arXiv:1512.03967v1 [math.CA] 12 Dec 2015

New fixed point theorems for set-valued contractions in b-metric spaces Radu MICULESCU and Alexandru MIHAIL Abstract. In this paper we indicate a way to generalize a series of fixed point results in the framework of b-metric spaces and we exemplify it by extending Nadler’s contraction principle for set-valued functions (see Multi-valued contraction mappings, Pac. J. Math., 30 (1969), 475-488) and a fixed point theorem for set-valued quasi-contractions functions due to H. Aydi, M.F. Bota, E. Karapinar and S. Mitrovi´c (see A fixed point theorem for set-valued quasi-contractions in b-metric spaces, Fixed Point Theory Appl. 2012, 2012:88).

2010 Mathematics Subject Classification: 54H25, 47H10 Key words and phrases: fixed point theorems, set-valued functions, bmetric spaces

1. Introduction In the last decades one can observe a remarkable amount of interest for the development of fixed point theory since it has a huge number of applications. Among the generalizations of the Banach-Caccioppoli-Picard principle one of the central results of the above mentioned theory, known also as the contraction principle - a central role is played by the following two: - the one due to S.B. Nadler [21] who extended the contraction principle to set-valued functions and generated in this way many applications in control theory, convex optimization etc; - the one due to I. A. Bakhtin [5] and S. Czerwik [13], [14] who, motivated by the problem of the convergence of measurable functions with respect to measure, introduced b-metric spaces (a generalization of metric spaces) and proved the contraction principle in this framework. In the last period many mathematicians obtained fixed point results for single-valued or set-valued functions, in the setting of b-metric spaces (see, for example, [1], [8], [9], [10], [17], [23], [24], [29], [30] and the references therein). In this paper we indicate a way (see Lemma 2.2) to generalize a series of fixed point results in the framework of b-metric spaces and we exemplify it by extending Nadler’s contraction principle for set-valued functions (see [21]) 1

and a fixed point theorem for set-valued quasi-contractions functions due to H. Aydi, M.F. Bota, E. Karapinar and S. Mitrovi´c (see [4]). 2. Preliminaries results In this section we sum up some basic facts that we are going to use later. Definition 2.1. Given a nonempty set X and a real number s ∈ [1, ∞), a function d : X × X → [0, ∞) is called b-metric if it satisfies the following properties: i) d(x, y) = 0 if and only if x = y; ii) d(x, y) = d(y, x) for all x, y ∈ X; iii) d(x, y) ≤ s(d(x, z) + d((z, y)) for all x, y, z ∈ X. The pair (X, d) is called b-metric space. Remark 2.1. As when s = 1, a b-metric space is a metric space, we infer that the family of b-metric spaces is larger than the one of metric spaces. In other words, every metric spaces is a b-metric space. Note that Czerwik proved that the converse need not be true (see also [4], [12], [18], [22] and [27]), so the family of b-metric spaces is effectively larger than the one of metric spaces. Definition 2.2. A sequence (xn )n∈N of elements from a b-metric space (X, d) is called: - convergent if there exists l ∈ R such that lim d(xn , l) = 0; n→∞

- Cauchy if

lim d(xm , xn ) = 0, i.e. for every ε > 0 there exists nε ∈ N

m,n→∞

such d(xm , xn ) < ε that for all m, n ∈ N, m, n ≥ nε . The b-metric space (X, d) is called complete if every Cauchy sequence of elements from (X, d) is convergent. Beside the classical spaces lp (R) and Lp [0, 1], where p ∈ (0, 1), one can find examples of b-metric spaces in [4], [6], [10], [13] and [14]. Remark 2.2. As in the case of metric spaces, a b-metric space can be endowed with the topology induced by its convergence and almost all the concepts and results which are valid for metric spaces can be extended to the framework of b-metric spaces. T.V. An, L.Q. Tuyen and N.V. Dung [3] proved that every b-metric space is a semi-metrizable space (i.e. there exists a function d : X × X → [0, ∞) 2

such that: i) d(x, y) = 0 if and only if x = y; ii) d(x, y) = d(y, x) for all x, y ∈ X; iii) x ∈ A if and only if d(x, A) = inf{d(x, y) | y ∈ A} = 0 for every x ∈ X and every A ⊆ X). Consequently many properties of bmetric spaces are obvious. In addition they provided a sufficient condition for a b-metric space to be metrizable and gave an example showing that, in the framework of a b-metric space (X, d), there exists an open ball (i.e. a set of the form {y ∈ X | d(x, y) < r}, where r > 0) which is not open. In a metric space (X, d), the functions d is continuous (i.e. lim d(xn , yn ) = n→∞

d(x, y) for all sequences (xn )n∈N and (yn )n∈N of elements from X and x, y ∈ X such that lim xn = x and lim yn = y). The fact that this property is n→∞

n→∞ 1 d(x, y) s2

not valid for b-metric spaces (as 2

s d(x, y) and

1 d(x, y) s

≤ lim d(xn , yn ) ≤ lim d(xn , yn ) ≤ n→∞

n→∞

≤ lim d(xn , y) ≤ lim d(xn , y) ≤ sd(x, y), see [20], [22] n→∞

n→∞

and [25]) is a motivation of our Definition 3.2. In the sequel, given a b-metric space (X, d): - by B(X) we denote the set of non-empty bounded closed subsets of X - for A, B ∈ B(X), we define the Hausdorff-Pompeiu distance between A and B by h(A, B) = max{supd(a, B), supd(b, A)}, where d(x, C) = inf d(x, c) a∈A

c∈C

b∈B

for every x ∈ X and every C ∈ B(X) - for c, d ∈ [0, 1] and x, y ∈ X, we shall use the following notation: d Nc,d (x, y) = max{d(x, y), cd(x, T (x)), cd(y, T (y)), (d(x, T (y))+d(y, T (x)))}. 2 Lemma 2.1. For every sequence (xn )n∈N of elements from a b-metric space (X, d), the inequality d(x0 , xk ) ≤ s

n

k−1 X

d(xi , xi+1 )

i=0

is valid for every n ∈ N and every k ∈ {1, 2, 3, ..., 2n − 1, 2n }. Proof. We are going to use the method of mathematical induction. Dek P noting by P (n) the statement: d(x0 , xk ) ≤ sn d(xi , xi+1 ) for every n ∈ N i=0

3

and every k ∈ {1, 2, 3, ..., 2n − 1, 2n }, as the statements P (0) and P (1) are obvious, it remains to prove that P (n) ⇒ P (n + 1). Indeed, the above mentioned implication is true since, on the one hand, for every k ∈ {1, 2, 3, ..., 2n − 1, 2n }, we have k k X X n+1 d(x0 , xk ) ≤ s d(xi , xi+1 ) ≤ s d(xi , xi+1 ). P (n)

n

i=0

i=0

On the other hand, for every k ∈ {2n + 1, 2n + 2, ..., 2n+1 − 1, 2n+1}, we have P (n)

d(x0 , xk ) ≤ s(d(x0 , x2n ) + d(x2n , xk )) ≤ ≤ s(s

n

n −1 2X

d(xi , xi+1 ) + s

n

k−1 X

d(xi , xi+1 )) = s

n+1

i=2n

i=0

k−1 X d(xi , xi+1 ).  i=0

Lemma 2.2. Every sequence (xn )n∈N of elements from a b-metric space (X, d), having the property that there exists γ ∈ [0, 1) such that d(xn+1 , xn ) ≤ γd(xn , xn−1 ), for every n ∈ N, is Cauchy. Proof. First let us note that d(xn+1 , xn ) ≤ γ n d(x1 , x0 ),

(1)

for every n ∈ N. For all m, k ∈ N, with the notation p = [log2 k], we have d(xm+1 , xm+k ) ≤ sd(xm+1 , xm+2 ) + sd(xm+2 , xm+k ) ≤ ≤ sd(xm+1 , xm+2 ) + s2 d(xm+2 , xm+22 ) + s2 d(xm+22 , xm+k ) ≤ ≤ sd(xm+1 , xm+2 )+s2 d(xm+2 , xm+22 )+s3 d(xm+22 , xm+23 )+s3 d(xm+23 , xm+k ) ≤ ... ≤

p X

sn d(xm+2n−1 , xm+2n ) + sp+1 d(xm+2p , xm+k )

Lemma 1



n=1

X p



n=1

2n

s (

n−1 m+2 X −1

d(x2n−1 +i , x2n−1 +i+1 ))+s

i=m

2(p+1)

(

p m+k−2 X −1

d(x2p +i , x2p +i+1 )) ≤

i=m

4



p+1 X

2n

s (

n−1 m+2 X −1

n=1

(1)

d(x2n−1 +i , x2n−1 +i+1 )) ≤ d(x0 , x1 )

n=1

i=m

2n

s (

2n−1 X−1

n−1 +i

γ m+2

)≤

i=0

d(x0 , x1 )γ m X 2n 2n−1 d(x0 , x1 ) X 2n logγ s+2n−1 s γ = γm γ . 1 − γ n=1 1 − γ n=1 p+1



p+1 X

p+1

Let us note that since lim (2n logγ s + 2n−1 − n) = ∞, there exists n0 ∈ N n→∞

n−1

such that 2n logγ s+2n−1 −n ≥ M, i.e. γ 2n logγ s+2 ≤ γ M γ n for each n ∈ N, ∞ P n−1 n ≥ n0 , hence the series γ 2n logγ s+2 is convergent and denoting by S its n=1

sum, we come to the conclusion that d(xm+1 , xm+k ) ≤ γ m

d(x0 , x1 )S , 1−γ

for all m, k ∈ N. Consequently, as lim γ n = 0, we infer that (xn )n∈N is n→∞ Cauchy.  Theorem 2.1. Let (X, d) be a b-metric space and T : X → B(X) having the property that there exist c, d ∈ [0, 1] and α ∈ [0, 1) such that: i) αds < 1; ii) h(T (x), T (y)) ≤ αNc,d (x, y) for all x, y ∈ X. Then, for every x0 ∈ X, there exist γ ∈ [0, 1) and a sequence (xn )n∈N of elements from X such that: a) xn+1 ∈ T (xn ) for every n ∈ N; b) d(xn+1 , xn ) ≤ γd(xn , xn−1 ) for every n ∈ N; c) (xn )n∈N is Cauchy. dsβ 1 Proof. Let us consider β ∈ (α, min(1, ds )), γ = max{β, 2−dsβ } < 1, x0 ∈ X and x1 ∈ T (x0 ). If x1 = x0 , then the sequence (xn )n∈N given by xn = x0 for every n ∈ N satisfies a), b) and c). ii)

Since d(x1 , T (x1 )) ≤ d(T (x0 ), T (x1 )) ≤ h(T (x0 ), T (x1 )) ≤ αNc,d(x0 , x1 ) < βNc,d (x0 , x1 ), there exists x2 ∈ T (x1 ) such that d(x1 , x2 ) < βNc,d(x0 , x1 ). If x2 = x1 , then the sequence (xn )n∈N given by xn = x1 for every n ∈ N, n ≥ 1, satisfies a), b) and c). By repeating this procedure we obtain a sequence (xn )n∈N of elements from X such that xn+1 ∈ T (xn ) and 0 < d(xn , xn+1 ) < βNc,d(xn−1 , xn ) for every n ∈ N, n ≥ 1. 5

With the notation dn = d(xn , xn+1 ), we have 0 < dn < βNc,d (xn−1 , xn ) = d = β max{dn−1 , cd(xn−1 , T (xn−1 )), cd(xn , T (xn )), (d(xn−1 , T (xn ))+d(xn , T (xn−1 )))} ≤ 2 d ds ≤ β max{dn−1, cdn , cdn−1, d(xn−1 , xn+1 )} ≤ β max{dn−1 , cdn , cdn−1 , (dn−1 +dn )} ≤ 2 2 ds ≤ β max{dn−1, (dn−1 + dn )}, 2 for every n ∈ N, where the justification of the last inequality is the following: (dn−1 + dn )} = cdn , if, by reduction ad absurdum, max{dn−1, cdn , cdn−1, ds 2 then we get that 0 < dn < βcdn ≤ βdn , so we obtain the contradiction 1 < β. Consequently dn < βdn−1 or dn < β ds (dn−1 + dn ), i.e. dn < βdn−1 2 dsβ dsβ or dn < 2−dsβ dn−1 for every n ∈ N. Thus dn ≤ max{β, 2−dsβ }dn−1 , i.e. d(xn+1 , xn ) ≤ γd(xn , xn−1 ) for every n ∈ N. Hence the sequence (xn )n∈N satisfies a) and b). From Lemma 2.2 we deduce that it also satisfies c).  3. Main results In this section, making use of Theorem 2.1, we present three fixed point theorems for set-valued functions. Definition 3.1. A function T : X → B(X), where (X, d) is a b-metric space, is called continuous if for all sequences (xn )n∈N and (yn )n∈N of elements from X and x, y ∈ X such that lim xn = x, lim yn = y and n→∞

n→∞

yn ∈ T (xn ) for every n ∈ N, we have y ∈ T (x). Theorem 3.1. A function T : X → B(X), where (X, d) is a complete b-metric space, has a fixed point, provided that it satisfies the following three conditions: i) T is continuous; ii) there exist c, d ∈ [0, 1] and α ∈ [0, 1) such that h(T (x), T (y)) ≤ αNc,d (x, y) for all x, y ∈ X; iii) αds < 1.

6

Proof. Taking into account ii) and iii), by virtue of Theorem 2.1, there exists a Cauchy sequence (xn )n∈N of elements of X such that xn+1 ∈ T (xn ),

(1)

for every n ∈ N. As the b-metric space (X, d) is complete, there exists u ∈ X such that lim xn = u (so lim xn+1 = u). We combine i) with 1) to see that u ∈ T (u), n→∞ n→∞ i.e. u is a fixed point of T .  Definition 3.2. Given a b-metric space (X, d), the b-metric d is called ∗-continuous if for every A ∈ B(X), every x ∈ X and every sequence (xn )n∈N of elements from X such that lim xn = x, we have lim d(xn , A) = d(x, A). n→∞

n→∞

Our notion of ∗-continuity of d is stronger than the continuity of d in the first variable. Theorem 3.2. A function T : X → B(X), where (X, d) is a complete b-metric space, has a fixed point, provided that it satisfies the following three conditions: i) d is ∗-continuous; ii) there exist c, d ∈ [0, 1] and α ∈ [0, 1) such that h(T (x), T (y)) ≤ αNc,d (x, y) for all x, y ∈ X; iii) αds < 1. Proof. Based on ii) and iii), according to Theorem 2.1, there exists a Cauchy sequence (xn )n∈N of elements of X such that xn+1 ∈ T (xn ),

(1)

for every n ∈ N. As the b-metric space (X, d) is complete, there exists u ∈ X such that lim xn = u. n→∞ Then we have (1)

i)

d(xn+1 , T (u)) ≤ d(T (xn ), T (u)) ≤ h(T (xn ), T (u)) ≤ ≤ αNc,d(xn , u) = (1) d = α max{d(xn , u), cd(xn , T (xn )), cd(u, T (u)), (d(xn , T (u))+d(u, T (xn)))} ≤ 2

7

d ≤ α max{d(xn , u), cd(xn , xn+1 ), cd(u, T (u)), (s(d(xn , u)+d(u, T (u)))+d(u, xn+1))}, 2 (2) for every n ∈ N. Since lim d(xn , u) = lim d(xn , xn+1 ) = 0 and lim d(xn+1 , T (u)) = d(u, T (u)) n→∞

n→∞

n→∞

(as d is ∗-continuous and lim xn+1 = u), upon passing to limit, as n → ∞, n→∞

in 2), we get ds αds d(u, T (u))} ≤ max{αc, }d(u, T (u)). 2 2 (3) As max{αc, αds } < 1 (see iii)), from 3), we conclude that d(u, T (u)) = 0, 2 i.e. u ∈ T (u). Hence T has a fixed point.  d(u, T (u) ≤ α max{0, cd(u, T (u)),

Theorem 3.3. A function T : X → B(X), where (X, d) is a complete b-metric space, has a fixed point, provided that it satisfies the following two conditions: i) there exist c, d ∈ [0, 1] and α ∈ [0, 1) such that h(T (x), T (y)) ≤ αNc,d (x, y) for all x, y ∈ X; ii) max{αcs, αds} < 1. Proof. Making use of i) and ii), according to Theorem 2.1, there exists a Cauchy sequence (xn )n∈N of elements from X such that xn+1 ∈ T (xn ), for every n ∈ N. As the b-metric space (X, d) is complete, there exists u ∈ X such that lim xn = u. n→∞

First let us note that, as we have seen in 2) from the proof of Theorem 3.2, we have d(xn+1 , T (u)) ≤ d ≤ α max{d(xn , u), cd(xn, T (xn )), cd(u, T (u)), (d(xn , T (u))+d(u, T (xn)))} ≤ 2 d ≤ α max{d(xn , u), cd(xn , T (xn )), cd(u, T (u)), (d(xn , T (u)) + d(u, xn+1 ))} ≤ 2 d ≤ α max{d(xn , u), cd(xn , xn+1 ), cd(u, T (u)), (s(d(xn , u)+d(u, T (u)))+d(u, xn+1)))}, 2 (1) for every n ∈ N. We divide the discussion into two cases: A. d(u, T (u)) ≤ lim d(xn , T (u)); n→∞

8

and B. d(u, T (u)) > lim d(xn , T (u)). n→∞

In case A, there exists a subsequence (xnk )k∈N of (xn )n∈N having the property that lim d(xnk +1 , T (u)) ≥ d(u, T (u)), so for every ε > 0 there k→∞

exists kε ∈ N such that (1)

d(u, T (u)) − ε ≤ d(xnk +1 , T (u)) ≤

d ≤ α max{d(xnk , u), cd(xnk , xnk +1 ), cd(u, T (u)), (s(d(xnk , u)+d(u, T (u)))+d(u, xnk+1 )))}, 2 for every k ∈ N, k ≥ kε . By passing to limit as k → ∞ in the above inequality, we get that αsd sd }, d(u, T (u))−ε ≤ α max{0, 0, cd(u, T (u)), d(u, T (u))} = d(u, T (u)) max{αc, 2 2 for every ε > 0, so αsd d(u, T (u)) ≤ d(u, T (u)) max{αc, }. 2 } < 1, from the above inequality, we conclude that Since max{αc, αsd 2 d(u, T (u)) = 0, i.e. u ∈ T (u), so T has a fixed point. In case B, there exists n0 ∈ N such that d(xn , T (u)) ≤ d(u, T (u)),

(2)

for every n ∈ N, n ≥ n0 . Since d(u, T (u)) ≤ s(d(u, xn+1) + d(xn+1 , T (u))), i.e. d(u,Ts (u)) − d(u, xn+1) ≤ d(xn+1 , T (u)), we get that (1) d(u, T (u)) − d(u, xn+1) ≤ d(xn+1 , T (u)) ≤ s (2) d ≤ α max{d(xn , u), cd(xn , xn+1 ), cd(u, T (u)), (d(xn , T (u)) + d(u, xn+1)))} ≤ 2 d ≤ α max{d(xn , u), cd(xn, xn+1 ), cd(u, T (u)), (d(u, T (u)) + d(u, xn+1)))}, 2 for every n ∈ N, n ≥ n0 . By passing to limit as n → ∞ in the above inequality, we obtain that d d d(u, T (u)) ≤ αs max{0, 0, cd(u, T (u)), d(u, T (u))} = αs max{c, }d(u, T (u)). 2 2 d As αs max{c, 2 } < 1 (see ii)), we infer that d(u, T (u)) = 0, so u ∈ T (u), i.e. T has a fixed point. 

9

4. Remarks and comments I. Let us recall the following result (see Lemma 3.1 from [28]): Lemma 4.1. Every sequence (xn )n∈N of elements from a b-metric space (X, d) is Cauchy provided that: i) there exists γ ∈ [0, 1) such that d(xn+1 , xn ) ≤ γd(xn , xn−1 ), for every n ∈ N; ii) sγ < 1. Obviously our Lemma 2.2 is a generalization of the above Lemma which is the corner stone of the results from [16], [18], [19], [22] and [28]. II. The following definition is inspired by the definition of a multi-valued weakly Picard operator in the setting of a metric space from [7]. Definition. A function T : X → B(X), where (X, d) is a b-metric space, is called a multi-valued weakly Picard operator if for each x ∈ X and each y ∈ T (x) there exists a sequence (xn )n∈N such that: i) x0 = x and x1 = y; ii) xn+1 ∈ T (xn ) for every n ∈ N; iii) the sequence (xn )n∈N is convergent and its limit is a fixed point of T . Let us mention that Theorems 3.1., 3.2 and 3.3 provide sufficient conditions for a function T to be multi-valued weakly Picard operator. III. For c = d = 0 in Theorem 3.3 we obtain Theorem 5 from [21], i.e. Nadler’s contraction principle for set-valued functions. IV. Let us recall the following result (see Theorem 2.2 from [4]) which is a generalization of Theorem 1.2 from [2] which improves Theorem 3.3 from [15], Corollary 3.3. from [26], Corollary 4.3 from [28] and Theorem 1 from [11]: Theorem 4.1. A function T : X → B(X), where (X, d) is a complete b-metric space, has a fixed point, provided that it satisfies the following two conditions: 10

i) there exists α ∈ [0, 1) such that h(T (x), T (y)) ≤ α max{d(x, y), d(x, T (x)), d(y, T (y)), d(x, T (y)), d(y, T (x))}, for all x, y ∈ X; 1 ii) α ≤ s+s 2. Our Theorem 3.3 is a generalization of Theorem 4.1. Indeed, on the one hand, i) from Theorem 4.1 is a particular case of i) from Theorem 3.3 since 1 N1,1 (x, y) = max{d(x, y), d(x, T (x)), d(y, T (y)), (d(x, T (y)) + d(y, T (x)))} 2 and

1 (d(x, T (y)) + d(y, T (x))) ≤ max{d(x, T (y)), d(y, T (x))}, 2 for all x, y ∈ X. One the other hand, ii) from Theorem 4.1 is a particular 1 case of ii) from Theorem 3.3 since α ≤ s+s 2 ⇒ max{αcs, αds} < 1. Now let us present a situation when Theorem 3.3 is applicable, but Theorem 4.1 is not. We consider the b-metric space (R, d), where d(x, y) = (x − y)2 for all x, y ∈ R, for which s = 2 and the function f : R → B(R) given by f (x) = 9 { 10 x} for every x ∈ R. On the one hand, Theorem 3.3. is applicable taking 9 c = d = 0 and α = 10 . On the other hand Theorem 4.1 is not applicable 9 since i) implies 10 ≤ α and ii) implies α ≤ 61 . References [1] A. Aghajani, M. Abbas and J.R. Roshan, Common fixed point of generalized weak contractive mappings in partially ordered b-metric spaces, Math. Slovaca, 64 (2014), 941-960. [2] A. Amini-Harandi, Fixed point theory for set-valued quasi-contraction maps in metric spaces, Appl. Math. Lett., 24 (2011), 1791-1794. [3] T.V. An, L.Q. Tuyen and N.V. Dung, Stone-type theorem on b-metric spaces and applications, Topology Appl., 185/186 (2015), 50-64. [4] H. Aydi, M.F. Bota, E. Karapinar and S. Mitrovi´c, A fixed point theorem for set-valued quasi-contractions in b-metric spaces, Fixed Point Theory Appl., 2012, 2012:88. 11

[5] I. A. Bakhtin, The contraction mapping principle in quasimetric spaces, Funct. Anal., Unianowsk Gos. Ped. Inst., 30 (1989), 26-37. [6] V. Berinde, Generalized contractions in quasimetric spaces, Seminar on Fixed Point Theory, 1993, 3-9. [7] M. Berinde and V. Berinde, On a general class of multi-valued weakly Picard mappings, J. Math. Anal. Appl., 326 (2007), 772-782 [8] M. Boriceanu, M. Bota and A. Petru¸sel, Multivalued fractals in bmetric spaces, Cent. Eur. J. Math., 8 (2010), 367-377. [9] M. Boriceanu, A. Petru¸sel and A.I. Rus, Fixed point theorems for some multivalued generalized contraction in b-metric spaces, Int. J. Math. Stat., 6 (2010), 65-76. [10] M. Bota, A. Moln´ar and C. Varga, On Ekeland’s variational principle in b-metric spaces, Fixed Point Theory, 12 (2011), 21-28. [11] L. Ciri´c, A generalization of Banach’s contraction principle, Proc. Amer. Math. Soc., 45 (1974), 267-273. [12] C. Chifu and G. Petru¸sel, Fixed points for multivalued contractions in b-metric spaces with applications to fractals, Taiwanese J. Math., 18 (2014), 1365-1375. [13] S. Czerwik, Contraction mappings in b-metric spaces, Acta Math. Inform. Univ. Ostraviensis, 1 (1993), 5-11. [14] S. Czerwik, Nonlinear set-valued contraction mappings in b-metric spaces, Atti Sem. Mat. Fis. Univ. Modena, 46 (1998), 263-276. [15] P.Z. Daffer and H. Kaneko, Fixed points of generalized contractive multi-valued mappings, J. Math. Anal. Appl., 192 (1995), 655-666. [16] A.K. Dubey, R. Shukla and R.P. Dubey, Some fixed point results in b-metric spaces, Asian Journal of Mathematics and Applications, 2014, Article ID ama0147. [17] M.A. Khamsi and N. Hussain, KKM mappings in metric type spaces, Nonlinear Anal., 73 (2010), 3123-3129. [18] M. Kir, H. Kizitune, On some well known fixed point theorems in b-metric spaces, Turkish Journal of Analysis and Number Theory, 1 (2013), 13-16. [19] P. K. Mishra, S. Sachdeva and S.K. Banerjee, Some fixed point theorems in b-metric space, Turkish Journal of Analysis and Number Theory, 2 (2014), 19-22. [20] S. K. Mohanta, Some fixed point theorems using wt-distance in bmetric spaces, Fasc. Math., 54 (2015), 125-140.

12

[21] S.B. Nadler, Multi-valued contraction mappings, Pac. J. Math., 30 (1969), 475-488. [22] H.N. Nashine and Z. Kadelburg, Cyclic generalized ϕ-contractions in b-metric spaces and an application to integral equations, Filomat, 28 (2014), 2047-2057. [23] M.O. Olatinwo, A fixed point theorem for multi-valued weakly Picard operators in b-metric spaces, Demonstratio Math., 42 (2009), 599-606. [24] M. P˘acurar, Sequences of almost contractions and fixed points in bmetric spaces, An. Univ. Vest Timi¸s. Ser. Mat.-Inform., 48 (2010), 125-137. [25] J.R. Roshan, N. Hussain, S. Sedghi and N. Shobkolaei, Suzuki-type fixed point results in b-metric spaces, Math. Sci. (Springer), 9 (2015), 153– 160. [26] B.D. Rouhani and S. Moradi, Common fixed point of multivalued generalized ϕ-weak contractive mappings, Fixed Point Theory Appl., 2010, Article ID 708984. [27] M. Sarwar and M.U. Rahman, Fixed point theorems for Ciric’s and generalized contractions in b-metric spaces, International Journal of Analysis and Applications, 7 (2015), 70-78. [28] S.L. Singh, S. Czerwik, K. Kr´ol and A. Singh, Coincidences and fixed points of hybrid contractions, Tamsui Oxf. J. Math. Sci., 24 (2008), 401-416. [29] S. Shukla, Partial b-metric spaces and fixed point theorems, Mediterr. J. Math., 11 (2014), 703-711. [30] H. Yingtaweesittikul, Suzuki type fixed point for generalized multivalued mappings in b-metric spaces, Fixed Point Theory and Applications, 2013, 2013:215. University of Bucharest Faculty of Mathematics and Computer Science Str. Academiei 14, 010014 Bucharest, Romania E-mail: [email protected], mihail [email protected]

13