Hindawi Publishing Corporation International Journal of Stochastic Analysis Volume 2014, Article ID 793275, 22 pages http://dx.doi.org/10.1155/2014/793275 Research Article SPDEs with πΌ-Stable Lévy Noise: A Random Field Approach Raluca M. Balan Department of Mathematics and Statistics, University of Ottawa, 585 King Edward Avenue, Ottawa, ON, Canada K1N 6N5 Correspondence should be addressed to Raluca M. Balan; rbalan@uottawa.ca Received 17 August 2013; Accepted 25 November 2013; Published 4 February 2014 Academic Editor: H. Srivastava Copyright © 2014 Raluca M. Balan. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. This paper is dedicated to the study of a nonlinear SPDE on a bounded domain in π π , with zero initial conditions and Dirichlet boundary, driven by an πΌ-stable LeΜvy noise π with πΌ ∈ (0, 2), πΌ =ΜΈ 1, and possibly nonsymmetric tails. To give a meaning to the concept of solution, we develop a theory of stochastic integration with respect to this noise. The idea is to first solve the equation with “truncated” noise (obtained by removing from π the jumps which exceed a fixed value πΎ), yielding a solution π’πΎ , and then show that the solutions π’πΏ , πΏ > πΎ coincide on the event π‘ ≤ ππΎ , for some stopping times ππΎ converging to infinity. A similar idea was used in the setting of Hilbert-space valued processes. A major step is to show that the stochastic integral with respect to ππΎ satisfies a πth moment inequality. This inequality plays the same role as the Burkholder-Davis-Gundy inequality in the theory of integration with respect to continuous martingales. 1. Introduction Modeling phenomena which evolve in time or space-time and are subject to random perturbations are a fundamental problem in stochastic analysis. When these perturbations are known to exhibit an extreme behavior, as seen frequently in finance or environmental studies, a model relying on the Gaussian distribution is not appropriate. A suitable alternative could be a model based on a heavy-tailed distribution, like the stable distribution. In such a model, these perturbations are allowed to have extreme values with a probability which is significantly higher than in a Gaussianbased model. In the present paper, we introduce precisely such a model, given rigorously by a stochastic partial differential equation (SPDE) driven by a noise term which has a stable distribution over any space-time region and has independent values over disjoint space-time regions (i.e., it is a LeΜvy noise). More precisely, we consider the SPDE: πΏπ’ (π‘, π₯) = π (π’ (π‘, π₯)) πΜ (π‘, π₯) , π‘ > 0, π₯ ∈ O (1) with zero initial conditions and Dirichlet boundary conditions, where π is a Lipschitz function, πΏ is a second-order pseudo-differential operator on a bounded domain O ⊂ Rπ , Μ π₯) = ππ+1 π/ππ‘ππ₯1 , . . . , ππ₯π is the formal derivative and π(π‘, of an πΌ-stable LeΜvy noise with πΌ ∈ (0, 2), πΌ =ΜΈ 1. The goal is to find sufficient conditions on the fundamental solution πΊ(π‘, π₯, π¦) of the equation πΏπ’ = 0 on R+ × O, which will ensure the existence of a mild solution of (1). We say that a predictable process π’ = {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O} is a mild solution of (1) if for any π‘ > 0, π₯ ∈ O, π‘ π’ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’ (π , π¦)) π (ππ , ππ¦) 0 O a.s. (2) We assume that πΊ(π‘, π₯, π¦) is a function in π‘, which excludes from our analysis the case of the wave equation with π ≥ 3. To explain the connections with other works, we describe briefly the construction of the noise (the details are given in Section 2). This construction is similar to that of a classical πΌ-stable LeΜvy process and is based on a Poisson random measure (PRM) π on R+ × Rπ × (R \ {0}) of intensity ππ‘ππ₯]πΌ (ππ§), where ]πΌ (ππ§) = [ππΌπ§−πΌ−1 1(0,∞) (π§) + ππΌ(−π§)−πΌ−1 1(−∞,0) (π§)] ππ§ (3) 2 International Journal of Stochastic Analysis for some π, π ≥ 0 with π + π = 1. More precisely, for any set π΅ ∈ Bπ (R+ × Rπ ), π (π΅) = ∫ π΅×{|π§|≤1} +∫ Μ (ππ , ππ₯, ππ§) π§π π΅×{|π§|>1} (4) π§π (ππ , ππ₯, ππ§) − π |π΅| , Μ where π(π΅×⋅) = π(π΅×⋅)−|π΅|]πΌ (⋅) is the compensated process and π is a constant (specified by Lemma 3). Here, Bπ (R+ × Rπ ) is the class of bounded Borel sets in R+ × Rπ and |π΅| is the Lebesgue measure of π΅. As the term on the right-hand side of (2) is a stochastic integral with respect to π, such an integral should be constructed first. Our construction of the integral is an extension to random fields of the construction provided by GineΜ and Marcus in [1] in the case of an πΌ-stable LeΜvy process {π(π‘)}π‘∈[0,1] . Unlike these authors, we do not assume that the measure ]πΌ is symmetric. Since any LeΜvy noise is related to a PRM, in a broad sense, one could say that this problem originates in ItoΜ’s papers [2, 3] regarding the stochastic integral with respect to a Poisson noise. SPDEs driven by a compensated PRM were considered for the first time in [4], using the approach based on Hilbert-space-valued solutions. This study was motivated by an application to neurophysiology leading to the cable equation. In the case of the heat equation, a similar problem was considered in [5–7] using the approach based on random-field solutions. One of the results of [6] shows that the heat equation: 1 ππ’ (π‘, π₯) = Δπ’ (π‘, π₯) ππ‘ 2 Μ (π‘, π₯, ππ§) + ∫ π (π‘, π₯, π’ (π‘, π₯) ; π§) π (5) π + π (π‘, π₯, π’ (π‘, π₯)) has a unique solution in the space of predictable processes π’ satisfying sup(π‘,π₯)∈[0,π]×Rπ πΈ|π’(π‘, π₯)|π < ∞, for any π ∈ (1 + Μ is the compensated process corre2/π, 2]. In this equation, π sponding to a PRM π on R+ × Rπ × π of intensity ππ‘ππ₯](ππ§), for an arbitrary π-finite measure space (π, B(π), ]) with measure ] satisfying ∫π |π§|π ](ππ§) < ∞. Because of this later condition, this result cannot be used in our case with π = R \ {0} and ] = ]πΌ . For similar reasons, the results of [7] also do not cover the case of an πΌ-stable noise. However, in the case πΌ > 1, we will be able to exploit successfully some ideas of [6] for treating the equation with “truncated” noise ππΎ , obtained by removing from π the jumps exceeding a value πΎ (see Section 5.2). The heat equation with the same type of noise as in the present paper was examined in [8, 9] in the cases πΌ < 1 and πΌ > 1, respectively, assuming that the noise has only positive jumps (i.e., π = 0). The methods used by these authors are different from those presented here, since they investigate the more difficult case of a non-Lipschitz function π(π’) = π’πΏ with πΏ > 0. In [8], Mueller removes the atoms of π of mass smaller than 2−π and solves the equation driven by the noise obtained in this way; here we remove the atoms of π of mass larger than πΎ and solve the resulting equation. In [9], Mytnik uses a martingale problem approach and gives the existence of a pair (π’, π) which satisfies the equation (the so-called “weak solution”), whereas in the present paper we obtain the existence of a solution π’ for a given noise π (the so-called “strong solution”). In particular, when πΌ > 1 and πΏ = 1/πΌ, the existence of a “weak solution” of the heat equation with πΌ-stable LeΜvy noise is obtained in [9] under the condition πΌ<1+ 2 π (6) which we encounter here as well. It is interesting to note that (6) is the necessary and sufficient condition for the existence of the density of the super-Brownian motion with “πΌ − 1”stable branching (see [10]). Reference [11] examines the heat equation with multiplicative noise (i.e., π(π’) = π’), driven by an πΌ-stable LeΜvy noise π which does not depend on time. To conclude the literature review, we should point out that there are many references related to stochastic differential equations with πΌ-stable LeΜvy noise, using the approach based on Hilbert-space valued solutions. We refer the reader to Section 12.5 of the monograph [12] and to [13–16] for a sample of relevant references. See also the survey article [17] for an approach based on the white noise theory for LeΜvy processes. This paper is organized as follows. (i) In Section 2, we review the construction of the πΌstable LeΜvy noise π, and we show that this can be viewed as an independently scattered random measure with jointly πΌ-stable distributions. (ii) In Section 3, we consider the linear equation (1) (with π(π’) = 1) and we identify the necessary and sufficient condition for the existence of the solution. This condition is verified in the case of some examples. (iii) Section 4 contains the construction of the stochastic integral with respect to the πΌ-stable noise π, for πΌ ∈ (0, 2). The main effort is dedicated to proving a maximal inequality for the tail of the integral process, when the integrand is a simple process. This extends the construction of [1] to the case random fields and nonsymmetric measure ]πΌ . (iv) In Section 5, we introduce the process ππΎ obtained by removing from π the jumps exceeding a fixed value πΎ, and we develop a theory of integration with respect to this process. For this, we need to treat separately the cases πΌ < 1 and πΌ > 1. In both cases, we obtain a πth moment inequality for the integral process for π ∈ (πΌ, 1) if πΌ < 1 and π ∈ (πΌ, 2) if πΌ > 1. This inequality plays the same role as the Burkholder-Davis-Gundy inequality in the theory of integration with respect to continuous martingales. (v) In Section 6 we prove the main result about the existence of the mild solution of (1). For this, we first solve the equation with “truncated” noise ππΎ using a Picard iteration scheme, yielding a solution π’πΎ . International Journal of Stochastic Analysis 3 We then introduce a sequence (ππΎ )πΎ≥1 of stopping times with ππΎ ↑ ∞ a.s. and we show that the solutions π’πΏ , πΏ > πΎ coincide on the event π‘ ≤ ππΎ . For the definition of the stopping times ππΎ , we need again to consider separately the cases πΌ < 1 and πΌ > 1. (vi) Appendix A contains some results about the tail of a nonsymmetric stable random variable and the tail of an infinite sum of random variables. Appendix B gives an estimate for the Green function associated with the fractional power of the Laplacian. Appendix C gives a local property of the stochastic integral with respect to π (or ππΎ ). 2. Definition of the Noise In this section we review the construction of the πΌ-stable LeΜvy noise on R+ × Rπ and investigate some of its properties. Let π = ∑π≥1 πΏ(ππ ,ππ ,ππ ) be a Poisson random measure on R+ × Rπ × (R \ {0}), defined on a probability space (Ω, F, π), with intensity measure ππ‘ππ₯]πΌ (ππ§), where ]πΌ is given by (3). Let (ππ )π≥0 be a sequence of positive real numbers such that ππ → 0 as π → ∞ and 1 = π0 > π1 > π2 > ⋅ ⋅ ⋅ . Let Γπ = {π§ ∈ R; ππ < |π§| ≤ ππ−1 } , π ≥ 1, (7) Γ0 = {π§ ∈ R; |π§| > 1} . π΅×Γπ = πΈ (πππ’π(π΅) ) = exp {|π΅| ∫ (πππ’π§ − 1 − ππ’π§1{|π§|≤1} ) ]πΌ (ππ§)} , π’ ∈ R. Hence, πΈ(π(π΅)) = |π΅| ∫R π§1{|π§|>1} ]πΌ (ππ§) and Var(π(π΅)) = |π΅| ∫R π§2 ]πΌ (ππ§). Lemma 2. The family {π(π΅); π΅ ∈ Bπ (R+ × Rπ )} defined by (10) is an independently scattered random measure; that is, (a) for any disjoint sets π΅1 , . . . , π΅π in Bπ (R+ × Rπ ), π(π΅1 ), . . . , π(π΅π ) are independent; (b) for any sequence (π΅π )π≥1 of disjoint sets in Bπ (R+ × Rπ ) such that βπ≥1 π΅π is bounded, π(βπ≥1 π΅π ) = ∑π≥1 π(π΅π ) a.s. Proof. (a) Note that for any function π ∈ πΏ2 (R+ × Rπ ) with compact support πΎ, we can define the random variable π(π) = ∑π≥1 [πΏ π (π) − πΈ(πΏ π (π))] + πΏ 0 (π) where πΏ π (π) = ∫πΎ×Γ π(π‘, π₯)π§ π(ππ‘, ππ₯, ππ§). For any π’ ∈ R, we have π = exp {∫ R+ ×Rπ ×R π§π (ππ‘, ππ₯, ππ§) ∑ ππ 1{ππ ∈Γπ } , (ππ ,ππ )∈π΅ π ≥ 0. For any π ≥ 0, the variable πΏ π (π΅) has a compound Poisson distribution with jump intensity measure |π΅| ⋅ ]πΌ |Γπ ; that is, πΈ [πππ’πΏ π (π΅) ] = exp {|π΅| ∫ (πππ’π§ − 1) ]πΌ (ππ§)} , (πππ’π§π(π‘,π₯) − 1 −ππ’π§π (π‘, π₯) 1{|π§|≤1} ) ππ‘ ππ₯ ]πΌ (ππ§) } . (8) Remark 1. The variable πΏ 0 (π΅) is finite since the sum above contains finitely many terms. To see this, we note that πΈ[π(π΅ × Γ0 )] = |π΅|]πΌ (Γ0 ) < ∞, and hence π(π΅ × Γ0 ) = card{π ≥ 1; (ππ , ππ , ππ ) ∈ π΅ × Γ0 } < ∞. Γπ π’ ∈ R. (12) For any disjoint sets π΅1 , . . . , π΅π and for any π’1 , . . . , π’π ∈ R, we have π πΈ [exp (π ∑ π’π π (π΅π ))] π=1 π = πΈ [exp (ππ ( ∑ π’π 1π΅π ))] π=1 = exp {∫ π R+ ×Rπ ×R (πππ§ ∑π=1 π’π 1π΅π (π‘,π₯) − 1 − ππ§1{|π§|≤1} (9) π × ∑ π’π 1π΅π (π‘, π₯)) ππ‘ ππ₯ ]πΌ (ππ§)} It follows that πΈ(πΏ π (π΅)) = |π΅| ∫Γ π§]πΌ (ππ§) and Var(πΏ π (π΅)) = π |π΅| ∫Γ π§2 ]πΌ (ππ§) for any π ≥ 0. Hence, Var(πΏ π (π΅)) < ∞ for any π π ≥ 1 and Var(πΏ 0 (π΅)) = ∞. If πΌ > 1, then πΈ(πΏ 0 (π΅)) is finite. Define π (π΅) = ∑ [πΏ π (π΅) − πΈ (πΏ π (π΅))] + πΏ 0 (π΅) . π≥1 (11) R πΈ (πππ’π(π) ) For any set π΅ ∈ Bπ (R+ × Rπ ), we define πΏ π (π΅) = ∫ From (9) and (10), it follows that π(π΅) is an infinitely divisible random variable with characteristic function: π=1 π σ΅¨ σ΅¨ = exp { ∑ σ΅¨σ΅¨σ΅¨π΅π σ΅¨σ΅¨σ΅¨ ∫ (πππ’π π§ − 1 π=1 −ππ’π π§1{|π§|≤1} ) ]πΌ (ππ§) } (10) This sum converges a.s. by Kolmogorov’s criterion since {πΏ π (π΅) − πΈ(πΏ π (π΅))}π≥1 are independent zero-mean random variables with ∑π≥1 Var(πΏ π (π΅)) < ∞. R π = ∏πΈ [exp (ππ’π π (π΅π ))] , π=1 (13) 4 International Journal of Stochastic Analysis using (12) with π = ∑ππ=1 π’π 1π΅π for the second equality and (9) for the last equality. This proves that π(π΅1 ), . . . , π(π΅π ) are independent. (b) Let ππ = ∑ππ=1 π(π΅π ) and π = π(π΅), where π΅ = βπ≥1 π΅π . By LeΜvy’s equivalence theorem, (ππ )π≥1 converges a.s. if and only if it converges in distribution. By (13), with π’π = π’ for all π = 1, . . . , π, we have σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ ππ’ππ πΈ (π ) = exp {σ΅¨σ΅¨σ΅¨ β π΅π σ΅¨σ΅¨σ΅¨ ∫ (πππ’π§ − 1 − ππ’π§1{|π§|≤1} ) ]πΌ (ππ§)} . σ΅¨σ΅¨π=1 σ΅¨σ΅¨ R σ΅¨ σ΅¨ (14) This clearly converges to πΈ(πππ’π ) = exp{|π΅| ∫R (πππ’π§ − 1 − ππ’π§1{|π§|≤1} )]πΌ (ππ§)}, and hence (ππ )π≥1 converges in distribution to π. Recall that a random variable π has an πΌ-stable distribution with parameters πΌ ∈ (0, 2), π ∈ [0, ∞), π½ ∈ [−1, 1], and π ∈ R if, for any π’ ∈ R, ππΌ ) + ππ’π} , πΈ (πππ’π ) = exp {−|π’|πΌ ππΌ (1 − π sgn (π’) π½ tan 2 if πΌ =ΜΈ 1, (15) From Lemmas 2 and 3, it follows that π = {π (π΅) = π (π΅) − π |π΅| ; π΅ ∈ Bπ (R+ × Rπ )} (19) is an πΌ-stable random measure, in the sense of Definition 3.3.1 of [18], with control measure π(π΅) = ππΌ |π΅| and constant skewness intensity π½. In particular, π(π΅) has a ππΌ (π|π΅|1/πΌ , π½, 0) distribution. We say that π is an πΌ-stable LeΜvy noise. Coming back to the original construction (10) of π(π΅) and noticing that π |π΅| = − |π΅| ∫ π§1{|π§|≤1} ]πΌ (ππ§) = −∑ πΈ (πΏ π (π΅)) , R π≥1 if πΌ < 1, (20) π |π΅| = |π΅| ∫ π§1{|π§|>1} ]πΌ (ππ§) = πΈ (πΏ 0 (π΅)) , R if πΌ > 1, it follows that π(π΅) can be represented as π (π΅) = ∑ πΏ π (π΅) =: ∫ π΅×(R\{0}) π≥0 π§π (ππ‘, ππ₯, ππ§) , if πΌ < 1, (21) or 2 πΈ (πππ’π ) = exp {− |π’| π (1 + π sgn (π’) π½ ln |π’|) + ππ’π} , π if πΌ = 1 (16) (see Definition 1.1.6 of [18]). We denote this distribution by ππΌ (π, π½, π). Lemma 3. π(π΅) has a ππΌ (π|π΅|1/πΌ , π½, π|π΅|) distribution with π½ = π − π, πΌ π =∫ ∞ 0 ππΌ Γ (2 − πΌ) { { { 1 − πΌ cos 2 , sin π₯ ππ₯ = { { π₯πΌ {π, {2 πΌ {π½ , ππ πΌ =ΜΈ 1, π={ πΌ−1 ππ πΌ = 1, π½π , { 0 ππ πΌ =ΜΈ 1, ππ πΌ = 1, Proof. We first express the characteristic function (11) of π(π΅) in Feller’s canonical form (see Section XVII.2 of [19]): =: ∫ π΅×(R\{0}) Μ (ππ‘, ππ₯, ππ§) , π§π if πΌ > 1. Μ is the compensated Poisson measure associated with Here π Μ π; that is, π(π΄) = π(π΄)−πΈ(π(π΄)) for any relatively compact set π΄ in R+ × Rπ × (R \ {0}). In the case πΌ = 1, we will assume that π = π so that ]πΌ is symmetric around 0, πΈ(πΏ π (π΅)) = 0 for all π ≥ 1, and π(π΅) admits the same representation as in the case πΌ < 1. As a preliminary investigation, we consider first equation (1) with π = 1: (18) with ππΌ (ππ§) = π§2 ]πΌ (ππ§) and π = ∫R (sin π§ − π§1{|π§|≤1} )]πΌ (ππ§). Then the result follows from the calculations done in Example XVII.3.(g) of [19]. π‘ > 0, π₯ ∈ O (23) with zero initial conditions and Dirichlet boundary conditions. In this section O is a bounded domain in Rπ or O = Rπ . By definition, the process {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O} given by π‘ π’ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (ππ , ππ¦) 0 ) πππ’π§ − 1 − ππ’ sin π§ = exp {ππ’π |π΅| + |π΅| ∫ ππΌ (ππ§)} π§2 R (22) πΏπ’ (π‘, π₯) = πΜ (π‘, π₯) , and π0 = ∫0 (sin π§ − π§1{π§≤1} )π§−2 ππ§. If πΌ > 1, then πΈ(π(π΅)) = π|π΅|. πΈ (π π≥0 3. The Linear Equation (17) ∞ ππ’π(π΅) π (π΅) = ∑ [πΏ π (π΅) − πΈ (πΏ π (π΅))] O (24) is a mild solution of (23), provided that the stochastic integral on the right-hand side of (24) is well defined. We define now the stochastic integral of a deterministic function π: ∞ π (π) = ∫ ∫ π (π‘, π₯) π (ππ‘, ππ₯) . 0 Rπ (25) International Journal of Stochastic Analysis 5 If π ∈ πΏπΌ (R+ × Rπ ), this can be defined by approximation with simple functions, as explained in Section 3.4 of [18]. The process {π(π); π ∈ πΏπΌ (R+ × Rπ )} has jointly πΌ-stable finite dimensional distributions. In particular, each π(π) has a ππΌ (ππ , π½, 0)-distribution with scale parameter: ∞ σ΅¨πΌ σ΅¨ ππ = π(∫ ∫ σ΅¨σ΅¨σ΅¨π (π‘, π₯)σ΅¨σ΅¨σ΅¨ ππ₯ ππ‘) 0 Rπ 1/πΌ . Condition (29) can be easily verified in the case of several examples. Example 6 (heat equation). Let πΏ = π/ππ‘ − (1/2)Δ. Assume first that O = Rπ . Then πΊ(π‘, π₯, π¦) = πΊ(π‘, π₯ − π¦), where 1 πΊ (π‘, π₯) = (26) (2ππ‘) π/2 exp (− |π₯|2 ), 2π‘ (30) More generally, a measurable function π : R+ × Rπ → R is integrable with respect to π if there exists a sequence (ππ )π≥1 of simple functions such that ππ → π a.e., and, for any π΅ ∈ Bπ (R+ × Rπ ), the sequence {π(ππ 1π΅ )}π converges in probability (see [20]). The next results show that condition π ∈ πΏπΌ (R+ × Rπ ) is also necessary for the integrability of π with respect to π. Due to Lemma 2, this follows immediately from the general theory of stochastic integration with respect to independently scattered random measures developed in [20]. and condition (29) is equivalent to (6). In this case, πΌπΌ (π‘) = ππΌ,π π‘π(1−πΌ)/2+1 . If O is a bounded domain in Rπ , then πΊ(π‘, π₯, π¦) ≤ πΊ(π‘, π₯ − π¦) (see page 74 of [11]) and condition (29) is implied by (6). Lemma 4. A deterministic function π is integrable with respect to π if and only if π ∈ πΏπΌ (R+ × Rπ ). is the generator of a Markov process with values in Rπ , without jumps (a diffusion). Assume that O is a bounded domain in Rπ or O = Rπ . By Aronson estimate (see, e.g., Theorem 2.6 of [12]), under some assumptions on the coefficients πππ , ππ , there exist some constants π1 , π2 > 0 such that Proof. We write the characteristic function of π(π΅) in the form used in [20]: πΈ (πππ’π(π΅) ) Example 7 (parabolic equation). Let πΏ = π/ππ‘ − L where π π,π=1 σ΅¨2 σ΅¨σ΅¨ σ΅¨π₯ − π¦σ΅¨σ΅¨σ΅¨ ) πΊ (π‘, π₯, π¦) ≤ π1 π‘−π/2 exp (− σ΅¨ π2 π‘ = exp {∫ [ππ’π π΅ + ∫ (πππ’π§ − 1 − ππ’π (π§)) ]πΌ (ππ§)] ππ‘ππ₯} R (27) with π = π½ − π, π(π§) = π§ if |π§| ≤ 1 and π(π§) = sgn(π§) if |π§| > 1. By Theorem 2.7 of [20], π is integrable with respect to π if and only if ∫ R+ ×Rπ π π2 π ππ (π₯) + ∑ππ (π₯) (π₯) ππ₯π ππ₯π ππ₯ π π=1 Lπ (π₯) = ∑ πππ (π₯) σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨π (π (π‘, π₯))σ΅¨σ΅¨σ΅¨ ππ‘ ππ₯ < ∞, for all π‘ > 0 and π₯, π¦ ∈ O. In this case, condition (29) is implied by (6). Example 8 (heat equation with fractional power of the Laplacian). Let πΏ = π/ππ‘ + (−Δ)πΎ for some πΎ > 0. Assume that O is a bounded domain in Rπ or O = Rπ . Then (see, e.g., Appendix B.5 of [12]) πΊ (π‘, π₯, π¦) = ∫ G (π , π₯, π¦) ππ‘,πΎ (π ) ππ 0 (33) ∞ π (π (π‘, π₯)) ππ‘ ππ₯ < ∞, R+ ×Rπ 1/πΎ = ∫ G (π‘ 0 where π(π¦) = ππ¦ + ∫R (π(π¦π§) − π¦π(π§))]πΌ (ππ§) and π(π¦) = ∫R (1∧|π¦π§|2 )]πΌ (ππ§). Direct calculations show that, in our case, π(π¦) = −(π½/(πΌ − 1))π¦πΌ if πΌ =ΜΈ 1, π(π¦) = 0 if πΌ = 1, and π(π¦) = (2/(2 − πΌ))π¦πΌ . The following result follows immediately from (24) and Lemma 4. Proposition 5. Equation (23) has a mild solution if and only if for any π‘ > 0, π₯ ∈ O π‘ πΌ πΌπΌ (π‘) = ∫ ∫ πΊ(π , π₯, π¦) ππ¦ ππ < ∞. 0 O (32) ∞ (28) ∫ (31) (29) In this case, {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O} has jointly πΌ-stable finite-dimensional distributions. In particular, π’(π‘, π₯) has a ππΌ (ππΌπΌ (π‘)1/πΌ , π½, 0) distribution. π , π₯, π¦) π1,πΎ (π ) ππ , where G(π‘, π₯, π¦) is the fundamental solution of ππ’/ππ‘ − Δπ’ = 0 on O and ππ‘,πΎ is the density of the measure ππ‘,πΎ , (ππ‘,πΎ )π‘≥0 being a convolution semigroup of measures on [0, ∞) whose Laplace transform is given by ∞ ∫ π−π’π ππ‘,πΎ (π ) ππ = exp (−π‘π’πΎ ) , 0 ∀π’ > 0. (34) Note that if πΎ < 1, ππ‘,πΎ is the density of ππ‘ , where (ππ‘ )π‘≥0 is a πΎ-stable subordinator with LeΜvy measure ππΎ (ππ₯) = (πΎ/Γ(1 − πΎ))π₯−πΎ−1 1(0,∞) (π₯)ππ₯. Assume first that O = Rπ . Then πΊ(π‘, π₯, π¦) = πΊ(π‘, π₯ − π¦), where 2πΎ πΊ (π‘, π₯) = ∫ πππ⋅π₯ π−π‘|π| ππ. Rπ (35) 6 International Journal of Stochastic Analysis If πΎ < 1, then πΊ(π‘, ⋅) is the density of ππ‘ , with (ππ‘ )π‘≥0 being a symmetric (2πΎ)-stable LeΜvy process with values in Rπ defined by ππ‘ = πππ‘ , with (ππ‘ )π‘≥0 a Brownian motion in Rπ with variance 2. By Lemma B.1 (Appendix B), if πΌ > 1, then (29) holds if and only if 2πΎ . π πΌ<1+ (36) If O is a bounded domain in Rπ , then πΊ(π‘, π₯, π¦) ≤ πΊ(π‘, π₯ − π¦) (by Lemma 2.1 of [8]). In this case, if πΌ > 1, then (29) is implied by (36). Example 9 (cable equation in R). Let πΏπ’ = ππ’/ππ‘−π2 π’/ππ₯2 +π’ and O = R. Then πΊ(π‘, π₯, π¦) = πΊ(π‘, π₯ − π¦), where πΊ (π‘, π₯) = 2 1 |π₯| exp (− − π‘) , √4ππ‘ 4π‘ (37) π if π = 2. (38) βπβπΌπΌ 4. Stochastic Integration In this section we construct a stochastic integral with respect to π by generalizing the ideas of [1] to the case of random fields. Unlike these authors, we do not assume that π(π΅) has a symmetric distribution, unless πΌ = 1. Let Fπ‘ = Fπ π‘ ∨ N where N is the π-field of negligible sets in (Ω, F, π) and Fπ π‘ is the π-field generated by π([0, π ]× π΄ × Γ) for all π ∈ [0, π‘], π΄ ∈ Bπ (Rπ ) and for all Borel sets Γ ⊂ π π R \ {0} bounded away from 0. Note that Fπ π‘ ⊂ Fπ‘ where Fπ‘ is the π-field generated by π([0, π ] × π΄), π ∈ [0, π‘], and π΄ ∈ Bπ (Rπ ). A process π = {π(π‘, π₯)}π‘≥0,π₯∈Rπ is called elementary if it is of the form π (π‘, π₯) = 1(π,π] (π‘) 1π΄ (π₯) π, (39) π where 0 ≤ π < π, π΄ ∈ Bπ (R ), and π is Fπ -measurable and bounded. A simple process is a linear combination of elementary processes. Note that any simple process π can be written as (41) =∑ π≥1 1 ∧ βπβπΌ,π,πΈπ 2π 1 ∧ βπβπΌπΌ,π,πΈπ 2π , if πΌ > 1, (42) , if πΌ ≤ 1. We identify two processes π and π for which βπ − πβπΌ = 0; that is, π = π] a.e., where ] = πππ‘ππ₯. In particular, we identify two processes π and π if π is a modification of π; that is, π(π‘, π₯) = π(π‘, π₯) a.s. for all (π‘, π₯) ∈ R+ × Rπ . The space LπΌ becomes a metric space endowed with the metric ππΌ : ππΌ (π, π) = βπ − πβπΌ , if πΌ > 1, ππΌ (π, π) = βπ − πβπΌπΌ , if πΌ ≤ 1. (43) This follows using Minkowski’s inequality if πΌ > 1 and the inequality |π + π|πΌ ≤ |π|πΌ + |π|πΌ if πΌ ≤ 1. The following result can be proved similarly to Proposition 2.3 of [21]. Proposition 12. For any π ∈ LπΌ there exists a sequence (ππ )π≥1 of bounded simple processes such that βππ − πβπΌ → 0 as π → ∞. By Proposition 5.7 of [22], the πΌ-stable LeΜvy process {π(π‘, π΅) = π([0, π‘] × π΅); π‘ ≥ 0} has a caΜdlaΜg modification, for any π΅ ∈ Bπ (Rπ ). We work with these modifications. If π is a simple process given by (40), we define π−1 ππ πΌ (π) (π‘, π΅) = ∑ ∑πππ π ((π‘π ∧ π‘, π‘π+1 ∧ π‘] × (π΄ ππ ∩ π΅)) . π−1 π (π‘, π₯) = 1{0} (π‘) π0 (π₯) + ∑ 1(π‘π ,π‘π+1 ] (π‘) ππ (π₯) π΅ for all π > 0 and π΅ ∈ Bπ (Rπ ). Note that LπΌ is a linear space. Let (πΈπ )π≥1 be an increasing sequence of sets in Bπ (Rπ ) such that βπ πΈπ = Rπ . We define π≥1 Condition (29) holds for any πΌ ∈ (0, 2). In this case, πΌπΌ (π‘) = 2−πΌ π‘2 if π = 1 and πΌπΌ (π‘) = ((2π)1−πΌ /(2−πΌ)(3−πΌ))π‘3−πΌ if π = 2. π=0 Let LπΌ be the class of all predictable processes π such that βπβπΌ = ∑ if π = 1, 1 1 1{|π₯|<π‘} , ⋅ 2π √ 2 π‘ − |π₯|2 Remark 11. One can show that the predictable π-field P is the π-field generated by the class C of processes π such that π‘ σ³¨→ π(π, π‘, π₯) is left continuous for any π ∈ Ω, π₯ ∈ Rπ and (π, π₯) σ³¨→ π(π, π‘, π₯) is Fπ‘ × B(Rπ )-measurable for any π‘ > 0. 0 Example 10 (wave equation in Rπ with π = 1, 2). Let πΏ = π2 /ππ‘2 − Δ and O = Rπ with π = 1 or π = 2. Then πΊ(π‘, π₯, π¦) = πΊ(π‘, π₯ − π¦), where πΊ (π‘, π₯) = disjoint sets in Bπ (Rπ ). Without loss of generality, we assume that π0 = 0. We denote by P the predictable π-field on Ω × R+ × Rπ , that is, the π-field generated by all simple processes. We say that a process π = {π(π‘, π₯)}π‘≥0,π₯∈Rπ is predictable if the map (π, π‘, π₯) σ³¨→ π(π, π‘, π₯) is P-measurable. βπβπΌπΌ,π,π΅ := πΈ ∫ ∫ |π (π‘, π₯)|πΌ ππ₯ ππ‘ < ∞, and condition (29) holds for any πΌ ∈ (0, 2). 1 πΊ (π‘, π₯) = 1{|π₯|<π‘} , 2 π π 1π΄ ππ (π₯)πππ , with 0 = π‘0 < π‘1 < ⋅ ⋅ ⋅ < π‘π < ∞ and ππ (π₯) = ∑π=1 where (πππ )π=1,...,ππ are Fπ‘π -measurable and (π΄ ππ )π=1,...,ππ are (40) π=0 π=1 (44) International Journal of Stochastic Analysis 7 Note that, for any π΅ ∈ Bπ (Rπ ), πΌ(π) (π‘, π΅) is Fπ‘ -measurable for any π‘ ≥ 0, and {πΌ(π)(π‘, π΅)}π‘≥0 is caΜdlaΜg. We write Consequently, for any π = 1, . . . , π π−1 πΌ (π) (π π , π΅) = ∑ (πΌ (π) (π π+1 , π΅) − πΌ (π) (π π , π΅)) π‘ πΌ (π) (π‘, π΅) = ∫ ∫ π (π , π₯) π (ππ , ππ₯) . 0 π΅ (45) π=0 (51) π−1 ππ = ∑ ∑π½ππ πππ . The following result will be used for the construction of the integral. This result generalizes Lemma 3.3 of [1] to the case of random fields and nonsymmetric measures ]πΌ . Theorem 13. If π is a bounded simple process then π=0 π=1 0, Using (47) and (51), it is enough to prove that for any π > σ΅¨σ΅¨ π ππ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π ( max σ΅¨σ΅¨σ΅¨σ΅¨∑ ∑π½ππ πππ σ΅¨σ΅¨σ΅¨σ΅¨ > π) π=0,...,π−1 σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨π=0 π=1 σ΅¨ sup ππΌ π ( sup |πΌ (π) (π‘, π΅)| > π) π‘∈[0,π] π>0 (46) π πΌ ≤ ππΌ πΈ ∫ ∫ |π (π‘, π₯)| ππ₯ ππ‘, 0 ≤ ππΌ π−πΌ πΈ ∫ ∫ |π(π , π₯)|πΌ ππ₯ ππ . π΅ 0 for any π > 0 and π΅ ∈ Bπ (Rπ ), where ππΌ is a constant depending only on πΌ. π ( sup |πΌ (π) (π‘, π΅)| > π) (47) = lim π (max |πΌ (π) (π‘, π΅)| > π) . π→∞ π‘∈πΉπ Fix π ≥ 1. Denote by 0 = π 0 < π 1 < ⋅ ⋅ ⋅ < π π = π the points of the set πΉπ . Say π‘π = π ππ for some 0 = π0 < π1 < ⋅ ⋅ ⋅ < ππ. Then each interval (π‘π , π‘π+1 ] can be written as the union of some intervals of the form (π π , π π+1 ]: (π‘π , π‘π+1 ] = β (π π , π π+1 ] , (48) π∈πΌπ where πΌπ = {π; ππ ≤ π < ππ+1 }. By (44), for any π = 0, . . . , π − 1 and π ∈ πΌπ , πΌ (π) (π π+1 , π΅) − πΌ (π) (π π , π΅) ππ = ∑πππ π ((π π , π π+1 ] × (π΄ ππ ∩ π΅)) . (49) π πΈ ∫ ∫ |π (π , π₯)|πΌ ππ₯ ππ 0 π΅ π−1 ππ π=0 π=1 σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ = ∑ (π π+1 − π π ) ∑ πΈσ΅¨σ΅¨σ΅¨σ΅¨π½ππ σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ . π π π½ππ πππ . For the event We now prove (52). Let ππ = ∑π=1 on the left-hand side, we consider its intersection with the event {max0≤π≤π−1 |ππ | > π} and its complement. Hence, the probability of this event can be bounded by π−1 σ΅¨ σ΅¨ ∑ π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π) π=0 σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ + π ( max σ΅¨σ΅¨σ΅¨∑ππ 1{|ππ |≤π} σ΅¨σ΅¨σ΅¨ > π) =: πΌ + πΌπΌ. σ΅¨σ΅¨ 0≤π≤π−1 σ΅¨σ΅¨ σ΅¨π=0 σ΅¨ ππ πΌ (π) (π π+1 , π΅) − πΌ (π) (π π , π΅) = ∑π½ππ πππ , (54) We treat separately the two terms. For the first term, we note that π½π = (π½ππ )1≤π≤ππ is Fπ π measurable and ππ = (πππ )1≤π≤ππ is independent of Fπ π . By Fubini’s theorem πΌ= ∑∫ For any π ∈ πΌπ , let ππ = ππ , and, for any π = 1, . . . , ππ , define π½ππ = πππ , π»ππ = π΄ ππ , and πππ = π((π π , π π+1 ] × (π»ππ ∩ π΅)). With this notation, we have (53) This follows from the definition (40) of π and (48), since ππ π(π‘, π₯) = ∑π−1 π=0 ∑π∈πΌπ 1(π π ,π π+1 ] (π‘) ∑π=1 π½ππ 1π»ππ (π₯). π−1 π=1 π΅ First, note that Proof. Suppose that π is of the form (40). Since {πΌ(π) (π‘, π΅)}π‘∈[0,π] is caΜdlaΜg, it is separable. Without loss of generality, we assume that its separating set π· can be written as π· = ∪π πΉπ where (πΉπ )π is an increasing sequence of finite sets containing the points (π‘π )π=0,...,π. Hence, π‘∈[0,π] (52) π π=0 Rππ σ΅¨σ΅¨ σ΅¨σ΅¨ ππ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨∑ π₯π πππ σ΅¨σ΅¨σ΅¨σ΅¨ > π) ππ½ (ππ₯) , π σ΅¨σ΅¨ σ΅¨σ΅¨π=1 σ΅¨ σ΅¨ (55) where π₯ = (π₯π )1≤π≤ππ and ππ½ is the law of π½π . π π π We examine the tail of ππ = ∑π=1 π₯π πππ for a fixed ∀π = 0, . . . , π. π=1 (50) π₯ ∈ Rππ . By Lemma 3, πππ has a ππΌ (π(π π+1 − π π )1/πΌ |π»ππ ∩ π΅|1/πΌ , π½, 0) distribution. Since the sets (π»ππ )1≤π≤ππ are disjoint, the variables (πππ )1≤π≤ππ are independent. Using elementary 8 International Journal of Stochastic Analysis properties of the stable distribution (Properties 1.2.1 and 1.2.3 of [18]), it follows that ππ has a ππΌ (ππ , π½π∗ , 0) distribution with parameters: π Hence, 1 1−πΌ σ΅¨ σ΅¨ π (π π+1 − π π ) πΈ [σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1|ππ |≤π ] ≤ ππΌ∗ ππΌ 1−πΌ π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ πππΌ = ππΌ (π π+1 − π π ) ∑σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ , ππ σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ × ∑ πΈσ΅¨σ΅¨σ΅¨σ΅¨π½ππ σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ . π=1 π½π∗ = ππ ∑π=1 π=1 (56) ππ π½ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ σ΅¨σ΅¨ σ΅¨σ΅¨ ∑ sgn (π₯π ) σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨ . σ΅¨σ΅¨π₯π σ΅¨σ΅¨ σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨ π=1 σ΅¨ σ΅¨ σ΅¨ σ΅¨ From (59), (60), and (63), it follows that By Lemma A.1 (Appendix A), there exists a constant ππΌ∗ > 0 such that ππ σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ∗ π−πΌ ππΌ (π π+1 − π π ) ∑σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ (57) π=1 for any π > 0. Hence, πΌπΌ ≤ ππΌ∗ ππΌ π 1 −πΌ π πΈ ∫ ∫ |π (π , π₯)|πΌ ππ₯ ππ . 1−πΌ 0 π΅ Case 2 (πΌ > 1). We have σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π π πΌπΌ ≤ π ( max σ΅¨σ΅¨σ΅¨∑ππ σ΅¨σ΅¨σ΅¨ > ) + π ( max ππ > ) σ΅¨ 0≤π≤π−1 σ΅¨σ΅¨ 0≤π≤π−1 2 σ΅¨π=0 σ΅¨σ΅¨ 2 π π=0 π=1 (58) π = ππΌ∗ π−πΌ ππΌ πΈ ∫ ∫ |π (π , π₯)|πΌ ππ₯ ππ . 0 π΅ We now treat πΌπΌ. We consider three cases. For the first two cases we deviate from the original argument of [1] since we do not require that π½ = 0. πΌπΌ ≤ π ( max ππ > π) , 0≤π≤π−1 (59) where {ππ = ∑ππ=0 |ππ |1{|ππ |≤π} , Fπ π+1 ; 0 ≤ π ≤ π − 1} is a submartingale. By the submartingale maximal inequality (Theorem 35.3 of [23]), π ( max ππ > π) ≤ 4 π−1 σ΅¨ σ΅¨ π πΌπΌσΈ = π ( max σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > ) ≤ 2 ∑ πΈ (ππ2 ) 0≤π≤π−1 2 π π=0 4 π−1 ≤ 2 ∑ πΈ [ππ2 1{|ππ |≤π} ] . π π=0 π πΈ [ππ2 1{|ππ |≤π} ] ≤ 2ππΌ∗ ππΌ 1 2−πΌ π (π π+1 − π π ) 2−πΌ π π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ × ∑ σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ . As in Case 1, we obtain that πΈ [ππ2 1{|ππ |≤π} ] ≤ ππΌ∗ ππΌ σ΅¨ σ΅¨ πΈ [σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1|ππ |≤π ] 2 2−πΌ π (π π+1 − π π ) 2−πΌ ππ 1 1−πΌ σ΅¨ σ΅¨ πΈ [σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1{|ππ |≤π} ] ≤ ππΌ∗ ππΌ π (π π+1 − π π ) 1−πΌ π=1 (68) π=1 and hence Let ππ = π₯π πππ . Using (57) and Remark A.2 (Appendix A), we get π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ × ∑ πΈσ΅¨σ΅¨σ΅¨σ΅¨π½ππ σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ , (61) ππ ∑π=1 π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ × ∑ σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ . (67) π=1 (60) Using the independence between π½π and ππ it follows that σ΅¨σ΅¨ ππ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ πΈ [σ΅¨σ΅¨σ΅¨σ΅¨∑ π₯π πππ σ΅¨σ΅¨σ΅¨σ΅¨ 1{| ∑ππ π₯ π |≤π} ] ππ½ (ππ₯) . π π=1 π ππ σ΅¨ σ΅¨σ΅¨ σ΅¨ [σ΅¨σ΅¨π=1 ] (66) π = ∑π=1 π₯π πππ . Using (57) and Remark A.2 Let ππ (Appendix A), we get 1 πΈ (ππ−1 ) π 1 π−1 σ΅¨ σ΅¨ = ∑ πΈ (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1|ππ |≤π ) . π π=0 Rππ (65) where ππ = ππ 1{|ππ |≤π} − πΈ[ππ 1{|ππ |≤π} | Fπ π ] and ππ = |πΈ[ππ 1{|ππ |≤π} | Fπ π ]|. We first treat the term πΌπΌσΈ . Note that {ππ = π ∑π=0 ππ , Fπ π+1 ; 0 ≤ π ≤ π − 1} is a zero-mean square integrable martingale, and Case 1 (πΌ < 1). Note that =∫ (64) =: πΌπΌσΈ + πΌπΌσΈ σΈ , π−1 π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ πΌ ≤ ππΌ∗ π−πΌ ππΌ ∑ (π π+1 − π π ) ∑ πΈσ΅¨σ΅¨σ΅¨σ΅¨π½ππ σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ 0≤π≤π−1 (63) πΌπΌσΈ ≤ 8ππΌ∗ ππΌ π 1 −πΌ π πΈ ∫ ∫ |π (π , π₯)|πΌ ππ₯ ππ . 2−πΌ 0 π΅ (69) We now treat πΌπΌσΈ σΈ . Note that {ππ = ∑ππ=0 ππ , Fπ π+1 ; 0 ≤ π ≤ π − 1} is a semimartingale and hence, by the submartingale inequality, (62) πΌπΌσΈ σΈ ≤ 2 2 π−1 πΈ (ππ−1 ) = ∑ πΈ (ππ ) . π π π=0 (70) International Journal of Stochastic Analysis 9 To evaluate πΈ(ππ ), we note that, for almost all π ∈ Ω, Moreover, there exists a subsequence (ππ )π such that σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ (πππ ) (π‘, π΅) − π (π‘, π΅)σ΅¨σ΅¨σ΅¨ σ³¨→ 0 σ΅¨ σ΅¨ π‘∈[0,π] πΈ [ππ 1{|ππ |≤π} | Fπ π ] (π) sup ππ = πΈ [ ∑π½ππ (π) πππ 1{| ∑ππ π½ (π)π |≤π} ] , ππ π=1 ππ [π=1 ] (71) due to the independence between π½π and ππ . We let ππ = ππ ∑π=1 π₯π πππ with π₯π = π½ππ (π). Since πΌ > 1, πΈ(ππ ) = 0. Using (57) and Remark A.2, we obtain σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨σ΅¨πΈ [ππ 1{|ππ |≤π} ]σ΅¨σ΅¨σ΅¨ = σ΅¨σ΅¨σ΅¨πΈ [ππ 1{|ππ |>π} ]σ΅¨σ΅¨σ΅¨ ≤ πΈ [σ΅¨σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨σ΅¨ 1{|ππ |>π} ] σ΅¨ σ΅¨ σ΅¨ σ΅¨ πΌ ≤ ππΌ∗ ππΌ π1−πΌ (π π+1 − π π ) πΌ−1 (72) ππ σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ × ∑σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ . πΌπΌσΈ σΈ ≤ ππΌ∗ ππΌ (78) as π → ∞. Hence, π(π‘, π΅) is Fπ‘ -measurable for any π‘ ∈ [0, π]. The process π(⋅, π΅) does not depend on the sequence (ππ )π and can be extended to a caΜdlaΜg process on [0, ∞), which is unique up to indistinguishability. We denote this extension by πΌ(π)(⋅, π΅) and we write π‘ πΌ (π) (π‘, π΅) = ∫ ∫ π (π , π₯) π (ππ , ππ₯) . 0 (79) π΅ If π΄ and π΅ are disjoint sets in Bπ (Rπ ), then πΌ (π) (π‘, π΄ ∪ π΅) = πΌ (π) (π‘, π΄) + πΌ (π) (π‘, π΅) a.s. (80) Lemma 14. Inequality (46) holds for any π ∈ LπΌ . π=1 Hence, πΈ(ππ ) ≤ ππΌ∗ ππΌ (πΌ/(πΌ ππ πΌ π π ) ∑π=1 πΈ|π½ππ | |π»ππ ∩ π΅| and a.s. − 1))π1−πΌ (π π+1 π 2πΌ −πΌ π πΈ ∫ ∫ |π (π‘, π₯)|πΌ ππ₯ ππ‘. πΌ−1 0 π΅ − (73) Case 3 (πΌ = 1). In this case we assume that π½ = 0. Hence, ππ π₯π πππ has a symmetric distribution for any π₯ ∈ Rππ . ππ = ∑π=1 Using (71), it follows that πΈ[ππ 1{|ππ |≤π} | Fπ π ] = 0 a.s. for all π = 0, . . . , π − 1. Hence, {ππ = ∑ππ=0 ππ 1{|ππ |≤π} , Fπ π+1 ; 0 ≤ π ≤ π − 1} is a zero-mean square integrable martingale. By the martingale maximal inequality, Proof. Let (ππ )π be a sequence of simple functions such that βππ − πβπΌ → 0. For fixed π΅, we denote πΌ(π) = πΌ(π)(⋅, π΅). We let β ⋅ β∞ be the sup-norm on π·[0, π]. For any π > 0, we have σ΅© σ΅© π (βπΌ (π)β∞ > π) ≤ π (σ΅©σ΅©σ΅©πΌ (π) − πΌ (ππ )σ΅©σ΅©σ΅©∞ > ππ) (81) σ΅© σ΅© + π (σ΅©σ΅©σ΅©πΌ (ππ )σ΅©σ΅©σ΅©∞ > π (1 − π)) . Multiplying by ππΌ and using Theorem 13, we obtain sup ππΌ π (βπΌ (π)β∞ > π) π>0 σ΅© σ΅© ≤ π−πΌ sup ππΌ π (σ΅©σ΅©σ΅©πΌ(π) − πΌ(ππ )σ΅©σ΅©σ΅©∞ > π) π−1 1 1 2 πΌπΌ ≤ 2 πΈ [ππ−1 ] = 2 ∑ πΈ [ππ2 1{|ππ |≤π} ] . π π π=0 π>0 (74) The result follows using (68). (82) σ΅© σ΅©πΌ + (1 − π)−πΌ ππΌ σ΅©σ΅©σ΅©ππ σ΅©σ΅©σ΅©πΌ,π,π΅ . We now proceed to the construction of the stochastic integral. If π = {π(π‘)}π‘≥0 is a jointly measurable random process, we define Let π → ∞. Using (76) one can prove that supπ>0 ππΌ π(βπΌ(ππ ) − πΌ(π)β∞ > π) → 0. We obtain that supπ>0 ππΌ π(βπΌ(π)β∞ > π) ≤ (1 − π)−πΌ ππΌ βπβπΌπΌ,π,π΅ . The conclusion follows letting π → 0. βπβπΌπΌ,π = sup ππΌ π ( sup |π (π‘)| > π) . For an arbitrary Borel set O ⊂ Rπ (possibly O = Rπ ), we assume, in addition, that π ∈ LπΌ satisfies the condition: π>0 π‘∈[0,π] (75) Let π ∈ LπΌ be arbitrary. By Proposition 12, there exists a sequence (ππ )π≥1 of simple functions such that βππ − πβπΌ → 0 as π → ∞. Let π > 0 and π΅ ∈ Bπ (Rπ ) be fixed. By linearity of the integral and Theorem 13, σ΅© σ΅©πΌ σ΅©πΌ σ΅©σ΅© σ΅©σ΅©πΌ (ππ ) (⋅, π΅) − πΌ (ππ ) (⋅, π΅)σ΅©σ΅©σ΅©πΌ,π ≤ ππΌ σ΅©σ΅©σ΅©ππ − ππ σ΅©σ΅©σ΅©πΌ,π,π΅ σ³¨→ 0, (76) as π, π → ∞. In particular, the sequence {πΌ(ππ )(⋅, π΅)}π is Cauchy in probability in the space π·[0, π] equipped with the sup-norm. Therefore, there exists a random element π(⋅, π΅) in π·[0, π] such that, for any π > 0, σ΅¨ σ΅¨ π ( sup σ΅¨σ΅¨σ΅¨πΌ (ππ ) (π‘, π΅) − π (π‘, π΅)σ΅¨σ΅¨σ΅¨ > π) σ³¨→ 0. π‘∈[0,π] (77) π πΈ ∫ ∫ |π (π‘, π₯)|πΌ ππ₯ ππ‘ < ∞, 0 O ∀π > 0. (83) Then we can define πΌ(π)(⋅, O) as follows. Let Oπ = O ∩ πΈπ where (πΈπ )π is an increasing sequence of sets in Bπ (Rπ ) such that βπ πΈπ = Rπ . By (80), Lemma 14, and (83), σ΅¨ σ΅¨ sup ππΌ π (sup σ΅¨σ΅¨σ΅¨πΌ (π) (π‘, Oπ ) − πΌ (π) (π‘, Oπ )σ΅¨σ΅¨σ΅¨ > π) π>0 π‘≤π (84) π ≤ ππΌ πΈ ∫ ∫ 0 Oπ \Oπ πΌ |π (π‘, π₯)| ππ₯ ππ‘ σ³¨→ 0, as π, π → ∞. This shows that {πΌ(π)(⋅, Oπ )}π is a Cauchy sequence in probability in the space π·[0, π] equipped with 10 International Journal of Stochastic Analysis the sup-norm. We denote by πΌ(π)(⋅, O) its limit. As above, this process can be extended to [0, ∞) and πΌ(π)(π‘, O) is Fπ‘ measurable for any π‘ > 0. We denote π‘ πΌ (π) (π‘, O) = ∫ ∫ π (π , π₯) π (ππ , ππ₯) . 0 (85) O Similarly, to Lemma 14, one can prove that, for any π ∈ LπΌ satisfying (83), π‘≤π (86) π ≤ ππΌ πΈ ∫ ∫ |π (π‘, π₯)|πΌ ππ₯ ππ‘. 0 Lemma 16. If πΌ < 1, then σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) ≤ O πΈ (πππ’ππ,πΎ (π΅) ) = exp {|π΅| ∫ {π−1 <|π§|≤πΎ} ππΎ (π΅) = ∫ π§π (ππ , ππ₯, ππ§) , if πΌ ≤ 1, (87) ππΎ (π΅) = ∫ Μ (ππ , ππ₯, ππ§) , π§π if πΌ > 1. (88) We treat separately the cases πΌ ≤ 1 and πΌ > 1. 5.1. The Case πΌ ≤ 1. Note that {ππΎ (π΅); π΅ ∈ Bπ (R+ × Rπ )} is an independently scattered random measure on R+ ×Rπ with characteristic function given by πΈ (πππ’ππΎ (π΅) ) = exp {|π΅| ∫ |π§|≤πΎ (πππ’π§ − 1) ]πΌ (ππ§)} , ∀π’ ∈ R. (89) We first examine the tail of ππΎ (π΅). Lemma 15. For any set π΅ ∈ Bπ (R+ × Rπ ), σ΅¨ σ΅¨ sup ππΌ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ |π΅| , π>0 (90) where ππΌ > 0 is a constant depending only on πΌ (given by Lemma A.3). Proof. This follows from Example 3.7 of [1]. We denote by ]πΌ,πΎ the restriction of ]πΌ to {π§ ∈ R; 0 < |π§| ≤ πΎ}. Note that π‘−πΌ − πΎ−πΌ , ]πΌ,πΎ ({π§ ∈ R; |π§| > π‘}) = { 0, (92) (πππ’π§ − 1) ]πΌ (ππ§)} . (93) For the study of nonlinear equations, we need to develop a theory of stochastic integration with respect to another process ππΎ which is defined by removing from π the jumps whose modulus exceeds a fixed value πΎ > 0. More precisely, for any π΅ ∈ Bπ (R+ × Rπ ), we define π΅×{0<|π§|≤πΎ} ∀π’ > πΎ. Proof. We use the same idea as in Example 3.7 of [1]. For each π ≥ 1, let ππ,πΎ (π΅) be a random variable with characteristic function: 5. The Truncated Noise π΅×{0<|π§|≤πΎ} πΌ πΎ1−πΌ |π΅| π’−1 , 1−πΌ If πΌ = 1, then π(|ππΎ (π΅)| > π’) ≤ πΎ|π΅|π’−2 for all π’ > πΎ. sup ππΌ π (sup |πΌ (π) (π‘, O)| > π) π>0 In fact, since the tail of ]πΌ,πΎ vanishes if π‘ > πΎ, we can obtain another estimate for the tail of ππΎ (π΅) which, together with (90), will allow us to control its πth moment for π ∈ (πΌ, 1). This new estimate is given below. if 0 < π‘ ≤ πΎ, if π‘ > πΎ, (91) and hence supπ‘>0 π‘πΌ ]πΌ,πΎ ({π§ ∈ R; |π§| > π‘}) = 1. Next we observe that we do not need to assume that the measure ]πΌ,πΎ is symmetric since we use a modified version of Lemma 2.1 of [24] given by Lemma A.3 (Appendix A). Since {ππ,πΎ (π΅)}π converges in distribution to ππΎ(π΅), it suffices to prove the lemma for ππ,πΎ (π΅). Let ππ be the restriction of ]πΌ to {π§; π−1 < |π§| ≤ πΎ}. Since ππ is finite, ππ,πΎ (π΅) has a compound Poisson distribution with π |π΅| ∗π σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨ππ,πΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) = π−|π΅|ππ (R) ∑ ππ ({π§; |π§| > π’}) , π≥0 π! (94) where ππ∗π denotes the π-fold convolution. Note that σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π ππ∗π ({π§; |π§| > π’}) = [ππ (R)] π (σ΅¨σ΅¨σ΅¨∑ππ σ΅¨σ΅¨σ΅¨ > π’) , σ΅¨σ΅¨π=1 σ΅¨σ΅¨ σ΅¨ σ΅¨ (95) where (ππ )π≥1 are i.i.d. random variables with law ππ /ππ (R). Assume first that πΌ < 1. To compute π(| ∑ππ=1 ππ | > π’) we consider the intersection with the event {max1≤π≤π |ππ | > π’} and its complement. Note that π(|ππ | > π’) = 0 for any π’ > πΎ. Using this fact and Markov’s inequality, we obtain that, for any π’ > πΎ, σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π (σ΅¨σ΅¨σ΅¨∑ππ σ΅¨σ΅¨σ΅¨ > π’) ≤ π (σ΅¨σ΅¨σ΅¨∑ππ 1{|ππ |≤π’} σ΅¨σ΅¨σ΅¨ > π’) σ΅¨σ΅¨π=1 σ΅¨σ΅¨ σ΅¨σ΅¨π=1 σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ (96) π 1 σ΅¨σ΅¨ σ΅¨σ΅¨ ≤ ∑πΈ (σ΅¨σ΅¨ππ σ΅¨σ΅¨ 1{|ππ |≤π’} ) . π’ π=1 Note that π(|ππ | > π ) ≤ (π −πΌ − πΎ−πΌ )/ππ (R) if π ≤ πΎ. Hence, for any π’ > πΎ π’ πΎ 0 0 σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ πΈ (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1{|ππ |≤π’} ) ≤ ∫ π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π ) ππ = ∫ π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π ) ππ 1 πΌ ≤ πΎ1−πΌ . ππ (R) 1 − πΌ (97) Combining all these facts, we get that for any π’ > πΎ ππ∗π ({π§; |π§| > π’}) ≤ [ππ (R)] π−1 πΌ πΎ1−πΌ ππ’−1 , 1−πΌ and the conclusion follows from (94). (98) International Journal of Stochastic Analysis 11 Assume now that πΌ = 1. In this case, πΈ(ππ 1{|ππ |≤π’} ) = 0 since ππ has a symmetric distribution. Using Chebyshev’s inequality this time, we obtain σ΅¨σ΅¨σ΅¨ π σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨ π σ΅¨σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨∑ππ σ΅¨σ΅¨σ΅¨ > π’) ≤ π (σ΅¨σ΅¨σ΅¨∑ππ 1{|ππ |≤π’} σ΅¨σ΅¨σ΅¨ > π’) σ΅¨σ΅¨π=1 σ΅¨σ΅¨ σ΅¨σ΅¨π=1 σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ (99) π 1 2 ≤ 2 ∑πΈ (ππ 1{|ππ |≤π’} ) . π’ π=1 The result follows as above using the fact that, for any π’ > πΎ, π’ σ΅¨ σ΅¨ ≤ 2 ∫ π π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π ) ππ πΈ (ππ2 1{|ππ |≤π’} ) 0 πΎ σ΅¨ σ΅¨ = 2 ∫ π π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π ) ππ ≤ 0 1 πΎ. ππ (R) (100) We now proceed to the construction of the stochastic integral with respect to ππΎ . For this, we use the same method π π as for π. Note that Fπ‘ πΎ ⊂ Fπ‘ , where Fπ‘ πΎ is the π-field generated by ππΎ ([0, π ] × π΄) for all π ∈ [0, π‘] and π΄ ∈ Bπ (Rπ ). For any π΅ ∈ Bπ (Rπ ), we will work with a caΜdlaΜg modification of the LeΜvy process {ππΎ (π‘, π΅) = ππΎ ([0, π‘] × π΅); π‘ ≥ 0}. If π is a simple process given by (40), we define π‘ πΌπΎ (π) (π‘, π΅) = ∫ ∫ π (π , π₯) ππΎ (ππ , ππ₯) 0 by the same formula (44) with π replaced by ππΎ . The following result shows that πΌπΎ (π)(π‘, π΅) has the same tail behavior as πΌ(π)(π‘, π΅). σ΅¨ σ΅¨ sup ππΌ π ( sup σ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, π΅)σ΅¨σ΅¨σ΅¨ > π) π‘∈[0,π] π>0 πππ πππ¦ π ∈ (1, 2) , (102) where πΆπ is a constant depending on π. Proof. Note that ∞ σ΅¨π σ΅¨π σ΅¨ σ΅¨ πΈσ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ = ∫ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π‘) ππ‘ 0 ∞ σ΅¨ σ΅¨ = π ∫ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’. 0 (103) πΎ 0 0 σ΅¨ σ΅¨ ∫ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’ ≤ ππΌ |π΅| ∫ π’−πΌ+π−1 ππ’ 1 = ππΌ |π΅| πΎπ−πΌ . π−πΌ (104) ∞ (105) and if πΌ = 1, then ∞ ≤ πΎ |π΅| ∫ π’ πΎ π−3 1 ππ’ = |π΅| πΎπ−1 . 2−π 0 π΅ for any π > 0 and π΅ ∈ Bπ (Rπ ), where ππΌ is a constant depending only on πΌ. Proof. As in the proof of Theorem 13, it is enough to prove that σ΅¨σ΅¨ π ππ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π ( max σ΅¨σ΅¨σ΅¨σ΅¨∑ ∑π½ππ πππ∗ σ΅¨σ΅¨σ΅¨σ΅¨ > π) π=0,...,π−1 σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨π=0 π=1 σ΅¨ (109) ππ π−1 σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ ≤ ππΌ π−πΌ ∑ (π π+1 − π π ) ∑πΈσ΅¨σ΅¨σ΅¨σ΅¨π½ππ σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ , where πππ∗ that ππ∗ = π=1 = ππΎ ((π π , π π+1 ] × (π»ππ ∩ π΅)). This reduces to showing π π ∑π=1 π₯π πππ∗ satisfies an inequality similar to (57) for any π₯ ∈ Rππ ; that is, π π σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨ππ∗ σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ∗ π−πΌ (π π+1 − π π ) ∑ σ΅¨σ΅¨σ΅¨σ΅¨π₯π σ΅¨σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨σ΅¨π»ππ ∩ π΅σ΅¨σ΅¨σ΅¨σ΅¨ , σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨σ΅¨πππ∗ σ΅¨σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ (π π+1 − π π ) πΎππ π−πΌ , πΎ ∞ ≤ ππΌ πΈ ∫ ∫ |π (π‘, π₯)| ππ₯ ππ‘, (110) for any π > 0, for some ππΌ∗ > 0. We first examine the tail of πππ∗ . By (90), σ΅¨ σ΅¨ ∫ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’ σ΅¨ σ΅¨ ∫ π (σ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’ πΎ πΌ π=1 For the second one we use Lemma 16: if πΌ < 1 then ∞ πΌ ≤ πΎ1−πΌ |π΅| ∫ π’π−2 ππ’ 1−πΌ πΎ πΌ = |π΅| πΎπ−πΌ , (1 − πΌ) (1 − π) (108) π π=0 We consider separately the integrals for π’ ≤ πΎ and π’ > πΎ. For the first integral we use (90): πΎ (107) Proposition 18. If π is a bounded simple process then Lemma 17. If πΌ < 1 then σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ ≤ πΆπΌ,π πΎπ−πΌ |π΅| πππ πππ¦ π ∈ (πΌ, 1) , (101) where πΆπΌ,π is a constant depending on πΌ and π. If πΌ = 1, then σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨ππΎ (π΅)σ΅¨σ΅¨σ΅¨ ≤ πΆπ πΎπ−1 |π΅| π΅ (106) (111) where πΎππ = |π»ππ ∩ π΅|. Letting πππ = πΎππ−1/πΌ πππ∗ , we obtain that, for any π’ > 0, σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨σ΅¨πππ σ΅¨σ΅¨σ΅¨σ΅¨ > π’) ≤ ππΌ (π π+1 − π π ) π’−πΌ , ∀π = 1, . . . , ππ . (112) By Lemma A.3 (Appendix A), it follows that, for any π > 0, σ΅¨σ΅¨ σ΅¨σ΅¨ ππ ππ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨ σ΅¨πΌ σ΅¨ σ΅¨ π (σ΅¨σ΅¨ ∑ ππ πππ σ΅¨σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ2 (π π+1 − π π ) ∑σ΅¨σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨σ΅¨ π−πΌ , (113) σ΅¨σ΅¨ σ΅¨σ΅¨π=1 π=1 σ΅¨ σ΅¨ for any sequence (ππ )π=1,...,ππ of real numbers. Inequality (110) (with ππΌ∗ = ππΌ2 ) follows by applying this to ππ = π₯π πΎππ1/πΌ . 12 International Journal of Stochastic Analysis In view of the previous result and Proposition 12, for any process π ∈ LπΌ , we can construct the integral π‘ πΌπΎ (π) (π‘, π΅) = ∫ ∫ π (π , π₯) ππΎ (ππ , ππ₯) 0 π΅ (114) in the same manner as πΌ(π)(π‘, π΅), and this integral satisfies (108). If in addition the process π ∈ LπΌ satisfies (83), then we can define the integral πΌπΎ (π)(π‘, O) for an arbitrary Borel set O ⊂ Rπ (possibly O = Rπ ). This integral will satisfy an inequality similar to (108) with π΅ replaced by O. The appealing feature of πΌπΎ (π)(π‘, π΅) is that we can control its moments, as shown by the next result. Theorem 19. If πΌ < 1, then for any π ∈ (πΌ, 1) and for any π ∈ Lπ , π‘ σ΅¨σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, π΅)σ΅¨σ΅¨ ≤ πΆπΌ,π πΎπ−πΌ πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ , (115) 0 π΅ for any π‘ > 0 and π΅ ∈ Bπ (Rπ ), where πΆπΌ,π is a constant depending on πΌ, π. If O ⊂ Rπ is an arbitrary Borel set and we assume, in addition, that the process π ∈ Lπ satisfies π πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ < ∞, 0 O ∀π > 0, (116) then inequality (115) holds with π΅ replaced by O. Proof. Consider the following steps. Step 1. Suppose that π is an elementary process of the form (39). Then πΌπΎ (π)(π‘, π΅) = πππΎ (π») where π» = (π‘ ∧ π, π‘ ∧ π] × (π΄∩π΅). Note that ππΎ (π») is independent of Fπ . Hence, ππΎ (π») is independent of π. Let ππ denote the law of π. By Fubini’s theorem, σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨πππΎ (π»)σ΅¨σ΅¨σ΅¨ ∞ σ΅¨ σ΅¨ = π ∫ π (σ΅¨σ΅¨σ΅¨πππΎ (π»)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’ 0 (117) Step 2. Suppose now that π is a simple process of the form ππ (40). Then π(π‘, π₯) = ∑π−1 π=0 ∑π=1 πππ (π‘, π₯) where πππ (π‘, π₯) = 1(π‘π ,π‘π+1 ] (π‘)1π΄ ππ (π₯)πππ . Using the linearity of the integral, the inequality |π+π|π ≤ π |π| + |π|π , and the result obtained in Step 1 for the elementary processes πππ , we get σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, π΅)σ΅¨σ΅¨σ΅¨ π π−1 π σ΅¨π σ΅¨ ≤ πΈ ∑ ∑ σ΅¨σ΅¨σ΅¨σ΅¨πΌπΎ (πππ ) (π‘, π΅)σ΅¨σ΅¨σ΅¨σ΅¨ π=0 π=1 ≤ πΆπΌ,π πΎ π−πΌ π−1 ππ σ΅¨π σ΅¨ πΈ ∑ ∑ ∫ ∫ σ΅¨σ΅¨σ΅¨σ΅¨πππ (π , π₯)σ΅¨σ΅¨σ΅¨σ΅¨ ππ₯ ππ 0 π΅ π‘ = πΆπΌ,π πΎπ−πΌ πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ . 0 R 0 σ΅¨ σ΅¨πΌ ππΌ σ΅¨σ΅¨σ΅¨π¦σ΅¨σ΅¨σ΅¨ |π»| ∫ πΎ|π¦| 0 + π’−πΌ+π−1 ππ’ πΌ σ΅¨σ΅¨ σ΅¨σ΅¨ 1−πΌ σ΅¨π¦σ΅¨ πΎ |π»| 1−πΌσ΅¨ σ΅¨ ×∫ ∞ πΎ|π¦| σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, π΅)σ΅¨σ΅¨σ΅¨ σ΅¨π σ΅¨ ≤ lim inf πΈσ΅¨σ΅¨σ΅¨σ΅¨πΌπΎ (πππ ) (π‘, π΅)σ΅¨σ΅¨σ΅¨σ΅¨ π→∞ π‘ σ΅¨π σ΅¨ ≤ πΆπΌ,π πΎπ−πΌ lim inf πΈ ∫ ∫ σ΅¨σ΅¨σ΅¨σ΅¨πππ (π , π₯)σ΅¨σ΅¨σ΅¨σ΅¨ ππ₯ ππ π→∞ 0 π΅ σ΅¨ σ΅¨π σΈ π’π−2 ππ’ = πΆπΌ,π πΎπ−πΌ σ΅¨σ΅¨σ΅¨π¦σ΅¨σ΅¨σ΅¨ |π»| , = πΆπΌ,π πΎ π‘ π‘ = πΆπΌ,π πΎπ−πΌ πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ . Step 4. Suppose that π ∈ Lπ satisfies (116). Let Oπ = O ∩ πΈπ where (πΈπ )π is an increasing sequence of sets in Bπ (Rπ ) such that βπ≥1 πΈπ = Rπ . By the definition of πΌπΎ (π)(π‘, O), there exists a subsequence (ππ )π such that {πΌπΎ (π)(π‘, Oππ )}π converges to πΌπΎ (π)(π‘, O) a.s. Using Fatou’s lemma, the result obtained in Step 3 (for π΅ = Oππ ) and the monotone convergence theorem, we get σ΅¨ σ΅¨π ≤ lim inf πΈσ΅¨σ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, Oππ )σ΅¨σ΅¨σ΅¨σ΅¨ π→∞ π‘ π→∞ π π΅ π΅ ≤ πΆπΌ,π πΎπ−πΌ lim inf πΈ ∫ ∫ 0 π‘ πΈ ∫ ∫ |π (π , π₯)| ππ₯ ππ . 0 (120) σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, O)σ΅¨σ΅¨σ΅¨ σ΅¨π σ΅¨ σΈ πΈσ΅¨σ΅¨σ΅¨πππΎ (π»)σ΅¨σ΅¨σ΅¨ ≤ ππΆπΌ,π πΎπ−πΌ |π»| πΈ|π|π π−πΌ π΅ Step 3. Let π ∈ Lπ be arbitrary. By Proposition 12, there exists a sequence (ππ )π of bounded simple processes such that βππ − πβπ → 0. Since πΌ < π, it follows that βππ − πβπΌ → 0. By the definition of πΌπΎ (π)(π‘, π΅) there exists a subsequence {ππ }π such that {πΌπΎ (πππ )(π‘, π΅)}π converges to πΌπΎ (π)(π‘, π΅) a.s. Using Fatou’s lemma and the result obtained in Step 2 (for the simple processes πππ ), we get 0 We evaluate the inner integral. We split this integral into two parts, for π’ ≤ πΎ|π¦| and π’ > πΎ|π¦|, respectively. For the first integral, we use (90). For the second one, we use Lemma 16. Therefore, the inner integral is bounded by (119) π=0 π=1 ∞ σ΅¨ σ΅¨ = π ∫ (∫ π (σ΅¨σ΅¨σ΅¨π¦ππΎ (π»)σ΅¨σ΅¨σ΅¨ > π’) π’π−1 ππ’) ππ (ππ¦) . π‘ (118) Oππ |π (π , π₯)|π ππ₯ ππ = πΆπΌ,π πΎπ−πΌ πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ . 0 O (121) International Journal of Stochastic Analysis 13 Remark 20. Finding a similar moment inequality for the cases πΌ = 1 and π ∈ (1, 2) remains an open problem. The argument used in Step 2 above relies on the fact that π < 1. Unfortunately, we could not find another argument to cover the case π > 1. By the usual procedure, the integral can be extended to the class of all P × B(R \ {0})-measurable processes π such that [π]π,π,E < ∞. The integral is defined as an element in the space πΏπ (Ω; π·[0, π]) and will be denoted by π‘ Μ 5.2. The Case πΌ>1. In this case, the construction of the integral with respect to ππΎ relies on an integral with respect Μ which exists in the literature. We recall briefly the to π definition of this integral. For more details, see Section 1.2.2 of [6], Section 24.2 of [25], or Section 8.7 of [12]. Let E = Rπ × (R \ {0}) endowed with the measure π(ππ₯, ππ§) = ππ₯]πΌ (ππ§) and let Bπ (E) be the class of bounded Borel sets in E. For a simple process π = {π(π‘, π₯, π§); π‘ ≥ 0, Μ (π₯, π§) ∈ E}, the integral πΌπ(π)(π‘, π΅) is defined in the usual Μ way, for any π‘ > 0, π΅ ∈ Bπ (E). The process πΌπ(π)(⋅, π΅) is a (caΜdlaΜg) zero-mean square-integrable martingale with quadratic variation π‘ Μ [πΌπ (π) (⋅, π΅)] = ∫ ∫ |π (π , π₯, π§)|2 π (ππ , ππ₯, ππ§) π‘ 0 π΅ (122) and predictable quadratic variation 2 β¨πΌ (π) (⋅, π΅)β© = ∫ ∫ |π (π , π₯, π§)| ]πΌ (ππ§) ππ₯ ππ . π‘ 0 π΅ (123) π βπβ22,π,π΅ := πΈ ∫ ∫ |π (π , π₯, π§)|2 ]πΌ (ππ§) ππ₯ ππ < ∞. π΅ (124) The integral is a martingale with the same quadratic variations as above and has the isometry property: Μ πΈ|πΌπ(π)(π‘, π΅)|2 = βπβ22,π,π΅ . If, in addition, βπβ2,π,E < ∞, then the integral can be extended to E. By the Burkholder-DavisGundy inequality for discontinuous martingales, for any π ≥ 1, π/2 σ΅¨σ΅¨π σ΅¨σ΅¨ Μ Μ πΈ supσ΅¨σ΅¨σ΅¨πΌπ (π) (π‘, E)σ΅¨σ΅¨σ΅¨ ≤ πΆπ πΈ[πΌπ (π) (⋅, E)] . σ΅¨ π π‘≤π σ΅¨ σ΅¨σ΅¨π σ΅¨σ΅¨ Μ πΈ supσ΅¨σ΅¨σ΅¨πΌπ (π) (π‘, E)σ΅¨σ΅¨σ΅¨ σ΅¨ σ΅¨ π‘≤π π 0 Rπ Μ (ππ , ππ₯, ππ§) . π (π , π₯, π§) π (127) Its appealing feature is that it satisfies inequality (126). From now on, we fix π ∈ [1, 2]. Based on (88), for any π΅ ∈ Bπ (Rπ ), we let π‘ πΌπΎ (π) (π‘, π΅) = ∫ ∫ π (π , π₯) ππΎ (ππ , ππ₯) 0 π΅ (128) π‘ =∫ ∫ ∫ 0 π΅ {|π§|≤πΎ} Μ (ππ , ππ₯, ππ§) , π (π , π₯) π§π for any predictable process π = {π(π‘, π₯); π‘ ≥ 0, π₯ ∈ Rπ } for which the rightmost integral is well defined. Letting π(π‘, π₯, π§) = π(π‘, π₯)π§1{0<|π§|≤πΎ} , we see that this is equivalent to saying that π > πΌ and π ∈ Lπ . By (126), R\{0} σ΅¨π σ΅¨ πΈ supσ΅¨σ΅¨σ΅¨πΌπΎ (π) (π‘, π΅)σ΅¨σ΅¨σ΅¨ ≤ πΆπΌ,π πΎπ−πΌ πΈ ∫ ∫ |π (π , π₯)|π ππ₯ ππ , 0 π‘≤π π΅ where πΆπΌ,π = πΆπ πΌ/(π−πΌ). If, in addition, the process π ∈ Lπ satisfies (116) then (129) holds with π΅ replaced by O, for an arbitrary Borel set O ⊂ Rπ . Note that (129) is the counterpart of (115) for the case πΌ > 1. Together, these two inequalities will play a crucial role in Section 6. Table 1 summarizes all the conditions. 6. The Main Result In this section, we state and prove the main result regarding the existence of a mild solution of (1). For this result, O is a bounded domain in Rπ . For any π‘ > 0, we denote (125) The previous inequality is not suitable for our purposes. A more convenient inequality can be obtained for another stochastic integral, constructed for π ∈ [1, 2] fixed, as suggested on page 293 of [6]. More precisely, one can show that, for any bounded simple process π, ≤ πΆπ πΈ ∫ ∫ ∫ R\{0} (129) By approximation, this integral can be extended to the class of all P × B(R \ {0})-measurable processes π such that for any π > 0 and π΅ ∈ Bπ (E) 0 Rπ 0 π π‘ Μ π πΌπ,π (π) (π‘, E) = ∫ ∫ ∫ π π½π (π‘) = sup ∫ πΊ(π‘, π₯, π¦) ππ¦. π₯∈O O Theorem 21. Let πΌ ∈ (0, 2), πΌ =ΜΈ 1. Assume that for any π > 0 π σ΅¨π σ΅¨ lim ∫ ∫ σ΅¨σ΅¨πΊ (π‘, π₯, π¦) − πΊ (π‘ + β, π₯, π¦)σ΅¨σ΅¨σ΅¨ ππ¦ ππ‘ = 0, β→0 0 O σ΅¨ π |π (π‘, π₯, π§)|π ]πΌ (ππ§) ππ₯ ππ‘ (126) π =: |π|π,π,E , where πΆπ is the constant appearing in (125) (see Lemma 8.22 of [12]). (130) σ΅¨π σ΅¨ lim ∫ ∫ σ΅¨σ΅¨πΊ(π‘, π₯, π¦) − πΊ(π‘, π₯ + β, π¦)σ΅¨σ΅¨σ΅¨ ππ¦ ππ‘ = 0, |β| → 0 0 O σ΅¨ ∀π₯ ∈ O, (131) ∀π₯ ∈ O, (132) π ∫ π½π (π‘) ππ‘ < ∞, 0 (133) for some π ∈ (πΌ, 1) if πΌ < 1, or for some π ∈ (πΌ, 2] if πΌ > 1. Then (1) has a mild solution. Moreover, there exists a sequence 14 International Journal of Stochastic Analysis time and space increments of πΊ can be obtained directly from (35), as on page 196 of [26]. These arguments cannot be used when O is a bounded domain.) Table 1: Conditions for πΌπΎ (π)(π‘, π΅) to be well defined. πΌ<1 π΅ is bounded π΅ = O is unbounded π ∈ LπΌ π ∈ LπΌ and π satisfies (83) πΌ>1 π ∈ Lπ for some π ∈ (πΌ, 2] π ∈ Lπ and π satisfies (116) for some π ∈ (πΌ, 2] (ππΎ )πΎ≥1 of stopping times with ππΎ ↑ ∞ a.s. such that, for any π > 0 and πΎ ≥ 1, sup (π‘,π₯)∈[0,π]×O πΈ (|π’(π‘, π₯)|π 1{π‘≤ππΎ } ) < ∞. (134) Example 22 (heat equation). Let πΏ = π/ππ‘ − (1/2)Δ. Then πΊ(π‘, π₯, π¦) ≤ πΊ(π‘, π₯ − π¦) where πΊ(π‘, π₯) is the fundamental solution of πΏπ’ = 0 on Rπ . Condition (133) holds if π < 1 + 2/π. If πΌ < 1, this condition holds for any π ∈ (πΌ, 1). If πΌ > 1, this condition holds for any π ∈ (πΌ, 1 + 2/π], as long as πΌ satisfies (6). Conditions (131) and (132) hold by the continuity of the function πΊ in π‘ and π₯, by applying the dominated convergence theorem. To justify the application of this theorem, we use the trivial bound (2ππ‘)−ππ/2 for both πΊ(π‘ + β, π₯, π¦)π and πΊ(π‘, π₯ + β, π¦)π , which introduces the extra condition ππ < 2. Unfortunately, we could not find another argument for proving these two conditions (In the case of the heat equation on Rπ , Lemmas A.2 and A.3 of [6] estimate the integrals appearing in (132) and (131), with π = 1 in (131). These arguments rely on the structure of πΊ and cannot be used when O is a bounded domain.). Example 23 (parabolic equations). Let πΏ = π/ππ‘−L where L is given by (31). Assuming (32), we see that (133) holds if π < 1 + 2/π. The same comments as for the heat equation apply here as well (Although in a different framework, a condition similar to (131) was probably used in the proof of Theorem 12.11 of [12] (page 217) for the claim limπ → π‘ πΈ|π½3 (π)(π ) − π π½3 (π)(π‘)|πΏπ (O) = 0. We could not see how to justify this claim, unless ππ < 2.). Example 24 (heat equation with fractional power of the Laplacian). Let πΏ = π/ππ‘ + (−Δ)πΎ for some πΎ > 0. By Lemma B.23 of [12], if πΌ > 1, then condition (133) holds for any π ∈ (πΌ, 1+2πΎ/π), provided that πΌ satisfies (36) (This condition is the same as in Theorem 12.19 of [12], which examines the same equation using the approach based on Hilbert-space valued solution.). To verify conditions (131) and (132), we use the continuity of πΊ in π‘ and π₯ and apply the dominated convergence theorem. To justify the application of this theorem, we use the trivial bound πΆπ,πΎ π‘−ππ/(2πΎ) for both πΊ(π‘ + β, π₯, π¦)π and πΊ(π‘, π₯ + β, π¦)π , which introduces the extra condition ππ < 2πΎ. This bound can be seen from (33), using the fact that G(π‘, π₯, π¦) ≤ G(π‘, π₯ − π¦) where G and G are the fundamental solutions of ππ’/ππ‘ − Δπ’ = 0 on O and Rπ , respectively. (In the case of the same equation on Rπ , elementary estimates for the The remaining part of this section is dedicated to the proof of Theorem 21. The idea is to solve first the equation with the truncated noise ππΎ (yielding a mild solution π’πΎ ) and then identify a sequence (ππΎ )πΎ≥1 of stopping times with ππΎ ↑ ∞ a.s. such that, for any π‘ > 0, π₯ ∈ O, and πΏ > πΎ, π’πΎ (π‘, π₯) = π’πΏ (π‘, π₯) a.s. on the event {π‘ ≤ ππΎ }. The final step is to show that process π’ defined by π’(π‘, π₯) = π’πΎ (π‘, π₯) on {π‘ ≤ ππΎ } is a mild solution of (1). A similar method can be found in Section 9.7 of [12] using an approach based on stochastic integration of operator-valued processes, with respect to Hilbert-space-valued processes, which is different from our approach. Since π is a Lipschitz function, there exists a constant πΆπ > 0 such that |π (π’) − π (V)| ≤ πΆπ |π’ − V| , ∀π’, V ∈ R. (135) In particular, letting π·π = πΆπ ∨ |π(0)|, we have |π (π’)| ≤ π·π (1 + |π’|) , ∀π’ ∈ R. (136) For the proof of Theorem 21, we need a specific construction of the Poisson random measure π, taken from [13]. We review briefly this construction. Let (Oπ )π≥1 be a partition of Rπ with sets in Bπ (Rπ ) and let (ππ )π≥1 be a partition of R \ {0} such that ]πΌ (ππ ) < ∞ for all π ≥ 1. We may take ππ = Γπ−1 for all π ≥ 1. π,π π,π π,π Let (πΈπ , ππ , ππ )π,π,π≥1 be independent random variables defined on a probability space (Ω, F, π), such that σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨π΅ ∩ Oπ σ΅¨σ΅¨σ΅¨ π,π π,π −π π,π π‘ , π (ππ ∈ π΅) = σ΅¨σ΅¨ σ΅¨σ΅¨ , π (πΈπ > π‘) = π σ΅¨σ΅¨Oπ σ΅¨σ΅¨ (137) σ΅¨σ΅¨ σ΅¨ σ΅¨σ΅¨Γ ∩ ππ σ΅¨σ΅¨σ΅¨ π,π σ΅¨ σ΅¨ π (ππ ∈ Γ) = σ΅¨σ΅¨ σ΅¨σ΅¨ , σ΅¨σ΅¨ππ σ΅¨σ΅¨ σ΅¨ σ΅¨ π,π where π π,π = |Oπ |]πΌ (ππ ). Let ππ Then π,π = ∑ππ=1 πΈπ for all π ≥ 1. π = ∑ πΏ(ππ,π ,ππ,π ,ππ,π ) π,π,π≥1 π π π (138) is a Poisson random measure on R+ × Rπ × (R \ {0}) with intensity ππ‘ππ₯]πΌ (ππ§). This section is organized as follows. In Section 6.1 we prove the existence of the solution of the equation with truncated noise ππΎ . Sections 6.2 and 6.3 contain the proof of Theorem 21 when πΌ < 1 and πΌ > 1, respectively. 6.1. The Equation with Truncated Noise. In this section, we fix πΎ > 0 and we consider the equation: πΏπ’ (π‘, π₯) = π (π’ (π‘, π₯)) πΜ πΎ (π‘, π₯) , π‘ > 0, π₯ ∈ O (139) with zero initial conditions and Dirichlet boundary conditions. A mild solution of (139) is a predictable process π’ which International Journal of Stochastic Analysis 15 satisfies (2) with π replaced by ππΎ . For the next result, O can be a bounded domain in Rπ or O = Rπ (with no boundary conditions). Theorem 25. Under the assumptions of Theorem 21, (139) has a unique mild solution π’ = {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O}. For any π > 0, sup (π‘,π₯)∈[0,π]×O πΈ|π’(π‘, π₯)|π < ∞, (140) and the map (π‘, π₯) σ³¨→ π’(π‘, π₯) is continuous from [0, π] × O into πΏπ (Ω). Proof. We use the same argument as in the proof of Theorem 13 of [27], based on a Picard iteration scheme. We define π’0 (π‘, π₯) = 0 and π‘ π’π+1 (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’π (π , π¦)) ππΎ (ππ , ππ¦) 0 O (141) for any π ≥ 0. We prove by induction on π ≥ 0 that (i) π’π (π‘, π₯) is well defined; (ii) πΎπ (π‘) := sup(π‘,π₯)∈[0,π]×O πΈ|π’π (π‘, π₯)|π < ∞ for any π > 0; (iii) π’π (π‘, π₯) is Fπ‘ -measurable for any π‘ > 0 and π₯ ∈ O; (iv) the map (π‘, π₯) σ³¨→ π’π (π‘, π₯) is continuous from [0, π] × O into πΏπ (Ω) for any π > 0. The statement is trivial for π = 0. For the induction step, assume that the statement is true for π. By an extension to random fields of Theorem 30, Chapter IV of [28], π’π has a jointly measurable modification. Since this modification is (Fπ‘ )π‘ -adapted (in the sense of (iii)), it has a predictable modification (using an extension of Proposition 3.21 of [12] to random fields). We work with this modification, that we call also π’π . We prove that (i)–(iv) hold for π’π+1 . To show (i), it suffices to prove that ππ ∈ Lπ , where ππ (π , π¦) = 1[0,π‘] (π )πΊ(π‘ − π , π₯, π¦)π(π’π (π , π¦)). By (136) and (133), π‘ σ΅¨ σ΅¨π πΈ ∫ ∫ σ΅¨σ΅¨σ΅¨ππ (π , π¦)σ΅¨σ΅¨σ΅¨ ππ¦ ππ 0 O (142) π‘ ≤ π·ππ 2π−1 (1 + πΎπ (π‘)) ∫ π½π (π‘ − π ) ππ < ∞. 0 In addition, if O = Rπ , we have to prove that ππ satisfies (83) if πΌ < 1, or (116) if πΌ > 1 (see Table 1). If πΌ < 1, this follows as above, since πΌ < π and hence sup(π‘,π₯)∈[0,π]×O πΈ|π’(π‘, π₯)|πΌ < ∞; the argument for πΌ > 1 is similar. Combined with the moment inequality (115) (or (129)), this proves (ii), since and (π‘, π‘ + β], we obtain that πΈ|π’π+1 (π‘ + β, π₯) − π’π+1 (π‘, π₯)|π ≤ 2π−1 (πΌ1 (β) + πΌ2 (β)), where σ΅¨σ΅¨σ΅¨ π‘ πΌ1 (β) = πΈ σ΅¨σ΅¨σ΅¨∫ ∫ (πΊ (π‘ + β − π , π₯, π¦) − πΊ (π‘ − π , π₯, π¦)) σ΅¨σ΅¨ 0 O σ΅¨σ΅¨σ΅¨π × π (π’π (π , π¦)) ππΎ (ππ , ππ¦) σ΅¨σ΅¨σ΅¨ , σ΅¨σ΅¨ (144) σ΅¨σ΅¨σ΅¨ π‘+β σ΅¨ σ΅¨ πΌ2 (β) = πΈ σ΅¨σ΅¨∫ ∫ πΊ (π‘ + β − π , π₯, π¦) π σ΅¨σ΅¨ π‘ O σ΅¨σ΅¨π σ΅¨ × (π’π (π , π¦)) ππΎ (ππ , ππ¦) σ΅¨σ΅¨σ΅¨σ΅¨ . σ΅¨σ΅¨ Using again (136) and the moment inequality (115) (or (129)), we obtain πΌ1 (β) ≤ π·ππ 2π−1 (1 + πΎπ (π‘)) π‘ σ΅¨π σ΅¨ × ∫ ∫ σ΅¨σ΅¨σ΅¨πΊ (π + β, π₯, π¦) − πΊ (π , π₯, π¦)σ΅¨σ΅¨σ΅¨ ππ¦ ππ , 0 O πΌ2 (β) ≤ π·ππ 2π−1 (1 + πΎπ (π‘)) β π × ∫ ∫ πΊ(π , π₯, π¦) ππ¦ ππ . 0 O It follows that both πΌ1 (β) and πΌ2 (β) converge to 0 as β → 0, using (131) for πΌ1 (β) and the Dominated Convergence Theorem and (133) for πΌ2 (β), respectively. The left continuity in π‘ is similar, by writing the interval [0, π‘−β] as the difference between [0, π‘] and (π‘ − β, π‘] for β > 0. For the continuity in π₯, similarly as above, we see that πΈ|π’π+1 (π‘, π₯ + β) − π’π+1 (π‘, π₯)|π is bounded by π·ππ 2π−1 (1 + πΎπ (π‘)) π‘ σ΅¨π σ΅¨ × ∫ ∫ σ΅¨σ΅¨σ΅¨πΊ (π , π₯ + β, π¦) − πΊ (π , π₯, π¦)σ΅¨σ΅¨σ΅¨ ππ¦ ππ , 0 O π‘ ππ (π‘) ≤ πΆ1 ∫ (1 + ππ−1 (π )) π½π (π‘ − π ) ππ , 0 π‘ (143) 0 for any π₯ ∈ O. Property (iii) follows by the construction of the integral πΌπΎ . To prove (iv), we first show the right continuity in π‘. Let β > 0. Writing the interval [0, π‘ + β] as the union of [0, π‘] ∀π ≥ 1, (147) where πΆ1 = πΆπΌ,π πΎπ−πΌ π·ππ 2π−1 . By applying Lemma 15 of Erratum to [27] with ππ = ππ , π1 = 0, π2 = 1, and π(π ) = πΆπ½π (π ), we obtain that π≥0 π‘∈[0,π] ≤ πΆπΌ,π πΎπ−πΌ π·ππ 2π−1 (1 + πΎπ (π‘)) ∫ π½π (π‘ − π ) ππ , (146) which converges to 0 as |β| → 0 due to (132). This finishes the proof of (iv). We denote ππ (π‘) = supπ₯∈O πΈ|π’π (π‘, π₯)|π . Similarly to (143), we have sup sup ππ (π‘) < ∞, σ΅¨π σ΅¨ πΈσ΅¨σ΅¨σ΅¨π’π+1 (π‘, π₯)σ΅¨σ΅¨σ΅¨ (145) ∀π > 0. (148) We now prove that {π’π (π‘, π₯)}π converges in πΏπ (Ω), uniformly in (π‘, π₯) ∈ [0, π] × O. To see this, let ππ (π‘) = supπ₯∈O πΈ|π’π+1 (π‘, π₯) − π’π (π‘, π₯)|π for π ≥ 0. Using the moment inequality (115) (or (129)) and (135), we have π‘ ππ (π‘) ≤ πΆ2 ∫ ππ−1 (π ) π½π (π‘ − π ) ππ , 0 (149) 16 International Journal of Stochastic Analysis where πΆ2 = πΆπΌ,π πΎπ−πΌ πΆππ . By Lemma 15 of Erratum to [27], ∑π≥0 ππ (π‘)1/π converges uniformly on [0, π] (Note that this lemma is valid for all π > 0.). We denote by π’(π‘, π₯) the limit of π’π (π‘, π₯) in πΏπ (Ω). One can show that π’ satisfies properties (ii)–(iv) listed above. So π’ has a predictable modification. This modification is a solution of (139). To prove uniqueness, let V be another solution and denote π»(π‘) = supπ₯∈O πΈ|π’(π‘, π₯) − V(π‘, π₯)|π . Then π‘ π» (π‘) ≤ πΆ2 ∫ π» (π ) π½π (π‘ − π ) ππ . (150) 0 for any πΏ > πΎ, π‘ ∈ [0, π], and π΄ ∈ Bπ (Rπ ) with π΄ ⊂ π΅. Using an approximation argument and the construction of the integrals πΌ(π) and πΌπΎ (π), it follows that, for any π ∈ LπΌ and for any πΏ > πΎ, a.s. on {ππΎ (π΅) > π}, we have πΌ (π) (π, π΅) = πΌπΎ (π) (π, π΅) = πΌπΏ (π) (π, π΅) . π}. The next result gives the probability of the event {ππΎ (π΅) > Lemma 26. For any π > 0 and π΅ ∈ Bπ (Rπ ), π (ππΎ (π΅) > π) = exp (−π |π΅| πΎ−πΌ ) . Using (133), it follows that π»(π‘) = 0 for all π‘ > 0. 6.2. Proof of Theorem 21: Case πΌ<1. In this case, for any π‘ > 0 and π΅ ∈ Bπ (Rπ ), we have (see (21)) π (π‘, π΅) = ∫ [0,π‘]×π΅×(R\{0}) π§π (ππ , ππ₯, ππ§) . (151) The characteristic function of π(π‘, π΅) is given by ππ’π(π‘,π΅) πΈ (π ππ’π§ ) = exp {π‘ |π΅| ∫ R\{0} (π Consequently, limπΎ → ∞ π(ππΎ (π΅) limπΎ → ∞ ππΎ (π΅) = ∞ a.s. − 1) ]πΌ (ππ§)} , (152) Note that {π(π‘, π΅)}π‘≥0 is not a compound Poisson process since ]πΌ is infinite. We introduce the stopping times (ππΎ )πΎ≥1 , as on page 239 of [13]: ππΎ (π΅) = inf {π‘ > 0; |π (π‘, π΅) − π (π‘−, π΅)| > πΎ} , (153) where π(π‘−, π΅) = limπ ↑π‘ π(π , π΅). Clearly, ππΏ (π΅) ≥ ππΎ (π΅) for all πΏ > πΎ. We first investigate the relationship between π and ππΎ and the properties of ππΎ (π΅). Using construction (138) of π and definition (87) of ππΎ , we have Since ]πΌ ({π§; |π§| > πΎ}) = πΎ−πΌ and (ππΎ (π΅))π,π≥1 are independent, it is enough to prove that, for any π, π ≥ 1, π,π σ΅¨ σ΅¨ π (ππΎ (π΅) > π) = exp {−π σ΅¨σ΅¨σ΅¨π΅ ∩ Oπ σ΅¨σ΅¨σ΅¨ ]πΌ ({π§; |π§| > πΎ} ∩ ππ )} . (160) π,π π,π Note that {ππΎ (π΅) > π} = {π; |ππ (π)| ≤ πΎ for all π π,π π,π for which ππ ≤ π and ππ ∈ π΅} and (πππ,π )π≥1 are the jump times of a Poisson process with intensity π π,π . Hence, π,π π (ππΎ (π΅) > π) π =∑∑ π,π ∑ π≥0 π=0 πΌ⊂{1,...,π},card(πΌ)=π π,π π∈πΌ (154) π,π π,π,π≥1 σ΅¨σ΅¨ π,π σ΅¨σ΅¨ × π (β {σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ ≤ πΎ}) σ΅¨ σ΅¨ π∈πΌ π π,π We observe that {π (π‘, π΅)}π‘≥0 is a compound Poisson process with π,π πΈ (πππ’π (π‘,π΅) ) π,π π∈πΌπ ππ = ∑π ∀π’ ∈ R. (155) Note that ππΎ (π΅) > π means that all the jumps of {π(π‘, π΅)}π‘≥0 in [0, π] are smaller than πΎ in modulus; that is, π,π {ππΎ (π΅) > π} = {π; |ππ (π)| ≤ πΎ for all π, π, π ≥ 1 for which π,π π,π ππ (π) ≤ π and ππ (π) ∈ π΅}. Hence, on {ππΎ (π΅) > π}, π ([0, π‘] × π΄) = ππΎ ([0, π‘] × π΄) = ππΏ ([0, π‘] × π΄) , (156) (161) × π (β {ππ ∉ π΅}) −π π,π π σ΅¨ σ΅¨ = exp {π‘ σ΅¨σ΅¨σ΅¨Oπ ∩ π΅σ΅¨σ΅¨σ΅¨ ∫ (πππ’π§ − 1) ]πΌ (ππ§)} , π (πππ,π ≤ π < ππ+1 ) × π (β {ππ ∈ π΅}) π,π≥1 π 1 and σ΅¨ σ΅¨ π,π ππΎ (π΅) = inf {π‘ > 0; σ΅¨σ΅¨σ΅¨σ΅¨ππ,π (π‘, π΅) − ππ,π (π‘−, π΅)σ΅¨σ΅¨σ΅¨σ΅¨ > πΎ} . (159) ππΎ (π‘, π΅) = ∑ ππ 1{|ππ,π |≤πΎ} 1{ππ,π ≤π‘} 1{ππ,π ∈π΅} . π = π,π π,π π π) Proof. Note that {ππΎ (π΅) > π} = βπ,π≥1 {ππΎ (π΅) > π}, where π (π‘, π΅) = ∑ ππ 1{ππ,π ≤π‘} 1{ππ,π ∈π΅} =: ∑ ππ,π (π‘, π΅) , π > (158) π,π ∀π’ ∈ R. π,π,π≥1 (157) π≥0 (π π,π π) π π! π σ΅¨σ΅¨ π,π σ΅¨σ΅¨ π,π × [1 − π (π1 ∈ π΅) π (σ΅¨σ΅¨σ΅¨π1 σ΅¨σ΅¨σ΅¨ > πΎ)] σ΅¨ σ΅¨ σ΅¨σ΅¨ π,π σ΅¨σ΅¨ π,π = exp {−π π,π ππ (π1 ∈ π΅) π (σ΅¨σ΅¨σ΅¨π1 σ΅¨σ΅¨σ΅¨ > πΎ)} , σ΅¨ σ΅¨ which yields (160). To prove the last statement, let π΄(π) π = {ππΎ (π΅) > π}. Then (π) ) ≥ lim π(π΄ ) = 1 for any π ≥ 1, and hence π(limπΎ π΄(π) πΎ πΎ πΎ (π) π(βπ≥1 limπΎ π΄ πΎ ) = 1. Hence, with probability 1, for any International Journal of Stochastic Analysis 17 π, there exists some πΎπ such that ππΎπ > π. Since (ππΎ )πΎ is nondecreasing, this proves that ππΎ → ∞ with probability 1. Remark 27. The construction of ππΎ (π΅) given above is due to [13] (in the case of a symmetric measure ]πΌ ). This construction relies on the fact that π΅ is a bounded set. Since π(π‘, Rπ ) (and consequently ππΎ (Rπ )) is not well defined, we could not see why this construction can also be used when π΅ = Rπ , as it is claimed in [13]. To avoid this difficulty, one could try to use an increasing sequence (πΈπ )π of sets in Bπ (Rπ ) with βπ πΈπ = Rπ . Using (157) with π΅ = πΈπ and letting π → ∞, we obtain that πΌ(π)(π‘, Rπ ) = πΌπΎ (π‘, Rπ ) a.s. on {π‘ ≤ ππΎ }, where ππΎ = inf π≥1 ππΎ (πΈπ ). But π(ππΎ > π‘) ≤ π(limπ {ππΎ (πΈπ ) > π‘}) ≤ limπ π(ππΎ (πΈπ ) > π‘) = limπ exp(−π‘|πΈπ |πΎ−πΌ ) = 0 for any π‘ > 0, which means that ππΎ = 0 a.s. Finding a suitable sequence (ππΎ )πΎ of stopping times which could be used in the case O = Rπ remains an open problem. In what follows, we denote ππΎ = ππΎ (O). Let π’πΎ be the solution of (139), whose existence is guaranteed by Theorem 25. Lemma 28. Under the assumptions of Theorem 21, for any π‘ > 0, π₯ ∈ O, and πΏ > πΎ, π’πΎ (π‘, π₯) = π’πΏ (π‘, π₯) π.π . ππ {π‘ ≤ ππΎ } . (162) Proof. By the definition of π’πΏ and (157), Proposition 29. Under the assumptions of Theorem 21, the process π’ = {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O} defined by π’ (π, π‘, π₯) = π’πΎ (π, π‘, π₯) , π’ (π, π‘, π₯) = 0, π’πΏ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’πΏ (π , π¦)) ππΏ (ππ , ππ¦) π‘ = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’πΏ (π , π¦)) ππΎ (ππ , ππ¦) O (163) a.s. on the event {π‘ ≤ ππΎ }. Using the definition of π’πΎ and Proposition C.1 (Appendix C), we obtain that, with probability 1, π’ (π‘, π₯) = lim (π’πΎ (π‘, π₯) 1{π‘≤ππΎ } ) 1Ω∗π‘,π₯ . πΎ→∞ O The process π(π, π‘, π₯) = 1{π‘≤ππΎ } (π) is clearly predictable, being in the class C defined in Remark 11. By the definition of Ωπ‘,π₯ , since π’πΎ , π’πΏ are predictable, it follows that (π, π‘, π₯) σ³¨→ 1Ω∗π‘,π₯ (π) is P-measurable. Hence, π’ is predictable. We now prove that π’ satisfies (2). Let π‘ > 0 and π₯ ∈ O be arbitrary. Using (157) and Proposition C.1 (Appendix C), with probability 1, we have 1{π‘≤ππΎ } π’ (π‘, π₯) = 1{π‘≤ππΎ } π’πΎ (π‘, π₯) = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) π 0 O × (π’πΎ (π , π¦)) ππΎ (ππ , ππ¦) π‘ = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) π 0 O × (π’πΎ (π , π¦)) π (ππ , ππ¦) π‘ = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) π 0 π‘ (164) = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) π 0 O × (π’ (π , π¦)) 1{π ≤ππΎ } π (ππ , ππ¦) × 1{π ≤ππΎ } ππΎ (ππ , ππ¦) . π‘ Let π(π‘) = supπ₯∈O πΈ(|π’πΎ (π‘, π₯) − π’πΏ (π‘, π₯)|π 1{π‘≤ππΎ } ). Using the moment inequality (115) and the Lipschitz condition (135), we get π‘ 0 (168) O × (π (π’πΎ (π , π¦)) − π (π’πΏ (π , π¦))) π (π‘) ≤ πΆ ∫ π½π (π‘ − π ) π (π ) ππ , (167) × (π’πΎ (π , π¦)) 1{π ≤ππΎ } π (ππ , ππ¦) π‘ 0 (166) Proof. We first prove that π’ is predictable. Note that (π’πΎ (π‘, π₯) − π’πΏ (π‘, π₯)) 1{π‘≤ππΎ } = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) ππ π ∉ Ω∗π‘,π₯ is a mild solution of (1). O 0 ππ π ∈ Ω∗π‘,π₯ , π‘ ≤ ππΎ (π) π‘ π‘ 0 For any π‘ > 0 and π₯ ∈ O, let Ωπ‘,π₯ = βπΏ>πΎ {π‘ ≤ ππΎ (π‘), π’πΎ (π‘, π₯) =ΜΈ π’πΏ (π‘, π₯)}, where πΏ and πΎ are positive integers. Let Ω∗π‘,π₯ = Ωπ‘,π₯ ∩ {limπΎ → ∞ ππΎ = ∞}. By Lemmas 26 and 28, π(Ω∗π‘,π₯ ) = 1. The next result concludes the proof of Theorem 21. (165) where πΆ = πΆπΌ,π πΎπ−πΌ πΆππ . Using (133), it follows that π(π‘) = 0 for all π‘ > 0. = 1{π‘≤ππΎ } ∫ ∫ πΊ (π‘ − π , π₯, π¦) π 0 O × (π’ (π , π¦)) π (ππ , ππ¦) . For the second last equality, we used the fact that processes π(π , π¦) = 1[0,π‘] (π )πΊ(π‘ − π , π₯, π¦)π(π’πΎ (π , π¦))1{π ≤ππΎ } and π(π , π¦) = 1[0,π‘] (π )πΊ(π‘ − π , π₯, π¦)π(π’(π , π¦))1{π ≤ππΎ } are modifications of each other (i.e., π(π , π¦) = π(π , π¦) a.s. for all π > 0, π¦ ∈ O), and, hence, [π − π]πΌ,π‘,O = 0 and πΌ(π)(π‘, O) = πΌ(π)(π‘, O) a.s. The conclusion follows letting πΎ → ∞, since ππΎ → ∞ a.s. 18 International Journal of Stochastic Analysis 6.3. Proof of Theorem 21: Case πΌ>1. In this case, for any π‘ > 0 and π΅ ∈ Bπ (Rπ ), we have (see (22)) π (π‘, π΅) = ∫ [0,π‘]×π΅×(R\{0}) Μ (ππ , ππ₯, ππ§) . π§π (169) π (π‘, π΅) = π (π‘, π΅) + π (π‘, π΅) − π‘ |π΅| π. (170) We let ππΎ (π‘, π΅) = ππΎ (π‘, π΅) − πΈ[ππΎ (π‘, π΅)] = ππΎ (π‘, π΅) − π‘|π΅|ππΎ , where [0,π‘]×π΅×(R\{0}) π πΌ (π) (π, π΅) = πΌπΎ (π) (π, π΅) − ππΎ ∫ ∫ π (π , π¦) ππ¦ ππ 0 π§1{1<|π§|≤πΎ} π (ππ , ππ₯, ππ§) ππΎ (π‘, π΅) = π (π‘, π΅) + ππΎ (π‘, π΅) − π‘ |π΅| ππΎ . (172) 0 πΏπ’ (π‘, π₯) = π (π’ (π‘, π₯)) πΜ πΎ (π‘, π₯) − ππΎ π (π’ (π‘, π₯)) , π‘ > 0, π₯ ∈ O where π(π‘−, π΅) = limπ ↑π‘ π(π , π΅). Lemma 26 holds again, but its proof is simpler than in the case πΌ < 1, since {π(π‘, π΅)}π‘≥0 is a compound Poisson process. By (138), π‘ π’ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’ (π , π¦)) ππΎ (ππ , ππ¦) 0 π,π,π≥1 π‘ − ππΎ ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’ (π , π¦)) ππ¦ ππ 0 O π π (174) π,π π π (179) for any π‘ > 0, π₯ ∈ O. The existence and uniqueness of a mild solution of (178) can be proved similarly to Theorem 25. We omit these details. We denote this solution by VπΎ . Lemma 30. Under the assumptions of Theorem 21, for any π‘ > 0, π₯ ∈ O, and πΏ > πΎ, Proof. By the definition of VπΏ and (177), a.s. on the event {π‘ ≤ ππΎ }, VπΏ (π‘, π₯) is equal to π‘ ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (VπΏ (π , π¦)) ππΏ (ππ , ππ¦) 0 O π‘ − ππΏ ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (VπΏ (π , π¦)) ππ¦ ππ O (181) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (VπΏ (π , π¦)) ππΎ (ππ , ππ¦) 0 O 0 (175) O Using the definition of VπΎ and Proposition C.1 (Appendix C), we obtain that, with probability 1, (VπΎ (π‘, π₯) − VπΏ (π‘, π₯)) 1{π‘≤ππΎ } π‘ Let ππΎ = π − ππΎ = ∫|π§|>πΎ π§]πΌ (ππ§). Using (170) and (172), it follows that = ππΏ ([0, π‘] × π΄) − π‘ |π΄| ππΏ (180) − ππΎ ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (VπΏ (π , π¦)) ππ¦ ππ . Hence, on {ππΎ (π΅) > π}, for any πΏ > πΎ, π‘ ∈ [0, π], and π΄ ∈ Bπ (Rπ ) with π΄ ⊂ π΅, π ([0, π‘] × π΄) = ππΎ ([0, π‘] × π΄) − π‘ |π΄| ππΎ π.π . ππ {π‘ ≤ ππΎ } . π‘ π π ([0, π‘] × π΄) = ππΎ ([0, π‘] × π΄) = ππΏ ([0, π‘] × π΄) . a.s. π‘ ππΎ (π‘, π΅) = ∑ ππ 1{1<|ππ,π |≤πΎ} 1{ππ,π ≤π‘} 1{ππ,π ∈π΅} . π,π,π≥1 O 0 π,π π (π‘, π΅) = ∑ ππ 1{|ππ,π |>1} 1{ππ,π ≤π‘} 1{ππ,π ∈π΅} , (178) with zero initial conditions and Dirichlet boundary conditions. A mild solution of (178) is a predictable process π’ which satisfies VπΎ (π‘, π₯) = VπΏ (π‘, π₯) (173) O We denote ππΎ = ππΎ (O). We consider the following equation: For any πΎ > 0, we let ππΎ (π΅) = inf {π‘ > 0; |π (π‘, π΅) − π (π‘−, π΅)| > πΎ} , (177) = πΌπΏ (π) (π, π΅) − ππΏ ∫ ∫ π (π , π¦) ππ¦ ππ . (171) and ππΎ = ∫1<|π§|≤πΎ π§]πΌ (ππ§). Recalling definition (88) of ππΎ , it follows that π O π To introduce the stopping times (ππΎ )πΎ≥1 we use the same idea as in Section 9.7 of [12]. Let π(π‘, π΅) = ∑π≥1 (πΏ π (π‘, π΅) − πΈπΏ π (π‘, π΅)) and π(π‘, π΅) = πΏ 0 (π‘, π΅), where πΏ π (π‘, π΅) = πΏ π ([0, π‘] × π΅) was defined in Section 2. Note that {π(π‘, π΅)}π‘≥0 is a zero-mean squareintegrable martingale and {π(π‘, π΅)}π‘≥0 is a compound Poisson process with πΈ[π(π‘, π΅)] = π‘|π΅|π where π = ∫|π§|>1 π§]πΌ (ππ§) = π½(πΌ/(πΌ − 1)). With this notation, ππΎ (π‘, π΅) = ∫ for any π ∈ LπΌ and for any πΏ > πΎ, a.s. on {ππΎ (π΅) > π}, we have (176) for any πΏ > πΎ, π‘ ∈ [0, π], and π΄ ∈ Bπ (Rπ ) with π΄ ⊂ π΅. Let π ∈ (πΌ, 2] be fixed. Using an approximation argument and the construction of the integrals πΌ(π) and πΌπΎ (π), it follows that, = 1{π‘≤ππΎ } (∫ ∫ πΊ (π‘ − π , π₯, π¦) (π (VπΎ (π , π¦))) 0 O − π (VπΏ (π , π¦)) 1{π ≤ππΎ } ππΎ (ππ , ππ¦) π‘ − ∫ ∫ πΊ (π‘ − π , π₯, π¦) (π (VπΎ (π , π¦))) 0 O −π (VπΏ (π , π¦)) 1{π ≤ππΎ } ππ¦ ππ ) . (182) International Journal of Stochastic Analysis 19 Letting π(π‘) = supπ₯∈O πΈ(|VπΎ (π‘, π₯) − VπΏ (π‘, π₯)|π 1{π‘≤ππΎ } ), we see that π(π‘) ≤ 2π−1 (πΈ|π΄(π‘, π₯)|π + πΈ|π΅(π‘, π₯)|π ) where π‘ 1{π‘≤ππΎ } π’ (π‘, π₯) π΄ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) (π (VπΎ (π , π¦))) 0 O − π (VπΏ (π , π¦)) 1{π ≤ππΎ } ππΎ (ππ , ππ¦) , π‘ Proof. We proceed as in the proof of Proposition 29. In this case, with probability 1, we have π‘ (183) = 1{π‘≤ππΎ } (∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’ (π , π¦)) π (ππ , ππ¦) 0 π‘ π΅ (π‘, π₯) = ∫ ∫ πΊ (π‘ − π , π₯, π¦) (π (VπΎ (π , π¦))) 0 − ππΎ ∫ ∫ πΊ (π‘ − π , π₯, π¦) π (π’ (π , π¦)) ππ¦ ππ ) . O 0 − π (VπΏ (π , π¦)) 1{π ≤ππΎ } ππ¦ ππ . We estimate separately the two terms. For the first term, we use the moment inequality (129) and the Lipschitz condition (135). We get π‘ sup πΈ|π΄ (π‘, π₯)|π ≤ πΆ ∫ π½π (π‘ − π ) π (π ) ππ , 0 π₯∈O (184) where πΆ = πΆπΌ,π πΎπ−πΌ πΆππ . For the second term, we use HoΜlder’s inequality | ∫ ππππ| ≤ (∫ |π|π ππ)1/π (∫ |π|π ππ)1/π with π(π , π¦) = πΊ(π‘ − π , π₯, π¦)1/π (π(VπΎ (π , π¦))) − π(VπΏ (π , π¦))1{π ≤ππΎ } and π(π , π¦) = πΊ(π‘ − π , π₯, π¦)1/π , where π−1 + π−1 = 1. Hence, π‘ O σ΅¨π σ΅¨ × σ΅¨σ΅¨σ΅¨VπΎ (π , π¦) − VπΏ (π , π¦)σ΅¨σ΅¨σ΅¨ 1{π ≤ππΎ } ππ¦ ππ , (185) π‘ where πΎπ‘ = ∫0 π½1 (π )ππ < ∞ (Since O is a bounded set, π½1 (π ) ≤ πΆπ½π (π )1/π where πΆ is a constant depending on |O| and π. Since π‘ π‘ π > 1, ∫0 π½π (π )1/π ππ ≤ ππ‘ (∫0 π½π (π )ππ )1/π < ∞ by (133). This shows that πΎπ‘ < ∞.). Therefore, π‘ sup πΈ|π΅ (π‘, π₯)|π ≤ πΆπ‘ ∫ π½1 (π‘ − π ) π (π ) ππ , 0 π₯∈O (186) π/π where πΆπ‘ = πΆππ πΎπ‘ . From (184) and (186), we obtain that π‘ π (π‘) ≤ πΆπ‘σΈ ∫ (π½π (π‘ − π ) + π½1 (π‘ − π )) π (π ) ππ , 0 where π‘ > 0. π−1 =2 (189) The conclusion follows letting πΎ → ∞, since ππΎ → ∞ a.s. and ππΎ → 0. Appendices A. Some Auxiliary Results The following result is used in the proof of Theorem 13. Lemma A.1. If π has a ππΌ (π, π½, 0) distribution then For any π‘ > 0 and π₯ ∈ O, we let Ωπ‘,π₯ = βπΏ>πΎ {π‘ ≤ ππΎ , VπΎ (π‘, π₯) =ΜΈ VπΏ (π‘, π₯)} where πΎ and πΏ are positive integers, and Ω∗π‘,π₯ = Ωπ‘,π₯ ∩ {limπΎ → ∞ ππΎ = ∞}. By Lemma 30, π(Ω∗π‘,π₯ ) = 1. Proposition 31. Under the assumptions of Theorem 21, the process π’ = {π’(π‘, π₯); π‘ ≥ 0, π₯ ∈ O} defined by π’ (π, π‘, π₯) = 0, is a mild solution of (1). ππ π ∈ Ω∗π‘,π₯ , π‘ ≤ ππΎ (π) , ππ π ∉ Ω∗π‘,π₯ (A.1) Proof. Consider the following steps. Step 1. We first prove the result for π = 1. We treat only the right tail, with the left tail being similar. We denote π by ππ½ to emphasize the dependence on π½. By Property 1.2.15 of [18], limπ → ∞ ππΌ π(ππ½ > π) = πΆπΌ ((1 + π½)/2), where πΆπΌ = ∞ (∫0 π₯−πΌ sin π₯ππ₯)−1 . We use the fact that, for any π½ ∈ [−1, 1], π (ππ½ > π) ≤ π (π1 > π) , ∀π > π πΌ (A.2) for some π πΌ > 0 (see Property 1.2.14 of [18] or Section 1.5 of [29]). Since limπ → ∞ ππΌ π(π1 > π) = πΆπΌ , there exists π∗πΌ > π πΌ such that (187) (πΆ ∨ πΆπ‘ ). This implies that π(π‘) = 0 for all π’ (π, π‘, π₯) = VπΎ (π, π‘, π₯) , ∀π > 0, where ππΌ∗ > 0 is a constant depending only on πΌ. × ∫ ∫ πΊ (π‘ − π , π₯, π¦) πΆπ‘σΈ O ππΌ π (|π| > π) ≤ ππΌ∗ ππΌ , π/π |π΅ (π‘, π₯)|π ≤ πΆππ πΎπ‘ 0 O (188) ππΌ π (π1 > π) < 2πΆπΌ , ∀π > π∗πΌ . (A.3) It follows that ππΌ π(ππ½ > π) < 2πΆπΌ for all π > π∗πΌ and π½ ∈ [−1, 1]. Clearly, for all π ∈ (0, π∗πΌ ] and π½ ∈ [−1, 1], ππΌ π(ππ½ > π) ≤ ππΌ ≤ (π∗πΌ )πΌ . Step 2. We now consider the general case. Since π/π has a ππΌ (1, π½, 0) distribution, by Step 1, it follows that ππΌ π(|π| > ππ) ≤ ππΌ∗ for any π > 0. The conclusion follows multiplying by ππΌ . In the proof of Theorem 13 and Lemma A.3 below, we use the following remark, due to Adam Jakubowski (personal communication). 20 International Journal of Stochastic Analysis Remark A.2. Let π be a random variable such that π(|π| > π) ≤ πΎπ−πΌ for all π > 0, for some πΎ > 0 and πΌ ∈ (0, 2). Then, for any π΄ > 0, Case 2 (πΌ > 1). Let π = ∑π≥1 ππ ππ 1{|ππ ππ |≤π} . Since πΈ(∑π≥1 ππ ππ ) = 0, σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ |πΈ (π)| = σ΅¨σ΅¨σ΅¨πΈ (∑ ππ ππ 1{|ππ ππ |>π} )σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ π≥1 σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ ≤ ∑ σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ πΈ (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ 1{|ππ ππ |>π} ) π΄ πΈ (|π| 1{|π|≤π΄} ) ≤ ∫ π (|π| > π‘) ππ‘ 0 ≤πΎ 1 π΄1−πΌ , 1−πΌ if πΌ < 1, π≥1 ∞ πΈ (|π| 1{|π|>π΄} ) ≤ ∫ π (|π| > π‘) ππ‘ + π΄π (|π| > π΄) π΄ ≤πΎ πΌ π΄1−πΌ , πΌ−1 if πΌ > 1, π΄ πΈ (π2 1{|π|≤π΄} ) ≤ 2 ∫ π‘π (|π| > π‘) ππ‘ 0 ≤πΎ 2 π΄2−πΌ , 2−πΌ ≤ (A.4) πΎπΌ 1−πΌ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ π ∑ σ΅¨σ΅¨ππ σ΅¨σ΅¨ , πΌ−1 π≥1 where we used Remark A.2 for the last inequality. From here, we infer that for any πΌ ∈ (0, 2) . The next result is a generalization of Lemma 2.1 of [24] to the case of nonsymmetric random variables. This result is used in the proof of Lemma 15 and Proposition 18. Lemma A.3. Let (ππ )π≥1 be independent random variables such that σ΅¨ σ΅¨ sup ππΌ π (σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ > π) ≤ πΎ, ∀π ≥ 1 (A.5) π>0 for some πΎ > 0 and πΌ ∈ (0, 2). If πΌ > 1, we assume that πΈ(ππ ) = 0 for all π, and, if πΌ = 1, we assume that ππ has a symmetric distribution for all π. Then for any sequence (ππ )π≥1 of real numbers, we have σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨πΌ sup ππΌ π (σ΅¨σ΅¨σ΅¨ ∑ ππ ππ σ΅¨σ΅¨σ΅¨ > π) ≤ ππΌ πΎ∑ σ΅¨σ΅¨σ΅¨ππ σ΅¨σ΅¨σ΅¨ , (A.6) σ΅¨σ΅¨π≥1 σ΅¨σ΅¨ π>0 π≥1 σ΅¨ σ΅¨ where ππΌ > 0 is a constant depending only on πΌ. Proof. We consider the intersection of the event on the lefthand side of (A.6) with the event {supπ≥1 |ππ ππ | > π} and its complement. Hence, σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨σ΅¨ σ΅¨ σ΅¨ π (σ΅¨σ΅¨σ΅¨ ∑ ππ ππ σ΅¨σ΅¨σ΅¨ > π) σ΅¨σ΅¨ σ΅¨σ΅¨π≥1 σ΅¨ σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ (A.7) σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨ ≤ ∑ π (σ΅¨σ΅¨ππ ππ σ΅¨σ΅¨ > π) + π (σ΅¨σ΅¨ ∑ ππ ππ 1{|ππ ππ |≤π} σ΅¨σ΅¨σ΅¨ > π) σ΅¨σ΅¨π≥1 σ΅¨σ΅¨ π≥1 σ΅¨ σ΅¨ =: πΌ + πΌπΌ. (A.9) |πΈ (π)| < π , 2 for any π > π πΌ , (A.10) where ππΌπΌ = 2πΎ(πΌ/(πΌ − 1)) ∑π≥1 |ππ |πΌ . By Chebyshev’s inequality, for any π > π πΌ , πΌπΌ = π (|π| > π) ≤ π (|π − πΈ (π)| > π − |πΈ (π)|) ≤ 4 4 πΈ|π − πΈ (π)|2 ≤ 2 ∑ ππ2 πΈ (ππ2 1{|ππ ππ |≤π} ) 2 π π π≥1 ≤ 8πΎ −πΌ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ π ∑ σ΅¨σ΅¨ππ σ΅¨σ΅¨ , 2−πΌ π≥1 (A.11) using Remark A.2 for the last inequality. On the other hand, if π ∈ (0, π πΌ ], πΌπΌ = π (|π| > π) ≤ 1 ≤ ππΌπΌ π−πΌ = 2πΎ πΌ −πΌ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ π ∑ σ΅¨σ΅¨ππ σ΅¨σ΅¨ . πΌ−1 π≥1 (A.12) Case 3 (πΌ = 1). Since ππ has a symmetric distribution, we can use the original argument of [24]. B. Fractional Power of the Laplacian Using (A.5), we have πΌ ≤ πΎπ−πΌ ∑π≥1 |ππ |πΌ . To treat πΌπΌ, we consider 3 cases. Let πΊ(π‘, π₯) be the fundamental solution of ππ’/ππ‘+(−Δ)πΎ π’ = 0 on Rπ , πΎ > 0. Case 1 (πΌ < 1). By Markov’s inequality and Remark A.2, we have Lemma B.1. For any π > 1, there exist some constants π1 , π2 > 0 depending on π, π, and πΎ such that πΌπΌ ≤ 1 −πΌ σ΅¨σ΅¨ σ΅¨σ΅¨πΌ 1 σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨ )≤πΎ π ∑ σ΅¨σ΅¨ππ σ΅¨σ΅¨ . ∑ σ΅¨π σ΅¨ πΈ (σ΅¨π σ΅¨ 1 π π≥1 σ΅¨ π σ΅¨ σ΅¨ π σ΅¨ {|ππ ππ |≤π} 1−πΌ π≥1 (A.8) π1 π‘−(π/2πΎ)(π−1) ≤ ∫ πΊ(π‘, π₯)π ππ₯ ≤ π2 π‘−(π/2πΎ)(π−1) . Rπ (B.1) International Journal of Stochastic Analysis 21 Proof. The upper bound is given by Lemma B.23 of [12]. For the lower bound, we use the scaling property of the functions (ππ‘,πΎ )π‘>0 . We have πΊ (π‘, π₯) =∫ ∞ π/2 (4ππ‘1/πΎ π) 0 ≥∫ ∞ 1 π/2 (4ππ‘1/πΎ π) 1 ≥ 1 1 π/2 (4ππ‘1/πΎ ) |π₯|2 exp (− 1/πΎ ) π1,πΎ (π) ππ 4π‘ π exp (− |π₯|2 ) π1,πΎ (π) ππ 4π‘1/πΎ π (B.2) → By the construction of the integral, πΌ(πππ )(π, O) πΌ(π)(π, O) a.s. for a subsequence {ππ }. It suffices to show that πΌ(ππ )(π, O) = 0 a.s. on π΄ for all π. But πΌ(ππ )(π, O) = πΌ(ππ )(π, Oπ ) and Oπ is bounded. Step 2. We show that the proof can be reduced to the case of a bounded processes. For this, let ππ (π‘, π₯) = π(π‘, π₯)1{|π(π‘,π₯)|≤π} . Clearly, ππ ∈ LπΌ is bounded and satisfies (C.1) for all π. By the dominated convergence theorem, [ππ − π]πΌ → 0, and hence πΌ(πππ )(π, O) → πΌ(π)(π, O) a.s. for a subsequence {ππ }. It suffices to show that πΌ(ππ )(π, O) = 0 a.s. on π΄ for all π. Step 3. We show that the proof can be reduced to the case of bounded continuous processes. Assume that π ∈ LπΌ is bounded and satisfies (C.1). For any π‘ > 0 and π₯ ∈ Rπ , we define |π₯|2 exp (− 1/πΎ ) πΆπ,πΎ 4π‘ ∞ with πΆπ,πΎ := ∫ π−π/2 π1,πΎ (π) ππ < ∞, 1 ππ (π‘, π₯) = ππ+1 ∫ and hence π‘ ∫ (π‘−1/π)∨0 (π₯−1/π,π₯]∩O σΈ π‘−ππ/2πΎ ∫ πΊ(π‘, π₯)π ππ₯ ≥ ππ,πΎ,π Rπ × ∫ exp (− Rπ π|π₯|2 ) ππ₯ = ππ,π,πΎ π‘−(π/2πΎ)(π−1) . 4π‘1/πΎ (B.3) where (π, π] = {π¦ ∈ Rπ ; ππ < π¦π ≤ ππ for all π = 1, . . . , π}. Clearly, ππ is bounded and satisfies (C.1). We prove that ππ ∈ LπΌ . Since ππ is bounded, [ππ ]πΌ < ∞. To prove that ππ is predictable, we consider π‘ πΉ (π‘, π₯) = ∫ ∫ 0 C. A Local Property of the Integral The following result is the analogue of Proposition 8.11 of [12]. Proposition C.1. Let π > 0 and O ⊂ Rπ be a Borel set. Let π = {π(π‘, π₯); π‘ ≥ 0, π₯ ∈ Rπ } be a predictable process such that π ∈ LπΌ if πΌ < 1, or π ∈ Lπ for some π ∈ (πΌ, 2] if πΌ > 1. If O is unbounded, assume in addition that π satisfies (83) if πΌ < 1, or π satisfies (116) for some π ∈ (πΌ, 2), if πΌ > 1. Suppose that there exists an event π΄ ∈ Fπ such that π (π, π‘, π₯) = 0, ∀π ∈ π΄, π‘ ∈ [0, π] , π₯ ∈ O. (C.1) Then for any πΎ > 0, πΌ(π)(π, O) = πΌπΎ (π)(π, O) = 0 a.s. on π΄. Proof. We only prove the result for πΌ(π), with the proof for πΌπΎ (π) being the same. Moreover, we include only the argument for πΌ < 1; the case πΌ > 1 is similar. The idea is to reduce the argument to the case when π is a simple process, as in the proof Proposition of 8.11 of [12]. Step 1. We show that the proof can be reduced to the case of a bounded set O. Let ππ (π‘, π₯) = π(π‘, π₯)1Oπ (π₯) where Oπ = O ∩ πΈπ and (πΈπ )π is an increasing sequence of sets in Bπ (Rπ ) such that βπ πΈπ = Rπ . Then ππ ∈ LπΌ satisfies (C.1). By the dominated convergence theorem, π σ΅¨πΌ σ΅¨ πΈ ∫ ∫ σ΅¨σ΅¨σ΅¨ππ (π‘, π₯) − π (π‘, π₯)σ΅¨σ΅¨σ΅¨ σ³¨→ 0. 0 O (C.2) π (π , π¦) ππ¦ ππ , (C.3) (0,π₯]∩O π (π , π¦) ππ¦ ππ . (C.4) Since π is predictable, it is progressively measurable; that is, for any π‘ > 0, the map (π, π , π₯) σ³¨→ π(π, π , π₯) from Ω × [0, π‘] × Rπ to R is Fπ‘ × B([0, π‘]) × B(Rπ )-measurable. Hence, πΉ(π‘, ⋅) is Fπ‘ × B(Rπ )-measurable for any π‘ > 0. Since the map π‘ σ³¨→ πΉ(π, π‘, π₯) is left continuous for any π ∈ Ω, π₯ ∈ Rπ , it follows that πΉ is predictable, being in the class C defined in Remark 11. Hence, ππ is predictable, being a sum of 2π+1 terms involving πΉ. Since πΉ is continuous in (π‘, π₯), ππ is continuous in (π‘, π₯). By Lebesgue differentiation theorem in Rπ+1 , ππ (π, π‘, π₯) → π(π, π‘, π₯) for any π ∈ Ω, π‘ > 0, and π₯ ∈ O. By the bounded convergence theorem, [ππ − π]πΌ → 0. Hence, πΌ(πππ )(π, O) → πΌ(π)(π, O) a.s. for a subsequence {ππ }. It suffices to show that πΌ(ππ )(π, O) = 0 a.s. on π΄ for all π. Step 4. Assume that π ∈ LπΌ is bounded, continuous, and satisfies (C.1). Let (ππ(π) )π=1,...,ππ be a partition of O in Borel sets with Lebesgue measure smaller than 1/π. Let π₯ππ ∈ ππ(π) be arbitrary. Define π−1 ππ ππ (π‘, π₯) = ∑ ∑ π ( π=0 π=1 ππ π ,π₯ )1 (π‘) 1π(π) (π₯) . π π π (ππ/π,(π+1)π/π] (C.5) Since π is continuous in (π‘, π₯), ππ (π‘, π₯) → π(π‘, π₯). By the bounded convergence theorem, [ππ − π]πΌ → 0, and hence 22 International Journal of Stochastic Analysis πΌ(πππ )(π, O) → πΌ(π)(π, O) a.s. for a subsequence {ππ }. Since on the event π΄, πΌ (ππ ) (π, O) π−1 ππ = ∑ ∑π ( π=0 π=1 ππ π ππ (π + 1) π , π₯π ) π (( , ] × ππ(π) ) = 0, π π π (C.6) it follows that πΌ(π)(π, O) = 0 a.s. on π΄. Conflict of Interests The author declares that there is no conflict of interests regarding the publication of this paper. Acknowledgments The author is grateful to Robert Dalang for suggesting this problem. This research is supported by a grant from the Natural Sciences and Engineering Research Council of Canada. References [1] E. GineΜ and M. B. Marcus, “The central limit theorem for stochastic integrals with respect to LeΜvy processes,” The Annals of Probability, vol. 11, no. 1, pp. 58–77, 1983. [2] K. ItoΜ, “Stochastic integral,” Proceedings of the Imperial Academy, vol. 20, pp. 519–524, 1944. [3] K. ItoΜ, “On a stochastic integral equation,” Proceedings of the Japan Academy, vol. 22, no. 2, pp. 32–35, 1946. [4] G. Kallianpur, J. Xiong, G. Hardy, and S. Ramasubramanian, “The existence and uniqueness of solutions of nuclear spacevalued stochastic differential equations driven by Poisson random measures,” Stochastics and Stochastics Reports, vol. 50, no. 1-2, pp. 85–122, 1994. [5] S. Albeverio, J. L. Wu, and T. S. Zhang, “Parabolic SPDEs driven by Poisson white noise,” Stochastic Processes and their Applications, vol. 74, no. 1, pp. 21–36, 1998. [6] E. S. L. BieΜ, “EΜtude d’une EDPS conduite par un bruit poissonnien,” Probability Theory and Related Fields, vol. 111, no. 2, pp. 287–321, 1998. [7] D. Applebaum and J. L. Wu, “Stochastic partial differential equations driven by LeΜvy space-time white noise,” Random Operators and Stochastic Equations, vol. 8, no. 3, pp. 245–259, 2000. [8] C. Mueller, “The heat equation with LeΜvy noise,” Stochastic Processes and Their Applications, vol. 74, no. 1, pp. 67–82, 1998. [9] L. Mytnik, “Stochastic partial differential equation driven by stable noise,” Probability Theory and Related Fields, vol. 123, no. 2, pp. 157–201, 2002. [10] D. A. Dawson, “Infinitely divisible random measures and superprocesses,” in Stochastic Analysis and Related Topics, H. KoΜrezliogΜlu and A. UΜstuΜnel, Eds., vol. 31, pp. 1–129, BirkhaΜuser Boston, Boston, Mass, USA, 1992. [11] C. Mueller, L. Mytnik, and A. Stan, “The heat equation with time-independent multiplicative stable LeΜvy noise,” Stochastic Processes and Their Applications, vol. 116, no. 1, pp. 70–100, 2006. [12] S. Peszat and J. Zabczyk, Stochastic Partial Differential Equations with LeΜvy Noise, vol. 113 of Encyclopedia of Mathematics and Its Applications, Cambridge University Press, Cambridge, UK, 2007. [13] S. Peszat and J. Zabczyk, “Stochastic heat and wave equations driven by an impulsive noise,” in Stochastic Partial Differential Equations and Applications VII, G. Da Prato and L. Tubaro, Eds., vol. 245, pp. 229–242, Chapman & Hall/CRC, Boca Raton, Fla, USA, 2006. [14] S. Albeverio, V. Mandrekar, and B. RuΜdiger, “Existence of mild solutions for stochastic differential equations and semilinear equations with non-Gaussian LeΜvy noise,” Stochastic Processes and Their Applications, vol. 119, no. 3, pp. 835–863, 2009. [15] C. Marinelli and M. RoΜckner, “Well-posedness and asymptotic behavior for stochastic reaction-diffusion equations with multiplicative Poisson noise,” Electronic Journal of Probability, vol. 15, pp. 1528–1555, 2010. [16] E. Priola and J. Zabczyk, “Structural properties of semilinear SPDEs driven by cylindrical stable processes,” Probability Theory and Related Fields, vol. 149, no. 1-2, pp. 97–137, 2011. [17] B. Øksendal, “Stochastic partial differential equations driven by multi-parameter white noise of LeΜvy processes,” Quarterly of Applied Mathematics, vol. 66, no. 3, pp. 521–537, 2008. [18] G. Samorodnitsky and M. S. Taqqu, Stable Non-Gaussian Random Processes, Chapman & Hall, New York, NY, USA, 1994. [19] W. Feller, An Introduction to Probability Theory and Its Applications, vol. 2, John Wiley & Sons, New York, NY, USA, 2nd edition, 1971. [20] B. S. Rajput and J. RosinΜski, “Spectral representations of infinitely divisible processes,” Probability Theory and Related Fields, vol. 82, no. 3, pp. 451–487, 1989. [21] J. B. Walsh, “An introduction to stochastic partial differential equations,” in EΜcole d’EΜteΜ de ProbabiliteΜs de Saint-Flour XIV, vol. 1180 of Lecture Notes in Mathematics, pp. 265–439, Springer, Berlin, Germany, 1986. [22] S. I. Resnick, Heavy-Tail Phenomena: Probabilistic and Statistical Modelling, Springer Series in Operations Research and Financial Engineering, Springer, New York, NY, USA, 2007. [23] P. Billingsley, Probability and Measure, Wiley Series in Probability and Mathematical Statistics, John Wiley & Sons, New York, NY, USA, 3rd edition, 1995. [24] E. GineΜ and M. B. Marcus, “Some results on the domain of attraction of stable measures on πΆ(πΎ),” Probability and Mathematical Statistics, vol. 2, no. 2, pp. 125–147, 1982. [25] A. Truman and J. L. Wu, “Fractal Burgers’ equation driven by LeΜvy noise,” in Stochastic Partial Differential Equations and Applications VII, G. Da Prato and L. Tubaro, Eds., vol. 245, pp. 295–310, Chapman & Hall/CRC, Boca Raton, Fla, USA, 2006. [26] P. Azerad and M. Mellouk, “On a stochastic partial differential equation with non-local diffusion,” Potential Analysis, vol. 27, no. 2, pp. 183–197, 2007. [27] R. C. Dalang, “Extending the martingale measure stochastic integral with applications to spatially homogeneous s.p.d.e.’s,” Electronic Journal of Probability, vol. 4, no. 6, 29 pages, 1999, Erratum in Electronic Journal of Probability, vol. 6, 5 pages, 2001. [28] C. Dellacherie and P. A. Meyer, ProbabiliteΜs et Potentiel, vol. 1, Hermann, Paris, France, 1975. [29] J. P. Nolan, Stable Distributions: Models for Heavy Tailed Data, chapter 1, BirkhaΜauser, Boston, Mass, USA, 2013, http://academic2.american.edu/∼jpnolan/stable/chap1.pdf. Advances in Operations Research Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Advances in Decision Sciences Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Mathematical Problems in Engineering Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Algebra Hindawi Publishing Corporation http://www.hindawi.com Probability and Statistics Volume 2014 The Scientific World Journal Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Differential Equations Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Volume 2014 Submit your manuscripts at http://www.hindawi.com International Journal of Advances in Combinatorics Hindawi Publishing Corporation http://www.hindawi.com Mathematical Physics Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Complex Analysis Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Mathematics and Mathematical Sciences Journal of Hindawi Publishing Corporation http://www.hindawi.com Stochastic Analysis Abstract and Applied Analysis Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com International Journal of Mathematics Volume 2014 Volume 2014 Discrete Dynamics in Nature and Society Volume 2014 Volume 2014 Journal of Journal of Discrete Mathematics Journal of Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Applied Mathematics Journal of Function Spaces Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Optimization Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014