Skip to main content

Fixed point results via α-admissible mappings and cyclic contractive mappings in partial b-metric spaces

Abstract

Considering α-admissible mappings in the setup of partial b-metric spaces, we establish some fixed and common fixed point results for ordered cyclic weakly (ψ,φ,L,A,B)-contractive mappings in complete ordered partial b-metric spaces. Our results extend several known results in the literature. Examples are also provided in support of our results.

MSC:47H10, 54H25.

1 Introduction

There are a lot of generalizations of the concept of metric space. The concepts of b-metric space and partial metric space were introduced by Czerwik [1] and Matthews [2], respectively. Combining these two notions, Shukla [3] introduced another generalization which is called a partial b-metric space. Also, in [4], Mustafa et al. introduced a modified version of partial b-metric spaces. In fact, the advantage of their definition of partial b-metric is that by using it one can define a dependent b-metric which is called the b-metric associated with the partial b-metric.

Definition 1.1 [4]

Let X be a (nonempty) set and s1 be a given real number. A function p b :X×X R + is a partial b-metric if, for all x,y,zX, the following conditions are satisfied:

( p b 1 ) x=y p b (x,x)= p b (x,y)= p b (y,y),

( p b 2 ) p b (x,x) p b (x,y),

( p b 3 ) p b (x,y)= p b (y,x),

( p b 4 ) p b (x,y)s( p b (x,z)+ p b (z,y) p b (z,z))+( 1 s 2 )( p b (x,x)+ p b (y,y)).

The pair (X, p b ) is called a partial b-metric space.

Example 1.2 [3]

Let X= R + , q>1 be a constant, and p b :X×X R + be defined by

p b (x,y)= [ max { x , y } ] q +|xy | q for all x,yX.

Then (X, p b ) is a partial b-metric space with the coefficient s= 2 q 1 >1, but it is neither a b-metric nor a partial metric space.

Some more examples of partial b-metrics can be constructed with the help of following propositions.

Proposition 1.3 [3]

Let X be a nonempty set and let p be a partial metric and d be a b-metric with the coefficient s1 on X. Then the function p b :X×X R + defined by p b (x,y)=p(x,y)+d(x,y), for all x,yX, is a partial b-metric on X with the coefficient s.

Proposition 1.4 [3]

Let (X,p) be a partial metric space and q1. Then (X, p b ) is a partial b-metric space with the coefficient s= 2 q 1 , where p b is defined by p b (x,y)= [ p ( x , y ) ] q .

Proposition 1.5 [4]

Every partial b-metric p b defines a b-metric d p b , where

d p b (x,y)=2 p b (x,y) p b (x,x) p b (y,y)

for all x,yX.

Now, we recall some definitions and propositions in a partial b-metric space.

Definition 1.6 [4]

Let (X, p b ) be a partial b-metric space. Then for an xX and an ϵ>0, the p b -ball with center x and radius ϵ is

B p b (x,ϵ)= { y X p b ( x , y ) < p b ( x , x ) + ϵ } .

Proposition 1.7 [4]

Let (X, p b ) be a partial b-metric space, xX, and r>0. If y B p b (x,r) then there exists a δ>0 such that B p b (y,δ) B p b (x,r).

Thus, from the above proposition the family of all p b -balls

Δ= { B p b ( x , r ) x X , r > 0 }

is a base of a T 0 topology τ p b on X which we call the p b -metric topology.

The topological space (X, p b ) is T 0 , but it does not need to be T 1 .

Definition 1.8 [4]

A sequence { x n } in a partial b-metric space (X, p b ) is said to be:

  1. (i)

    p b -convergent to a point xX if lim n p b (x, x n )= p b (x,x).

  2. (ii)

    A p b -Cauchy sequence if lim n , m p b ( x n , x m ) exists (and is finite).

  3. (iii)

    A partial b-metric space (X, p b ) is said to be p b -complete if every p b -Cauchy sequence { x n } in X p b -converges to a point xX such that lim n , m p b ( x n , x m )= lim n , m p b ( x n ,x)= p b (x,x).

Lemma 1.9 [4]

  1. (1)

    A sequence { x n } is a p b -Cauchy sequence in a partial b-metric space (X, p b ) if and only if it is a b-Cauchy sequence in the b-metric space (X, d p b ).

  2. (2)

    A partial b-metric space (X, p b ) is p b -complete if and only if the b-metric space (X, d p b ) is b-complete. Moreover, lim n d p b (x, x n )=0 if and only if

    lim n p b (x, x n )= lim n , m p b ( x n , x m )= p b (x,x).

Definition 1.10 [4]

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

Proposition 1.11 [4]

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

Definition 1.12 A triple (X,, p b ) is called an ordered partial b-metric space if (X,) is a partially ordered set and p b is a partial b-metric on X.

The following crucial lemma is useful in proving our main results.

Lemma 1.13 [4]

Let (X, p b ) be a partial b-metric space with the coefficient s>1 and suppose that { x n } and { y n } are convergent to x and y, respectively. Then we have

1 s 2 p b ( x , y ) 1 s p b ( x , x ) p b ( y , y ) lim inf n p b ( x n , y n ) lim sup n p b ( x n , y n ) s p b ( x , x ) + s 2 p b ( y , y ) + s 2 p b ( x , y ) .

In particular, if p b (x,y)=0, then we have lim n p b ( x n , y n )=0.

Moreover, for each zX we have

1 s p b ( x , z ) p b ( x , x ) lim inf n p b ( x n , z ) lim sup n p b ( x n , z ) s p b ( x , z ) + s p b ( x , x ) .

In particular, if p b (x,x)=0, then we have

1 s p b (x,z) lim inf n p b ( x n ,z) lim sup n p b ( x n ,z)s p b (x,z).

One of the interesting generalizations of the Banach contraction principle was given by Kirk et al. [5] in 2003 by introducing the notion of cyclic representation.

Definition 1.14 [5]

Let A and B be nonempty subsets of a metric space (X,d) and T:ABAB. Then T is called a cyclic map if T(A)B and T(B)A.

The following interesting theorem for a cyclic map was given in [5].

Theorem 1.15 [5]

Let A and B be nonempty closed subsets of a complete metric space (X,d). Suppose that T:ABAB is a cyclic map such that

d(Tx,Ty)kd(x,y)

for all xA and yB, where k[0,1) is a constant. Then T has a unique fixed point u and uAB.

Berinde initiated in [6, 7] the concept of almost contractions and obtained several interesting fixed point theorems for Ćirić strong almost contractions. Babu et al. introduced in [8] the class of mappings which satisfy ‘condition (B)’. Moreover, they proved the existence of fixed points for such mappings on complete metric spaces. Finally, Ćirić et al. in [9], and Aghajani et al. in [10] introduced the concept of almost generalized contractive conditions (for two, resp. four mappings) and proved some important results in ordered metric spaces. Let us recall one of these definitions.

Definition 1.16 [9]

Let f and g be two self-mappings on a metric space (X,d). They are said to satisfy almost generalized contractive condition, if there exist a constant δ(0,1) and some L0 such that

d ( f x , g y ) δ max { d ( x , y ) , d ( x , f x ) , d ( y , g y ) , d ( x , g y ) + d ( y , f x ) 2 } + L min { d ( x , f x ) , d ( y , g y ) , d ( x , g y ) , d ( y , f x ) } ,

for all x,yX.

Definition 1.17 [11]

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.

Definition 1.18 [12]

Let (X,) be a partially ordered set and A and B be closed subsets of X with X=AB. Let f,g:XX be two mappings. The pair (f,g) is said to be (A,B)-weakly increasing if fxgfx, for all xA and gyfgy, for all yB.

In [13], Hussain et al. introduced the notion of ordered cyclic weakly (ψ,φ,L,A,B)-contractive pair of self-mappings as follows.

Definition 1.19 [13]

Let (X,,d) be an ordered b-metric space, let f,g:XX be two mappings, and let A and B be nonempty closed subsets of X. The pair (f,g) is called an ordered cyclic weakly (ψ,φ,L,A,B)-contraction if

  1. (1)

    X=AB is a cyclic representation of X w.r.t. the pair (f,g); that is, fAB and gBA;

  2. (2)

    there exist two altering distance functions ψ, φ and a constant L0, such that for arbitrary comparable elements x,yX with xA and yB, we have

    ψ ( s 2 d ( f x , g y ) ) ψ ( M s ( x , y ) ) φ ( M s ( x , y ) ) +Lψ ( N ( x , y ) ) ,

where

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

and

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

Also, in [13] the authors proved the following results.

Theorem 1.20 [13]

Let (X,,d) be a complete ordered b-metric space and A and B be closed subsets of X. Let f,g:XX be two (A,B)-weakly increasing mappings with respect to . Suppose that:

  1. (a)

    the pair (f,g) is an ordered cyclic weakly (ψ,φ,L,A,B)-contraction;

  2. (b)

    f or g is continuous.

Then f and g have a common fixed point uAB.

An ordered b-metric space (X,,d) is called regular if for any nondecreasing sequence { x n } in X such that x n xX, as n, one has x n x for all nN.

Theorem 1.21 [13]

Let the hypotheses of Theorem  1.20 be satisfied, except that condition (b) is replaced by the assumption

(b′) the space (X,,d) is regular.

Then f and g have a common fixed point in X.

In this paper, first we prove some fixed point results for α-admissible mappings in the context of partial b-metric spaces. Then we express some common fixed point results for cyclic generalized almost contractive mappings. Our results extend and generalize some recent results in [4] and [13]. In fact, they are cyclic variants of the results in [4].

2 Fixed point results via α-admissible mappings in partial b-metric spaces

Samet et al. [14] defined the notion of α-admissible mappings and proved the following result.

Definition 2.1 [14]

Let T be a self-mapping on X and α:X×X[0,) be a function. We say that T is an α-admissible mapping if

x,yX,α(x,y)1α(Tx,Ty)1.

Denote by Ψ the family of all nondecreasing functions ψ:[0,)[0,) such that n = 1 ψ n (t)< for all t>0, where ψ n is the n th iterate of ψ.

Theorem 2.2 [14]

Let (X,d) be a complete metric space and T be an α-admissible mapping. Assume that

α(x,y)d(Tx,Ty)ψ ( d ( x , y ) )
(2.1)

where ψ Ψ . Also, suppose that the following assertions hold:

  1. (i)

    there exists x 0 X such that α( x 0 ,T x 0 )1;

  2. (ii)

    either T is continuous or for any sequence { x n } in X with α( x n , x n + 1 )1 for all nN{0} such that x n x as n, we have α( x n ,x)1 for all nN{0}.

Then T has a fixed point.

We now recall the concept of (c)-comparison function which was introduced by Berinde [15].

Definition 2.3 (Berinde [15])

A function φ:[0,)[0,) is said to be a (c)-comparison function if

( c 1 ) φ is increasing,

( c 2 ) there exist k 0 N, a(0,1), and a convergent series of nonnegative terms k = 1 v k such that φ k + 1 (t)a φ k (t)+ v k , for k k 0 and any t[0,).

Later, Berinde [16] introduced the notion of (b)-comparison function as a generalization of a (c)-comparison function.

Definition 2.4 (Berinde [16])

Let s1 be a real number. A mapping φ:[0,)[0,) is called a (b)-comparison function if the following conditions are fulfilled:

  1. (1)

    φ is monotone increasing;

  2. (2)

    there exist k 0 N, a(0,1), and a convergent series of nonnegative terms k = 1 v k such that s k + 1 φ k + 1 (t)a s k φ k (t)+ v k , for k k 0 and any t[0,).

Let Ψ b be the class of (b)-comparison functions φ:[0,)[0,). It is clear that the notion of (b)-comparison function coincides with (c)-comparison function for s=1.

We now recall the following lemma, which will simplify the proofs.

Lemma 2.5 (Berinde [17])

If φ:[0,)[0,) is a (b)-comparison function, then we have the following.

  1. (1)

    the series k = 0 s k φ k (t) converges for any t R + ;

  2. (2)

    the function b s :[0,)[0,), defined by b s (t)= k = 0 s k φ k (t), t[0,), is increasing and continuous at 0.

Theorem 2.6 Let (X, p b ) be a p b -complete partial b-metric space, f be a continuous α-admissible mapping on X, there exists x 0 X such that α( x 0 ,f x 0 )1 and if any sequence { x n } in X p b -converges to a point x, where α( x n , x n + 1 )1 for all n, then we have α(x,x)1. Assume that

sα(x,y) p b (fx,fy)ψ ( M s ( x , y ) )
(2.2)

for all x,yX, where ψ Ψ b and

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

Then f has a fixed point.

Proof Let x 0 X be such that α( x 0 ,f x 0 )1. Define a sequence { x n } by x n = f n x 0 for all nN. Since f is an α-admissible mapping and α( x 0 , x 1 )=α( x 0 ,f x 0 )1, we deduce that α( x 1 , x 2 )=α(f x 0 ,f x 1 )1. Continuing this process, we get that α( x n , x n + 1 )1 for all nN{0}.

Now, we will finish the proof in the following steps.

First, we prove that

p b ( x n , x n + 1 )ψ ( p b ( x n 1 , x n ) ) ,
(2.3)

for each n=1,2,3, .

If x n = x n + 1 , for some nN, then x n =f x n . Thus, x n is a fixed point of f. Therefore, we assume that x n x n + 1 , for all nN.

Using condition (2.2) as α( x n 1 , x n )1 for all nN{0}, we obtain

s p b ( x n , x n + 1 )sα( x n 1 , x n ) p b (f x n 1 ,f x n )ψ ( M s ( x n 1 , x n ) ) .

Here,

M s ( x n 1 , x n ) = max { p b ( x n 1 , x n ) , p b ( x n 1 , f x n 1 ) , p b ( x n , f x n ) , 1 2 s [ p b ( x n 1 , f x n ) + p b ( x n , f x n 1 ) ] } = max { p b ( x n 1 , x n ) , p b ( x n 1 , x n ) , p b ( x n , x n + 1 ) , 1 2 s [ p b ( x n 1 , x n + 1 ) + p b ( x n , x n ) ] } max { p b ( x n 1 , x n ) , p b ( x n , x n + 1 ) } .

If p b ( x n , x n + 1 ) p b ( x n 1 , x n ), then

M s ( x n 1 , x n ) p b ( x n , x n + 1 ),

which yields

s p b ( x n , x n + 1 )ψ ( p b ( x n , x n + 1 ) ) < p b ( x n , x n + 1 ),

a contradiction.

Hence,

p b ( x n , x n + 1 )ψ ( p b ( x n 1 , x n ) ) .

So (2.3) holds.

By induction, we get

p b ( x n , x n + 1 ) ψ ( p b ( x n 1 , x n ) ) ψ 2 ( p b ( x n 2 , x n 1 ) ) ψ n ( p b ( x 0 , x 1 ) ) .
(2.4)

Then, by the triangular inequality and (2.4), we get

p b ( x n , x m ) s p b ( x n , x n + 1 ) + s 2 p b ( x n + 1 , x n + 2 ) + + s m n 1 p b ( x m 1 , x m ) k = n m 2 s k n + 1 ψ k ( p b ( x 0 , x 1 ) ) k = n s k ψ k ( p b ( x 0 , x 1 ) ) 0 ,

as n.

Since { x n } is a p b -Cauchy sequence in the p b -complete partial b-metric space X, from Lemma 1.9, { x n } is a b-Cauchy sequence in the b-metric space (X, d p b ). p b -Completeness of (X, p b ) shows that (X, d p b ) is also b-complete. Then there exists zX such that

lim n d p b ( x n ,z)=0.
(2.5)

Since lim m , n p b ( x n , x m )=0, from Lemma 1.9

lim n p b ( x n ,z)= lim m , n p b ( x n , x m )= p b (z,z)=0.
(2.6)

From the continuity of f we have

lim n p b ( x n + 1 ,fz)= p b (fz,fz)

and hence we get

p b (z,fz) lim n s p b (z, x n + 1 )+ lim n s p b ( x n + 1 ,fz)=s p b (fz,fz).

So, we get p b (z,fz)s p b (fz,fz). As α(z,z)1, we have

p b (z,fz)sα(z,z) p b (fz,fz)ψ ( max { p b ( z , z ) , p b ( z , f z ) , p b ( z , f z ) , p b ( z , f z ) + p b ( f z , z ) 2 s } ) .

Hence, p b (z,fz)ψ( p b (z,fz)). Thus, p b (z,fz)=0, that is, z=fz. □

In Theorem 2.6, we omit the continuity of the mapping f and we replace α( x n ,x)1 instead of α(x,x)1 and rearrange it as follows.

Theorem 2.7 Let (X, p b ) be a p b -complete partial b-metric space and f be an α-admissible mapping on X such that

sα(x,y) p b (fx,fy)ψ ( M s ( x , y ) )
(2.7)

for all x,yX, where ψ Ψ b . Assume that the following conditions hold:

  1. (i)

    there exists x 0 X such that α( x 0 ,f x 0 )1;

  2. (ii)

    if { x n } is a sequence in X such that α( x n , x n + 1 )1 for all n and x n x as n, then α( x n ,x)1 for all nN{0}.

Then f has a fixed point.

Proof Let x 0 X be such that α( x 0 ,f x 0 )1 and define a sequence { x n } in X by x n = f n x 0 =f x n 1 for all nN. Following the proof of Theorem 2.6, we have α( x n , x n + 1 )1 for all nN{0} and there exists zX such that x n z as n which p b (z,z)=0. Hence, from (ii) we deduce that α( x n ,z)1 for all nN{0}. Therefore, by (2.7), we obtain

s p b (fz, x n + 1 )sα( x n ,z) p b (fz,f x n )ψ ( M s ( z , x n ) ) .

Here,

M s ( z , x n ) = max { p b ( z , x n ) , p b ( z , f z ) , p b ( x n , f x n ) , 1 2 s [ p b ( z , f x n ) + p b ( x n , f z ) ] } = max { p b ( z , x n ) , p b ( z , f z ) , p b ( x n , x n + 1 ) , 1 2 s [ p b ( z , x n + 1 ) + p b ( x n , f z ) ] } .

Taking the upper limit as n in the above inequality from Lemma 1.13 we obtain

s [ 1 s p b ( f z , z ) ] s lim sup n p b (fz,f x n )ψ ( lim sup n M s ( z , x n ) ) ψ ( p b ( z , f z ) ) ,

which implies that z=fz. □

Definition 2.8 [18]

Let f:XX and α:X×XR. We say that f is a triangular α-admissible mapping if

(T1) α(x,y)1 implies α(fx,fy)1, x,yX,

(T2) { α ( x , z ) 1 α ( z , y ) 1 implies α(x,y)1, x,y,zX.

Example 2.9 [18]

Let X=R, fx= x 3 , and α(x,y)= e x y , then f is a triangular α-admissible mapping. Indeed, if α(x,y)= e x y 1, then xy which implies that fxfy, that is, α(fx,fy)= e f x f y 1. Also, if { α ( x , z ) 1 α ( z , y ) 1 , then { x z 0 z y 0 , that is, xy0 and therefore α(x,y)= e x y 1.

Example 2.10 [18]

Let X=R, fx= e x 7 , and α(x,y)= x y 5 +1. Hence, f is a triangular α-admissible mapping. Indeed, if α(x,y)= x y 5 +11 then xy which implies that fxfy, that is, α(fx,fy)1.

Moreover, if { α ( x , z ) 1 α ( z , y ) 1 , then xy0 and hence α(x,y)1.

Example 2.11 [18]

Let X=[0,), fx= x 4 +ln( x 2 +1), and

α(x,y)= x 3 1 + x 3 y 3 y 3 + 1 +1.

Then f is a triangular α-admissible mapping. In fact, if

α(x,y)= x 3 1 + x 3 y 3 y 3 + 1 +11,

then xy. Hence, fxfy, that is, α(fx,fy)1. Also,

α ( x , z ) + α ( z , y ) = x 3 1 + x 3 z 3 z 3 + 1 + 1 + z 3 1 + z 3 y 3 y 3 + 1 + 1 = x 3 1 + x 3 y 3 y 3 + 1 + 2 2 ( x 3 1 + x 3 y 3 y 3 + 1 + 1 ) = 2 α ( x , y ) .

Thus, α(x,z)+α(z,y)2α(x,y). Now, if { α ( x , z ) 1 α ( z , y ) 1 , then α(x,y)1.

Example 2.12 [18]

Let X=R, fx= x 3 + x 7 , and α(x,y)= x 5 y 5 +1. Then f is a triangular α-admissible mapping.

Lemma 2.13 [18]

Let f be a triangular α-admissible mapping. Assume that there exists x 0 X such that α( x 0 ,f x 0 )1. Define the sequence { x n } by x n = f n x 0 . Then

α( x m , x n )1for all m,nN with m<n.

A mapping ψ:[0,)[0,) is called a comparison function if it is increasing and ψ n (t)0, as n for any t[0,).

Lemma 2.14 (Berinde [15], Rus [19])

If ψ:[0,)[0,) is a comparison function, then:

  1. (1)

    each iterate ψ k of ψ, k1, is also a comparison function;

  2. (2)

    ψ is continuous at 0;

  3. (3)

    ψ(t)<t, for any t>0.

Denote by Ψ the family of all continuous comparison functions ψ:[0,)[0,).

In the sequel, ψΨ, α:X×X[0,) is a function and

M s (x,y)=max { p b ( x , y ) , p b ( x , f x ) , p b ( y , f y ) , 1 2 s [ p b ( x , f y ) + p b ( y , f x ) ] } .

Theorem 2.15 Let (X, p b ) be a p b -complete partial b-metric space, f be a continuous triangular α-admissible mapping on X, there exists x 0 X such that α( x 0 ,f x 0 )1 and if any sequence { x n } in X p b -converges to a point x, where α( x n , x n + 1 )1 for all n, then we have α(x,x)1. Assume that

sα(x,y) p b (fx,fy)ψ ( M s ( x , y ) )
(2.8)

for all x,yX. Then f has a fixed point.

Proof Let x 0 X be such that α( x 0 ,f x 0 )1. Define a sequence { x n } by x n = f n x 0 for all nN. Since f is an α-admissible mapping and α( x 0 , x 1 )=α( x 0 ,f x 0 )1, we deduce that α( x 1 , x 2 )=α(f x 0 ,f x 1 )1. Continuing this process, we get α( x n , x n + 1 )1 for all nN{0}.

Now, we will finish the proof in the following steps.

Step I. We will prove that

lim n p b ( x n , x n + 1 )=0.

First, we prove that

p b ( x n , x n + 1 )ψ ( p b ( x n 1 , x n ) ) ,
(2.9)

for each n=1,2,3, .

If x n = x n + 1 , for some nN, then x n =f x n . Thus, x n is a fixed point of f. Therefore, we assume that x n x n + 1 , for all nN.

Using condition (2.8) as α( x n 1 , x n )1 for all nN{0}, we obtain

s p b ( x n , x n + 1 )sα( x n 1 , x n ) p b (f x n 1 ,f x n )ψ ( M s ( x n 1 , x n ) ) .

Here,

M s ( x n 1 , x n ) = max { p b ( x n 1 , x n ) , p b ( x n 1 , f x n 1 ) , p b ( x n , f x n ) , 1 2 s [ p b ( x n 1 , f x n ) + p b ( x n , f x n 1 ) ] } = max { p b ( x n 1 , x n ) , p b ( x n 1 , x n ) , p b ( x n , x n + 1 ) , 1 2 s [ p b ( x n 1 , x n + 1 ) + p b ( x n , x n ) ] } max { p b ( x n 1 , x n ) , p b ( x n , x n + 1 ) } .

If p b ( x n , x n + 1 ) p b ( x n 1 , x n ), then

M s ( x n 1 , x n ) p b ( x n , x n + 1 ),

which yields

s p b ( x n , x n + 1 )ψ ( p b ( x n , x n + 1 ) ) < p b ( x n , x n + 1 ),

a contradiction.

Hence,

p b ( x n , x n + 1 )ψ ( p b ( x n 1 , x n ) ) .

So (2.9) holds.

By induction, we get

p b ( x n , x n + 1 )ψ ( p b ( x n 1 , x n ) ) ψ 2 ( p b ( x n 2 , x n 1 ) ) ψ n ( p b ( x 0 , x 1 ) ) .
(2.10)

As ψΨ, we conclude that

lim n p b ( x n , x n + 1 )=0.
(2.11)

So by ( p b 2 ) we get

lim n p b ( x n , x n )=0.
(2.12)

Step II. We will show that { x n } is a p b -Cauchy sequence in X. For this, we have to show that { x n } is a b-Cauchy sequence in (X, d p b ) (see Lemma 1.9). Suppose the contrary; that is, { x n } is not a b-Cauchy sequence. Then there exists 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 >iand d p b ( x m i , x n i )ε.
(2.13)

This means that

d p b ( x m i , x n i 1 )<ε.
(2.14)

From (2.13) and using the triangular inequality, we get

ε d p b ( x m i , x n i )s d p b ( x m i , x n i 1 )+s d p b ( x n i 1 , x n i ).

Using (2.11), (2.12), and from the definition of d p b and (2.14), and taking the upper limit as i, we get

ε s lim sup i d p b ( x m i , x n i 1 )ε.
(2.15)

Also,

ε lim inf i d p b ( x m i , x n i ) lim sup i d p b ( x m i , x n i )sε.
(2.16)

Further,

ε s lim sup i d p b ( x m i + 1 , x n i ) s 2 ε
(2.17)

and

lim sup i d p b ( x m i + 1 , x n i 1 )sε.
(2.18)

On the other hand, by the definition of d p b and (2.12)

lim sup i d p b ( x m i , x n i 1 )=2 lim sup i p b ( x m i , x n i 1 ).

Hence, by (2.15),

ε 2 s lim sup i p b ( x m i , x n i 1 ) ε 2 .
(2.19)

Similarly,

ε 2 lim inf i p b ( x m i , x n i ) lim sup i p b ( x m i , x n i ) s ε 2 ,
(2.20)
ε 2 s lim sup i p b ( x m i + 1 , x n i ) s 2 ε 2 ,
(2.21)

and

lim sup i p b ( x m i + 1 , x n i 1 ) s ε 2 .
(2.22)

From (2.8) and Lemma 2.13 as α( x m i , x n i 1 )1, we have

s p b ( x m i + 1 , x n i )sα( x m i , x n i 1 ) p b (f x m i ,f x n i 1 )ψ ( M s ( x m i , x n i 1 ) ) ,
(2.23)

where

M s ( x m i , x n i 1 ) = max { p b ( x m i , x n i 1 ) , p b ( x m i , f x m i ) , p b ( x n i 1 , f x n i 1 ) , p b ( x m i , f x n i 1 ) + p b ( f x m i , x n i 1 ) 2 s } = max { p b ( x m i , x n i 1 ) , p b ( x m i , x m i + 1 ) , p b ( x n i 1 , x n i ) , p b ( x m i , x n i ) + p b ( x m i + 1 , x n i 1 ) 2 s } .
(2.24)

Taking the upper limit as i in (2.24) and using (2.11), (2.19), (2.20), and (2.22), we get

lim sup i M s ( x m i , x n i 1 ) = max { lim sup i p b ( x m i , x n i 1 ) , 0 , 0 , lim sup i p b ( x m i , x n i ) + lim sup i p b ( x m i + 1 , x n i 1 ) 2 s } max { ε 2 , ε s + ε s 2 2 s } = ε 2 .
(2.25)

Now, taking the upper limit as i in (2.23) and using (2.21) and (2.25), we have

s ε 2 s s lim sup i p b ( x m i + 1 , x n i )ψ ( lim sup i M s ( x m i , x n i 1 ) ) < ε 2 ,

a contradiction.

Step III. There exists z such that fz=z.

Since { x n } is a p b -Cauchy sequence in the p b -complete partial b-metric space X, from Lemma 1.9, { x n } is a b-Cauchy sequence in the b-metric space (X, d p b ). p b -Completeness of (X, p b ) shows that (X, d p b ) is also b-complete. Then there exists zX such that

lim n d p b ( x n ,z)=0.
(2.26)

Since lim m , n d p b ( x n , x m )=0, from the definition of d p b and (2.12), we get

lim m , n p b ( x n , x m )=0.

Again, from Lemma 1.9,

lim n p b (z, x n )= lim m , n p b ( x n , x m )= p b (z,z)=0.
(2.27)

From the continuity of f we have

lim n p b ( x n + 1 ,fz)= p b (fz,fz)

and hence we get

p b (z,fz) lim n s p b (z, x n + 1 )+ lim n s p b ( x n + 1 ,fz)=s p b (fz,fz).

So, we get p b (z,fz)s p b (fz,fz). As α(z,z)1, we have

p b ( z , f z ) s α ( z , z ) p b ( f z , f z ) ψ ( max { p b ( z , z ) , p b ( z , f z ) , p b ( z , f z ) , p b ( z , f z ) + p b ( f z , z ) 2 s } ) .

Hence, p b (z,fz)ψ( p b (z,fz)). Thus, p b (z,fz)=0, that is, z=fz. □

If in Theorem 2.15 we take α(x,y)=1 then we deduce the following corollary.

Corollary 2.16 Let (X, p b ) be a p b -complete partial b-metric space and f be a continuous mapping on X. Assume that

s p b (fx,fy)ψ ( M s ( x , y ) )
(2.28)

for all x,yX. Then f has a fixed point.

In Theorem 2.15, we omit the continuity of the mapping f and we replace α( x n ,x)1 instead of α(x,x)1 and rearrange it as follows.

Theorem 2.17 Let (X, p b ) be a p b -complete partial b-metric space and f be a triangular α-admissible mapping on X such that

sα(x,y) p b (fx,fy)ψ ( M s ( x , y ) )
(2.29)

for all x,yX, where ψΨ. Assume that the following conditions hold:

  1. (i)

    there exists x 0 X such that α( x 0 ,f x 0 )1;

  2. (ii)

    if { x n } is a sequence in X such that α( x n , x n + 1 )1 for all n and x n x as n, then α( x n ,x)1 for all nN{0}.

Then f has a fixed point.

Proof Let x 0 X be such that α( x 0 ,f x 0 )1 and define a sequence { x n } in X by x n = f n x 0 =f x n 1 for all nN. Following the proof of Theorem 2.15, we have α( x n , x n + 1 )1 for all nN{0} and there exists zX such that x n z as n which p b (z,z)=0. Hence, from (ii) we deduce that α( x n ,z)1 for all nN{0}. Therefore, by (2.29), we obtain

s p b (fz, x n + 1 )sα( x n ,z) p b (fz,f x n )ψ ( M s ( z , x n ) ) .

Here,

M s ( z , x n ) = max { p b ( z , x n ) , p b ( z , f z ) , p b ( x n , f x n ) , 1 2 s [ p b ( z , f x n ) + p b ( x n , f z ) ] } = max { p b ( z , x n ) , p b ( z , f z ) , p b ( x n , x n + 1 ) , 1 2 s [ p b ( z , x n + 1 ) + p b ( x n , f z ) ] } .

Taking the upper limit as n in the above inequality from Lemma 1.13 we obtain

s [ 1 s p b ( f z , z ) ] s lim sup n p b (fz,f x n )ψ ( lim sup n M s ( z , x n ) ) ψ ( p b ( z , f z ) ) ,

which implies that z=fz. □

Example 2.18 Let X=[0,1] and p b (x,y)=|xy | 2 be a p b -metric on X. Define f:XX by fx=ln( x 4 +1) and α:X×X[0,) by

α(x,y)={ 1 , if  ( x , y ) [ 0 , 1 4 ] × [ 0 , 1 4 ] , 0 , otherwise ,

and ψ(t)= t 8 for all t[0,). Now, we prove that all the hypotheses of Theorem 2.17 are satisfied and hence f has a fixed point.

First, we see that (X, p b ) is a p b -complete partial b-metric space. Let x,yX. If α(x,y)1, then x,y[0, 1 4 ]. On the other hand, for all x[0,1], we have fx x 4 1 4 and hence α(fx,fy)=1. This implies that f is a triangular α-admissible mapping on X. Obviously, α(0,f0)=1.

Now, if { x n } is a sequence in X such that α( x n , x n + 1 )=1 for all nN{0} and x n x as n, it is easy to see that α( x n ,x)=1.

Using the Mean Value Theorem for the function fx=ln( x 4 +1) for any x,yX, we have

s α ( x , y ) p b ( f x , f y ) s p b ( f x , f y ) = 2 | f x f y | 2 = 2 | ln ( x 4 + 1 ) ln ( y 4 + 1 ) | 2 1 8 | x y | 2 = ψ ( p b ( x , y ) ) ψ ( M s ( x , y ) ) .

Thus, all the conditions of Theorem 2.17 are satisfied and therefore f has a fixed point (z=0).

3 Common fixed points of generalized almost cyclic weakly (ψ,φ,L,A,B)-contractive mappings

In this section, we consider the notion of ordered cyclic weakly (ψ,φ,L,A,B)-contractions in the setup of ordered partial b-metric spaces and then obtain some common fixed point theorems for these cyclic contractions in the setup of complete ordered partial b-metric spaces. Our results extend some fixed point theorems from the framework of ordered metric spaces and ordered b-metric spaces, in particular Theorems 1.20 and 1.21.

We shall call an ordered partial b-metric space (X,, p b ) regular if for any nondecreasing sequence { x n } in X such that x n xX, as n, one has x n x, for all nN.

Definition 3.1 Let (X,, p b ) be an ordered partial b-metric space, let f,g:XX be two mappings, and let A and B be nonempty closed subsets of X. The pair (f,g) is called an ordered cyclic almost generalized weakly (ψ,φ,L,A,B)-contraction if

  1. (1)

    X=AB is a cyclic representation of X w.r.t. the pair (f,g); that is, fAB and gBA;

  2. (2)

    there exist two altering distance functions ψ, φ and a constant L0, such that for arbitrary comparable elements x,yX with xA and yB, we have

    ψ ( s 2 p b ( f x , g y ) ) ψ ( M s ( x , y ) ) φ ( M s ( x , y ) ) +Lψ ( N ( x , y ) ) ,
    (3.1)

where

M s (x,y)=max { p b ( x , y ) , p b ( x , f x ) , p b ( y , g y ) , p b ( x , g y ) + p b ( y , f x ) 2 s }
(3.2)

and

N(x,y)=min { d p b ( x , f x ) , d p b ( x , g y ) , d p b ( y , f x ) , d p b ( y , g y ) } .
(3.3)

Theorem 3.2 Let (X,, p b ) be a p b -complete ordered partial b-metric space and A and B be two nonempty closed subsets of X. Let f,g:XX be two (A,B)-weakly increasing mappings with respect to . Suppose that the pair (f,g) is an ordered cyclic almost generalized weakly (ψ,φ,L,A,B)-contraction. Then f and g have a common fixed point zAB.

Proof First, note that uAB is a fixed point of f if and only if u is a fixed point of g. Indeed, suppose that u is a fixed point of f. As uu and uAB, by (3.1), we have

ψ ( s 2 p b ( u , g u ) ) = ψ ( s 2 p b ( f u , g u ) ) ψ ( max { p b ( u , u ) , p b ( u , f u ) , p b ( u , g u ) , 1 2 s ( p b ( u , g u ) + p b ( u , f u ) ) } ) φ ( max { p b ( u , u ) , p b ( u , f u ) , p b ( u , g u ) , 1 2 s ( p b ( u , g u ) + p b ( u , f u ) ) } ) + L min { d p b ( u , g u ) , d p b ( u , f u ) } = ψ ( p b ( u , g u ) ) φ ( p b ( u , g u ) ) ψ ( s 2 p b ( u , g u ) ) φ ( p b ( u , g u ) ) .

It follows that φ( p b (u,gu))=0. Therefore, p b (u,gu)=0 and hence gu=u. Similarly, we can show that if u is a fixed point of g, then u is a fixed point of f.

Let x 0 A and let x 1 =f x 0 . Since fAB, we have x 1 B. Also, let x 2 =g x 1 . Since gBA, we have x 2 A. Continuing this process, we can construct a sequence { x n } in X such that x 2 n + 1 =f x 2 n , x 2 n + 2 =g x 2 n + 1 , x 2 n A and x 2 n + 1 B. Since f and g are (A,B)-weakly increasing, we have

x 1 =f x 0 gf x 0 = x 2 =g x 1 fg x 1 = x 3 x 2 n + 1 =f x 2 n gf x 2 n = x 2 n + 2 .

If x 2 n = x 2 n + 1 , for some nN, then x 2 n =f x 2 n . Thus x 2 n is a fixed point of f. By the first part of the proof, we conclude that x 2 n is also a fixed point of g. Similarly, if x 2 n + 1 = x 2 n + 2 , for some nN, then x 2 n + 1 =g x 2 n + 1 . Thus, x 2 n + 1 is a fixed point of g. By the first part of the proof, we conclude that x 2 n + 1 is also a fixed point of f. Therefore, we assume that x n x n + 1 , for all nN. Now, we complete the proof in the following steps.

Step 1. We will prove that

lim n p b ( x n , x n + 1 )=0.

As x 2 n and x 2 n + 1 are comparable and x 2 n A and x 2 n + 1 B, by (3.1), we have

ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) ψ ( s 2 p b ( x 2 n + 1 , x 2 n + 2 ) ) = ψ ( s 2 p b ( f x 2 n , g x 2 n + 1 ) ) ψ ( M s ( x 2 n , x 2 n + 1 ) ) φ ( M s ( x 2 n , x 2 n + 1 ) ) + L ψ ( N ( x 2 n , x 2 n + 1 ) ) ,

where

M s ( x 2 n , x 2 n + 1 ) = max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n , f x 2 n ) , p b ( x 2 n + 1 , g x 2 n + 1 ) , p b ( f x 2 n , x 2 n + 1 ) + p b ( x 2 n , g x 2 n + 1 ) 2 s } = max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) , p b ( x 2 n + 1 , x 2 n + 1 ) + p b ( x 2 n , x 2 n + 2 ) 2 s } max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) , s [ p b ( x 2 n , x 2 n + 1 ) + p b ( x 2 n + 1 , x 2 n + 2 ) ] 2 s } = max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) }

and

N ( x 2 n , x 2 n + 1 ) = min { d p b ( x 2 n , f x 2 n ) , d p b ( x 2 n , g x 2 n + 1 ) , d p b ( x 2 n + 1 , f x 2 n ) , d p b ( x 2 n + 1 , g x 2 n + 1 ) } = min { d p b ( x 2 n , x 2 n + 1 ) , d p b ( x 2 n , x 2 n + 2 ) , d p b ( x 2 n + 1 , x 2 n + 1 ) , d p b ( x 2 n + 1 , x 2 n + 2 ) } = 0 .

Hence, we have

ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) ψ ( max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) } ) φ ( max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) } ) .
(3.4)

If

max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) } = p b ( x 2 n + 1 , x 2 n + 2 ),

then (3.4) becomes

ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) φ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) < ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) ,

which gives a contradiction. So,

max { p b ( x 2 n , x 2 n + 1 ) , p b ( x 2 n + 1 , x 2 n + 2 ) } = p b ( x 2 n , x 2 n + 1 )

and hence (3.4) becomes

ψ ( p b ( x 2 n + 1 , x 2 n + 2 ) ) ψ ( p b ( x 2 n , x 2 n + 1 ) ) φ ( p b ( x 2 n , x 2 n + 1 ) ) < ψ ( p b ( x 2 n , x 2 n + 1 ) ) .
(3.5)

Similarly, we can show that

ψ ( p b ( x 2 n + 1 , x 2 n ) ) <ψ ( p b ( x 2 n , x 2 n 1 ) ) .
(3.6)

By (3.5) and (3.6), we see that {d( x n , x n + 1 ):nN} is a nonincreasing sequence of positive numbers. Hence, there is r0 such that

lim n p b ( x n , x n + 1 )=r.

Letting n in (3.5), we get

ψ(r)ψ(r)φ(r),

which implies that φ(r)=0 and hence r=0. So, we have

lim n p b ( x n , x n + 1 )=0.
(3.7)

Step 2. We will prove that { x n } is a p b -Cauchy sequence. Because of (3.7), it is sufficient to show that { x 2 n } is a p b -Cauchy sequence. By Lemma 1.9, we should show that { x 2 n } is b-Cauchy in (X, d p b ). Suppose the contrary, i.e., that { x 2 n } is not a b-Cauchy sequence in (X, d p b ). Then there exists ε>0 for which we can find two subsequences { x 2 m i } and { x 2 n i } of { x 2 n } such that n i is the smallest index for which

n i > m i >iand d p b ( x 2 m i , x 2 n i )ε.
(3.8)

This means that

d p b ( x 2 m i , x 2 n i 2 )<ε.
(3.9)

From (3.8) and using the triangular inequality, we get

ε d p b ( x 2 m i , x 2 n i )s d p b ( x 2 m i , x 2 m i + 1 )+s d p b ( x 2 m i + 1 , x 2 n i ).

Using (3.7) and from the definition of d p b and taking the upper limit as i, we get

ε s lim sup i d p b ( x 2 m i + 1 , x 2 n i ).
(3.10)

On the other hand, we have

d p b ( x 2 m i , x 2 n i 1 )s d p b ( x 2 m i , x 2 n i 2 )+s d p b ( x 2 n i 2 , x 2 n i 1 ).

Using (3.7), (3.9), and taking the upper limit as i, we get

lim sup i d p b ( x 2 m i , x 2 n i 1 )εs.
(3.11)

Again, using the triangular inequality, we have

d p b ( x 2 m i , x 2 n i ) s d p b ( x 2 m i , x 2 n i 2 ) + s d p b ( x 2 n i 2 , x 2 n i ) s d p b ( x 2 m i , x 2 n i 2 ) + s 2 d p b ( x 2 n i 2 , x 2 n i 1 ) + s 2 d p b ( x 2 n i 1 , x 2 n i )

and

d p b ( x 2 m i + 1 , x 2 n i 1 )s d p b ( x 2 m i + 1 , x 2 m i )+s d p b ( x 2 m i , x 2 n i 1 ).

Taking the upper limit as i in the above inequalities, and using (3.7), (3.9), and (3.11) we get

lim sup i d p b ( x 2 m i , x 2 n i )εs
(3.12)

and

lim sup i d p b ( x 2 m i + 1 , x 2 n i 1 )ε s 2 .
(3.13)

From the definition of d p b and (3.7), (3.10), (3.11), (3.12), and (3.13) we have the following relations:

ε 2 s lim inf i p b ( x 2 m i + 1 , x 2 n i ),
(3.14)
lim sup i p b ( x 2 m i , x 2 n i 1 ) s ε 2 ,
(3.15)
lim sup i p b ( x 2 m i , x 2 n i ) s ε 2 ,
(3.16)
lim sup i p b ( x 2 m i + 1 , x 2 n i 1 ) s 2 ε 2 .
(3.17)

Since x 2 m i A and x 2 n i 1 B are comparable, using (3.1) we have

ψ ( s 2 p b ( x 2 m i + 1 , x 2 n i ) ) = ψ ( s 2 p b ( f x 2 m i , g x 2 n i 1 ) ) ψ ( M s ( x 2 m i , x 2 n i 1 ) ) φ ( M s ( x 2 m i , x 2 n i 1 ) ) + L ψ ( N ( x 2 m i , x 2 n i 1 ) ) ,
(3.18)

where

M s ( x 2 m i , x 2 n i 1 ) = max { p b ( x 2 m i , x 2 n i 1 ) , p b ( x 2 m i , x 2 m i + 1 ) , p b ( x 2 n i 1 , x 2 n i ) , p b ( x 2 m i , x 2 n i ) + p b ( x 2 m i + 1 , x 2 n i 1 ) 2 s }
(3.19)

and

N ( x 2 m i , x 2 n i 1 ) = min { d p b ( x 2 m i , f x 2 m i ) , d p b ( x 2 m i , g x 2 n i 1 ) , d p b ( x 2 n i 1 , f x 2 m i ) , d p b ( x 2 n i 1 , g x 2 n i 1 ) } = min { d p b ( x 2 m i , x 2 m i + 1 ) , d p b ( x 2 m i , x 2 n i ) , d p b ( x 2 n i 1 , x 2 m i + 1 ) , d p b ( x 2 n i 1 , x 2 n i ) } .
(3.20)

Taking the upper limit in (3.19) and (3.20), and using (3.7) and (3.14)-(3.17), we get

lim sup i M s ( x 2 m i , x 2 n i 1 ) = max { lim sup i p b ( x 2 m i , x 2 n i 1 ) , 0 , 0 , lim sup i p b ( x 2 m i , x 2 n i ) + lim sup i p b ( x 2 m i + 1 , x 2 n i 1 ) 2 s } max { s ε 2 , ε s + ε s 2 2 2 s } = s ε 2
(3.21)

and

lim sup i N( x 2 m i , x 2 n i 1 )=0.
(3.22)

Now, taking the upper limit as i in (3.18) and using (3.14), (3.21), and (3.22), we have

ψ ( s ε 2 ) = ψ ( s 2 ε 2 s ) ψ ( s 2 lim sup i p b ( x 2 m i + 1 , x 2 n i ) ) ψ ( lim sup i M s ( x 2 m i , x 2 n i 1 ) ) φ ( lim inf i M s ( x 2 m i , x 2 n i 1 ) ) ψ ( s ε 2 ) φ ( lim inf i M s ( x 2 m i , x 2 n i 1 ) ) ,

which implies that φ( lim inf i M s ( x 2 m i , x 2 n i 1 ))=0. By (3.19), it follows that

lim inf i p b ( x 2 m i , x 2 n i )=0,

which is in contradiction with (3.8). Thus, we have proved that { x n } is a b-Cauchy sequence in the metric space (X, d p b ). Since (X, p b ) is p b -complete, from Lemma 1.9, (X, d p b ) is a b-complete b-metric space. Therefore, the sequence { x n } converges to some zX, that is, lim n d p b ( x n ,z)=0. Since lim m , n d p b ( x n , x m )=0, from the definition of d p b and (3.7), we get

lim m , n p b ( x n , x m )=0.

Again, from Lemma 1.9,

lim n p b (z, x n )= lim m , n p b ( x n , x m )= p b (z,z)=0.

Step 3. In the above steps, we constructed an increasing sequence { x n } in X such that x n z, for some zX. As A and B are closed subsets of X, we have zAB. Using the regularity assumption on X, we have x n z, for all nN. Now, we show that fz=gz=z. By (3.1), we have

ψ ( s 2 p b ( x 2 n + 1 , g z ) ) = ψ ( s 2 p b ( f x 2 n , g z ) ) ψ ( M s ( x 2 n , z ) ) φ ( M s ( x 2 n , z ) ) + L ψ ( N ( x 2 n , z ) ) ,
(3.23)

where

M s ( x 2 n , z ) = max { p b ( x 2 n , z ) , p b ( x 2 n , f x 2 n ) , p b ( z , g z ) , p b ( x 2 n , g z ) + p b ( f x 2 n , z ) 2 s } = max { p b ( x 2 n , z ) , p b ( x 2 n , x 2 n + 1 ) , p b ( z , g z ) , p b ( x 2 n , g z ) + p b ( x 2 n + 1 , z ) 2 s }
(3.24)

and

N ( x 2 n , z ) = min { d p b ( x 2 n , f x 2 n ) , d p b ( z , g z ) , d p b ( z , f x 2 n ) , d p b ( x 2 n , g z ) } = min { d p b ( x 2 n , x 2 n + 1 ) , d p b ( z , g z ) , d p b ( z , x 2 n + 1 ) , d p b ( x 2 n , g z ) } .
(3.25)

Letting n in (3.24) and (3.25), and using Lemma 1.13, we get

lim sup i M s ( x 2 n ,z)max { p b ( z , g z ) , s p b ( z , g z ) 2 s } = p b (z,gz),
(3.26)

and N( x 2 n ,z)0. Now, taking the upper limit as n in (3.23), and using Lemma 1.13 and (3.26) we get

ψ ( s p b ( z , g z ) ) = ψ ( s 2 1 s p b ( z , g z ) ) ψ ( s 2 lim sup n p b ( x 2 n + 1 , g z ) ) ψ ( lim sup n M s ( x 2 n , z ) ) φ ( lim inf n M s ( x 2 n , z ) ) ψ ( s p b ( z , g z ) ) φ ( lim inf n M s ( x 2 n , z ) ) .

It follows that φ( lim inf n M s ( x 2 n ,z))=0, and hence, by (3.24), that p b (z,gz)=0. Thus, z is a fixed point of g. On the other hand, from the first part of the proof, fz=z. Hence, z is a common fixed point of f and g. □

Theorem 3.3 Let (X,, p b ) be a p b -complete ordered partial b-metric space and A and B be nonempty closed subsets of X. Let f,g:XX be two (A,B)-weakly increasing mappings with respect to . Suppose that

ψ ( s 2 p b ( f x , g y ) ) ψ ( M s ( x , y ) ) φ ( M s ( x , y ) ) .
(3.27)

Also, let f and g be continuous. Then f and g have a common fixed point zAB.

Proof Repeating the proof of Theorem 3.2, we construct an increasing sequence { x n } in X such that x n z, for some zX. As A and B are closed subsets of X, we have zAB. Now, we show that fz=gz=z.

Using the triangular inequality, we get

p b (z,fz)s p b (z,f x 2 n )+s p b (f x 2 n ,fz)

and

p b (z,gz)s p b (z,g x 2 n + 1 )+s p b (g x 2 n + 1 ,gz).

Letting n and using continuity of f and g, we get

p b ( z , f z ) s lim n p b ( z , f x 2 n ) + s lim n p b ( f x 2 n , f z ) = s p b ( f z , f z ) , p b ( z , g z ) s lim n p b ( z , g x 2 n + 1 ) + s lim n p b ( g x 2 n + 1 , g z ) = s p b ( g z , g z ) .

Therefore,

max { p b ( z , f z ) , p b ( z , g z ) } max { s p b ( f z , f z ) , s p b ( g z , g z ) } s 2 p b (gz,fz).
(3.28)

From (3.27) as zAB, we have

ψ ( s 2 p b ( f z , g z ) ) ψ ( M s ( z , z ) ) φ ( M s ( z , z ) ) ,
(3.29)

where

M s ( z , z ) = max { p b ( z , z ) , p b ( z , f z ) , p b ( z , g z ) , p b ( z , g z ) + p b ( z , f z ) 2 s } = max { p b ( z , f z ) , p b ( z , g z ) } .

As ψ is nondecreasing, we have s 2 p b (fz,gz)max{ p b (z,fz), p b (z,gz)}. Hence, by (3.28) we obtain s 2 p b (fz,gz)=max{ p b (z,fz), p b (z,gz)}. But then, using (3.29), we get φ( M s (z,z))=0. Thus, we have fz=gz=z and z is a common fixed point of f and g. □

As consequences, we have the following results.

By putting A=B=X in Theorems 3.2 and 3.3 and L=0 in Theorem 3.2, we obtain the main results (Theorems 3 and 4) of Mustafa et al. [4].

Taking φ=(1δ)ψ, 0<δ<1 in Theorem 3.2, we get the following.

Corollary 3.4 Let (X,, p b ) be a p b -complete ordered partial b-metric space and A and B be closed subsets of X. Let f,g:XX be two (A,B)-weakly increasing mappings with respect to . Suppose that:

  1. (a)

    X=AB is a cyclic representation of X w.r.t. the pair (f,g);

  2. (b)

    there exist 0<δ<1, L0, and an altering distance function ψ such that for any comparable elements x,yX with xA and yB, we have

    ψ ( s 2 p b ( f x , g y ) ) δψ ( M s ( x , y ) ) +Lψ ( N ( x , y ) ) ,
    (3.30)

where M s (x,y) and N(x,y) are given by (3.2) and (3.3), respectively;

  1. (c)

    f and g are continuous, or

(c′) the space (X,, p b ) is regular.

Then f and g have a common fixed point zAB.

Taking s=1 and L=0 in Corollary 3.4, we obtain the partial version of Theorems 2.1 and 2.2 of Shatanawi and Postolache [12].

In Definitions 1.18 and 3.1 and Theorems 3.2 and 3.3, if we take f=g, then we have the following definitions and results.

Definition 3.5 Let (X,) be a partially ordered set and A and B be closed subsets of X with X=AB. The mapping f:XX is said to be (A,B)-weakly increasing if fx f 2 x, for all xA and fy f 2 y, for all yB.

Definition 3.6 Let (X,, p b ) be an ordered partial b-metric space, let f:XX be a mapping, and let A and B be nonempty closed subsets of X. The mapping f is called an ordered cyclic almost generalized weakly (ψ,φ,L,A,B)-contraction if

  1. (1)

    X=AB is a cyclic representation of X w.r.t. f; that is, fAB and fBA;

  2. (2)

    there exist two altering distance functions ψ, φ and a constant L0, such that for arbitrary comparable elements x,yX with xA and yB, we have

    ψ ( s 2 p b ( f x , f y ) ) ψ ( M s ( x , y ) ) φ ( M s ( x , y ) ) +Lψ ( N ( x , y ) ) ,

where

M s (x,y)=max { p b ( x , y ) , p b ( x , f x ) , p b ( y , f y ) , p b ( x , f y ) + p b ( y , f x ) 2 s }

and

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

Corollary 3.7 Let (X,, p b ) be a p b -complete ordered partial b-metric space and A and B be two nonempty closed subsets of X. Let f:XX be a (A,B)-weakly increasing mapping with respect to . Suppose that the mapping f is an ordered cyclic almost generalized weakly (ψ,φ,L,A,B)-contraction. Then f has a fixed point zAB.

Corollary 3.8 Let (X,, p b ) be a p b -complete ordered partial b-metric space and A and B be nonempty closed subsets of X. Let f:XX be a (A,B)-weakly increasing mapping with respect to . Suppose that

ψ ( s 2 p b ( f x , f y ) ) ψ ( M s ( x , y ) ) φ ( M s ( x , y ) ) .

Also, let f be continuous. Then f has a fixed point zAB.

We illustrate our results with the following example.

Example 3.9 Consider the partial b-metric space X=[0,6] by p b (x,y)= [ max { x , y } ] 2 . Define an order on X by

xyx=y ( x , y [ 0 , 1 ] x y ) .

Obviously, (X,, p b ) is a p b -complete ordered p b -metric space. Indeed, if we have lim n , m p b ( x n , x m )=u, for some u[0,), then we have

lim m , n ( max { x n , x m } ) 2 = u max { ( lim n x n ) 2 , ( lim m x m ) 2 } = u ( lim n x n ) 2 = ( lim m x m ) 2 = u .

So, we have lim n x n = u , which convergence holds in the case of the usual metric in X. Now, it is easy to see that lim n , m p b ( x n , x m )= lim n p b ( x n , u )= p b ( u , u )=u.

Let f:XX be given by

fx={ x 2 3 ( 1 + x ) , x [ 0 , 1 ] , x 6 , x > 1 ,

ψ(t)=t and φ(t)= 8 9 t for all t[0,). Also, let A=[0,1] and B=[0,6]. In order to check the conditions of Corollary 3.8, take x,yX such that xy and consider the following two possible cases.

1 x1. Then obviously also y1 and xy. It is easy to check that

2 2 p b ( f x , f y ) = 4 [ max { x 2 3 ( 1 + x ) , y 2 3 ( 1 + y ) } ] 2 = 4 [ x 2 3 ( 1 + x ) ] 2 = 4 [ x 3 ( 1 + x ) x ] 2 4 [ x 6 ] 2 = 1 9 p b ( x , y ) M s ( x , y ) φ ( M s ( x , y ) ) .

2 x>1. Then x=y>1 and

2 2 p b ( f x , f y ) = 4 [ max { x 6 , y 6 } ] 2 = 4 [ y 6 ] 2 = 1 9 p b ( x , y ) p b ( x , y ) M s ( x , y ) φ ( M s ( x , y ) ) .

Hence, all the conditions of Corollary 3.8 are satisfied and f has a fixed point (which is z=0).

References

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

    MathSciNet  MATH  Google Scholar 

  2. Matthews SG: Partial metric topology. Annals of the New York Academy of Sciences 728. General Topology and Its Applications Proc. 8th Summer Conf., Queen’s College 1992 1994, 183–197.

    Google Scholar 

  3. Shukla S: Partial b -metric spaces and fixed point theorems. Mediterr. J. Math. 2014. 10.1007/s00009-013-0327-4

    Google Scholar 

  4. Mustafa Z, Roshan JR, Parvaneh V, Kadelburg Z: Some common fixed point results in ordered partial b -metric spaces. J. Inequal. Appl. 2013., 2013: Article ID 562

    Google Scholar 

  5. Kirk WA, Srinivasan PS, Veeramani P: Fixed points for mappings satisfying cyclical contractive conditions. Fixed Point Theory 2003,4(1):79–89.

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  8. 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  MATH  Google Scholar 

  9. 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  MATH  Google Scholar 

  10. Aghajani A, Radenović S, Roshan JR:Common fixed point results for four mappings satisfying almost generalized (S,T)-contractive condition in partially ordered metric spaces. Appl. Math. Comput. 2012, 218: 5665–5670. 10.1016/j.amc.2011.11.061

    Article  MathSciNet  MATH  Google Scholar 

  11. 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  MATH  Google Scholar 

  12. Shatanawi W, Postolache M: Common fixed point results of mappings for nonlinear contraction of cyclic form in ordered metric spaces. Fixed Point Theory Appl. 2013., 2013: Article ID 60 10.1186/1687-1812-2013-60

    Google Scholar 

  13. 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 

  14. Samet B, Vetro C, Vetro P: Fixed point theorem for α - ψ -contractive type mappings. Nonlinear Anal. 2012, 75: 2154–2165. 10.1016/j.na.2011.10.014

    Article  MathSciNet  MATH  Google Scholar 

  15. Berinde V: Contracţii generalizate şi aplicaţii. Editura Club Press 22, Baia Mare; 1997.

    Google Scholar 

  16. Berinde V: Sequences of operators and fixed points in quasimetric spaces. Stud. Univ. Babeş-Bolyai, Math. 1996,16(4):23–27.

    MathSciNet  MATH  Google Scholar 

  17. Berinde V: Generalized contractions in quasimetric spaces. Preprint 3. Seminar on Fixed Point Theory 1993, 3–9.

    Google Scholar 

  18. Karapinar E, Kumam P, Salimi P: On α - ψ -Meir-Keeler contractive mappings. Fixed Point Theory Appl. 2013., 2013: Article ID 94

    Google Scholar 

  19. Rus IA: Generalized Contractions and Applications. Cluj University Press, Cluj-Napoca; 2001.

    MATH  Google Scholar 

Download references

Acknowledgements

This article was funded by the Deanship of Scientific Research (DSR), King Abdulaziz University. The authors, therefore, acknowledge with thanks DSR for technical and financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Abdul Latif.

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 4.0 International License (https://creativecommons.org/licenses/by/4.0), which permits use, duplication, adaptation, distribution, and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Latif, A., Roshan, J.R., Parvaneh, V. et al. Fixed point results via α-admissible mappings and cyclic contractive mappings in partial b-metric spaces. J Inequal Appl 2014, 345 (2014). https://doi.org/10.1186/1029-242X-2014-345

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1029-242X-2014-345

Keywords