Altering distance functions and fixed point theorems through rational ...

3 downloads 1862 Views 99KB Size Report
Jan 25, 2012 - n→∞. Snx = z0. Fixed point theorems involving the notion of ..... each n ∈ N. Fix n > 1 and we assume that z ∈ FSn we want to show that.
arXiv:1201.5189v1 [math.FA] 25 Jan 2012

Altering distance functions and fixed point theorems through rational expression J.R. Moralesa and E.M. Rojasb a

Departamento de Matem´ aticas, Universidad de Los Andes, M´erida-5101, Venezuela. [email protected] b Departamento de Matem´ aticas, Pontificia Universidad Javeriana, Bogot´ a, Colombia [email protected]

Abstract In this paper, using altering distance functions we obtain a generalization of the results due to B.K. Das and S. Gupta [3] as well as the results given by B. Samet and H. Yazid [11]. Moreover, we study the so-called property P for the contraction mappings considered in this article.

1

Introduction and preliminary facts

In 1984, M.S. Khan, M. Swalech and S. Sessa [7] expanded the research of the metric fixed point theory to a new category by introducing a control function which they called an altering distance function. Definition 1.1 ([7]). A function ψ : R+ −→ R+ := [0, +∞) is called an altering distance function if the following properties are satisfied: (Ψ1 ) ψ(t) = 0 ⇔ t = 0. (Ψ2 ) ψ is monotonically non-decreasing. (Ψ3 ) ψ is continuous. By Ψ we denote the set of all altering distance functions.

1

Using those control functions the authors extend the Banach Contraction Principle by taking ψ = Id, (the identity mapping), in the inequality contraction (1.1) of the following theorem. Theorem 1.1 ([7]). Let (M, d) be a complete metric space, let ψ ∈ Ψ and let S : M −→ M be a mapping which satisfies the following inequality ψ[d(Sx, Sy)] ≤ aψ[d(x, y)]

(1.1)

for all x, y ∈ M and for some 0 < a < 1. Then. S has a unique fixed point z0 ∈ M and moreover for each x ∈ M , lim S n x = z0 . n→∞

Fixed point theorems involving the notion of altering distance functions has been widely studied, see for instance [1, 9, 11] and references therein. On the other hand, in 1975, B.K. Das and S. Gupta [3] proved the following result. Theorem 1.2. Let (M, d) be a metric space and let S : M −→ M be a given mapping such that, (i) d(Sx, Sy) ≤ ad(x, y) + bm(x, y)

(1.2)

for all x, y ∈ M, a > 0, b > 0, a + b < 1 where m(x, y) = d(y, Sy)

1 + d(x, Sx) 1 + d(x, y)

(1.3)

for all x, y ∈ M . (ii) For some x0 ∈ M , the sequence of iterates (S n x0 ) has a subsequence (S nk x0 ) with lim S nk x0 = z0 . Then z0 is the unique fixed point of S. k→∞

It is important to indicate that B.K. Das and S. Gupta [3] did not hypothesize that M is a complete metric space but they used this fact in their proof. Finally, in this paper we will study the property introduced by G.S. Jeong and B.E. Rhoades in [5] which they called the property P in metric spaces: Definition 1.2. Let S be a self mapping of a metric space (M, d) with a nonempty fixed point set F (S). Then S is said to satisfy the property P if F (S) = F (S n ) for each n ∈ N. 2

An interesting fact about mappings satisfying the property P is that they have no nontrivial periodic points. For more information about this property see e.g., [5, 6, 10]. The following lemma given by G.U. Babu and P.P. Sailaja [1] will be used in the sequel in order to prove our main results. Lemma 1.3. Let (M, d) be a metric space. Let (xn ) be a sequence in M such that lim d(xn , xn+1 ) = 0. (1.4) n→∞

If (xn ) is not a Cauchy sequence in M , then there exist an ε0 > 0 and sequences of integers positive (m(k)) and (n(k)) with m(k) > n(k) > k such that, d(xm(k) , xn(k) ) ≥ ε0 ,

d(xm(k)−1 , xn(k) ) < ε0

and (i) lim d(xm(k)−1 , xn(k)+1 ) = ε0 , k→∞

(ii) lim d(xm(k) , xn(k) ) = ε0 , k→∞

(iii) lim d(xm(k)−1 , xn(k) ) = ε0 . k→∞

Remark 1.4. From Lemma 1.3 is easy to get lim d(xm(k)+1 , xn(k)+1 ) = ε0 .

k→∞

In this paper, we consider the altering distance functions to generalize the results given in [3] and also we will obtain an extension of Theorem 2 given in [11]. Moreover, we study the property P for the contraction mappings considered in this work.

2

Fixed point theorems

This section is devoted to generalize the Theorem 1.2, as well as generalize Theorem 2 of [11]. Theorem 2.1. Let (M, d) be a complete metric space, let ψ ∈ Ψ and let S : M −→ M be a mapping which satisfies the following condition: ψ[d(Sx, Sy)] ≤ aψ[d(x, y)] + bψ(m(x, y)) 3

(2.1)

for all x, y ∈ M, a > 0, b > 0, a + b < 1 and m(x, y) is given by (1.3). Then S has a unique fixed point z0 ∈ M , and moreover for each x ∈ M lim S n x = z0 . n→∞

Proof. Let x ∈ M be an arbitrary point and let (xn ) be a sequence defined as follows: xn+1 = Sxn = S n+1 x, for each n ≥ 1. Now, ψ[d(xn , xn+1 )] = ψ[d(Sxn−1 , Sxn )] ≤ aψ[d(xn−1 , xn )] + bψ(m(xn−1 , xn ))

(2.2)

using (1.3), m(xn−1 , xn ) = d(xn , xn+1 )

1 + d(xn−1 , xn ) = d(xn , xn+1 ) 1 + d(xn−1 , xn )

substituting it into (2.2), one obtains ψ[d(xn , xn+1 )] ≤ aψ(d(xn−1 , xn )) + bψ(d(xn , xn+1 )) it follows that, a ψ(d(xn−1 , xn )) 1−b   2 a ψ(d(xn−2 , xn−1 )) ≤ . . . ≤  1 − b n a ≤ ψ(d(x0 , x1 )). 1−b

ψ[d(xn , xn+1 )] ≤

Since

a 1−b

(2.3)

∈ (0, 1), from (2.3) we obtain lim ψ[d(xn , xn+1 )] = 0.

n→∞

From the fact that ψ ∈ Ψ, we have lim d(xn , xn+1 ) = 0.

n→∞

(2.4)

Now, we will show that (xn ) is a Cauchy sequence in M. Suppose that (xn ) is not a Cauchy sequence, which means that there is a constant ε0 > 0 such that for each positive integer k, there are positive integers m(k) and n(k) with m(k) > n(k) > k such that d(xm(k) , xn(k) ) ≥ ε0 , d(xm(k)−1 , xn(k) ) < ε0 . From Lemma 1.3 and Remark 1.4 we obtain 4

lim d(xm(k) , xn(k) ) = ε0

(2.5)

lim d(xm(k)+1 , xn(k)+1 ) = ε0 .

(2.6)

k→∞

and k→∞

For x = xm(k) and y = yn(k) from (2.1) we have, ψ[d(xm(k)+1 , xn(k)+1 )] = ψ[d(Sx   m(k) , xn(k) )] ≤ aψ[d(xm(k) , xn(k) )] 1 + d(xm(k) , xm(k)+1 ) +bψ d(xn(k) , xn(k)+1 ) 1 + d(xm(k) , xn(k) ) using (2.4), (2.5) and (2.6) we obtain ψ(ε) =

lim ψ[d(xm(k)+1 , xn(k)+1 )]

k→∞

≤ a lim ψ[d(xm(k) , xn(k) )] k→∞

≤ aψ(ε), since a ∈ (0, 1), we get a contradiction. Thus (xn ) is a Cauchy sequence in the complete metric space M , thus there exists z0 ∈ M such that lim xn = z0 .

n→∞

Setting x = xn and y = z0 in (2.1) we have ψ[d(xn+1 , Sz0 )] = ψ[d(Sxn , Sz0 )] 

 1 + d(xn , Sxn ) ≤ aψ[d(xn , z0 )] + bψ d(z0 , Sz0 ) . 1 + d(xn , z0 ) Therefore, lim ψ[d(xn+1 , Sz0 )] ≤ bψ[d(z0 , Sz0 )]

n→∞

i.e., ψ[d(z0 , Sz0 )] ≤ bψ[d(z0 , Sz0 )] since b ∈ (0, 1), then ψ[d(z0 , Sz0 )] = 0 which implies that d(z0 , Sz0 ) = 0 thus z0 = Sz0 . Now we are going to establish the uniqueness of the fixed point. Let y0 , z0 be two fixed points of S such that y0 6= z0 . Putting x = y0 and y = z0 in (2.1) we get   1 + d(z0 , Sz0 ) ψ[d(Sz0 , Sy0 )] ≤ aψ[d(z0 , y0 )] + bψ d(y0 , Sy0 ) 1 + d(z0 , y0 ) = aψ[d(z0 , y0 )], which implies that ψ[d(z0 , y0 )] = 0, so d(z0 , y0 ) = 0. Thus z0 = y0 . 5



Corollary 2.2 ([11], Theorem 2). Let (M, d) be a complete metric space and let T : M −→ M be a mapping. We assume that for each x, y ∈ M, d(Sx,Sy)

Z

ϕ(t)dt ≤

0

d(y,Sy)

d(x,y) Z

ϕ(t)dt + b

0

1+d(x,Sx)

Z1+d(x,y)

ϕ(t)dt

(2.7)

0

where 0 < a + b < 1 and ϕ : R+ −→ R+ is a Lebesgue integrable mapping which is summable on each compact subset of [0, +∞), non negative and Rε such that ϕ(t)dt > 0 for all ε > 0. Then S admits a unique fixed point 0

z0 ∈ M such that for each x ∈ M

lim S n x = z0 .

n→∞

Proof. Let ϕ : R+ −→ R+ be as in the hypothesis, we define ψ0 (t) = Z t ϕ(t)dt, t ∈ R+ . It is clear that ψ0 (0) = 0. ψ0 is monotonically non 0

decreasing and by hypothesis ψ0 is absolutely continuous, hence ψ0 is continuous. Therefore, ψ0 ∈ Ψ. So (2.1) becomes   1 + d(x, Sx) . ψ0 (d(Sx, Sy)) ≤ aψ0 (d(x, y)) + bψ0 d(y, Sy) 1 + d(x, y)

Hence from Theorem 2.1 there exists a unique fixed point z0 ∈ M such that for each x ∈ M , lim S n x = z0 .  n→∞

Remark 2.3. (1) If we take b = 0, then (2.1) reduces to (1.2), thus the Theorem 1.1 is a corollary of Theorem 2.1. (2) If we take ψ = Id in (2.1), then we obtain (1.2). Therefore the Theorem 2.1 is a generalization of Theorem 1.2.

3

The property P .

In this section we are going to prove that the mappings satisfying the contractive conditions(1.1), (1.2), (2.1) and (2.7) fulfill the property P .

6

Theorem 3.1. Let (M, d) be a complete metric space, let ψ ∈ Ψ and let S : M −→ M be a mapping which satisfies the following inequality: ψ[d(Sx, Sy)] ≤ aψ[d(x, y)] for all x, y ∈ M and for some 0 < a < 1. Then FS 6= ∅ and S has the property P . Proof. From Theorem 1.1, S has a fixed point. Therefore FS n 6= ∅ for each n ∈ N. Fix n > 1 and we assume that z ∈ FS n we want to show that z ∈ FS . Suppose that z 6= Sz, using (1.1) ψ[d(z, Sz)] = ψ[d(S n z, S n+1 z)] ≤ aψ[d(S n−1 z, S n z)] ≤ . . . ≤ an ψ[d(z, Sz)], since a ∈ (0, 1), lim ψ[d(z, Sz)] = 0. From the fact that ψ ∈ Ψ, we get n→∞ z = Sz which is a contraction. Therefore z ∈ FS i.e., S has the property P .  Theorem 3.2. Let (M, d) be a complete metric space and let S : M −→ M be a mapping which satisfies the contractive condition (1.2), then FS 6= ∅ and S has the property P . Proof. From Theorem 1.2, FS 6= ∅. Therefore FS n 6= ∅ for each n ∈ N. Fix n > 1 and we assume that z ∈ FS n . We want to show that z ∈ FS . Suppose that z 6= Sz. Using (1.2), d(z, Sz) = d(S n z, S n+1 z) ≤ ad(S n−1 z, S n z) 1 + d(S n−1 z, S n z) +bd(S n z, S n+1 z) 1 + d(S n−1 z, S n z) n−1 n = ad(S z, S z) + bd(S n z, S n+1 z). Therefore n

d(z, Sz) = d(S z, S

n+1

a z) ≤ d(S n−1 z, S n z) ≤ . . . ≤ 1−b



a 1−b

n

d(z, Sz),

which is a contradiction. Consequently, z ∈ FS and S has the property P .  Theorem 3.3. Let (M, d) be a complete metric space, let ψ ∈ Ψ and let S : M −→ M be a mapping which satisfies the condition (2.1). Then FS 6= ∅ and S has a property P. 7

Proof. From Theorem 2.1, S has a fixed point. Therefore FS n 6= ∅ for each n ∈ N. Fix n > 1 and assume that z ∈ FS n we wish to show that z ∈ FS . Suppose that z 6= Sz, using (2.1), ψ[d(z, Sz)] = ψ[d(Sn z, S n+1 z)] ≤ aψ[d(S n−1 z, S n z)]  1 + d(S n−1 z, S n z) +bψ d(S n z, S n+1 z) 1 + d(S n−1 z, S n z) n−1 n ≤ aψ[d(S z, S z)] + bψ[d(S n z, S n+1 z)] hence (1 − b)ψ[d(S n z, S n+1 z)] ≤ aψ[d(S n−1 z, S n z)] a ψ[d(z, Sz)] = ψ(d(S n z, S n+1 z)) ≤ ψ[d(S n−1 z, S n z)] 1 − b n  a ψ(d(z, Sz)), ≤ 1−b thus ψ[d(z, Sz)] ≤



a 1−b

n

ψ(d(z, Sz))

which is a contradiction, therefore ψ[d(z, Sz)] = 0, since ψ ∈ Ψ, we conclude that d(z, Sz) = 0, thus z ∈ FS and S has the property P . 

References [1] G. U. R. Babu and P.P. Sailaja, A fixed point theorem of generalized weakly contractive maps in orbitally complete metric space, Thai Journal of Math., 9 1 (2011) 1–10. [2] R. Chugh, T. Kadian, A. Rani and B. E. Rhoades, Property P in GMetric Spaces, Fixed Point Theory and Applications, vol. 2010, Article ID 401684, 12 pages, 2010. doi:10.1155/2010/401684 [3] B.K. Das and S. Gupta, An extension of Banach contractive principle through rational expression, Indian Jour. Pure and Applied Math., 6 (1975) 1455–1458. [4] P.N. Dulta and B.S. Choudhury, A Generalisation of Contraction Principle in Metric Spaces, Fixed Point Theory and Applications, vol. 2008, Article ID 406368, 8 pages, 2008. doi:10.1155/2008/406368 [5] G.S. Jeong and B.E. Rhoades, Maps for which F (T ) = F (T n ), Fixed Point Theory and Appl., 6 (2005) 87–131. 8

[6] G.S. Jeong and B.E. Rhoades, More maps for which F (T ) = F (T n ), Demostratio Math., 40 (2007) 671–680. [7] M. S. Khan, M. Swalech and S. Sessa, Fixed point theorems by altering distances between the points, Bull. Austral Math. Soc., 30 (1984) 1–9. [8] S.V.R. Naidu, Some fixed point theorems in metric spaces by altering distances, Czechoslovak Math. Jour. 53 1 (2003) 205–212. [9] V. Popa and M. Mocanu, Altering distance and common fixed points under implicit relations, Hacettepe Jour. Math. and Stat., 38 3 (2009) 329–337. [10] B.E. Rhoades and M. Abbas, Maps satisfying generalized contractive conditions of integral type for which F (T ) = F (T n ), Int. Jour. of Pure and Applied Math. 45 2 (2008) 225–231. [11] B. Samet and H. Yazidi, An extension of Banach fixed point theorem for mappings satisfying a contractive condition of integral type, J. Nonlinear Sci. Appl., accepted, 2011.

9