• Nie Znaleziono Wyników

arXiv:1211.3870v2 [math.PR] 18 Feb 2013

N/A
N/A
Protected

Academic year: 2021

Share "arXiv:1211.3870v2 [math.PR] 18 Feb 2013"

Copied!
25
0
0

Pełen tekst

(1)

arXiv:1211.3870v2 [math.PR] 18 Feb 2013

INTEGRABILITY AND CONCENTRATION OF SAMPLE PATHS’ TRUNCATED VARIATION OF FRACTIONAL

BROWNIAN MOTIONS, DIFFUSIONS AND L´EVY PROCESSES

WITOLD BEDNORZ AND RAFA L LOCHOWSKI

Abstract. For a real c`adl`ag function f defined on a compact interval, its truncated variation at the level c > 0 is the infimum of total vari- ations of functions uniformly approximating f with accuracy c/2 and (in opposite to total variation) is alway finite. In this paper we discuss exponential integrability and concentration properties of the truncated variation of fractional Brownian motions, diffusions and L´evy processes.

We develop a special technique based on chaining approach and using it we prove Gaussian concentration of the truncated variation for certain class of diffusions. We also give sufficient and necessary condition for the existence of exponential moment of order α > 0 of truncated variation of L´evy process in terms of its L´evy triplet.

1. Introduction

Let X = (X(t))t≥0 be a real valued stochastic process with c`adl`ag tra- jectories. In general, the total path variation of X on the compact interval [a; b]⊂ [0; +∞), defined as

T V (X, [a; b]) = sup

n sup

a≤t0<t1<...<tn≤b n

X

i=1

|X(ti)− X(ti−1)| ,

may be (and in many most important cases is) a.s. infinite. However, in the neighborhood of every c`adl`ag path we may easily find a function with finite total variation.

Let f be a c`adl`ag function f : [a; b]→ R and let c > 0. The natural ques- tion arises, what is the smallest possible (or the greatest lower bound for) total variation of functions from the ball B(f, c/2) ={g : kf − gk≤ c/2} , where kf − gk := sups∈[a;b]|f (s) − g (s)| . Some bound from below reads as

T V (g, [a; b])≥ T Vc(f, [a; b]) ,

2010 Mathematics Subject Classification. 60G17, 60G15, 60G22, 60G48, 60G51.

Key words and phrases. sample boundedness, Gaussian processes, diffusions, L´evy pro- cesses, truncated variation.

The research of the first and the second author was supported by the National Science Center in Poland under decision no. DEC-2011/01/B/ST1/05089.

1

(2)

where (1.1)

T Vc(f, [a; b]) := sup

n sup

a≤t0<t1<...<tn≤b n

X

i=1

max{|f (ti)− f (ti−1)| − c, 0}

and follows immediately from the inequality

|g (ti)− g (ti−1)| ≥ max {|f (ti)− f (ti−1)| − c, 0} . It is possible to show (cf. [9]) that in fact we have equality (1.2) inf{T V (g, [a; b]) : kf − gk≤ c/2} = T Vc(f, [a; b]) attained for some function fc from the ball B(f, c/2).

Remark 1. Since we deal with c`adl`ag functions, more natural setting of our problem would be the investigation of

inf{T V (g, [a; b]) : g - c`adl`ag, dD(f, g)≤ c/2} ,

where dD denotes Skorohod metric. Since the total variation does not de- pend on the (continuous and strictly increasing) change of argument and the function fc minimizing T V (g, [a; b]) appears to be a c`adl`ag one, solutions of both problems coincide.

The bound (1.1) is called truncated variation and it is finite for any c`adl`ag function, since every such function may be uniformly approximated by step functions. Moreover, truncated variation is a continuous and convex function of the parameter c > 0 (cf. [9]), and it obviously tends to the total variation as c↓ 0. For a process with paths with a.s. infinite total variation it may be of interest to assess the rate of this convergence.

This was done so far for continuous semimartingales and it appears (cf.

[12]) that for any continuous semimartingale X we have that c· T Vc(X, [a; b])→c↓0 hXib− hXia a.s., whereh.i denotes the quadratic variation of X.

For t ≥ 0 denote T Vc(X, t) = T Vc(X, [0; t]) . For X being the unique strong solution of the equation X0 = 0, dXt = µ (Xt) dt + σ (Xt) dBt, t ∈ [0; S] , driven by a standard Brownian motion B, with µ and σ satisfying some linear growth conditions, we have even stronger result (cf. [12, Theo- rem 10]).

T Vc(X, t)−hXit

c ⇒c↓0hXi

t/√ 3,

where ˜B is a standard Brownian motion independent from B and the conver- gence ”⇒” is understood as the weak functional convergence in C([0; S], R) topology.

For a standard Brownian motion and fixed S > 0 this result seems to in- dicate very strong concentration of its truncated variation on the interval [0; S] around S/c, but it still does not tell anything about the tail proba- bilities of the functional considered. This observation inclined us to study

(3)

the integrability and concentration properties of the truncated variation in greater detail.

Some investigation into this direction was undertaken in [10], were the existence of moment generating function of the truncated variation of Brow- nian motion with drift on the whole real line was proven, but in this paper we obtain much stronger - Gaussian concentration result.

Another incentive for the study of the magnitude of truncated variation for possibly broad class of processes is the pathwise approach to stochastic integration. In [11] it was shown that it is possible to define the stochastic integral with some correction term as an a.s. limit of pathwise Lebesgue- Stieltjes integrals, when the semimartingale integrator is uniformly approx- imated on compacts by a finite variation process. Thus the truncated vari- ation gives the magnitude of such integrals, more precisely

kX−Xinfck≤c/2 sup

kY k≤1

Z S 0

YdXc = T Vc(X, S) .

Thus, in this paper we study the magnitude of the truncated variation for Gaussian processes, among them for fractional Brownian motions, and for diffusions. Further we also consider L´evy processes. Our main goal it to describe the tail behavior of T Vc(X, S) assuming that X satisfies some in- crement condition. We use various techniques depending on the assumption we make. At the beginning we use the chaining concept, i.e. we assume that X satisfies some exponential integrability condition on increments and deduce the exponential integrability of the truncated variation. The chain- ing approach was first used to study problems of sample boundedness of processes on the general index space [5, 6]. The method was developed to give the full description of classes of processes that are sample bounded un- der certain integrability condition [1, 2, 3, 8, 15] as well as the small ball probability [4]. For a comprehensive study where many analytical examples are given see [16]. In our study we need some modification of the idea, since we are interested in bounding the supremum of special sums of increments rather then the supremum over increments itself. Therefore we have to in- vent a special random variable of exponential integrability that bounds the truncated variation.

Our main toy example is the class of fractional Brownian motions, i.e. cen- tered Gaussian processes WH, H ∈ (0, 1), starting from 0 and such that E|WH(t)− WH(s)|2=|t − s|2H. One of the corollaries we get is the follow- ing concentration inequality

P(T Vc(WH, S)≥ cH−1H S(AH + BHu))≤ CHexp(−u2H), for u≥ 0,

(4)

where AH, BH, CH are constants, moreover one can set CH = 1 for H ≥ 12. By the homogeneity of increments we deduce that ET Vc(WH, S) is compa- rable with cH−1H S and in this way we prove that

(1.3)

P(T Vc(WH, S)≥ ET Vc(WH, S)( ¯AH+ ¯BHu))≤ ¯CHexp(−u2H), for u≥ 0, for some constants ¯AH, ¯BH, ¯CH (again ¯CH = 1 for H ≥ 12). In fact any process with similar properties as the fractional Brownian motion, i.e. with homogeneity of increments and exponential integrability of increments can be treated by our method.

Next we turn to investigate the specific fractional Brownian motion, i.e.

W = W1/2and diffusions driven by it. Here we can improve our result using the Markov property. It occurs that for Markov processes with moderate growth some local exponential integrability can be extended to the global one. Note that (1.3) implies the existence of the Laplace transform for sufficiently small λ > 0; assuming the Markov property for diffusions with moderate growth we get the estimate for the Laplace transform on the whole real line. The main result we get in this way is Theorem 2, which for standard Brownian motion implies the following concentration inequality

P(T Vc(W, S)≥ ¯AET Vc(W, S) + ¯B√

Su)≤ exp(−u2), for u≥ 0, where ¯A, ¯B are universal constants. Therefore the Gaussian concentration inequality holds for the truncated variation of the standard Brownian mo- tion. Our result gives better understanding of the already mentioned phe- nomenon that S12(T Vc(W, S)− S/c) converges in distribution to N (0, 1/3) as c↓ 0.

We continue the study proving sufficient and necessary condition for the finiteness of E exp (αT Vc(X, S)) for a L´evy process X, in terms of its gen- erating triplet. Here we apply the method of level crossing stopping times.

The structure of the paper is as follows. In Section 2 we introduce the chaining approach which will lead us to the main result on the concentra- tion for processes with increments of exponential decay. Then in Section 3 we discuss the application of the developed methodology to the fractional Brownian motions and then, in Section 4 its improvement for a standard Wiener process and diffusions with moderate growth. In Section 5 we deal with truncated variation of a L´evy process.

Remark 2. In the whole paper, any dependence of a nonnegative constant on some parameters is always indicated by listing them in brackets or in subscripts, for example, C(n, ε) or Cn,ε.

(5)

2. The chaining approach

In this section we prove fundamental Theorem 1, which will allow us to establish integrability and concentration properties of the truncated varia- tion for a broad class of processes satisfying some increment condition. For simplicity, in this paper we consider processes indexed by a parameter from the metric space (T, d), where T is the compact interval [0, S] equipped with the distance d(s, t) = |s − t|q for s, t ∈ T , where 0 < q < 1. Further, let X(t), t∈ T, be a stochastic process. We assume that there exists p > 0 and the function ϕp(x) = 2xp− 1 for x ≥ 0, such that

(2.1) Eϕp

|X(s) − X(t)|

Cd(s, t)



≤ 1 for s, t ∈ T, s 6= t,

where 0 ≤ C < ∞ is a universal constant. Condition (2.1) enables us to control the magnitude of the increments of process X, while the truncated variation takes into account only increments greater than c (cf. formula (1.1)). Note that as the consequence of (2.1) and the compactness of T we obtain the existence of a separable modification of X(t), t∈ T . Then by the linear order of T we can define the c`adl`ag modification of X which we refer to from now on. Note that the exponential growth of ϕp implies that (2.2) ϕ−1p (xy)≤ Lp−1p (x) + ϕ−1p (y)),

for x, y ≥ 0, where Lp ∈ (0; 2]. Furthermore observe that if p ≥ 1, ϕp(x) is convex on the whole interval [0; +∞) and if 0 < p < 1, ϕp(x) is convex on the interval [Cp; +∞) where Cp =

1−p p ln 2

1/p

. In the same way ϕp(xq) = ϕpq(x) is convex on the whole interval [0; +∞) if pq ≥ 1 and convex on the interval [Cp,q; +∞), where Cp,q= Cpq, if pq < 1. We use the notation Cp, Cp,qfor all p > 0, 0 < q < 1 setting Cp = 0 for p≥ 1 and Cp,q= 0 for pq≥ 1. Further, we denote Dp = ϕp(Cp), Dp,q = ϕp(Cp,qq ) = ϕpq(Cpq). Note that Dp,q = 0 for pq≥ 1.

The fundamental result of this paper, from which exponential integrability and concentration properties will follow is

Theorem 1. Let X(t), t∈ T , satisfies (2.1) for ϕp = 2xp− 1 and d(s, t) =

|s − t|q, where where p > 0, 0 < q < 1. Then there exist random vari- ables Z1, Z2 ≥ 0 such that EZ1, EZ2 ≤ 1 and for some universal constants K1(p, q), K2(p, q) <∞ the following estimate holds

T Vc(X, S)≤ cq−1q S[K1(p, q)ϕ−1p (Z1+ Dp) + K2(p, q)[ϕ−1p ]1q(Z2+ Dp,q)].

Remark 3. The main reason why the result holds is that (2.1) gives an exponential decay of increments with large jumps. Therefore we can show a global upper bound on increments in the defined set approximation of the truncated variation. Such an idea is used to bound suprema of processes e.g.

[1, 5, 7, 15]. In this paper the main question is to invent common upper bound for an arbitrary sum of truncated increments.

(6)

Remark 4. The more general setting is possible, namely one may consider exponential like Young functions ϕ, i.e. positive, increasing, with ϕ(0) = 0, ϕ(1) = 1 and such that the following condition holds

(2.3) ϕ−1(xy)≤ L(ϕ−1(x) + ϕ−1(y)), for x, y > 0,

where L <∞ is a constant. On the other hand we can consider any distance d of the form d(s, t) = η(|s − t|), where η is positive, concave, increasing to

∞ and such that η(0) = 0.

We start with the construction of finite sets approximating T .

2.1. Approximating sequence. In order to control increments of the pro- cess X we approximate space T by a sequence of finite sets (Tn)n=0, Tn⊂ T . The approximation must be efficient and here we use the entropy of (T, d).

Let N (T, d, ε) be the smallest number of closed balls of radius ε that cov- ers T . More precisely, let us fix r ≥ 4. We require that Tn is such that

|Tn| = N(T, d, r−nSq) andS

t∈TnBd(t, r−nS)= T, where Bd(t, r) ={s ∈ T : d(s, t)≤ r}. By the linear order of points on the interval it is clear that (2.4) 2−1rnq ≤ |Tn| < 2−1rnq + 1.

Moreover we may require that d(s, t)≥ r−nSq for any s, t ∈ Tn, s 6= t and d(t, Tn)≤ r−nSq for all t∈ T . Clearly, for any m = 1, 2, . . .

(2.5)

m

X

n=0

r−n|Tn+1| ≤ 2−1

m

X

n=0

r−n(rn+1q + 2)≤ A(r, q)rm1−qq ,

where A(r, q) := r2−qq (r1−qq − 1)−1 (note that rn+1q ≥ 2). For each t ∈ Tn+1

let In+1(t) denote the set of s∈ Tn+1 such that d(s, t)≤ 2r−nSq, i.e.

(2.6) In+1(t) =s ∈ Tn+1: d(s, t)≤ 2r−nSq . Observe that since|s − t| ≥ rn+1q S for s, t∈ Tn+1, s6= t, (2.7) |In+1(t)| ≤ 21qrnqS

rn+1q S

+ 1 = 21qr1q + 1 =: B(r, q).

Let πn(t) denote the projection of T on Tn, i.e. we require that πn: T → Tn

satisfies d(t, πn(t)) = d(t, Tn). Note that we can define πn in such a way that it preserves the order on real line i.e. πn(s) ≤ πn(t) whenever s ≤ t.

Moreover by the construction d(t, πn(t))≤ r−nSq for all t∈ T .

2.2. Proof of the main theorem. Our first step is to describe two classes of possible points in a given partition. We fix Πn = {t0, t1, . . . , tn}, where 0 ≤ t0 < t1 < · · · < tn ≤ S. Let Jm = {i ∈ N : r−m−1Sq < d(ti−1, ti) ≤ r−mSq} for m ≥ 0. The crucial level is m0 ≥ 0 such that r−m0−1Sq < c≤

(7)

r−m0Sq. For i∈ Jm with m > m0 we will apply exponential concentration.

We have

n

X

i=1

(|X(ti)− X(ti−1)| − c)+

m0

X

m=0

X

i∈Jm

(|X(ti)− X(ti−1)| − c)+

+

X

m=m0+1

X

i∈Jm

(|X(ti)− X(ti−1)| − c)+

(2.8)

The main tool we will use is the chaining method. We turn to describe the path approximation of ti for i ∈ {0, 1, . . . , n}. Therefore we fix N ≥ 0 large enough and define tN +1i = πN +1(ti), then for l∈ {0, 1, . . . , N} we put by the reverse induction tli = πl(tl+1i ). Note that by the construction of πl we preserve the order of the projections, namely tl0 ≤ tl1 ≤ ... ≤ tln for any 0 ≤ l ≤ N + 1. We require that N > m0 and tNi for i ∈ {0, 1, . . . , n}

are separated. For a given i ∈ Jm the level m is the the best to stop the approximation for ti and move to ti−1. For i ∈ Jm we split c among all increments. If m > m0 then

(|X(ti)− X(ti−1)| − c)+

≤ (|X(tm+1i )− X(tm+1i−1 )| − c

3)++ X

s∈{i−1,i}

|X(tN +1s )− X(ts)|

+

N

X

l=m+1

X

s∈{i−1,i}

[(|X(tli)− X(tl+1i )| − 2−l+mc 3)+ (2.9)

and if m≤ m0 then

(|X(ti)− X(ti−1)| − c)+≤ |X(tm+1i )− X(tm+1i−1 )|

+ X

s∈{i−1,i}

|X(tN +1s )− X(ts)| +

m0

X

l=m+1

X

s∈{i−1,i}

|X(tls)− X(tl+1s )|

+

N

X

l=m0+1

X

s∈{i−1,i}

(|X(tls)− X(tl+1s )| − 2−l+m0c 3)+. (2.10)

Putting together estimates (2.8), (2.9) and (2.10) we obtain the following decomposition lemma.

Lemma 1. For any partition Πn={t0, ..., tn}, where n ≥ 0, 0 ≤ t0 < t1 <

... < tn≤ S and N > m0 the following estimate holds

n

X

i=1

(|X(ti)− X(ti−1)| − c)+≤ V1+ V2+ W1+ W2

+

n

X

i=1

X

s∈{i−1,i}

|X(ts)− X(tN +1s )|,

(8)

where

V1 :=

m0

X

m=0

X

i∈Jm

m0

X

l=m+1

X

s∈{i−1,i}

|X(tls)− X(tl+1s )|;

W1 :=

m0

X

m=0

X

i∈Jm

|X(tm+1i )− X(tm+1i−1 )|;

V2 :=

m0

X

m=0

X

i∈Jm

N

X

l=m0+1

X

s∈{i−1,i}

(|X(tls)− X(tl+1s )| − 2−l+m0c 3)+

+

X

m=m0+1

X

i∈Jm

N

X

l=m+1

X

s∈{i−1,i}

(|X(tls)− X(tl+1s )| − 2−l+mc 3)+;

W2 :=

X

m=m0+1

X

i∈Jm

(|X(tm+1i )− X(tm+1i−1 )| − c 3)+.

In the sequel we will use two simple observations concerning increasing function ψ that is convex starting from some C0 ≥ 0, i.e. convex for x ≥ C0. Fact 1. Let ψ : [0; +∞) → [0; +∞) be a strictly increasing function. As- sume that ψ is convex on the interval [C0; +∞) where C0≥ 0, then for any nonnegative x1, ..., xkand positive α1, ..., αk such thatPk

i=1αi≤ M we have (2.11)

k

X

i=1

αixi≤ Mψ−1(M−1

k

X

i=1

αiψ(xi) + ψ(C0)).

Proof. Observe that the function ¯ψ(x) = ψ(x + C0)− ψ(C0) for x≥ 0 is convex, strictly increasing and such that ¯ψ(0) = 0. Consequently ¯ψ−1(y) = ψ−1(y+ψ(C0))−C0is concave with ¯ψ−1(0) = 0 and using Jensen’s inequality we have

k

X

i=1

αixi

k

X

i=1

αiψ−1(ψ(xi) + ψ(C0))

=

k

X

i=1

αi( ¯ψ−1(ψ(xi)) + C0)≤ MC0+ M

k

X

i=1

αi

Mψ¯−1(ψ(xi))

≤ MC0+ M ¯ψ−1

k

X

i=1

αi Mψ(xi)

!

= M ψ−1(M−1

k

X

i=1

αiψ(xi) + ψ(C0)).

 Further, we also have

(9)

Fact 2. For any strictly increasing function ψ : [0; +∞) → [0; +∞) such that ψ is convex on the interval [C0; +∞) where C0 ≥ 0 and any M > 0 we have

(2.12) ψ−1(y + ψ (C0))≤ max{M, 1}ψ−1(y/M + ψ (C0)) .

Proof. Again, we consider the function ¯ψ−1. If M < 1, then the thesis follows from the monotonicity of ¯ψ−1. Now assume that M ≥ 1. By convexity and ¯ψ−1(0) = 0, for y≥ 0 and M ≥ 1 we get

M ¯ψ−1(y/M )≥ ¯ψ−1(y) , which reads as

M ψ−1(y/M + ψ (C0))− C0

 ≥ ψ−1(y + ψ (C0))− C0,

M ψ−1(y/M + ψ (C0)) ≥ ψ−1(y + ψ (C0)) + (M− 1) C0

and which gives

ψ−1(y + ψ (C0))≤ Mψ−1(y/M + ψ (C0)) .

 Now we turn to estimate the increments for i∈ Jm, m≤ m0. We have Lemma 2. There exists a universal constant K1(r, q) < ∞ and a random variable Z1 ≥ 0 independent from the partition Πn, such that EZ1 ≤ 1 and

V1+ W1≤ K1(r, q)cq−1q−1p (Z1+ Dp).

Proof. Recall that r ≥ 4. Note that for i ∈ Jm, m ≤ N the level m + 1 is the largest l ≤ N + 1 such that tli may be equal tli−1 (i.e. tli−1 < tli for l > m + 1). The reason is that d(ti, ti−1) > r−m−1Sq and therefore for l≥ m + 1

d(tl+1i , tl+1i−1)≥ r−m−1Sq− d(tl+1i−1, ti−1)− d(tl+1i , ti)

≥ r−m−1Sq− 2

X

j=l+1

r−jSq≥ r−m−1Sq− 2r−m−2Sq 1− r−1 > 0.

Consequently in the path approximation of Πnthe increment|X(u)−X(πl(u))| with some u ∈ Tl+1 can be used at most twice and only for a single ti from the partition. Moreover, by the entropy construction of Tl we have d(u, πl(u))≤ r−lSq. It implies that

V1≤ 2

m0

X

l=0

X

u∈Tl+1

|X(u) − X(πl(u))|

≤ V1 := 2C

m0

X

l=0

r−lSq X

u∈Tl+1

|X(u) − X(πl(u))| Cd(u, πl(u)) . (2.13)

(10)

On the other hand d(ti, ti−1)≤ r−mSq for i∈ Jm and d(tm+1i , tm+1i−1 )≤ d(ti, ti−1) + d(tN +1i , ti) + d(tN +1i−1 , ti−1)

+

N

X

l=m+1

[d(tl+1i , tli) + d(tl+1i−1, tli−1)]≤ r−mSq+ 2

X

l=m+1

r−lSq≤ 2r−mSq. Therefore using the defined sets Im(u) = {v ∈ Tm+1 : d(u, v)≤ 2r−mSq}, we obtain that

W1

m0

X

l=0

X

u∈Tl+1

X

v∈Il+1(u)

|X(u) − X(v)|

≤ W1:= C

m0

X

l=0

2r−lSq X

u∈Tl+1

X

v∈Il+1(u)

|X(u) − X(v)|

Cd(u, v) . (2.14)

We calculate the sum of all weights appearing in (2.13) and (2.14). By (2.7) for each u∈ Tm+1 we have|Il+1(u)| ≤ B(r, q) and hence, using also (2.5)

M1 :=

m0

X

l=0

r−lSq[|Tl+1| + X

u∈Tl+1

|Il+1(u)|]

≤ [1 + B(r, q)]Sq

m0

X

l=0

r−l|Tl+1| ≤ A(r, q)[1 + B(r, q)]rm01−qq Sq.

Therefore by c ≤ r−m0Sq we get rm01−qq Sq ≤ c1−qq S and hence M1 ≤ A(r, q)[1 + B(r, q)]cq−1q S. Using Fact 1 for ϕp which is convex above Cp we get

(2.15)

V1+W1 ≤ 2CM1ϕ−1p (Z1+ ϕp(Cp))≤ K1(r, q)cq−1q−1p (Z1+ ϕp(Cp)), where K1(r, q) := 2A(r, q)[1 + B(r, q)]C and

Z1= M1−1

m0

X

l=0

r−lSq X

u∈Tl+1

p(|X(u) − X(πl(u))| Cd(u, πl(u)) )

+ X

v∈Il+1(u)

ϕp(|X(u) − X(v)|

Cd(u, v) )].

Obviously Z1 ≥ 0 and EZ1≤ 1 by (2.1). Combining (2.13),(2.14) and (2.15) we get the result.

 Our second goal is to prove a bound for increments above the level m0. Lemma 3. There exists a universal constant K2(p, q, r) <∞ and a random variable Z2 ≥ 0 independent from the partition Πn such that EZ2 ≤ 1 and the following inequality holds

V2+ W2 ≤ K2(p, q, r)[ϕp−1]1q(Z2+ Dp,q).

(11)

Proof. First we prove a bound for V2. Our main tool is (2.2), which implies that

(2.16) ϕp([x− y]+)≤ ϕp(Lpx)

ϕp(y) , x, y≥ 0.

We analyze the increment

(|X(tl+1i )− X(tli)| − 2m−lc

3)+, l > m, i∈ Jm, m > m0. Using (2.16) and d(tl+1i , tli)≤ r−lSq we obtain that

|X(tl+1i )− X(tli)| − 2−l+mc 3 ≤

"

r−lSq|X(tl+1i )− X(tli)|

d(tl+1i , tli) − 2−l+mc 6

#

+

− 2−l+mc 6

= CLpr−lSqϕ−1pp "

|X(tl+1i )− X(tli)|

CLpd(tl+1i , tli) −2−l+mrlc 6CLpSq

#

+

!

)− 2−l+mc 6

≤ CLpr−lSqϕ−1p (a−1l,mϕp(|X(tl+1i )− X(tli)|

Cd(tl+1i , tli) ))− bl,mc, where

al,m= ϕp(2−l+mrlc

6CLpSq ), bl,m= 1

62−l+m for l > m.

Then we apply [z− 1]+ ≤ z1/q for z≥ 0 and obtain

|X(tl+1i )− X(tli)| − 2−l+mc 3

≤ bl,mc

"

(bl,mc)−1CLpr−lSqϕ−1p (a−1l,mϕp(|X(tl+1i )− X(tli)| Cd(tl+1i , tli) ))− 1

#

+

≤ (bl,mc)q−1q (CLpr−l)1qS[ϕ−1p ]1q(a−1l,mϕp(|X(tl+1i )− X(tli)| Cd(tl+1i , tli) ))

= Bl,mcq−1q S[ϕ−1p ]1q(al,m−1ϕp(|X(tl+1i )− X(tli)| Cd(tl+1i , tli) )), (2.17)

where Bl,m= b

q−1 q

l,m(CLpr−l)1q. Note that similar estimate holds for|X(tl+1i−1)− X(tli−1)| − 2−l+mc/3.

Now observe that for i ∈ Jm, m > m0, using that |ti− ti−1| ≥ rm+1q S we get

(2.18)

N

X

l=m+1

Bl,mS =

N

X

l=m+1

b

q−1 q

l,m(CLpr−l)1qS≤ M2rm+1q S ≤ M2|ti− ti−1|,

(12)

where M2 = M2(r, p, q) is defined by

N

X

l=m+1

b

q−1 q

l,m(CLpr−l+m+1)1q =

N

X

l=m+1

(1

62−l+m)q−1q (CLpr−l+m+1)1q

≤ (121−qCLp)1q

X

l=0

(21−qq r1q)l = (121−qCLp)1q(1− 21−qq r1q)−1=: M2. If i∈ Jm, m≤ m0, l > m0 then the same argument works and we get

|X(tl+1i )− X(tli)| − 2−l+m0c 3

≤ ¯Blcq−1q S[ϕ−1p ]1q(¯a−1l ϕp(|X(tl+1i )− X(tli)| Cd(tl+1i , tli) )), (2.19)

where

¯

al = ϕp(2−l+m0rlc

6CLpSq ), ¯bl= 1

62−l+m0, ¯Bl= ¯b

q−1 q

l (CLpr−l)1q for l > m0. Moreover in this case |ti− ti−1| ≥ rm0+1q S and hence again

(2.20)

N

X

l=m0+1

lS ≤ M2rm0+1q S ≤ M2|ti− ti−1|.

Using (2.17) and (2.19) we estimate V2. Denoting

X

m=m0+1

X

i∈Jm

N

X

l=m+1

X

s∈{i−1,i}

=:X

1

and

m0

X

m=0

X

i∈Jm

N

X

l=m0+1

X

s∈{i−1,i}

=:X

2

we have

V2 ≤ cq−1q SX

1

Bl,m−1p ]1q(a−1l,mϕp(|X(tl+1s )− X(tls)| Cd(tl+1i , tli) )) + cq−1q SX

2

l−1p ]1q(¯al−1ϕp(|X(tl+1s )− X(tls)| Cd(tl+1i , tli) )).

Using (2.18) and (2.20) we estimate the sum of weights appearing in the expression estimating V2

X

1

Bl,m+X

2

l = S−1X

1

Bl,mS + S−1X

2

lS

≤ S−12M2

n

X

i=1

|ti− ti−1| + S−12M2

n

X

i=1

|ti− ti−1| ≤ 4M2.

(13)

Using Fact 1 for ϕp(xq) which is convex above Cp,q we get (2.21) V2 ≤ 4M2cq−1q S[ϕp−1]1q( ¯V2+ Dp,q), where

2= (4M2)−1X

1

Bl,ma−1l,mϕp(|X(tl+1s )− X(tls)| Cd(tl+1s , tls) ) + (4M2)−1X

2

l¯a−1l ϕp(|X(tl+1s )− X(tls)| Cd(tl+1s , tls) ).

Now note that for m≥ m0

Bl,ma−1l,m≤ Bl,m0a−1l,m0 = ¯Bla−1l .

Moreover again we use the fact that for the path approximation of Πn and given u ∈ Tl+1 the increment X(u)− X(πl(u)) can be used only twice and for a single ti from the partition. It implies that

(2.22) V¯2 ≤ V2 := (4M2)−1

X

l=m0+1

la−1l X

u∈Tl+1

ϕ(|X(u) − X(πl(u))| Cd(u, πl(u)) ).

Observe that by (2.1) and (2.4) EV2≤ (4M2)−1

X

l=m0+1

la−1l |Tl+1|

≤ (4M2)−1

X

l=m0+1

¯

a−1l ¯blq−1q (CLpr−l)1q2−1(rl+1q + 1)

≤ (4M2)−1(61−qCLp)1q

X

l=m0+1

2(−l+m0)q−1q ϕp(2−l+m0rlc 6CLpSq )−1. Therefore due to rm0+1c/Sq > 1 we obtain that

EV2≤ (4M2)−1(61−qCLpr)1q

X

l=1

2l′ 1−qq ϕp(2−lrm0+1+lc 6CLpSqr )−1

≤ (4M2)−1(61−qCLpr)1q

X

l=1

2l′ 1−qq ϕp (r/2)l 6CLpr

!−1

:= M3.

Finally, by (2.21), (2.22) and Fact 2, we get

(2.23) V2 ≤ 4M2max{M3, 1}cq−1q S[ϕ−1p ]1q(V2/M3+ Dp,q).

Note that EV2/M3≤ 1.

(14)

A similar argument can be used to bound increments in W2, namely for m > m0

|X(tm+1i )− X(tm+1i−1 )| −1 3c

≤ CLpr−m−1Sqϕ−1pp([|X(tm+1i )− X(tm+1i−1 )|

LpCd(tm+1i , tm+1i−1 ) − rm+1c

6CLpSq]+))− c 6

≤ CLpr−m−1Sqϕ−1p (¯am−1ϕp(|X(tm+1i )− X(tm+1i−1 )| Cd(tm+1i , tm+1i−1 ) ))− c

6, where

ˆ

am = ϕp( rm+1c

6CLpSq) for m > m0. Then by [z− 1]+≤ z1q for z≥ 0

|X(tm+1i )− X(tm+1i−1 )| − c 3

≤ c 6

"

6c−1CLpr−m−1Sqϕ−1p (ˆa−1m ϕp(|X(tm+1i )− X(tm+1i−1 )| Cd(tm+1i , tm+1i−1 ) ))− 1

#

+

≤c 6

q−1q

(CLpr−m−1)1qS[ϕ−1p ]1q(ˆa−1m ϕp(|X(tm+1i )− X(tm+1i−1 )| Cd(tm+1i , tm+1i−1 ) )).

Let M4 := (61−qCLp)1q. Using that rm+1q S≤ |ti− ti−1| ≤ rmqS we get

X

m=m0+1

X

i∈Jm

(61−qCLpr−m−1)1qS≤ M4 n

X

i=1

|ti− ti−1| = M4S.

Therefore using Fact 1 for ϕp(xq) we get

(2.24) W2≤ M4cq−1q S[ϕp−1]1q( ¯W2+ Dp,q), where

2 := M4−1

N

X

m=m0+1

X

i∈Jm

ˆ

a−1m (61−qCLpr−m−1)1qϕp(|X(tm+1i )− X(tm+1i−1 )| Cd(tm+1i , tm+1i−1 ) ).

Clearly W¯2 ≤ W2

:= M4−1

X

m=m0+1

ˆ

a−1m(61−qCLpr−m−1)1q X

u∈Tm+1

X

v∈Im+1(u)

ϕp(|X(u) − X(v)|

Cd(u, v) ).

Note that by (2.4), (2.7) X

u∈Tm+1

|Im+1(u)| ≤ 2−1B(r, q)(rm+1q + 1)≤ B(r, q)rm+1q .

(15)

Moreover rm0+1c/Sq > 1 and hence EW2 = M4−1B(r, q)

X

m=m0+1

ˆ

a−1m (61−qCLpr−m−1)1qrm+1q

≤ M4−1(61−qCLp)1qB(r, q)

X

m=m0+1

ϕp( rm+1c 6CLpSq)−1

≤ B(r, q)

X

m=1

ϕp rm 6CLp

!−1

=: M5. By (2.24) and Fact 2 we get

(2.25) W2 ≤ M4max{M5, 1}cq−1q S[ϕ−1p ]1q(W2/M5+ Dp,q).

Since EW2/M5 ≤ 1 by (2.23), (2.25) and Jensen’s inequality we obtain the thesis.

 Now we are ready to finish the proof of Theorem 1.

Proof.(of Theorem 1) It suffices to use Lemma 1, then universal bounds given in Lemmas 2, 3 and finally let N → ∞. Recall that by the construction limN →∞d(t, πN +1(t)) = 0 for any t ∈ T . Note that we can optimize the inequalities with respect to r≥ 4 and therefore constants K1 and K2 depend on p and q only.

 The meaning of the result the is that for p > 0 and 0 < q < 1 there holds some concentration inequality. To formulate the results in the elegant way observe that there exists Eq ∈ [0; 1] such that Eq+ x1q ≥ x for x ≥ 0 and hence, due to Dp,q ≤ Dp, we get

(2.26) Eq+ [ϕ−1p ]1q(x + Dp,q)≥ ϕ−1p (x + Dp) for x≥ 0.

As the consequence of (2.26) and Jensen’s inequality we get

Corollary 1. Under the assumptions of Theorem 1 there exist r.v. Z such that Z ≥ 0, EZ ≤ 1 and for some constants A(p, q), B(p, q) the following estimate holds

T Vc(X, S)≤ cq−1q S[A(p, q) + B(p, q)[ϕ−1p ]1q(Z + Dp,q)].

Application of Chebyshev’s inequality immediately gives Corollary 2. The following inequality holds

P

T Vc(X, S)≥ cq−1q S[ ¯A(p, q) + ¯B(p, q)u]

≤ ¯Dp,qexp(−upq), for u > 0.

where ¯A(p, q), ¯B(p, q) are universal constants, e.g. ¯A(p, q) = A(p, q)+(2/ ln 2)pq1 B(p, q), B(p, q) = (2/ ln 2)¯ pq1B(p, q) and ¯Dp,q = Dp,q+ 1. In particular ¯Dp,q = 1 for

pq≥ 1.

(16)

3. Application to the fractional Brownian motion

Now we consider T = [0, S] with distance d(s, t) = |t − s|H, where H ∈ (0, 1). Let WH(t), t ∈ T, be a fractional Brownian motion with the Hurst coefficient H. Then for some constant C(H)

2(|WH(t)− WH(s)|

C(H)|t − s|H )≤ 1, for s, t ∈ T.

Consequently all the assumption of Theorem 1 are satisfied and we get Corollary 3. For any fractional Brownian motion WH(t), t ∈ T, the fol- lowing inequality holds

P

T Vc(WH, S)≥ cH−1H S(AH + BHu)

≤ CHexp(−u2H), for u > 0.

where AH, BH, CH are universal constants and CH = 1 for H ≥ 1/2.

Note that Corollary 3 implies that ET Vc(WH, S)≤ KHcH−1H S. On the other hand cH−1H S is also the proper lower bound for ET Vc(X, S). Indeed, let us consider the partition Π given by the entropy N (T, d, c), i.e. Π such that |Π| = N(T, d, c) and S

t∈ΠBd(t, c) = T . In particular the construction gives that d(s, t)≥ c for s 6= t, s, t ∈ Π. Thus denoting Π = {t1, ..., tN} and 0 = t0 < t1< ... < tN < S we have

T Vc(WH, S)≥

N

X

i=1

(|X(ti)− X(ti−1)| − c)+.

Clearly N ∼ cH1S and E(|WH(ti) − WH(ti−1)| − c)+ ∼ c due to c ≤ d(ti, ti−1)≤ 2Hc. It proves that cH−1H S is comparable with ET Vc(WH, S) up to a constant depending only on H. Therefore we have another formulation of Corollary 1.

Corollary 4. For any fractional Brownian motion WH(t), t ∈ T, the fol- lowing inequality holds

P T Vc(WH, S)≥ ET Vc(WH, S)( ¯AH + ¯BHu) ≤ ¯CHexp(−u2H), for u > 0, where ¯AH, ¯BH, ¯CH are constants. Moreover ¯CH = 1, for H ≥ 1/2.

4. Application to the standard Brownian motion and diffusions For a standard Brownian motion W = W1/2, which is the only fractional Brownian motion with independent increments one may, using this property, strengthen the results obtained for general fBm and obtain Gaussian con- centration of T Vc(W, S). Generalization for diffusions driven by W, with moderate growth is also possible.

Let us assume that Xt, t≥ 0, is a one-dimensional diffusion satisfying (4.1) X (t) = x0+

Z t 0

µ (s, X (s)) ds + Z t

0

σ (s, X (s)) dW (s)

(17)

We assume that σ : [0; +∞) × R → [−R; R] is measurable and bounded (i.e.

0 < R < +∞) and µ : [0; +∞) × R → R is measurable and satisfying the following linear growth condition: there exists C, D ≥ 0 such that for all t≥ 0

(4.2) |µ (t, x)| ≤ C + D |x| .

We will also need the natural assumption that X is a Markov process. With this assumption we have

Theorem 2. For X being a Markov process satisfying (4.1) with µ and σ as above and λ≥ 0 one has

Eexp (λT Vc(X, S)) ≤ 2 exp λ2R+ λSc−1βR+ λγx0,C,D,S ×

× 1 + 8ληD,R,Sexp λ2η2D,R,S ,

where γx0,C,D,S = (C + D|x0|) SeDS, δD,S = DSeDSand ηD,R,S = δD,SRpS/2.

In particular, for D = 0 we get

Eexp (λT Vc(X, S))≤ 2 exp λ2R+ λS c−1βR+ C

and for the standard Brownian motion X = W we get

(4.3) Eexp (λT Vc(W, S))≤ 2 exp λ2Sα + λSc−1β , where α, β are universal constants.

Proof. We have Let us define M (t) :=

Z t 0

µ (s, X (s)) ds, Y (t) :=

Z t 0

σ (s, X (s)) dW (s)

and Y= sup0≤s≤S|Y (s)| . We have X (t) = x0+ M (t) + Y (t) , and due to (4.2) we estimate

|M (t)| ≤ Z t

0 |µ (s, X (s))| ds ≤ Z t

0

C + D|X (s)| ds

≤ Z t

0

C + D|x0| + D |M (s)| + DYds

≤ (C + D |x0| + DY) S + D Z t

0 |M (s)| ds.

(4.4)

Hence, from Gronwall’s lemma we get

(4.5) |M (t)| ≤ (C + D |x0| + DY) SeDt.

Notice that due to (4.5) M is adapted, absolute continuous process with locally bounded total variation. Indeed, repeating estimates (4.4) and using

Cytaty

Powiązane dokumenty

The parameter σ α has appeared in many papers on exponential sums but we are not aware of an upper bound of the type (1.13) ever appearing before, even for the case of

As mentioned in Section 5, the plan is to apply Theorem 3.1 to equations of the form (3.2) defined by certain minimal forms L ∈ L(T ).. Since we must apply transformations to the

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

Key words and phrases: graded-commutative algebras, supermanifolds, Levi flat su- per CR structure, locally direct sheaf, super real integrable distribution, super complex

The two ways of regularizing the integral (26), described above, coincide if and only if we apply formula (24) when we perform differentiation before integration.. Integrals of

The volume of a simplex in a hyperbolic space is a hyperlogarithmic function of basic algebraic invariants, as a simple consequence of the Schl¨ afli formula.. However, there

M u sialek, The Green's function and the solutions of the Neumann and Dirichlet problem,