• Nie Znaleziono Wyników

For this case Erd˝os and Tur´an [4] proved the following upper bound: (1.1) DN(F, xn

N/A
N/A
Protected

Academic year: 2021

Share "For this case Erd˝os and Tur´an [4] proved the following upper bound: (1.1) DN(F, xn"

Copied!
8
0
0

Pełen tekst

(1)

LXXXV.4 (1998)

On the Erd˝os–Tur´an inequality for balls

by

Glyn Harman (Cardiff)

1. Introduction. Let (Ω, M, µ) be a probability space. Given a family of sets F ⊆ M, we define the discrepancy of a set of points {x1, . . . , xN} ⊆ X with respect to F by

DN(F, xn) = sup

A∈F

X

xn∈A n≤N

1 − N µ(A) .

The most well known case of this has Ω = T = R/Z, µ(·) the usual Haar measure (equivalent to Lebesgue measure on [0, 1)) and F the family of intervals (mod 1). For this case Erd˝os and Tur´an [4] proved the following upper bound:

(1.1) DN(F, xn) ≤ C1N L + C2

XL m=1

1 m

XN n=1

e(mxn)

for all positive integers L, where e(y) = exp(2πiy). Here C1 and C2 are absolute, and the best known values are C1= 1, C2= 2 + 2/π (see [1], p. 20, or [10], p. 8). The Erd˝os–Tur´an theorem was generalized independently to Tkby Koksma [9] and Sz¨usz [13]. They took F to be the family of boxes with sides parallel to the axes (see also [3] and [6], §5.4). It is possible to prove an analogue when F is the family of all boxes. As is well known, allowing the boxes to be “tilted” greatly increases the discrepancy in general (see [6],

§5.4, [2] and [10]).

In [7] Holt has given an Erd˝os–Tur´an inequality for balls. This was based on an approximation to the characteristic function of a ball which is in some sense optimal [8]. It is the purpose of this note to prove a stronger version of this result. It may seem paradoxical that we will not employ the “optimal”

functions used by Holt. We will explain below the difference between Holt’s work and our own. Suffice it to say now that his approach optimized the analogue of C1 above, but gave the wrong size coefficients in the analogue

1991 Mathematics Subject Classification: Primary 11K38.

[389]

(2)

of the sum over m. We will not try to optimize C1 and C2, but we will get the correct order of diminution for the coefficients. If we write B(r) for the family of balls with radius r, Holt’s result is as follows.

Theorem 1. Let r ≥ 0 and {x1, . . . , xN} ⊂ Tk. Then, for all s > 0, (1.2) DN(B(r), xn) ≤ N Ak(r, s)

+ X

0<|m|<s



Ak(r, s) +

 r

|m|

k/2

|Jk/2(2π|m|)|



XN n=1

e(m · xn) . In the above Jν denotes the Bessel function of order ν, and, when rs > 1,

Ak(r, s)  rk−1s−1.

The fault with (1.2) is that the coefficients involving m should have size

 r

|m|

k/2

|Jk/2(2π|m|)|  min



rk, r(k−1)/2

|m|(k+1)/2

 .

Our main result establishes this. We will write Cj(t), Cj(a, b), etc. for con- stants depending only on given parameters (t, a, b in the examples quoted).

Theorem 2. Under the hypotheses of Theorem 1 we have (1.3) DN(B(r), xn) ≤ C1(k)N (rk−1s−1+ s−k)

+ C2(k) X

0<|m|<s

 1

sk + min



rk, r(k−1)/2

|m|(k+1)/2



XN n=1

e(m · xn) . In particular , writing C for the family of all balls of radius less than one half , we have

(1.4) DN(C, xn) ≤ C3(k)N

s + C4(k) X

0<|m|<s

1

|m|(k+1)/2

XN n=1

e(m · xn) . We state two applications of the above result as further theorems.

Theorem 3. Let p be a prime number. Then there is a lattice point h ∈ Zk such that |h| < p and the sequence xn = nh/p satisfies, for all r ∈ (0, 1),

(1.5) Dp(B(r), xn)  (prk)(k−1)/(k+1)+ 1.

Theorem 4. For almost all α ∈ Tk we have (1.6) DN(C, nα)  N1−2/(k+1)log2N.

We note that in (1.5) we have improved Holt’s exponent (k − 1)/k (The- orem 2 of [7]). We have also relaxed the condition rkp > 1, but this only involved the addition of the term 1 in (1.5). We remark that the correspond- ing bounds for tilted boxes have the form N1−1/k, a result which is known

(3)

to be best possible (see [14] and §5.4 of [6]). The author has not been able to determine whether (1.5) and (1.6) are quite sharp in the case k = 2 (exponent 1/3), or whether the universal lower bound for (1.6)

DN(C, xn)  N1/4 may be nearer the truth.

2. Preparatory lemmas. A real even entire function is an entire func- tion which is even and only takes real values on the real axis. Let E(s) be the set of all such functions which satisfy

f (z) = O(exp(2πs|z|)) as |z| → ∞.

Lemma 1. Let r, s > 0. Then there are functions F1(z), F2(z) ∈ E(s) such that

(2.1) F1(x) ≤ χ(x) ≤ F2(x), x ∈ R,

where χ(x) is the characteristic function of the interval [−r, r], and (2.2) Fj(x) = χ(x) + O

 min



1, 1

|s(x − r)|2



, where the constant implied by the O notation is absolute.

P r o o f. See [8], Chapter 1 of [10], or Chapter 2 of [1]. Usually the close- ness of approximation of Fj(x) to χ(x) is given by calculating

\

−∞

(F2(x) − F1(x)) dx or

\

−∞

|Fj(x) − χ(x)| dx.

The estimate (2.2) will prove more helpful in our context for reasons which should become clear.

Lemma 2. Let r, s > 0 and k ∈ N. Then there are functions G1(z), G2(z)

∈ E(s) such that

(2.3) G1(x) ≤ χ(x) ≤ G2(x)

and

(2.4) |Gj(x) − χ(x)| ≤ C(k)h(x), where

(2.5) h(x) = min(1, |s(x − r)|−2j(x,r)), j(x, r) =

k if |x| ≥ r, 1 if |x| < r.

P r o o f. We appeal to Lemma 1 with s replaced by s/k. We take G2(x) = F2(x)k, which clearly satisfies all our demands. For odd k we can take

(4)

G1(x) = F1(x)k. For k even, either replace k by k + 1 (since the truth of (2.5) with k + 1 implies its truth for k), or take

G1(x) = kF1(x)F2(x)k−1− (k − 1)F2(x)k.

See Lemma 6 of [5] for the motivation and proof that this is a lower bound function.

Remark. The advantage of the functions constructed above is that the errors are given a suitable bound both for small and large values of x, unlike the situation in [8] where the term of importance is

\

0

|Fj(x) − χ(x)|xudx,

and the xu factor cancels out the errors for small x. To obtain our results, essentially we will need to estimate such integrals for two different values of u, as becomes clear in the next lemma.

Lemma 3. Let r, s > 0 and k ∈ N. Let G1(x), G2(x) be the functions given by Lemma 2. Then, for all m with 1 ≤ m ≤ 2k − 1, and all v with 0 < v ≤ s, we have

(2.6)

\

0

|Gj(x) − χ(x)| min

 xm−1,

x v

(m−1)/2 dx

≤ Cj(k, m)

 1

sm + min



rm,r(m−1)/2 v(m+1)/2



for j = 1, 2. Moreover , (2.7)

\

0

|Gj(x) − χ(x)|xm−1dx ≤ Cj+2(k, m)

 1

sm +rm−1 s

 .

P r o o f. These results follow from Lemma 2 after splitting [0, ∞) into appropriate regions. For example, if v > 1/r we have to consider, for (2.6), the four integrals (assuming v−1+ s−1< r for clarity)

1/v\

0

xm−1 s2(r − x)2dx,

r−1/s\

1/v

x v

(m−1)/2 1

s2(r − x)2dx,

r+1/s\

r−1/s

x v

(m−1)/2 dx,

\

r+1/s

x v

(m−1)/2 1

s2k(x − r)2k dx.

The reader will have no difficulty in verifying that each of these integrals has the right size.

(5)

3. Proof of Theorem 2. Theorem 2 follows immediately from the following lemma which may have other applications. We write Vk for the volume of a ball belonging to B(1) in Rk.

Lemma 4. Let B ∈ B(r) in Tk, and let χ(x) be its characteristic function.

Let s > 0. Then there are functions H1(x), H2(x) such that

(3.1) H1(x) ≤ χ(x) ≤ H2(x)

and

(3.2) Hj(x) = Vkrk+ O

1

sk +rk−1 s



+ X

0<|m|<s

cj(m)e(m · x),

where

(3.3) |cj(m)|  1

sk + min



rk, r(k−1)/2

|m|(k+1)/2

 .

P r o o f. We deal with the construction of H1(x) only; the choice of H2is analogous. Also, without loss of generality, we can suppose that B is centred at the origin: this only affects cj(m) by a factor e(−m · a), where a is the centre of the ball. We pick G1(x) from Lemma 2, and let χ(x) be as defined in Lemma 1. There should be no confusion in the following between χ(x) the characteristic function of the ball, and χ(x) the characteristic function of the interval. Since G1 is an even entire function we may write

G1(x) = X n=0

cnx2n.

Also, if we write z = (z1, . . . , zk), the function U (z) =

X n=0

cn(z12+ . . . + zk2)n is an entire function of k complex variables, and

(3.4) |U (z)| = O(exp(2πskzk))

where

kzk = sup

x∈S

|z · x|,

and S ∈ B(1) is centred at the origin (see [8] for details).

Write

U (y) =b \

Rk

U (x)e(−x · y) dx.

Then, using (3.4) with Theorem 4.9 of Chapter 3 in [12], we obtain bU (y) = 0

(6)

for |y| ≥ s. Also U (0) =b \

Rk

U1(x) dx = kVk

\

0

G1(t)tk−1dt (3.5)

= kVk

\

0

χ(t)tk−1dt −

\

0

(χ(t) − G1(t))tk−1dt

= Vkrk+ O

rk−1 s + 1

sk



by Lemma 3. Moreover, U (v) =b \

Rk

χ(x)e(−x · v) dx − \

Rk

(χ(x) − U (x))e(−x · v) dx

= I1(v) − I2(v) say.

Now, as is well known (it follows from Theorem 3.3 of [12], Chapter 4, for example),

(3.6) I1(v) =

 r

|v|

k/2

Jk/2(2π|v|r).

Also, since χ and G1 are both radial functions, we can apply Theorem 3.3 of [12], Chapter 4, to I2 to obtain

(3.7) I2(v) = 2π|v|1−k/2

\

0

(χ(x) − G1(x))Jk/2−1(2π|v|x)xk/2dx.

Since, for u > 0,

Ju(x) ≤ C(u) min(x−1/2, xu) for x > 0, (3.6) and (3.7) with Lemma 3 (with m = k say) give, for |v| > 0, (3.8) | bU (v)| 

 1

sk + min



rk, r(k−1)/2

|v|(k+1)/2



. Hence, if we write

H1(x) =X

n∈Z

U (x + n),

we have (3.1)–(3.3) as desired by the k-dimensional Poisson summation for- mula (see [12], pp. 251–252, for example).

4. Proof of Theorem 3. By the argument on page 65 of [7] there is a lattice point h with |h| < p, and, using our Theorem 2 in place of Holt’s

(7)

Theorem 1, we obtain

Dp(B(r), xn)  ps−1rk−1+ X

|m|<s

min



rk, r(k−1)/2

|m|(k+1)/2

 (4.1)

 ps−1rk−1+ 1 + (rs)(k−1)/2

after an easy calculation. If we then take the optimal choice s = (p2rk−1)1/(k+1)

to balance the first and last terms in (4.1) we thereby obtain (1.5).

5. Proof of Theorem 4. We note that, if k · k now denotes distance to a nearest integer, then

X

n≤N

e(nα · m) ≤ min



N, 1

kα · mk

 .

If m = (m1, . . . , mk), write d(m) =

Yk j=1

max(1, |mj|).

By Theorem 2 of [11] (applied in 1, . . . , k dimensions), we have, for almost all α and all ε > 0,

X

0<|m|<s

(d(m)kα · mk)−1 (log s)k+1+ε. From this it easily follows that, for almost all α,

X

0<|m|<s

(|m|(k+1)/2kα · mk)−1 s(k−1)/2(log s)k+1+ε. Theorem 4 then follows from Theorem 2 with the choice

s = N2/(k+1)(log N )−ε−2.

References

[1] R. C. B a k e r, Diophantine Inequalities, Oxford Univ. Press, 1986.

[2] J. B e c k and W. W. L. C h e n, Irregularities of Distribution, Cambridge Univ. Press, 1987.

[3] T. C o c h r a n e, Trigonometric approximation and uniform distribution modulo 1, Proc. Amer. Math. Soc. 103 (1988), 695–702.

[4] P. E r d ˝o s and P. T u r ´a n, On a problem in the theory of uniform distribution, I , Indag. Math. 10 (1948), 370–378.

[5] G. H a r m a n, Small fractional parts of additive forms, Philos. Trans. Roy. Soc.

London Ser. A 345 (1993), 339–347.

(8)

[6] G. H a r m a n, Metric Number Theory, Oxford Univ. Press, 1998.

[7] J. J. H o l t, On a form of the Erd˝os–Tur´an inequality, Acta Arith. 74 (1996), 61–66.

[8] J. J. H o l t and J. D. V a a l e r, The Beurling–Selberg extremal functions for a ball in Euclidean space, Duke Math. J. 83 (1996), 203–248.

[9] J. F. K o k s m a, Some theorems on Diophantine inequalities, Math. Centrum Ams- terdam Scriptum no. 5.

[10] H. L. M o n t g o m e r y, Ten Lectures on the Interface Between Analytic Number The- ory and Harmonic Analysis, Amer. Math. Soc., Providence, R.I., 1994.

[11] W. M. S c h m i d t, Metrical theorems on fractional parts of sequences, Trans. Amer.

Math. Soc. 110 (1964), 493–518.

[12] E. M. S t e i n and G. W e i s s, Introduction to Fourier Analysis on Euclidean Spaces, Princeton Univ. Press, Princeton, N.J., 1971.

[13] P. S z ¨u s z, ¨Uber ein Problem der Gleichverteilung, in: Comptes Rendus du Premier Congr`es des Math´ematiciens Hongrois, 1950, 461–472.

[14] S. K. Z a r e m b a, Good lattice points in the sense of Hlawka and Monte Carlo inte- gration, Monatsh. Math. 72 (1968), 264–269.

School of Mathematics Cardiff University P.O. Box 926 Cardiff CF2 4YH Wales, U.K.

E-mail: Harman@cf.ac.uk

Received on 15.12.1997 (3314)

Cytaty

Powiązane dokumenty

We also prove that the affirmative answer to Wall’s question implies the first case of FLT (Fermat’s last theorem); from this it follows that the first case of FLT holds for

The proofs of these results depend on the method of Roth and Halber- stam on difference between consecutive ν-free integers, the results of Baker [1] on the approximations of

In the proofs of Lemmas 1 and 2 the authors combined various powerful results and methods in number theory, including Wiles’ proof of most cases of the Shimura–Taniyama

We note that, at first glance, the results Carlitz achieves in [1] do not appear to be the same as Theorem 1 with α = 1.. It can be checked, however, that they are

In 1842 Dirichlet proved that for any real number ξ there exist infinitely many rational numbers p/q such that |ξ−p/q| &lt; q −2.. This problem has not been solved except in

In this way (Kohnen–) Kloosterman sums for weight 1/2 come into play by using the discrete Fourier transform identity (Theorem A below), and we will have a possibility of

(A similar application is given in [8], pp. 154–157.) Although our result is not as sharp as Beck’s, our proof is much simpler and we do not have the requirement that r ≤ 1.. We use

(c) For positive diagonal quaternary forms Halmos [3] (with a final touch added by Pall [5]) found all near misses, relative to all positive integers.. (d) I have additional