Dedication

Dedicated to Professor Wataru Takahashi on the occasion of his seventieth birthday

1 Introduction and preliminaries

In metric fixed point theory, the contractive conditions on underlying functions play an important role for finding solution of fixed point problems. Banach contraction principle is a remarkable result in metric fixed point theory. Over the years, it has been generalized in different directions by several mathematicians (see [121]). In 2012, Samet et al. [15] introduced the concepts of α-ψ-contractive and α-admissible mappings and established various fixed point theorems for such mappings in complete metric spaces. Afterwards, Karapinar and Samet [13] generalized these notions to obtain fixed point results. More recently, Salimi et al. [14] modified the notions of α-ψ-contractive and α-admissible mappings and established fixed point theorems, which are proper generalizations of the recent results in [13, 15]. Here, we continue to utilize these modified notions for single-valued Geraghty and Meir-Keeler-type contractions, as well as multivalued contractive mappings. Presented theorems provide main results of Hussain et al. [9], Karapinar et al. [11] and Asl et al. [12] as corollaries. Moreover, some examples are given here to illustrate the usability of the obtained results.

Denote with Ψ the family of 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.

The following lemma is obvious.

Lemma 1.1 If ψΨ, then ψ(t)<t for all t>0.

Samet et al. [15] defined the notion of α-admissible mappings as follows.

Definition 1.1 Let T be a self-mapping on X, and let 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.

Theorem 1.1 [15]

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

α(x,y)d(Tx,Ty)ψ ( d ( x , y ) )
(1.1)

for all x,yX, where ψΨ. Also, suppose that

  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} and x n x as n+, we have α( x n ,x)1 for all nN{0}.

Then T has a fixed point.

Very recently Salimi et al. [14] modified the notions of α-admissible and α-ψ-contractive mappings as follows.

Definition 1.2 [14]

Let T be a self-mapping on X, and let α,η:X×X[0,+) be two functions. We say that T is an α-admissible mapping with respect to η if

x,yX,α(x,y)η(x,y)α(Tx,Ty)η(Tx,Ty).

Note that if we take η(x,y)=1, then this definition reduces to Definition 1.1. Also, if we take α(x,y)=1, then we say that T is η-subadmissible mapping.

The following result properly contains Theorem 1.1 and Theorems 2.3 and 2.4 of [13].

Theorem 1.2 [14]

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

x,yX,α(x,y)η(x,y)d(Tx,Ty)ψ ( M ( x , y ) ) ,
(1.2)

where ψΨ and

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

Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

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

Then T has a fixed point.

2 Modified α-η-Geraghty type contractions

Our first main result of this section is concerning α-η-Geraghty-type [4] contractions.

Theorem 2.1 Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping with respect to η. Assume that there exists a function β:[0,)[0,1) such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

x , y X , α ( x , f x ) α ( y , f y ) η ( x , f x ) η ( y , f y ) d ( f x , f y ) β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } } .
(2.1)

Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, α( x n , x n + 1 )η( x n , x n + 1 ) for all n, then α(x,fx)η(x,fx).

If there exists x 0 X such that α( x 0 ,f x 0 )η( x 0 ,f x 0 ), then f has a fixed point.

Proof Let x 0 X such that α( x 0 ,f x 0 )η( x 0 ,f x 0 ). Define a sequence { x n } in X by x n = f n x 0 =f x n 1 for all nN. If x n + 1 = x n for some nN, then x= x n is a fixed point for f, and the result is proved. Hence, we suppose that x n + 1 x n for all nN. Since f is an α-admissible mapping with respect to η and α( x 0 ,f x 0 )η( x 0 ,f x 0 ), we deduce that α( x 1 , x 2 )=α(f x 0 , f 2 x 0 )η(f x 0 , f 2 x 0 )=η( x 1 , x 2 ). By continuing this process, we get α( x n ,f x n )η( x n ,f x n ) for all nN{0}. Then,

α( x n 1 ,f x n 1 )α( x n ,f x n )η( x n 1 ,f x n 1 )η( x n ,f x n ).

Now from (2.1), we have

d ( x n , x n + 1 ) β ( d ( x n 1 , x n ) ) max { d ( x n 1 , x n ) , min { d ( x n 1 , f x n 1 ) , d ( x n , f x n ) } } = β ( d ( x n 1 , x n ) ) max { d ( x n 1 , x n ) , min { d ( x n 1 , x n ) , d ( x n , x n + 1 ) } } .

Now, if d( x n 1 , x n )<d( x n , x n + 1 ) for some nN, then

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

Also, if d( x n , x n + 1 )d( x n 1 , x n ) for some nN, then

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

That is, for all nN, we have

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

Hence,

d( x n , x n + 1 )β ( d ( x n 1 , x n ) ) d( x n 1 , x n )
(2.2)

for all nN, which implies that d( x n , x n + 1 )d( x n 1 , x n ). It follows that the sequence {d( x n , x n + 1 )} is decreasing. Thus, there exists d R + such that lim n d( x n , x n + 1 )=d. We shall prove that d=0. From (2.2), we have

d ( x n , x n + 1 ) d ( x n 1 , x n ) β ( d ( x n 1 , x n ) ) 1,

which implies that lim n β(d( x n 1 , x n ))=1. Regarding the property of the function β, we conclude that

lim n d( x n , x n + 1 )=0.
(2.3)

Next, we shall prove that { x n } is a Cauchy sequence. Suppose, to the contrary, that { x n } is not a Cauchy sequence. Then there is ε>0 and sequences {m(k)} and {n(k)} such that for all positive integers k, we have

n(k)>m(k)>k,d( x n ( k ) , x m ( k ) )εandd( x n ( k ) , x m ( k ) 1 )<ε.

By the triangle inequality, we derive that

ε d ( x n ( k ) , x m ( k ) ) d ( x n ( k ) , x m ( k ) 1 ) + d ( x m ( k ) 1 , x m ( k ) ) < ε + d ( x m ( k ) 1 , x m ( k ) )

kN. Taking the limit as k+ in the inequality above, and regarding the limit in (2.3), we get

lim k + d( x n ( k ) , x m ( k ) )=ε.
(2.4)

Again, by the triangle inequality, we find that

d( x n ( k ) , x m ( k ) )d( x m ( k ) , x m ( k ) + 1 )+d( x m ( k ) + 1 , x n ( k ) + 1 )+d( x n ( k ) + 1 , x n ( k ) )

and

d( x n ( k ) + 1 , x m ( k ) + 1 )d( x m ( k ) , x m ( k ) + 1 )+d( x m ( k ) , x n ( k ) )+d( x n ( k ) + 1 , x n ( k ) ).

Taking the limit in inequality above as k+, together with (2.3) and (2.4), we deduce that

lim k + d( x n ( k ) + 1 , x m ( k ) + 1 )=ε.
(2.5)

Now, since

α( x n ( k ) ,f x n ( k ) )α( x m ( k ) ,f x m ( k ) )η( x n ( k ) ,f x n ( k ) )η( x m ( k ) ,f x m ( k ) ),

then from (2.1), (2.4) and (2.5), we have

d ( x n ( k ) + 1 , x m ( k ) + 1 ) β ( d ( x n ( k ) , x m ( k ) ) ) max { d ( x n ( k ) , x m ( k ) ) , min { d ( x n ( k ) , f x n ( k ) ) , d ( x m ( k ) , f x m ( k ) ) } } = β ( d ( x n ( k ) , x m ( k ) ) ) max { d ( x n ( k ) , x m ( k ) ) , min { d ( x n ( k ) , x n ( k ) + 1 ) , d ( x m ( k ) , x m ( k ) + 1 ) } } .

Hence,

d ( x n ( k ) + 1 , x m ( k ) + 1 ) max { d ( x n ( k ) , x m ( k ) ) , min { d ( x n ( k ) , x n ( k ) + 1 ) , d ( x m ( k ) , x m ( k ) + 1 ) } } β ( d ( x n ( k ) , x m ( k ) ) ) 1.

Letting k in the inequality above, we get

lim n β ( d ( x n ( k ) , x m ( k ) ) ) =1.

That is, lim k d( x n ( k ) , x m ( k ) )=0, which is a contradiction. Hence { x n } is a Cauchy sequence. Since X is complete, then there is zX such that x n z. First, we suppose that f is continuous. Since f is continuous, then we have

fz= lim n f x n = lim n x n + 1 =z.

So z is a fixed point of f. Next, we suppose that (b) holds. Then, α(z,fz)η(z,fz), and so, α(z,fz)α( x n ,f x n )η(z,fz)η( x n ,f x n ). Now by (2.1), we have

d(fz, x n + 1 )β ( d ( z , x n ) ) max { d ( z , x n ) , min { d ( z , f z ) , d ( x n , x n + 1 ) } } ,

and hence

d ( f z , z ) d ( f z , x n + 1 ) + d ( z , x n + 1 ) β ( d ( z , x n ) ) max { d ( z , x n ) , min { d ( z , f z ) , d ( x n , x n + 1 ) } } + d ( z , x n + 1 ) .

Letting n in the inequality above, we get d(fz,z)=0, that is, z=fz. □

If in Theorem 2.1 we take, η(x,y)=1, then we have the following corollary.

Corollary 2.1 Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

x , y X , α ( x , f x ) α ( y , f y ) 1 d ( f x , f y ) β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } } .

Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.2 Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

( d ( f x , f y ) + ) α ( x , f x ) α ( y , f y ) β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } } +

for all x,yX, where >0. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.3 Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

( α ( x , f x ) α ( y , f y ) + 1 ) d ( f x , f y ) 2 β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } }

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.4 Let (X,d) be a metric space such that (X,d) is complete and f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

α(x,fx)α(y,fy)d(fx,fy)β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } }

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, α( x n ,f x n )1 for all n, then α(x,fx)1.

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Further, if in Theorem 2.1 we take α(x,y)=1, then we have the following corollary.

Corollary 2.5 Let (X,d) be a complete metric space, and let f:XX be a η-subadmissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

x , y X , η ( x , f x ) η ( y , f y ) 1 d ( f x , f y ) β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } } .

Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, η( x n , x n + 1 )1 for all n, then η(x,fx)1.

If there exists x 0 X such that η( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.6 Let (X,d) be a complete metric space, and let f:XX be a η-subadmissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

d(fx,fy)+ [ β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } } + ] η ( x , f x ) η ( y , f y )

for all x,yX, where >0. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, η( x n , x n + 1 )1 for all n, then η(x,fx)1.

If there exists x 0 X such that η( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.7 Let (X,d) be a complete metric space, and let f:XX be a η-subadmissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

2 d ( f x , f y ) ( η ( x , f x ) η ( y , f y ) + 1 ) β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } }

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, η( x n , x n + 1 )1 for all n, then η(x,fx)1.

If there exists x 0 X such that η( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.8 Let (X,d) be a metric space such that (X,d) is complete, and let f:XX be η-subadmissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

d(fx,fy)η(x,fx)η(y,fy)β ( d ( x , y ) ) max { d ( x , y ) , min { d ( x , f x ) , d ( y , f y ) } }

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, η( x n ,f x n )1 for all n, then η(x,fx)1.

If there exists x 0 X such that η( x 0 ,f x 0 )1, then f has a fixed point.

From Corollary 2.1, we can deduce the following corollary.

Corollary 2.9 Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

x,yX,α(x,fx)α(y,fy)1d(fx,fy)β ( d ( x , y ) ) d(x,y).

Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Also, from the corollary above, we can deduce the following corollaries.

Corollary 2.10 (Theorem 4 of [9])

Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

( d ( f x , f y ) + ) α ( x , f x ) α ( y , f y ) β ( d ( x , y ) ) d(x,y)+

for all x,yX, where 1. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.11 (Theorem 6 of [9])

Let (X,d) be a complete metric space, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

( α ( x , f x ) α ( y , f y ) + 1 ) d ( f x , f y ) 2 β ( d ( x , y ) ) d ( x , y )

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

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

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Corollary 2.12 (Theorem 8 of [9])

Let (X,d) be a metric space such that (X,d) is complete, and let f:XX be an α-admissible mapping. Assume that there exists a function β:[0,)[0,1] such that for any bounded sequence { t n } of positive reals, β( t n )1 implies that t n 0 and

α(x,fx)α(y,fy)d(fx,fy)β ( d ( x , y ) ) d(x,y)

for all x,yX. Suppose that either

  1. (a)

    f is continuous, or

  2. (b)

    if { x n } is a sequence in X such that x n x, α( x n ,f x n )1 for all n, then α(x,fx)1.

If there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Example 2.1 Let X=[0,) be endowed with the usual metric d(x,y)=|xy| for all x,yX, and let f:XX be defined by

fx={ 1 4 x if  x [ 0 , 1 ] , ln ( x 2 + x + 3 ) if  x ( 1 , ) .

Define also α:X×X[0,+) and ψ:[0,)[0,) by

α(x,y)={ 6 if  x , y [ 0 , 1 ] , 0 otherwise andβ(t)= 1 2 .

We prove that Corollary 2.9 can be applied to f, but Corollaries 2.10, 2.11 and 2.12 (Theorem 4, 6 and 8 of [9]) cannot be applied to f.

Clearly, (X,d) is a complete metric space. We show that f is an α-admissible mapping. Let x,yX with α(x,y)1, then x,y[0,1]. On the other hand, for all x[0,1], we have fx1. It follows that α(fx,fy)1. Hence, the assertion holds. Also, 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+, then { x n }[0,1], and hence x[0,1]. This implies that α( x n ,x)1 for all nN.

Let α(x,y)1. Then x,y[0,1]. We get,

d(fx,fy)=|fyfx|= | 1 4 x 1 4 y | = 1 4 |xy| 1 2 |xy|=β ( d ( x , y ) ) d(x,y).

That is,

α(x,y)1d(fx,fy)β ( d ( x , y ) ) d(x,y),

then the conditions of Corollary 2.1 hold, and f has a fixed point.

Let x=0, y=1, and let =1, then

( d ( f 0 , f 1 ) + 1 ) α ( 0 , f 0 ) α ( 1 , f 1 ) = ( 1 / 4 + 1 ) 36 >1/2+1=β ( d ( 0 , 1 ) ) d(0,1)+1.

That is, Corollary 2.10 (Theorem 4 of [9]) cannot be applied for this example.

Let, x=0, and let y=1, then

( α ( 0 , f 0 ) α ( 1 , f 1 ) + 1 ) d ( f 0 , f 1 ) = 37 4 > 2 = 2 β ( d ( 0 , 1 ) ) d ( 0 , 1 ) .

That is, Corollary 2.11 (Theorem 6 of [9]) cannot be applied for this example.

Let, x=0, and let y=1, then

α(0,f0)α(1,f1)d(f0,f1)=9>1/2=β ( d ( 0 , 1 ) ) d(0,1).

That is, Corollary 2.12 (Theorem 8 of [9]) cannot be applied for this example.

3 Modified α-ψ-Meir-Keeler contractive mappings

Recently, Karapinar et al. [11] introduced the notion of a triangular α-admissible mapping as follows.

Definition 3.1 [11]

Let f:XX, and let α:X×X(,+). We say that f is a triangular α-admissible mapping if

  1. (T1)

    α(x,y)1 implies that α(fx,fy)1, x,yX;

  2. (T2)

    { α ( x , z ) 1 , α ( z , y ) 1 implies that α(x,y)1.

Lemma 3.1 [11]

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 sequence { x n } by x n = f n x 0 . Then

α( x m , x n )1for all m,nN with m<n.

Denote with Ψ the family of nondecreasing functions ψ:[0,+)[0,+) continuous at t=0 such that

  • ψ(t)=0 if and only if t=0,

  • ψ(t+s)ψ(t)+ψ(s).

Definition 3.2 [11]

Let (X,d) be a metric space, and let ψΨ. Suppose that f:XX is a triangular α-admissible mapping satisfying the following condition:

for each ε>0 there exists δ>0 such that

εψ ( d ( x , y ) ) <ε+δimplies thatα(x,y)ψ ( d ( f x , f y ) ) <ε
(3.1)

for all x,yX. Then f is called an α-ψ-Meir-Keeler contractive mapping.

Now, we modify Definition 3.2 as follows.

Definition 3.3 Let (X,d) be a metric space, and let ψΨ. Suppose that f:XX is a triangular α-admissible mapping satisfying the following condition:

for each ε>0 there exists δ>0 such that

εψ ( d ( x , y ) ) <ε+δimplies thatψ ( d ( f x , f y ) ) <ε
(3.2)

for all x,yX with α(x,y)1. Then f is called a modified α-ψ-Meir-Keeler contractive mapping.

Remark 3.1 Let f be a modified α-ψ-Meir-Keeler contractive mapping. Then

ψ ( d ( f x , f y ) ) <ψ ( d ( x , y ) )

for all x,yX when xy and α(x,y)1. Also, if x=y and α(x,y)1, then d(fx,fy)=0, i.e.,

ψ ( d ( f x , f y ) ) ψ ( d ( x , y ) ) .

Theorem 3.1 Let (X,d) be a complete metric space. Suppose that f is a continuous modified α-ψ-Meir-Keeler contractive mapping, and that there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Proof Let x 0 X and define a sequence { x n } by x n = f n x 0 for all nN. If x n 0 = x n 0 + 1 for some n 0 N{0}, then, obviously, f has a fixed point. Hence, we suppose that

x n x n + 1
(3.3)

for all nN{0}. We have d( x n , x n + 1 )>0 for all nN{0}. Now, define s n =ψ(d( x n , x n + 1 )). By Remark 3.1, we deduce that for all nN{0} ψ(d( x n + 1 , x n + 2 ))=ψ(d(f x n ,f x n + 1 ))<ψ(d( x n , x n + 1 )). By applying Lemma 3.1 for

α( x m , x n )1for all m,nN with m<n,

we have

ψ ( d ( x n + 1 , x n + 2 ) ) <ψ ( d ( x n , x n + 1 ) ) .

Hence, the sequence { s n } is decreasing in R + , and so, it is convergent to s R + . We will show that s=0. Suppose, to the contrary, that s>0. Note that

0<s<ψ ( d ( x n , x n + 1 ) ) for all nN{0}.
(3.4)

Let ε=s>0. Then by hypothesis, there exists a δ(ε)>0 such that (3.2) holds. On the other hand, by the definition of ε, there exists n 0 N such that

ε< s n 0 =ψ ( d ( x n 0 , x n 0 + 1 ) ) <ε+δ.

Now by (3.2), we have

s n 0 + 1 =ψ ( d ( x n 0 + 1 , x n 0 + 2 ) ) ψ ( d ( f x n 0 , f x n 0 + 1 ) ) <ε=s,

which is a contradiction. Hence s=0, that is, lim n + s n =0. Now, by the continuity of ψ at t=0, we have lim n + d( x n , x n + 1 )=0. For given ε>0, by the hypothesis, there exists a δ=δ(ε)>0 such that (3.2) holds. Without loss of generality, we assume that δ<ε. Since s=0, then there exists NN such that

s n 1 =ψ ( d ( x n 1 , x n ) ) <δfor all nN.
(3.5)

We will prove that for any fixed k N 0 ,

ψ ( d ( x k , x k + l ) ) εfor all lN,
(3.6)

holds. Note that (3.6) holds for l=1 by (3.5). Suppose that condition (3.2) is satisfied for some mN. For l=m+1, by (3.5), we get

ψ ( d ( x k 1 , x k + m ) ) ψ ( d ( x k 1 , x k ) + d ( x k , x k + m ) ) ψ ( d ( x k 1 , x k ) ) + ψ ( d ( x k , x k + m ) ) < ε + δ .
(3.7)

If ψ(d( x k 1 , x k + m ))ε, then by (3.2), we get

ψ ( d ( x k , x k + m + 1 ) ) =ψ ( d ( f x k 1 , f x k + m ) ) <ε,

and hence (3.6) holds.

If ψ(d( x k 1 , x k + m ))<ε, by Remark 3.1, we get

ψ ( d ( x k , x k + m + 1 ) ) ψ ( d ( x k 1 , x k + m ) ) <ε.

Consequently, (3.6) holds for l=m+1. Hence, ψ(d( x k , x k + l ))ε for all k N 0 and l1, which means

d( x n , x m )<εfor all mn N 0 .
(3.8)

Hence { x n } is a Cauchy sequence. Since (X,d) is complete, there exists zX such that x n z as n. Now, since f is continuous, then

fz=f ( lim n x n ) = lim n x n + 1 =z,

that is, f has a fixed point. □

Corollary 3.1 (Theorem 10 of [11])

Let (X,d) be a complete metric space. Suppose that f is a continuous α-ψ-Meir-Keeler contractive mapping, and that there exists x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Proof Let εψ(d(x,y))<ε+δ, where α(x,y)1. Then by εψ(d(x,y))<ε+δ and Definition 3.2, we deduce that α(x,y)ψ(d(fx,fy))<ε. On the other hand, since α(x,y)1, then we have

ψ ( d ( f x , f y ) ) α(x,y)ψ ( d ( f x , f y ) ) <ε.

That is, conditions of Theorem 3.1 hold, and f has a fixed point. □

Theorem 3.2 Let (X,d) be a complete metric space, and let f be a modified α-ψ-Meir-Keeler contractive mapping. If 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 n.

Then f has a fixed point.

Proof Following the proof of Theorem 3.1, we say that α( x n , x n + 1 )1 for all nN{0}, and that there exist zX such that x n z as n+. Hence, from (ii) α( x n ,z)1. By Remark 3.1, we have

ψ ( d ( f z , z ) ) ψ ( d ( f z , f x n ) + d ( f x n , z ) ) ψ ( d ( f z , f x n ) ) + ψ ( d ( f x n , z ) ) ψ ( d ( z , x n ) ) + ψ ( d ( x n + 1 , z ) ) .

By taking limit as n+, in the inequality above, we get ψ(d(fz,z))0, that is, d(fz,z)=0. Hence fz=z. □

Corollary 3.2 (Theorem 11 of [11])

Let (X,d) be a complete metric space, and let f be a α-ψ-Meir-Keeler contractive mapping. If 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 n.

Then f has a fixed point.

Example 3.1 Let X=[0,), and let d(x,y)=|xy| be a metric on X. Define f:XX by

fx={ x 5 if  x [ 0 , 1 ] , x x 2 + 1 x ( 1 , ) , andα(x,y)={ 10 if  x , y [ 0 , 1 ] , 2 otherwise ,

and ψ(t)= 1 4 t. Clearly, (X,d) is a complete metric space. We show that f is a triangular α-admissible mapping. Let x,yX, if α(x,y)1, then x,y[0,1]. On the other hand, for all x,y[0,1], we have fx1 and fy1. It follows that α(fx,fy)1. Also, if α(x,z)1 and α(z,y)1, then x,y,z[0,1], and hence, α(x,y)1. Thus the assertion holds by the same arguments. Notice that 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+, then { x n }[0,1], and hence x[0,1]. This implies that α( x n ,x)1 for all nN{0}. Let α(x,y)1, then x,y[0,1]. Without loss of generality, take xy. Then

ψ ( d ( f x , f y ) ) = y 20 x 20 , ψ ( d ( x , y ) ) = y 4 x 4 .

Clearly, by taking δ=4ε, the condition (3.2) holds. Hence, conditions of Theorem 3.2 hold, and f has a fixed point. But if x,y[0,1] and

εd(x,y)<ε+δ,

where ε>0 and δ>0. Then

α(x,y)d(fx,fy)=2|xy|=2d(x,y)2ε.

That is, Corollary 3.2 (Theorem 11 of [11]) cannot be applied for this example.

Denote with Ψ st the family of strictly nondecreasing functions ψ st :[0,+)[0,+) continuous at t=0 such that

  • ψ st (t)=0 if and only if t=0,

  • ψ st (t+s) ψ st (t)+ ψ st (s).

Definition 3.4 [11]

Let (X,d) be a metric space, and let ψ st Ψ st . Suppose that f:XX is a triangular α-admissible mapping satisfying the following condition:

for each ε>0, there exists δ>0 such that

ε ψ st ( M ( x , y ) ) <ε+δimplies thatα(x,y) ψ st ( d ( f x , f y ) ) <ε
(3.9)

for all x,yX, where

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

Then f is called a generalized α- ψ st -Meir-Keeler contractive mapping.

Definition 3.5 Let (X,d) be a metric space, and let ψ st Ψ st . Suppose that f:XX is a triangular α-admissible mapping satisfying the following condition:

for each ε>0 there exists δ>0 such that

ε ψ st ( M ( x , y ) ) <ε+δimplies that ψ st ( d ( f x , f y ) ) <ε
(3.10)

for all x,yX, where α(x,y)1 and

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

Then f is called a modified generalized α- ψ st -Meir-Keeler contractive mapping.

Remark 3.2 Let f be a modified generalized α- ψ st -Meir-Keeler contractive mapping. Then

ψ st ( d ( f x , f y ) ) < ψ st ( M ( x , y ) )

for all x,yX, where α(x,y)1 when M(x,y)>0. Also, if M(x,y)=0 and α(x,y)1, then x=y, which implies that ψ(d(fx,fy))=0, i.e.,

ψ st ( d ( f x , f y ) ) ψ st ( M ( x , y ) ) .

Proposition 3.1 Let (X,d) be a metric space, and let f:XX be a modified generalized α- ψ st -Meir-Keeler contractive mapping. If there exists x 0 X such that α( x 0 ,f x 0 )1, then lim n d( f n + 1 x 0 , f n x 0 )=0.

Proof Define a sequence { x n } by x n = f n x 0 for all nN. If x n 0 = x n 0 + 1 for some n 0 N{0}, then, obviously, the conclusion holds. Hence, we suppose that

x n x n + 1
(3.11)

for all nN{0}. Then we have M( x n + 1 , x n )>0 for every n0. Then by Lemma 3.1 and Remark 3.2, we have

ψ st ( d ( x n + 1 , x n + 2 ) ) = ψ st ( d ( f x n , f x n + 1 ) ) < ψ st ( M ( x n , x n + 1 ) ) = ψ st ( max { d ( x n , x n + 1 ) , d ( f x n , x n ) , d ( f x n + 1 , x n + 1 ) , 1 2 [ d ( f x n , x n + 1 ) + d ( x n , f x n + 1 ) ] } ) ψ st ( max { d ( x n , x n + 1 ) , d ( x n + 1 , x n + 2 ) } ) .

Now, since ψ st is strictly nondecreasing, then we get

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

Hence the case, where

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

is not possible. Therefore, we deduce that

d( x n + 2 , x n + 1 )<d( x n + 1 , x n )
(3.12)

for all n. That is, { d ( x n + 1 , x n ) } n = 0 is a decreasing sequence in R + , and it converges to ε R + , that is,

lim n ψ st ( d ( x n + 1 , x n ) ) = lim n ψ st ( M ( x n + 1 , x n ) ) = ψ st (ε).
(3.13)

Notice that ε=inf{d( x n , x n + 1 ):nN}. Let us prove that ε=0. Suppose, to the contrary, that ε>0. Then ψ(ε)>0. Considering (3.13) together with the assumption that f is a generalized α- ψ st -Meir-Keeler contractive mapping, for ψ st (ε), there exists δ>0 and a natural number m such that

ψ st (ε) ψ st ( M ( x m , x m + 1 ) ) < ψ st (ε)+δ

implies that

ψ st ( d ( x m + 1 , x m + 2 ) ) = ψ st ( d ( f x m , f x m + 1 ) ) < ψ st (ε).

Now, since ψ st is strictly nondecreasing, then we get

d( x m + 2 , x m + 1 )<ε,

which is a contradiction, since ε=inf{d( x n , x n + 1 ):nN}. Then ε=0, and so,

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

 □

Theorem 3.3 Let (X,d) be a complete metric space, and let f:XX be an orbitally continuous modified generalized α- ψ st -Meir-Keeler contractive mapping. If there exist x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Proof Define x n + 1 = f n + 1 x 0 for all n0. We want to prove that lim m , n d( x n , x m )=0. If this is not so, then there exist ε>0 and a subsequence { x n ( i ) } of { x n } such that

d( x n ( i ) , x n ( i + 1 ) )>2ε.
(3.14)

For this ε>0, there exists δ>0 such that ε ψ st (M(x,y))<ε+δ implies that α(x,y) ψ st (d(fx,fy))<ε. Put r=min{ε,δ} and s n =d( x n , x n + 1 ) for all n1. From Proposition 3.1, there exists n 0 such that

s n =d( x n , x n + 1 )< r 4
(3.15)

for all n n 0 . Let n(i)> n 0 . We get n(i)n(i+1)1. If d( x n ( i ) , x n ( i + 1 ) 1 )ε+ r 2 , then

d ( x n ( i ) , x n ( i + 1 ) ) d ( x n ( i ) , x n ( i + 1 ) 1 ) + d ( x n ( i + 1 ) 1 , x n ( i + 1 ) ) d ( x n ( i ) , x n ( i + 1 ) 1 ) + d ( x n ( i + 1 ) 1 , x n ( i + 1 ) ) < ε + r 2 + s n ( i + 1 ) 1 < ε + 3 r 4 < 2 ε ,

which contradicts the assumption (3.14). Therefore, there are values of k such that n(i)kn(i+1) and d( x n ( i ) , x k )>ε+ r 2 . Now if d( x n ( i ) , x n ( i ) + 1 )ε+ r 2 , then

s n ( i ) =d( x n ( i ) , x n ( i ) + 1 )ε+ r 2 >r+ r 2 > r 4 ,

which is a contradiction to (3.15). Hence, there are values of k with n(i)kn(i+1) such that d( x n ( i ) , x k )<ε+ r 2 . Choose the smallest integer k with kn(i) such that d( x n ( i ) , x k )ε+ r 2 . Thus, d( x n ( i ) , x k 1 )<ε+ r 2 , and so,

d ( x n ( i ) , x k ) d ( x n ( i ) , x k 1 ) + d ( x k 1 , x k ) d ( x n ( i ) , x k 1 ) + d ( x k 1 , x k ) < ε + r 2 + r 4 = ε + 3 r 4 .

Now, we can choose a natural number k satisfying n(i)kn(i+1) such that

ε+ r 2 d( x n ( i ) , x k )<ε+ 3 r 4 .
(3.16)

Therefore, we obtain

d( x n ( i ) , x k )<ε+ 3 r 4 <ε+r,
(3.17)
d( x n ( i ) , x n ( i ) + 1 )= d n ( i ) < r 4 <ε+r,
(3.18)

and

d( x k , x k + 1 )= d k < r 4 <ε+r.
(3.19)

Thus, we have

1 2 [ d ( x n ( i ) , x k + 1 ) + d ( x n ( i ) + 1 , x k ) ] 1 2 [ d ( x n ( i ) , x k ) + d ( x k , x k + 1 ) + d ( x n ( i ) + 1 , x n ( i ) ) + d ( x n ( i ) , x k ) ] 1 2 [ d ( x n ( i ) , x k ) + d ( x k , x k + 1 ) + d ( x n ( i ) + 1 , x n ( i ) ) + d ( x n ( i ) , x k ) ] = d ( x n ( i ) , x k ) + 1 2 [ s k + s n ( i ) ] < ε + 3 r 4 + 1 2 [ r 4 + r 4 ] = ε + r .
(3.20)

Now, inequalities (3.17)-(3.20) imply that M( x n ( i ) , x k )<ε+rε+δ, and so, ψ st (M( x n ( i ) , x k ))< ψ st (ε+δ) ψ st (ε)+ ψ st (δ); the fact that f is a modified generalized α- ψ st -Meir-Keeler contractive mapping yields that

ψ st ( d ( x n ( i ) + 1 , x k + 1 ) ) < ψ st (ε).

Then d( x n ( i ) + 1 , x k + 1 )<ε. We deduce

d ( f n ( i ) x 0 , f k x 0 ) d ( f n ( i ) x 0 , f n ( i ) + 1 x 0 ) + d ( f n ( i ) + 1 x 0 , f k x 0 ) d ( f n ( i ) x 0 , f n ( i ) + 1 x 0 ) + d ( f n ( i ) + 1 x 0 , f k x 0 ) d ( f n ( i ) x 0 , f n ( i ) + 1 x 0 ) + d ( f n ( i ) + 1 x 0 , f k + 1 x 0 ) + d ( f k + 1 x 0 , f k x 0 ) .

From (3.16), (3.18) and (3.19), we obtain

d ( x n ( i ) + 1 , x k + 1 ) d ( x n ( i ) , x k ) d ( x n ( i ) , x n ( i ) + 1 ) d ( x k , x k + 1 ) > ε + r 2 r 4 r 4 = ε ,

which is a contradiction. We obtained that lim m , n d( x n , x m )=0, and so, { x n = f n x 0 } is a Cauchy sequence. Since X is complete, then there exists zX such that f n x 0 z as n. As f is orbitally continuous, so z=fz. □

Corollary 3.3 (Theorem 17 of [11])

Let (X,d) be a complete metric space, and let f:XX be an orbitally continuous generalized α- ψ st -Meir-Keeler contractive mapping. If there exist x 0 X such that α( x 0 ,f x 0 )1, then f has a fixed point.

Example 3.2 Let X=[0,), and let d(x,y)=|xy| be a metric on X. Define f:XX by

fx={ x 7 if  x [ 0 , 1 ] , x x 3 + 6 x ( 1 , )

and ψ st (t)= 1 2 t,

α(x,y)={ 28 if  x , y [ 0 , 1 ] , 8 otherwise .

Clearly, f is a triangular α-admissible mapping, and it is orbitally continuous. Let α(x,y)1, then x,y[0,1]. Without loss of generality, take xy. Then

ψ st ( d ( f x , f y ) ) = y 14 x 14 , ψ st ( M ( x , y ) ) = ψ st ( max { y x , 6 7 y , x y 7 , y x 7 } ) ψ st ( M ( x , y ) ) = max { y 2 x 2 , 6 14 y , x 2 y 14 , y 2 x 14 } .

Clearly, by taking δ=6ε, the condition (3.10) holds. Hence, all conditions of Theorem 3.3 are satisfied, and f has a fixed point. But if x=0 and y=1

εM(0,1)<δ+ε

for δ>0 and ε>0, then

ε1<δ+ε,

and so,

α(0,1) ψ st ( d ( f 0 , f 1 ) ) =21ε.

That is, Corollary 3.3 (Theorem 17 of [11]) cannot be applied for this example.

4 Modified α-η-contractive multifunction

Recently, Asl et al. [12] introduced the following notion.

Definition 4.1 Let T:X 2 X , and let α:X×X R + . We say that T is an α -admissible mapping if

α(x,y)1implies that α (Tx,Ty)1,x,yX,

where

α (A,B)= inf x A , y B α(x,y).

We generalize this concept as follows.

Definition 4.2 Let T:X 2 X be a multifunction, and let α,η:X×X R + be two functions, where η is bounded. We say that T is an α -admissible mapping with respect to η if

α(x,y)η(x,y)implies that α (Tx,Ty) η (Tx,Ty),x,yX,

where

α (A,B)= inf x A , y B α(x,y)and η (A,B)= sup x A , y B η(x,y).

If we take η(x,y)=1 for all x,yX, then this definition reduces to Definition 4.1. In case α(x,y)=1 for all x,yX, then T is called an η -subadmissible mapping.

Notice that Ψ is the family of 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.

As an application of our new concept, we develop now a fixed point result for a multifunction, which generalizes Theorem 1.1.

Theorem 4.1 Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible, with respect to η, and closed-valued multifunction on X. Assume that for ψΨ,

x,yX, α (Tx,Ty) η (Tx,Ty)H(Tx,Ty)ψ ( d ( x , y ) ) .
(4.1)

Also, suppose that the following assertions hold:

  1. (i)

    there exist x 0 X and x 1 T x 0 such that α( x 0 , x 1 )η( x 0 , x 1 );

  2. (ii)

    for a sequence { x n }X converging to xX and α( x n , x n + 1 )η( x n , x n + 1 ) for all nN, we have α( x n ,x)η( x n ,x) for all nN.

Then T has a fixed point.

Proof Let x 1 T x 0 be such that α( x 0 , x 1 )η( x 0 , x 1 ). Since T is an α -admissible mapping, then α (T x 0 ,T x 1 ) η (T x 0 ,T x 1 ). Therefore, from (4.1), we have

H(T x 0 ,T x 1 )ψ ( d ( x 0 , x 1 ) ) .
(4.2)

If x 0 = x 1 , then x 0 is a fixed point of T. Hence, we assume that x 0 x 1 . Also, if x 1 T x 1 , then x 1 is a fixed point of T. Assume that x 1 T x 1 and q>1. Then we have

0<d( x 1 ,T x 1 )H(T x 0 ,T x 1 )<qH(T x 0 ,T x 1 ),

and so, by (4.2), we get

0<d( x 1 ,T x 1 )<qH(T x 0 ,T x 1 )qψ ( d ( x 0 , x 1 ) ) .

This implies that there exists x 2 T x 1 such that

0<d( x 1 , x 2 )<qH(T x 0 ,T x 1 )qψ ( d ( x 0 , x 1 ) ) .
(4.3)

Note that x 1 x 2 (since x 1 T x 1 ). Also, since α (T x 0 ,T x 1 ) η (T x 0 ,T x 1 ), x 1 T x 0 and x 2 T x 1 , then α( x 1 , x 2 )η( x 1 , x 2 ). So α (T x 1 ,T x 2 ) η (T x 1 ,T x 2 ). Therefore, from (4.1), we have

H(T x 1 ,T x 2 )ψ ( d ( x 1 , x 2 ) ) .
(4.4)

Put t 0 =d( x 0 , x 1 ). Then from (4.3), we have d( x 1 , x 2 )<qψ( t 0 ), where t 0 >0. Now, since ψ is strictly increasing, then ψ(d( x 1 , x 2 ))<ψ(qψ( t 0 )). Put

q 1 = ψ ( q ψ ( t 0 ) ) ψ ( d ( x 1 , x 2 ) ) ,

and so q 1 >1. If x 2 T x 2 , then x 2 is a fixed point of T. Hence, we suppose that x 2 T x 2 . Then

0<d( x 2 ,T x 2 )H(T x 1 ,T x 2 )< q 1 H(T x 1 ,T x 2 ).

So there exists x 3 T x 2 such that

0<d( x 2 , x 3 )< q 1 H(T x 1 ,T x 2 ),

and then from (4.4), we get

0<d( x 2 , x 3 )< q 1 H(T x 1 ,T x 2 ) q 1 ψ ( d ( x 1 , x 2 ) ) =ψ ( q ψ ( t 0 ) ) .

Again, since ψ is strictly increasing, then ψ(d( x 2 , x 3 ))<ψ(ψ(qψ( t 0 ))). Put

q 2 = ψ ( ψ ( q ψ ( t 0 ) ) ) ψ ( d ( x 2 , x 3 ) ) .

So, q 2 >1. If x 3 T x 3 , then x 3 is a fixed point of T. Hence, we assume that x 3 T x 3 . Then

0<d( x 3 ,T x 3 )H(T x 2 ,T x 3 )< q 2 H(T x 2 ,T x 3 ),

and so, there exists x 4 T x 3 such that

0<d( x 3 , x 4 )H(T x 2 ,T x 3 )< q 2 H(T x 2 ,T x 3 ).
(4.5)

Clearly, x 2 x 3 . Also again, since α (T x 1 ,T x 2 ) η (T x 1 ,T x 2 ), x 2 T x 1 and x 3 T x 2 , then α( x 2 , x 3 )η( x 2 , x 3 ), and so, α (T x 2 ,T x 3 ) η (T x 2 ,T x 3 ). Then from (4.1), we have

H(T x 2 ,T x 3 )ψ ( d ( x 2 , x 3 ) ) ,

and so, from (4.5), we deduce that

d( x 3 , x 4 )< q 2 H(T x 2 ,T x 3 ) q 2 ψ ( d ( x 2 , x 3 ) ) =ψ ( ψ ( q ψ ( t 0 ) ) ) .

By continuing this process, we obtain a sequence { x n } in X such that x n T x n 1 , x n x n 1 , α ( x n , x n + 1 ) η ( x n , x n + 1 ) and d( x n , x n + 1 ) ψ n 1 (qψ( t 0 )) for all nN. Now, for all m>n, we can write

d( x n , x m ) k = n m 1 d( x k , x k + 1 ) k = n m 1 ψ k 1 ( q ψ ( t 0 ) ) .

Therefore, { x n } is a Cauchy sequence. Since (X,d) is a complete metric space, then there exists zX such that x n z as n. Now, since α( x n ,z)η( x n ,z) for all nN, then α (T x n ,Tz) η (T x n ,Tz), and so, from (4.1), we have

d(z,Tz)H(T x n ,Tz)+d( x n ,z)ψ ( d ( x n , z ) ) +d( x n ,z)

for all nN. Taking limit as n in the inequality above, we get d(z,Tz)=0, i.e., zTz. □

If in Theorem 4.1 we take η(x,y)=1, we have the following corollary.

Corollary 4.1 Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible and closed-valued multifunction on X. Assume that

x,yX, α (Tx,Ty)1H(Tx,Ty)ψ ( d ( x , y ) ) .

Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and α( x n , x n + 1 )1 for all nN, we have α( x n ,x)1 for all nN.

Then T has a fixed point.

If in Theorem 4.1 we take α(x,y)=1, then we have the following result.

Corollary 4.2 Let (X,d) be a complete metric space, and let T:X 2 X be an η -subadmissible and closed-valued multifunction on X. Assume that

x,yX, η (Tx,Ty)1H(Tx,Ty)ψ ( d ( x , y ) ) .

Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and η( x n , x n + 1 )1 for all nN, we have η( x n ,x)1 for all nN.

Then T has a fixed point.

Corollary 4.3 (Theorem 2.1 and 2.3 of [12])

Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible and closed-valued multifunction on X. Assume that

α (Tx,Ty)H(Tx,Ty)ψ ( d ( x , y ) )
(4.6)

for all x,yX. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and α( x n , x n + 1 )1 for all nN, we have α( x n ,x)1 for all nN.

Then T has a fixed point.

Proof Suppose that α (Tx,Ty)1 for x,yX. Then by (4.6), we have

H(Tx,Ty)ψ ( d ( x , y ) ) .

That is, conditions of Corollary 4.1 hold, and T has a fixed point. □

Similarly, we can deduce the following corollaries.

Corollary 4.4 Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible and closed-valued multifunction on X. Assume that

( α ( T x , T y ) + 1 ) H ( T x , T y ) 2 ψ ( d ( x , y ) )

for all x,yX. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and α( x n , x n + 1 )1 for all nN, we have α( x n ,x)1 for all nN.

Then T has a fixed point.

Corollary 4.5 Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible and closed-valued multifunction on X. Assume that

( H ( T x , T y ) + ) α ( T x , T y ) ψ ( d ( x , y ) ) +

for all x,yX, where >0. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and α( x n , x n + 1 )1 for all nN, we have α( x n ,x)1 for all nN.

Then T has a fixed point.

Corollary 4.6 Let (X,d) be a complete metric space, and let T:X 2 X be an η -subadmissible and closed-valued multifunction on X. Assume that

H(Tx,Ty) η (Tx,Ty)ψ ( d ( x , y ) )

for all x,yX. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and η( x n , x n + 1 )1 for all nN, we have η( x n ,x)1 for all nN.

Then T has a fixed point.

Corollary 4.7 Let (X,d) be a complete metric space, and let T:X 2 X be an η -subadmissible and closed-valued multifunction on X. Assume that

2 H ( T x , T y ) ( η ( T x , T y ) + 1 ) ψ ( d ( x , y ) )

for all x,yX. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and η( x n , x n + 1 )1 for all nN, we have η( x n ,x)1 for all nN.

Then T has a fixed point.

Corollary 4.8 Let (X,d) be a complete metric space, and let T:X 2 X be an α -admissible and closed-valued multifunction on X. Assume that

H(Tx,Ty)+ ( ψ ( d ( x , y ) ) + ) η ( T x , T y )

for all x,yX, where >0. Also, suppose that the following assertions hold:

  1. (i)

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

  2. (ii)

    for a sequence { x n }X converging to xX and η( x n , x n + 1 )1 for all nN, we have η( x n ,x)1 for all nN.

Then T has a fixed point.