Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

STATISTICAL APPROXIMATION BY POSITIVE LINEAR

OPERATORS ON MODULAR SPACES

SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

Abstract. In this paper, we investigate the problem of statistical ap-


proximation to a function by means of positive linear operators de…ned
on a modular space. Especially, in order to get more powerful results
than the classical aspects we mainly use the concept of statistical con-
vergence. A non-trivial application is also presented.

1. Introduction
In this paper, we investigate the problem of statistical approximation
to a function f by means of positive linear operators de…ned on a modu-
lar space. So, we obtain various Korovkin-type approximation theorems in
modular spaces via the concept of A-statistical convergence, where A is a
non-negative regular summability matrix. With the help of the notions of
modular convergence and strong convergence, some approximation theorems
have recently been introduced by Bardaro and Mantellini [3]. However, in
this paper, we replace the strong (or modular) convergence by the more
general A-statistical convergence with respect to a …xed summability ma-
trix A: When A is the identity matrix, our results reduce to the results in
[3]. Also, we display an example satisfying all conditions of our statistical
approximation result in a modular space but not the classical one in [3].
We now recall some basic de…nitions and notations used in the paper.
Let
A := [ajn ] (j; n 2 N := f1; 2; 3; :::g)
be an in…nite summability matrix. For a given sequence x := fxn g, the
A-transform of x, denoted by
Ax := f(Ax)j g;
is given by
1
X
(Ax)j = ajn xn ;
n=1

Key words and phrases. Positive linear operators, modular space, regular matrix, sta-
tistical convergence, Korovkin theorem.
2000 Mathematics Subject Classi…cation. 41A25, 41A36.
1
2 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

provided the series converges for each j 2 N. We say that A is regular (see
[12]) if
lim Ax = L whenever lim x = L:
Assume that A is a non-negative regular summability matrix. Then the
sequence x = fxn g is called A-statistically convergent to L provided that,
for every " > 0;
X
(1.1) lim ajn = 0:
j
n: jxn Lj "

We denote this limit as follows (cf. [8]; see also [5, 6, 13])
stA lim x = L:
Actually, this convergence method is based on the concept of A-density.
Recall the A-density of a subset K N, denoted by
A fKg;
is given by
1
X
A fKg = lim ajn K (n) ;
j
n=1
provided the limit exists, where K is the characteristic function of K; or
equivalently X
A fKg = lim ajn :
n
n2K
So, by (1.1), we easily see that
stA lim x = L i¤ A fn : jxn Lj "g = 0 for every " > 0:
We should note that if we take A = C1 := [cjn ], the Cesáro matrix de…ned
by 8
< 1
; if 1 n j;
cjn := j
: 0; otherwise,
then A-statistical convergence reduces to the concept of statistical conver-
gence (cf. [7]; see also [9, 10, 11]). In this case, we write
st lim x = L instead of stC1 lim x = L:
Further, taking A = I, the identity matrix, A-statistical convergence coin-
cides with the ordinary convergence, i.e.,
stI lim x = lim x = L:
Observe that every convergent sequence (in the usual sense) is A-statistically
convergent to the same value for any non-negative regular matrix A, but its
converse is not always true. Actually, in [13], Kolk proved that A-statistical
convergence is stronger than convergence when A = [ajn ] is a non-negative
regular summability matrix such that
lim maxfajn g = 0:
j n
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 3

The concepts of statistical limit superior and limit inferior have been
introduced by Fridy and Orhan [11]. A-statistical analogs of these concepts
have been examined by Connor and Kline [5], and Demirci [6] as follows.
The A-statistical limit superior of a number sequence x = fxn g, denoted by
stA lim sup x;
is de…ned by
sup Bx , if Bx 6= ;
stA lim sup x =
1, if Bx = ;
where Bx := fb 2 R : A fn : xn > bg = 6 0g and denotes the empty set. We
note that by A fKg = 6 0 we mean either A fKg > 0 or K fails to have
A-density. Similarly, the A-statistical limit inferior of fxn g; denoted by
stA lim inf x;
is de…ned by
inf Cx , if Cx 6= ;
stA lim inf x =
+1 , if Cx = ;
where Cx := fc 2 R : A fn : xn < cg = 6 0g. Of course, if we take A = C1 ;
then the above de…nitions reduce to the concepts of st lim sup x and st
lim inf x given in [11], respectively. As in the ordinary limit superior or
inferior, it was proved that
stA lim inf x stA lim sup x
and also that, for any sequence x = fxn g satisfying A fn : jxn j > M g = 0
for some M > 0;
stA lim x = L i¤ stA lim inf x = stA lim sup x = L:
We now focus on modular spaces.
Let I = [a; b] be a bounded interval of the real line R provided with the
Lebesgue measure. Then, by X (I) we denote the space of all real-valued
measurable functions on I provided with equality a.e. As usual, let C (I)
denote the space of all continuous real-valued functions, and C 1 (I) denote
the space of all in…nitely di¤erentiable functions on I. In this case, we say
that a functional : X (I) ! [0; +1] is a modular on X (I) provided that
the following conditions hold:
(i) (f ) = 0 if and only if f = 0 a.e. in I,
(ii) ( f ) = (f ) for every f 2 X (I),
(iii) ( f + g) (f )+ (g) for every f; g 2 X(I) and for any ; 0
with + = 1.
Recall that a modular is called N -quasi convex if the following is satis…ed:
there exists a constant N 1 such that
( f + g) N (N f ) + N (N g)
holds for every f; g 2 X (I), ; 0 with + = 1. In particular,
if N = 1; then is called convex.
4 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

Furthermore, a modular is called N -quasi semiconvex if it holds:


there exists a constant N 1 such that
(af ) N a (N f )
holds for every f 2 X (I) and a 2 (0; 1]:
It is clear that every N -quasi semiconvex modular is N -quasi convex. We
should recall that the above two concepts were introduced and discussed in
details by Bardaro et. al. [4].
We now consider some appropriate vector subspaces of X(I) by means of
a modular as follows:

L (I) := f 2 X (I) : lim ( f) = 0


!0+

and
E (I) := ff 2 L (I) : ( f ) < +1 for all > 0g :
Here, L (I) is called the modular space generated by ; and also E (I) is
called the space of the …nite elements of L (I) : Observe that if is N -quasi
semiconvex, then the space
ff 2 X (I) : [ f ] < +1 for some > 0g
coincides with L (I). The notions about modulars are introduced in [18]
and widely discussed in [4] (see also [15, 17]).
Now we recall the convergence methods in modular spaces.
Let ffn g be a function sequence whose terms belong to L (I) : Then,
ffn g is modularly convergent to a function f 2 L (I) i¤
(1.2) lim ( 0 (fn f )) = 0 for some 0 > 0:
n
Also, ffn g is F -norm convergent (or, strongly convergent) to f i¤
(1.3) lim ( (fn f )) = 0 for every > 0:
n
It is known from [17] that (1.2) and (1.3) are equivalent if and only if the
modular satis…es the 2 -condition, i.e. there exists a constant M > 0
such that [2f ] M [f ] for every f 2 X (I).
In this paper, we will need the following assumptions on a modular :
if (f ) (g) for jf j jgj ; then is called monotone,
if the characteristic function I of the interval I belongs to L (I) ;
is called …nite,
if is …nite and, for every " > 0, > 0; there exists a > 0 such
that ( B ) < " for any measurable subset B I with jBj < ,
then is called absolutely …nite,
if I 2 E (I) ; then is called strongly …nite
is called absolutely continuous provided that there exists an > 0
such that, for every f 2 X (I) with (f ) < +1, the following
condition holds: for every " > 0 there is > 0 such that ( f B ) <
" whenever B is any measurable subset of I with jBj < .
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 5

Observe now that (see [3]) if a modular is monotone and …nite, then
we have C(I) L (I) : In a similar manner, if is monotone and strongly
…nite, then C(I) E (I). Some important relations between the above
properties may be found in [2, 4, 16, 18].

2. Statistical Korovkin Theorems in Modular Spaces


Let be a monotone and …nite modular on X (I), and let A = [ajn ] be a
non-negative regular summability matrix. Assume that D is a set satisfying
C 1 (I) D L (I) : We can construct such a subset D since is monotone
and …nite (see [3]). Assume further that T := fTn g is a sequence of positive
linear operators from D into X (I) for which there exists a subset XT D
containing C 1 (I) such that

(2.1) stA lim sup ( (Tn h)) P ( h)


n

holds for every h 2 XT ; > 0 and for an absolute positive constant P . We


also use the test functions ei de…ned by

ei (x) = xi (i = 0; 1; 2; :::):

Theorem 2.1. Let A = [ajn ] be a non-negative regular summability matrix


and let be a monotone, strongly …nite, absolutely continuous and N -quasi
semiconvex modular on X (I). Let T := fTn g be a sequence of positive linear
_ Assume that
operators from D into X (I) satisfying (2:1):

(2.2) stA lim ( (Tn ei ei )) = 0 for every > 0 and i = 0; 1; 2:


n

Now let f be any function belonging to L (I) such that f g 2 XT for every
g 2 C 1 (I) : Then, we have

(2.3) stA lim ( 0 (Tn f f )) = 0 for some 0 > 0:


n

Proof. We …rst claim that

(2.4) stA lim ( (Tn g g)) = 0 for every g 2 C(I) and every > 0:
n

To see this assume that g belongs to C (I) and is any positive number.
Then, by the uniform continuity of g on I and by the linearity and positivity
of the operators Tn ; we easily see that (see, for instance, [1, 14]), for a given
" > 0; there exists a > 0 such that

jTn (g; x) g(x)j "Tn (e0 ; x) + M jTn (e0 ; x) e0 (x)j


2M
+ 2 Tn (y x)2 ; x
6 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

holds for every x 2 I and n 2 N, where M := supx2I jg(x)j. The above


inequality implies that
2M c2
jTn (g; x) g(x)j " + (" + M + 2 ) jTn (e0 ; x) e0 (x)j
4M c
+ 2 jTn (e1 ; x) e1 (x)j
2M
+ 2 jTn (e2 ; x) e2 (x)j ;

where c := maxfjaj ; jbjg: So, the last inequality gives, for any > 0; that
jTn (g; x) g(x)j " + K fjTn (e0 ; x) e0 (x)j + jTn (e1 ; x) e1 (x)j
+ jTn (e2 ; x) e2 (x)jg;
2M c2 4M c 2M
where K := max " + M + 2 ; 2 ; 2 : Now, applying the modular
in the both-sides of the above inequality, since is monotone, we have
( (Tn g g)) ( " + K jTn e0 e0 j + K jTn e1 e1 j + K jTn e2 e2 j) :
Hence, we may write that
( (Tn g g)) (4 ") + (4 K (Tn e0 e0 )) + (4 K (Tn e1 e1 ))
+ (4 K (Tn e2 e2 ))
Since is N -quasi semiconvex and strongly …nite, we have, assuming 0 <
" 1
( (Tn g g)) N " (4 N ) + (4 K (Tn e0 e0 ))
+ (4 K (Tn e1 e1 )) + (4 K (Tn e2 e2 )) .
For a given r > 0; choose an " 2 (0; 1] such that N " (4 N ) < r: Now de…ne
the following sets:
G : = fn : ( (Tn g g)) rg
r N " (4 N )
G ;1 : = n : (4 K (Tn e0 e0 ))
3
r N " (4 N )
G ;2 : = n : (4 K (Tn e1 e1 ))
3
r N " (4 N )
G ;3 : = n : (4 K (Tn e2 e2 )) :
3
Then, it is easy to see that G G ;1 [ G ;2 [ G ;3 . So, we can write, for
all j 2 N, that
X X X X
(2.5) ajn ajn + ajn + ajn .
n2G n2G ;1 n2G ;2 n2G ;3
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 7

Taking limit as j ! 1 in (2.5) and using the hypothesis (2.2), we get


X
lim ajn = 0;
j
n2G

which proves our claim (2.4). Observe that (2.4) also holds for every g 2
C 1 (I) because of C 1 (I) C(I): Now let f 2 L (I) satisfying f g 2 XT
for every g 2 C 1 (I). Since jIj < 1 and is strongly …nite and absolutely
continuous, we can see that is also absolutely …nite on X(I) (see [2]). Using
these properties of the modular ; it is known from [4, 16] that the space
C 1 (I) is modularly dense in L (I) ; i.e., there exists a sequence fgk g
C 1 (I) such that
lim [3 0 (gk f )] = 0 for some 0 > 0:
k
This means that, for every " > 0; there is a positive number k0 = k0 (") so
that
(2.6) [3 0 (gk f )] < " for every k k0 :
On the other hand, by the linearity and positivity of the operators Tn , we
may write that
0 jTn (f ; x) f (x)j 0 jTn (f gk0 ; x)j + 0 jTn (gk0 ; x) gk0 (x)j
+ 0 jgk0 (x) f (x)j
holds for every x 2 I and n 2 N. Applying the modular in the last
inequality and using the monotonicity of ; we have
( 0 (Tn f f )) (3 0 Tn (f gk0 )) + (3 0 (Tn gk0 gk0 ))
(2.7)
+ (3 0 (gk0 f )) .
Then, it follows from (2.6) and (2.7) that
(2.8) ( 0 (Tn f f )) " + (3 0 Tn (f gk0 )) + (3 0 (Tn gk0 gk0 )) :
So, taking A-statistical limit superior as n ! 1 in the both-sides of (2.8)
and also using the facts that gk0 2 C 1 (I) and f gk0 2 XT ; we obtained
from (2.1) that
stA lim sup ( 0 (Tn f f )) " + P (3 0 (f gk0 ))
n
+stA lim sup (3 0 (Tn gk0 gk0 )) ;
n
which gives
(2.9)
stA lim sup ( 0 (Tn f f )) "(P +1)+stA lim sup (3 0 (Tn gk0 gk0 )) :
n n
By (2.4), since
stA lim (3 0 (Tn gk0 gk0 )) = 0;
n
we get
(2.10) stA lim sup (3 0 (Tn gk0 gk0 )) = 0:
n
8 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

Combining (2.9) with (2.10), we conclude that

stA lim sup ( 0 (Tn f f )) "(P + 1):


n

Since " > 0 was arbitrary, we …nd

stA lim sup ( 0 (Tn f f )) = 0:


n

Furthermore, since ( 0 (Tn f f )) is non-negative for all n 2 N, we can


easily show that
stA lim ( 0 (Tn f f )) = 0;
n

which completes the proof.

If the modular satis…es the 2 -condition, then one can get the following
result from Theorem 2.1 at once.

Theorem 2.2. Let A = [ajn ]; T := fTn g; be the same as in Theorem 2:1:


If satis…es the 2 -condition, then the following statements are equivalent:
(a) stA lim ( (Tn ei ei )) = 0 for every > 0 and i = 0; 1; 2;
n
(b) stA lim ( (Tn f f )) = 0 for every > 0 provided that f is
n
any function belonging to L (I) such that f g 2 XT for every
g 2 C 1 (I).

If one replaces the matrix A by the identity matrix, then the condition
(2.1) reduces to

(2.11) lim sup ( (Tn h)) P ( h)


n

for every h 2 XT ; > 0 and for an absolute positive constant P: In this


case, the next results which were obtained by Bardaro and Mantellini [3]
immediately follows from our Theorems 2.1 and 2.2.

Corollary 2.3 ([3]). Let be a monotone, strongly …nite, absolutely con-


tinuous and N -quasi semiconvex modular on X (I). Let T := fTn g be a
sequence of positive linear operators from D into X (I) satisfying (2:11): _ If
fTn ei g is strongly convergent to ei for each i = 0; 1; 2; then fTn f g is mod-
ularly convergent to f provided that f is any function belonging to L (I)
such that f g 2 XT for every g 2 C 1 (I).

Corollary 2.4 ([3]). T := fTn g and be the same as in Corollary 2:3: If


satis…es the 2 -condition, then the following statements are equivalent:
(a) fTn ei g is strongly convergent to ei for each i = 0; 1; 2;
(b) fTn f g is strongly convergent to f provided that f is any function
belonging to L (I) such that f g 2 XT for every g 2 C 1 (I).
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 9

3. Concluding Remarks
In this section, we display an example such that our Korovkin-type statis-
tical approximation results in modular spaces are stronger than the classical
ones in [3].

Example. Take I = [0; 1] and let ' : [0; 1) ! [0; 1) be a continuous


function for which the following conditions hold:
' is convex,
' (0) = 0, ' (u) > 0 for u > 0 and limn!1 ' (u) = 1.
Hence, consider the functional ' on X(I) de…ned by
Z1
'
(3.1) (f ) := ' (jf (x)j) dx for f 2 X (I) :
0
In this case, ' is a convex modular on X (I) ; which satis…es all assump-
tions listed in Section 1 (see [3]). Consider the Orlicz space generated by '
as follows:
'
L' (I) := ff 2 X (I) : ( f ) < +1 for some > 0g :
Then consider the following classical Bernstein-Kantorovich operator U :=
fUn g on the space L' (I) (see [3]) which is de…ned by:

(3.2)
(k+1)=(n+1)
Z
n
X n k
Un (f ; x) := x (1 x)n k
(n + 1) f (t) dt for x 2 I:
k
k=0 k=(n+1)

Observe that the operators Un map the Orlicz space L' (I) into itself.
Moreover, the property (2.11) is satis…ed with the choice of XU := L' (I).
Then, by Corollary 2.3, we know that, for any function f 2 L' (I) such that
f g 2 XU for every g 2 C 1 (I), fUn f g is modularly convergent to f:
Now take A = C1 ; the Cesáro matrix of order one, and de…ne a sequence
fsn g by
0; if n is square
(3.3) sn =
1; otherwise.
In this case, we know that C1 -statistical convergence coincides with statis-
tical convergence, and its limit is denoted by st lim : Observe now that,
for the sequence fsn g given by (3.3),
st lim sn = 1:
n
However, the sequence fsn g does not convergent in the usual sense. Then,
using the operators Un , we de…ne the sequence of positive linear operators
V := fVn g on L' (I) as follows:
(3.4) Vn (f ; x) = sn Un (f ; x) for f 2 L' (I) , x 2 [0; 1] and n 2 N,
10 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

where s = fsn g is the same as in (3.3). By Lemma 5.1 of [3], we get, for
every h 2 XV := L' (I); > 0 and for an absolute positive constant P; that
'
( Vn h) = ' ( sn Un h)
= sn ' ( Un h)
P sn ' ( h) :

Now taking st lim sup as n ! 1 on the both-sides of the above inequality


and using the fact that st lim sup sn = st lim sn = 1; we obtain
n n

' '
(3.5) st lim sup ( Vn f ) P ( f ):
n

Therefore, the condition (2.1) works for our operators Vn given by (3.4) with
the choice of XV = XU = L' (I):
We now claim that
'
(3.6) st lim ( (Vn ei ei )) = 0 for every > 0 and i = 0; 1; 2:
n

Indeed, …rst observe that (see [1])

Vn (e0 ; x) = sn ;
nx 1
Vn (e1 ; x) = sn + ;
n + 1 2 (n + 1)
n (n 1) x2 2nx 1
Vn (e2 ; x) = sn 2 + 2 + :
(n + 1) (n + 1) 3 (n + 1)2

So, we can see, for any > 0; that

jVn (e0 ; x) e0 (x)j = (1 sn );

which implies
' '
( (Vn e0 e0 )) = ( (1 sn ))
Z1
= ' ( (1 sn )) dx
0
= ' ( (1 sn ))
= (1 sn )' ( )

because of the de…nition of (sn ): Now, since st limn (1 sn ) = 0; we get


'
st lim ( (Vn e0 e0 )) = 0 for every > 0;
n
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 11

which guarantees that (3.6) holds true for i = 0: Also, since


nsn sn
jVn (e1 ; x) e1 (x)j = x 1 +
n+1 2(n + 1)
nsn sn
1 +
n+1 2(n + 1)
3sn
= (1 sn ) + ;
2(n + 1)
by the de…nitions of (sn ) and '; we may write that
' ' 3sn
( (Vn e1 e1 )) (1 sn ) +
2(n + 1)
' 3 sn
(2 (1 sn )) + '
n+1
3 sn
= ' (2 (1 sn )) + ' ;
n+1
which implies, for any > 0; that
' 3
(3.7) ( (Vn e1 e1 )) (1 sn )' (2 ) + sn '
n+1
3 3
Since ' is continuous, we have lim ' = ' lim = '(0) = 0:
n+1
n n n+1
Also considering st lim sn = 1; we obtain from (3.7) that
n
'
st lim ( (Vn e1 e1 )) = 0 for every > 0:
n
Hence (3.6) is valid for i = 1: Finally, since
n (n 1) sn
jVn (e2 ; x) e2 (x)j = x2 1
(n + 1)2
2nsn 1
+x 2 +
(n + 1) 3 (n + 1)2
n (n 1) sn
1
(n + 1)2
2nsn sn
+ 2 +
(n + 1) 3 (n + 1)2
15n + 4
= (1 sn ) + sn
3(n + 1)2
we get
' ' ' 15n + 4
( (Vn e2 e2 )) (2 (1 sn )) + 2 sn
3(n + 1)2
15n + 4
= ' (2 (1 sn )) + ' 2 sn ;
3(n + 1)2
12 SEVDA KARAKUŞ, KAMIL DEMIRCI AND OKTAY DUMAN

which yields
' 15n + 4
(3.8) ( (Vn e2 e2 )) (1 sn )' (2 ) + sn ' 2
3(n + 1)2
Considering the continuity of '; it follows from (3.8) that
'
st lim ( (Vn e2 e2 )) = 0 for every > 0:
n
So, our claim (3.6) holds true for each i = 0; 1; 2 and for any > 0: Now,
from (3.5) and (3.6), we can say that our sequence V := fVn g de…ned by
(3.4) satisfy all assumptions of Theorem 2.1. Therefore, we conclude that
'
st lim ( 0 (Vn f f )) = 0 for some 0 >0
n
holds for any f 2 L' (I) such that f g 2 XV = L' (I) for every g 2 C 1 (I).
However, for the same f ’s and for any > 0; since
' '
( (Vn f f )) = ( (sn Un f f ))
' ( f ); if n is square
= ' ( (U f
n f )) ; otherwise,
the sequence f ' ( (Vn f f ))g has two subsequences with di¤erent limit
points. In other words, there is not any > 0 such that the sequence
f ' ( (Vn f f ))g goes to zero as n ! 1: This clearly shows that fVn f g
is not modularly convergent to f: So, Corollary 2.3 does not work for the
sequence V := fVn g:
With this example, we can immediately say that our Theorem 2.1 is a non-
trivial generalization of the classical Korovkin-type results in the modular
spaces.

Acknowledgement. The authors would like to thank the referee for care-
fully reading the manuscript.

References
[1] F. Altomare and M. Campiti, Korovkin Type Approximation Theory and Its Appli-
cation, Walter de Gruyter Publ., Berlin, 1994.
[2] C. Bardaro and I. Mantellini, Approximation properties in abstract modular spaces
for a class of general sampling-type operators, Appl. Anal. 85 (2006), 383-413.
[3] C. Bardaro and I. Mantellini, Korovkin’s theorem in modular spaces, Commentationes
Math. 47 (2007), 239-253.
[4] C. Bardaro, J. Musielak and G. Vinti, Nonlinear Integral Operators and Applications,
de Gruyter Series in Nonlinear Analysis and Appl. Vol. 9, Walter de Gruyter Publ.,
Berlin, 2003.
[5] J.S. Connor and J. Kline, On statistical limit points and the consistency of statistical
convergence, J. Math. Anal. Appl. 197 (1996), 392-399.
[6] K. Demirci, A-statistical core of a sequence, Demonstratio Math. 33 (2006), 343-353.
[7] H. Fast, Sur la convergence statistique, Colloq, Math. 2 (1951), 241-244.
[8] A.R. Freedman and J.J. Sember, Densities and summability, Paci…c. J. Math. 95
(1981), 293-305.
[9] J.A. Fridy, On statistical convergence, Analysis 5 (1985), 301-313.
STATISTICAL APPROXIMATION BY POSITIVE LINEAR OPERATORS 13

[10] J.A. Fridy, Statistical limit points, Proc. Amer. Math. Soc. 118 (1993), 1187-1192.
[11] J.A. Fridy and C. Orhan, Statistical limit superior and limit inferior, Proc. Amer.
Math. Soc. 125 (1997), 3625-3631.
[12] G.H. Hardy, Divergent Series, Oxford Univ. Press, London, 1949.
[13] E. Kolk, Matrix summability of statistically convergent sequences, Analysis 13 (1993),
77-83.
[14] P.P. Korovkin, Linear Operators and Approximation Theory, Hindustan Publ. Corp.,
Delhi, 1960.
[15] W.M. Kozlowski, Modular Function Spaces, Pure Appl. Math., Vol. 122, Marcel
Dekker, Inc., New York, 1988.
[16] I. Mantellini, Generalized sampling operators in modular spaces, Commentationes
Math., 38 (1998), 77-92.
[17] J. Musielak, Orlicz Spaces and Modular Spaces, Lecture Notes in Mathematics, Vol.
1034, Springer-Verlag, Berlin, 1983.
[18] J. Musielak, Nonlinear approximation in some modular function spaces I, Math.
Japon. 38 (1993), 83-90.

Sevda Karakuş
Sinop University, Faculty of Arts and Sciences,
Department of Mathematics, 57000 Sinop, Turkey
E-Mail: skarakus@omu.edu.tr

Kamil Demirci
Sinop University, Faculty of Arts and Sciences,
Department of Mathematics, 57000 Sinop, Turkey
E-Mail: kamild@omu.edu.tr

Oktay Duman (corresponding author )


TOBB Economics and Technology University,
Faculty of Arts and Sciences, Department of Mathematics,
Sö¼
gütözü 06530, Ankara, Turkey
E-Mail: oduman@etu.edu.tr

You might also like