T:XX be a given mapping satisfying

G(Tx,Ty,Tz)G(x,y,z)φ ( G ( x , y , z ) )
(9)

for all x,yX, where φ:[0,)[0,) is continuous with φ 1 ({0})=0. Then there is a unique xX such that Tx=x.

Definition 2.6 (See, e.g., [38])

A quasi-metric on a nonempty set X is a mapping p:X×X[0,) such that

(p1) x=y if and only if p(x,y)=0,

(p2) p(x,y)p(x,z)+p(z,y),

for all x,y,zX. A pair (X,p) is said to be a quasi-metric space.

Samet et al. [37] and Jleli-Samet [38] noticed that p(x,y)= p G (x,y)=G(x,y,y) is a quasi-metric whenever G:X×X×X[0,) is a G-metric. It is well known that each quasi-metric induces a metric. Indeed, if (X,p) is a quasi-metric space, then the function defined by

d(x,y)= d G (x,y)=max { p ( x , y ) , p ( y , x ) } for all x,yX

is a metric on X.

Theorem 2.9 Let (Xd) be a complete metric space and let T:XX be a mapping with the property

d(Tx,Ty)qmax { d ( x , y ) , d ( x , T x ) , d ( y , T y ) , d ( x , T y ) , d ( y , T x ) }
(10)

for all xX, where q is a constant such that q[0,1). Then T has a unique fixed point.

Samet et al. [37] proved that Theorem 2.4-Theorem 2.7 are the consequences of Theorem 2.9 by using the following proposition.

Proposition 2.3

  1. (A)

    If (X,G) is a complete G-metric space, then (X,d) is a complete metric space.

  2. (B)

    If (X,G) is a sequentially G-compact G-metric space, then (X,d) is a compact metric space.

3 Main results

We first state the following theorem about the existence and uniqueness of a common fixed point, which is a generalization of Theorem 2.7. Furthermore, the techniques of the papers [37, 38] are not applicable to this theorem.

Theorem 3.1 Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)kM(x,y,z)
(11)

for all x, y, z, where k[0, 1 2 ) and

M ( x , y , z ) = max { G ( x , T x , y ) , G ( y , T 2 x , T y ) , G ( T x , T 2 x , T y ) , G ( y , T x , T y ) , G ( x , T x , z ) , G ( z , T 2 x , T z ) , G ( T x , T 2 x , T z ) , G ( z , T x , T y ) , G ( x , y , z ) , G ( x , T x , T x ) , G ( y , T y , T y ) , G ( z , T z , T z ) , G ( z , T x , T x ) , G ( x , T y , T y ) , G ( y , T z , T z ) } .

Then there is a unique xX such that Tx=x.

Proof Let x 0 X. We define a sequence { x n } in the following way:

x n + 1 =T x n ,nN.
(12)

Taking x= x n , z=y= x n + 1 in (11), we find

G(T x n ,T x n + 1 ,T x n + 1 )kM( x n , x n + 1 , x n + 1 ),
(13)

where

M ( x n , x n + 1 , x n + 1 ) = max { G ( x n , T x n , x n + 1 ) , G ( x n + 1 , T 2 x n , T x n + 1 ) , G ( T x n , T 2 x n , T x n + 1 ) , G ( x n + 1 , T x n , T x n + 1 ) , G ( x n , T x n , x n + 1 ) , G ( x n + 1 , T 2 x n , T x n + 1 ) , G ( T x n , T 2 x n , T x n + 1 ) , G ( x n + 1 , T x n , T x n + 1 ) , G ( x n , x n + 1 , x n + 1 ) , G ( x n , T x n , T x n ) , G ( x n + 1 , T x n + 1 , T x n + 1 ) , G ( x n + 1 , T x n + 1 , T x n + 1 ) , G ( x n + 1 , T x n , T x n ) , G ( x n , T x n + 1 , T x n + 1 ) , G ( x n + 1 , T x n + 1 , T x n + 1 ) } = max { G ( x n , x n + 1 , x n + 1 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 1 , x n + 2 ) , G ( x n , x n + 1 , x n + 1 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 1 , x n + 2 ) , G ( x n , x n + 1 , x n + 1 ) , G ( x n , x n + 1 , x n + 1 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 1 , x n + 1 ) , G ( x n , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 2 , x n + 2 ) } = max { G ( x n , x n + 1 , x n + 1 ) , G ( x n + 1 , x n + 2 , x n + 2 ) , G ( x n + 1 , x n + 1 , x n + 2 ) , G ( x n , x n + 2 , x n + 2 ) } .
(14)

Now, we have to examine four cases in (14). For the first case, assume that M( x n , x n + 1 , x n + 1 )=G( x n + 1 , x n + 2 , x n + 2 ). Then the expression (13) turns into

G ( x n + 1 , x n + 2 , x n + 2 ) = G ( T x n , T x n + 1 , T x n + 1 ) k M ( x n , x n + 1 , x n + 1 ) = k G ( x n + 1 , x n + 2 , x n + 2 ) .
(15)

It is a contradiction since 0k< 1 2 . For the second case, assume that M( x n , x n + 1 , x n + 1 )=G( x n + 1 , x n + 1 , x n + 2 ). Regarding (G5) together with the inequality (13), we derive that

G ( x n + 1 , x n + 2 , x n + 2 ) = G ( T x n , T x n + 1 , T x n + 1 ) k G ( x n + 1 , x n + 1 , x n + 2 ) k [ G ( x n + 1 , x n + 2 , x n + 2 ) + G ( x n + 2 , x n + 1 , x n + 2 ) ] ,
(16)

a contradiction since 0k< 1 2 .

For the third case, assume that M( x n , x n + 1 , x n + 1 )=G( x n , x n + 2 , x n + 2 ). By (G5) and the inequality (13), we have

G ( x n + 1 , x n + 2 , x n + 2 ) = G ( T x n , T x n + 1 , T x n + 1 ) k G ( x n , x n + 2 , x n + 2 ) k [ G ( x n , x n + 1 , x n + 1 ) + G ( x n + 1 , x n + 2 , x n + 2 ) ] ,
(17)

which is equivalent to

G( x n + 1 , x n + 2 , x n + 2 )hG( x n , x n + 1 , x n + 1 ),
(18)

where h= k 1 k <1 since 0k< 1 2 .

For the last case, assume that M( x n , x n + 1 , x n + 1 )=G( x n , x n + 1 , x n + 1 ). Then the inequality (13) turns into

G( x n + 1 , x n + 2 , x n + 2 )kG( x n , x n + 1 , x n + 1 ),
(19)

where 0k< 1 2 .

As a result, from (15)-(19) we conclude that

G( x n + 2 , x n + 2 , x n + 1 ) r n + 1 G( x 1 , x 1 , x 0 ),
(20)

where r{h,k} and hence r<1. We show that the sequence { x n } is G-Cauchy. By the rectangle inequality (G5), we have for m>n

G ( x m , x m , x n ) G ( x n + 1 , x n + 1 , x n ) + G ( x n + 2 , x n + 2 , x n + 1 ) + + G ( x m 1 , x m 1 , x m 2 ) + G ( x m , x m , x m 1 ) r n G ( x 1 , x 1 , x 0 ) + r n + 1 G ( x 1 , x 1 , x 0 ) + + r m 2 G ( x 1 , x 1 , x 0 ) + r m 1 G ( x 1 , x 1 , x 0 ) ( i = n m 1 r i ) G ( x 1 , x 1 , x 0 ) .
(21)

Letting n,m in (21), we get that G( x m , x m , x n )0. Hence, { x n } is a G-Cauchy sequence in X. Since (X,G) is G-complete, then there exists x X such that { x n } is G-convergent to x . We shall show that x =T x . Suppose, on the contrary, that x T x . On the other hand, we have x n + 1 =T x n and hence

G ( x n + 1 , T x , T x ) = G ( T x n , T x , T x ) k M ( x n , x , x ) ,
(22)

where

M ( x n , x , x ) = max { G ( x n , T x n , x ) , G ( x , T 2 x n , T x ) , G ( T x n , T 2 x n , T x ) , G ( x , T x n , T x ) , G ( x n , T x n , x ) , G ( x , T 2 x n , T x ) , G ( T x n , T 2 x n , T x ) , G ( x , T x n , T x ) , G ( x n , x , x ) , G ( x n , T x n , T x n ) , G ( x , T x , T x ) , G ( x , T x , T x ) , G ( x , T x n , T x n ) , G ( x n , T x , T x ) , G ( x , T x , T x ) } = max { G ( x n , x n + 1 , x ) , G ( x , x n + 2 , T x ) , G ( x n + 1 , x n + 2 , T x ) , G ( x , x n + 1 , T x ) , G ( x n , x n + 1 , x ) , G ( x , x n + 2 , T x ) , G ( x n + 1 , x n + 2 , T x ) , G ( x , x n + 1 , T x ) , G ( x n , x , x ) , G ( x n , x n + 1 , x n + 1 ) , G ( x , T x , T x ) , G ( x , T x , T x ) , G ( x , x n + 1 , x n + 1 ) , G ( x n , T x , T x ) , G ( x , T x , T x ) } .

Letting n in (22) and using the fact that the metric G is continuous, we get that either

G ( x , T x , T x ) kG ( x , T x , T x )
(23)

or

G ( x , T x , T x ) kG ( x , x , T x ) k [ 2 G ( x , T x , T x ) ]
(24)

by the rectangular property (G5). Since 0k< 1 2 , the inequalities above yield contradictions. Hence we have G( x ,T x ,T x )=0, that is, x =T x .

Finally, we shall show that x is the unique fixed point of T. Suppose that contrary to our claim, there exists another common fixed point t X with t x . From (4) we have

G ( t , t , x ) =G ( T t , T t , T x ) kM ( t , t , x ) ,
(25)

where

M ( t , t , x ) =max { G ( t , t , x ) , G ( t , x , x ) } .

Hence, the inequality (25) is equal to either

G ( t , t , x ) kG ( t , t , x )
(26)

or

G ( t , t , x ) kG ( t , x , x ) 2kG ( t , t , x ) .
(27)

Since 0k< 1 2 , the expressions (26) and (27) yield contradictions. Thus, x is the unique fixed point of T. □

In Theorem 3.1, the interval of constant of the contractive condition can be extended to the interval [0,1) by eliminating the same terms. Since the proof is the mimic of Theorem 3.1, we omit it.

Theorem 3.2 Let (X,G) be a G-metric space. Let T:XX be a mapping such that

G(Tx,Ty,Tz)kM(x,y,z)
(28)

for all x, y, z, where k[0,1) and

M ( x , y , z ) = max { G ( x , T x , y ) , G ( y , T 2 x , T y ) , G ( T x , T 2 x , T y ) , G ( x , T x , z ) , G ( z , T 2 x , T z ) , G ( T x , T 2 x , T z ) , G ( x , y , z ) , G ( x , T x , T x ) , G ( y , T y , T y ) , G ( z , T z , T z ) , G ( z , T x , T x ) , G ( y , T z , T z ) } .

Then there is a unique xX such that Tx=x.

Remark 3.1 Theorem 2.1-Theorem 2.6 are the consequences of Theorem 3.1 and Theorem 3.2.

Inspired by Theorem 2.8, we state the following theorem for which the methods in [37, 38] are not applicable.

Theorem 3.3 Let (X,G) be a complete G-metric space and let T:XX be a given mapping satisfying

G ( T x , T 2 x , T y ) G(x,Tx,y)φ ( G ( x , T x , y ) )
(29)

for all x,yX, where φ:[0,)[0,) is continuous with φ 1 ({0})=0. Then there is a unique xX such that Tx=x.

Proof We first show that if the fixed point of the operator T exists, then it is unique. Suppose, on the contrary, that z, w are two fixed points of T such that zw. Hence, G(z,z,w)0. By (29), we have

G ( T z , T 2 z , T w ) G(z,Tz,w)φ ( G ( z , T z , w ) ) ,
(30)

which is equivalent to

G(z,z,w)G(z,z,w)φ ( G ( z , z , w ) ) ,
(31)

a contradiction. Hence, T has a unique fixed point.

Let x 0 X. We define a sequence { x n } in the following way:

x n + 1 =T x n ,nN.
(32)

If x n 0 = x n 0 + 1 for some n 0 N, then we get the desired result. From now on, we assume that x n = x n + 1 for some nN. Taking x= x n , z=y= x n + 1 in (29), we find

G ( x n + 1 , x n + 2 , x n + 2 ) = G ( T x n , T x n + 1 , T x n + 1 ) G ( x n , T x n , x n + 1 ) φ ( G ( x n , T x n , x n + 1 ) ) = G ( x n , x n + 1 , x n + 1 ) φ ( G ( x n , x n + 1 , x n + 1 ) ) < G ( x n , x n + 1 , x n + 1 ) .
(33)

Hence, {G( x n , x n + 1 , x n + 1 )} is a positive decreasing sequence. Thus, the sequence {G( x n , x n + 1 , x n + 1 )} converges to L0. We shall show that L=0. Suppose, on the contrary, that L>0. Letting n in (33), we find that

LLφ(L).
(34)

It is a contradiction. Hence, we conclude that

lim n { G ( x n , x n + 1 , x n + 1 ) } =0.
(35)

Moreover, by Lemma 2.1, we derive that

lim n { G ( x n , x n , x n + 1 ) } =0.
(36)

Now, we demonstrate that the sequence { x n } is G-Cauchy. Suppose that { x n } is not G-Cauchy. So, there exists ε>0 and subsequences { x n ( k ) } and { x m ( k ) } of { x n } with n(k)>m(k)>k such that

G( x n ( k ) , x m ( k ) , x m ( k ) )εfor all kN.
(37)

Furthermore, corresponding to m(k), one can choose n(k) such that it is the smallest integer with n(k)>m(k) satisfying (37). Thus, we have

G( x n ( k ) 1 , x m ( k ) , x m ( k ) )<εfor all kN.
(38)

By the triangle inequality, we get

εG( x n ( k ) , x m ( k ) , x m ( k ) )G( x n ( k ) , x n ( k ) 1 , x n ( k ) 1 )+G( x n ( k ) 1 , x m ( k ) , x m ( k ) ).
(39)

Letting k in the expression (39) and keeping (36) in mind, we find

lim n G( x n ( k ) , x m ( k ) , x m ( k ) )=ε.
(40)

On the other hand, we have

G ( x n ( k ) + 1 , x m ( k ) + 1 , x m ( k ) + 1 ) G ( x n ( k ) + 1 , x n ( k ) , x n ( k ) ) + G ( x n ( k ) , x m ( k ) , x m ( k ) ) + G ( x m ( k ) , x m ( k ) + 1 , x m ( k ) + 1 )
(41)

and

G ( x n ( k ) , x m ( k ) , x m ( k ) ) G ( x n ( k ) , x n ( k ) + 1 , x n ( k ) + 1 ) + G ( x n ( k ) + 1 , x m ( k ) + 1 , x m ( k ) + 1 ) + G ( x m ( k ) + 1 , x m ( k ) , x m ( k ) ) .
(42)

Letting k in the expression (41)-(42) and regarding (35), (36) and (40), we derive

lim n G( x n ( k ) + 1 , x m ( k ) + 1 , x m ( k ) + 1 )=ε.
(43)

Further, we have

G ( x n ( k ) , x m ( k ) , x m ( k ) ) G ( x n ( k ) , x m ( k ) , x m ( k ) + 1 ) G ( x n ( k ) , x m ( k ) , x m ( k ) ) + G ( x m ( k ) , x m ( k ) , x m ( k ) + 1 )
(44)

by (G3) and the triangle inequality. Letting k in (44) and using (35), (36) and (40), we conclude that

lim n G( x n ( k ) , x m ( k ) , x m ( k ) + 1 )=ε.
(45)

Analogously, we have

G ( x n ( k ) + 1 , x m ( k ) + 1 , x m ( k ) + 1 ) G ( x n ( k ) + 1 , x m ( k ) + 2 , x m ( k ) + 1 ) G ( x n ( k ) + 1 , x m ( k ) + 1 , x m ( k ) + 1 ) + G ( x m ( k ) + 1 , x m ( k ) + 2 , x m ( k ) + 1 )
(46)

by (G3) and the triangle inequality. Letting k in (44) and using (35), (36) and (43), we conclude that

lim n G( x n ( k ) + 1 , x m ( k ) + 2 , x m ( k ) + 1 )=ε.
(47)

Due to (33) and regarding (G4), we obtain

G ( T x m ( k ) , T 2 x m ( k ) , T x n ( k ) ) = G ( x m ( k ) + 1 , x m ( k ) + 2 , x n ( k ) + 1 ) = G ( x n ( k ) + 1 , x m ( k ) + 2 , x m ( k ) + 1 ) G ( x n ( k ) , x m ( k ) , x m ( k ) + 1 ) φ ( G ( x n ( k ) , x m ( k ) , x m ( k ) + 1 ) )
(48)

for all kN. Letting k in the inequality (48) and keeping (45) and (47) in mind, we get

εεϕ(ε),
(49)

a contradiction. Hence, { x n } is a G-Cauchy sequence. Since (X,G) is G-complete, there is zX such that x n z.

We claim that Tz=z. From (33), we have

G ( x n + 1 , x n + 2 , T z ) = G ( T x n , T 2 x n , T z ) G ( x n , T x n , z ) φ ( G ( x n , T x n , z ) ) = G ( x n , x n + 1 , z ) φ ( G ( x n , x n + 1 , z ) ) .
(50)

Letting k in (50), regarding the continuity of G, we get that

G(z,z,Tz)G(z,z,z)φ ( G ( z , z , z ) ) =0.

Hence G(z,z,Tz)=0, that is, Tz=z. □

Remark 3.2 Let X be a nonempty set. We define functions p,d:X×X[0,) in the following way:

d(x,y)= ( y , T 2 x , T y ) andp(x,y)=G ( T x , T 2 x , T y )

for all x,yX. It is easy to see that both mappings p and q do not satisfy the conditions of Definition 2.6. Hence, Theorem 3.1 and Theorem 3.3 cannot be characterized in the context of quasi-metric as it is suggested in [37, 38].

Example 3.1 Let X=[0,), G:X×X×XR be defined by

G(x,y,z)={ 0 if  x = y = z , max { x , y , z } otherwise .

Then (X,G) is a G-complete G-metric space. Let T:XX be defined by

Tx={ 1 4 x if  0 x < 1 / 3 , 1 8 x 4 if  1 / 3 x 1

and φ(t)= 3 4 t for all t[0,+).

Proof For the proof the Example 3.1, we examine the following cases:

  • Let 0x,y<1/3. Then

    G ( T x , T 2 x , T y ) = max { 1 4 x , 1 16 x , 1 4 y } 1 4 max { x , 1 4 x , y } = G ( x , T x , y ) φ ( G ( x , T x , y ) ) .
  • Let 1/3x,y<1. Then

    G ( T x , T 2 x , T y ) = max { 1 8 x 4 , 1 64 x 16 , 1 8 y 4 } 1 4 max { x , 1 8 x 4 , y } = G ( x , T x , y ) φ ( G ( x , T x , y ) ) .
  • Let 0x<1/3y<1. Then

    G ( T x , T 2 x , T y ) = max { 1 4 x , 1 16 x , 1 8 y 4 } 1 4 max { x , 1 4 x , y } = G ( x , T x , y ) φ ( G ( x , T x , y ) ) .
  • Let 0y<1/3x<1. Then

    G ( T x , T 2 x , T y ) = max { 1 8 x 4 , 1 64 x 16 , 1 4 y } 1 4 max { x , 1 8 x 4 , y } = G ( x , T x , y ) φ ( G ( x , T x , y ) ) .

Then

G ( T x , T 2 x , T y ) G(x,Tx,y)φ ( G ( x , T x , y ) ) .

Then the conditions of Theorem 3.3 hold and T has a unique fixed point. Notice that (0,0,0) is the desired fixed point of T. □