Convergence analysis of a finite difference method for stochastic Cahn–Hilliard equation
Abstract.
This paper presents the convergence analysis of the spatial finite difference method (FDM) for the stochastic Cahn–Hilliard equation with Lipschitz nonlinearity and multiplicative noise. Based on fine estimates of the discrete Green function, we prove that both the spatial semi-discrete numerical solution and its Malliavin derivative have strong convergence order . Further, by showing the negative moment estimates of the exact solution, we obtain that the density of the spatial semi-discrete numerical solution converges in to the exact one. Finally, we apply an exponential Euler method to discretize the spatial semi-discrete numerical solution in time and show that the temporal strong convergence order is nearly , where a difficulty we overcome is to derive the optimal Hölder continuity of the spatial semi-discrete numerical solution.
Key words and phrases:
strong convergence rate, finite difference method, exponential Euler method, stochastic Cahn–Hilliard equation2010 Mathematics Subject Classification:
65C30, 60H35, 60H15, 60H071. Introduction
Consider the following stochastic Cahn–Hilliard equation
(1.1) |
with initial condition and homogeneous Dirichlet boundary conditions (DBCs) on . Here, , , and is a Brownian sheet defined on some probability space . Assume that is a deterministic continuous function, and is bounded and globally Lipschitz continuous. Eq. (1.1) is a well-known phenomenological model to describe the complicated phase separation. In the original form, is the derivative of the homogeneous free energy which contains a logarithmic term and in some cases can be approximated by an even-degree polynomial with a positive dominant coefficient [4], for example, . In this case, the truncation technique is usually used to localize (1.1) such that the sequence given by
(1.2) |
can approximate in some sense (see e.g., [4, 8]), where satisfies the global Lipschitz condition. Hence, an effective numerical method applied to Eq. (1.2) is expected to approximate Eq. (1.1) well. With this consideration, the present work investigates numerical methods for stochastic Cahn–Hilliard equations with Lipschitz nonlinearity (i.e., is globally Lipschitz continuous), which includes the linearized Cahn–Hilliard equation ().
The existing results on the numerical methods for stochastic Cahn–Hilliard equations mainly focus on the strong convergence analysis. Without being too exhaustive, we mention [5, 21] on the finite element approximation for the case of . For the case of polynomial nonlinearity and additive noise, [19] and [14] respectively obtain the strong convergence of a spatial semi-discretization and a full discretization; [10, 24] establish the strong convergence rates of full discretizations based on the finite element method and spectral Galerkin method in space, respectively. Concerning the case of multiplicative noise, [8] presents the sharp strong convergence rate for a full discretization by using the spectral Galerkin method in space. In addition to the strong convergence analysis, the convergence analysis of densities of numerical solutions is also meaningful, which provides a theoretical foundation to approximate the density of the exact solution of the original system by means of numerical methods. There have been plenty results on density convergence of numerical solutions for various stochastic systems (see e.g., [3, 6, 9, 16, 18, 22]), of which we have not yet found relevant results for stochastic Cahn-Hilliard equations. The present paper aims to approximate the density of the exact solution of Eq. (1.1) via a spatial finite difference method (FDM) and present the strong convergence rate and density convergence of the associated numerical solution.
The spatial FDM has been employed to numerically solve, for instance, stochastic heat equations [15, 11, 1] and stochastic wave equations [7]. First, we give subtle error estimates between the discrete Green function of the spatial FDM and the exact one. Then under the globally Lipschitz condition on , we obtain the strong convergence order for the spatial semi-discrete numerical solution based on the FDM, with being the spatial stepsize. Further, it is shown that the Malliavin derivative of has the strong convergence order as well. Combining the above results with the negative moment estimates of the exact solution, we deduce that the spatial semi-discrete numerical solution admits a density, which converges in to the density of the exact solution.
For more effective computation, we further discretize via an exponential Euler method in time and obtain the full discretization , where denotes the temporal stepsize. As an explicit method, the exponential Euler method is more computationally efficient than the implicit method and does not suffer from the CFL condition. By investigating the temporal Hölder continuity of the spatial semi-discrete numerical solution, we attain the strong convergence rate of the proposed method for Eq. (1.1) with Lipschitz nonlinearity, namely
(1.3) |
where . The spatial convergence order and temporal convergence order nearly in (1.3) are optimal in the sense that they coincide with the mean-square spatial and temporal Hölder continuity exponents of the exact solution, respectively. On the basis of (1.3), a localized argument leads to an convergence order localized on a set of arbitrarily large probability for Eq. (1.1) with polynomial nonlinearity. With the independent interest, when is a polynomial of degree with a positive dominant coefficient, we also establish the Hölder continuity and the uniform moment estimate of the exact solution. These are prepared for the density convergence analysis of the numerical solution of the spatial FDM for Eq. (1.1) with polynomial nonlinearity in our future work.
The rest of this paper is organized as follows. Section 2 gives the Hölder continuity and the uniform moment estimate of the exact solution. Then we introduce the spatial FDM and study its strong convergence order in Section 3. Section 4 presents the density convergence of the spatial semi-discrete numerical solution. In Section 5, we further apply an exponential Euler method to obtain a full discretization and obtain its strong convergence order.
2. Preliminaries
Let be the space of -Hölder continuous functions on if , and be the space of times continuously differentiable functions on if . For , we denote by and the Euclidean norm and inner product of , respectively. For , we denote by the usual norm of the space . For and a Banach space , let be the space of -valued random variables with bounded th moment, endowed with the norm Especially, we write for short. Hereafter, we use to denote a generic positive constant that may change from one place to another and depend on several parameters but never on the space and time stepsizes. Without illustrated, the supremum with respect to (respectively, and ) is denoted by (respectively, and ). In this section, we present the regularity estimate of the exact solution of Eq. (1.1) under Assumption 1 or 2.
Assumption 1.
satisfies the globally Lipschitz condition, i.e., there is such that
Assumption 2.
is a polynomial of degree with a positive dominant coefficient, i.e., with .
The physical importance of the Dirichlet problem is pointed out to us by M. E. Gurtin: it governs the propagation of a solidification front into an ambient medium which is at rest relative to the front [12]; see [13, 8] and references therein for the study of Cahn–Hilliard equation with DBCs. In this case, the Green function associated to is given by , , where , . It is known that forms an orthonormal basis of . Denote , Similar to [4, Lemma 1.2], there exist such that
(2.1) | |||
(2.2) |
The well-posedness of the stochastic Cahn–Hilliard equation under Assumption 2 with NBCs has been established in [4]. Since the Green function with DBCs and NBCs share similar properties, the existence and uniqueness of the solution to Eq. (1.1) under Assumption 2 can be obtained in an almost same way, and we present an outline of the idea here. For , let be an even smooth cut-off function satisfying
(2.3) |
and , . Consider a sequence of SPDEs
(2.4) |
with DBCs and for some . Define the stopping times
Using the uniqueness of the solution of Eq. (2.4), it is concluded from the local property of stochastic integrals that for , for , so that a process can be defined by for . Set . Then is the unique solution of (1.1) on the interval . Further, are -adapted stochastic processes such that for ,
(2.5) |
(see the second inequality in P794 of [4]). Based on (2.5), a.s. (see [4, (2.36)]), and thus under Assumption 2, Eq. (1.1) admits a global solution, i.e.,
It follows from Fatou’s lemma and (2.5) that for ,
(2.6) |
which also holds for any and in view of the Hölder inequality and the continuity of . Besides, under Assumption 1, a standard Picard approximation argument shows that Eq. (1.1) admits a unique solution satisfying (2.6).
Similar to [4, Lemma 1.8], we have the following regularity of .
Lemma 2.1.
For , there exists such that for and ,
Based on (2.6) and Lemma 2.1, we present the Hölder continuity of the exact solution. Under either Assumption 1 or Assumption 2, there exists some constant such that
(2.7) |
Lemma 2.2.
Proof.
We first prove
(2.9) |
To this end, we write with
(2.10) | |||
(2.11) | |||
(2.12) |
Let us state a useful property in [4, Lemma 1.6]. For any , , and , the linear operator defined by
is a mapping from to with
(2.13) |
It follows from (2.6) and (2.7) that
(2.14) |
Applying (2.13) with , and leads to
which combined with the Minkowski inequality and (2.14) implies
(2.15) |
Since is bounded, the Burkholder inequality [17, Theorem B.1] and (2.1) yield
(2.16) |
for . In addition, (2.1) implies for , which together with (2.15) and (2.16) completes the proof of (2.9).
Without loss of generality, assume that . Notice that with
and with
where the explicit dependence of , , , , , on is dropped for simplicity. Using [4, Lemma 2.3] and the assumption , we get By the Burkholder inequality, the boundedness of , and Lemma 2.1, we obtain For any ,
(2.17) |
By the orthogonality of in , (2.17), and for
with . Choosing and using the Cauchy–Schwarz inequality,
(2.18) |
Taking advantage of (2) and (2.9), we obtain
Similarly, the orthogonality of in , the Cauchy–Schwarz inequality, and (2.17) yield that for ,
(2.19) |
which together with (2.9) implies that for ,
Observe that for any ,
(2.20) |
Then the orthogonality of in , the Cauchy–Schwarz inequality, (2.20) and (2.17) imply that for and with , it holds that
(2.21) |
Hence, it follows from (2.9) that with . Finally, combining the above estimates, we obtain (2.8), which completes the proof. ∎
Lemma 2.3.
Under the same conditions of Lemma 2.2, for any ,
3. Finite difference method
In this section, we introduce the spatial FDM method for Eq. (1.1) and derive its strong convergence rate. Given a function on the mesh , we define the difference operator
for , where . The compatibility conditions and are direct results of DBCs and the initial condition. One can approximate via , where and
(3.1) |
for and , under the boundary conditions
for . For , we use the polygonal interpolation
To solve (3), we introduce
with and for , where the explicit dependence of and on is omitted. Let
Then (3) can be rewritten into an -dimensional SDE
(3.2) |
with the initial condition and the coefficients
Under Assumption 1, (3.2) admits a unique strong solution which satisfies
(3.3) |
For , given by
(3.4) |
is an eigenvector of associated with the eigenvalue , where
satisfies The vectors form an orthonormal basis of . In particular,
(3.5) |
where with being the floor function (see e.g., [15]). It is verified that for all , which indicates that for ,
(3.6) |
Introduce the discrete kernel
where for , . Define the discrete Dirichlet Laplacian by for ,
(3.7) |
where with . Since ,
Similar to [15, Section 2], based on (3), the diagonalization of the matrix , (3.4) and , one has
(3.8) |
The follow lemma characterizes the error between and .
Lemma 3.1.
There exists some constant such that for any and ,
(3.9) | ||||
(3.10) |
Proof.
For and , denote , and . Since is an orthonormal basis of , it holds that
(3.11) |
It follows from the boundedness of and (2.17) that for ,
(3.12) |
By (3.6), we have and thus, , which along with (2.20) yields . Besides, it can be verified that Therefore, for
Choosing , we obtain
(3.13) |
We recall the following inequality in [15, Lemma 3.2]:
(3.14) |
Introducing and making use of (3.14), we obtain
Hence, it follows from (2.17) that for ,
(3.15) |
Combining (3)-(3.13) with (3.15) allows us to deduce
which proves (3.9). It remains to prove (3.10). Set Then where
For the first term, we have . For , which implies that with arbitrarily small . Since , it holds that for , and thus Using (3.14), we arrive at
Combining the above estimates completes the proof of (3.10). ∎
For , denote by the exact solution of Eq. (1.1) on spatial grid points, where the explicit dependence of on is omitted. We introduce the following auxiliary process by
with initial value . Let satisfy
(3.16) |
Then for and . In order to estimate , it suffices to estimate and , where the first term is tackled as follows.
Lemma 3.2.
Proof.
Recall that , where , , are defined in (2.10)-(2.12), respectively. Similarly, for , we introduce ,
and divide the proof into three parts.
Part 1: Following the proof of [4, Lemma 2.3], we use the PDE satisfied by to write As a numerical counterpart,
(3.17) |
where in the last step we have used the fact that
for with on . Here, for , and for , . In particular, when , it holds that
(3.18) |
By and (3.7), there exist such that for ,
and for , Therefore, using (2.2) and (3.9), a direct calculation gives
(3.19) |
By Lemma 3.2, we present the strong convergence rate of the spatial FDM for Eq. (1.1). We would like to mention that Theorem 3.3 also holds for stochastic Cahn–Hilliard equations with NBCs.
Theorem 3.3.
Suppose that Assumption 1 holds and . Then for every , there exists some constant such that for any ,
Proof.
Denote . In view of (3) and (3),
By the expressions of and and (2.17), we arrive at that for ,
(3.21) |
Hence, the Cauchy–Schwarz inequality with respect to the measure , Lemma 3.2, Assumption 1, and the Minkowski and Burkholder inequalities yield that for ,
in which the second step used (3.21). Taking the supremum over produces
(3.22) |
which along with the Gronwall lemma with weak singularities (see e.g., [15, Lemma 3.4]) completes the proof. ∎
4. Convergence of density
For real-valued random variables , we write to indicate the total variation distance between and , i.e.,
where is the set of continuous functions which are bounded by , and is the Borel -algebra of . Furthermore, if and have the densities and respectively, then
(4.1) |
In this section, we show that for , the spatial semi-discrete numerical solution admits a density, which converges in to the density of the exact solution .
4.1. Malliavin calculus
We start with introducing some notations in the context of the Malliavin calculus with respect to the space-time white noise (see e.g., [23]). The isonormal Gaussian family corresponding to is given by the Wiener integral Denote by the class of smooth real-valued random variables of the form
(4.2) |
where Here is the space of all -valued smooth functions on whose partial derivatives have at most polynomial growth. The Malliavin derivative of of the form (4.2) is an -valued random variable given by which is also a random field with for almost everywhere . For any , we denote the domain of in by , meaning that is the closure of with respect to the norm
We define the iteration of the operator in such a way that for , the iterated derivative is an -valued random variable. More precisely, for , is a measurable function on the product space . Then for , , denote by the completion of with respect to the norm Define and to be topological projective limits.
We close this part by the following proposition, which allows us to obtain the convergence of density of a sequence of random variables from the convergence in .
Proposition 4.1.
[22, Theorem 4.2] Let be a sequence in such that each admits a density. Let and let be such that . If in , then there exists a constant depending only on such that for any ,
4.2. Convergence in
In this part, we extend the strong convergence of the spatial FDM to the convergence in . It is shown in [8, Proposition 3.1] or [4, Lemma 3.2] that if , then for any , and satisfies
(4.3) |
if , and , if . Their proofs rely on the global Lipschitz continuity of , and thus (4.2) holds naturally whenever satisfies Assumption 1. Further, we impose Assumption 3 to study the regularity of the exact solution in the Malliavin Sobolev space.
Assumption 3.
For some integer , and have bounded derivatives up to order .
Proof.
Define the Picard approximation by , , and for ,
Fix . In view of [23, Lemma 1.5.3], the proof of boils down to proving that
(i) converges to in for every .
(ii) for any ,
Property (i) and property (ii) with and can be obtained in the same way as in [4, Lemma 3.2] (the sequence corresponds to in [4]). The proof of property (ii) with general is omitted since it is standard and similar to those for other kinds of SPDEs with Lipschitz continuous coefficients; see [2, Proposition 4.3] for the case of stochastic heat equations, [25, Theorem 1] for the case of stochastic wave equations. ∎
Similar to properties (i) and (ii), the standard Picard approximation also shows that for any , .
Proposition 4.3.
Proof.
By the chain rule and (3), we obtain
(4.4) |
if , and , if . Combining (4.2) and (4.2), we write
(4.5) |
where for , , and for ,
When , we always set for and . Hereafter, let be an arbitrarily fixed positive number. A combination of Lemma 2.2 and Theorem 3.3 reveals that for any ,
(4.6) |
for all . Then the Lipschitz continuity of , the Minkowski and Cauchy-Schwarz inequalities and (3.21) produce
where (4.2) and Lemma 4.2 were used in the second line. Similarly, by the boundedness of , Lemma 2.3, Lemma 4.2, and (3.9),
Since is bounded and Lipschitz continuous, it follows from the Minkowski inequality, (4.2), (3.10), and (2.1) that for ,
Replacing by in the above inequality, we also have
which along with the Burkholder inequality for Hilbert space valued martingales (see e.g. [2, (4.18)]), the Hölder inequality and Lemma 4.2 indicates
In order to estimate and , we claim that for , there exists some constant such that for any and ,
(4.7) |
Indeed, from (4.2), we have that for ,
For , let , . The boundedness of and Lemma 2.1 indicate
Since is bounded, it follows from (2) and Lemma 4.2 that
Similarly, it follows from the Burkholder inequality, Lemmas 2.1 and 4.2 that Gathering the above estimates of , and , we obtain (4.7). By means of (4.7) and (3.21), it can be verified that Substituting the above estimates of , , , and , , into (4.5), we deduce that for ,
(4.8) |
where the last step used the Hölder inequality and (3.21). Similar to (3.22), by taking the supremum over on both sides of (4.2) and applying the Gronwall lemma with weak singularities (see e.g., [15, Lemma 3.4]), we complete the proof. ∎
4.3. Convergence of density
In this part, we present the convergence of density of the numerical solution for . In order to apply Proposition 4.1 with , we impose Assumption 4 and investigate the negative moment estimate of .
Assumption 4.
There exists some such that , for any .
Proof.
We are ready to give the main result of this section, which states that for , the density of the numerical solution exists and converges in to the density of the exact solution. The readers are referred to [4, 8] for the existence of the density of the exact solution for any .
Theorem 4.5.
Proof.
By [23, Theorem 2.3.3] and (3.2), we obtain that under Assumption 4, for any , the law of is absolutely continuous with respect to the Lebesgue measure on . Thus, admits a density. Theorem 3.3, Lemma 4.2, Proposition 4.3, and Lemma 4.4 indicate that the conditions of Proposition 4.1 are fulfilled for , and . As a result,
which together to (4.1) completes the proof. ∎
5. Full discretization
For the purpose of effective computation, we combine the spatial FDM with a temporal exponential Euler method to obtain the full discretization of Eq. (1.1), and give the strong convergence rate of the fully discrete numerical solution in this section.
Let (with ) be a uniform partition of , where is the uniform time stepsize. Denote by the largest time grid point smaller than . By replacing in (3) by , we obtain the full discretization given by
(5.1) |
The discrete Green function satisfies the following estimates.
Lemma 5.1.
Let . Then for any and with ,
Proof.
In a similar way, one can prove the following lemma.
Lemma 5.2.
Let . Then for any and with ,
(5.4) | |||
(5.5) |
Proof.
Proposition 5.3.
Let Assumption 1 hold and . Then for any and , there exists such that for any ,
Proof.
As a result of Theorem 3.3 and Lemma 2.3, for and ,
(5.6) |
Recall that . It follows from (3) and (3.18) that
Since , for , and hence (5.5) implies Similarly, based on (3), (5.4), and (5.5), one also has that for any , Based on a standard argument as in the proof of Lemma 2.2, it follows from (5.6) and Lemmas 5.1 and 5.2 that . The proof is completed. ∎
Theorem 5.4.
Suppose that Assumption 1 holds and . Then for every and , there exists some constant such that for any ,
Proof.
Let and . By virtue of Theorem 3.3, it remains to show
By (3), (5), the Minkowski inequality, the Burkholder inequality, and the Lipschitz continuity of and , we obtain that for any ,
where
Taking advantage of Corollary 5.3 and (3.21), we obtain
Similar to the proof of (5), one has that for ,
which along with the boundedness of shows that . Similar to (5.5) with , we also have that for ,
which along with (5.6) reveals that . Gathering the above estimates together yields that for any ,
Letting , and using (3.21), we obtain
(5.7) |
Hence, for any ,
which together with the discrete Gronwall lemma (see e.g. [20, Lemma A.4]) implies that Finally, taking (5.7) into account gives
Thus the proof is complete. ∎
Remark 5.5.
The application of the orthogonality of plays a key role to obtain the temporal convergence order nearly of the full discretization in Theorem 5.4. For example, if the left hand of (5.4) is estimated in the following way
with , then we can only obtain with . As a result, the temporal Hölder continuity exponent of is only nearly , which leads to that the temporal convergence order of the exponential Euler method is only nearly .
Combining Theorem 5.4 with a localized argument, we show an convergence order localized on a set of arbitrarily large probability for Eq. (1.1) with polynomial nonlinearity.
Corollary 5.6.
Suppose that Assumption 2 hold and . Then for any , and , there exists such that
Proof.
For , denote Set with defined by (2.3). Consider the localized Cahn-Hilliard equation
(5.8) |
with and DBCs. Then the local property of stochastic integrals shows (i.e., for any , ) on a.s. Consider the numerical solution of (5.8) based on the FDM in space and the exponential Euler method in time, i.e.,
(5.9) |
for and . Seting , and comparing (5) with (5), it follows from the local property of stochastic integrals that on a.s. For fixed , since satisfies Assumption 1, Theorem 5.4 indicates that there exists some constant such that for and ,
Since and have almost surely continuous trajectories, we have , which implies . By on a.s., and on a.s., we obtain
The proof is completed. ∎
Remark 5.7.
Theorems 3.3 and 4.5 indicate that when applying the spatial FDM to the localized Cahn–Hilliard equation (1.2), the associated numerical solution is strongly convergent and the density of the numerical solution converges in . In addition, Section 2 gives the uniform moment estimate and Hölder continuity of the exact solution for Eq. (1.1) with being a polynomial of degree with a positive dominant coefficient. We expect to combine the above results with the localization technique to study the strong convergence of the spatial FDM and the density convergence of the associated numerical solution for the stochastic Cahn–Hilliard equation with polynomial nonlinearity and multiplicative noise in the future.
References
- [1] Rikard Anton, David Cohen, and Lluis Quer-Sardanyons. A fully discrete approximation of the one-dimensional stochastic heat equation. IMA J. Numer. Anal., 40(1):247–284, 2020.
- [2] Vlad Bally and Etienne Pardoux. Malliavin calculus for white noise driven parabolic SPDEs. Potential Anal., 9(1):27–64, 1998.
- [3] Vlad Bally and Denis Talay. The law of the Euler scheme for stochastic differential equations. II. Convergence rate of the density. Monte Carlo Methods Appl., 2(2):93–128, 1996.
- [4] Caroline Cardon-Weber. Cahn-Hilliard stochastic equation: existence of the solution and of its density. Bernoulli, 7(5):777–816, 2001.
- [5] Shimin Chai, Yanzhao Cao, Yongkui Zou, and Wenju Zhao. Conforming finite element methods for the stochastic Cahn-Hilliard-Cook equation. Appl. Numer. Math., 124:44–56, 2018.
- [6] Chuchu Chen, Jianbo Cui, Jialin Hong, and Derui Sheng. Convergence of Density Approximations for Stochastic Heat Equation. arXiv:2007.12960.
- [7] David Cohen and Lluís Quer-Sardanyons. A fully discrete approximation of the one-dimensional stochastic wave equation. IMA J. Numer. Anal., 36(1):400–420, 2016.
- [8] Jianbo Cui and Jialin Hong. Absolute continuity and numerical approximation of stochastic Cahn-Hilliard equation with unbounded noise diffusion. J. Differential Equations, 269(11):10143–10180, 2020.
- [9] Jianbo Cui, Jialin Hong, and Derui Sheng. Convergence in Density of Splitting AVF Scheme for Stochastic Langevin Equation. arXiv:1906.03439.
- [10] Jianbo Cui, Jialin Hong, and Liying Sun. Strong convergence of full discretization for stochastic Cahn-Hilliard equation driven by additive noise. SIAM J. Numer. Anal., 59(6):2866–2899, 2021.
- [11] A. M. Davie and J. G. Gaines. Convergence of numerical schemes for the solution of parabolic stochastic partial differential equations. Math. Comp., 70(233):121–134, 2001.
- [12] Qiang Du and R. A. Nicolaides. Numerical analysis of a continuum model of phase transition. SIAM J. Numer. Anal., 28(5):1310–1322, 1991.
- [13] Charles M. Elliott and Stig Larsson. Error estimates with smooth and nonsmooth data for a finite element method for the Cahn-Hilliard equation. Math. Comp., 58(198):603–630, S33–S36, 1992.
- [14] Daisuke Furihata, Mihály Kovács, Stig Larsson, and Fredrik Lindgren. Strong convergence of a fully discrete finite element approximation of the stochastic Cahn-Hilliard equation. SIAM J. Numer. Anal., 56(2):708–731, 2018.
- [15] István Gyöngy. Lattice approximations for stochastic quasi-linear parabolic partial differential equations driven by space-time white noise. I. Potential Anal., 9(1):1–25, 1998.
- [16] Y. Hu and S. Watanabe. Donsker’s delta functions and approximation of heat kernels by the time discretization methods. J. Math. Kyoto Univ., 36(3):499–518, 1996.
- [17] Davar Khoshnevisan. Analysis of Stochastic Partial Differential Equations, volume 119 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2014.
- [18] A. Kohatsu-Higa. High order Itô-Taylor approximations to heat kernels. J. Math. Kyoto Univ., 37(1):129–150, 1997.
- [19] Mihály Kovács, Stig Larsson, and Ali Mesforush. Finite element approximation of the Cahn-Hilliard-Cook equation. SIAM J. Numer. Anal., 49(6):2407–2429, 2011.
- [20] R. Kruse. Strong and Weak Approximation of Semilinear Stochastic Evolution Equations, volume 2093 of Lecture Notes in Mathematics. Springer, Cham, 2014.
- [21] Stig Larsson and Ali Mesforush. Finite-element approximation of the linearized Cahn-Hilliard-Cook equation. IMA J. Numer. Anal., 31(4):1315–1333, 2011.
- [22] Ivan Nourdin and Guillaume Poly. Convergence in total variation on Wiener chaos. Stochastic Process. Appl., 123(2):651–674, 2013.
- [23] David Nualart. The Malliavin Calculus and Related Topics. Probability and its Applications (New York). Springer-Verlag, Berlin, second edition, 2006.
- [24] Ruisheng Qi and Xiaojie Wang. Error estimates of semidiscrete and fully discrete finite element methods for the Cahn-Hilliard-Cook equation. SIAM J. Numer. Anal., 58(3):1613–1653, 2020.
- [25] Lluís Quer-Sardanyons and Marta Sanz-Solé. A stochastic wave equation in dimension 3: smoothness of the law. Bernoulli, 10(1):165–186, 2004.