• Nie Znaleziono Wyników

On Harnack inequality for α-stable Ornstein-Uhlenbeck processes.

N/A
N/A
Protected

Academic year: 2022

Share "On Harnack inequality for α-stable Ornstein-Uhlenbeck processes."

Copied!
25
0
0

Pełen tekst

(1)

On Harnack inequality for α-stable Ornstein-Uhlenbeck processes.

Tomasz Jakubowski

Institute of Mathematics and Computer Science Wroc law University of Technology

Wybrze˙ze Wyspia´ nskiego 27, 50-370 Wroc law, Poland e-mail: Tomasz.Jakubowski@pwr.wroc.pl

Abstract

We consider the α-stable Ornstein-Uhlenbeck process in Rd with the generator L = ∆α/2− λx · ∇x. We show that if 2 > α ≥ 1 or α < 1 = d the Harnack inequality holds. For α < 1 < d we construct a counterexample that shows that the Harnack inequality doesn’t hold.

keywords: α-stable Ornstein-Uhlenbeck process, harmonic func- tions, Harnack inequality.

AMS Subject Classification: 60J75 60J45

1 Introduction

We consider the α-stable Ornstein-Uhlenbeck process in Rd with 0 < α < 2 given by the stochastic integral

Xt= e−λtX0+ Z t

0

e−λ(t−s)d bXs,

partially supported by KBN and MEN

(2)

where bXt is the isotropic α-stable process (for more details see section Pre- liminaries). The process Xtwas considered by Janicki and Weron in [12] and Samorodnitsky and Taqqu [13]. In the previous paper [11] the author gives the estimates of the mean first exit time of Xtfrom the ball. In this paper we are interested in the behaviour of harmonic functions related to the process Xt, which will be called L-harmonic (for definition see section Preliminar- ies). One of the fundamental questions concerning harmonic functions of the process Xt is the existence of the Harnack inequality which says

Harnack inequality. Let M > 0. There exists a constant C1 depend- ing only on d, α, λ, M such that for any x0 ∈ B(0, M ), r ∈ (0, 1/2) and any bounded function h nonnegative on Rd and L-harmonic in B(x0, 2r) it holds h(y) ≤ C1h(x), x, y ∈ B(x0, r/2) . (1) The answer to this question is positive for many Markov processes. It is well known that for Brownian motion Harnack inequality holds (see for example [8]). The first result on Harnack inequality for processes with jumps was obtained for the process bXt (for the proof see for example the paper of Bogdan, Byczkowski [3]). The proof bases on exact formula of density of harmonic measure of the ball. In paper [1] Bass and Levin prove Har- nack inequality for processes with symmetric L´evy measure comparable with L´evy measure of bXt. Completely different methods based on so-called reg- ularization of Poisson kernel were used by Bogdan, St´os and Sztonyk in [4]

for α-stable processes on d-sets with some conditions on α. In [6] Chen and Kumagai generalize results from [4] on α ∈ (0, 2). In recent paper [14] Song and Vondracek applying methods used in [1] and [6] prove Harnack inequal- ity for a class of jump Markov process satisfying certain conditions. We note that in all papers above the constant C1 doesn’t depend on M . Since the process Xt is not space invariant we cannot avoid this dependence which will be justified in section 5. The main results of this paper are

Theorem 1. For α < 1 < d Harnack inequality (1) doesn’t hold. It means that there exist a sequence of nonnegative functions hn, L-harmonic in B(x0, 1/n) and a sequence of points yn such that hn(yn) → 0 while n → ∞ and hn(x0) >

c > 0 for some constant c.

Theorem 2. For α ≥ 1 or α < 1 = d Harnack inequality (1) holds.

(3)

In [5] and [7] the authors consider the pure jump Girsanov transform and drift transform to show that the Green functions of the initial process and transformed process are comparable. From Theorem 1 we deduce that we cannot obtain the α-stable Ornstein-Uhlenbeck process from the process bXt by pure jump Girsanov transform and drift transform at least in the case d > 1 > α and the mentioned papers do not apply to the process Xt.

The work is organized as follows. In section 2 and 3 we introduce basic definitions and facts. In section 4 we deal with the case α < 1. First, for d > 1, we construct a counterexample and show Theorem 1. Next, adapting some methods from [4], we show Theorem 2 for d = 1 > α. The section 5 is devoted to the case α ≥ 1. We prove Theorem 2 for α ≥ 1 by showing that Xt satisfies Song-Vondracek conditions from [14].

2 Preliminaries

By Rd we denote the d-dimensional Euclidean space with norm | · |. We denote a ∧ b = min(a, b). All constants are positive. In this paper we use a convention that constants denoted by small letters may differ in each lemma, while constants denoted by capital letters don’t change.

We denote by ( bXt, Px) the isotropic α-stable L´evy process in Rd with index of stability α ∈ (0, 2) and characteristic function

E0ei bXt·ξ = e−t|ξ|α, ξ ∈ Rd, t ≥ 0 . (2) As usual, Ex denotes the expectation with respect to the distribution Px of the process starting from x ∈ Rd. We always assume that sample paths of bXt are right continuous and have left limits almost surely. ( bXt, Px) is a Markov process with transition densities p(t, x, y) satisfying Pb x( bXt ∈ D) = R

Dp(t, x, y)dy, and is strong Markov with respect to the so-called ”standardb filtration” (Ft, F ) (see [2]).

We consider the α-stable Ornstein-Uhlenbeck process in Rd with 0 < α <

2 given by the stochastic integral Xt= e−λtX0+

Z t 0

e−λ(t−s)d bXs, Xb0 = 0 ,

where the integral is in Stieltjes sense. Hence we may rewrite this equation

(4)

in the form (see Corollary 5 in [11]) Xt= e−λtX0+ bXt− λ

Z t 0

e−λ(t−s)Xbsds , Xb0 = 0 , (3)

Xt is a Markov process (see [12], [13]) with transition density (see [11]) p(t, x, y) = bp

1−e−αλt

αλ , e−λtx, y

, t > 0, x, y ∈ Rd. (4) We have the following estimates for the function p(t, x, y) from [4]b

C2−1

 t

|x − y|d+α ∧ t−d/α



≤p(t, x, y) ≤ Cb 2

 t

|x − y|d+α ∧ t−d/α



, (5) for t > 0, x, y ∈ Rd and some constant C2 depending only on d and α.

Proposition 3. Xt is a Feller and strong Feller process. In particular the transition semigroup Pt generated by the process Xt is a strongly continuous contraction semigroup on C0(Rd).

Proof. Let f ∈ L(Rd) and xn → x. Then

|Ptf (xn) − Ptf (x)| = Z

Rd

(p(t, xn, y) − p(t, x, y))f (y) dy

≤ ||f || Z

B(xe−λt,r)

|p(t, xn, y) − p(t, x, y)| dy + ||f ||

Z

B(xe−λt,r)c

|p(t, xn, y) − p(t, x, y)| dy

By (4) and (5) the second integral in the last inequality is arbitrary small for large r while the first integral tends to 0 by bounded convergence. Hence for f ∈ L(Rd) the function Ptf (x) is continuous in x. Obviously Ptf (x) is also bounded which gives the strong Feller property. To show the Feller property it suffices to show that for any f ∈ C0(Rd) it holds

lim

|x|→∞Ptf (x) = 0 (6)

limt→0||Ptf − f || = 0 (7)

(5)

For any ε > 0 there exists r > 0 such that |f (y)| < ε if |y| > r. Hence

|Ptf (x)| ≤ ε + ||f ||

Z

B(0,r)

p(t, x, y) dy.

The last integral tends to 0 while |x| → ∞, proving (6). Now we have

||Ptf − f || ≤ sup

x∈Rd

Z

Rd

p(t, x, y)|f (y) − f (x)| dy .

Since f ∈ C0(Rd), it is uniformly continous on Rd. For any ε > 0 there exists δ > 0 such that |f (x) − f (y)| < ε if |x − y| < δ. Moreover there is r > 0 such that |f (x)| < ε for |x| ≥ r. We note that for |x| < 2r and t < δ/(4rλ) we have |x − y| < δ if |xe−λt− y| < δ/2. Therefore

sup

x∈B(0,2r)

Z

Rd

p(t, x, y)|f (y) − f (x)| dy

≤ ε + 2||f || sup

x∈B(0,2r)

Z

|xe−λt−y|≥δ/2

t

|xe−λt− y|d+αdy

The second term in the last line tends to 0 when t → 0. For |x| > 2r and t ≤ ln(4/3)/λ we have e−λt|x| > 3r/2. Hence

sup

x∈B(0,2r)c

Z

Rd

p(t, x, y)|f (y) − f (x)| dy

≤ 2ε + (ε + ||f ||) sup

x∈B(0,2r)c

Z

B(0,r)

t

|xe−λt− y|d+αdy .

Like in the previous case the last term tends to 0 when t → 0, which gives (6).

For an open set D ⊂ Rd, we define a Markov time τD = inf{t ≥ 0 : Xt∈ Dc}, the first exit time from D. If |D| < ∞ then PxD < ∞} = 1, x ∈ Rd. In this case the Px distribution of XτD is a probability measure on Dc. We define the function pD(t, x, y) by

pD(t, x, y) = p(t, x, y) − Ex(p(t − τD, XτD, y)1D<t}) .

The function pD(t, x, y) is a density of the semigroup PtD of the process Xt killed on exiting the set D given by

PtDf (x) = Ex(f (Xt)1t<τD) .

(6)

By standard argument (see [8]) we obtain that for any open set A it holds PxD > t , Xt∈ A) =

Z

A

pD(t, x, y) dy , t > 0, x ∈ Rd (8) and the function pD satisfies the Chapman-Kolmogorov equation

pD(t, x, y) = Z

D

pD(t − s, x, z)pD(s, z, y) dz , t > s > 0, x, y ∈ Rd. (9) Using the method given in [4] we also get that pD(t, x, y) > 0 for all t > 0 and x, y ∈ Rd.

Since for any r > 0 and x, z ∈ Rd we have ExB(z,r)) < ∞ (see [11]), we get

ExB) = Z

0

PxB > t) dt = Z

0

Z

B

pB(t, z, y) dy dt

= Z

B

Z 0

pB(t, z, y) dt

 dy .

Hence we may define the Green function of the set B by GB(x, y) =

Z 0

pB(t, z, y) dt , and we have the equality

ExB) = Z

B

GB(x, y) dy . (10)

We will also use the estimates for ExB(x,r)) established in [11]:

Lemma 4. Let 0 < r < M . There exists a constant C3 depending on d, α, λ, M such that for all x ∈ Rd it holds

1 C3



rα∧ r

|x|



≤ ExB(x,r)) ≤ C3



rα∧ r

|x|



Since Xt is a Markov process, it may be identified by its infinitesimal generator denoted in this paper by L. We will show that L = ∆α/2− λx · ∇x, where

α/2ψ(x) = Ad,α P.V.

Z

Rd

ψ(y) − ψ(x)

|x − y|d+α dy

is the infinitesimal generator of the process bXt and Ad,α is some constant.

(7)

Lemma 5. For any bounded function f ∈ Cc(Rd) we have limt→0

1 t

Z

Rd

p(t, x, y)f (y) dy − f (x)



= ∆α/2f (x) − λx · ∇xf (x) , x ∈ Rd. Proof. Denote β = αλ and (t)β = 1−eβ−βt. By (2) and (4) we have

Z

Rd

eiy·ξp(t, x, y) dy = Z

Rd

eiy·ξp((t)b β, 0, y − e−λtx) dy = e−(t)β|ξ|α+ie−λtx·ξ. Hence

p(t, x, y) = 1 (2π)d

Z

Rd

exp −(t)β|ξ|α+ ie−λtx · ξ e−iξ·ydξ . (11) If we denote the Fourier transform of f by F (f )(ξ) = R

Rde−ixξf (x) dx we obtain

Z

Rd

p(t, x, y)f (y) dy − f (x)

= 1

(2π)d Z

Rd

Z

Rd

exp −(t)β|ξ|α+ ie−λtx · ξ e−iξ·yf (y) dξ dy − f (x)

= 1

(2π)d Z

Rd

e−(t)β|ξ|α+ie−λtx·ξF (f )(ξ) dξ − 1 (2π)d

Z

Rd

eiξ·xF (f )(ξ) dξ

Since f ∈ Cc(Rd) then |F (f )(ξ)| ≤ c1(1 + |ξ|)−N for all N ∈ N and some constant c1 depending on N . We have estimates

e−(t)β|ξ|α+i(1−e−λt)x·ξ− 1 t

≤ |(t)β|ξ|α/t + i(1 − e−λt)x · ξ/t| ≤ |ξ|α+ λ|x||ξ| . Hence by dominated convergence theorem

limt→0

1 t

Z

Rd

p(t, x, y)f (y) dy − f (x)



= 1

(2π)dlim

t→0

1 t

Z

Rd



e−(t)β|ξ|α+ie−λtx·ξ − eiξ·x

F (f )(ξ) dξ

= 1

(2π)d Z

Rd

limt→0

1 t



e−(t)β|ξ|α+ie−λtx·ξ − eiξ·x

F (f )(ξ) dξ

= 1

(2π)d Z

Rd

(|ξ|α− iλx · ξ)eix·ξF (f )(ξ) dξ

= ∆α/2f (x) − λx · ∇xf (x) .

(8)

Therefore the infinitesimal generator of Xtis equal to L = ∆α/2− λx · ∇x. Definition 6. Let u ∈ B(Rd). We say that a function u is harmonic with respect to Xt in an open set D ⊂ Rd if

u(x) = Exu(X(τU)), x ∈ U, for every bounded open set U satisfying U ⊂ D.

Since L is the infinitesimal generator of the process Xt, harmonic func- tions with respect to Xt will be called L-harmonic.

3 L´ evy system and Ikeda-Watanabe formula

One of the main tools in this paper is the following Ikeda-Watanabe formula expressing the probability that the process on exiting set B hits set A Proposition 7 (Ikeda-Watanabe formula). Let B be an open nonempty bounded set in Rd. For any Borel set A ⊂ ¯Bc it holds

Px(XτB ∈ A) = Cd,α Z

A

Z

B

GB(x, z)

|y − z|d+αdz dy , (12) where Cd,α is some constant.

To prove it we need only to check the following assumptions from [10].

(A1) The semigroup Ptf (x) =R

Rdp(t, x, y)f (y) dy maps C0(Rd) into C0(Rd) and is strongly continuous in t ≥ 0.

(A2) There exists a positive kernel n(x, E), x ∈ Rd, E ⊂ Rd a Borel subset, such that

(a) If dist(x, E) > 0 then n(x, E) < ∞.

(b) For f ∈ C0(Rd) and a bounded open set D with dist(D, supp f ) >

0,

sup

x∈D,t>0

t−1Ptf (x) < ∞ and

lim

t→0+t−1Ptf (x) = Z

Rd

f (y)n(x, dy) for every x ∈ D.

(9)

A family of measures n(x, E) is called a L´evy system of Xt. For any x, y ∈ Rd define N (x, y) = limt→0+t−1p(t, x, y).

Lemma 8. We have N (x, y) = Cd,α|x − y|−d−α, for some constant Cd,α. Proof. It suffices to show that (p(t, x, y) − p(t, x, y))/t → 0 when t → 0.b Denote β = αλ and (t)β = 1−eβ−βt. By (11) we have

p(t, x, y) −p(t, x, y)b

t = 1

(2π)d Z

Rd

e−(t)β|ξ|α+ie−λtx·ξ− e−t|ξ|α+ix·ξ

t e−iξ·y

= Z

Rd

e−((t)β−t)|ξ|α−i(1−e−λt)x·ξ− 1

(2π)dt e−t|ξ|α+i(x−y)·ξ

Like in the proof of Lemma 5 we may enter with limit under the integral and we obtain the assertion of the Lemma.

For any Borel set E ⊂ Rd and x ∈ Rd we define n(x, E) =R

EN (x, y) dy, hence the function N (x, y) is a kernel of a L´evy system n(x, E). Now we prove Proposition 7.

Proof. Assumption (A1) is satisfied by Proposition 3. Clearly for any Borel set E ⊂ Rd it holds n(x, E) < ∞ if x 6∈ E. Let D be open bounded subset of Rd and let f ∈ C0(Rd) be such that dist(D, supp f ) ≥ δ > 0. There exists t0 > 0 such that dist(e−λtx, supp f ) > δ/2 for all t < t0 and x ∈ D. Hence by (4) and (5)

sup

x∈D,t0>t>0

t−1Ptf (x) ≤ c1 sup

x∈D,t0>t>0

Z

Rd

f (y) dy

|eλtx − y|d+α ≤ c2||f ||δ−α < ∞ Next we have supx∈D,t≥t0t−1Ptf (x) ≤ ||f ||/t0 < ∞. At the end, when x ∈ D, by bounded convergence

lim

t→0+t−1Ptf (x) = lim

t→0+t−1 Z

Rd

f (y)p(t, x, y) dy = Z

Rd

f (y)N (x, y) dy

= Z

Rd

f (y)n(x, dy) .

(10)

4 The case α < 1

In this section we will deal with α < 1. In this case the properties of the process Xt are diffrent from the case α ≥ 1. It may be seen from esti- mates of the mean first exit time from the ball (Lemma 4). When r → 0 then ExB(x,r)) is comparable with r for any x 6= 0, while in the case α ≥ 1 we have asymptotics rα. As a result of this fact we may construct a contrexemple to show the Theorem 1. To do it we will need some aux- iliary facts. Let a > 0 and x0 = (a, 0, 0, . . . , 0). Denote B = B(x0, r), S = {z ∈ Rd: pz22+ . . . + z2d < r/8} and S = {z ∈ S : z1 < a + r/8}.

Lemma 9. Let α < 1. There exists a constant C4 such that Px0(XτB∩S ∈ S) ≥ 1 − C4r1−α for 0 < r < a/4.

Proof. There exists a constant c1 independent from a such that |e−λc1r/ax0− x0| > 2r for r < a/4, which means that B(e−λc1r/ax0, r) ∩ B = ∅. Let t = c1r/a. By L´evy inequality and (5) we have

P0(sup

s≤t

| bXs| > r/16) ≤ 2P0(| bXt| > r/16)

≤ 2c2 Z

B(0,r/16)c

t

|y|d+α ≤ c3r/a rα = c3

ar1−α. Using the representation (3) we have

Px0(|Xs− eλsX0| < r/8, ∀s ≤ t)

= P0(| bXs− λ Z s

0

e−λ(s−u)Xbudu| < r/8, ∀s ≤ t)

≥ P0(| bXs| + λ Z s

0

e−λ(s−u)| bXu| du < r/8, ∀s ≤ t)

≥ P0(sup

s≤t

| bXs|(1 + λ Z s

0

e−λ(s−u)du) < r/8, ∀s ≤ t)

≥ P0(sup

s≤t

| bXs| < r/16) ≥ 1 − c3

ar1−α Therefore putting C4 = c3/a

Px0(XτB∩S ∈ S) ≥ Px0(Xs∈ S, ∀s ≤ t and Xt6∈ B)

≥ Px0(|Xs− e−λsX0| < r/8, ∀s ≤ t)

≥ 1 − C4r1−α.

(11)

Denote A = {z ∈ S∩ Bc: dist(z, ∂B) ≤ r/8}.

Corollary 10. There exists a constant C5 such that Px0(XτB ∈ A) ≥ 1 − C5r1−α for 0 < r < a/4.

Proof. By Ikeda-Watanabe formula and Lemma 4

Px0(XτB ∈ B(x0, 9r/8)c) = Cd,α Z

B(x0,9r/8)c

Z

B

GB(x0, z)

|z − y|d+αdz dy

≤ c1 Z

B(x0,9r/8)c

Z

B

GB(x0, z)

|z − x0|d+αdz dy

= c2

rαEx0B) ≤ c2C3/|x0|r1−α Therefore by Lemma 9 we obtain

Px0(XτB ∈ A) ≥ Px0(XτB∩S ∈ A)

= Px0(XτB∩S ∈ S) − Px0(XτB∩S ∈ S\ A)

≥ Px0(XτB∩S ∈ S) − Px0(XτB∩S ∈ B(x0, 9r/8)c)

≥ 1 − C4r1−α− c2C3/|x0|r1−α = 1 − (C4+ c2C3/|x0|)r1−α

We note that C5 = c/a, where c is independent from a.

Let d > 1. For any x ∈ Rd we consider the rotation Tx around 0 such that Txx = (|x|, 0, . . . , 0). Let y0 ∈ B(x0, r) be such that |y0| = a and

|y0 − x0| = r/3. We note that Ty−1

0 x0 = y0. By the same argument as in Lemma 9 we obtain that there is C6 such that

Py0(X(τB∩T−1

y0 S) ∈ Ty−10 S) ≥ 1 − C6r1−α (13) for 0 < r < a/4.

Lemma 11. For r < 2a/9 we have A ∩ Ty−10 S= ∅.

Proof. Let xt = (1 − t)x0 and yt = (1 − t)y0. It suffices to show that

|xt− yt| > r/4 for 0 ≤ t ≤ 9r/8a. Since |xt− yt| is decreasing as a function of

(12)

t it suffices to verify inequality |(1 − 9r/8a)x0− (1 − 9r/8a)y0| > r/4. Since

|x0− y0| = r/3 we have

|(1 − 9r/8a)x0− (1 − 9r/8a)y0| = a − 9r/8

a |x0− y0| > a − a/4 a

r 3 = r

4.

Now we prove Theorem 1.

Proof of Theorem 1. Let x0 = (1, 0, . . . , 0), rn = 1/n, Bn = B(x0, rn), Sn = {z ∈ Rd: pz22+ . . . + zd2 < rn/8} and Sn− = {z ∈ Sn: z1 < 1 + rn/8}. De- note An= {z ∈ Sn−∩Bnc: dist(z, ∂Bn) ≤ rn/8}. Let hn(x) = Px(XτBn ∈ An).

Each function hn is L-harmonic in Bn. Indeed by strong Markov property for any open set E such that E ⊂ Bn it holds

Ex(hn(XτE)) = Ex PXτE(XτBn ∈ An) = Px(XτBn ∈ An) = hn(x) . Let yn ∈ B(x0, rn) be such that |yn| = 1 and |yn− x0| = rn/3. By Lemma 11 and (13) we have

h(yn) = 1−Pyn(XτBn 6∈ An) ≤ 1−Pyn(X(τBn∩Ty−1n Sn) ∈ Ty−1n Sn−) ≤ C6r1−αn . But by Corollary 10 it holds

hn(x0) ≥ 1 − C5r1−αn .

If n → ∞ then hn(yn) → 0 and hn(x0) → 1, hence there is no constant c such that hn(yn) ≥ chn(x0) for all n ∈ N. This shows that the Harnack inequality (1) does not hold for α < 1 < d.

Now we will show that the process on exiting the ball B(x0, r) not con- taining the point 0 hits the boundary of B(x0, r) with the probability bigger than 0. It means that the harmonic measure of the ball not containing 0 has a nontrivial singular part on the boundary of the ball.

Proposition 12. Let 0 6∈ B(x0, r) and x ∈ B(x0, r). Then Px(X(τB(x0,r)) ∈

∂B(x0, r)) > 0.

(13)

Proof. Since the process is invariant by rotation around 0 we may and do suppose x0 = (|x0|, 0, . . . , 0). For any z ∈ Rddenote δ(z) = dist(z, ∂B(x0, r)).

Let S = S ∩ {z ∈ B(x0, r) : δ(z) < |z|/4, C5(δ(z)/2)1−α < 1}, where C5

is the constant from Corollary 10 corresponding to a = dist(0, B(x0, r)). For any x ∈ B(x0, r) denote

Bx = B(x, δ(x)/2), Ix = {tx : t ∈ [0, 1]},

Ax = {z ∈ Bxc: dist(z, Ix) < δ(x)/8, dist(z, ∂Bx) < δ(x)/8}.

First suppose that x ∈ S. By Corollary 10 and invariance by rotation around 0 it holds

Px(XτBx ∈ Ax) ≥ 1 − C5(δ(x)/2)1−α. (14) We note that C5 depends on dist(0, B(x0, r)). Define a stopping time

T =

BX0 if X(τBX0) ∈ AX0,

∞ if X(τBX0) 6∈ AX0,

We use a convention that X = ∂, where ∂ is a cementary point. Let T0 = 0, T1 = T and Tn+1 = T ◦ θTn, where θ is a shift operator acting on the space of trajectories by θt(ω)(s) = ω(t + s). Consider an event

n= {XT ∈ AX0, XT2 ∈ AXT, . . . , XTn ∈ AX

T n−1} . We claim that for some constant 1/2 < p < 1 it holds

Px(Ωn) ≥

n

Y

k=1

1 − C5



pk−1δ(x) 2

1−α!

. (15)

Actually p will be equal 5/8. We prove (15) by induction. For n = 1 (15) becomes (14). Now suppose that (15) holds for n. For given x ∈ B(x0, r) we define a set Anx = {y ∈ Az: z ∈ An−1x }, where A1x = Ax. We note that supz∈Axδ(z) ≤ 58δ(x). Hence by induction supz∈An

xδ(z) ≤ 58n

δ(x). Let p = 58. Since AXT n ⊂ AnX

0 by strong Markov property, (14) and (15) we

(14)

obtain

Px(Ωn+1) = Ex(PXT n(XT ∈ AX0)1n)

≥ Ex( inf

z∈AXT n Pz(XT ∈ AX0)1n)

≥ Ex( inf

z∈AnX0Pz(XτBz ∈ Az)1n)

= Px(Ωn) inf

z∈AnxPz(XτBz ∈ Az)

n

Y

k=1

1 − C5



pk−1δ(x) 2

1−α!

z∈Ainfnx(1 − C5(δ(z)/2)1−α)

n+1

Y

k=1

1 − C5



pk−1δ(x) 2

1−α!

Since supz∈An

xδ(z) ≤ 58n

δ(x), hence the event Ωn yields that the process Xt hasn’t exited the ball B(x0, r), but went out from B(x0, r − (5/8)nδ(x)).

Therefore

= {XTn ∈ AXT n−1, ∀n ≥ 0} ⊂ {XτB(x0,r) ∈ ∂B(x0, r)}

So we obtain that

Px(XτB(x0,r) ∈ ∂B(x0, r)) ≥ Px(Ω) ≥

Y

k=1

1 − C5



pk−1δ(x) 2

1−α!

It remains to prove that the last expression is bigger than 0. We calculate the logarithm of it. Since C5(δ(x)/2)α−1 < 1 it holds

ln

Y

k=1

1 − C5



pk−1δ(x) 2

1−α!!

=

X

k=1

ln 1 − C5



pk−1δ(x) 2

1−α!

X

k=1

−2C5



pk−1δ(x) 2

1−α

= −δ(x)1−αp > −∞ , where p = 2αC5/(1 − p1−α). Therefore for x ∈ S it holds

Px(XτB(x0,r) ∈ ∂B(x0, r)) > e−δ(x)1−αp > 0. (16)

(15)

Now let x 6∈ S. Let ρ = supz∈Sδ(z). By strong Markov property and (16)

Px(XτB ∈ ∂B) ≥ Ex(PX(τB\S∗)(XτB ∈ ∂B); XτB\S∗ ∈ S)

> e−ρ1−αpPx(XτB\S∗ ∈ S) > 0 .

Remark 13. In fact we proved that the process, on exiting a ball B(x0, r), hits the set ∂B(x0, r) ∩ Tx−1

0 {z ∈ Rd: |z1| ≤ |x0|} with probability bigger than 0.

Denote σD = τDc the first hitting time of the set D of the process Xt. Corollary 14. Let d = 1 > α. If y 6= 0 then Px{y} < ∞) > 0.

Proof. Since Xt is invariant by rotation around 0 we may and do assume that y > 0. If x > y then by Proposition 12 we have

Px{y} < ∞) ≥ Px{y} < τB(x,2|x−y|)) ≥ Px(XτB(x,x−y) = y) = c1 > 0 , Now let x < y. Let c2 = 2C51/(α−1). We note that if z ∈ (y, 5y/4 ∧ y + c2) then z − y < |z|/4 and C5((z − y)/2)1−α < 1. Hence by strong Markov property and (16)

Px{y} < τB(x,2|y−x|))

≥ Ex

PXτB(x,y−x)(Xτ

B(X0,|X0−y|) = y); XτB(x,y−x) ∈ (y, 5y/4 ∧ y + c2)

≥ e−c1−α2 pPx(XτB(x,y−x) ∈ (y, 5y/4 ∧ y + c2)) > e−c1−α2 pc3 > 0 , where p > 0 is defined in the proof of Proposition 12.

From Corollary 14 we deduce that for d = 1 > α the process Xt is point recurrent except the point 0. We may prove that the process Xt hits 0 in finite time with probability 0 by considering the λ-potentials.

(16)

Harnack inequality for d = 1 > α.

Now we consider the case d = 1 > α. In dimension 1 there is no such geometric construction like in the case d ≥ 2 and we may prove the Harnack inequality (1). We will adapt the proof of Theorem 7.1 (transient case) in [4].

Althougth our process is point recurrent we cannot adapt the method given in [4] for recurrent case because Px{y} < τB(x,r|x−y|)) → 0 for any r > 0, when y → x. First we will need some lemmas

Lemma 15. Let r < 1. There exists a constant C7 such that for all x, y ∈ B(x0, r/2) it holds

ExτB(x0,2r) ≤ C7EyτB(x0,2r) (17) Proof. By Lemma 4

ExτB(x0,2r) ≤ ExτB(x,5r/2) ≤ C2 5r

2|x| ∧ (5r/2)α



≤ c1 r

|x| ∧ rα



. (18) Since x, y ∈ B(x0, r/2) we have |y| ≤ |x| + r < |x| + r1−α ≤ 2(|x| ∨ r1−α).

Hence r

|y| ∧ rα ≥ r

2(|x| ∨ r1−α) ∧ rα ≥ r 2|x|∧ rα

2 = 1 2

 r

|x| ∧ rα

 . By (18) and Lemma 4 we obtain

ExτB(x0,2r) ≤ 2c1 r

|y| ∧ rα



≤ c2 3r

2|y| ∧ (3r/2)α



≤ c2C2EyτB(y,3r/2)

≤ c2C2EyτB(x0,2r)

For any open set D ⊂ Rd we define the function PD(x, y) by PD(x, y) =

Z

D

GD(x, z)

|y − z|d+αdz , x ∈ D, y ∈ Int(Dc) . If y ∈ D then we put PD(x, y) = 0.

Lemma 16. Let d = 1 > α. There exists a constant C8 such that for all r ∈ (0, 1), x ∈ B(x0, r/2) and y ∈ B(x0, r)c it holds

PB(x0,r)(x, y) ≤ C8ExB(x0,r))r−1δB(x0,r)(y)−α.

(17)

Proof. Without loss of generality we may suppose that x ≥ 0. Let y0 be the one of the points x0 + r, x0 − r lying on interval xy. Denote B = B(x0, r) and B = B ∩ B(y0, r/8). We note that

1 2C2

r

x ∧ rα

≤ ExB(x,r/2)) ≤ ExB) ≤ ExB(x,3r/2)) ≤ 3C2 2

r

x ∧ rα . For z ∈ B \ B we have |y − z| ≥ r/8 hence

Z

B\B

GB(x, z)

|z − y|1+α dz ≤ 1 δB(y)αr/8

Z

B

GB(x, z) dz = 8ExB)r−1δB(y)−α. So it suffices to show that GB(x, z) ≤ c1ExB)r−1 for z ∈ B. We estimate the integral R

0 pB(t, x, z) dt. Let c be some positive constant whose value will be determined later. In the estimates below we will often use the density estimates (4), (5). First we estimate

Z crα

pB(t, x, z) dz = Z

crα

Z

B

pB(t − crα, x, u)pB(crα, u, z) du dt

≤ Z

crα

Z

B

pB(t − crα, x, u)p(crα, u, z) du dt

≤ c2 Z

crα

PxB > t − crα) 1 − eαλcrα αλ

−1/α dt

≤ c3ExB)r−1.

where the constant c3 depends on c. For t ∈ (0, crα) we will estimate pB(t, x, z) ≤ p(t, x, z). We will consider two cases.

Let x < 2r1−α. For t < r/(8λx) it holds e−λtx ∈ B(x, r/4) (see i.e.

Lemma 11 in [11]), hence |xe−λt− z| > r/8. Since 2r/x > rα, for c = 1/(16λ) we have

Z crα 0

pB(t, x, z) dz ≤ Z crα

0

t

|xe−λt− z|1+α dt ≤ c4rr−1−α ≤ c5ExB)r−1. Now we suppose that x ≥ 2r1−α. We may and do assume that y < x.

For t > 3r/(λx) it holds e−λtx 6∈ B(x, 3κr/2) (see Lemma 11 in [11]), with a constant κ > 1. Hence e−λtx 6∈ B(x0, κr) and we have for c > 3/(2λ)

Z crα 3r/(λ|x|)

pB(t, x, z) dt ≤

Z crα 3r/(λ|x|)

c6 t

|xe−λt− x|1+α dt ≤ c7 x1+α

Z crα 0

t−αdt

≤ c8

rα(1−α) x1+α ≤ c8

r

2αxr−1 ≤ c9ExB)r−1.

(18)

It remains to estimate R3r/(λx)

0 p(t, x, z) dz. Let t0 = λ1ln(x/z) so e−λt0x = z.

We note that c10r/x ≤ 1λln(x/z) ≤ c11r/x hence t0 = axr, where c10≤ a ≤ c11

and c10, c11 do not depend on x, z, r. Let t1 = a21/α−1(r/x)1/αx−1. We note that t12xar = t20. We estimate

Z t0+t1

t0−t1

p(t, x, z) dz ≤ c12

Z t0+t1

t0−t1

t−1/αdt ≤ c13(r/x)−1/αt1 ≤ c14ExB)r−1. Next we estimate

Z t0−t1

0

p(t, x, z) dt ≤ c15

Z t0−t1

0

(t0− t1)dt x1+α(e−λt− e−λt0)1+α

≤ c16

r x2+α

Z t0−t1 0

dt (t0− t)1+α

≤ c16 r x2+α

Z t1

ds s1+α

= c16r

αx2+α(t1)−α ≤ c17ExB)r−1. In a similar way we estimateR3r/(λx)

t0+t1 p(t, x, z) dt. Therefore we obtained that for some constant c16 we have GB(x, z) ≤ c16ExB)r−1, where z ∈ B. Integrating R

B dz

|z−y|1+α we get the assertion of the lemma.

Denote by ωxD the distribution of the process starting from x and exit- ing the set D. From Ikeda-Watanabe formula ωxB on Int(Bc) has a density PB(x, y) = R

B

GB(x,z)

|y−z|d+αdz. Since Xt hits the boundary of Bt = B(x0, t) on exiting Bt the measure ωBxt has Lebesgue decomposition,

ωBxt(dy) = PBt(x, y)dy + ptδx0−t(dy) + qtδx0+t(dy) , (19) where pt = Px(Xτ

B(x0,t) = x0 − t) and qt = Px(Xτ

B(x0,t) = x0 + t). In next lemma we will show that if 0 ∈ B(x0, t) then pt= qt= 0.

Lemma 17. If 0 ∈ B(x0, r) then Px(XτB(x0,r) ∈ {x0+ r, x0− r}) = 0.

Proof. Since ωxD is absolutely continuous with respect to Lebesgue measure on Int(Dc) the process hits x0+ r on exiting B(x0, r) only if the event

0 = {Xτ((x0+r)/2,(x0+r+X0)/2) ∈ [(x0+ r + X0)/2, x0+ r) , X0 > 0}

(19)

happens infinitely many times. It means that the process Xt jumps out from every B(x0, t), t < r before it exits B(x0, r). By the same method as in he proof of Proposition 12 we show that Px(Xτ((x0+r)/2,(x0+r+X0)/2) ∈ (0, (x + r)/2)) > c1 > 0 for all x ∈ (0, x0 + r) and some constant c1 in- dependent of x. Hence Px(Ω0) < 1 − c1 for all x ∈ (0, x0 + r). Therefore P(Ω0 happens infinitely many times) = 0 and we get pr = 0. We use the same reasoning to obtain qr = 0.

Corollary 18. From Lemma 17 and Proposition 12 we deduce that for x0 >

t > 0 we have qt = 0 and pt > 0 in (19).

Using argument from [4] (proof of Theorem 7.1 in recurrent case) we prove (1) for |y| < |x|.

Lemma 19. Let d = 1 > α. There exists a constant C9 such that for all functions h nonnegative on Rdand L-harmonic in the ball B(x0, r) satisfying M > x0 > 4r it holds

h(x) ≥ C9h(y), x ∈ (x0− r, x0), y ∈ (x0− r, x) .

Proof. Denote δt(x) = dist(x, ∂B(x0, t)). At first let r < c1 = 2C51/(α−1). Since x0 > 4r then for x ∈ (x0− t, x0) and 0 < t ≤ r we have |x|/4 > δt(x) hence from (16) it holds

Px(Xτ

B(x0,t) = x0− t) > e−δt(x)1−αp > e−c1−α1 p > 0 , x ∈ (x0− t, x0) . Therefore for y ∈ (x0− r, x) we have

h(x) = Ex(h(XτB(x,|x−y|))) ≥ Ex(h(XτB(x,|x−y|)); XτB(x,|x−y|) = y)

≥ e−c1−α1 ph(y) .

Now suppose r > c1. Let n ∈ N be such c1n ≥ r > c1(n − 1). By the chain argument h(x) ≥ e−c1−α1 nph(y) for y ∈ (x0 − r, x). Since r < M we have n ≤ M/c1 + 1. Therefore we obtain assertion of the lemma with C9 = e−c−α1 (M +c1)p.

Now we may prove Theorem 2 for d = 1 > α.

(20)

proof of Theorem 2 for d = 1 > α. Denote B = B(x0, 2r). Let h be a non- negative function on Rd, L-harmonic in B(x0, 2r) and x, x ∈ B(x0, r/2). We assume that x0 > 0. Let

h1(z) = Ez(h(XτB); XτB ∈ B(x0, 3r)), h2(z) = Ez(h(XτB); XτB 6∈ B(x0, 3r)).

We have h = h1 + h2 and h1, h2 are L-harmonic in B. Hence by Ikeda- Watanabe formula and Lemma 15

h2(x) = Z

B(x0,3r)c

h2(y)PB(x, y) dy

≤ c1ExB) Z

B(x0,3r)c

h2(y)

|x − y|1+α dy

≤ c2Ex

B) Z

B(x0,3r)c

h2(y)

|x− y|1+α dy

≤ c3 Z

B(x0,3r)c

h2(y)PB(x, y) dy = c3h2(x) .

It is enough to prove an analogous inequality for h1. Let R = B(x0, 3r) \ B(x0, r). We note that supp h1 ⊂ B(x0, 3r) and for x ∈ B(x0, r/2) and y ∈ R we have |x − y| ≤ 4r. Denote Bt = B(x0, t). By (19) and Corollary 18 we have

ωxBt(dy) = PBt(x, y)dy + ptδx0−t(dy) . with pt = Px(Xτ

B(x0,t) = x0 − t). We note that if 0 ∈ B(x0, t) then pt = 0.

Hence we get the inequality h1(x) = Exh1(XτB(x0,3r/4))

= Z

B(x0,3r/4)c

PB(x0,3r/4)(x, y)h1(y) dy + p3r/4h1(x0− 3r/4)

≥ c4 Z

R

Z

B(x0,3r/4)

GB(x0,3r/4)(x, z)

|z − y|1+α dz h1(y)dy + p3r/4h1(x0− 3r/4)

≥ c5ExB(x0,r))r−1−α Z

R

h1(y) dy + p3r/4h1(x0− 3r/4) . (20) In the last line we use Lemma 4 to estimate ExB(x0,3r/4)) ≥ cExB(x0,r)).

Since ωBxt(A) = ωxBt(A ∩ Btc) for all measurable A we have for all t ∈ [r, 2r]

h1(x) = Z

B(x0,r)c

h1(y)ωBxt(dy) .

Cytaty

Powiązane dokumenty

Oleszkiewicz, Institute of Mathematics, University of Warsaw, Banacha 2, 02-097 Warszawa,

So, now you think you can solve every single problem involving systems of linear differential equations with constant coefficients, eh.. Not to let you down after all the work you

In the last ten years, much of the study has been focused upon finding conditions to ensure the existence of a solution in the case where T need not be upper semicontinuous, since

In this paper we present a result on convergence of approximate solutions of stochastic differential equations involving integrals with respect to α-stable L´evy motion.. We prove

While this result is a little weaker than the one due to Blanksby and Montgomery, the method used has since become quite important as it has produced the best results yet known,

Schoißengeier [14] expressed ν ∗ (α) in terms of the continued fraction expansion of α after he had obtained partial results in [13].. Em- ploying these

Equivalence of linear canonical transform domains to fractional Fourier domains and the bicanonical width product: A generalization of the space-bandwidth product, Journal of

When a diffusion is ergodic its transition density converges to its invariant density, see Durrett (1998).. This convergence enabled us to introduce a sample partitioning technique