• Nie Znaleziono Wyników

On the rates of convergence of

N/A
N/A
Protected

Academic year: 2021

Share "On the rates of convergence of"

Copied!
20
0
0

Pełen tekst

(1)

Paulina Pych-Taberska, Harun Karsli

On the rates of convergence of

Chlodovsky-Kantorovich operators and their B´ezier variant

Abstract. In the present paper we consider the B´ezier variant of Chlodovsky- Kantorovich operators Kn−1,αf for functions f measurable and locally bounded on the interval [0,∞). By using the Chanturiya modulus of variation we estimate the rate of pointwise convergence of Kn−1,αf(x) at those x > 0 at which the one-sided limits f(x+) , f(x−) exist.

2000 Mathematics Subject Classification: 41A25, 41A35.

Key words and phrases: rate of convergence, Chlodovsky-Kantorovich operator, B´ezier basis, Chanturiya’s modulus of variation, pth power variation.

1. Introduction. Let f be a function defined on the interval [0, ∞) and let ℕ := {1, 2, ...} . The Bernstein-Chlodovsky polynomials Cnf applied to f are defined as

(1) Cnf (x) :=

Xn k=0

f

kbn

n

 Pn,k

x bn



for x∈ [0, bn], n ∈ ℕ,

where Pn,k(t) =

 n k



tk(1 − t)n−k for t ∈ [0, 1] and (bn) is a positive increasing sequence satisfying the properties

(2) lim

n→∞bn= ∞ and limn→∞bn

n = 0

(see [5,11,15]). They were introduced by I. Chlodovsky [5] in 1937 as a generalization of the classical Bernstein polynomials Bnf of a function f on the interval [0, 1] ( defined by (1) in which bn = 1 for all n ∈ ℕ).

(2)

Recently, Butzer and Karsli [3] considered the Kantorovich variant of the Chlodovsky- Bernstein polynomials as

(3) Knf (x) := n + 1 bn+1

Xn k=0

Pn,k

 x bn+1

 (k+1)bn+1Zn+1

kbn+1 n+1

f (u)du if 0 ¬ x ¬ bn+1,

where f ∈ Lloc[0, ∞) i.e. f is locally integrable on the interval [0, ∞) and (bn) satisfies conditions (2). Note that the classical Kantorovich polynomials Knf of a function f integrable on the interval [0, 1] are defined as in (3) in which bn= 1 for all n ∈ ℕ. Several authors have studied approximation properties of those classical Kantorovich polynomials (see e.g. [1, 6, 9, 11, 12, 17, 18] etc.). It is easy to see that the operators Kn and Kn are described in the following way. If F denotes an indefinite integral of f, then Knf (x) = dxd Bn+1F (x) and Knf (x) = dxdCn+1F (x). As is well- known, if f is Lebesgue integrable on [0, 1] then

nlim→∞Knf (x) = f (x) almost everywhere on [0, 1].

Analogous relation holds true for operators (3). Namely (see [14]) if f ∈ Lloc[0, ∞) and if

nlim→∞

bn

Z

0

|f(u)| du exp(−σn bn

) = 0 for each σ > 0,

then

nlim→∞Knf (x) = f (x) almost everywhere on [0,∞),

i.e. at every x > 0 at which F0(x) = f(x). Some modified Chlodovsky-Kantorovich operators are considered also in [8].

In this paper we introduce the B´ezier variant of the Chlodovsky-Kantorovich operators (3) of order (n − 1) for f ∈ Lloc[0, ∞) as

(4) Kn−1,αf (x) := n bn

nX−1 k=0

Q(α)n−1,k

x bn

 (k+1)bnZn

kbnn

f (u)du if 0¬ x ¬ bn,

where α > 0, Q(α)n−1,k(t) = Jnα−1,k(t) − Jnα−1,k+1(t) and Jn−1,k(t) are the B´ezier basis functions defined for t ∈ [0, 1] as

Jn−1,k(t) =

n−1

X

j=k

Pn−1,j(t) if k = 0, 1, . . . , n − 1,

Jn−1,n(t) = 0 (see [2]). Clearly, if α = 1 then Kn−1,αf reduce to operators (3) with n replaced by n− 1. Our paper is concerned with the rate of pointwise convergence of operators (4) when f ∈ Mloc[0, ∞),i.e. f is measurable and locally bounded

(3)

on [0, ∞). By using the Chanturiya modulus of variation we present estimates for the rate of convergence of Kn−1,αf (x) at the points x of continuity of f and at the discontinuity points of the first kind of f. We will formulate our results for Kn−1,αf with α > 0. The corresponding estimates for the Chlodovsky-Kantorovich polynomials Kn−1f follow immediately as a special case α = 1.

At first, we give the following definition.

Definition 1.1 . Let g be a bounded function on a compact interval I = [a, b].

The modulus of variation νk(g; [a, b]) of the function g is defined for nonnegative integers k as follows:

ν0(g; [a, b]) := 0 and for k ­ 1

νk(g; [a, b]) := sup

Πk

k−1

X

j=0

|g(x2j+1) − g(x2j)| ,

where Πk is an arbitrary system of k disjoint intervals (x2j, x2j+1), j = 0, 1, ..., k −1 i.e., a ¬ x0< x1¬ x2< x3...¬ x2k−2< x2k−1¬ b ( see [4,10] ).

Some properties of the modulus of variation and its applications can be found e.g. in [4] ( see [10,13] also). In particular, νk(g; I) is a non-decreasing function of k∈ ℕ,

νk(g; I) ¬ 2k sup

t∈I|g(t)| , νk(g; Z) ¬ νk(g; I) for any compact interval Z contained in I.

Let p ­ 1. The pth power variation of a function g on a compact interval I = [a, b]

is denoted by Vp(g, I) and is defined as the upper bound of the set of numbers

X

j

|g(sj) − g(tj)|p

1/p

over all finite systems of non-overlapping intervals (sj, tj) ⊂ I. If f is of bounded pth power variation on I, then for every k∈ ℕ,

(5) νk(g; I) ¬ k1−1/pVp(g, I),

(see [4]). The class of all functions of bounded pth power variation on every compact interval contained in [0, ∞) will be denoted by BVlocp [0, ∞).

Throughout this paper we will assume that the sequence (bn) satisfies the funda- mental conditions (2). The symbols cγ and cγ,ρwill denote some positive constants, not necessarily the same at each occurrence, depending only on the indicated para- meter γ and parameters γ, ρ, respectively. The symbol [a] will denote the greatest integer not greater than a. Moreover, if x ∈ (0, ∞) the intervals Hx(u) = [x−u, x+u]

for 0 < u ¬ x will be used.

(4)

Theorem 1.2 Let f ∈ Mloc[0, ∞) and let the one-sided limits f(x+), f(x−) exist at a fixed point x ∈ (0, ∞). Then, for all integers n ­ 2 such that bn> 2x, n/bn­ N (α, x) one has

Kn−1,αf (x)− 1

2αf (x+)− (1 − 1

2α)f(x−)

¬ 2v1(gx; Hx(xp bn/n))

+cαϕn(x) x2

m−1X

j=1

vj(gx; Hx(jxp bn/n))

j3 +vm(gx; Hx(x)) m2

+cα,qM (bn; f) x2q

bn

n

q

ϕqn(x) + 4κα

rbn

n

s bn

x (bn− x)|f(x+) − f(x−)| , where m := [p

n/bn], M(b; f) := sup

0¬t¬b|f(t)| , ϕn(x) = x 1 − bxn

+bnn,

(6) gx(t) :=



f (t)− f(x+) if t > x,

0 if t = x,

f (t)− f(x−) if 0 ¬ t < x,

κα = max {1, α} , q is an arbitrary positive integer, N(α, x) = 4 if α ­ 1 and N (α, x) = max{4, 50/x} if 0 < α < 1. If α ­ 1 then cα= α and cα,q= αcq. Retaining the notation used in Theorem 1.2 and using inequality (5), we get

Theorem 1.3 Let f ∈ BVlocp [0, ∞), p ­ 1, and let x ∈ (0, ∞). Then, for all integers n­ 2 such that bn> 2x, n/bn­ N(α, x) one has

Kn−1,αf (x)− 1

2αf (x+)− (1 − 1

2α)f(x−) ¬



2 + cαϕn(x) x2



Vp(gx; Hx(xp bn/n))

+cαϕn(x) x2

rbn

n

!1+1/p [n/bn] X

k=1

Vp(gx; Hx(x/√

√ k)

k1−1/p +cα,qM (bn; f) x2q

bn

n

q

ϕqn(x)

+4κα

rbn

n

s bn

x (bn− x)|f(x+) − f(x−)| .

Here we note that in some subclasses of Mloc[0, ∞) the right-hand sides of the inequalities mentioned in the hypotheses of our theorems tend to zero as n → ∞.

In view of (3) we have m = hp n/bn

i → ∞ as n → ∞. So, in Theorem 1.2 it is enough to consider only the term

Λm(x) =

mX−1 j=1

vj(gx; Hx(jxdn))

j3 where dn =p

bn/n.

(5)

Clearly,

Λm(x) ¬

mX−1 j=1

v1(gx; Hx(jxdn)) j2 ¬ 4dn

mdZ n

dn

v1(gx; Hx(xt)) t2 dt

¬ 4dn

m+1Z

1

v1(gx; Hx(x

s))ds ¬ 4 m

Xm k=1

v1(gx; Hx(x k)).

Since the function gxis continuous at x and v1(gx; Hx(x/k)) denotes the oscillation of gxon the interval Hx(x/k), we have

klim→∞v1(gx; Hx(x k)) = 0 and consequently

mlim→∞Λm(x) = 0.

As regards Theorem 1.3, it is easy to verify that in view of continuity of gx at x,

mlim→∞

1 m1+1/p

(m+1)2

X

k=1

√ 1

k1−1/pVp(gx; Hx( x

√k)) = 0.

So, we get the following approximation theorem.

Corollary 1.4 Suppose that f ∈ Mloc[0, ∞) ( in particular, f ∈ BVlocp [0, ∞) , p­ 1) and that there exists a positive integer q such that

nlim→∞

bn

n

q

M (bn; f) = 0.

Then, at every point x ∈ (0, ∞) at which the limits f(x+), f(x−) exist, we have

n→∞limKn−1,αf (x) = 1

2αf (x+) + (1− 1

2α)f(x−).

Obviously, the above relations hold true for every measurable function f bounded on [0, ∞), in particular for every function f of bounded pth power variation (p ­ 1) on the whole interval [0, ∞).

2. Auxiliary results. We now mention certain results which are necessary to prove our main theorem. For this, let us introduce the following notation. Given any x ∈ [0, bn] and any non-negative integer q we will write

ψqx(t) := (t − x)q f or t∈ [0, ∞),

µn−1,q(x) := Kn−1ψqx(x) ≡ n bn

n−1

X

k=0

Pn−1,k

x bn

 (k+1)bnZn

kbnn

(t − x)qdt.

(6)

Lemma 2.1 If n ∈ ℕ (n ­ 2) , x ∈ [0, bn], then

µn−1,0(x) = 1, µn−1,1(x) =bn− 2x 2n , µn−1,2(x) = bn

n

n− 2 n x

 1 − x

bn

 + bn

3n



¬bn

n(x) and, for q > 1,

(7) µn−1,2q(x) ¬ cq

bn

n

q

ϕqn(x), where

ϕn(x) = x

 1 − x

bn

 +bn

n.

Proof Formulas for µn−1,0(x), µn−1,1(x), µn−1,2(x) follow by simple calculation.

Now, we will prove (7). Suppose q > 1 and put y := x/bn. Then y∈ [0, 1] and

µn−1,2q(x) = n bn

nX−1 k=0

Pn−1,k(y)

(k+1)bn

Zn

kbnn

(t − ybn)2qdt

= nb2qn

nX−1 k=0

Pn−1,k(y)

k+1

Zn k n

(s − y)2qds = b2qnKn−1ψ2qy (y) , (8)

where Kn−1 is the classical Kantorovich operator of order n − 1. Clearly, Kn−1ψ2qy (y) = d

dyBnG(y), where G is an indefinite integral of ψy, i.e.

G(y) = Zy

0

(t − y)2qdt for y ∈ [0, 1]

and BnG is the Bernstein polynomial of order n of the function G.

Hence (see [6, formula 9.5.14]), Kn−1 ψy2q(y) = 1

2q + 1 d

dyBnψy2q+1(y) + Bnψy2q(y)

= 1

n2q

 1

n (2q + 1) d

dyTn,2q+1(y) + Tn,2q(y)

 (9) ,

where

Tn,r(y) = Xn k=0

Pn,k(y) (k − ny)r, y∈ [0, 1] .

(7)

As is known (see [6, Lemma 9.4.3] or [7, Lemma 3.6]), for Tn,r(y) there holds the recursion relation

Tn,0(y) = 1, Tn,1(y) = 0,

Tn,r+1(y) = nry(1 − y)Tn,r−1(y) + y(1 − y) d

dyTn,r(y) for r ­ 2.

From this relation there follow by induction the following representation formulas

Tn,2q(y) = Xq j=1

βj,2q(n) (ny(1 − y))j,

Tn,2q+1(y) = (1 − 2y) Xq j=1

βj,2q+1(n) (ny(1 − y))j,

where βj,2q(n), βj,2q+1(n) are real numbers independent of y and bounded uniformly in n ( see [7, Corollary 3.7]).

Observing that

d

dyy(1− y) = 1 − 2y and that

(1 − 2y)2= 1 − 4y(1 − y) we easily get from (9) that

Kn−1 ψy2q(y) = 1 n2q

Xq j=0

ηj,q(n) (ny(1 − y))j,

where |ηj,q(n)| ¬ γj,q and γj,q are non-negative numbers depending only on j and q. Now, it is enough to use the following inequalities:

a ) if y ∈0,n1

or y ∈

1 − 1n, 1

, then ny(1− y) ¬ (n − 1)/n < 1 and

Xq j=0

ηj,q(n) (ny(1 − y))j ¬

Xq j=0

γj,q,

b ) if y ∈1

n, 1−n1

, then (ny(1− y))−1¬ n/(n − 1) ¬ 2 and

Xq j=0

ηj,q(n) (ny(1 − y))j

¬ (ny(1 − y))q Xq j=0

2q−jγj,q.

Consequently,

Kn−1 ψy2q(y) ¬ cq

n2q (1 + (ny(1 − y))q)

(8)

with

cq = Xq j=0

γj,q2q−j, and by (8)

µn−1,2q(x) ¬ cq

bn

n

2q 1 +nx

bn

 1 − x

bn

q

.

This inequality is equivalent to the one to be proved. Note that inequality (7) can be obtained also by using the representation formulas for Tn,2q(y) and Tn,2q+1(y) given in [6, Lemma 9.5.5].

Lemma 2.2 Let 0 < x < bn, α > 0and let

Φn−1,α(x, t) := n bn

nX−1 k=0

Q(α)n−1,k(x bn

n,k(t),

where χn,k(t)are the characteristic functions of the intervals In,k:= [kbnn,(k+1)bn n), k = 0, 1, ..., n − 1.

(i) If 0 < s < x then

(10)

Zs

0

Φn−1,α(x, t) dt ¬ καcq

(x − s)2q

bn

n

q ϕqn(x),

(ii) If x < s < bn then

(11)

bn

Z

s

Φn−1,α(x, t) dt ¬ cα,q

(s − x)2q

bn

n

q

ϕqn(x).

In the above inequalities

ϕn(x) = x

 1 − x

bn

 +bn

n,

qis an arbitrary positive integer, κα= max {1, α} . If α ­ 1 then cα,q = αcq, c1= 1.

Proof (i) Let 0 < s < x < bn. If α­ 1 then

Q(α)n−1,k

x bn



= Jnα−1,k

x bn



− Jnα−1,k+1

x bn



¬ αPn−1,k

x bn

 . Hence

Zs

0

Φn−1,α(x, t) dt ¬ αn bn

nX−1 k=0

Pn−1,k(x bn

) Zs

0

χn,k(t)dt ¬ α

(x − s)2qµn−1,2q(x).

(9)

Consider now the case 0 < α < 1. Let s ∈ [kbn/n, (k+ 1) bn/n). Then we can write s = (k+ ε) bn/n where 0¬ ε < 1 and

Zs

0

Φn−1,α(x, t) dt = n bn

nX−1 k=0

Q(α)n−1,k(x bn

) Zs

0

χn,k(t)dt

=

kX−1 k=0

Q(α)n−1,k(x bn

) + n bn

Q(α)n−1,k(x bn

)

(k+ε)bZ n/n

kbn/n

dt

=

kX−1 k=0

Q(α)n−1,k(x bn

) + εQ(α)n−1,k(x bn

)

= 1 −

n−1X

k=k

Q(α)n−1,k(x bn

) + εQ(α)n−1,k(x bn

)

= 1 − (1 − ε) Jnα−1,k(x

bn) − εJnα−1,k+1(x bn

)

¬ 1 − (1 − ε) Jn−1,k(x

bn) − εJn−1,k+1(x bn

)

=

kX−1 k=0

Q(1)n−1,k(x bn

) + εQ(1)n−1,k(x bn

)

= Zs

0

Φn−1,1(x, t) dt

¬ 1

(x − s)2qµn−1,2q(x)

(see [13,18]). Thus, estimate (10) follows by Lemma 2.1.

(ii) Let 0 < x < s < bn. If α­ 1, by using the same method in the proof of (i), one has the same bound, namely

bn

Z

s

Φn−1,α(x, t) dt ¬ α

(x − s)2qµn−1,2q(x) ¬ αcq

(x − s)2q

bn

n

q

ϕqn(x).

If 0 < α < 1 then retaining the symbol k used in the proof of (i) we obtain

(10)

bn

Z

s

Φn−1,α(x, t) dt = n bn

n−1

X

k=0

Q(α)n−1,k(x bn

)

bn

Z

(k+ε)bn/n

χn,k(t)dt

= n

bnQ(α)n−1,k(x

bn)(k+ 1) bn

n − s

+

n−1

X

k=k+1

Q(α)n−1,k(x bn)

¬ n

bn

Pn−1,kα (x bn

)(k+ 1) bn

n − s

+

n−1

X

k=k+1

Pn−1,k(x bn

)

!α

¬ 21−α n bn

Pn−1,k(x bn

)(k+ 1) bn

n − s

+

n−1

X

k=k+1

Pn−1,k(x bn

)

!α

¬ 21−α

(s − x)2qn

bn

α



Pn−1,k(

x bn)

(k∗+1)bn

Zn s

|t − x|2q/αdt

+

n−1

X

k=k+1

Pn−1,k(x bn

)

(k+1)bn

Zn

kbnn

|t − x|2q/αdt



α

¬ 21−α

(s − x)2q mn−1,2q/α(x)α

,

where

mn−1,γ(x) = n bn

n−1

X

k=0

Pn−1,k(x bn

)

(k+1)bn

Zn

kbnn

|t − x|γdt.

Write l = 2/α and denote by [l] the integral part of l. As in [18, Lemma 6] choose the numbers

p = 2[l]

2[l] + 2 − l, p0 = 2[l]

l− 2, r = 2

p, r0= 2v

p0, v = [l] + 1.

Clearly, l > 2, p > 1, p0 > 1, 1/p + 1/p0 = 1 and l = r + r0. Applying twice the

(11)

H¨older inequality (first for integrals, next for sums) we obtain

mn−1,2q/α(x) ¬ n bn

n−1X

k=0

Pn1/p−1,k(x bn

)Pn1/p−1,k0 (x bn

)

(k+1)bn

Zn

kbn n

|t − x|qr|t − x|qr0dt

¬

bnn

n−1

X

k=0

Pn−1,k(x bn

)

(k+1)bn

Zn

kbnn

|t − x|qrpdt



1/p

bnn

n−1

X

k=0

Pn−1,k(x bn)

(k+1)bn

Zn kbn

n

|t − x|qr0p0dt



1/p0

= n−1,2q(x))1/pn−1,2vq(x))1/p0.

Applying Lemma 2.1 and observing that 1/p + v/p0= 1/α we get

mn−1,2q/α(x) ¬

 cq

bn

n

q

ϕqn(x)

1/p cvq

bn

n

vq

ϕvqn (x)

1/p0

= cα,q

bn

n(x)

q/α

.

Thus inequality (11) follows.

In order to formulate the next lemma we introduce the following intervals. If x > 0 we will write

Ix(u) := [x + u, x] ∩ [0, ∞) if u < 0 Ix(u) := [x, x + u] if u > 0.

Lemma 2.3 Let f ∈ Mloc[0, ∞) and let at a fixed point x ∈ (0, ∞) the one-sided limits f(x+), f(x−) exist. Consider the function gx defined by (6) and write dn :=

pbn/n. If h = −x or h = x, then for all integers n such that n/bn­ 4 we have

Z

Ix(h)

gx(t)Φn−1,α(x, t) dt

¬ v1(gx; Ix(hdn))

+cαϕn(x) x2

m−1X

j=1

vj(gx; Ix(jhdn))

j3 +vm(gx; Ix(h)) m2

! ,

where m := [p

n/bn], ϕn(x) = x (1 − x/bn) + bn/n.If α ­ 1 then cα= 8α.

(12)

Proof Restricting the proof to h = −x we define the points tj = x + jhdn for j = 0, 1, 2, ..., m and we denote tm+1= 0. Put Tj= [tj, x] for j = 1, 2, ..., m + 1 and write

Z

Ix(h)

gx(t)Φn−1,α(x, t) dt = Zx

t1

gx(t)Φn−1,α(x, t) dt + Xm j=1

gx(tj)

tj

Z

tj+1

Φn−1,α(x, t) dt

+ Xm j=1

tj

Z

tj+1

(gx(t) − gx(tj)) Φn−1,α(x, t) dt

= : I1(n, x) + I2(n, x) + I3(n, x), say.

Clearly,

|I1(n, x)| ¬ Zx

t1

|gx(t) − gx(x)| Φn−1,α(x, t) dt

¬ v1(gx; T1)

bn

Z

0

Φn−1,α(x, t) dt = v1(gx; T1).

By the Abel lemma on summation by parts and by (10) (with q = 1) we have

|I2(n, x)| ¬ |gx(t1)|

t1

Z

0

Φn−1,α(x, t) dt

+

mX−1 j=1

|gx(tj+1) − gx(tj)|

tj+1

Z

0

Φn−1,α(x, t) dt

¬ καϕn(x) x2

|gx(t1) − gx(x)| +

mX−1 j=1

|gx(tj+1) − gx(tj)| 1 (j + 1)2

= καϕn(x) x2

|gx(t1) − gx(x)| +

mX−2 j=1

Xj k=1

|gx(tk+1) − gx(tk)|

 1

(j + 1)2 1 (j + 2)2

 + 1

m2 Xm k=1

|gx(tk+1) − gx(tk)|

!

¬ καϕn(x) x2

v1(gx; T1) + 2

m−2X

j=1

vj+1(gx; Tj+1)

(j + 1)3 +vm(gx; Tm) m2

¬ καϕn(x) x2

2

mX−1 j=1

vj(gx; Tj)

j3 +vm(gx; Tm+1) m2

 .

(13)

Next, in view of (10) (with q = 1) and the Abel transformation,

|I3(n, x)| ¬ Xm j=1

v1(gx; [tj+1,tj])

tj

Z

tj+1

Φn−1,α(x, t) dt

¬ καϕn(x) x2

Xm j=1

v1(gx; [tj+1,tj]) j2

= καϕn(x) x2

Xm k=1

v1(gx; [tk+1,tk]) m2

+

mX−1 j=1

Xj k=1

v1(gx; [tk+1,tk])

1

j2 1 (j + 1)2



¬ καϕn(x) x2

vm(gx; Tm+1)

m2 + 6

m−1X

j=1

vj(gx; Tj+1) (j + 1)3

¬ καϕn(x) x2

vm(gx; Tm+1)

m2 + 6

Xm j=2

vj(gx; Tj) j3

 .

Collecting the results and observing that Tj = Ix(jhdn) we get the desired estimate for h = −x. In case h = x the proof runs analogously; we use inequality (11) instead

of (10).

Lemma 2.4 Let α > 0 and let

sgn(α)x (t) :=



2α− 1 if t > x,

0 if t = x,

−1 if 0 ¬ t < x.

Then

Kn−1,αsgn(α)x (x) ¬ 2α

n−1X

k=k0+1

Pn−1,k(x bn

)

!α

1 2α

+ 2αQ(α)n−1,k(x bn

),

where k0= [(n − 1)x/bn], k= [nx/bn].

Proof Since x ∈ [kbn/n, (k+ 1) bn/n) we have

Kn−1,αsgn(α)x (x) = (2α− 1)

n−1

X

k=k+1

Q(α)n−1,k(x bn) −

kX−1 k=0

Q(α)n−1,k(x bn

)

+n bn

Q(α)n−1,k(x bn

)



(2α− 1)

(k+ 1) bn

n − x





x−kbn

n



,

(14)

i.e.

Kn−1,αsgn(α)x (x) = 2α

nX−1 k=k+1

Q(α)n−1,k(x

bn) − 1 + 2αQ(α)n−1,k(x bn

)n bn

(k+ 1) bn

n − x

 . The above identity can be rewritten in the form

Kn−1,αsgn(α)x (x) = 2α

n−1

X

k=k

Q(α)n−1,k(x

bn) − 1 − 2αQ(α)n−1,k(x bn

)



x−kbn

n

 .

Observing that k= k0 or k= k0+ 1 and that

nX−1 k=v

Q(α)n−1,k(x bn

) =

nX−1 k=v

Pn−1,k(x bn

)

!α

we easily get the estimate for Kn−1,αsgn(α)x (x) as desired.

Lemma 2.5 Let x > 0, α > 0 and let k0 = [(n − 1)x/bn]. Then

n−1X

k=k0+1

Pn−1,k

x bn

!α

1 2α

¬√

α

rbn

n

s bn

x (bn− x),

where κα = max {1, α} and n ­ 2 is so large that bn > 2x, n/bn > n0(α, x) with n0(α, x) = 0if α ­ 1, n0(α, x) = 21/x if 0 < α < 1.

Proof Let y = x/bn. Then k0 = [(n−1)y]. In view of the well-known Berry-Ess´een theorem

X

ny<k¬n

Pn,k(y) − 1 2

¬ 0.8 pny(1− y) (see e.g. [17, Lemma 2]). Consequently, assuming α ­ 1 we have

n−1X

k=k0+1

Pn−1,k

x bn

!α

1 2α

¬ α

X

k>(n−1)y

Pn−1,k(y) − 1 2

¬ 0.8α

p(n − 1) y(1 − y)¬√

rbn

n

s bn

x (bn− x). If 0 < α < 1 then

 X

ny<k¬n

Pn,k(y)

α

1 2α

¬ 1 pny(1− y)

(15)

for all n > 25y(1−y)256 (see [18, Lemma 1]). Hence

nX−1 k=k0+1

Pn−1,k(y)

!α

1 2α

=

 X

k>(n−1)y

Pn−1,k(y)

α

1 2α

¬ 1

p(n − 1) y(1 − y)¬√ 2

rbn

n

s bn

x (bn− x) for all n such that n − 1 > 25y(1−y)256 , i.e. for all n such that bn > 2x and

n/bn > 21/x.

Lemma 2.6 Let x > 0, α > 0 and let k= [nx/bn].Then

Q(α)n−1,k

x bn



¬ κα rbn

n

s bn

x (bn− x),

where κα = max {1, α} and n ­ 2 is so large that bn > 2x, n/bn > n1(α, x) with n1(α, x) = 0if α ­ 1, n1(α, x) = 50/x if 0 < α < 1.

Proof If α ­ 1 then Q(α)n−1,k(x/bn) ¬ αPn−1,k(x/bn) . The desired estimate follows with κα= α by the known inequality

(12) Pn,k(y) ¬ 1

√2ep

ny(1− y) f or y∈ (0, 1) , k ∈ ℕ (see [16, Theorem 1]). Indeed, putting y = x/bn we have

Q(α)n−1,k

x bn



¬ αPn−1,k(y) ¬ α

√2ep

(n − 1) y(1 − y) < α rbn

n

s bn

x (bn− x). Suppose now 0 < α < 1. Using the mean-value theorem we get

Q(α)n−1,k(y) = αγnα−1,k−1 (y) (Jn−1,k(y) − Jn−1,k+1(y))

= αγnα−1,k−1 (y)Pn−1,k(y)

where Jn−1,k+1(y) < γn−1,k(y) < Jn−1,k(y) . Denoting k0 = [(n − 1) y] we have k= k0 or k = k0+ 1.

If k= k0 then

γn−1,k(y) > Jn−1,k0+1(y) = X

k>(n−1)y

Pn−1,k(y) > 1 4

for all n such that n − 1 > 25y(1−y)256 (see [18, inequality (12)]).

Cytaty

Powiązane dokumenty

The next two results show that certain smaller radial growths can be associated with suitable exceptional sets..

In [6] Hibschweiler and MacGregor investigated membership in F α for univalent functions, in particular starlike and convex functions, with restricted growth.. The following result

In this paper we establish an estimation for the rate of pointwise convergence of the Chlodovsky-Kantorovich polynomials for functions f locally integrable on the interval [0, ∞).

120, which proves the ex- istence of cases of equality in Davenport’s bound, is completely contained in Stothers’s results, who actually finds a remarkable exact formula for the

Indeed the solutions of such problems are defined via linear combinations of Mellin convolution operators depending on a parameter, and the study of the approximation of these

In Section II we discuss the properties of this type of functions; Section III is devoted to the notion of characteristic interval and an extension theorem; finally, in Sec- tion

It is known that each simply connected domain, included in the disk U , symmetric about the real axis and containing 0, can be ap- proximated, in the sense of convergence towards

Generic and dense chaos have been characterized respectively in [Sn1] and [Sn2], and it is known that dense chaos need not imply generic chaos (cf.. In [Sn3], the definitions of