Introduction and preliminaries

The study of fixed points for multi-valued maps using a Hausdorff metric was initiated by Nadler [1] who proved the following:

Theorem 1.

Let (X,d) be a complete metric space and T:XC B(X) be a mapping satisfying H(T x,T y)≤k d(x,y), where k∈ [ 0,1) then there exists xX such that xT x.

Later, an interesting and rich fixed-point theory was developed and extended Theorem 1 using weak and generalized contraction mappings (see [27]). The theory of multi-valued maps has many applications in control theory, convex optimization, differential equations, and economics (see [8]). On the other hand, the basic notion of a partial metric space was introduced by Mathews [9] as a part of the study of denotational semantics of data flow networks. He presented a modified version of the Banach contraction principle, which is more suitable in this context (see also [10, 11]). In fact, the partial metric spaces constitute a suitable framework to model several distinguished examples of the theory of computation and also to model metric spaces via the domain theory (see [1231]). In this direction, Aydi et al. [32] introduced the concept of a partial Hausdorff metric and extended Nadler’s fixed-point theorem in the setting of partial metric spaces.

Consistent with [9, 32, 33], the following definitions and results will be needed in the sequel:

Definition 1.

([9]). A partial metric on a nonempty set X is a function p:X×X R + such that for all x,y,zX:

(p1) x=yp(x,x)=p(x,y)=p(y,y),

(p2) p(x,x)≤p(x,y),

(p3) p(x,y)=p(y,x),

(p4) p(x,y)≤p(x,z)+p(z,y)-p(z,z).

In this case, (X,p) is called a partial metric space.

It is clear that |p(x,y)-p(y,z)|≤p(x,z) ∀x,y,zX. It is also clear that p(x,y)=0 implies x=y from (p1) and (p2). However, if x=y, p(x,y) may not be zero. A basic example of a partial metric space is the pair ( R + ,p), where p(x,y)= max{x,y} for all x,y R + . Each partial metric p on X generates a τ0 topology τ p on X which has a base, the family of open p - balls {B p (x,ε)∣xX, ε>0} for all xX and ε>0, where B p (x,ε)={yXp(x,y)<p(x,x)+ε} for all xX and ε>0. If p is a partial metric on X, then the function p s :X×X R + given by ps(x,y)=2p(x,y)-p(x,x)-p(y,y) is a metric on X.

Definition 2.

([9]). Let (X,p) be a partial metric space:

  1. (i)

    A sequence {x n } in (X,p) is said to converge to a point xX if and only if p(x,x)= lim n p(x, x n ).

  2. (ii)

    A sequence {x n } in (X,p) is said to be a Cauchy sequence if lim n , m p( x n , x m ) exists and is finite.

  3. (iii)

    (X,p) is said to be complete if every Cauchy sequence {x n } in X converges, with respect to τ p , to a point xX such that p(x,x)= lim n , m p( x n , x m ).

Lemma 1.

([9]). Let (X,p) be a partial metric space:

(a) {x n } is a Cauchy sequence in (X,p) if and only if it is a Cauchy sequence in the metric space (X,ps).

(b) (X,p) is complete if the metric space (X,ps) is complete. Furthermore, lim n p s ( x n ,x)=0 if and only if

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

Lemma 2.

([33]). Let (X,p) be a partial metric space and A any nonempty set in X. Then, a A ¯ if and only if p(a,A)=p(a,a), where A ¯ denotes the closure of A with respect to the topology of the partial metric p.

Note that A is closed in (X,p) if and only if A= A ¯ .

Consistent with [32], let (X,p) be a partial metric space. Let CBp(X) be the family of all nonempty, closed, and bounded subsets of the partial metric space (X,p), induced by the partial metric p. For A,B∈ CBp(X) and xX, define

p ( A , B ) = inf p ( a , b ) : a A , b B ,
p ( x , A ) = inf p ( x , a ) : a A ,

and

δ p ( A , B ) = sup p ( a , B ) : a A , δ p ( B , A ) = sup p ( b , A ) : b B .

Also,

H p ( A , B ) = max δ p ( A , B ) , δ p ( B , A ) .

H p is called the partial Hausdorff metric induced by a partial metric p.

Also, Aydi et al. [32] proved that any Hausdorff metric is a partial Hausdorff metric and the converse is not true (see Example 2.6 in [32]):

Lemma 3.

([32]). Let (X,p) be a partial metric space. For any A,B,CC Bp(X), we have

(i) δ p (A,A)= sup{p(a,a):aA},

(ii) δ p (A,A)≤δ p (A,B),

(iii) δ p (A,B) = 0 implies that AB,

(iv) δ p (A,B) δ p (A,C)+ δ p (C,B)- inf c C p(c,c).

Lemma 4.

([32]). Let (X,p) be a partial metric space. For any A,B,CC Bp(X), we have

(i) H p (A,A)≤H p (A,B),

(ii) H p (A,B)=H p (B,A),

(iii) H p (A,B) H p (A,C)+ H p (C,B)- inf c C p(c,c).

Lemma 5.

([32]). Let (X,p) be a partial metric space. For any A,BC Bp(X), the following holds

H p ( A , B ) = 0 implies that A = B.

In [32], they also show that H p (A,A) need not be zero by an example.

Lemma 6.

([32]). Let (X,p) be a partial metric space, A,BC Bp(X), and h>1. For any aA, there exists bB such that p(a,b)≤hH p (A,B).

Theorem 2.

([32]). Let (X,p) be a complete partial metric space and T:XC Bp(X) is a multi-valued mapping such that for all x,yX

H p ( Tx , Ty ) k p ( x , y ) ,

where k∈(0,1), then T has a fixed point.

Very recently, Abbas et al. [34] generalized Theorem 2 by proving the following Suzuki type theorem:

Theorem 3.

Let (X,p) be a complete partial metric space. Take T:XC Bp(X) a multi-valued mapping and φ:[ 0,1)→(0,1] a nonincreasing function defined by

φ(r)= 1 if 0 r < 1 2 , 1 - r if 1 2 r < 1 .
(1)

If there exists r∈[0,1) such that T satisfies the condition

(A) φ(r)p(x,T x)≤p(x,y) implies

H p (Tx,Ty)rmax p ( x , y ) , p ( x , Tx ) , p ( y , Ty ) , 1 2 [ p ( x , Ty ) + p ( y , Tx ) ]

for all x,yX, then T has a fixed point, that is, there exists a point zX such that zT z.

Now, we give the following commutativity definitions mentioned in [35].

Definition 3.

([35]) Let (X,p) be a partial metric space. Let f:XX and S:XC Bp(X). The pair (f,S) is called

  1. (i)

    commuting if f S x=S f x,∀xX,

  2. (ii)

    weakly compatible if the pair (f,S) commutes at their coincidence points, that is, f S x=S f x whenever f xS x for xX,

  3. (iii)

    IS-commuting at xX if f S xS f x.

Generally, to prove a coincidence point or a common fixed-point theorem for hybrid mappings, one has to assume a commutativity condition and continuity of mappings. In this paper, we introduce a new condition and prove a unique common fixed-point theorem for hybrid mappings in partial metric spaces without using any standard arguments as commutativity and continuity conditions.

Main results

We start with the following lemma which is needed to prove our main results:

Lemma 7.

Let x n x as n in a partial metric space (X,p) such that p(x,x)=0, then lim n p( x n ,B)=p(x,B) for any BC Bp(X).

Proof.

Since x n x, we have lim n p( x n ,x)=p(x,x)=0. Applying a triangular inequality for x n X and yB, we get

p ( x n , B ) p ( x n , y ) p ( x n , x ) + p ( x , y ) - p ( x , x ) p ( x n , x ) + p ( x , y )

which implies that lim n p( x n ,B)p(x,y) for all yB. Therefore,

( i ) lim n p ( x n , B ) p ( x , B ) .

Similarly,

p ( x , y ) p ( x , x n ) + p ( x n , y ) - p ( x n , x n )

so p(x,y)≤p(x,x n )+p(x n ,y). Thus, p(x,B)≤p(x,x n )+p(x n ,B). Therefore,

( ii ) p ( x , B ) lim n p ( x n , B ) .

From (i) and (ii), we have lim n p( x n ,B)=p(x,B). □

Now, we introduce the following new condition, namely the W.C.C. condition, on mappings which are not necessarily continuous and commutative.

Definition 4.

Let (X,p) be a partial metric space. Let f:XX and S:XC Bp(X) be mappings. Then, the pair (f,S) is said to satisfy the W.C.C. condition if p(f x,f y)≤p(y,S x), ∀x,yX.

The following example illustrates the W.C.C. condition:

Example 1.

Let X= [ 0,1] and p(x,y)= max{x,y}, ∀x,yX. Let f:XX and S:XC Bp(X) be defined by

fx = 0 if x [ 0 , 1 2 ] 3 x 4 if x ( 1 2 , 1 ]

and Sx=[ 3 4 ,1], ∀x,yX. We consider the following four cases:

Case 1: x[0, 1 2 ] and y[0, 1 2 ]. Here, p(f x,f y)=0<p(y,S x).

Case 2: x[0, 1 2 ] and y( 1 2 ,1]. Then, p(fx,fy)= 3 y 4 3 4 =p(y,Sx).

Case 3: x( 1 2 ,1] and y[0, 1 2 ]. Here, p(fx,fy)= 3 x 4 3 4 =p(y,Sx).

Case 4: x( 1 2 ,1] and y( 1 2 ,1]. Then, p(fx,fy)=max{ 3 x 4 , 3 y 4 } 3 4 =p(y,Sx).

Thus (f,S) satisfies the W.C.C. condition. In this example, the pair (f,S) does not satisfy any type of commutativity mentioned in Definition 3.

The following example shows that the pair (f,S) satisfying the W.C.C condition need not be continuous even when S is a single-valued mapping:

Example 2.

Let X= [ 0,1] and p(x,y)= max{x,y}, ∀x,yX. Let f,S:XX be defined by

fx = x 6 if x 1 1 4 if x = 1

and

Sx = x if x 1 1 2 if x = 1 .

We distinguish the following cases:

Case (i): x≠1 and y≠1. We have p(fx,fy)=max{ x 6 , y 6 }= 1 6 max{x,y}= 1 6 p(y,Sx).

Case (ii): x≠1 and y=1. Then, p(fx,fy)=max{ x 6 , 1 4 }= 1 4 <1=p(y,Sx).

Case (iii): x=1 and y≠1. We have p(fx,fy)=max{ 1 4 , y 6 }= 1 4 < 1 2 p(y,Sx).

Case (iv): x=1 and y=1. Here, p(fx,fy)= 1 4 <1=p(y,Sx).

Thus (f,S) satisfies the W.C.C. condition.

In this example, note that f and S are discontinuous.

Now, we state and prove our main results.

Theorem 4.

Let (X,p) be a complete partial metric space. Let S,T:XC Bp(X) and f:XX. Assume that there exists r∈ [ 0,1) such that for every x,yX

(A 1) φ(r) min {p(f x,S x),p(f y,T y)}≤p(f x,f y) implies

H p ( Sx , Ty ) r max p ( fx , fy ) , p ( fx , Sx ) , p ( fy , Ty ) , 1 2 [ p ( fx , Ty ) + p ( fy , Sx ) ]

where φ is defined by (1),

(A 2) x X Sxf(X) and x X Txf(X),

(A 3) The pair (f,S) or the pair (f,T) satisfies the W.C.C condition.

Then f,S and T have a unique common fixed point in X.

Proof.

Let x0X and suppose that h= 1 r >0, y0=fx0. Now from (A 2), we have Sx0f(X), so there exists x1X such that y1=fx1Sx0.

By Lemma 6 with h= 1 r , there exists y2Tx1 such that

p ( fx 1 , y 2 ) 1 r H p ( Sx 0 , Tx 1 ) .

Since Tx1f(X), we may find a point x2X such that y2=fx2Tx1. Therefore,

p ( fx 1 , fx 2 ) 1 r H p ( Sx 0 , Tx 1 ) .

Since φ(r)p(fx0,Sx0)≤p(fx0,Sx0)≤p(fx0,fx1), we have

φ ( r ) min p ( fx 0 , Sx 0 ) , p ( fx 1 , Tx 1 ) p ( fx 0 , fx 1 ) .

By (A 1), we have

p ( fx 1 , fx 2 ) h H p ( Sx 0 , Tx 1 ) = 1 r H p ( Sx 0 , Tx 1 ) , r max p ( fx 0 , fx 1 ) , p ( fx 0 , Sx 0 ) , p ( fx 1 , Tx 1 ) , 1 2 p ( fx 0 , Tx 1 ) + p ( fx 1 , Sx 0 ) r max p ( y 0 , y 1 ) , p ( y 0 , y 1 ) , p ( y 1 , y 2 ) , 1 2 p ( y 0 , y 2 ) + p ( y 1 , y 1 ) p ( y 1 , y 2 ) r max p ( y 0 , y 1 ) , p ( y 1 , y 2 ) , 1 2 p ( y 0 , y 1 ) + p ( y 1 , y 2 ) , from ( p 4 ) r max p ( y 0 , y 1 ) , p ( y 1 , y 2 ) .

If p(y0,y1)<p(y1,y2) then p( y 1 , y 2 ) r p( y 1 , y 2 ) which is a contradiction. Hence, p(y0,y1)≥p(y1,y2. Thus, we have

p( y 1 , y 2 )βp( y 0 , y 1 ),
(2)

where β= r <1.

As fx2Tx1, from Lemma 6, we choose y3Sx2 such that

p ( fx 2 , y 3 ) 1 r H p ( Sx 2 , Tx 1 ) .

Since Sx2g(X), we find a point x3X such that y3=fx3Sx2. Therefore,

p ( fx 2 , fx 3 ) 1 r H p ( Sx 2 , Tx 1 ) .

Since φ(r)p(fx1,Tx1)≤p(fx1,Tx1)≤p(fx2,fx1), we have

φ ( r ) min p ( fx 2 , Sx 2 ) , p ( fx 1 , Tx 1 ) p ( fx 2 , fx 1 ) .

Hence, by (A 1), we have

p ( fx 2 , fx 3 ) 1 r H p ( Sx 2 , Tx 1 ) , r max p ( fx 2 , fx 1 ) , p ( fx 2 , Sx 2 ) , p ( fx 1 , Tx 1 ) , 1 2 p ( fx 2 , Tx 1 ) + p ( fx 1 , Sx 2 ) r max p ( y 2 , y 1 ) , p ( y 2 , y 3 ) , p ( y 1 , y 2 ) , 1 2 p ( y 2 , y 2 ) + p ( y 1 , y 3 ) p ( y 2 , y 3 ) r max p ( y 1 , y 2 ) , p ( y 2 , y 3 ) ] , from ( p 4 ) .

Thus, we have

p( y 2 , y 3 )βp( y 1 , y 2 ) β 2 p( y 0 , y 1 ).
(3)

Continuing in this way, we obtain a sequence {y n } in X such that for any nN,

y 2 n + 1 = fx 2 n + 1 Sx 2 n , y 2 n + 2 = fx 2 n + 2 Tx 2 n + 1

and

p( y n , y n + 1 ) β n p( y 0 , y 1 ).
(4)

Clearly,

p( y n + 1 , y n )0asn∞.
(5)

For m>n, we have

p ( y n , y m ) p ( y n , y n + 1 ) + p ( y n + 1 , y n + 2 ) + + p ( y m - 1 , y m ) , β n + β n + 1 + + β m - 1 p ( y 1 , y 0 ) , from ( 4 )
β n 1 - β p( y 1 , y 0 )0asn∞.
(6)

Thus, {y n } is a Cauchy sequence in X. Hence from Lemma 1, we have {y n } is a Cauchy sequence in the metric space (X,ps).

Since (X,p) is complete and again from Lemma 1, it follows that (X,ps) is complete. So, {y n } converges to some z in (X,ps). That is

lim n p s ( y n , z ) = 0 .

Now, from Lemma 1 and (6), we have

p(z,z)= lim n p( y n ,z)= lim n p( y n , y m )=0.
(7)

Suppose the pair (f,S) satisfies the W.C.C condition. Then,

p(fx,fy)p(y,Sx)for allx,yX.
(8)

From (8), we have

p ( fx 2 n , fz ) p ( z , Sx 2 n ) p ( z , fx 2 n + 1 ) .

Letting n and using Lemma 7 and (7), we can obtain

p(z,fz)0so thatfz=z.
(9)
Claim : p ( fz , Sx ) r max { p ( fx , fz ) , p ( fx , Sx ) } for any fx X - { fz } .
(10)

Let f xX-f z. Since y2n+1z=f z, y2n+2z=f z and p(z,z)= lim n p( y n ,z)=0, there exists a positive integer n0 such that for all nn0, we have

p ( fz , fx 2 n + 1 ) 1 3 p ( fz , fx )

and

p ( fz , fx 2 n + 2 ) 1 3 p ( fz , fx ) .

So, for any nn0, we have

φ ( r ) p ( fx 2 n + 1 , Tx 2 n + 1 ) p ( fx 2 n + 1 , Tx 2 n + 1 ) , p ( fx 2 n + 2 , fx 2 n + 1 ) , p ( fx 2 n + 2 , fz ) + p ( fz , fx 2 n + 1 ) , 2 3 p ( fz , fx ) , = p ( fx , fz ) - 1 3 p ( fx , fz ) , p ( fx , fz ) - p ( fz , fx 2 n + 1 ) , p ( fx , fx 2 n + 1 ) .

Hence, we have

φ ( r ) min p ( fx , Sx ) , p ( fx 2 n + 1 , Tx 2 n + 1 ) p ( fx , fx 2 n + 1 )

which implies that

p ( fx 2 n + 2 , Sx ) H p ( Sx , Tx 2 n + 1 ) , r max p ( fx , fx 2 n + 1 ) , p ( fx , Sx ) , p ( fx 2 n + 1 , Tx 2 n + 1 ) , 1 2 p ( fx , Tx 2 n + 1 ) + p ( fx 2 n + 1 , Sx ) .

Letting n, we get

p ( fz , Sx ) r max p ( fz , fx ) , p ( fx , Sx ) , p ( fz , fz ) , 1 2 [ p ( fx , fz ) + p ( fz , Sx ) ] r max p ( fx , fz ) , p ( fx , Sx ) , 1 2 [ p ( fx , fz ) + p ( fz , Sx ) ] , from ( p 2 ) .

If max p ( fx , fz ) , p ( fx , Sx ) , 1 2 [ p ( fx , fz ) + p ( fz , Sx ) ] =max{p(fx,fz),p(fx,Sx)}, then

p ( fz , Sx ) r max p ( fx , fz ) , p ( fx , Sx ) .

If max p ( fx , fz ) , p ( fx , Sx ) , 1 2 [ p ( fx , fz ) + p ( fz , Sx ) ] = 1 2 [p(fx,fz)+p(fz,Sx)], then

p ( fz , Sx ) 1 2 [ p ( fx , fz ) + p ( fz , Sx ) ]

which implies that

( 1 - r 2 ) p ( fz , Sx ) r 2 p ( fx , fz ) .

So,

p ( fz , Sx ) r 2 - r p ( fx , fz ) rp ( fx , fz ) r max p ( fx , fz ) , p ( fx , Sx ) .

Hence, (10) is proved.

Now, we will show that f zT z. First, consider the case 0r< 1 2 . On the contrary, suppose that fzTz= Tz ¯ as Tz is closed. Hence, by Lemma 2, together with (7) and (9), we have

p(fz,Tz)p(fz,fz)=p(z,z)=0.
(11)

Then, from (A2) and (11), we can choose f aT z such that

2rp(fa,fz)<p(fz,Tz).
(12)

Having f aT z and f zT z imply f af z, then by (10)

p(fz,Sa)rmax{p(fz,fa),p(fa,Sa)}.
(13)

Since φ(r)p(f z,T z)≤p(f z,T z)≤p(f a,f z), so it follows that

φ ( r ) min p ( fa , Sa ) , p ( fz , Tz ) p ( fa , fz ) .

Now by (A 1), we have

H p ( Sa , Tz ) r max p ( fa , fz ) , p ( fa , Sa ) , p ( fz , Tz ) , 1 2 [ p ( fa , Tz ) + p ( fz , Sa ) ] r max p ( fa , fz ) , p ( fa , Sa ) , p ( fz , fa ) , 1 2 [ p ( fa , fa ) + p ( fz , fa ) + p ( fa , Sa ) - p ( fa , fa ) ] r max p ( fa , fz ) , p ( fa , Sa ) , 1 2 [ p ( fa , Sa ) + p ( fz , fa ) ] r max p ( fa , fz ) , p ( fa , Sa ) .

Since f aT z, then p(f a,S a)≤H p (S a,T z). Therefore, we obtain

H p ( Sa , Tz ) r max p ( fa , fz ) , H p ( Sa , Tz ) .

Since r<1, it follows that

H p (Sa,Tz)rp(fa,fz)
(14)

Thus, we have

p(fa,Sa) H p (Sa,Tz)rp(fa,fz)<p(fa,fz).
(15)

By (13)

p(fz,Sa)rp(fa,fz).
(16)

Also,

p ( Sa , Tz ) = inf p ( x , y ) : x Sa , y Tz inf p ( x , fa ) : x Sa since fa Tz = p ( fa , Sa ) H p ( Sa , Tz )

and by (14) and (16), we have

p ( fz , Tz ) p ( fz , Sa ) + p ( Sa , Tz ) p ( fz , Sa ) + H p ( Sa , Tz ) rp ( fa , fz ) + rp ( fa , fz ) = 2 rp ( fa , fz ) < p ( fz , Tz ) , from ( 12 ) .

It is a contradiction, so f zT z. Thus, from (9)

z=fzTz.
(17)

Now, from (8), we have

p(fz,z)=p(fz,fz)p(z,Sz).
(18)

Since f zT z, so we have

φ ( r ) p ( fz , Tz ) p ( fz , Tz ) p ( fz , fz ) ,

which implies that

φ ( r ) min p ( fz , Sz ) , p ( fz , Tz ) p ( fz , fz ) .

Now, by (A 1)

p ( Sz , z ) H p ( Sz , Tz ) r max p ( fz , fz ) , p ( fz , Sz ) , p ( fz , Tz ) , 1 2 [ p ( fz , Tz ) + p ( fz , Sz ) ] = r max p ( z , z ) , p ( z , Sz ) , p ( z , Tz ) ] , 1 2 [ p ( z , Tz ) + p ( z , Sz ) ] rp ( z , Sz ) from ( 7 )

which in turn yields that p(z,S z)=0. By Lemma 2 and (7), we have zS z. Hence,

fz=z=Sz
(19)

From (17) and (19), z is a common fixed point of f, S, and T. Now, we consider the case 1 2 r<1. First, we prove that

H p (Sx,Tz)rmax p ( fx , fz ) , p ( fx , Sx ) , p ( fz , Tz ) , 1 2 [ p ( fx , Tz ) + p ( fz , Sx ) ]
(20)

for all xX such that f xf z.

Assume that f xf z. Then, for every nN, there exists z n S x such that

p ( fz , z n ) p ( fz , Sx ) + 1 n p ( fx , fz ) .

Therefore,

p ( fx , Sx ) p ( fx , z n ) p ( fx , fz ) + p ( fz , z n ) p ( fx , fz ) + p ( fz , Sx ) + 1 n p ( fx , fz ) p ( fx , fz ) + r max { p ( fz , fx ) , p ( fx , Sx ) } + 1 n p ( fx , fz ) , from ( 10 ) .

Hence, we have either p(fx,Sx)(1+r+ 1 n )p(fx,fz) or (1-r)p(fx,Sx)(1+ 1 n )p(fx,fz).

Letting n, we get

p ( fx , Sx ) ( 1 + r ) p ( fx , fz ) or ( 1 - r ) p ( fx , Sx ) p ( fx , fz ) .

Thus,

φ ( r ) p ( fx , Sx ) = ( 1 - r ) p ( fx , Sx ) 1 1 + r p ( fx , Sx ) p ( fx , fz ) ,

or

φ ( r ) p ( fx , Sx ) = ( 1 - r ) p ( fx , Sx ) p ( fx , fz ) .

Hence, we have

φ ( r ) min p ( fx , Sx ) , p ( fz , Tz ) p ( fx , fz ) .

Now, by (A 1), with y=z we get (20).

Since y n z, we may assume that y n z for any n. Taking x=x2n in (20), we get

p ( fx 2 n + 1 , Tz ) H p ( Sx 2 n , Tz ) r max { p ( fx 2 n , fz ) , p ( fx 2 n , Sx 2 n ) , p ( fz , Tz ) , 1 2 p ( fx 2 n , Tz ) + p ( fz , Sx 2 n ) } .

Letting n, using Lemma 7, (5), (7), and (9), we get

p ( z , Tz ) r max { 0 , 0 , p ( z , Tz ) , 1 2 [ p ( z , Tz ) + 0 ] } rp ( z , Tz )

which in turn yields that p(z,T z)=0 so that zT z. Thus f z=zT z.

Now, following as in the case 0r< 1 2 , and from (15) to (17), we have z=f zS z. Thus, z is a common fixed point of f,S and T. Thus, from the two cases above, we have z is a common fixed point of f, S, and T.

Suppose z is another common fixed point of f, S, and T.

By (8), we have

p(z, z )=p(fz,f z )p( z ,Sz) H p (Sz,T z ).
(21)

Using (p2)

φ ( r ) min p ( fz , Sz ) , p ( f z , T z ) p ( fz , f z ) .

Hence, by (A 1)

H p ( Sz , T z ) r max p ( fz , f z ) , p ( fz , Sz ) , p ( f z , T z ) , 1 2 p ( fz , T z ) + p ( f z , Sz ) r max H p ( Sz , T z ) , H p ( Sz , Sz ) , H p ( T z , T z ) , 1 2 H p ( Sz , T z ) + H p ( T z , Sz ) from ( 19 ) r H p ( Sz , T z ) from Lemma 4 (i).

Thus, H p (S z,T z)=0, so that from (21), we have z=z. Hence, z is the unique common fixed point of f, S, and T.

Similarly, we can prove the theorem when (f,T) satisfies the W.C.C. condition. □

Next, take f=I X (the identity map on X) in Theorem 4, we have the following corollary for two multi-valued maps.

Corollary 1.

Let (X,p) be a complete partial metric space and let S,T:XC Bp(X). Assume that there exists r∈ [ 0,1) such that for every x,yX,

(B 1) φ(r) min {p(x,S x),p(y,T y)}≤p(x,y) implies

H p ( Sx , Ty ) r max p ( x , y ) , p ( x , Sx ) , p ( y , Ty ) ] , 1 2 [ p ( x , Ty ) + p ( y , Sx ) ]

where φ is a function defined by (1).

(B 2) The pair (I X ,S) or the pair (I X ,T) satisfies the W.C.C. condition.

Then, S and T have a common fixed point in X, that is, there exists an element zX such that zS zT z.

Taking S=T in the above corollary, we get the following:

Corollary 2.

Let (X,p) be a complete partial metric space and let T:XC Bp(X). Assume that there exists r∈ [ 0,1) such that for every x,yX,

(C 1) φ(r) min {p(x,T x),p(y,T y)}≤p(x,y) implies

H p ( Tx , Ty ) r max p ( x , y ) , p ( x , Tx ) , p ( y , Ty ) ] , 1 2 [ p ( x , Ty ) + p ( y , Tx ) ]

where φ is a function defined by (1).

(C 2) The pair (I X ,T) satisfies the (W.C.C) condition.

Then, T has a unique fixed point in X, that is, there exists an element zX such that zT z.

In case of single-valued maps, Theorem 4 reduces to the following corollary:

Corollary 3.

Let (X,p) be a complete partial metric space and f,S,T:XX. Assume that there exists r∈ [ 0,1) such that for every x,yX every x,yX

(D 1) φ(r) min {p(f x,S x),p(f y,T y)}≤p(f x,f y) implies

H p ( Sx , Ty ) r max p ( fx , fy ) , p ( fx , Sx ) , p ( fy , Ty ) , 1 2 [ p ( fx , Ty ) + p ( fy , Sx ) ]

where φ is defined by (1).

(D 2) x X Sxf(X) and x X Txf(X).

(D 3) The pair (f,S) or the pair (f,T) satisfies the W.C.C. condition.

Then f, S, and T have a unique common fixed point in X.

We drop the W.C.C. condition in Corollary 2 to get a fixed-point result (without uniqueness):

Corollary 4.

Let (X,p) be a complete partial metric space and let T:XC Bp(X). Assume that there exists r∈ [ 0,1) such that for every x,yX,

(E 1) φ(r) min {p(x,T x),p(y,T y)}≤p(x,y) implies

H p ( Tx , Ty ) r max p ( x , y ) , p ( x , Tx ) , p ( y , Ty ) ] , 1 2 [ p ( x , Ty ) + p ( y , Tx ) ]

where φ is a function defined by (1).

Then, T has a fixed point in X, that is, there exists an element zX such that zT z.

Similarly, for single-valued maps we have

Corollary 5.

Let (X,p) be a complete partial metric space and T:XX. Assume that there exists r∈ [ 0,1) such that for every x,yX every x,yX

(F 1) φ(r) min {p(x,T x),p(y,T y)}≤p(x,y) implies

H p ( Tx , Ty ) r max p ( x , y ) , p ( x , Tx ) , p ( y , Ty ) , 1 2 [ p ( x , Ty ) + p ( y , Tx ) ]

where φ is defined by (1).

Then, T has a fixed point in X.

Remark 1.

Corollary 4 is a generalization of Theorem 3. Also, Corollary 4 improves and extends the main result of Doricć and Lazović [5] to partial metric spaces.