-strong convergence orders of fully discrete schemes for the SPDE driven by Lévy noise
Abstract.
It is well known that for a stochastic differential equation driven by Lévy noise, the temporal Hölder continuity in sense of the exact solution does not exceed . This leads to that the -strong convergence order of a numerical scheme will vanish as increases to infinity if the temporal Hölder continuity of the solution process is directly used. A natural question arises: can one obtain the -strong convergence order that does not depend on ? In this paper, we provide a positive answer for fully discrete schemes of the stochastic partial differential equation (SPDE) driven by Lévy noise. Two cases are considered: the first is the linear multiplicative Poisson noise with and the second is the additive Poisson noise with , where is the Lévy measure and is the mark set. For the first case, we present a strategy by employing the jump-adapted time discretization, while for the second case, we introduce the approach based on the recently obtained Lê’s quantitative John–Nirenberg inequality. We show that proposed schemes converge in sense with orders almost in both space and time for all , which contributes novel results in the numerical analysis of the SPDE driven by Lévy noise.
Key words and phrases:
Fully discrete scheme -strong convergence order Lévy noise Stochastic partial differential equation1. Introduction
Stochastic differential equations (SDEs) with Lévy noise are widely used to model sudden events and irregular changes in stochastic phenomenon, such as financial market crashes, abrupt phase transitions, and neural spiking patterns; see e.g. [3, 9, 12] and references therein. In numerical studies of SDEs, the -strong convergence analysis of numerical methods has received much attention, as it assesses the accuracy of simulating the paths of the underlying solution process. The -strong convergence analysis for SDEs with Lévy noise is different from that of the Gaussian noise case (see e.g. [5, 10]). For instance, there is an extra term appearing in the Burkholder–Davis–Gundy inequality of the stochastic integral with respect to the compensated Poisson measure compared with that in the Gaussian noise case. As a result, the temporal Hölder continuity in sense of the exact solution to an SDE with Lévy noise does not exceed , resulting in the deterioration of the -strong convergence order of a numerical scheme for large if one uses the temporal Hölder continuity directly. A natural question arises:
-
Q:
Can one obtain the -strong convergence order of a numerical scheme for the SDE with Lévy noise that does not depend on ?
A remarkable progress for this question has been made recently for the case of the stochastic ordinary differential equation (SODE) with Lévy noise: authors in [1] studied the -strong convergence order of the Euler–Maruyama scheme for a multidimensional SODE with irregular Hölder drift, driven by additive Lévy process with exponent , and the obtained convergence order does not depend on . For the case of SPDEs, to the best of our knowledge, this question is still open so far. The aim of this paper is to provide a positive answer for fully discrete schemes of the SPDE driven by Lévy noise.
We consider the following SPDE driven by Lévy noise:
(1) |
where and is the Laplacian with homogeneous Dirichlet boundary condition. Here with usual inner product and norm , and is the mark set. The process is an -valued Gaussian space-time white noise defined on a complete filtered probability space that satisfies usual conditions. Let stand for a compensated Poisson random measure, where is the Poisson random measure and is a Lévy measure on Borel -algebra satisfying and Here and are supposed to be independent. Precise assumptions on coefficients and will be given in Section 2.1. For the convergence analysis of numerical methods for (1), we refer interested readers to [2] for the -strong convergence and to [6] for the -strong convergence with .
We apply the spectral Galerkin method to approximate (1) in space, and further use the Euler-type methods in time to obtain the fully discrete schemes. The corresponding mapping for the one-step approximation is given as follows:
where is the subscript corresponding to the time node (i.e., ), and we mention that the time step size may be path-dependent. The main contribution of our work is that, we show for the first time that proposed schemes converge in sense with orders almost in both space and time for all . Indeed, the -strong convergence analysis in time is more technical compared with that in space. This is related to the aforementioned issue yielding poor order for large . To overcome this order barrier, we present two different strategies for the cases of the linear multiplicative Poisson noise with and the additive Poisson noise with , respectively. More precisely,
-
(i)
Case of the linear multiplicative Poisson noise with . In this case, there are a.s. finitely many jumps. We introduce the jump-adapted time discretization by a superposition of finitely many jump times to a deterministic equidistant grid. Then the jump-adapted fully discrete scheme is constructed as follows:
(2) Here is an -adapted Poisson point process with the intensity measure . In the convergence analysis of this scheme, there are two critical steps to overcome the order barrier. First, this scheme evolves without jump between time grid points, which implies that the stochastic integral with respect to the compensated Poisson random measure in (2) can be transformed into the integral with respect to the intensity measure. Second, it requires establishing the th moment estimates on the terms involving the multiplicative noise or the jump process when accumulating the local errors into the global one.
-
(ii)
Case of the additive Poisson noise with . In this case, there may be a.s. infinitely many jumps, in which the jump-adapted strategy is invalid. We consider the fully discrete scheme with constant step size . In the -strong convergence analysis of this scheme, new strategy is required to deal with the terms involving the temporal Hölder continuity of the stochastic convolution with respect to the compensated Poisson random measure of (1). Our idea is to bound the -norm of these terms by moment estimates of the nested conditional “-norm” based on the recently obtained Lê’s quantitative John–Nirenberg inequality. To this end, we carefully analyze some essential a priori estimates including the nested conditional “-norms” of both the stochastic convolutions and the solution of the perturbed stochastic equation, which help us to obtain the rigorous bound of the nested conditional “-norm” and then the desired convergence orders.
The rest of the paper is organized as follows. The preliminaries, including notations, assumptions, and the well-posedness of the exact solution are given in Section 2.1. Section 2.2 is devoted to the introduction of the fully discrete schemes and their -strong convergence order results. In Section 3, we present the -strong convergence analysis of the fully discrete scheme for (1) in the case of the linear multiplicative Poisson noise with . In Section 4, we present the corresponding -strong convergence analysis for the case of the additive Poisson noise with . Proofs of some essential propositions used in the convergence analysis are given in Section 5.
2. Fully discrete schemes and main results
In this section, we first give some preliminaries, including notations, assumptions, the well-posedness of the exact solution and some useful inequalities. Second, we present the fully discrete schemes of (1) for two different cases of noises, and respectively show their -strong convergence orders, as the main results of this paper.
2.1. Preliminaries
Throughout this paper, we let be a generic constant that may vary from one place to another. More specific constants which depend on certain parameters are numbered as . We use to denote the set of positive natural numbers and let be arbitrarily small parameters. Denote . The space of bounded linear operators on is denoted by with the operator norm . The subspace denotes the set of Hilbert–Schmidt operators on , with norm denoted by . Let , be the Sobolev space generated by the fractional power of , endowed with the inner product and the norm for all . We use the notation to denote the space of random variables satisfying for .
It is known that there is a sequence of real numbers and an orthonormal basis of with , , such that . In addition, operator generates the semigroup on . For all , is a bounded self-adjoint linear operator on , with . It is clear that, for all ,
Moreover, the semigroup satisfies the following properties:
(3) | ||||
(4) |
Here we impose some assumptions on coefficients and the initial value of (1), which will be used throughout this paper.
Assumption 1 (Nonlinearity).
The measurable mapping satisfies that , with some constant . In addition, is differentiable and there are constants such that
Remark 2.1.
The conditions in Assumption 1 are satisfied if is the Nemytskii operator defined by where is continuously differentiable with bounded first order derivative. Indeed, the mapping is then differentiable, and for all and , we have This implies with some constant In addition, for all and , we obtain
where we used [11, Lemma 4.4] in the last step.
Assumption 2 (Jump coefficient).
The coefficient is defined as with mapping being bounded by constant and with some mapping . In addition, for each there exists a constant such that
Assumption 3 (Initial value).
Let be an -measurable random variable, and for all .
We introduce the maximal inequality for the Lévy-type stochastic convolution as follows, see e.g. [8, Proposition 3.3] for details.
Lemma 2.2.
Let be a predictable process such that the expectation on the right hand side of the inequality below is finite. Then for all , there exists a constant such that
Under assumptions given above, the mild solution of (1) uniquely exists and the regularity estimate of the solution can be obtained, which are stated in the following proposition. The proof of the well-posedness is similar to that of [13, Theorem 2.1] and the one of the strong Markov property is similar to that of [13, Lemma 5.2], which are omitted here. For the proof of regularity estimate, we postpone it to Appendix.
Proposition 2.3.
2.2. Fully discrete schemes and -strong convergence orders
In this subsection, we first apply the spectral Galerkin method in the spatial direction and the Euler-type method in the temporal direction to obtain the fully discrete schemes for (1). Then we present the main results of this paper, namely, -strong convergence orders of proposed schemes in both space and time.
For fixed , let be the orthogonal projection operator from onto . Applying the spectral Galerkin method to the spatial direction of (1) yields
(5) |
with the initial value where and . Further, in the temporal direction, we use the Euler-type method to obtain the one-step approximation:
(6) |
where with being time grid points. With the one-step approximation at hand, below, we present the fully discrete schemes for two cases: the first is the linear multiplicative Poisson noise with , and the second is the additive Poisson noise with . We remark that the choices of time step sizes are different in these two cases.
(i) Case of the linear multiplicative Poisson noise with . In this case, since there are a.s. finitely many jumps. The corresponding non-decreasing jump times are denoted by , where the random variable denotes the number of jumps. We introduce the jump-adapted time partition by a superposition of finitely many jump times to a deterministic equidistant grid. More precisely, we first give a deterministic partition of the interval :
where is a constant step size and . For each sample path, we then merge and the partition from jump times to form a new partition
where denotes the subscript corresponding to the time node and . Note that in the new partition , step size is path-dependent, and the maximal time step size of the resulting jump-adapted time partition is no larger than .
The jump-adapted fully discrete scheme is proposed as
with being given by (2.2), which is equivalent to
(7) |
due to the following relation:
(8) |
for . We now give the -strong convergence result of the scheme (7), whose proof is presented in Section 3.
Theorem 2.4.
(ii) Case of the additive Poisson noise with . In this case when , we can obtain the same -strong convergence orders as shown in Theorem 2.4 by applying the scheme (7). However, this jump-adapted numerical scheme is invalid when , since there may be a.s. infinitely many jumps. In order to obtain the -strong convergence orders in a wider case: , we consider the fully discrete scheme on the deterministic partition with constant step size . For simplicity, we write instead of time grid point and Then the fully discrete scheme based on the one-step approximation (2.2) is given as
(9) |
New strategy is proposed in the -strong convergence analysis of this scheme. We now give the -strong convergence result of the scheme (9) for the case of the additive Poisson noise () as follows, and its proof is presented in Section 4.
3. Proof of Theorem 2.4
In this section, we present the -strong convergence analysis of fully discrete scheme (7) for (1) driven by the linear multiplicative Poisson noise with finite Lévy measure. Note that the stochastic integral with respect to the compensated Poisson random measure in (7) can be transformed into the integral with respect to the intensity measure, as shown in (2.2). This transformation is crucial to overcome the order barrier caused by the th moment estimates of the stochastic integral with respect to the compensated Poisson random measure.
Proof of Theorem 2.4.
The proof is split into two steps, based on the error between the solutions of (1) and the semi-discrete scheme (2.2), and the error between the solutions of (2.2) and the fully discrete scheme (7).
Step 1. Show that for all , there exists a constant independent of such that
It follows from (1) and (2.2) that, for all ,
where is an identical mapping. Denote the stochastic convolution . Owing to properties of projection operator , assumptions on coefficients and , Lemma 2.2, and Proposition 2.3, we obtain
Applying the Grönwall inequality finishes the proof of Step 1.
Step 2. Show that for all , there exists a constant independent of and such that
On the partition , the process can be rewritten as
(11) |
On the time interval , it follows from (7) and (11) that
and
Denote and . Then it holds that
(12) |
Using the definition of gives that for
(13) |
where , , and
By iteratively applying the relation between and (i.e., (13)), the local error is accumulated into the global one, namely, we have
where we used and set . Taking -norm yields
To proceed, we need to establish the th moment estimates on terms respectively, which involve the multiplicative noise or the jump process.
Let and for any , and use to denote the subscript corresponding to time (i.e., ). For the term , we have
Note that random variables , , and the -algebra are independent, which together with the Lipschitz condition of , (12), and yields
According to Assumption 2 on coefficient , it holds that with , and . Set . From Poisson distribution it follows that for all ,
(14) |
Hence we arrive at
For the term by Assumption 1, we have
To estimate for and , we note that
where we used the Burkholder–Davis–Gundy inequality (see e.g. [4, Theorem 4.36]) and the Hölder inequality. This, combining Proposition 2.3 shows
Hence, it follows from (3) that .
Combining estimates of terms above gives
which yields by using the Grönwall inequality. According to (12) and , we obtain
Combining Steps 1-2 finishes the proof. ∎
4. Proof of Theorem 2.5.
In this section, we present the proof of the -strong convergence orders of the fully discrete scheme (9) for (1) driven by the additive Poisson noise with , that is Theorem 2.5. The proof relies on estimates of the stochastic convolutions, the recently obtained Lê’s quantitative John–Nirenberg inequality and some a priori estimates on the nested conditional “-norms” of both stochastic convolutions and the solution of the perturbed stochastic equation.
For and , denote stochastic convolutions and with . Let , which satisfies the perturbed stochastic equation:
(15) |
with . It is known that satisfies
(16) |
We list properties of the stochastic convolution with respect to the Wiener process as follows, whose proofs are postponed to Appendix.
Lemma 4.1.
(i) For and , we have
(17) |
(ii) For and , there exists a constant independent of such that
(18) |
We then introduce the recently obtained Lê’s quantitative John–Nirenberg inequality, which provides a useful way to bound the -norm of a stochastic process by moment estimates of the nested conditional “-norm”. The conditional expectation given is denoted by .
Lemma 4.2.
[7, Theorem 1.1] Let be a metric space and be a right continuous with left limits and adapted integrable stochastic process. Then for every and , there exists a constant such that
Let denote . For a random variable , a sub--algebra , and , we set . For a measurable mapping , we set for any . Let , that is, is the smallest grid point strictly bigger than .
The following propositions give some a priori estimates including nested conditional “-norms” of both stochastic convolutions and the solution of the perturbed stochastic equation. The proofs are postponed to Section 5.
Proposition 4.3.
For each and any time grid point , , we have
(19) |
Proposition 4.4.
For , there exists a constant such that
(20) | |||
(21) | |||
(22) | |||
(23) |
Proposition 4.5.
(i) For and , there exists a constant such that
(24) |
(ii) For , there exists a constant such that
(25) |
With these preliminaries, we present the proof of Theorem 2.5.
Proof of Theorem 2.5.
The error between the exact solution and the semi-discrete numerical solution (i.e., ) can be proved by Step 1 in the proof of Theorem 2.4 and thus is omitted. Hence it suffices to estimate the error .
For the term , by Assumption 1, we have
Combining (24) and the same arguments as in the proof of Proposition 2.3 leads to
Similarly, by (18), we have
For the term , we claim that
(26) |
Estimate of term includes dealing with the temporal Hölder continuity of the stochastic convolution with respect to the compensated Poisson random measure to overcome the order barrier. To this end we aim to apply Lemma 4.2 to estimate the -norm of . For , define an integrable -adapted stochastic process:
with . By the globally Lipschitz condition of , we have
The Kolmogorov continuity theorem implies that is a.s. continuous. Thus by Lemma 4.2, we arrive at
which yields that
(27) |
Now we turn to estimating the term on the right hand side of the inequality in (27). We first show that, there exists a real-valued adapted stochastic process such that
(28) |
Indeed, using Assumption 1 and the conditional Hölder inequality gives
In the case of , i.e., , it follows from (19) and the Young inequality that almost surely with
In the case that is not a grid point, i.e., , we note that when ,
When , we observe that , , and
Hence, we derive that almost surely with
Therefore (28) holds with . To further estimate the term on the right hand side of (27), we use Propositions 4.3–4.5 to obtain
Similarly, we can derive that As a result, (26) is proved.
Combining estimates of terms , , and , and applying the Grönwall inequality complete the proof of the theorem. ∎
5. Proofs of some essential propositions
In this section, we give proofs of Propositions 4.3–4.5. Note that the process is progressively measurable on product space . For any interval , due to the independent increments property of the compensated Poisson measure, it holds that
-
(P1)
is -independent.
This property will be frequently utilized in the following proofs.
Proof of Proposition 4.3.
For any time grid point , , the property (P1) and the assumption on lead to
The proof is completed. ∎
Proof of Proposition 4.4.
-
(iii)
The proof of (22). Based on the property (P1), we have
And using Lemma 2.2 again, we obtain
(iv) The proof of (23). By using (17) in Lemma 4.1, the proof of (23) is similar to that of (22), which is omitted.
Combining (i)–(iv), the proof is completed. ∎
Proof of Proposition 4.5.
(i) Owing to the linear growth condition of , we derive that for any and ,
(ii) According to the estimates obtained in Propositions 4.3–4.4, we can derive that terms
are all bounded for with some constant .
Recall that satisfies the perturbed stochastic equation (15). Using the linear growth condition of , it implies that for all ,
Applying the Grönwall inequality leads to
In addition, for all ,
The Grönwall inequality leads to
Appendix.
Proof of Proposition 2.3.
For brevity, we set for all . Then satisfies
For all and all , we note that there exists a constant such that Using Assumption 1 and Lemma 2.2 leads to
When , owing to the moment estimates of , Assumption 2, and the Hölder inequality, we arrive at
Applying the Grönwall inequality yields that .
Then for , we have
Applying the Grönwall inequality yields the desired assertion. ∎
Proof of Lemma 4.1.
(i) Denote for all and . Based on the factorization method, we have expression with for ; see e.g. [4, Proposition 5.9, Theorem 5.10] for details. Note that for all ,
where we utilized the Hölder inequality with , in the last step. When , by the Hölder inequality and the Burkholder–Davis–Gundy inequality, we derive
which is finite if i.e., It means that . Hence, when positive parameters and are chosen such that and , we derive that Therefore, for all with large number , we obtain For , we have
References
- [1] O. Butkovsky, K. Dareiotis, and M. Gerencsér. Strong rate of convergence of the Euler scheme for SDEs with irregular drift driven by Lévy noise, arXiv preprint arXiv:2204.12926, 2024.
- [2] C. Chen, T. Dang, and J. Hong. On structure preservation for fully discrete finite difference schemes of stochastic heat equation with Lévy space-time white noise, arXiv preprint arXiv:2409.14064, 2024.
- [3] R. Cont and P. Tankov. Financial modelling with jump processes. Chapman & Hall/CRC Financial Mathematics Series. Chapman & Hall/CRC, Boca Raton, FL, 2004.
- [4] G. Da Prato and J. Zabczyk. Stochastic equations in infinite dimensions, volume 152 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, second edition, 2014.
- [5] T. Dunst, E. Hausenblas, and A. Prohl. Approximate Euler method for parabolic stochastic partial differential equations driven by space-time Lévy noise. SIAM J. Numer. Anal., 50(6):2873–2896, 2012.
- [6] E. Hausenblas. Finite element approximation of stochastic partial differential equations driven by Poisson random measures of jump type. SIAM J. Numer. Anal., 46(1):437–471, 2007.
- [7] K. Lê. Maximal inequalities and weighted BMO processes, arXiv preprint arXiv:2211.15550, 2022.
- [8] C. Marinelli, C. Prévôt, and M. Röckner. Regular dependence on initial data for stochastic evolution equations with multiplicative Poisson noise. J. Funct. Anal., 258(2):616–649, 2010.
- [9] S. Peszat and J. Zabczyk. Stochastic partial differential equations with Lévy noise: An evolution equation approach, volume 113 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2007.
- [10] P. Przybył owicz, M. Sobieraj, and Ł. Stepień. Efficient approximation of SDEs driven by countably dimensional Wiener process and Poisson random measure. SIAM J. Numer. Anal., 60(2):824–855, 2022.
- [11] X. Wang, S. Gan, and J. Tang. Higher order strong approximations of semilinear stochastic wave equation with additive space-time white noise. SIAM J. Sci. Comput., 36(6):A2611–A2632, 2014.
- [12] W. A. Woyczyński. Lévy processes in the physical sciences. In Lévy processes, pages 241–266. Birkhäuser Boston, Boston, MA, 2001.
- [13] B. Xie. Uniqueness of invariant measures of infinite dimensional stochastic differential equations driven by Lévy noises. Potential Anal., 36(1):35–66, 2012.