• 沒有找到結果。

and the packing dimension of the intersection E ∩ G, where G is an arbitrary Borel set on the circle

N/A
N/A
Protected

Academic year: 2022

Share "and the packing dimension of the intersection E ∩ G, where G is an arbitrary Borel set on the circle"

Copied!
17
0
0

加載中.... (立即查看全文)

全文

(1)

Bing Li, Narn-Rueih Shieh, and Yimin Xiao

Abstract. Let E be the Dvoretzky random covering sets on the circle. By applying the method of limsup type random fractals, as illustrated in Khosh- nevisan, Peres and Xiao [24], we determine the hitting probability P(E∩G 6= ∅) and the packing dimension of the intersection E ∩ G, where G is an arbitrary Borel set on the circle.

Introduction

We begin with a brief review on random coverings. Let {ωn}n≥1be a sequence of independent random variables on (Ω, B, P) which are uniformly distributed over the unit interval I = [0, 1). Let {ln}n≥1 be a sequence of positive real numbers which is decreasing to zero. For every n ≥ 1, denote by In := (ωn, ωn+ ln)(mod 1) the random interval whose starting point and lengthe are determined by ωn and ln

respectively. Define the random covering set as E := lim sup

n→∞

In= {t ∈ I : t ∈ In for infinitely many n ≥ 1}.

The set E consists of the points which are covered by {In} infinitely often (i.o. for short). The Borel-Cantelli Lemma implies that the Lebesgue measure of the random set E is either 1 or 0 almost surely according to the divergence or convergence of the seriesP

n=1ln.

It was Dvoretzky [5] who called the attention on study of such random sets; he raised the question that under what condition on {ln} one can have

(0.1) [0, 1) = lim sup

n→∞

In a.s.

In the literature this is referred to as the Dvoretzky covering problem and had attracted the attention of P. Billard, J.-P. Kahane, B. Mandelbrot, among others, before it was completely solved by L. A. Shepp in 1972. Shepp [31] provided a necessary and sufficient condition for (0.1) to hold, namely

(0.2)

X

n=1

1 n2exp

l1+ · · · + ln



= ∞.

1991 Mathematics Subject Classification. 60D05, 52A22, 28A78, 28A80.

Key words and phrases. Random covering sets, Dvoretzky random covering, hitting proba- bility, packing dimension, Hausdorff dimension, limsup random fractals.

Research supported in part by NSF of China Grant #11201155.

Research supported in part by NSF grant DMS-1006903.

1

(2)

Since then the topic has been under active development and there have been many extensions and refinements. We refer to [20, Chapter 11] for a systematic account on the Dvoretzky covering problem and its higher dimensional extensions, to the survey articles [8, 22, 23] for historical accounts and connections to multiplicative processes, and to [1, 4, 7, 8, 9, 10, 14, 21] and the references therein for further information. It should also be mentioned that Jonasson and Steif [16] (see also [15]) have recently extended the Dvoretzky covering model by including time dynamics (In the first variant, they identify I = [0, 1) with the unit circle C and allow the centers of In(n ≥ 1) perform independent Brownian motions on C, each with variance 1. In the second variant, they associate independent Poisson processes with the different intervals.) The work of Jonasson and Steif [16] has revealed rich structures in dynamical random coverings and raised more interesting questions about properties of the random covering sets, including their fractal dimensions and hitting probabilities.

This paper is concerned with the geometric and potential-theoretic properties of the Dvoretzky covering set E = lim sup

n→∞

In. It is known that the set E is a.s. dense in I and is of second category ([20, Chapter 5, Proposition 11]). Thus, the upper box dimension of the set E is 1 almost surely. Several authors have investigated the Hausdorff dimension and other fractal properties of E and/or its complement F = I\E (which is called the uncovered set). For example, Fan and Wu [10]

considered the Hausdorff dimension of the set E for the special case ln =naγ, where a > 0 and γ > 1 are constants, they proved that dimH(E) = 1γ a.s., where dimH denotes Hausdorff dimension. Durand [4] considered a general sequence {ln} with P

n=1ln< ∞ and proved, among other things,

(0.3) dimHE = α and dimPE = 1 a.s., where α is defined by

(0.4) α := inf

 s > 0 :

X

n=1

lsn< ∞



= sup

 s > 0 :

X

n=1

lsn= ∞



with the convention that sup ∅ = 0 and inf ∅ = 1.

The index α defined in (0.4) is known as the exponent of convergence of the sequence {ln} and can be calculated by using the following formula

(0.5) α = lim sup

n→∞

log n

− log ln

(see [27] p.285 or [30] p.26). Besicovitch and Taylor [3] applied the index α (they also introduced another index–the lower index for {ln}) to characterize the Haus- dorff measure and Hausdorff dimension of a linear compact set K whose complement forms a sequence of open intervals of lengths {ln}. Hawkes [13] showed that α is the upper box dimension of K, and the lower index of {ln} is the lower box dimension of K. Kahane [18, 19] called α the upper Besicovitch-Taylor index of {ln}. Some related indices for {ln} were also discussed in [1] for studying the Carleson problem and covering numbers for the Dvoretzky covering set E.

The following intersection problem is of intrinsic importance in the study of random coverings and other random fractals. For any given set A ⊂ [0, 1), we can ask whether or not it is a.s. covered infinitely often by {In}. That is, when

(3)

does P(A ⊂ E) = 1 hold? In the case

P

n=1

ln = ∞, which is opposite to what we are considering in this paper, the analogous problem for the uncovered set Fhas been investigated by several authors. For example, Kahane [20] considered the case ln =βn, 0 < β < 1 and showed that P(A ⊂ E) = 1 (equivalently, P A∩F6= ∅ = 0) if dimH(A) < β, whilst P(A ⊂ E) = 0 (equivalently, P A ∩ F 6= ∅

= 1) if dimH(A) > β. For a more general case, Hawkes [12] proved that, if the set A satisfies a regularity condition (which in particular requires dimH(A) = dimP(A)), then P(A ⊂ E) = 1 or 0 according as dimH(A) < τ and dimH(A) > τ , where τ is the index of {ln} defined by

τ = lim sup

n→∞

Pn i=1li

log n .

More precise hitting probability results for the Poisson covering (see Mandelbrot [26]) have been established by using the connection between Fand the range of a subordinator; see Fitzsimmons, et al. [11]. However, the problems for determining the hitting probabilities of the Dvoretzky random covering set E had never been studied.

The purpose of this paper is to study the hitting probabilities of the Dvoretzky covering set E, as well as fractal dimensions of the intersection E ∩ G, when it is nonempty. Our main result (Theorem 2.1 below) shows that the hitting probability P(E ∩ G 6= ∅) is determined by dimP(G), the packing dimension of G (see (0.7) below). This is in contrast with the hitting probability results for the random set F, where Hausdorff dimension plays the natural role. Theorem 2.1 will allow us to determine the packing dimension of E ∩ G for any analytic set G ⊆ [0, 1) and provide a refinement (under an extra condition) of (0.3) obtained by Durand [4].

Recall that packing dimension was introduced in the early 1980s by Tricot [33]

as follows. For any ε > 0 and any bounded set G ⊂ R, let N (G, ε) be the smallest number of balls of radius ε needed to cover G. The upper box dimension of G is defined as

(0.6) dimM(G) = lim sup

ε→0

log N (G, ε)

− log ε and the packing dimension of G is defined as

(0.7) dimP(G) = inf

 sup

n

dimMGn : G ⊂

[

n=1

Gn

 ,

where the infimum is taken over all countable coverings {Gn} of G. It is well known that 0 ≤ dimH(G) ≤ dimP(G) ≤ dimM(G) ≤ 1 for every set G ⊂ R. Similarly to Hausdorff dimension, packing dimension has been shown to be a useful tool for characterizing fractal sets and for studying “roughness” of stochastic processes. We refer to Falconer [6] and Mattila [28] for further properties of packing dimension and to Taylor [32] and Xiao [34] for extensive surveys on its applications to random fractals.

The rest of this paper is organized as follows. In Section 2 we state the main results and provide some discussions and examples. The proofs of the theorems are given in Section 3 and they rely on the general method on limsup random fractals in Khoshnevisan, Peres and Xiao [24]. We remark that our argument extends that in [24] and shows that their Theorems 3.1, 3.3 and corollaries still hold if their

(4)

Condition 4 is replaced by a weaker condition. Finally Section 4 contains some technical results on the upper Besicovitch-Taylor index. In particular we apply the results in Lapidus and van Frankenhuysen [25] to show that every sequence {ln} associated to a self-similar string has its upper Besicovitch-Taylor index equal to the self-similarity dimension and satisfies the condition (C) in this paper.

1. Main results and examples Throughout this paper we assumeP

n=1ln < ∞. Thus the Lebesgue measure of the Dvoretzky covering set E is 0 almost surely.

Let α be the upper Besicovitch-Taylor index of {ln}. Then by Proposition 3.1 below, we have

(1.1) α = lim sup

k→∞

log2nk

k , where log2is the logarithm in base 2 and nk is defined as

nk= #n

n ∈ N : ln ∈ [2−k+1, 2−k+2)o

(k ≥ 2).

Here #A denotes the cardinality of the set A.

To state the main results of this paper we will make use of the following con- dition (C):

(C) There exists an increasing sequence of positive integers {ki} such that

(1.2) lim

i→∞

ki+1

ki = 1 and

(1.3) lim

i→∞

log2nki

ki

= α < 1.

Theorem 1.1. Let E be the Dvoretzky covering set associated with the sequence {ln} whose upper Besicovitch-Taylor index is α. If the condition (C) holds, then for every analytic set G ⊂ [0, 1), we have

P E ∩ G 6= ∅ =

(0 if dimP(G) < 1 − α, 1 if dimP(G) > 1 − α.

Remark 1.2. Some remarks are in order.

(i) It is clear that if

(1.4) lim

k→∞

log2nk

k = α < 1,

then condition (C) holds. We will give several interesting examples of sequences {ln} that satisfy (1.4).

(ii) If G is regular in the sense that dimM(G) = dimM(G), where dimM(G) is the lower box dimension of G, which is defined by replacing limsup in (0.6) by liminf, then condition (C) is surplus. This follows from the proof of Theorem 1.1 below, in which we can take N = {ki0, ki0+1, . . . } for some i0≥ 1.

(iii) From the first part of the proof of Theorem 1.1, we see that the conclusion dimP(G) < 1 − α implies P(E ∩ G 6= ∅) = 0, even without the condition (C).

(5)

(iv) By Proposition 3.3 in Section 4, we see that (1.4) can be replaced by the following: there exists a constant b ∈ (1, 2] such that

(1.5) lim

k→∞

logbmk

k = α < 1, where mk = #n ∈ N : ln∈ [b−k+1, b−k+2) . (v) When P

n=1ln = ∞, but Shepp’s condition (0.2) is not satisfied, then E 6= [0, 1). One can consider the random set F = [0, 1)\E of the un- covered points. The fractal dimension and hitting probabilities have been studied by Hawkes [12] (see also Kahane [20, Chapter 11]) and have been shown to be very different from Theorem 1.1.

We can extend Theorem 1.1 to the following, which describes the intersection of two independent random covering sets of indices α, α0< 1.

Theorem 1.3. Let E and E0 be two independent Dvoretzky covering sets on the same probability space, associated to the sequences {ln} and {l0n} respectively.

Suppose both {ln} and {ln0} satisfy the condition (C) with the corresponding upper Besicovitch-Taylor indices α, α0 < 1 and possibly different subsequences {ki} and {ki0}. Then for any analytic set G ⊂ [0, 1) satisfying dimP(G) > 1 − min{α, α0}, we have

P E ∩ E0∩ G 6= ∅ = 1.

In particular, if dimP(G) > 1 − α, then

dimP(E ∩ G) = dimP(G) a.s.

In the following we provide an estimate on the Hausdorff dimension of the intersection E ∩ G for a given set G.

Theorem 1.4. Let E be the Dvoretzky covering set associated with the sequence {ln} which satisfies the condition (C). Then for any analytic set G ⊂ [0, 1), we have (1.6) dimH(G) − (1 − α) ≤ dimH(E ∩ G) ≤ dimP(G) − (1 − α) a.s.

By taking G = [0, 1) in Theorems 1.3 and 1.4, we obtain dimH(E) = α and dimP(E) = 1 almost surely. This recovers the result (0.3) of Durand [4], under the extra condition (C). We remark that our method is different from that of Durand [4].

Corollary 1.5. Assume the conditions of Theorem 1.4 hold. For any analytic set G ⊂ [0, 1) satisfying dimH(G) = dimP(G), we have

dimH(E ∩ G) = dimH(G) − (1 − α) a.s.

In particular dimH(E) = α almost surely.

We end this section with some examples.

Example 1.6. 1. If ln ∼ c n−γ, where c > 0 and γ > 1 are constants and ln ∼ jnmeans lim

n→∞

ln

jn = 1, then {ln} satisfies (1.4) with α = γ1. Hence Theorem 2.1 provides results on hitting probabilities for the associated Dvoretzky covering set E.

In particular, we have dimP(E) = 1. This complements the results in Fan and Wu (2004) on the Hausdorff dimension of E. More generally, we can take ln ∼ c1n−γ for even integers n, while ln ∼ c2n−γ0 for odd integers n, where both constants γ and γ0 are larger than 1. Such sequence satisfies (1.4) with α = max{γ−1, γ0−1}.

(6)

Thus, by Durand [4] or by our Corollary 1.5, dimH(E) = α < 1. On the other hand, by [4] or by our Theorem 1.3, dimP(E) = 1.

2. Let {ln} be the sequence corresponding to the complimentary open intervals of the tertiary Cantor set. That is, ln = 3−m when

m−1

P

i=0

2i < n ≤

m

P

i=0

2i (m = 1, 2, . . .). Then it can be verified directly that the upper Besicovitch-Taylor index α = log 2/ log 3 and, moreover, (1.4) holds. Hence our results are applicable to the corresponding random covering set E. Consequently, dimH(E) = log 2/ log 3 and dimP(E) = 1 almost surely. Similar results hold for the random covering sets associated with more general self-similar sets (or self-similar strings, in the terminology of Lapidus and van Frankenhuysen [25]). See Proposition 3.2.

3. If ln= a−n, where a > 1 is a constant, then {ln} satisfies the condition (1.5) with α = 0, by Remark 1.2 (iv), more generally, Proposition 3.3. Hence Theorem 1.1 holds for such {ln}. In particular, we have dimHE = 0 and dimP(E) = 1 almost surely.

4. Finally we provide a simple example of {ln} that satisfies condition (C), but not (1.4). Let β > log 3/ log 2 be a constant. We define

ln=





3−m if

m−1

P

i=0

2i< n ≤

m−1

P

i=0

2i+ 2m−1,

n−β if

m−1

P

i=0

2i+ 2m−1< n ≤

m

P

i=0

2i.

Then we can verify that condition (C) is satisfied with α = log 2/ log 3 and the sub- sequence ki= b(log23)ic, where bxc denotes the largest integer ≤ x. However, (1.4) fails. Nevertheless, the theorems in this section are applicable to the corresponding Dvoretzky covering set E.

2. Proofs of the theorems

In this section we prove Theorems 1.1, 1.3 and Theorem 1.4. It will be clear that the method for studying limsup random fractals in Khoshnevisan, Peres and Xiao [24] plays an essential role in our proofs. We remark that, even though the second half of the proof of Theorem 1.1 is a modification of the proof of Theorem 3.1 in [24], our argument is more general and proves that Theorems 3.1 and 3.2 and their corollaries in [24] still hold if their Condition 4 is replaced by

Condition 40: For some constant γ > 0, lim sup

k→∞

log2pk

k = −γ

and there exists an increasing sequence of positive integers {ki} satisfying (1.2) such that

i→∞lim

log2pki ki

= −γ.

For proving Theorem 1.1 (and for extending the results in [24]) we will use the following elementary lemma on upper box dimension.

Lemma 2.1. Let {ki} be an increasing sequence of positive integers which sat- isfies (1.2). Then for any bounded set G ⊂ R,

(2.1) dimM(G) = lim sup

i→∞

log2N (G, 2−ki) ki

.

(7)

Proof. For any ε > 0, there is an integer i such that 2−ki+1< ε ≤ 2−ki. Thus for any bounded set G ⊂ R we have

N (G, 2−ki) ≤ N (G, ε) ≤ N (G, 2−ki+1).

This implies that

log N (G, 2−ki) kilog 2

ki

ki+1 ≤log N (G, ε)

− log ε ≤log N (G, 2−ki+1) ki+1log 2

ki+1

ki .

It is clear that (2.1) follows from the above and (1.2).  Now we are ready to prove Theorem 1.1.

Proof of Theorem 1.1. Firstly we show that dimP(G) < 1−α implies P E∩

G 6= ∅ = 0. By (0.7), it suffices to show that whenever dimM(G) < 1 − α, then E ∩ G = ∅, a.s.

Fix an arbitrary but small η > 0 such that dimM(G) < 1 − α − η. For any r > 0, denote by Cr = Cr(G) a collection of the smallest number of the intervals with length r that cover the set G. Let Nr(G) = #Cr. Since

lim sup

n→∞

log Nln(G)

− log ln ≤ lim sup

r→0

log Nr(G)

− log r = dimM(G) < 1 − α − η, there exists an integer n0∈ N such that

(2.2) Nln(G) < l−(1−α−η)n

for all n ≥ n0. For any interval J in [0, 1) with length ln, since ωn is uniformly distributed on [0, 1), we have

PIn∩ J 6= ∅ ≤ 3ln. Note that

In∩ G 6= ∅ ⊂ [

J ∈Cln

In∩ J 6= ∅ , we derive from this and (2.2) that

PIn∩ G 6= ∅ ≤ X

J ∈Cln

PIn∩ J 6= ∅ ≤ Nln(G) · 3ln< 3lα+ηn for all n ≥ n0. Hence the seriesP

n=1P{In∩ G 6= ∅} converges by the definition of α and η > 0. By the Borel-Cantelli Lemma, we have

PIn∩ G 6= ∅ i.o. = 0.

That is, P∃n0, s.t. ∀n ≥ n0, In∩ G = ∅ = 1. Therefore, E ∩ G = ∅ a.s.

In the following, we prove that if dimP(G) > 1 − α, then P E ∩ G 6= ∅ = 1.

For this purpose, we construct a random subset E⊂ E and show that P E∩ G 6=

∅ = 1. The random subset E is a discrete limsup random fractal as in [24]. Our proof below is a modification and extension of the method in their Section 3 and is divided into two steps.

(i) Construction. For any k ≥ 2, letDkbe the collection of dyadic intervals of the form (2ik,i+12k ), i = 3, 4, . . . , 2k− 1. Denote by Tk = {n ∈ N : ln ∈ [2−k+1, 2−k+2)}

and let nk= #Tk.

(8)

For every J ∈Dk, define Zk(J ) =

(1 if ∃ n ∈ Tk such that J ⊂ In= (ωn, ωn+ ln), 0 otherwise.

Let

A(k) = [

J ∈Dk Zk(J)=1

J

be the union of open dyadic intervals of order k that are contained in some In with length ln∈ [2−k+1, 2−k+2). Observe that

A(k) ⊂ [

n∈Tk

In.

We define E:= lim supk→∞A(k). From the above, we have E⊂ E.

(ii) Hitting probability. Now let G ⊂ [0, 1) be an analytic set such that dimP(G) >

1 − α. Then by Joyce and Preiss [17], we can find a closed set G⊂ G, such that for all open set V , we have dimM(G∩ V ) > 1 − α, whenever G∩ V 6= ∅.

In the following, we show P E∩ G 6= ∅ = 1. Our method is a modification and extension of that in Section 3 of [24].

For every J ∈Dk, the probability

P Zk(J ) = 1 = P∃ n ∈ Tk such that J ⊂ (ωn, ωn+ ln)

does not depend on J due to our assumption on {ωn} and our definition of Dk. Denote the above probability by Pk. Then

(2.3) Pk ≤ nk(ln− 2−k) ≤ 3 nk2−k. On the other hand,

Pk = P [

n∈Tk

{J ⊂ In}

!

≥ X

n∈Tk

P(In ⊃ J ) − X

m∈Tk

X

n∈Tk n6=m

P Im⊃ J, In ⊃ J

≥ nk ln− 2−k − 9n2k2−2k

≥ nk2−k(1 − 9nk2−k).

(2.4)

In the above, we have used the independence of Im and In (m 6= n) to derive the second inequality. Combining (2.3) and (2.4), together with (1.1) and Condition (C), we derive that

(2.5) lim sup

k→∞

log2Pk

k = −(1 − α)

and there is an increasing sequence of integers {ki} that satisfies (1.2) such that

(2.6) lim

i→∞

log2Pki ki

= −(1 − α).

Hence E is a limsup random fractals which satisfies Condition 40 with γ = 1 − α (which is weaker than Condition 4 in Khoshnevisan, Peres and Xiao [24, p.11]).

Still using their terminology, we call E a limsup random fractal of index 1 − α.

(9)

Next we verify their Condition 5 regarding the correlation of Zk(J ) and Zk(J0) in [24]. Given J and J0 ∈Dk such that the distance d(J, J0) ≥ 2−k+2. Since

Cov Zk(J ), Zk(J0)

= E Zk(J )Zk(J0) − E Zk(J )

E Zk(J0)

= E Zk(J )Zk(J0) − Pk2, (2.7)

we estimate E(Zk(J )Zk(J0)) first,

E(Zk(J )Zk(J0)) = P Zk(J ) = 1, Zk(J0) = 1

= P∃ m, n ∈ Tk such that Im⊃ J and In⊃ J0

≤ X

m∈Tk

X

n∈Tk, n6=m

P Im⊃ J, In⊃ J0

=

 X

m∈Tk

P Im⊃ J



X

n∈Tk, n6=m

P In ⊃ J0

 . (2.8)

By (2.7), (2.8) and the first inequality in (2.4) we derive Cov(Zk(J ), Zk(J0)) ≤ 2

 X

m∈Tk

P Im⊃ J



· X

m∈Tk

X

n∈Tk n6=m

P Im⊃ J, In⊃ J0

≤ C nk2−k

E Zk(J )

E Zk(J0), (2.9)

where the last inequality follows from (2.3) and C > 0 is a finite constant. It follows from (2.9) and (1.1) that for any ε > 0

Cov Zk(J ), Zk(J0) < ε E Zk(J )

E Zk(J0) for all k large enough. This implies that f (k, ε) ≤ 8, where

f (k, ε) = max

J ∈Dk

#J0∈Dk: Cov(Zk(J ), Zk(J0)) ≥ εE(Zk(J ))E(Zk(J0)) . In particular,

lim

k→∞

log f (k, ε)

k = 0.

Thus we have shown that Condition 5 in [24] is satisfied with δ = 0.

The rest of the proof follows a similar line as in the proof of Theorem 3.1 in [24]. For convenience of the reader, we give it below. Notice that our set N is determined by Condition (C) and may be different from that in [24].

Fix an open set V ⊂ [0, 1) such that G∩ V 6= ∅. Let Nk be the number of dyadic intervals J ∈Dk such that

(2.10) J ∩ G∩ V 6= ∅.

Since dimM(G ∩ V ) > 1 − α, we use Lemma 2.1 to derive that, for any β ∈ 1 − α, dimM(G∩ V ), Nki ≥ 2kiβ for infinitely many integers i. This implies the set N defined as

(2.11) N:= {i ≥ 1 : Nki ≥ 2kiβ} satisfies #N = ∞. Similarly to [24], we define

Si= X

J ∈Dki J ∩G∗∩V 6=∅

Zki(J ).

(10)

Namely, Si is the total number of intervals J ∈Dki such that J ∩ G∩ V ∩ A(ki) 6= ∅.

We now show P Si> 0 i.o. = 1.

To this end, we estimate Var(Si) = X

J ∈Dki J ∩G∗∩V 6=∅

X

J 0 ∈Dki J 0 ∩G∗∩V 6=∅

Cov Zki(J ), Zki(J0).

Fix ε > 0, for each J ∈Dki which satisfies (2.10), letGki(J ) be the collection of all J0∈Dki such that

(i) J0∩ G∩ V 6= ∅, and

(ii) Cov Zki(J ), Zki(J0) ≤ εPk2i.

If J0∈Dki satisfies (i), but not (ii), then we say J0∈Bki(J ). Thus Var(Si) ≤ εNk2iPk2i+ X

J ∈Dki J 0 ∈Bki(J )

Cov(Zki(J ), Zki(J0))

≤ εNk2iPk2i+ Nki max

J ∈Dki#Bki(J )Pki,

where the last term comes from the fact that Cov(Zk(J ), Zk(J0)) ≤ E Zk(J ) = Pk. Since we have shown maxJ ∈Dk#Bk(J ) ≤ 8 for all k large enough, the above implies

lim sup

i→∞

i∈N

Var(Si)

[E(Si)]2 ≤ ε + lim sup

i→∞

i∈N

maxJ ∈Dki#Bki(J ) NkiPki

= ε.

In the above, we have used that facts that E(Si) = NkiPki and NkiPki → ∞ if i ∈ N and i → ∞ (recall (2.6) and (2.11)). Since ε > 0 is arbitrary, we have

(2.12) lim sup

i→∞

i∈N

Var(Si) [E(Si)]2 = 0.

It follows from the Paley-Zygmund inequality ([20, p.8]) that P Si> 0 ≥ (E(Si))2

E(Si2)

= 1 −Var(Si)

E(Si2) ≥ 1 − Var(Si)

E(Si)2.

Combining the above inequality, (2.12) and Fatou’s Lemma, we derive (2.13) P Si> 0 i.o. ≥ lim sup

i→∞ P Si > 0 = 1.

It follows from (2.13) that P

( [

k=n

A(k)



∩ G∩ V 6= ∅, ∀n ≥ 1 )

= 1

for every open set V with G∩ V 6= ∅. Letting V run over all open interval with rational endpoints, we obtain that (∪k=nA(k))∩Gis a.s. dense in Gfor all n ≥ 1.

Since

 [

k=n

A(k)



∩ G

(11)

is an open set in Gand Gis a complete metric space, by Baire’s category theorem (see Munkres [29]), we know ∩n=1(∪k=nA(k)) ∩ G is a.s. dense in G, that is, E∩ Gis a.s. dense in G. In particular, E∩ G6= ∅ a.s. This finishes the proof

of Theorem 1.1. 

Proof of Theorem 1.3. We use the same method as in the proof of Theorem 3.2 in [24]. Let G be the closed subset of G described in the proof of Theorem 1.1. Suppose dimP(G) > 1 − min{α, α0}, the proof of Theorem 1.1 shows that for any open set V such that V ∩ G6= ∅ we have

P

 [

k=n

Ik

∩ V ∩ G6= ∅, ∀n ≥ 1



= P

 [

k=n

Ik0

∩ V ∩ G6= ∅, ∀n ≥ 1



= 1.

By independence, there exists a single null probability event outside which for all open intervals V with rational endpoints satisfying V ∩ G6= ∅, we have

 [

k=n

Ik



∩ V ∩ G6= ∅ and

 [

k=n

Ik0



∩ V ∩ G6= ∅ for all n ≥ 1.

That is, 

k=nIk ∩ G}n≥1∪

k=nIk0) ∩ G

n≥1 is a countable collection of open, dense subsets of the complete metric space G. Again, Baire’s theorem implies that

P E ∩ E0∩ G is dense in G = 1.

In particular, E ∩ E0∩ G 6= ∅ a.s. That is, P(E ∩ E0∩ G 6= ∅) = 1. This proves the first part of Theorem 1.3.

In order to prove the second half, we regard the set E ∩ G as the target set with respect to the random covering set E0. By Theorem 1.1, we know that P(E0∩E∩G 6=

∅) = 1 implies dimP(E ∩ G) ≥ 1 − α0 a.s. Therefore, from the above we see that dimP(G) > 1 − min{α, α0} implies dimP(E ∩ G) ≥ 1 − α0 a.s.

Now we assume dimP(G) > 1 − α. For any α0with 1 − dimP(G) < α0 < α, that is, dimP(G) > 1 − min{α, α0}, we have dimP(E ∩ G) ≥ 1 − α0 a.s. Letting α0 tend to 1 − dimP(G) along rational numbers, we obtain

dimP(E ∩ G) ≥ dimP(G) a.s.

Therefore, dimP(E ∩ G) = dimP(G) a.s. 

Proof of Theorem 1.4. Firstly, we prove the right-hand inequality in (1.6).

By (0.7), it suffices to prove that

(2.14) dimH(E ∩ G) ≤ dimM(G) − (1 − α) a.s.

Denote by Cln a collection of the smallest number of the intervals with length ln, the union of such intervals covers the set G. Let Nln(G) = #Cln. Since ξ :=

dimM(G) ≥ lim sup

n→∞

log Nln(G)

− log ln , we have

Nln(G) < l−(ξ+ε)n

as n large enough, say n ≥ n1(ε), where ε > 0 is an arbitrary small real number.

LetGn be the collection of the intervals J ∈ Cln such that J ∩ In 6= ∅ and denote Tn= #Gn. For any J ∈Gn, P(In∩ J 6= ∅) ≤ 3ln. Thus

E(Tn) ≤ X

J ∈Gn

P(In∩ J 6= ∅) ≤ 3Nln(G)ln ≤ 3ln1−ξ−ε.

(12)

For any θ > ξ − (1 − α), we choose ε > 0 such that 2ε < θ − ξ + (1 − α), then

E



X

n=n1(ε)

Tnlnθ



< 3

X

n=n1(ε)

l1−ξ−εn lξ−(1−α)+2εn = 3

X

n=n1(ε)

lα+εn < ∞.

Thus E(P

n=1Tnlθn) < ∞. It follows thatP

n=1Tnlθn< ∞ a.s.

For any m ≥ 1, the collection {J ∈Gn}n≥m is a covering of the set E ∩ G, then Hθ E ∩ G ≤

X

n=m

Tnlθn< ∞ a.s.,

which implies dimH(E ∩ G) ≤ θ a.s. Since θ > dimM(G) − (1 − α) is arbitrary, this proves that (2.14) holds.

The left-hand inequality in (1.6) can be derived from Theorem 1.1 and the following Lemma, due to Khoshnevisan, Peres, and Xiao [24] (Lemma 3.4 with N = 1 and γ = 1 − α). The proof of Theorem 1.3 completed.  Lemma 2.2. Equip [0, 1] with the Borel σ-field. Suppose E = E(ω) is a random set in [0, 1] (i.e., the indicator function χE(ω)(x) is jointly measurable) such that for any compact set F ⊂ [0, 1] with dimH(F ) > γ, we have P(E ∩ F 6= ∅) = 1.

Then, for any analytic set F ⊂ [0, 1],

dimH(F ) − γ ≤ dimH(E ∩ F ) a.s.

3. Technical results

The upper Besicovitch-Taylor index (or the convergence exponent) of {ln} plays an essential role in this paper. In this section we provide some equivalent charac- terizations for this index and elaborate more on the condition (C) and (1.4).

First we show that

Proposition 3.1. For any constant a > 1, let nk = #{n ∈ N : ln∈ [a−k+1, a−k+2)}.

Then

(3.1) α = lim sup

k→∞

logank k . Proof. For any γ > α, we have P

n=1lγn < ∞ orP

k=1nka−γ(k−1) < ∞.

Thus

nka−γ(k−1) ≤ 1 for all k large, which implies

lim sup

k→∞

logank k ≤ γ.

Hence we have

lim sup

k→∞

logank

k ≤ α.

On the other hand, if γ > lim supk→∞ logaknk, we choose γ0 such that lim sup

k→∞

logank

k < γ0 < γ.

(13)

This implies nk≤ a0 for all k large enough, say k ≥ k0. Hence

X

k=k0

nka−kγ

X

k=k0

a−k(γ−γ0)< ∞.

This implies α ≤ γ, which proves α ≤ lim supk→∞logaknk. Therefore (3.1) holds.  For any decreasing sequence {ln} of positive numbers such that

P

n=1

ln < ∞, one can associate the following Dirichlet series

ζ(s) =

X

n=1

lns =

X

n=1

e−s ln(l−1n ),

which is called the geometric zeta function in Lapidus and van Franhenhuysen [25]. Then the upper Besicovitch-Taylor index α defined by (0.4) is the abscissa of convergence of the above Dirichlet series. On the other hand, denote by N (x) the counting function defined by

N (x) = #n

n : l−1n ≤ xo ,

see [25, p.8]. Then nkin Proposition 3.1 can be written as nk= N (ak−1)−N (ak−2) for a > 1, hence the index α can also be determined by N (x) (we take a = 2):

(3.2) α = lim sup

k→∞

log2 N (2k−1) − N (2k−2)

k .

Thanks to the above we can also apply the results in [25] to calculate the upper Besicovitch-Taylor index of a sequence {ln}. In the following we focus on sequences which are associated to self-similar sets (or self-similar strings in [25]).

Given an integer M ≥ 2 and constants r1, . . . , rM ∈ (0, 1) such that 1 ≥ r1≥ r2≥ · · · ≥ rM > 0 and R =

M

X

i=1

ri< 1,

one can construct self-similar sets in [0, 1] with scaling ratios r1, . . . , rM (cf. [6, 25, 28]). Similarly to the tertiary Cantor set in Section 2, we denote the corresponding sequence by {ln}, where each ln is of the form

r1k1· · · rkMM, where k1, . . . , kM ∈ N.

It can be verified that the multiplicity of the length rk11· · · rMkMin {ln} is the multino- mial coefficient k q

1··· kM, where q = PM

i=1

ki; see [25, p.24].

By the proof of Theorem 2.3 in [25] we see that the geometric zeta function of {ln} is

(3.3) ζ(s) =

X

q=0

 M X

i=1

rsi

q

, ∀s ∈ C.

This, together with (0.4), implies the first assertion of Proposition 3.2 below.

The asymptotic behavior of the counting function N (x) for a sequence {ln} associated to a self-similar set has been studied in [25] (see also the references therein for further information). We notice that the zeta function ζ(s) in (3.3)

(14)

satisfies conditions (H1) and (H02) in [25, p.80] with κ = 0 and A = rM (see [25, pp.121–122]). Hence we can apply Theorem 4.8 in [25, p.88] to obtain that

(3.4) N (x) = X

ω∈D(C)

res xsζ(s) s ; ω



+ constant

for all x > rM. In the above D(C) denotes the set of complex dimensions of {ln} (i.e., the set of poles of ζ(s) or equivalently the set of solutions of the equation PM

i=1riω= 1) and res(g(s); ω) denotes the residue of a meromorphic function g(s) at s = ω.

To obtain more explicit information about the terms on the right hand side of (3.4), we distinguish two cases:

Nonlattice case: The additive group generated by log r1, . . . , log rM is dense in R.

Lattice case: There exists some number δ > 0 such that log r1, . . . , log rM ∈ δZ. The largest such δ is called the additive generator and is denoted by r [25, p.34]. The positive constant p = (2π)/(log r−1) is called the oscil- latory period.

In the nonlattice case, it follows from (5.44) in [25, p.126] that (3.5) N (x) = res(ζ; α)xα

α + o(xα), as x → ∞.

The lattice case is much simpler since the complex dimension of {ln} are located on finitely many vertical lines [25, Theorem 2.13]. It follows from (5.33) and (5.34) in [25, pp.122-123] that

(3.6) N (x) = res(ζ; α)b1−{u}

b − 1 2π

p xα+ o(xα), as x → ∞,

where log b = 2πα/p, u = p log x/2π, {x} = x − bxc is the fractional part of x.

By (3.5), (3.6) and (3.2) we derive

(3.7) lim

k→∞

log2 N (2k−1) − N (2k−2)

k = α.

In other words, (1.4) always holds for a self-similar sequence {ln}.

Hence we have proved the following proposition.

Proposition 3.2. Let {ln} be the sequence associated to a self-similar set with scaling ratios r1, . . . , rM. Then the upper Besicovitch-Taylor index α of {ln} coin- cides with the self-similarity dimension D, which is the unique constant satisfying

M

X

i=1

riD= 1.

Moreover, (1.4) holds.

As an example, we mention the Fibonacci sequence, which is obtained by taking M = 2, r1 = 1/2 and r2= 1/4. Then it can be verified directly that α = log2φ, where φ = 1+

5

2 is the golden ratio, and its geometric counting function is given by

NFib(x) = 3 + 4φ

5 φ−{log2x}xα− 1 + 7 − 4φ

5 φ{log2x}x−α(−1)blog2xc, see [25, p.124]. It can be verified directly that (1.4) holds.

(15)

Finally we show that condition (1.4) can be replaced by (1.5), as stated in Remark 1.2 (iv).

Proposition 3.3. For any constants a > b > 1, let mk = #n : ln ∈ [b−k+1, b−k+2)

and let nk = #n : ln ∈ [a−k+1, a−k+2) . If lim

k→∞

logbmk

k = α, then

lim

k→∞

logank

k = α.

Proof. We state the elementary fact that if lim

k→∞

logbmk

k = α, then for any fixed integer τ0≥ 1, we have

(3.8) lim

k→∞

logb(mk+ mk+1+ · · · + mk+τ0)

k = α.

This can be verified by the fact that b(α−)k< mk< b(α+)k for all k large implies b(α−)k< mk+ mk+1+ · · · + mk+τ0 < (τ0+ 1)b(α+)(k+τ0)

for all k large.

To prove the lemma, observe that

ln ∈ [a−k+1, a−k+2) ⇐⇒ ln ∈ [b−(logba)(k−1), b−(logba)(k−2)).

Hence

nk ≤ #n : ln∈ [b−b(logba)(k−1)c−1, b−b(logba)(k−2)c)

= mb(logba)(k−1)c+2+ · · · + mb(logba)(k−2)c+2, (3.9)

where bxc denotes the largest integer ≤ x, and note that a > b, we have nk ≥ #n : ln∈ [b−b(logba)(k−1)c, b−b(logba)(k−2)c−1)

= mb(log

ba)(k−1)c+1+ · · · + mb(log

ba)(k−2)c+3. (3.10)

Since limk→∞

logbmb(logb a)kc

(logba)k = α, we derive from (3.8), (3.9) and (3.10) that lim

k→∞

logank

k = lim

k→∞

logbnk (logba)k = α.

This proves the lemma. 

Acknowledgement. This paper was developed and finished when Bing Li did his post-doc research at National Taiwan University, under a grant from NCTS Taipei Office, and during his visit to Michigan State University. The hospitality of the hosts is appreciated. The authors thank Prof. Ai Hua Fan for his helpful comments.

References

[1] J. Barral and A. H. Fan, Covering numbers of different points in Dvoretzky covering, Bull.

Sci. Math. 129 (2005), no. 4, 275–317.

[2] V. Beresnevich, and S. Velani, A mass transference principle and the Duffin-Schaeffer con- jecture for Hausdorff measures, Ann. of Math. (2) 164 (2006), no. 3, 971–992.

[3] A. S. Besicovitch and S. J. Taylor, On the complementary intervals of linesr closed set of zero Lebesgue measure, J. London Math. Soc. 29 (1954), 520–526.

[4] A. Durand, On randomly placed arcs on the circle, in: Recent Developments in Fractals and Related Fields (Applied and Numerical Harmonic Analysis), 343–352 (edited by J. Barral and S. Seuret), Birkh¨auser, Boston, 2010.

(16)

[5] A. Dvoretzky, On covering a circle by randomly placed arcs, Proc. Nat. Acad. Sci. U.S.A. 42 (1956), 199–203.

[6] K. J. Falconer, Fractal geometry – mathematical foundations and applications, Wiley & Sons Ltd., Chichester, 1990.

[7] A. H. Fan, How many intervals cover a point in Dvoretzky covering? Israel J. Math. 131 (2002), 157–184.

[8] A. H. Fan, Some topics in the theory of multiplicative chaos, Fractal geometry and stochastics III, 119–134, Progr. Probab., 57, Birkh¨auser, Basel, 2004.

[9] A. H. Fan, J. P. Kahane, Raret´e des intervalles recouvrant un point dans un recouvrement al´eatoire. Ann. Inst. H. Poincar´e Probab. Statist. 29 (1993), no. 3, 453–466.

[10] A. H. Fan and J. Wu, On the covering by small random intervals, Ann. Inst. H. Poincar´e Probab. Statist. 40 (2004), 125–131.

[11] P. J. Fitzsimmons, B. Fristedt and L. A. Shepp, The set of real numbers left uncovered by random covering intervals, Z. Wahr. Verw. Gebiete 70 (1985), 175–189.

[12] J. Hawkes, On the covering of small sets by random intervals, Quart. J. Math. Oxford (2) 24 (1973), 427–432.

[13] J. Hawkes, Random re-orderings of intervals complementary to a linear set, Quart. J. Math.

Oxford (2) 35 (1984), 165–172.

[14] S. Janson, Random covering in several dimensions, Acta Math. 156 (1986), 83–118.

[15] J. Jonasson, Dynamical circle covering with homogeneous Poisson updating, Statist. Probab.

Lett. 78 (2008), 2400-2403.

[16] J. Jonasson and J. E. Steif, Dynamical models for circle covering: Brownian motion and Poisson updating, Ann. Probab. 36 (2008), no. 2, 739–764.

[17] H. Joyce and D. Preiss, On the existence of subsets of finite positive packing measure, Math- ematika 42 (1995), no. 1, 15–24.

[18] J.-P. Kahane, The technique of using random measures and random sets in harmonic analy- sis, Advances in probability and related topics, Vol. 1, pp. 65–101. Dekker, New York, 1971.

[19] J.-P. Kahane, Ensembles parfaits et processus de L´evy, Periodica Math. Hungarica 2 (1972), 49–59.

[20] J.-P. Kahane, Some random series of functions, Second edition. Cambridge Studies in Ad- vanced Mathematics, 5. Cambridge University Press, Cambridge, 1985.

[21] J.-P. Kahane, Recouvrements al´eatoires et theorie du potentiel, Colloq. Math. LX/LXI (1990), 387–411.

[22] J.-P. Kahane, Recouvrements al´eatoires, Gaz. Math. 53 (1992), 115–129.

[23] J.-P. Kahane, Random coverings and multiplicative processes, In: Fractal geometry and sto- chastics, II (Greifswald/Koserow, 1998), 125–146, Progr. Probab., 46, Birkha¨user, Basel, 2000.

[24] D. Khoshnevisan, Y. Peres, and Y. Xiao, Limsup random fractals, Electron. J. Probab. 5 (2000), no. 5, 24 pp. (electronic).

[25] M. L. Lapidus and M. van Frankenhuysen, Fractal geometry and number theory, Complex dimensions of fractal strings and zeros of zeta functions, Birkh¨auser, Boston, 2000.

[26] B. B. Mandelbrot, Renewal sets and random cutouts, Z. Wahr. Verw. Gebibte 22 (1972), 145–157.

[27] A. I. Markushevich, Theory of functions of a complex variable, Vol. II. Revised English edition translated and edited by Richard A. Silverman. Prentice-Hall, Inc., Englewood Cliffs, N.J. 1965.

[28] P. Mattila, Geometry of sets and measures in Euclidean spaces, Fractals and rectifiability, Cambridge Studies in Advanced Mathematics, 44. Cambridge University Press, Cambridge, 1995.

[29] J. R. Munkres, Topology: a first course, Prentice-Hall, Inc., Englewood Cliffs, N. J., 1975.

[30] G. P´olya and G. Szeg¨o, Problems and theorems in analysis. I. Series, integral calculus, theory of functions, Classics in Mathematics. Springer-Verlag, Berlin, 1998.

[31] L. A. Shepp, Covering the circle with random arcs, Israel J. Math. 11 (1972), 328–345.

[32] S. J. Taylor, The measure theory of random fractals, Math. Proc. Cambridge Philo. Soc. 100 (1986), 383–406.

[33] C. Tricot, Two definitions of fractional dimension, Math. Proc. Cambridge Philo. Soc. 91 (1982), 57–74.

(17)

[34] Y. Xiao, Random fractals and Markov processes, In: Fractal Geometry and Applications: A Jubilee of Benoit Mandelbrot, (Michel L. Lapidus and Machiel van Frankenhuijsen, editors), pp. 261–338, American Mathematical Society.

Department of Mathematics, South China University of Technology. 510640, Guangzhou, P. R. China.

E-mail address: libing0826@gmail.com

Mathematics Department, Honorary Faculty, National Taiwan University. Taipei 10617, Taiwan.

E-mail address: shiehnr@ntu.edu.tw

Department of Statistics and Probability, 619 Red Cedar Road, Michigan State University, East Lansing, MI 48824, U.S.A.

E-mail address: xiao@stt.msu.edu

參考文獻

相關文件

Given an undirected graph with nonnegative costs on the edges, the routing cost of any of its spanning trees is the sum over all pairs of vertices of the cost of the path between

That the sequence is increasing can be shown by induction... If not,

To do (9), you need to recall the exercise from hw 1 and hw 2 in Calculus I: (you do not need to turn in the following exercises) If you are not familiar with the exercises below,

Juang has received numerous distinctions and recognitions, including Bell Labs' President Gold Award, IEEE Signal Processing Society Technical Achievement Award, the IEEE

Walker provides all past (and present) flower children an opportunity to get in touch with their roots in her article, &#34;Where Have All the Flowers Gone?&#34;. Walker uses

The first typ e is the unsafe drivers, who disregard traffic regulations and make life difficult for other drivers and people walking in the street.. Unsafe drivers do

Chun-Chuan Chou, A Study Of Stamp problem, Thesis for Master of Science, De- partment of Applied Mathematics, National Chiao Tung University, Hsinchu,

When a solution curve crosses one of these lines, it has a local maximum or