Skip to main content

b 2 -Metric spaces and some fixed point theorems

Abstract

The aim of this paper is to establish the structure of b 2 -metric spaces, as a generalization of 2-metric spaces. Some fixed point results for various contractive-type mappings in the context of ordered b 2 -metric spaces are presented. We also provide examples to illustrate the results presented herein, as well as an application to integral equations.

MSC:47H10, 54H25.

1 Introduction

The concept of metric spaces has been generalized in many directions.

The notion of a b-metric space was studied by Czerwik in [1, 2] and many fixed point results were obtained for single and multivalued mappings by Czerwik and many other authors.

On the other hand, the notion of a 2-metric was introduced by Gähler in [3], having the area of a triangle in R 2 as the inspirative example. Similarly, several fixed point results were obtained for mappings in such spaces. Note that, unlike many other generalizations of metric spaces introduced recently, 2-metric spaces are not topologically equivalent to metric spaces and there is no easy relationship between the results obtained in 2-metric and in metric spaces.

In this paper, we introduce a new type of generalized metric spaces, which we call b 2 -metric spaces, as a generalization of both 2-metric and b-metric spaces. Then we prove some fixed point theorems under various contractive conditions in partially ordered b 2 -metric spaces. These include Geraghty-type conditions, conditions using comparison functions and almost generalized weakly contractive conditions. We illustrate these results by appropriate examples, as well as an application to integral equations.

2 Mathematical preliminaries

The notion of a b-metric space was studied by Czerwik in [1, 2].

Definition 1 [1]

Let X be a nonempty set and s1 be a given real number. A function d:X×X R + is a b-metric on X if, for all x,y,zX, the following conditions hold:

(b1) d(x,y)=0 if and only if x=y,

(b2) d(x,y)=d(y,x),

(b3) d(x,z)s[d(x,y)+d(y,z)].

In this case, the pair (X,d) is called a b-metric space.

Note that a b-metric is not always a continuous function of its variables (see, e.g., [[4], Example 2]), whereas an ordinary metric is.

On the other hand, the notion of a 2-metric was introduced by Gähler in [3].

Definition 2 [3]

Let X be a nonempty set and let d: X 3 R be a map satisfying the following conditions:

  1. 1.

    For every pair of distinct points x,yX, there exists a point zX such that d(x,y,z)0.

  2. 2.

    If at least two of three points x, y, z are the same, then d(x,y,z)=0.

  3. 3.

    The symmetry: d(x,y,z)=d(x,z,y)=d(y,x,z)=d(y,z,x)=d(z,x,y)=d(z,y,x) for all x,y,zX.

  4. 4.

    The rectangle inequality: d(x,y,z)d(x,y,t)+d(y,z,t)+d(z,x,t) for all x,y,z,tX.

Then d is called a 2-metric on X and (X,d) is called a 2-metric space.

Definition 3 [3]

Let (X,d) be a 2-metric space, a,bX and r0. The set B(a,b,r)={xX:d(a,b,x)<r} is called a 2-ball centered at a and b with radius r.

The topology generated by the collection of all 2-balls as a subbasis is called a 2-metric topology on X.

Note that a 2-metric is not always a continuous function of its variables, whereas an ordinary metric is.

Remark 1

  1. 1.

    [5] It is straightforward from Definition 2 that every 2-metric is non-negative and every 2-metric space contains at least three distinct points.

  2. 2.

    A 2-metric d(x,y,z) is sequentially continuous in each argument. Moreover, if a 2-metric d(x,y,z) is sequentially continuous in two arguments, then it is sequentially continuous in all three arguments; see [6].

  3. 3.

    A convergent sequence in a 2-metric space need not be a Cauchy sequence; see [6].

  4. 4.

    In a 2-metric space (X,d), every convergent sequence is a Cauchy sequence if d is continuous; see [6].

  5. 5.

    There exists a 2-metric space (X,d) such that every convergent sequence in it is a Cauchy sequence but d is not continuous; see [6].

For some fixed point results on 2-metric spaces, the readers may refer to [515].

Now, we introduce new generalized metric spaces, called b 2 -metric spaces, as a generalization of both 2-metric and b-metric spaces.

Definition 4 Let X be a nonempty set, s1 be a real number and let d: X 3 R be a map satisfying the following conditions:

  1. 1.

    For every pair of distinct points x,yX, there exists a point zX such that d(x,y,z)0.

  2. 2.

    If at least two of three points x, y, z are the same, then d(x,y,z)=0.

  3. 3.

    The symmetry: d(x,y,z)=d(x,z,y)=d(y,x,z)=d(y,z,x)=d(z,x,y)=d(z,y,x) for all x,y,zX.

  4. 4.

    The rectangle inequality: d(x,y,z)s[d(x,y,t)+d(y,z,t)+d(z,x,t)] for all x,y,z,tX.

Then d is called a b 2 -metric on X and (X,d) is called a b 2 -metric space with parameter s.

Obviously, for s=1, b 2 -metric reduces to 2-metric.

Definition 5 Let { x n } be a sequence in a b 2 -metric space (X,d).

  1. 1.

    { x n } is said to be b 2 -convergent to xX, written as lim n x n =x, if for all aX, lim n d( x n ,x,a)=0.

  2. 2.

    { x n } is said to be a b 2 -Cauchy sequence in X if for all aX, lim n d( x n , x m ,a)=0.

  3. 3.

    (X,d) is said to be b 2 -complete if every b 2 -Cauchy sequence is a b 2 -convergent sequence.

The following are some easy examples of b 2 -metric spaces.

Example 1 Let X=[0,+) and d(x,y,z)= [ x y + y z + z x ] p if xyzx, and otherwise d(x,y,z)=0, where p1 is a real number. Evidently, from convexity of function f(x)= x p for x0, then by Jensen inequality we have

( a + b + c ) p 3 p 1 ( a p + b p + c p ) .

So, one can obtain the result that (X,d) is a b 2 -metric space with s 3 p 1 .

Example 2 Let a mapping d: R 3 [0,+) be defined by

d(x,y,z)=min { | x y | , | y z | , | z x | } .

Then d is a 2-metric on , i.e., the following inequality holds:

d(x,y,z)d(x,y,t)+d(y,z,t)+d(z,x,t),

for arbitrary real numbers x, y, z, t. Using convexity of the function f(x)= x p on [0,+) for p1, we obtain that

d p (x,y,z)= [ min { | x y | , | y z | , | z x | } ] p

is a b 2 -metric on with s< 3 p 1 .

Definition 6 Let (X,d) and ( X , d ) be two b 2 -metric spaces and let f:X X be a mapping. Then f is said to be b 2 -continuous at a point zX if for a given ε>0, there exists δ>0 such that xX and d(z,x,a)<δ for all aX imply that d (fz,fx,a)<ε. The mapping f is b 2 -continuous on X if it is b 2 -continuous at all zX.

Proposition 1 Let (X,d) and ( X , d ) be two b 2 -metric spaces. Then a mapping f:X X is b 2 -continuous at a point xX if and only if it is b 2 -sequentially continuous at x; that is, whenever { x n } is b 2 -convergent to x, {f x n } is b 2 -convergent to f(x).

We will need the following simple lemma about the b 2 -convergent sequences in the proof of our main results.

Lemma 1 Let (X,d) be a b 2 -metric space and suppose that { x n } and { y n } are b 2 -convergent to x and y, respectively. Then we have

1 s 2 d(x,y,a) lim inf n d( x n , y n ,a) lim sup n d( x n , y n ,a) s 2 d(x,y,a),

for all a in X. In particular, if y n =y is constant, then

1 s d(x,y,a) lim inf n d( x n ,y,a) lim sup n d( x n ,y,a)sd(x,y,a),

for all a in X.

Proof Using the rectangle inequality in the given b 2 -metric space, it is easy to see that

d ( x , y , a ) = d ( x , a , y ) s d ( x , a , x n ) + s d ( a , y , x n ) + s d ( y , x , x n ) s d ( x , a , x n ) + s 2 [ d ( a , y , y n ) + d ( y , x n , y n ) + d ( x n , a , y n ) ] + s d ( y , x , x n )

and

d ( x n , y n , a ) = d ( x n , a , y n ) s d ( x n , a , x ) + s d ( a , y n , x ) + s d ( y n , x , x n ) s d ( x n , a , x ) + s 2 [ d ( a , y n , y ) + d ( y n , x , y ) + d ( x , a , y ) ] + s d ( y n , x , x n ) .

Taking the lower limit as n in the first inequality and the upper limit as n in the second inequality we obtain the desired result.

If y n =y, then

d(x,y,a)sd(x,y, x n )+sd(y,a, x n )+sd(a,x, x n )

and

d( x n ,y,a)sd( x n ,y,x)+sd(y,a,x)+sd(a, x n ,x).

 □

3 Main results

3.1 Results under Geraghty-type conditions

In 1973, Geraghty [16] proved a fixed point result, generalizing the Banach contraction principle. Several authors proved later various results using Geraghty-type conditions. Fixed point results of this kind in b-metric spaces were obtained by Ðukić et al. in [17].

Following [17], for a real number s1, let F s denote the class of all functions β:[0,)[0, 1 s ) satisfying the following condition:

β( t n ) 1 s as nimplies t n 0as n.

Theorem 1 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be an increasing mapping with respect to such that there exists an element x 0 X with x 0 f x 0 . Suppose that

sd(fx,fy,a)β ( d ( x , y , a ) ) M(x,y,a)
(3.1)

for all aX and for all comparable elements x,yX, where

M(x,y,a)=max { d ( x , y , a ) , d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) } .

If f is b 2 -continuous, then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Starting with the given x 0 , put x n = f n x 0 . Since x 0 f x 0 and f is an increasing function we obtain by induction that

x 0 f x 0 f 2 x 0 f n x 0 f n + 1 x 0 .

Step I: We will show that lim n d( x n , x n + 1 ,a)=0. Since x n x n + 1 for each nN, then by (3.1) we have

s d ( x n , x n + 1 , a ) = s d ( f x n 1 , f x n , a ) β ( d ( x n 1 , x n , a ) ) M ( x n 1 , x n , a ) 1 s d ( x n 1 , x n , a ) d ( x n 1 , x n , a ) ,
(3.2)

because

M ( x n 1 , x n , a ) = max { d ( x n 1 , x n , a ) , d ( x n 1 , f x n 1 , a ) d ( x n , f x n , a ) 1 + d ( f x n 1 , f x n , a ) } = max { d ( x n 1 , x n , a ) , d ( x n 1 , x n , a ) d ( x n , x n + 1 , a ) 1 + d ( x n , x n + 1 , a ) } = d ( x n 1 , x n , a ) .

Therefore, the sequence {d( x n , x n + 1 ,a)} is decreasing. Then there exists r0 such that lim n d( x n , x n + 1 ,a)=r. Suppose that r>0. Then, letting n, from (3.2) we have

1 s rsr lim n β ( d ( x n 1 , x n , a ) ) rr.

So, we have lim n β(d( x n 1 , x n ,a)) 1 s and since β F s we deduce that lim n d( x n 1 , x n ,a)=0 which is a contradiction. Hence, r=0, that is,

lim n d( x n , x n + 1 ,a)=0.
(3.3)

Step II: As {d( x n , x n + 1 ,a)} is decreasing, if d( x n 1 , x n ,a)=0, then d( x n , x n + 1 ,a)=0. Since from part 2 of Definition 4, d( x 0 , x 1 , x 0 )=0, we have d( x n , x n + 1 , x 0 )=0 for all nN. Since d( x m 1 , x m , x m )=0, we have

d( x n , x n + 1 , x m )=0
(3.4)

for all nm1. For 0n<m1, we have m1n+1, and from (3.4) we have

d( x m 1 , x m , x n + 1 )=d( x m 1 , x m , x n )=0.
(3.5)

It implies that

d ( x n , x n + 1 , x m ) s d ( x n , x n + 1 , x m 1 ) + s d ( x n + 1 , x m , x m 1 ) + s d ( x m , x n , x m 1 ) = s d ( x n , x n + 1 , x m 1 ) .

Since d( x n , x n + 1 , x n + 1 )=0, from the above inequality, we have

d( x n , x n + 1 , x m ) s m n 1 d( x n , x n + 1 , x n + 1 )=0
(3.6)

for all 0n<m1. From (3.4) and (3.6), we have

d( x n , x n + 1 , x m )=0
(3.7)

for all n,mN.

Now, for all i,j,kN with i<j, we have

d( x j 1 , x j , x i )=d( x j 1 , x j , x k )=0.
(3.8)

Therefore, from (3.8) and triangular inequality

d ( x i , x j , x k ) s [ d ( x i , x j , x j 1 ) + d ( x j , x k , x j 1 ) + d ( x k , x i , x j 1 ) ] = s d ( x i , x j 1 , x k ) s j i d ( x i , x i , x k ) = 0 .

This proves that for all i,j,kN

d( x i , x j , x k )=0.
(3.9)

Step III: Now, we prove that the sequence { x n } is a b 2 -Cauchy sequence. Using the rectangle inequality and by (3.1) we have

d ( x n , x m , a ) s d ( x n , x m , x n + 1 ) + s d ( x m , a , x n + 1 ) + s d ( a , x n , x n + 1 ) s d ( x n , x n + 1 , x m ) + s 2 [ d ( x m , x m + 1 , a ) + d ( x n + 1 , x m + 1 , a ) + d ( x m , x m + 1 , x n + 1 ) ] + s d ( x n , x n + 1 , a ) s d ( x n , x n + 1 , x m ) + s 2 d ( x m , x m + 1 , a ) + s β ( d ( x n , x m , a ) ) M ( x n , x m , a ) + s 2 d ( x m , x m + 1 , x n + 1 ) + s d ( x n , x n + 1 , a ) .

Letting m,n in the above inequality and applying (3.3) and (3.7) we have

lim m , n d( x n , x m ,a)s lim m , n β ( d ( x n , x m , a ) ) lim m , n M( x n , x m ,a).
(3.10)

Here

d ( x n , x m , a ) M ( x n , x m , a ) = max { d ( x n , x m , a ) , d ( x n , f x n , a ) d ( x m , f x m , a ) 1 + d ( f x n , f x m , a ) } = max { d ( x n , x m , a ) , d ( x n , x n + 1 , a ) d ( x m , x m + 1 , a ) 1 + d ( x n + 1 , x m + 1 , a ) } .

Letting m,n in the above inequality we get

lim m , n M( x n , x m ,a)= lim m , n d( x n , x m ,a).
(3.11)

Hence, from (3.10) and (3.11), we obtain

lim m , n d( x n , x m ,a)s lim m , n β ( d ( x n , x m , a ) ) lim m , n d( x n , x m ,a).
(3.12)

Now we claim that lim m , n d( x n , x m ,a)=0. If, to the contrary, lim m , n d( x n , x m ,a)0, then we get

1 s lim m , n β ( d ( x n , x m , a ) ) .

Since β F s we deduce that

lim m , n d( x n , x m ,a)=0,
(3.13)

which is a contradiction. Consequently, { x n } is a b 2 -Cauchy sequence in X. Since (X,d) is b 2 -complete, the sequence { x n } b 2 -converges to some zX, that is, lim n d( x n ,z,a)=0.

Step IV: Now, we show that z is a fixed point of f.

Using the rectangle inequality, we get

d(fz,z,a)sd(fz,f x n ,z)+sd(z,a,f x n )+sd(a,fz,f x n ).

Letting n and using the continuity of f, we have fz=z. Thus, z is a fixed point of f.

Step V: Finally, suppose that the set of fixed point of f is well ordered. Assume, to the contrary, that u and v are two distinct fixed points of f. Then by (3.1), we have

s d ( u , v , a ) = s d ( f u , f v , a ) β ( d ( u , v , a ) ) M ( u , v , a ) = β ( d ( u , v , a ) ) d ( u , v , a ) < 1 s d ( u , v , a ) ,
(3.14)

because

M ( u , v , a ) = max { d ( u , v , a ) , d ( u , f u , a ) d ( v , f v , a ) 1 + d ( f u , f v , a ) } = max { d ( u , v , a ) , 0 } = d ( u , v , a ) .

Thus, we get sd(u,v,a)< 1 s d(u,v,a), a contradiction. Hence, f has a unique fixed point. The converse is trivial. □

Note that the continuity of f in Theorem 1 can be replaced by certain property of the space itself.

Theorem 2 Under the hypotheses of Theorem  1, without the b 2 -continuity assumption on f, assume that whenever { x n } is a nondecreasing sequence in X such that x n u, one has x n u for all nN. Then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Repeating the proof of Theorem 1, we construct an increasing sequence { x n } in X such that x n zX. Using the assumption on X we have x n z. Now, we show that z=fz. By (3.1) and Lemma 1,

s [ 1 s d ( z , f z , a ) ] s lim sup n d ( x n + 1 , f z , a ) lim sup n β ( d ( x n , z , a ) ) lim sup n M ( x n , z , a ) ,

where

lim n M ( x n , z , a ) = lim n max { d ( x n , z , a ) , d ( x n , f x n , a ) d ( z , f z , a ) 1 + d ( f x n , f z , a ) } = lim n max { d ( x n , z , a ) , d ( x n , x n + 1 , a ) d ( z , f z , a ) 1 + d ( x n + 1 , f z , a ) } = 0 ( see  (3.3) ) .

Therefore, we deduce that d(z,fz,a)0. As a is arbitrary, hence, we have z=fz.

The proof of uniqueness is the same as in Theorem 1. □

If in the above theorems we take β(t)=r, where 0r< 1 s , then we have the following corollary.

Corollary 1 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be an increasing mapping with respect to such that there exists an element x 0 X with x 0 f x 0 . Suppose that for some r, with 0r< 1 s ,

sd(fx,fy,a)rM(x,y,a)

holds for each aX and all comparable elements x,yX, where

M(x,y,a)=max { d ( x , y , a ) , d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) } .

If f is continuous, or, for any nondecreasing sequence { x n } in X such that x n uX one has x n u for all nN, then f has a fixed point. Additionally, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Corollary 2 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be an increasing mapping with respect to such that there exists an element x 0 X with x 0 f x 0 . Suppose that

d(fx,fy,a)αd(x,y,a)+β d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a )

for each aX and all comparable elements x,yX, where α,β0 and α+β 1 s .

If f is continuous, or, for any nondecreasing sequence { x n } in X such that x n uX one has x n u for all nN, then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Since

α d ( x , y , a ) + β d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) ( α + β ) max { d ( x , y , a ) , d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) } .

Putting r=α+β, the conditions of Corollary 1 are satisfied and f has a fixed point. □

Example 3 Let X={(α,0):α[0,+)}{(0,2)} R 2 and let d(x,y,z) denote the square of the area of triangle with vertices x,y,zX, e.g.,

d ( ( α , 0 ) , ( β , 0 ) , ( 0 , 2 ) ) = ( α β ) 2 .

It is easy to check that d is a b 2 -metric with parameter s=2. Introduce an order in X by

(α,0)(β,0)αβ,

with all other pairs of distinct points in X incomparable.

Consider the mapping f:XX given by

f(α,0)= ( α 3 , 0 ) for α[0,+) and f(0,2)=(0,2),

and the function β F 2 given as

β(t)= 1 + t 2 + 4 t for t[0,+).

Then f is an increasing mapping with (α,0)f(α,0) for each α0. If { x n }={( α n ,0)} is a nondecreasing sequence in X, converging to some z=(γ,0), then ( α n ,0)(γ,0) for all nN. Finally, in order to check the contractive condition (3.1), only the case when x=(α,0), y=(β,0), a=(0,2) is nontrivial. But then d(x,y,a)= ( α β ) 2 and

s d ( f x , f y , a ) = 2 d ( ( 1 3 α , 0 ) , ( 1 3 β , 0 ) , ( 0 , 2 ) ) = 2 1 9 ( α β ) 2 1 4 ( α β ) 2 β ( d ( x , y , a ) ) d ( x , y , a ) β ( d ( x , y , a ) ) M ( x , y , a ) .

All the conditions of Theorem 2 are satisfied and f has two fixed points, (0,0) and (0,2). Note that the condition (stated in Theorem 1 and Theorem 2) for the uniqueness of a fixed point is here not satisfied.

3.2 Results using comparison functions

Let Ψ denote the family of all nondecreasing and continuous functions ψ:[0,)[0,) such that lim n ψ n (t)=0 for all t>0, where ψ n denotes the n th iterate of ψ. It is easy to show that, for each ψΨ, the following are satisfied:

  1. (a)

    ψ(t)<t for all t>0;

  2. (b)

    ψ(0)=0.

Theorem 3 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be an increasing mapping with respect to such that there exists an element x 0 X with x 0 f x 0 . Suppose that

sd(fx,fy,a)ψ ( M ( x , y , a ) ) ,
(3.15)

where

M(x,y,a)=max { d ( x , y , a ) , d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) } ,

for some ψΨ and for all elements x,y,aX, with x, y comparable. If f is b 2 -continuous, then f has a fixed point. In addition, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Since x 0 f x 0 and f is an increasing function, we obtain by induction that

x 0 f x 0 f 2 x 0 f n x 0 f n + 1 x 0 .

By letting x n = f n x 0 , we have

x 0 x 1 x 2 x n x n + 1 .

If there exists n 0 N such that x n 0 = x n 0 + 1 , then x n 0 =f x n 0 and so we have nothing to prove. Hence, we assume that x n x n + 1 for all nN.

Step I. We will prove that lim n d( x n , x n + 1 ,a)=0. Using condition (3.15), we obtain

d( x n + 1 , x n ,a)sd( x n + 1 , x n ,a)=sd(f x n ,f x n 1 ,a)ψ ( M ( x n , x n 1 , a ) ) .

Here

M ( x n 1 , x n , a ) = max { d ( x n 1 , x n , a ) , d ( x n 1 , f x n 1 , a ) d ( x n , f x n , a ) 1 + d ( f x n 1 , f x n , a ) } = d ( x n 1 , x n , a ) .

Hence,

d( x n , x n + 1 ,a)sd( x n , x n + 1 ,a)ψ ( d ( x n 1 , x n , a ) ) <d( x n 1 , x n ,a).
(3.16)

By induction, we get

d(a, x n + 1 , x n )ψ ( d ( a , x n , x n 1 ) ) ψ 2 ( d ( a , x n 1 , x n 2 ) ) ψ n ( d ( a , x 1 , x 0 ) ) .

As ψΨ, we conclude that

lim n d( x n , x n + 1 ,a)=0.
(3.17)

From similar arguments as in Theorem 1, since {d( x n , x n + 1 ,a)} is decreasing, we can conclude that

d( x i , x j , x k )=0
(3.18)

for all i,j,kN.

Step II. We will prove that { x n } is a b 2 -Cauchy sequence. Suppose the contrary. Then there exist aX and ε>0 for which we can find two subsequences { x m i } and { x n i } of { x n } such that n i is the smallest index for which

n i > m i >iandd( x m i , x n i ,a)ε.
(3.19)

This means that

d( x m i , x n i 1 ,a)<ε.
(3.20)

From (3.19) and using the rectangle inequality, we get

εd( x m i , x n i ,a)sd( x m i , x n i , x m i + 1 )+sd( x m i + 1 , x n i ,a)+sd( x m i + 1 , x m i ,a).

Taking the upper limit as i, from (3.17) and (3.18) we get

ε s lim sup i d( x m i + 1 , x n i ,a).
(3.21)

From the definition of M(x,y,a) we have

M ( x m i , x n i 1 , a ) = max { d ( x m i , x n i 1 , a ) , d ( x m i , f x m i , a ) d ( x n i 1 , f x n i 1 , a ) 1 + d ( f x m i , f x n i 1 , a ) } = max { d ( x m i , x n i 1 , a ) , d ( x m i , a , x m i + 1 ) d ( x n i 1 , a , x n i ) 1 + d ( x m i + 1 , x n i , a ) }

and if i, by (3.17) and (3.20) we have

lim sup i M( x m i , x n i 1 ,a)ε.

Now, from (3.15) we have

sd( x m i + 1 , x n i ,a)=sd(f x m i ,f x n i 1 ,a)ψ ( M ( x m i , x n i 1 , a ) ) .

Again, if i by (3.21) we obtain

ε=s ε s s lim sup i d( x m i + 1 , x n i ,a)ψ(ε)<ε,

which is a contradiction. Consequently, { x n } is a b 2 -Cauchy sequence in X. Therefore, the sequence { x n } b 2 -converges to some zX, that is, lim n d( x n ,z,a)=0 for all aX.

Step III. Now we show that z is a fixed point of f.

Using the rectangle inequality, we get

d(z,fz,a)sd(z,fz,f x n )+sd(f x n ,fz,a)+sd(f x n ,z,a).

Letting n and using the continuity of f, we get

d(z,fz,a)0.

Hence, we have fz=z. Thus, z is a fixed point of f.

The uniqueness of the fixed point can be proved in the same manner as in Theorem 1. □

Theorem 4 Under the hypotheses of Theorem  3, without the b 2 -continuity assumption on f, assume that whenever { x n } is a nondecreasing sequence in X such that x n uX, one has x n u for all nN. Then f has a fixed point. In addition, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Following the proof of Theorem 3, we construct an increasing sequence { x n } in X such that x n zX. Using the given assumption on X we have x n z. Now, we show that z=fz. By (3.15) we have

sd(fz, x n ,a)=sd(fz,f x n 1 ,a)ψ ( M ( z , x n 1 , a ) ) ,
(3.22)

where

M(z, x n 1 ,a)=max { d ( z , x n 1 , a ) , d ( z , f z , a ) d ( x n 1 , f x n 1 , a ) 1 + d ( f z , f x n 1 , a ) } .

Letting n in the above relation, we get

lim sup n M(z, x n 1 ,a)=0.
(3.23)

Again, taking the upper limit as n in (3.22) and using Lemma 1 and (3.23) we get

s [ 1 s d ( z , f z , a ) ] s lim sup n d ( x n , f z , a ) lim sup n ψ ( M ( z , x n 1 , a ) ) = 0 .

So we get d(z,fz,a)=0, i.e., fz=z. □

Corollary 3 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be an increasing mapping with respect to such that there exists an element x 0 X with x 0 f x 0 . Suppose that

sd(fx,fy,a)rM(x,y,a),

where 0r<1 and

M(x,y,a)=max { d ( x , y , a ) , d ( x , f x , a ) d ( y , f y , a ) 1 + d ( f x , f y , a ) } ,

for all elements x,y,aX with x, y comparable. If f is continuous, or, whenever { x n } is a nondecreasing sequence in X such that x n uX, one has x n u for all nN, then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Example 4 Let X={A,B,C,D} be ordered by ABC, with all other pairs of distinct points incomparable. Define d: X 3 R by

d(A,B,C)=0,d(A,B,D)=1,d(A,C,D)=4,d(B,C,D)=2,

with symmetry in all variables and with d(x,y,z)=0 when at least two of the arguments are equal. Then it is easy to check that (X,d) is a complete b 2 -metric space with s= 4 3 .

Consider the mapping f:XX given as

f= ( A B C D A A B D )

and a comparison function ψ(t)= 2 3 t. Then f is a nondecreasing mapping w.r.t. and there exists x 0 X such that x 0 f x 0 . The only nontrivial cases for checking the contractive condition (3.15) are when a=D and x=A, y=C or x=B, y=C (or vice versa). Then we have

sd(fA,fC,D)= 4 3 d(A,B,D)= 4 3 < 2 3 4=ψ(4)=ψ ( d ( A , C , D ) ) ψ ( M ( A , C , D ) ) ,

resp.

sd(fB,fC,D)= 4 3 d(A,B,D)= 4 3 = 2 3 2=ψ(2)=ψ ( d ( B , C , D ) ) ψ ( M ( B , C , D ) ) .

Hence, all the conditions of Theorem 3 are fulfilled. The mapping f has two fixed points (A and D).

3.3 Results for almost generalized weakly contractive mappings

Berinde in [1821] initiated the concept of almost contractions and obtained many interesting fixed point theorems. Results with similar conditions were obtained, e.g., in [22] and [23]. In this section, we define the notion of almost generalized ( ψ , φ ) s , a -contractive mapping and we prove some new results. In particular, we extend Theorems 2.1, 2.2 and 2.3 of Ćirić et al. in [24] to the setting of b 2 -metric spaces.

Recall that Khan et al. introduced in [25] the concept of an altering distance function as follows.

Definition 7 [25]

A function φ:[0,+)[0,+) is called an altering distance function, if the following properties hold:

  1. 1.

    φ is continuous and nondecreasing.

  2. 2.

    φ(t)=0 if and only if t=0.

Let (X,d) be a b 2 -metric space and let f:XX be a mapping. For x,y,aX, set

M a (x,y)=max { d ( x , y , a ) , d ( x , f x , a ) , d ( y , f y , a ) , d ( x , f y , a ) + d ( y , f x , a ) 2 s }

and

N a (x,y)=min { d ( x , f x , a ) , d ( x , f y , a ) , d ( y , f x , a ) , d ( y , f y , a ) } .

Definition 8 Let (X,d) be a b 2 -metric space. We say that a mapping f:XX is an almost generalized ( ψ , φ ) s , a -contractive mapping if there exist L0 and two altering distance functions ψ and φ such that

ψ ( s d ( f x , f y , a ) ) ψ ( M a ( x , y ) ) φ ( M a ( x , y ) ) +Lψ ( N a ( x , y ) )
(3.24)

for all x,y,aX.

Now, let us prove our new result.

Theorem 5 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be a continuous mapping, nondecreasing with respect to . Suppose that f satisfies condition (3.24), for all elements x,y,aX, with x, y comparable. If there exists x 0 X such that x 0 f x 0 , then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Starting with the given x 0 , define a sequence { x n } in X such that x n + 1 =f x n , for all n0. Since x 0 f x 0 = x 1 and f is nondecreasing, we have x 1 =f x 0 x 2 =f x 1 , and by induction

x 0 x 1 x n x n + 1 .

If x n = x n + 1 , for some nN, then x n =f x n and hence x n is a fixed point of f. So, we may assume that x n x n + 1 , for all nN. By (3.24), we have

ψ ( d ( x n , x n + 1 , a ) ) ψ ( s d ( x n , x n + 1 , a ) ) = ψ ( s d ( f x n 1 , f x n , a ) ) ψ ( M a ( x n 1 , x n ) ) φ ( M a ( x n 1 , x n ) ) + L ψ ( N a ( x n 1 , x n ) ) ,
(3.25)

where

M a ( x n 1 , x n ) = max { d ( x n 1 , x n , a ) , d ( x n 1 , f x n 1 , a ) , d ( x n , f x n , a ) , d ( x n 1 , f x n , a ) + d ( x n , f x n 1 , a ) 2 s } = max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , a ) 2 s } max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 }
(3.26)

and

N a ( x n 1 , x n ) = min { d ( x n 1 , f x n 1 , a ) , d ( x n 1 , f x n , a ) , d ( x n , f x n 1 , a ) , d ( x n , f x n , a ) } = min { d ( x n 1 , x n , a ) , d ( x n 1 , x n + 1 , a ) , 0 , d ( x n , x n + 1 , a ) } = 0 .
(3.27)

From (3.25)–(3.27) and the properties of ψ and φ, we get

ψ ( d ( x n , x n + 1 , a ) ) ψ ( max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 } ) φ ( max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , a ) 2 s } ) .
(3.28)

If

max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 } = d ( x n , x n + 1 , a ) ,

then by (3.28) we have

ψ ( d ( x n , x n + 1 , a ) ) ψ ( d ( x n , x n + 1 , a ) ) φ ( max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , a ) 2 s } ) ,

which gives a contradiction.

If d( x n 1 , x n + 1 , x n )=0, then

max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 } = d ( x n 1 , x n , a ) ,

therefore (3.28) becomes

ψ ( d ( x n , x n + 1 , a ) ) ψ ( d ( x n , x n 1 , a ) ) φ ( max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , a ) 2 s } ) ψ ( d ( x n , x n 1 , a ) ) .
(3.29)

Thus, {d( x n , x n + 1 ,a):nN{0}} is a nonincreasing sequence of positive numbers. Hence, there exists r0 such that

lim n d( x n , x n + 1 ,a)=r.

Letting n in (3.29), we get

ψ(r)ψ(r)φ ( max { r , r , lim n d ( x n 1 , x n + 1 , a ) 2 s } ) ψ(r).

Therefore,

φ ( max { r , r , lim n d ( x n 1 , x n + 1 , a ) 2 s } ) =0,

and hence r=0. Thus, we have

lim n d( x n , x n + 1 ,a)=0,
(3.30)

for each aX.

Note that if d( x n 1 , x n + 1 , x n )0 and

max { d ( x n 1 , x n , a ) , d ( x n , x n + 1 , a ) , d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 } = d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , a , x n ) + d ( a , x n 1 , x n ) 2 .

Then, by (3.28) and taking a= x n 1 , we have

ψ ( d ( x n , x n + 1 , x n 1 ) ) ψ ( d ( x n 1 , x n + 1 , x n ) + d ( x n + 1 , x n 1 , x n ) + d ( x n 1 , x n 1 , x n ) 2 ) φ ( max { d ( x n 1 , x n , x n 1 ) , d ( x n , x n + 1 , x n 1 ) , d ( x n 1 , x n + 1 , x n 1 ) 2 s } ) ,

which gives d( x n 1 , x n + 1 , x n )=0, a contradiction.

Next, we show that { x n } is a b 2 -Cauchy sequence in X. For this purpose, we use the following relation (see (3.9) and (3.18)):

d( x i , x j , x k )=0,
(3.31)

for all i,j,kN (note that this can obtained as {d( x n , x n + 1 ,a):nN{0}} is a nonincreasing sequence of positive numbers).

Suppose the contrary, that is, { x n } is not a b 2 -Cauchy sequence. Then there exist aX and ε>0 for which we can find two subsequences { x m i } and { x n i } of { x n } such that n i is the smallest index for which

n i > m i >i,d( x m i , x n i ,a)ε.
(3.32)

This means that

d( x m i , x n i 1 ,a)<ε.
(3.33)

Using (3.33) and taking the upper limit as i, we get

lim sup n d( x m i , x n i 1 ,a)ε.
(3.34)

On the other hand, we have

d( x m i , x n i ,a)sd( x m i , x n i , x m i + 1 )+sd( x n i ,a, x m i + 1 )+sd(a, x m i , x m i + 1 ).

Using (3.30), (3.31), (3.32), and taking the upper limit as i, we get

ε s lim sup n d( x m i + 1 , x n i ,a).
(3.35)

Again, using the rectangular inequality, we have

d( x m i + 1 , x n i 1 ,a)sd( x m i + 1 , x n i 1 , x m i )+sd( x n i 1 ,a, x m i )+sd(a, x m i + 1 , x m i ),

and

d( x m i , x n i ,a)sd( x m i , x n i , x n i 1 )+sd( x n i ,a, x n i 1 )+sd(a, x m i , x n i 1 ).

Taking the upper limit as i in the first inequality above, and using (3.30), (3.31), and (3.34) we get

lim sup n d( x m i + 1 , x n i 1 ,a)εs.
(3.36)

Similarly, taking the upper limit as i in the second inequality above, and using (3.30), (3.31), and (3.33), we get

lim sup n d( x m i , x n i ,a)εs.
(3.37)

From (3.24), we have

ψ ( s d ( x m i + 1 , x n i , a ) ) = ψ ( s d ( f x m i , f x n i 1 , a ) ) ψ ( M a ( x m i , x n i 1 ) ) φ ( M a ( x m i , x n i 1 ) ) + L ψ ( N a ( x m i , x n i 1 ) ) ,
(3.38)

where

M a ( x m i , x n i 1 ) = max { d ( x m i , x n i 1 , a ) , d ( x m i , f x m i , a ) , d ( x n i 1 , f x n i 1 , a ) , d ( x m i , f x n i 1 , a ) + d ( f x m i , x n i 1 , a ) 2 s } = max { d ( x m i , x n i 1 , a ) , d ( x m i , x m i + 1 , a ) , d ( x n i 1 , x n i , a ) , d ( x m i , x n i , a ) + d ( x m i + 1 , x n i 1 , a ) 2 s } ,
(3.39)

and

N a ( x m i , x n i 1 ) = min { d ( x m i , f x m i , a ) , d ( x m i , f x n i 1 , a ) , d ( x n i 1 , f x m i , a ) , d ( x n i 1 , f x n i 1 , a ) } = min { d ( x m i , x m i + 1 , a ) , d ( x m i , x n i , a ) , d ( x n i 1 , x m i + 1 , a ) , d ( x n i 1 , x n i , a ) } .
(3.40)

Taking the upper limit as i in (3.39) and (3.40) and using (3.30), (3.34), (3.36), and (3.37), we get

lim sup n M a ( x m i 1 , x n i 1 ) = max { lim sup n d ( x m i , x n i 1 , a ) , 0 , 0 , lim sup n d ( x m i , x n i , a ) + lim sup n d ( x m i + 1 , x n i 1 , a ) 2 s } max { ε , ε s + ε s 2 s } = ε .
(3.41)

So, we have

lim sup n M a ( x m i 1 , x n i 1 )ε,
(3.42)

and

lim sup n N a ( x m i , x n i 1 )=0.
(3.43)

Now, taking the upper limit as i in (3.38) and using (3.35), (3.42), and (3.43) we have

ψ ( s ε s ) ψ ( s lim sup n d ( x m i + 1 , x n i , a ) ) ψ ( lim sup n M a ( x m i , x n i 1 ) ) lim inf n φ ( M a ( x m i , x n i 1 ) ) ψ ( ε ) φ ( lim inf n M a ( x m i , x n i 1 ) ) ,

which further implies that

φ ( lim inf n M a ( x m i , x n i 1 ) ) =0,

so lim inf n M a ( x m i , x n i 1 )=0, a contradiction to (3.32). Thus, { x n + 1 =f x n } is a b 2 -Cauchy sequence in X.

As X is a b 2 -complete space, there exists uX such that x n u as n, that is,

lim n x n + 1 = lim n f x n =u.

Now, using continuity of f and the rectangle inequality, we get

d(u,fu,a)sd(u,fu,f x n )+sd(fu,a,f x n )+sd(a,u,f x n ).

Letting n, we get

d(u,fu,a)s lim n d(u,fu,f x n )+s lim n d(fu,a,f x n )+s lim n d(a,u,f x n )=0.

Therefore, we have fu=u. Thus, u is a fixed point of f.

The uniqueness of fixed point can be proved as in Theorem 1. □

Note that the continuity of f in Theorem 5 can be replaced by a property of the space.

Theorem 6 Under the hypotheses of Theorem  5, without the continuity assumption on f, assume that whenever { x n } is a nondecreasing sequence in X such that x n xX, one has x n x, for all nN. Then f has a fixed point in X. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Following similar arguments to those given in the proof of Theorem 5, we construct an increasing sequence { x n } in X such that x n u, for some uX. Using the assumption on X, we have x n u, for all nN. Now, we show that fu=u. By (3.24), we have

ψ ( s d ( x n + 1 , f u , a ) ) = ψ ( s d ( f x n , f u , a ) ) ψ ( M a ( x n , u ) ) φ ( M a ( x n , u ) ) + L ψ ( N a ( x n , u ) ) ,
(3.44)

where

M a ( x n , u ) = max { d ( x n , u , a ) , d ( x n , f x n , a ) , d ( u , f u , a ) , d ( x n , f u , a ) + d ( f x n , u , a ) 2 s } = max { d ( x n , u , a ) , d ( x n , x n + 1 , a ) , d ( u , f u , a ) , d ( x n , f u , a ) + d ( x n + 1 , u , a ) 2 s }
(3.45)

and

N a ( x n , u ) = min { d ( x n , f x n , a ) , d ( x n , f u , a ) , d ( u , f x n , a ) , d ( u , f u , a ) } = min { d ( x n , x n + 1 , a ) , d ( x n , f u , a ) , d ( u , x n + 1 , a ) , d ( u , f u , a ) } .
(3.46)

Letting n in (3.45) and (3.46) and using Lemma 1, we get

1 s d ( u , f u , a ) 2 s lim inf n M a ( x n , u ) lim sup n M a ( x n , u ) max { d ( u , f u , a ) , s d ( u , f u , a ) 2 s } = d ( u , f u , a ) ,
(3.47)

and

N a ( x n ,u)0.

Again, taking the upper limit as i in (3.44) and using Lemma 1 and (3.47) we get

ψ ( d ( u , f u , a ) ) = ψ ( s 1 s d ( u , f u , a ) ) ψ ( s lim sup n d ( x n + 1 , f u , a ) ) ψ ( lim sup n M a ( x n , u ) ) lim inf n φ ( M a ( x n , u ) ) ψ ( d ( u , f u , a ) ) φ ( lim inf n M a ( x n , u ) ) .

Therefore, φ( lim inf n M a ( x n ,u))0, equivalently, lim inf n M a ( x n ,u)=0. Thus, from (3.47) we get u=fu and hence u is a fixed point of f. □

Corollary 4 Let (X,) be a partially ordered set and suppose that there exists a b 2 -metric d on X such that (X,d) is a b 2 -complete b 2 -metric space. Let f:XX be a nondecreasing continuous mapping with respect to . Suppose that there exist k[0,1) and L0 such that

d ( f x , f y , a ) k s max { d ( x , y , a ) , d ( x , f x , a ) , d ( y , f y , a ) , d ( x , f y , a ) + d ( y , f x , a ) 2 s } + L s min { d ( x , f x , a ) , d ( y , f x , a ) } ,

for all elements x,y,aX with x, y comparable. If there exists x 0 X such that x 0 f x 0 , then f has a fixed point. Moreover, the set of fixed points of f is well ordered if and only if f has one and only one fixed point.

Proof Follows from Theorem 5 by taking ψ(t)=t and φ(t)=(1k)t, for all t[0,+). □

Corollary 5 Under the hypotheses of Corollary  4, without the continuity assumption of f, let for any nondecreasing sequence { x n } in X such that x n xX we have x n x, for all nN. Then f has a fixed point in X.

4 An application to integral equations

As an application of our results, inspired by [26], we will consider the following integral equation:

x(t)=h(t)+ 0 T g(t,s)F ( s , x ( s ) ) ds,tI=[0,T].
(4.1)

Consider the set X= C R (I) of all real continuous functions on I, ordered by the natural relation

xyx(t)y(t)for all tI,

and take arbitrary real p>1. We will use the following assumptions.

  1. (I)

    h:IR, g:I×R[0,+) and F:I×RR are continuous functions;

  2. (II)

    for x,yX,

    xy 0 T g(,s)F ( s , x ( s ) ) ds 0 T g(,s)F ( s , y ( s ) ) ds;
  3. (III)

    for some 0r<1 and all x,y,aX, with x and y comparable (w.r.t. ),

    3 p 1 [ max 0 t T min { | 0 T g ( t , s ) [ F ( s , x ( s ) ) F ( s , y ( s ) ) ] d s | , | h ( t ) + 0 T g ( t , s ) F ( s , y ( s ) ) d s a ( t ) | , | h ( t ) + 0 T g ( t , s ) F ( s , x ( s ) ) d s a ( t ) | } ] p r [ max 0 t T min { | x ( t ) y ( t ) | , | y ( t ) a ( t ) | , | x ( t ) a ( t ) | } ] p ;
  4. (IV)

    there exists x 0 X such that x 0 (t)h(t)+ 0 T g(t,s)F(s, x 0 (s))ds for all tI.

Let d:X×X×X[0,) be defined by

d(x,y,z)= [ max 0 t T min { | x ( t ) y ( t ) | , | y ( t ) z ( t ) | , | x ( t ) z ( t ) | } ] p .

Then (X,d) is a b 2 -complete b 2 -metric space, with s 3 p 1 (similarly as in Example 2). We have the following result.

Theorem 7 Let the functions h, g, F satisfy conditions (I)-(IV) and let the space (X,,d) satisfy the requirement that if { x n } is a sequence in X, nondecreasing w.r.t. , and converging (in d) to some uX, then x n u for all nN. Then the integral equation (4.1) has a solution in X.

Proof Define the mapping f:XX by

fx(t)=h(t)+ 0 T g(t,s)F ( s , x ( s ) ) ds,tI.

Then all the conditions of Corollary 3 are fulfilled. In particular, condition (III) implies that, for all x,y,aX, with x, y comparable, we have

sd(fx,fy,a) 3 p 1 d(fx,fy,a)rd(x,y,a)rM(x,y,a).

Hence, using Corollary 3, we conclude that there exists a fixed point xX of f, which is obviously a solution of (4.1). □

References

  1. Czerwik S: Contraction mappings in b -metric spaces. Acta Math. Inform. Univ. Ostrav. 1993, 1: 5–11.

    MathSciNet  Google Scholar 

  2. Czerwik S: Nonlinear set-valued contraction mappings in b -metric spaces. Atti Semin. Mat. Fis. Univ. Modena 1998, 46: 263–276.

    MathSciNet  Google Scholar 

  3. Gähler VS: 2-metrische Räume und ihre topologische Struktur. Math. Nachr. 1963, 26: 115–118. 10.1002/mana.19630260109

    Article  MathSciNet  Google Scholar 

  4. Hussain N, Parvaneh V, Roshan JR, Kadelburg Z: Fixed points of cyclic weakly (ψ,φ,L,A,B)-contractive mappings in ordered b -metric spaces with applications. Fixed Point Theory Appl. 2013., 2013: Article ID 256

    Google Scholar 

  5. Dung NV, Le Hang VT: Fixed point theorems for weak C -contractions in partially ordered 2-metric spaces. Fixed Point Theory Appl. 2013., 2013: Article ID 161

    Google Scholar 

  6. Naidu SVR, Prasad JR: Fixed point theorems in 2-metric spaces. Indian J. Pure Appl. Math. 1986, 17(8):974–993.

    MathSciNet  Google Scholar 

  7. Aliouche A, Simpson C: Fixed points and lines in 2-metric spaces. Adv. Math. 2012, 229: 668–690. 10.1016/j.aim.2011.10.002

    Article  MathSciNet  Google Scholar 

  8. Deshpande B, Chouhan S: Common fixed point theorems for hybrid pairs of mappings with some weaker conditions in 2-metric spaces. Fasc. Math. 2011, 46: 37–55.

    MathSciNet  Google Scholar 

  9. Freese RW, Cho YJ, Kim SS: Strictly 2-convex linear 2-normed spaces. J. Korean Math. Soc. 1992, 29(2):391–400.

    MathSciNet  Google Scholar 

  10. Iseki K: Fixed point theorems in 2-metric spaces. Math. Semin. Notes 1975, 3: 133–136.

    Google Scholar 

  11. Iseki K: Mathematics on 2-normed spaces. Bull. Korean Math. Soc. 1976, 13(2):127–135.

    Google Scholar 

  12. Lahiri BK, Das P, Dey LK: Cantor’s theorem in 2-metric spaces and its applications to fixed point problems. Taiwan. J. Math. 2011, 15: 337–352.

    MathSciNet  Google Scholar 

  13. Lai SN, Singh AK: An analogue of Banach’s contraction principle in 2-metric spaces. Bull. Aust. Math. Soc. 1978, 18: 137–143. 10.1017/S0004972700007887

    Article  Google Scholar 

  14. Popa V, Imdad M, Ali J: Using implicit relations to prove unified fixed point theorems in metric and 2-metric spaces. Bull. Malays. Math. Soc. 2010, 33: 105–120.

    MathSciNet  Google Scholar 

  15. Ahmed MA: A common fixed point theorem for expansive mappings in 2-metric spaces and its application. Chaos Solitons Fractals 2009, 42(5):2914–2920. 10.1016/j.chaos.2009.04.034

    Article  MathSciNet  Google Scholar 

  16. Geraghty M: On contractive mappings. Proc. Am. Math. Soc. 1973, 40: 604–608. 10.1090/S0002-9939-1973-0334176-5

    Article  MathSciNet  Google Scholar 

  17. Ðukić D, Kadelburg Z, Radenović S: Fixed points of Geraghty-type mappings in various generalized metric spaces. Abstr. Appl. Anal. 2011., 2011: Article ID 561245

    Google Scholar 

  18. Berinde V: On the approximation of fixed points of weak contractive mappings. Carpath. J. Math. 2003, 19: 7–22.

    MathSciNet  Google Scholar 

  19. Berinde V: Approximating fixed points of weak contractions using the Picard iteration. Nonlinear Anal. Forum 2004, 9: 43–53.

    MathSciNet  Google Scholar 

  20. Berinde V: General contractive fixed point theorems for Ćirić-type almost contraction in metric spaces. Carpath. J. Math. 2008, 24: 10–19.

    MathSciNet  Google Scholar 

  21. Berinde V: Some remarks on a fixed point theorem for Ćirić-type almost contractions. Carpath. J. Math. 2009, 25: 157–162.

    MathSciNet  Google Scholar 

  22. Babu GVR, Sandhya ML, Kameswari MVR: A note on a fixed point theorem of Berinde on weak contractions. Carpath. J. Math. 2008, 24: 8–12.

    MathSciNet  Google Scholar 

  23. Roshan JR, Parvaneh V, Sedghi S, Shobkolaei N, Shatanawi W: Common fixed points of almost generalized ( ψ , φ ) s -contractive mappings in ordered b -metric spaces. Fixed Point Theory Appl. 2013., 2013: Article ID 159

    Google Scholar 

  24. Ćirić L, Abbas M, Saadati R, Hussain N: Common fixed points of almost generalized contractive mappings in ordered metric spaces. Appl. Math. Comput. 2011, 217: 5784–5789. 10.1016/j.amc.2010.12.060

    Article  MathSciNet  Google Scholar 

  25. Khan MS, Swaleh M, Sessa S: Fixed point theorems by altering distances between the points. Bull. Aust. Math. Soc. 1984, 30: 1–9. 10.1017/S0004972700001659

    Article  MathSciNet  Google Scholar 

  26. Fathollahi S, Hussain N, Khan LA: Fixed point results for modified weak and rational α - ψ -contractions in ordered 2-metric spaces. Fixed Point Theory Appl. 2014., 2014: Article ID 6

    Google Scholar 

Download references

Acknowledgements

The authors are highly indebted to the referees of this paper who helped us to improve it in several places. The fourth author is thankful to the Ministry of Education, Science and Technological Development of Serbia.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zead Mustafa.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

All authors contributed equally and significantly in writing this paper. All authors read and approved the final manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mustafa, Z., Parvaneh, V., Roshan, J.R. et al. b 2 -Metric spaces and some fixed point theorems. Fixed Point Theory Appl 2014, 144 (2014). https://doi.org/10.1186/1687-1812-2014-144

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1687-1812-2014-144

Keywords