This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Analysis of Chorin-Type Projection Methods for the Stochastic Stokes Equations with General Multiplicative Noises†

Xiaobing Feng Department of Mathematics, The University of Tennessee, Knoxville, TN 37996, U.S.A. [email protected]  and  Liet Vo Department of Mathematics, The University of Tennessee, Knoxville, TN 37996, U.S.A. [email protected]
Abstract.

This paper is concerned with numerical analysis of two fully discrete Chorin-type projection methods for the stochastic Stokes equations with general non-solenoidal multiplicative noise. The first scheme is the standard Chorin scheme and the second one is a modified Chorin scheme which is designed by employing the Helmholtz decomposition on the noise function at each time step to produce a projected divergence-free noise and a “pseudo pressure” after combining the original pressure and the curl-free part of the decomposition. An O(k14)O(k^{\frac{1}{4}}) rate of convergence is proved for the standard Chorin scheme, which is sharp but not optimal due to the use of non-solenoidal noise, where kk denotes the time mesh size. On the other hand, an optimal convergence rate O(k12)O(k^{\frac{1}{2}}) is established for the modified Chorin scheme. The fully discrete finite element methods are formulated by discretizing both semi-discrete Chorin schemes in space by the standard finite element method. Suboptimal order error estimates are derived for both fully discrete methods. It is proved that all spatial error constants contain a growth factor k12k^{-\frac{1}{2}}, where kk denotes the time step size, which explains the deteriorating performance of the standard Chorin scheme when k0k\to 0 and the space mesh size is fixed as observed earlier in the numerical tests of [8]. Numerical results are also provided to guage the performance of the proposed numerical methods and to validate the sharpness of the theoretical error estimates.

Key words and phrases:
Stochastic Stokes equations, multiplicative noise, Wiener process, Itô stochastic integral, Chorin projection scheme, inf-sup condition, error estimates
2010 Mathematics Subject Classification:
Primary 65N12, 65N15, 65N30,
†This work was partially supported by the NSF grants DMS-1620168 and DMS-2012414.

1. Introduction

This paper is concerned with developing and analyzing Chorin-type projection finite element methods for the following time-dependent stochastic Stokes problem:

(1.1a) d𝐮\displaystyle d{\bf u} =[νΔ𝐮p+𝐟]dt+𝐁(𝐮)d𝐖(t)\displaystyle=\bigl{[}\nu\Delta{\bf u}-\nabla p+{{\bf f}}\bigr{]}dt+{\bf B}({\bf u})d{\bf W}(t) a.s. inDT:=(0,T)×D,\displaystyle\qquad\mbox{a.s. in}\,D_{T}:=(0,T)\times D,
(1.1b) div 𝐮\displaystyle\mbox{\rm div\,}{\bf u} =0\displaystyle=0 a.s. inDT,\displaystyle\qquad\mbox{a.s. in}\,D_{T},
(1.1c) 𝐮(0)\displaystyle{\bf u}(0) =𝐮0\displaystyle={\bf u}_{0} a.s. inD,\displaystyle\qquad\mbox{a.s. in}\,D,

where D=(0,L)2d(d=2,3)D=(0,L)^{2}\subset\mathbb{R}^{d}\,(d=2,3) represents a period of the periodic domain in d\mathbb{R}^{d}, 𝐮{\bf u} and pp stand for respectively the velocity field and the pressure of the fluid, 𝐁{\bf B} is an operator-valued random field, {𝐖(t);t0}\{{\bf W}(t);t\geq 0\} denotes an 𝐋2(D){\bf L}^{2}(D)-valued QQ-Wiener process, and 𝐟{\bf f} is a body force function (see Section 2 for their precise definitions). Here we seek periodic-in-space solutions (𝐮,p)({\bf u},p) with period LL, that is, 𝐮(t,𝐱+L𝐞i)=𝐮(t,𝐱){\bf u}(t,{\bf x}+L{\bf e}_{i})={\bf u}(t,{\bf x}) and p(t,𝐱+L𝐞i)=p(t,𝐱)p(t,{\bf x}+L{\bf e}_{i})=p(t,{\bf x}) almost surely and for any (t,𝐱)(0,T)×d(t,{\bf x})\in(0,T)\times\mathbb{R}^{d} and 1id1\leq i\leq d, where {𝐞𝐢}𝐢=𝟏𝐝\{\bf e_{i}\}_{i=1}^{d} denotes the canonical basis of d\mathbb{R}^{d}.

The system (1.1a) is a stochastic perturbation of the deterministic Stokes system by introducing a multiplicative noise force term 𝐁()d𝐖(s){\bf B}(\cdot)d{\bf W}(s) and it has been used to model turbulent fluids (cf. [1, 2, 17, 21]). The stochastic Stokes system is a simplified model of the full stochastic Navier-Stokes equations by omitting the nonlinear term (𝐮)𝐮({\bf u}\cdot\nabla){\bf u} in the drift part of the stochastic Navier-Stokes equations. Although the deterministic Stokes equations is a linear PDE system which has been well studied in the literature (cf. [14, 21] and the references therein), the stochastic Stokes system (1.1a) is intrinsically nonlinear because the diffusion coefficient 𝐁{\bf B} is nonlinear in the velocity 𝐮{\bf u}. Due to the introduction of random forces it has been well known that the solution of problem (1.1) has very low regularities in time. We refer the reader to [1, 18, 11] and the references therein for a detailed account about the well-posedness and regularities of the solution for system (1.1).

Besides their mathematical and practical importance, the stochastic Stokes (and Navier-Stokes) equations have been used as prototypical stochastic PDEs for developing efficient numerical methods and general numerical analysis techniques for analyzing numerical methods for stochastic PDEs. In that regard several works have been reported in the literature [9, 12, 13, 5, 8, 3]. Euler-Maruyama time dsicretization and divergence-free finite element space discretization was proposed and analyzed in [9] in the case of divergence-free noises (i.e., 𝐁(𝐮)\mathbf{B}({\bf u}) is divergence-free). Optimal order error estimates in strong norm for the velocity approximation were obtained. In [12, 13] the authors considered the general noise and analyzed the standard and a modified mixed finite element methods as well as pressure stabilized methods for space discretization, suboptimal order error estimates were proved in [12] for the velocity approximation in strong norm and for the pressure approximation in a time-averaged norm, all these suboptimal order error estimates were improved to optimal order for a Helmholtz projection-enhanced mixed finite element in [13] (also see [5] for a similar approach). It should be noted that the reason for measuring the pressure errors in a time-averaged norm is because the low regularity of the pressure field which is only a distribution in general and the numerical tests of [12, 13] suggest that these error estimates are sharp. In [8] the authors proposed a Chorin time-splitting finite element method for problem (1.1) and proved a suboptimal convergence rate in strong norm for the velocity approximation in the case of divergence-free noises. In [3] the authors proposed an iterative splitting scheme for stochastic Navier-Stokes equations and a strong convergence in probability was established in the 2-D case for the velocity approximation. In a recent work [4], the authors proposed another time-splitting scheme and proved its strong L2L^{2} convergence for the velocity approximation.

Compared to the recent advances on mixed finite element methods [9, 12, 13], the numerical analysis of the well-known Chorin projection/splitting scheme for the stochastic Stokes equations lags behind. To the best of our knowledge, the only analysis result obtained in [8] is the optimal convergence in the energy norm for the velocity approximation in the case of divergence-free noises (i.e., 𝐁(𝐮)\mathbf{B}({\bf u}) is divergence-free). Several natural and important questions arise and must be addressed for a better understanding of the Chorin projection scheme for problem (1.1). Among them are (i) Does the pressure approximation converge even when the noise is divergence-free? If so, in what sense and what rate? (ii) Does the Chorin projection scheme converge (for both the velocity and pressure approximations) for general noises? If so, in what sense and what rate? (iii) Could the performance of the standard Chorin projection scheme be improved one way or another in the case of general noises? The primary objective this paper is to provide a positive answer to each of the above questions.

As it was shown in [8], the adaptation of the standard deterministic Chorin projection scheme to problem (1.1) is straightforward (see Algorithm 1 of Section 3). The idea of the Chorin scheme is to separate the computation of the velocity and pressure at each time step which is done by solving two decoupled Poisson problems and the divergence-free constraint for the velocity approximation is enforced by a Helmholtz projection technique which can be easily obtained using the solutions of the two Poisson problems. The Chorin scheme also can be compactly rewritten as a pressure stabilization scheme at each time step as follows (cf. [8]):

(1.2a) 𝐮~n+1𝐮~nkνΔ𝐮~n+1+kpn\displaystyle\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}-k\nu\Delta\tilde{{\bf u}}^{n+1}+k\nabla p^{n} =k𝐟n+1+𝐁(𝐮~n)ΔWn+1\displaystyle=k{\bf f}^{n+1}+{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1} a.s. in DT,\displaystyle\qquad\mbox{a.s. in }D_{T},
(1.2b) div 𝐮~n+1kΔpn+1\displaystyle\mbox{\rm div\,}\tilde{{\bf u}}^{n+1}-k\Delta p^{n+1} =0\displaystyle=0 a.s. in DT,\displaystyle\qquad\mbox{a.s. in }D_{T},
(1.2c) 𝐧pn+1\displaystyle\partial_{\bf n}p^{n+1} =0\displaystyle=0 a.s. on DT,\displaystyle\qquad\mbox{a.s. on }\partial D_{T},
where 𝐧pn+1\partial_{\bf n}p^{n+1} denotes the normal derivative of pn+1p^{n+1} and kk is the time step size.

One of advantages of the above Chorin scheme is that the spatial approximation spaces for 𝐮~n+1\tilde{{\bf u}}^{n+1} and pn+1p^{n+1} can be chosen independently, so unlike in the mixed finite element method, they are not required to satisfy an inf-sup condition. Notice that a time lag on pressure appears in equation (1.2a) which causes most of difficulties in the convergence analysis (cf. [20, 15, 19, 8]). We also note that the term kΔpn+1-k\Delta p^{n+1} in equation (1.2b) is known as a pressure stabilization term.

To improve the convergence of the standard Chorin scheme, we adopt a Helmholtz projection technique as used in [13] (also see [5]). At each time step we first perform the Helmholtz decomposition 𝐁(𝐮~n)=𝜼n+ξn{\bf B}(\tilde{{\bf u}}^{n})=\boldsymbol{\eta}^{n}+\nabla\xi^{n} and then rewrite (1.2a) as

(1.3) 𝐮~n+1𝐮~nkνΔ𝐮~n+1+krn=k𝐟n+1+𝜼nΔWn+1a.s. in DT,\displaystyle\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}-k\nu\Delta\tilde{{\bf u}}^{n+1}+k\nabla r^{n}=k{\bf f}^{n+1}+\boldsymbol{\eta}^{n}\Delta W_{n+1}\qquad\mbox{a.s. in }D_{T},

where rn=pnk1ξnΔWn+1r^{n}=p^{n}-k^{-1}\xi^{n}\Delta W_{n+1}. Our modified Chorin scheme consists of (1.3), (1.2b)–(1.2c) and the Helmholtz decomposition 𝐁(𝐮~n)=𝜼n+ξn{\bf B}(\tilde{{\bf u}}^{n})=\boldsymbol{\eta}^{n}+\nabla\xi^{n}. Since 𝜼n\boldsymbol{\eta}^{n} is divergence-free, it turns out that the finite element approximation of the modified Chorin scheme has better convergence properties. Notice that pnp^{n} can be recovered from rnr^{n} via the simple algebraic relation pn=rn+k1ξnΔWn+1p^{n}=r^{n}+k^{-1}\xi^{n}\Delta W_{n+1}.

The main contributions of this paper are summarized below.

  • We proved the following error estimates in strong norms for the Chorin-P1P_{1} finite element method (see Algorithm 3) for problem (1.1) with general multiplicative noises:

    (𝔼[km=0M𝐮(tm)𝐮~hm2])12\displaystyle\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\|{\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h}\|^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}} +max0M(𝔼[km=0(𝐮(tm)𝐮~hm)2])12\displaystyle+\max_{0\leq\ell\leq M}\biggl{(}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{m=0}^{\ell}\nabla({\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h})\Bigr{\|}^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}
    C(k14+hk12),\displaystyle\qquad\qquad\qquad\leq C\Bigl{(}k^{\frac{1}{4}}+hk^{-\frac{1}{2}}\Bigr{)},
    (𝔼[km=0MP(tm)\displaystyle\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\Bigl{\|}P(t_{m}) kn=0mphn2])12C(k14+hk12),\displaystyle-k\sum_{n=0}^{m}p^{n}_{h}\Bigr{\|}^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}\leq C\Bigl{(}k^{\frac{1}{4}}+hk^{-\frac{1}{2}}\Bigr{)},

    where (𝐮(tm),P(tm))({\bf u}(t_{m}),P(t_{m})) are the solution to problem (1.1) while (𝐮~hm,phm)(\tilde{{\bf u}}_{h}^{m},p_{h}^{m}) are the discrete solution of Algorithm 3, see Sections 2 and 4 for their precise definitions.

  • We proposed a modified Chorin-P1P_{1} finite element method (see Algorithm 4) and proved the following error estimates in strong norms for problem (1.1) with general multiplicative noises:

    max1mM(𝔼[𝐮(tm)𝐮~hm2])12+(𝔼[km=1M(𝐮(tn)𝐮hn)2])12\displaystyle\max_{1\leq m\leq M}\Bigl{(}\mathbb{E}\bigl{[}\bigl{\|}{\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h}\bigr{\|}^{2}\,\bigl{]}\Bigr{)}^{\frac{1}{2}}+\biggl{(}\mathbb{E}\biggl{[}k\sum_{m=1}^{M}\bigl{\|}\nabla({\bf u}(t_{n})-{\bf u}^{n}_{h})\bigr{\|}^{2}\,\biggr{]}\biggr{)}^{\frac{1}{2}}
    C(k12+h+k12h2),\displaystyle\hskip 137.31255pt\leq C\Bigl{(}k^{\frac{1}{2}}+h+k^{-\frac{1}{2}}h^{2}\Bigr{)},
    (𝔼[R(tm)kn=1mrhn2])12+(𝔼[P(tm)kn=1mphn2])12\displaystyle\biggl{(}\mathbb{E}\biggl{[}\bigg{\|}R(t_{m})-k\sum^{m}_{n=1}r^{n}_{h}\biggr{\|}^{2}\,\biggr{]}\biggr{)}^{\frac{1}{2}}+\biggl{(}\mathbb{E}\biggl{[}\bigg{\|}P(t_{m})-k\sum^{m}_{n=1}p^{n}_{h}\biggr{\|}^{2}\,\biggr{]}\biggr{)}^{\frac{1}{2}}
    C(k12+h+k12h2).\displaystyle\hskip 137.31255pt\leq C\Bigl{(}k^{\frac{1}{2}}+h+k^{-\frac{1}{2}}h^{2}\Bigr{)}.

    where (𝐮(tm),P(tm))({\bf u}(t_{m}),P(t_{m})) is the solution to problem (1.1) and R(t)R(t) is defined as the time-average of the pseudo pressure r(t)r(t) while (𝐮hm,rhm,phm)({\bf u}^{m}_{h},r^{m}_{h},p^{m}_{h}) is the solution of Algorithm 4, see Sections 2 and 4 for their precise definitions.

We note that all spatial error constants contain a growth factor k12k^{-\frac{1}{2}}, which explains the deteriorating performance of the standard (and modified) Chorin scheme when k0k\to 0 and the mesh size hh is fixed as observed in the numerical tests of [8]. The numerical experiments to be given in Section 5 indicate that the dependence on factor k12k^{-\frac{1}{2}} is sharp.

The remainder of this paper is organized as follows. In Section 2, we first introduce some space notations and state the assumptions on the initial data and on 𝐁{\bf B} as well as recall the definition of solutions to (1.1). We then state and prove a Hölder continuity property for the pressure pp in a time-averaged norm. In Section 3, we define the standard Chorin projection scheme as Algorithm 1 for problem (1.1) in Subsection 3.1 and the modified Chorin scheme as Algorithm 2 in Subsection 3.2. The highlights of this section are to prove some uniform (in kk) stability estimates which are very useful for error analysis later. In Section 4, we formulate the finite element spatial discretization for both the standard Chorin and modified Chorin schemes in Algorithm 3 and 4, respectively and prove the quasi-optimal error estimates for both algorithms as summarized above. In Section 5, we present several numerical experiments to gauge the performance of the proposed numerical methods and to validate the sharpness of the proved error estimates.

2. Preliminaries

Standard function and space notation will be adopted in this paper. Let 𝐇01(D){\bf H}^{1}_{0}(D) denote the subspace of 𝐇1(D){\bf H}^{1}(D) whose d{\mathbb{R}}^{d}-valued functions have zero trace on D\partial D, and (,):=(,)D(\cdot,\cdot):=(\cdot,\cdot)_{D} denote the standard L2L^{2}-inner product, with induced norm \|\cdot\|. We also denote 𝐋perp(D){\bf L}^{p}_{per}(D) and 𝐇perk(D){\bf H}^{k}_{per}(D) as the Lebesgue and Sobolev spaces of the functions that are periodic and have vanishing mean, respectively. Let (Ω,,{t},)(\Omega,\mathcal{F},\{\mathcal{F}_{t}\},\mathbb{P}) be a filtered probability space with the probability measure \mathbb{P}, the σ\sigma-algebra \mathcal{F} and the continuous filtration {t}\{\mathcal{F}_{t}\}\subset\mathcal{F}. For a random variable vv defined on (Ω,,{t},)(\Omega,\mathcal{F},\{\mathcal{F}_{t}\},\mathbb{P}), 𝔼[v]{\mathbb{E}}[v] denotes the expected value of vv. For a vector space XX with norm X\|\cdot\|_{X}, and 1p<1\leq p<\infty, we define the Bochner space (Lp(Ω,X);vLp(Ω,X))\bigl{(}L^{p}(\Omega,X);\|v\|_{L^{p}(\Omega,X)}\bigr{)}, where vLp(Ω,X):=(𝔼[vXp])1p\|v\|_{L^{p}(\Omega,X)}:=\bigl{(}{\mathbb{E}}[\|v\|_{X}^{p}]\bigr{)}^{\frac{1}{p}}. We also define

\displaystyle{\mathbb{H}} :={𝐯𝐋per2(D);div 𝐯=0 in D},\displaystyle:=\bigl{\{}{\bf v}\in{\bf L}^{2}_{per}(D);\,\mbox{\rm div\,}{\bf v}=0\mbox{ in }D\,\bigr{\}}\,,
𝕍\displaystyle{\mathbb{V}} :={𝐯𝐇per1(D);div 𝐯=0 in D}.\displaystyle:=\bigl{\{}{\bf v}\in{\bf H}^{1}_{per}(D);\,\mbox{\rm div\,}{\bf v}=0\mbox{ in }D\bigr{\}}\,.

We recall from [14] that the (orthogonal) Helmholtz projection 𝐏:𝐋per2(D){\bf P}_{{\mathbb{H}}}:{\bf L}^{2}_{per}(D)\rightarrow{\mathbb{H}} is defined by 𝐏𝐯=𝜼{\bf P}_{{\mathbb{H}}}{\bf v}=\boldsymbol{\eta} for every 𝐯𝐋per2(D){\bf v}\in{\bf L}^{2}_{per}(D), where (𝜼,ξ)×Hper1(D)/(\boldsymbol{\eta},\xi)\in{\mathbb{H}}\times H^{1}_{per}(D)/\mathbb{R} is a unique tuple such that

𝐯=𝜼+ξ,{\bf v}=\boldsymbol{\eta}+\nabla\xi\,,

and ξHper1(D)/\xi\in H^{1}_{per}(D)/\mathbb{R} solves the following Poisson problem with the homogeneous Neumann boundary condition:

(2.1) Δξ=div 𝐯.\Delta\xi=\mbox{\rm div\,}{\bf v}.

We also define the Stokes operator 𝐀:=𝐏Δ:𝕍𝐇per2(D){\bf A}:=-{\bf P}_{\mathbb{H}}\Delta:{\mathbb{V}}\cap{\bf H}^{2}_{per}(D)\rightarrow{\mathbb{H}}.

Throughout this paper we assume that 𝐁:𝐋per2(D)𝐋per2(D){\bf B}:{\bf L}^{2}_{per}(D)\rightarrow{\bf L}^{2}_{per}(D) is a Lipschitz continuous mapping and has linear growth, that is, there exists a constant C>0C>0 such that for all 𝐯,𝐰𝐋per2(D){\bf v},{\bf w}\in{\bf L}^{2}_{per}(D)

(2.2a) 𝐁(𝐯)𝐁(𝐰)\displaystyle\|{\bf B}({\bf v})-{\bf B}({\bf w})\| C𝐯𝐰,\displaystyle\leq C\|{\bf v}-{\bf w}\|\,,
(2.2b) 𝐁(𝐯)\displaystyle\|{\bf B}({\bf v})\| C(1+𝐯),\displaystyle\leq C\bigl{(}1+\|{\bf v}\|\bigr{)}\,,

Since we shall not explicitly track the dependence of all constants on ν\nu, for ease of the presentation, unless it is stated otherwise, we shall set ν=1\nu=1 in the rest of the paper and assume that 𝐟L2(Ω;Lper2(D)){\bf f}\in L^{2}(\Omega;L^{2}_{per}(D)). In addition, we shall use CC to denote a generic positive constant which may depend on TT, the datum functions 𝐮0{\bf u}_{0} and 𝐟{\bf f}, and the domain DD but is independent of the mesh parameter hh and kk.

2.1. Variational formulation of the stochastic Stokes equations

We first define the variational solution concept for (1.1) and refer the reader to [10, 11] for a proof of its existence and uniqueness.

Definition 2.1.

Given (Ω,,{t},)(\Omega,\mathcal{F},\{\mathcal{F}_{t}\},\mathbb{P}), let WW be an {\mathbb{R}}-valued Wiener process on it. Suppose 𝐮0L2(Ω,𝕍){\bf u}_{0}\in L^{2}(\Omega,{\mathbb{V}}) and 𝐟L2(Ω;L2((0,T);Lper2(D))){\bf f}\in L^{2}(\Omega;L^{2}((0,T);L^{2}_{per}(D))). An {t}\{\mathcal{F}_{t}\}-adapted stochastic process {𝐮(t);0tT}\{{\bf u}(t);0\leq t\leq T\} is called a variational solution of (1.1) if 𝐮L2(Ω;C([0,T];𝕍))L2(Ω;0,T;𝐇per2(D)){\bf u}\in L^{2}\bigl{(}\Omega;C([0,T];{\mathbb{V}}))\cap L^{2}\bigl{(}\Omega;0,T;{\bf H}^{2}_{per}(D)\bigr{)}, and satisfies \mathbb{P}-a.s. for all t(0,T]t\in(0,T]

(2.3) (𝐮(t),𝐯)+0t(𝐮(s),𝐯)𝑑s\displaystyle\bigl{(}{\bf u}(t),{\bf v}\bigr{)}+\int_{0}^{t}\bigl{(}\nabla{\bf u}(s),\nabla{\bf v}\bigr{)}\,ds =(𝐮0,𝐯)+0t(𝐟(s),𝐯)𝑑s\displaystyle=({\bf u}_{0},{\bf v})+\int_{0}^{t}\big{(}{\bf f}(s),{\bf v}\big{)}\,ds
+0t(𝐁(𝐮(s)),𝐯)𝑑W(s)𝐯𝕍.\displaystyle\qquad+{\int_{0}^{t}\Bigl{(}{\bf B}\bigl{(}{\bf u}(s)\bigr{)},{\bf v}\Bigr{)}\,dW(s)}\qquad\forall\,{\bf v}\in{\mathbb{V}}\,.

We cite the following Hölder continuity estimates for the variational solution whose proofs can be found in [8, 12].

Lemma 2.1.

Suppose 𝐮0L2(Ω;𝕍𝐇per2(D)){\bf u}_{0}\in L^{2}\bigl{(}\Omega;{\mathbb{V}}\cap{\bf H}^{2}_{per}(D)\bigr{)} and 𝐟L2(Ω;C12([0,T]);Hper1(D)){\bf f}\in L^{2}(\Omega;C^{\frac{1}{2}}([0,T]);H^{1}_{per}(D)). Then there exists a constant CC(DT,𝐮0,𝐟)>0C\equiv C(D_{T},{\bf u}_{0},{\bf f})>0, such that the variational solution to problem (1.1) satisfies for s,t[0,T]s,t\in[0,T]

(2.4a) 𝔼[𝐮(t)𝐮(s)2]+𝔼[st(𝐮(τ)𝐮(s))2𝑑τ]C|ts|,\displaystyle{\mathbb{E}}\bigl{[}\|{\bf u}(t)-{\bf u}(s)\|^{2}\bigr{]}+{\mathbb{E}}\Bigl{[}\int_{s}^{t}\|\nabla\bigl{(}{\bf u}(\tau)-{\bf u}(s)\bigr{)}\|^{2}\,d\tau\Bigr{]}\leq C|t-s|\,,
(2.4b) 𝔼[(𝐮(t)𝐮(s))2]+𝔼[st𝐀(𝐮(τ)𝐮(s))2𝑑τ]C|ts|.\displaystyle{\mathbb{E}}\bigl{[}\|\nabla\bigl{(}{\bf u}(t)-{\bf u}(s)\bigr{)}\|^{2}\bigr{]}+{\mathbb{E}}\Bigl{[}\int_{s}^{t}\|{\bf A}\bigl{(}{\bf u}(\tau)-{\bf u}(s)\bigr{)}\|^{2}\,d\tau\Bigr{]}\leq C|t-s|\,.
Remark 2.1.

We note that to ensure the Hölder continuity estimage (2.4b) is the only reason for restricting to the periodic boundary condition in this paper.

Definition 2.1 only defines the velocity 𝐮\mathbf{u} for (1.1), its associated pressure pp is subtle to define. In that regard we quote the following theorem from [13].

Theorem 2.1.

Let {𝐮(t);0tT}\{{\bf u}(t);0\leq t\leq T\} be the variational solution of (1.1). There exists a unique adapted process PL2(Ω,L2(0,T;Hper1(D)/))P\in{L^{2}\bigl{(}\Omega,L^{2}(0,T;H^{1}_{per}(D)/{\mathbb{R}})\bigr{)}} such that (𝐮,P)(\mathbf{u},P) satisfies \mathbb{P}-a.s. for all t(0,T]t\in(0,T]

(2.5a) (𝐮(t),𝐯)+0t(𝐮(s),𝐯)𝑑s(div 𝐯,P(t))\displaystyle\bigl{(}{\bf u}(t),{\bf v}\bigr{)}+\int_{0}^{t}\bigl{(}\nabla{\bf u}(s),\nabla{\bf v}\bigr{)}\,ds-\bigl{(}\mbox{\rm div\,}\mathbf{v},P(t)\bigr{)}
=(𝐮0,𝐯)+0t(𝐟(s),𝐯)𝑑s+0t(𝐁(𝐮(s)),𝐯)𝑑W(s)𝐯𝐇per1(D;d),\displaystyle=({\bf u}_{0},{\bf v})+\int_{0}^{t}\big{(}{\bf f}(s),{\bf v}\big{)}\,ds+{\int_{0}^{t}\bigl{(}{\bf B}\bigl{(}{\bf u}(s)\bigr{)},{\bf v}\bigr{)}\,dW(s)}\,\,\,\forall\,{\bf v}\in{\bf H}^{1}_{per}(D;\mathbb{R}^{d})\,,
(2.5b) (div 𝐮,q)=0qL02(D):={qLper2(D):(q,1)=0}.\displaystyle\bigl{(}\mbox{\rm div\,}{\bf u},q\bigr{)}=0\qquad\forall\,q\in L^{2}_{0}(D):=\{q\in L^{2}_{per}(D):\,(q,1)=0\}\,.

System (2.5) is a mixed formulation for the stochastic Stokes system (1.1), where the (time-averaged) pressure PP is defined. The distributional derivative p:=Ptp:=\frac{\partial P}{\partial t}, which was shown to belong to L1(Ω;W1,(0,T;Hper1(D)/))L^{1}\bigl{(}\Omega;W^{-1,\infty}(0,T;H^{1}_{per}(D)/{\mathbb{R}})\bigr{)}, was defined as the pressure. Below, we also define another time-averaged “pressure”

(2.6) R(t):=P(t)0tξ(s)𝑑W(s),R(t):=P(t)-\int_{0}^{t}\xi(s)\,dW(s),

using the Helmholtz decomposition 𝐁(𝐮(t))=𝜼(t)+ξ(t){\bf B}({\bf u}(t))=\boldsymbol{\eta}(t)+\nabla\xi(t), where ξHper1(D)/\xi\in H^{1}_{per}(D)/\mathbb{R} -a.s.{\mathbb{P}}\mbox{-a.s.} such that

(2.7) (ξ(t),ϕ)=(𝐁(𝐮(t)),ϕ)ϕHper1(D).\bigl{(}\nabla\xi(t),\nabla\phi\bigr{)}=\bigl{(}{\bf B}({\bf u}(t)),\nabla\phi\bigr{)}\qquad\forall\,\phi\in H^{1}_{per}(D)\,.

Then, it is easy to check that -a.s.{\mathbb{P}}\mbox{-a.s.}

(2.8) R(t)=𝐮(t)+0t𝐮(s)𝑑s+𝐮0+0t𝐟(s)𝑑s+0t𝜼(s)𝑑W(s)t(0,T).\nabla R(t)=-\mathbf{u}(t)+\int_{0}^{t}{\bf u}(s)\,ds+\mathbf{u}_{0}+\int_{0}^{t}{\bf f}(s)\,ds+\int_{0}^{t}\boldsymbol{\eta}(s)\,dW(s)\quad\forall\,t\in(0,T).

The process {R(t);0tT}\{R(t);0\leq t\leq T\} will also be approximated in our numerical methods.

Next, we establish some stability estimates for the velocity 𝐮{\bf u} and the pressure PP in the following lemma.

Lemma 2.2.

Suppose that 𝐮0L2(Ω;𝕍){\bf u}_{0}\in L^{2}(\Omega;\mathbb{V}). Let (𝐮,P)({\bf u},P) solve (2.5). Then there exists a constant CC(DT,𝐮0,𝐟)C\equiv C(D_{T},{\bf u}_{0},{\bf f}) such that

(2.9) 𝔼[sup0tT𝐮(t)2]+𝔼[0T𝐀𝐮(s)2𝑑s]\displaystyle\mathbb{E}\Bigl{[}\sup_{0\leq t\leq T}\|\nabla{\bf u}(t)\|^{2}\Bigr{]}+\mathbb{E}\biggl{[}\int_{0}^{T}\|{\bf A}{\bf u}(s)\|^{2}\,ds\biggr{]} C,\displaystyle\leq C,
(2.10) sup0tT𝔼[P(t)2]\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\Bigl{[}\|\nabla P(t)\|^{2}\Bigr{]} C.\displaystyle\leq C.
Proof.

The proof of (2.9) is standard by using Itô’s formula and can be found in [8, Theorem 4.4 and Section 5]. To prove (2.10), using (2.5a) can easily get

(2.11) 𝔼[P(t)2]\displaystyle\mathbb{E}\big{[}\|\nabla P(t)\|^{2}\big{]} C(𝔼[𝐮(t)2]+𝔼[0tΔ𝐮(s)2ds]+𝔼[0t𝐮(s)2ds]\displaystyle\leq C\bigg{(}\mathbb{E}\big{[}\|{\bf u}(t)\|^{2}\big{]}+\mathbb{E}\bigg{[}\int_{0}^{t}\|\Delta{\bf u}(s)\|^{2}\,ds\bigg{]}+\mathbb{E}\bigg{[}\int_{0}^{t}\|{\bf u}(s)\|^{2}\,ds\bigg{]}
+𝔼[0t𝐟(s)2ds]).\displaystyle\qquad+\mathbb{E}\bigg{[}\int_{0}^{t}\|{\bf f}(s)\|^{2}\,ds\bigg{]}\bigg{)}.

Then, the desired estimate from immediately after taking supreme over all t[0,T]t\in[0,T] and using (2.9). ∎

We finish this section by establishing the following Hölder continuity result for PP, which is used for the error analysis in sections.

Lemma 2.3.

Suppose that 𝐮L2(Ω;C([0,T];𝕍))L2(Ω;0,T;𝐇per2(D)){\bf u}\in L^{2}(\Omega;C([0,T];\mathbb{V}))\cap L^{2}(\Omega;0,T;{\bf H}^{2}_{per}(D)), 𝐟L2(Ω;{\bf f}\in L^{2}(\Omega; 0,T;𝐋per2(D))0,T;{\bf L}^{2}_{per}(D)) and PL2(Ω;L2(0,T;Hper1(D/)))P\in L^{2}(\Omega;L^{2}(0,T;H^{1}_{per}(D/\mathbb{R}))). Then, there holds

(2.12) 𝔼[(P(s)P(t))2]C|st|s,t[0,T],\displaystyle\mathbb{E}\big{[}\|\nabla\big{(}P(s)-P(t)\big{)}\|^{2}\big{]}\leq C|s-t|\qquad\forall s,t\in[0,T],

where the constant C>0C>0 depends on DTD_{T}, 𝐮0{\bf u}_{0} and 𝐟{\bf f}.

Proof.

From its definition of P(t)P(t), we get

(2.13) P(t)=𝐮(t)+0tΔ𝐮(s)𝑑s+0t𝐟(s)𝑑s+0t𝐁(𝐮(s))𝑑𝐖(s)\displaystyle\nabla P(t)=-{\bf u}(t)+\int_{0}^{t}\Delta{\bf u}(s)\,ds+\int_{0}^{t}{\bf f}(s)\,ds+\int_{0}^{t}{\bf B}({\bf u}(s))\,d{\bf W}(s)

Therefore, for any s,t[0,T]s,t\in[0,T], we have

(2.14) (P(t)P(s))\displaystyle\nabla\big{(}P(t)-P(s)\big{)} =(𝐮(t)𝐮(s))+stΔ𝐮(τ)𝑑τ+st𝐟(τ)𝑑τ\displaystyle=-\big{(}{\bf u}(t)-{\bf u}(s)\big{)}+\int_{s}^{t}\Delta{\bf u}(\tau)\,d\tau+\int_{s}^{t}{\bf f}(\tau)\,d\tau
+st𝐁(𝐮(τ))𝑑W(τ).\displaystyle\qquad+\int_{s}^{t}{\bf B}({\bf u}(\tau))\,dW(\tau).

Thus,

(2.15) 𝔼[(P(t)P(s))2]\displaystyle\mathbb{E}\big{[}\|\nabla\big{(}P(t)-P(s)\big{)}\|^{2}\big{]} 4𝔼[𝐮(t)𝐮(s)2]+4𝔼[stΔ𝐮(τ)𝑑τ2]\displaystyle\leq 4\mathbb{E}\big{[}\|{\bf u}(t)-{\bf u}(s)\|^{2}\big{]}+4\mathbb{E}\bigg{[}\bigg{\|}\int_{s}^{t}\Delta{\bf u}(\tau)\,d\tau\bigg{\|}^{2}\bigg{]}
+4𝔼[st𝐟(τ)𝑑τ2+st𝐁(𝐮(τ))𝑑W(τ)2]\displaystyle\quad+4\mathbb{E}\bigg{[}\bigg{\|}\int_{s}^{t}{\bf f}(\tau)\,d\tau\bigg{\|}^{2}+\bigg{\|}\int_{s}^{t}{\bf B}({\bf u}(\tau))\,dW(\tau)\bigg{\|}^{2}\bigg{]}
=I+II+III+IV.\displaystyle=\mbox{\tt I}+\mbox{\tt II}+\mbox{\tt III}+\mbox{\tt IV}.

The first term I can be controlled by using (2.5a). For II and III, by Schwarz inequality, we have

(2.16) II+III4𝔼[stΔ𝐮(τ)2𝑑τ]|ts|+4𝔼[st𝐟(τ)2𝑑τ]|ts|.\displaystyle\mbox{\tt II}+\mbox{\tt III}\leq 4\mathbb{E}\bigg{[}\int_{s}^{t}\|\Delta{\bf u}(\tau)\|^{2}\,d\tau\bigg{]}|t-s|+4\mathbb{E}\bigg{[}\int_{s}^{t}\|{\bf f}(\tau)\|^{2}\,d\tau\bigg{]}|t-s|.

In addition, using Itô isometry and (2.2b), we obtain

(2.17) IV=4𝔼[st𝐁(𝐮(τ))2𝑑τ]C𝔼[st𝐮(τ)2𝑑τ].\displaystyle\mbox{\tt IV}=4\mathbb{E}\bigg{[}\int_{s}^{t}\|{\bf B}({\bf u}(\tau))\|^{2}\,d\tau\bigg{]}\leq C\mathbb{E}\bigg{[}\int_{s}^{t}\|{\bf u}(\tau)\|^{2}\,d\tau\bigg{]}.

In summary, we have

(2.18) 𝔼[(P(t)P(s))2]\displaystyle\mathbb{E}\big{[}\|\nabla\big{(}P(t)-P(s)\big{)}\|^{2}\big{]} C|ts|+4𝔼[stΔ𝐮(τ)2𝑑τ]|ts|\displaystyle\leq C|t-s|+4\mathbb{E}\bigg{[}\int_{s}^{t}\|\Delta{\bf u}(\tau)\|^{2}\,d\tau\bigg{]}|t-s|
+4𝔼[st𝐟(τ)2𝑑τ|ts|+Cst𝐮(τ)2𝑑τ]\displaystyle\quad+4\mathbb{E}\bigg{[}\int_{s}^{t}\|{\bf f}(\tau)\|^{2}\,d\tau\,|t-s|+C\int_{s}^{t}\|{\bf u}(\tau)\|^{2}\,d\tau\bigg{]}
(C+4𝔼[stΔ𝐮(τ)2dτ+st𝐟(τ)2dτ]\displaystyle\leq\bigg{(}C+4\mathbb{E}\bigg{[}\int_{s}^{t}\|\Delta{\bf u}(\tau)\|^{2}\,d\tau+\int_{s}^{t}\|{\bf f}(\tau)\|^{2}\,d\tau\bigg{]}
+Cmaxτ[0,T]𝔼[𝐮(τ)2])|ts|.\displaystyle\quad+C\max_{\tau\in[0,T]}\mathbb{E}\big{[}\|{\bf u}(\tau)\|^{2}\big{]}\bigg{)}|t-s|.

Finally, the desired estimate follows from the assumptions on 𝐮{\bf u} and 𝐟{\bf f}. ∎

3. Two Chorin-type time-stepping schemes

In this section, we first formulate two Chorin-type semi-discrete-in-time schemes for problem (1.1). The first scheme is the standard Chorin scheme and the second one is a Helmholtz decomposition enhanced nonstandard Chorin scheme. We then present a complete convergence analysis for each scheme which include establishing their stability and error estimates in strong norms for both velocity and pressure approximations.

3.1. Standard Chorin projection scheme

We first consider the standard Chorin scheme for (1.1), its formulation is a straightforward adaptation of the well-known scheme for the deterministic Stokes problem and is given in Algorithm 1 below. As mentioned earlier, the standard Chorin scheme for (1.1) was already studied in [8] in the special case when the noise is divergence-free and error estimates were only obtained for the velocity approximation. In contrast, here we consider the Chorin scheme for general multiplicative noise and to derive error estimates in strong norms not only for the velocity but also for pressure approximations and to achieve a full understanding about the scheme.

3.1.1. Formulation of the standard Chorin scheme

Let MM be a (large) positive integer and k:=T/Mk:=T/M be the time step size. Set tj=jkt_{j}=jk for j=0,1,2,,Mj=0,1,2,\cdots,M, then {tj}j=0M\{t_{j}\}_{j=0}^{M} forms a uniform mesh for (0,T)(0,T). The standard Chorin projection scheme is given as follows (cf. [8, 21, 14]):

Algorithm 1

Let 𝐮~0=𝐮0=𝐮0\tilde{{\bf u}}^{0}={\bf u}^{0}={\bf u}_{0}. For n=0,1,2,,M1n=0,1,2,\cdots,M-1, do the following three steps.

Step 1: Given 𝐮nL2(Ω;){\bf u}^{n}\in L^{2}(\Omega;\mathbb{H}) and 𝐮~nL2(Ω;𝐇per1(D))\tilde{{\bf u}}^{n}\in L^{2}(\Omega;{\bf H}^{1}_{per}(D)), find 𝐮~n+1L2(Ω;𝐇per1(D))\tilde{{\bf u}}^{n+1}\in L^{2}(\Omega;{\bf H}_{per}^{1}(D)) such that \mathbb{P}-a.s.

(3.1) 𝐮~n+1kΔ𝐮~n+1\displaystyle\tilde{{\bf u}}^{n+1}-k\Delta\tilde{{\bf u}}^{n+1} =𝐮n+k𝐟n+1+𝐁(𝐮~n)ΔWn+1in D.\displaystyle={\bf u}^{n}+k{\bf f}^{n+1}+{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1}\qquad\mbox{in }D.

Step 2: Find pn+1L2(Ω;Hper1(D)/)p^{n+1}\in L^{2}(\Omega;H^{1}_{per}(D)/\mathbb{R}) such that \mathbb{P}-a.s.

(3.2) Δpn+1\displaystyle-\Delta p^{n+1} =1kdiv 𝐮~n+1in D.\displaystyle=-\frac{1}{k}\mbox{\rm div\,}\tilde{{\bf u}}^{n+1}\qquad\mbox{in }D.

Step 3: Define 𝐮n+1L2(Ω;){\bf u}^{n+1}\in L^{2}(\Omega;\mathbb{H}) by

(3.3) 𝐮n+1\displaystyle{\bf u}^{n+1} =𝐮~n+1kpn+1.\displaystyle=\tilde{{\bf u}}^{n+1}-k\nabla p^{n+1}.
Remark 3.1.

(a) The above formulation is written in the way in which the scheme is implemented, it is slightly different from the traditional writing which combines Step 2 and 3 together.

(b) It is easy to check (𝐮~n+1,𝐮n+1,pn+1)(\tilde{{\bf u}}^{n+1},{{\bf u}}^{n+1},p^{n+1}) satisfies the following system:

(3.4a) 𝐮~n+1𝐮~nkΔ𝐮~n+1+kpn\displaystyle\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}-k\Delta\tilde{{\bf u}}^{n+1}+k\nabla p^{n} =k𝐟n+1+𝐁(𝐮~n)ΔWn+1\displaystyle=k{\bf f}^{n+1}+{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1} in D,\displaystyle\qquad\mbox{in }D,
(3.4b) div 𝐮~n+1kΔpn+1\displaystyle\mbox{\rm div\,}\tilde{{\bf u}}^{n+1}-k\Delta p^{n+1} =0\displaystyle=0 in D,\displaystyle\qquad\mbox{in }D,

where u~0=u0\tilde{u}^{0}=u_{0}.

3.1.2. Stability estimates for the standard Chorin method

The goal of this subsection is to establish some stability estimates for Algorithm 1 in strong norms.

Lemma 3.1.

The discrete processes {(𝐮~n,pn)}n=0M\{(\tilde{{\bf u}}^{n},p^{n})\}_{n=0}^{M} defined in (3.4) satisfy

(3.5a) max0nM𝔼[𝐮~n2]+𝔼[n=1M𝐮~n𝐮~n12]+𝔼[kn=0M𝐮~n2]\displaystyle\max_{0\leq n\leq M}\mathbb{E}[\|\tilde{{\bf u}}^{n}\|^{2}]+\mathbb{E}\bigg{[}\sum_{n=1}^{M}\|\tilde{{\bf u}}^{n}-\tilde{{\bf u}}^{n-1}\|^{2}\bigg{]}+\mathbb{E}\bigg{[}k\sum_{n=0}^{M}\|\nabla\tilde{{\bf u}}^{n}\|^{2}\bigg{]} C,\displaystyle\leq C,
(3.5b) 𝔼[kn=0Mpn2]\displaystyle\mathbb{E}\bigg{[}k\sum_{n=0}^{M}\|\nabla p^{n}\|^{2}\bigg{]} Ck,\displaystyle\leq\frac{C}{k},
(3.5c) max0nM𝔼[𝐮~n2]+𝔼[kn=0MΔ𝐮~n2]\displaystyle\max_{0\leq n\leq M}\mathbb{E}[\|\nabla\tilde{{\bf u}}^{n}\|^{2}]+\mathbb{E}\bigg{[}k\sum_{n=0}^{M}\|\Delta\tilde{{\bf u}}^{n}\|^{2}\bigg{]} Ck,\displaystyle\leq\frac{C}{k},

where C>0C>0 depends only on DT,u0,𝐟D_{T},u_{0},{\bf f}.

Proof.

We first prove (3.5a) and (3.5b). Testing (3.4a) by u~n+1\tilde{u}^{n+1} and (3.4b) by pnp^{n}, and integrating by parts, we obtain

(3.6) (𝐮~n+1𝐮~n,𝐮~n+1)+k𝐮~n+12+k(pn,𝐮~n+1)\displaystyle\big{(}\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n},\tilde{{\bf u}}^{n+1}\big{)}+k\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}+k\big{(}\nabla p^{n},\tilde{{\bf u}}^{n+1}\big{)}
=k(𝐟n+1,𝐮~n+1)+(𝐁(𝐮~n)ΔWn+1,𝐮~n+1),\displaystyle\hskip 86.72377pt=k\big{(}{\bf f}^{n+1},\tilde{{\bf u}}^{n+1}\big{)}+\big{(}{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\tilde{{\bf u}}^{n+1}\big{)},
(3.7) (𝐮~n+1,pn)k(pn+1,pn)=0.\displaystyle\big{(}\tilde{{\bf u}}^{n+1},\nabla p^{n}\big{)}-k\big{(}\nabla p^{n+1},\nabla p^{n}\big{)}=0.

Substituting (3.7) into (3.6) yields

(3.8) (𝐮~n+1𝐮~n,𝐮~n+1)\displaystyle\big{(}\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n},\tilde{{\bf u}}^{n+1}\big{)} +k𝐮~n+12+k2(pn+1,pn)\displaystyle+k\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}+k^{2}\big{(}\nabla p^{n+1},\nabla p^{n}\big{)}
=k(𝐟n+1,𝐮~n+1)+(𝐁(𝐮~n)ΔWn+1,𝐮~n+1).\displaystyle=k\big{(}{\bf f}^{n+1},\tilde{{\bf u}}^{n+1}\big{)}+\big{(}{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\tilde{{\bf u}}^{n+1}\big{)}.

Using the identity 2a(ab)=a2b2+(ab)22a(a-b)=a^{2}-b^{2}+(a-b)^{2} and the following identity from (3.4b)

k(pn+1,(pn+1pn))=(𝐮~n+1𝐮~n,pn+1),k\big{(}\nabla p^{n+1},\nabla(p^{n+1}-p^{n})\big{)}=\big{(}\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n},\nabla p^{n+1}\big{)},

and taking expectations on both sides of (3.8), we get

(3.9) 12𝔼[𝐮~n+12𝐮~n2+𝐮~n+1𝐮~n2]+𝔼[k𝐮~n+12+k2pn+12]\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}\|^{2}-\|\tilde{{\bf u}}^{n}\|^{2}+\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+\mathbb{E}\bigl{[}k\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}+k^{2}\|\nabla p^{n+1}\|^{2}\bigr{]}
=k𝔼[(𝐮~n+1𝐮~n,pn+1)]+k𝔼[(𝐟n+1,𝐮~n)]\displaystyle\quad=k\mathbb{E}[\big{(}\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n},\nabla p^{n+1}\big{)}]+k\mathbb{E}\big{[}\big{(}{\bf f}^{n+1},\tilde{{\bf u}}^{n}\big{)}\big{]}
+𝔼[(𝐁(𝐮~n)ΔWn+1,𝐮~n+1)].\displaystyle\qquad+\mathbb{E}\bigl{[}\big{(}{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\tilde{{\bf u}}^{n+1}\big{)}\bigr{]}.

Next, we bound each term on the right-hand side of (3.9). First, by Schwarz and Young’s inequalities, we have

(3.10) k𝔼[(𝐮~n+1𝐮~n,pn+1)]\displaystyle k\mathbb{E}\big{[}\big{(}\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n},\nabla p^{n+1}\big{)}\big{]} 310𝔼[𝐮~n+1𝐮~n2]+56k2𝔼[pn+12],\displaystyle\leq\frac{3}{10}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+\frac{5}{6}k^{2}\mathbb{E}\bigl{[}\|\nabla p^{n+1}\|^{2}\bigr{]},
(3.11) k𝔼[(𝐟n+1,𝐮~n)]\displaystyle k\mathbb{E}\big{[}\big{(}{\bf f}^{n+1},\tilde{{\bf u}}^{n}\big{)}\big{]} Ck𝔼[𝐟n+12]+k4𝔼[𝐮~n+12].\displaystyle\leq Ck\mathbb{E}\big{[}\|{\bf f}^{n+1}\|^{2}\big{]}+\frac{k}{4}\mathbb{E}\big{[}\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}\big{]}.

In addition, by Itô isometry, we have

(3.12) 𝔼[(𝐁(𝐮~n)ΔWn+1,𝐮~n+1)]\displaystyle\mathbb{E}\bigl{[}\big{(}{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\tilde{{\bf u}}^{n+1}\big{)}\bigr{]} =𝔼[(𝐁(𝐮~n)ΔWn+1,𝐮~n+1𝐮~n)]\displaystyle=\mathbb{E}\bigl{[}\big{(}{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\big{)}\bigr{]}
2𝔼[𝐁(𝐮~n)ΔWn+12]+18𝔼[𝐮~n+1𝐮~n2]\displaystyle\leq 2\mathbb{E}\bigl{[}\|{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1}\|^{2}\bigr{]}+\frac{1}{8}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}
=2k𝔼[𝐁(𝐮~n)2]+18𝔼[𝐮~n+1𝐮~n2]\displaystyle=2k\mathbb{E}\bigl{[}\|{\bf B}(\tilde{{\bf u}}^{n})\|^{2}\bigr{]}+\frac{1}{8}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}
Ck𝔼[𝐮~n2]+18𝔼[𝐮~n+1𝐮~n2].\displaystyle\leq Ck\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+\frac{1}{8}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}.

Substituting (3.10)–(3.12) into (3.9) gives

(3.13) 12𝔼[𝐮~n+12𝐮~n2]+120𝔼[𝐮~n+1𝐮~n2]+3k4𝔼[𝐮~n+12]\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}\|^{2}-\|\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+\frac{1}{20}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+\frac{3k}{4}\mathbb{E}\bigl{[}\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}\bigr{]}
+16k2𝔼[pn+12]Ck𝔼[𝐟n+12]+Ck𝔼[𝐮~n2].\displaystyle\qquad\qquad\qquad+\frac{1}{6}k^{2}\mathbb{E}\bigl{[}\|\nabla p^{n+1}\|^{2}\bigr{]}\leq Ck\mathbb{E}\big{[}\|{\bf f}^{n+1}\|^{2}\big{]}+Ck\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n}\|^{2}\bigr{]}.

Applying the summation operator n=0m\sum_{n=0}^{m} to (3.13) for any 0mM10\leq m\leq M-1 yields

(3.14) 12𝔼[𝐮~m+12𝐮~02]+120n=0m𝔼[𝐮~n+1𝐮~n2]+kn=0m𝔼[𝐮~n+12]\displaystyle\frac{1}{2}\mathbb{E}\big{[}\|\tilde{{\bf u}}^{m+1}\|^{2}-\|\tilde{{\bf u}}^{0}\|^{2}\big{]}+\frac{1}{20}\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n}\|^{2}\bigr{]}+k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}\bigr{]}
+16k2n=0m𝔼[pn+12]Ckn=0m𝔼[u~nL22]+Ckn=0m𝔼[𝐟n+12].\displaystyle\qquad+\frac{1}{6}k^{2}\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\nabla p^{n+1}\|^{2}\bigr{]}\leq Ck\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\tilde{u}^{n}\|_{L^{2}}^{2}\bigr{]}+Ck\sum_{n=0}^{m}\mathbb{E}\big{[}\|{\bf f}^{n+1}\|^{2}\big{]}.

Finally, by the discrete Gronwall’s inequality we obtain

12max1nM𝔼[𝐮~n2]\displaystyle\frac{1}{2}\max_{1\leq n\leq M}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n}\|^{2}\bigr{]} +120n=1M𝔼[𝐮~n𝐮~n12]+3k4n=1M𝔼[𝐮~n2]\displaystyle+\frac{1}{20}\sum_{n=1}^{M}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{n}-\tilde{{\bf u}}^{n-1}\|^{2}\bigr{]}+\frac{3k}{4}\sum_{n=1}^{M}\mathbb{E}\bigl{[}\|\nabla\tilde{{\bf u}}^{n}\|^{2}\bigr{]}
+16k2n=1M𝔼[pn2]C(𝔼[𝐮0L22]+kn=0m𝔼[𝐟n+12]),\displaystyle+\frac{1}{6}k^{2}\sum_{n=1}^{M}\mathbb{E}\bigl{[}\|\nabla p^{n}\|^{2}\bigr{]}\leq C\bigg{(}\mathbb{E}\bigl{[}\|{\bf u}_{0}\|_{L^{2}}^{2}\bigr{]}+k\sum_{n=0}^{m}\mathbb{E}\big{[}\|{\bf f}^{n+1}\|^{2}\big{]}\bigg{)},

which implies the desired estimate (3.5a) and (3.5b).

It remains to show (3.5c). To the end, testing (3.4a) by Δ𝐮~n+1-\Delta\tilde{{\bf u}}^{n+1} we get

(3.15) 12[𝐮~n+12𝐮~n2+(𝐮~n+1𝐮~n)2]+kΔ𝐮~n+12\displaystyle\frac{1}{2}\bigl{[}\|\nabla\tilde{{\bf u}}^{n+1}\|^{2}-\|\nabla\tilde{{\bf u}}^{n}\|^{2}+\|\nabla(\tilde{{\bf u}}^{n+1}-\tilde{{\bf u}}^{n})\|^{2}\bigr{]}+k\|\Delta\tilde{{\bf u}}^{n+1}\|^{2}
=k(pn,Δ𝐮~n+1)k(𝐟n+1,Δ𝐮~n+1)+(𝐁(𝐮~n)ΔWn+1,𝐮~n+1).\displaystyle\qquad=k\bigl{(}\nabla p^{n},\Delta\tilde{{\bf u}}^{n+1}\bigr{)}-k\bigl{(}{\bf f}^{n+1},\Delta\tilde{{\bf u}}^{n+1}\bigr{)}+\bigl{(}\nabla{\bf B}(\tilde{{\bf u}}^{n})\Delta W_{n+1},\nabla\tilde{{\bf u}}^{n+1}\bigr{)}.

Proceeding with similar arguments as used above yields

(3.16) 12𝔼[𝐮~m+12]+12kn=0m𝔼[Δ𝐮~n+12]\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|\nabla\tilde{{\bf u}}^{m+1}\|^{2}\bigr{]}+\frac{1}{2}k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\Delta\tilde{{\bf u}}^{n+1}\|^{2}\bigr{]}
12𝔼[𝐮02]+kn=0m𝔼[pn2+𝐟n+12+C𝐮~n2].\displaystyle\hskip 57.81621pt\leq\frac{1}{2}\mathbb{E}[\|\nabla{\bf u}_{0}\|^{2}]+k\sum_{n=0}^{m}\mathbb{E}\Bigl{[}\|\nabla p^{n}\|^{2}+\|{\bf f}^{n+1}\|^{2}+C\|\nabla\tilde{{\bf u}}^{n}\|^{2}\Bigr{]}.

Then by the discrete Gronwall inequality we get

12𝔼[𝐮~m+12]+12kn=0m𝔼[Δ𝐮~n+12]\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|\nabla\tilde{{\bf u}}^{m+1}\|^{2}\bigr{]}+\frac{1}{2}k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\Delta\tilde{{\bf u}}^{n+1}\|^{2}\bigr{]}
(12𝔼[𝐮02]+kn=0m𝔼[pn2]+kn=0m𝔼[𝐟n+12])exp(Ctm),\displaystyle\qquad\leq\Bigl{(}\frac{1}{2}\mathbb{E}[\|\nabla{\bf u}_{0}\|^{2}]+k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|\nabla p^{n}\|^{2}\bigr{]}+k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|{\bf f}^{n+1}\|^{2}\bigr{]}\Bigr{)}\exp(Ct_{m}),

which and (3.5b) immediately infer (3.5c). The proof is complete. ∎

3.1.3. Error estimates for the standard Chorin scheme

In this subsection we shall derive some error estimates for the time-discrete processes generated by Algorithm 1. To the best of our knowledge, these are the first error estimate results for the standard Chorin scheme in the case general multiplicative noises. For the sake of brevity, but without loss of the generality, we set 𝐟=0{\bf f}=0 in this subsection.

First, we state the following error estimate result for the velocity.

Theorem 3.1.

Let {(𝐮~n,pn)}n=0M\{(\tilde{{\bf u}}^{n},p^{n})\}_{n=0}^{M} be generated by Algorithm 1, then there exists a positive constant CC which depends on DT,𝐮0,D_{T},{\bf u}_{0}, and 𝐟{\bf f} such that

(3.17) (𝔼[kn=0M𝐮(tn)\displaystyle\biggl{(}\mathbb{E}\biggl{[}k\sum_{n=0}^{M}\|{\bf u}(t_{n}) 𝐮~n2])12\displaystyle-\tilde{{\bf u}}^{n}\|^{2}\biggr{]}\biggr{)}^{\frac{1}{2}}
+max0M(𝔼[kn=0(𝐮(tn)𝐮~n)2])12Ck14.\displaystyle+\max_{0\leq\ell\leq M}\biggl{(}\mathbb{E}\biggl{[}\Bigl{\|}k\sum_{n=0}^{\ell}\nabla({\bf u}(t_{n})-\tilde{{\bf u}}^{n})\Bigr{\|}^{2}\biggr{]}\biggr{)}^{\frac{1}{2}}\leq Ck^{\frac{1}{4}}.
Proof.

Let 𝐞𝐮~m=𝐮(tm)𝐮~m{\bf e}_{\tilde{{\bf u}}}^{m}={\bf u}(t_{m})-\tilde{{\bf u}}^{m} and pm=P(tm)kn=0mpn\displaystyle\mathcal{E}_{p}^{m}=P(t_{m})-k\sum_{n=0}^{m}p^{n}. Obviously, 𝐞𝐮~mL2(Ω;𝐇per1(D)){\bf e}_{\tilde{{\bf u}}}^{m}\in L^{2}\big{(}\Omega;{\bf H}^{1}_{per}(D)\big{)} and pmL2(Ω;Hper1(D)/)\mathcal{E}_{p}^{m}\in L^{2}(\Omega;H^{1}_{per}(D)/\mathbb{R}). In addition, from (2.5a), we have

(3.18) (𝐮(tm+1),𝐯)+n=0mtntn+1(𝐮(s),𝐯)𝑑s+(𝐯,P(tm+1))\displaystyle\big{(}{\bf u}(t_{m+1}),{\bf v}\big{)}+\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla{\bf u}(s),\nabla{\bf v}\big{)}\,ds+\big{(}{\bf v},\nabla P(t_{m+1})\big{)}
=(𝐮0,𝐯)+(n=0mtntn+1(𝐁(𝐮(s))dW(s),𝐯)𝐯𝐇per1(D)a.s.\displaystyle\quad=\bigl{(}{\bf u}_{0},{\bf v}\bigr{)}+\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}{\bf B}({\bf u}(s))\,dW(s),{\bf v}\biggr{)}\quad\forall{\bf v}\in{\bf H}^{1}_{per}(D)\,\,a.s.

Applying the summation operator n=0m\sum_{n=0}^{m} to (3.4a) yields

(3.19) (𝐮~m+1,𝐯)\displaystyle\bigl{(}\tilde{{\bf u}}^{m+1},{\bf v}\bigr{)} +kn=0m(𝐮~n+1,𝐯)+(kn=0mpn,𝐯)\displaystyle+k\sum_{n=0}^{m}\bigl{(}\nabla\tilde{{\bf u}}^{n+1},\nabla{\bf v}\bigr{)}+\Bigl{(}k\sum_{n=0}^{m}\nabla p^{n},{\bf v}\Bigr{)}
=(𝐮~0,𝐯)+(n=0mtntn+1𝐁(𝐮~n)𝑑W(s),𝐯).\displaystyle=\bigl{(}\tilde{{\bf u}}_{0},{\bf v}\bigr{)}+\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}{\bf B}(\tilde{{\bf u}}^{n})\,dW(s),{\bf v}\biggr{)}.

Subtracting (3.19) from (3.18) we get

(3.20) (𝐞𝐮~m+1,𝐯)+k(n=0m𝐞𝐮~n+1,𝐯)+(pm,𝐯)\displaystyle\big{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},{\bf v}\big{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1},\nabla{\bf v}\Bigr{)}+\big{(}\nabla\mathcal{E}_{p}^{m},{\bf v}\big{)}
=n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐯)𝑑s((P(tm+1)P(tm)),𝐯)\displaystyle\quad=\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf v}\big{)}\,ds-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),{\bf v}\bigr{)}
+(n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s),𝐯)𝐯𝐇per1(D).\displaystyle\qquad+\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s),{\bf v}\biggr{)}\quad\forall{\bf v}\in{\bf H}_{per}^{1}(D).

Setting 𝐯=𝐞𝐮~m+1{\bf v}={\bf e}_{\tilde{{\bf u}}}^{m+1} in (3.20) we obtain

(3.21) 𝐞𝐮~m+12+k(n=0m+1𝐞𝐮~n,𝐞𝐮~m+1)+(pm,𝐞𝐮~m+1)\displaystyle\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{n},\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\Bigr{)}+\big{(}\nabla\mathcal{E}_{p}^{m},{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}
=n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s\displaystyle\quad=\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds
((P(tm+1)P(tm)),𝐞𝐮~m+1)\displaystyle\qquad-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),{\bf e}_{\tilde{{\bf u}}}^{m+1}\bigr{)}
+(n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s),𝐞𝐮~m+1).\displaystyle\qquad+\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s),{\bf e}_{\tilde{{\bf u}}}^{m+1}\biggr{)}.

Similarly, by (2.5b) and (3.4b) we get

(3.22) div 𝐞𝐮~n+kΔpn=0.\displaystyle\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}}^{n}+k\Delta p^{n}=0.

Applying the summation n=0m+1\sum_{n=0}^{m+1} to (3.22) and then adding ±ΔP(tm+1)\pm\Delta P(t_{m+1}) yield

div (n=0m+1𝐞𝐮~n)Δpm+1=ΔP(tm+1).\displaystyle\mbox{\rm div\,}\biggl{(}\sum_{n=0}^{m+1}{\bf e}_{\tilde{{\bf u}}}^{n}\biggr{)}-\Delta\mathcal{E}_{p}^{m+1}=-\Delta P(t_{m+1}).

Therefore,

(3.23) div 𝐞𝐮~m+1Δ(pm+1pm)=Δ(P(tm+1)P(tm)).\displaystyle\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}}^{m+1}-\Delta(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})=-\Delta(P(t_{m+1})-P(t_{m})).

Testing (3.23) by any qL2(Ω;Hper1(D)/)q\in L^{2}(\Omega;H^{1}_{per}(D)/\mathbb{R}), we have

(3.24) (𝐞𝐮~m+1,q)=((pm+1pm),q)((P(tm+1)P(tm)),q).\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla q\bigr{)}=\bigl{(}\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m}),\nabla q\bigr{)}-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla q\bigr{)}.

Choosing q=pmq=\mathcal{E}_{p}^{m} in (3.24) gives

(3.25) (𝐞𝐮~m+1,pm)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla\mathcal{E}_{p}^{m}\bigr{)} =((pm+1pm),pm)((P(tm+1)P(tm)),pm)\displaystyle=\bigl{(}\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m}),\nabla\mathcal{E}_{p}^{m}\bigr{)}-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}
=((pm+1pm),pm+1)(pm+1pm)2\displaystyle=\bigl{(}\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m}),\nabla\mathcal{E}_{p}^{m+1}\bigr{)}-\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}
((P(tm+1)P(tm)),pm).\displaystyle\qquad\qquad-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}.

Substituting (3.25) into (3.21) we obtain

𝐞𝐮~m+12+k(n=0m+1𝐞𝐮~n,𝐞𝐮~m+1)+((pm+1pm),pm+1)\displaystyle\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}+k\biggl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{n},\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\biggr{)}+\bigl{(}\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m}),\nabla\mathcal{E}_{p}^{m+1}\bigr{)}
=n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s\displaystyle\quad=\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds
+(pm+1pm)2+((P(tm+1)P(tm)),pm)\displaystyle\qquad+\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}+\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}
((P(tm+1)P(tm)),𝐞𝐮~m+1)\displaystyle\qquad-\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),{\bf e}_{\tilde{{\bf u}}}^{m+1}\bigr{)}
+(n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s),𝐞𝐮~m+1)\displaystyle\qquad+\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s),{\bf e}_{\tilde{{\bf u}}}^{m+1}\biggr{)}
n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s\displaystyle\quad\leq\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds
+(pm+1pm)2+((P(tm+1)P(tm)),pm)\displaystyle\qquad+\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}+\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}
+(P(tm+1)P(tm))2+14𝐞𝐮~m+12\displaystyle\qquad+\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}+\frac{1}{4}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}
+n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2+14𝐞𝐮~m+12.\displaystyle\qquad+\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}^{2}+\frac{1}{4}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}.

Therefore,

(3.26) 12𝐞𝐮~m+12+k(n=0m+1𝐞𝐮~n,𝐞𝐮~m+1)+((pm+1pm),pm+1)\displaystyle\frac{1}{2}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}+k\biggl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{n},\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\biggr{)}+\bigl{(}\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m}),\nabla\mathcal{E}_{p}^{m+1}\bigr{)}
n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s\displaystyle\quad\leq\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds
+(P(tm+1)P(tm))2+(pm+1pm)2\displaystyle\qquad+\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}+\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}
+((P(tm+1)P(tm)),pm)\displaystyle\qquad+\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}
+n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2.\displaystyle\qquad+\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}^{2}.

Using the identity 2a(ab)=a2b2+(ab)22a(a-b)=a^{2}-b^{2}+(a-b)^{2} in (3.26) yields

(3.27) 12𝐞𝐮~m+12\displaystyle\frac{1}{2}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2} +k2[n=0m+1𝐞𝐮~n2n=0m𝐞𝐮~n2+𝐞𝐮~m+12]\displaystyle+\frac{k}{2}\biggl{[}\biggl{\|}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{n}\biggr{\|}^{2}-\biggl{\|}\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n}\biggr{\|}^{2}+\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\biggr{]}
+12[pm+12pm2+(pm+1pm)2]\displaystyle+\frac{1}{2}\bigl{[}\|\nabla\mathcal{E}_{p}^{m+1}\|^{2}-\|\nabla\mathcal{E}_{p}^{m}\|^{2}+\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\bigr{]}
n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s\displaystyle\leq\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds
+(P(tm+1)P(tm))2+(pm+1pm)2\displaystyle\qquad+\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}+\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}
+((P(tm+1)P(tm)),pm)\displaystyle\qquad+\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}
+n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2.\displaystyle\qquad+\Bigl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\Bigr{\|}^{2}.

Next, we apply the summation operator km=0k\sum_{m=0}^{\ell} for 0M10\leq\ell\leq M-1, followed by applying the expectation operator 𝔼[]\mathbb{E}[\cdot], to (3.27) to obtain

(3.28) km=0𝔼[𝐞𝐮~m+12]+𝔼[km=0+1𝐞𝐮~m2]+k2m=0𝔼[𝐞𝐮~m+12]\displaystyle k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigr{]}+\mathbb{E}\biggl{[}\biggl{\|}k\sum_{m=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{m}\biggr{\|}^{2}\biggr{]}+k^{2}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigr{]}
+k𝔼[p+12]+km=0𝔼[(pm+1pm)2]\displaystyle\qquad+k\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}+k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\bigr{]}
2𝔼[km=0n=0mtntn+1((𝐮(tn+1)𝐮(s)),𝐞𝐮~m+1)𝑑s]\displaystyle\quad\leq 2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\big{(}\nabla({\bf u}(t_{n+1})-{\bf u}(s)),\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\big{)}\,ds\biggr{]}
+2𝔼[km=0(P(tm+1)P(tm))2]\displaystyle\qquad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\biggr{]}
+2𝔼[km=0(pm+1pm)2]\displaystyle\qquad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\biggr{]}
+2𝔼[km=0((P(tm+1)P(tm)),pm)]\displaystyle\qquad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}\biggr{]}
+2𝔼[km=0n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2]\displaystyle\qquad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\Bigr{\|}^{2}\biggr{]}
:=𝙸+𝙸𝙸+𝙸𝙸𝙸+𝙸𝚅+𝚅.\displaystyle:={\tt I+II+III+IV+V}.

Now we estimate each term on the right-hand side of (3.28) as follows.

By using the discrete and continuous Hölder inequality estimates (2.4b), (2.9) and (3.5a), we obtain

(3.29) 𝙸\displaystyle{\tt I} =2𝔼[km=0(n=0mtntn+1(𝐮(tn+1)𝐮(s))ds,𝐞𝐮~m+1)]\displaystyle=2\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds,\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\Bigr{)}\Bigr{]}
2𝔼[km=0n=0mtntn+1(𝐮(tn+1)𝐮(s))ds𝐞𝐮~m+1]\displaystyle\leq 2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}\,\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|\biggr{]}
2𝔼[(km=0n=0mtntn+1(𝐮(tn+1)𝐮(s))ds2)12(km=0𝐞𝐮~n2)12]\displaystyle\leq 2\mathbb{E}\biggl{[}\Bigl{(}k\sum_{m=0}^{\ell}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}^{2}\Bigr{)}^{\frac{1}{2}}\,\Bigl{(}k\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{n}\|^{2}\Bigr{)}^{\frac{1}{2}}\biggr{]}
C𝔼[(km=0n=0mtntn+1(𝐮(tn+1)𝐮(s))2𝑑s)12(km=0𝐞𝐮~n2)12]\displaystyle\leq C\mathbb{E}\biggl{[}\Bigl{(}k\sum_{m=0}^{\ell}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|\nabla({\bf u}(t_{n+1})-{\bf u}(s))\|^{2}\,ds\Bigr{)}^{\frac{1}{2}}\,\Bigl{(}k\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{n}\|^{2}\Bigr{)}^{\frac{1}{2}}\biggr{]}
C(𝔼[km=0n=0mtntn+1(𝐮(tn+1)𝐮(s))2𝑑s])12(𝔼[km=0𝐞𝐮~n2])\displaystyle\leq C\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|\nabla({\bf u}(t_{n+1})-{\bf u}(s))\|^{2}\,ds\Bigr{]}\Bigr{)}^{\frac{1}{2}}\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{n}\|^{2}\Bigr{]}\Bigr{)}
Ck12.\displaystyle\leq Ck^{\frac{1}{2}}.

Next, by using (2.12) we have

(3.30) 𝙸𝙸\displaystyle{\tt II} =2𝔼[km=0(P(tm+1)P(tm))2]Ck.\displaystyle=2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\biggr{]}\leq Ck.

By using (2.12) and the stability estimate (3.5b) we obtain

(3.31) 𝙸𝙸𝙸\displaystyle{\tt III} =2𝔼[km=0(pm+1pm)2]\displaystyle=2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\biggr{]}
=2𝔼[km=0(P(tm+1)P(tm))kpm+12]\displaystyle=2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(P(t_{m+1})-P(t_{m}))-k\nabla p^{m+1}\|^{2}\biggr{]}
C𝔼[km=0(P(tm+1)P(tm))2]+Ck𝔼[k2m=0pm+12]\displaystyle\leq C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\biggr{]}+Ck\mathbb{E}\biggl{[}k^{2}\sum_{m=0}^{\ell}\|\nabla p^{m+1}\|^{2}\biggr{]}
Ck.\displaystyle\leq Ck.

It follows from the Itô isometry and (2.4a) that

(3.32) 𝚅\displaystyle{\tt V} =2km=0𝔼[n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2]\displaystyle=2k\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\bigl{(}{\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n})\bigr{)}\,dW(s)\biggr{\|}^{2}\biggr{]}
=2km=0𝔼[n=0mtntn+1𝐁(𝐮(s))𝐁(𝐮~n)2𝑑s]\displaystyle=2k\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|{\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n})\|^{2}\,ds\biggr{]}
Ckm=0n=0mtntn+1𝔼[𝐮(s)𝐮~n2]𝑑s\displaystyle\leq Ck\sum_{m=0}^{\ell}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\mathbb{E}[\|{\bf u}(s)-\tilde{{\bf u}}^{n}\|^{2}]\,ds
Ckm=0kn=0m𝔼[𝐞𝐮n2]+Ckm=0n=0mtntn+1𝔼[𝐮(s)𝐮(tn)2]𝑑s\displaystyle\leq Ck\sum_{m=0}^{\ell}k\sum_{n=0}^{m}\mathbb{E}[\|{\bf e}_{{\bf u}}^{n}\|^{2}]+Ck\sum_{m=0}^{\ell}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\mathbb{E}[\|{\bf u}(s)-{\bf u}(t_{n})\|^{2}]\,ds
Ckm=0kn=0m𝔼[𝐞𝐮n2]+Ck.\displaystyle\leq Ck\sum_{m=0}^{\ell}k\sum_{n=0}^{m}\mathbb{E}[\|{\bf e}_{{\bf u}}^{n}\|^{2}]+Ck.

To bound term IV, we first derive its rewriting as follows:

(3.33) 𝙸𝚅\displaystyle{\tt IV} =2𝔼[km=0((P(tm+1)P(tm)),pm)]\displaystyle=2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m}\bigr{)}\biggr{]}
=2𝔼[km=0((P(tm+1)P(tm)),pm+1)]\displaystyle=2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m+1}\bigr{)}\biggr{]}
+2𝔼[km=0((P(tm+1)P(tm)),(pmpm+1))].\displaystyle\quad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla(\mathcal{E}_{p}^{m}-\mathcal{E}_{p}^{m+1})\bigr{)}\biggr{]}.

By using the summation by parts, the first term above can be rewritten as

(3.34) 2𝔼[km=0((P(tm+1)P(tm)),pm+1)]\displaystyle 2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla\mathcal{E}_{p}^{m+1}\bigr{)}\biggr{]}
=2k𝔼[(P(t+1),p+1)]2k𝔼[m=0(P(tm),(pm+1pm))].\displaystyle\quad=2k\mathbb{E}\bigl{[}\bigl{(}\nabla P(t_{\ell+1}),\nabla\mathcal{E}_{p}^{\ell+1}\bigr{)}\bigr{]}-2k\mathbb{E}\biggl{[}\sum_{m=0}^{\ell}\bigl{(}\nabla P(t_{m}),\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\bigr{)}\biggr{]}.

Substituting (3.34) into (3.33) yields

(3.35) 𝙸𝚅\displaystyle{\tt IV} =2k𝔼[(P(t+1),p+1)]2k𝔼[m=0(P(tm),(pm+1pm))]\displaystyle=2k\mathbb{E}\bigl{[}\bigl{(}\nabla P(t_{\ell+1}),\nabla\mathcal{E}_{p}^{\ell+1}\bigr{)}\bigr{]}-2k\mathbb{E}\biggl{[}\sum_{m=0}^{\ell}\bigl{(}\nabla P(t_{m}),\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\bigr{)}\biggr{]}
+2𝔼[km=0((P(tm+1)P(tm)),(pmpm+1))]\displaystyle\quad+2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),\nabla(\mathcal{E}_{p}^{m}-\mathcal{E}_{p}^{m+1})\bigr{)}\biggr{]}
:=𝙸𝚅𝟷+𝙸𝚅𝟸+𝙸𝚅𝟹.\displaystyle:={\tt IV_{1}+IV_{2}+IV_{3}}.

We now bound each term above. Using the stability (2.10) we get

(3.36) 𝙸𝚅𝟷\displaystyle{\tt IV_{1}} Ck𝔼[P(t+1)2]+k4𝔼[p+12]\displaystyle\leq Ck\mathbb{E}\bigl{[}\|\nabla P(t_{\ell+1})\|^{2}\bigr{]}+\frac{k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}
Ck+k4𝔼[p+12].\displaystyle\leq Ck+\frac{k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}.

Expectedly, the term k4𝔼[p+12]\frac{k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]} will be absorbed to the left side of (3.28) later.

To bound term 𝙸𝚅𝟸{\tt IV_{2}}, we reuse the estimation from III in (3.31) together with the stability of PP given in (2.10) to get

(3.37) 𝙸𝚅𝟸\displaystyle{\tt IV_{2}} 2𝔼[km=0P(tm)(pm+1pm)]\displaystyle\leq 2\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla P(t_{m})\|\,\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|\biggr{]}
C𝔼[(km=0P(tm)2)12(km=0(pm+1pm)2)12]\displaystyle\leq C\mathbb{E}\biggl{[}\Bigl{(}k\sum_{m=0}^{\ell}\|\nabla P(t_{m})\|^{2}\Bigr{)}^{\frac{1}{2}}\,\Bigl{(}k\sum_{m=0}^{\ell}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\Bigr{)}^{\frac{1}{2}}\biggr{]}
C(𝔼[km=0P(tm)2])12(𝔼[km=0(pm+1pm)2])12\displaystyle\leq C\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\nabla P(t_{m})\|^{2}\Bigr{]}\Bigr{)}^{\frac{1}{2}}\,\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\Bigr{]}\Bigr{)}^{\frac{1}{2}}
Ck12.\displaystyle\leq Ck^{\frac{1}{2}}.

Using again (3.31) and (2.12) we have

(3.38) 𝙸𝚅𝟹\displaystyle{\tt IV_{3}} Ckm=0(P(tm+1)P(tm))2+C𝔼[km=0(pm+1pm)2]\displaystyle\leq Ck\sum_{m=0}^{\ell}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}+C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\biggr{]}
Ck.\displaystyle\leq Ck.

Substituting estimates (3.36)–(3.38) into (3.35) yields

(3.39) 𝙸𝚅Ck12+k4𝔼[p+12].\displaystyle{\tt IV}\leq Ck^{\frac{1}{2}}+\frac{k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}.

Now, substituting the estimates for I, II, III, IV, V into (3.28) and using the notation X=km=0𝔼[𝐞𝐮~m2]X^{\ell}=k\sum_{m=0}^{\ell}\mathbb{E}[\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}] we obtain

X+1+𝔼[km=0+1𝐞𝐮~m2]+k2m=0𝔼[𝐞𝐮~m+12]\displaystyle X^{\ell+1}+\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{m=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{m}\Bigr{\|}^{2}\Bigr{]}+k^{2}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigr{]}
+3k4𝔼[p+12]+km=0𝔼[(pm+1pm)2]\displaystyle\quad+\frac{3k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}+k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\bigr{]}
Ck12+Ckm=0Xm.\displaystyle\leq Ck^{\frac{1}{2}}+Ck\sum_{m=0}^{\ell}X^{m}.

Thus, it follows from the discrete Gronwall inequality that

X+1+𝔼[km=0+1𝐞𝐮~m2]+k2m=0𝔼[𝐞𝐮~m+12]\displaystyle X^{\ell+1}+\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{m=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}}^{m}\Bigr{\|}^{2}\Bigr{]}+k^{2}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigr{]}
+3k4𝔼[p+12]+km=0𝔼[(pm+1pm)2]\displaystyle\quad+\frac{3k}{4}\mathbb{E}\bigl{[}\|\nabla\mathcal{E}_{p}^{\ell+1}\|^{2}\bigr{]}+k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(\mathcal{E}_{p}^{m+1}-\mathcal{E}_{p}^{m})\|^{2}\bigr{]}
Ck12exp(Ct),\displaystyle\leq Ck^{\frac{1}{2}}\exp(Ct_{\ell}),

which yields the desired error estimate for the velocity approximation. ∎

Next, we derive an error estimate for the pressure approximation. Our main idea is to estimate the pressure error in a time averaged fashion.

Theorem 3.2.

Let {(𝐮~m,pm)}m=0M\{(\tilde{{\bf u}}^{m},p^{m})\}_{m=0}^{M} be generated by Algorithm 1. Then, there exists a positive constant CC which depends on DT,𝐮0,𝐟,D_{T},{\bf u}_{0},{\bf f}, and β\beta such that

(3.40) (𝔼[km=0MP(tm)kn=0mpn2])12Ck14,\displaystyle\bigg{(}\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\Bigl{\|}P(t_{m})-k\sum_{n=0}^{m}p^{n}\Bigr{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq Ck^{\frac{1}{4}},

where β\beta denotes the stochastic inf-sup constant (see below).

Proof.

We first recall the following inf-sup condition (cf. [6]):

(3.41) sup𝐯𝐇per1(D)(q,div 𝐯)𝐯βqqLper2(D)/,\displaystyle\sup_{{\bf v}\in{\bf H}^{1}_{per}(D)}\frac{\bigl{(}q,\mbox{\rm div\,}{\bf v}\bigr{)}}{\|\nabla{\bf v}\|}\geq\beta\,\|q\|\qquad\forall q\in L^{2}_{per}(D)/\mathbb{R},

where β\beta is a positive constant.

Below we reuse all the notations from Theorem 3.1. From the error equation (3.20) we obtain for all 𝐯𝐇per1(D){\bf v}\in{\bf H}^{1}_{per}(D)

(3.42) (pm,div 𝐯)\displaystyle\bigl{(}\mathcal{E}_{p}^{m},\mbox{\rm div\,}{\bf v}\bigr{)} =(𝐞𝐮~m+1,𝐯)+(kn=0m𝐞𝐮~n+1,𝐯)\displaystyle=\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},{\bf v}\bigr{)}+\biggl{(}k\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1},\nabla{\bf v}\biggr{)}
(n=0mtntn+1(𝐮(tn+1)𝐮(s))ds,𝐯)\displaystyle\quad-\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds,\nabla{\bf v}\biggr{)}
+((P(tm+1)P(tm)),𝐯)\displaystyle\quad+\bigl{(}\nabla(P(t_{m+1})-P(t_{m})),{\bf v}\bigr{)}
(n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s),𝐯).\displaystyle\quad-\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s),{\bf v}\biggr{)}.

By using Schwarz and Poincaré inequalities on the right side of (3.42), we get

(3.43) (pm,div 𝐯)\displaystyle\bigl{(}\mathcal{E}_{p}^{m},\mbox{\rm div\,}{\bf v}\bigr{)} C𝐞𝐮~m+1𝐯+kn=0m𝐞𝐮~n+1𝐯\displaystyle\leq C\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|\|\nabla{\bf v}\|+\biggl{\|}k\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1}\biggr{\|}\|\nabla{\bf v}\|
+n=0mtntn+1(𝐮(tn+1)𝐮(s))ds𝐯\displaystyle\quad+\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}\|\nabla{\bf v}\|
+C(P(tm+1)P(tm))𝐯\displaystyle\quad+C\|\nabla(P(t_{m+1})-P(t_{m}))\|\|\nabla{\bf v}\|
+Cn=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)𝐯\displaystyle\quad+C\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}\|\nabla{\bf v}\|

Applying (3.41) yields

(3.44) βpm\displaystyle\beta\|\mathcal{E}_{p}^{m}\| sup𝐯𝐇per1(D)(pm,div 𝐯)𝐯\displaystyle\leq\sup_{{\bf v}\in{\bf H}^{1}_{per}(D)}\frac{\bigl{(}\mathcal{E}_{p}^{m},\mbox{\rm div\,}{\bf v}\bigr{)}}{\|\nabla{\bf v}\|}
C𝐞𝐮~m+1+kn=0m𝐞𝐮~n+1\displaystyle\leq C\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|+\biggl{\|}k\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1}\biggr{\|}
+n=0mtntn+1(𝐮(tn+1)𝐮(s))ds\displaystyle\quad+\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}
+C(P(tm+1)P(tm))\displaystyle\quad+C\|\nabla(P(t_{m+1})-P(t_{m}))\|
+Cn=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s).\displaystyle\quad+C\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}.

Next, squaring both sides of (3.44) followed by applying operators km=0k\sum_{m=0}^{\ell} and 𝔼[]\mathbb{E}[\cdot], we obtain

(3.45) β2𝔼[km=0pm2]\displaystyle\beta^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\mathcal{E}_{p}^{m}\|^{2}\biggr{]} C𝔼[km=0𝐞𝐮~m+12]+Ckm=0𝔼[kn=0m𝐞𝐮~n+12]\displaystyle\leq C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\biggr{]}+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1}\biggr{\|}^{2}\biggr{]}
+Ckm=0𝔼[n=0mtntn+1(𝐮(tn+1)𝐮(s))ds2]\displaystyle\,\,+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}^{2}\biggr{]}
+Ckm=0𝔼[(P(tm+1)P(tm))2]\displaystyle\,\,+Ck\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\bigr{]}
+Ckm=0𝔼[n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2]\displaystyle\,\,+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}^{2}\biggr{]}
:=𝙸+𝙸𝙸+𝙸𝙸𝙸+𝙸𝚅+𝚅.\displaystyle\,\,:={\tt I+II+III+IV+V}.

We now bound each term above as follows. Using Theorem 3.1 we get

(3.46) 𝙸+𝙸𝙸\displaystyle{\tt I+II} =C𝔼[km=0𝐞𝐮~m+12]+Ckm=0𝔼[kn=0m𝐞𝐮~n+12]\displaystyle=C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\biggr{]}+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}}^{n+1}\biggr{\|}^{2}\biggr{]}
Ck12.\displaystyle\leq Ck^{\frac{1}{2}}.

By the Hölder continuities given in (2.4b) and (2.12), we have

(3.47) 𝙸𝙸𝙸+𝙸𝚅\displaystyle{\tt III+IV} =Ckm=0𝔼[n=0mtntn+1(𝐮(tn+1)𝐮(s))ds2]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\nabla({\bf u}(t_{n+1})-{\bf u}(s))\,ds\biggr{\|}^{2}\biggr{]}
+Ckm=0𝔼[(P(tm+1)P(tm))2]\displaystyle\qquad+Ck\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\bigr{]}
Ckm=0𝔼[n=0mtntn+1(𝐮(tn+1)𝐮(s))2𝑑s]\displaystyle\leq Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|\nabla({\bf u}(t_{n+1})-{\bf u}(s))\|^{2}\,ds\biggr{]}
+Ckm=0𝔼[(P(tm+1)P(tm))2]\displaystyle\quad+Ck\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla(P(t_{m+1})-P(t_{m}))\|^{2}\bigr{]}
Ck.\displaystyle\leq Ck.

By using the Itô isometry and Theorem 3.1 and (2.4a), we conclude that

(3.48) 𝚅\displaystyle{\tt V} =Ckm=0𝔼[n=0mtntn+1(𝐁(𝐮(s))𝐁(𝐮~n))𝑑W(s)2]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}({\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n}))\,dW(s)\biggr{\|}^{2}\biggr{]}
=Ckm=0𝔼[n=0mtntn+1𝐁(𝐮(s))𝐁(𝐮~n)2𝑑s]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|{\bf B}({\bf u}(s))-{\bf B}(\tilde{{\bf u}}^{n})\|^{2}\,ds\biggr{]}
Ckm=0𝔼[n=0mtntn+1𝐮(s)𝐮~n2𝑑s]\displaystyle\leq Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|{\bf u}(s)-\tilde{{\bf u}}^{n}\|^{2}\,ds\biggr{]}
Ckm=0𝔼[n=0mtntn+1𝐮(s)𝐮(tn)2𝑑s]+Ckm=0𝔼[kn=0m𝐞𝐮~n2]\displaystyle\leq Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}\|{\bf u}(s)-{\bf u}(t_{n})\|^{2}\,ds\biggr{]}+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}k\sum_{n=0}^{m}\|{\bf e}_{\tilde{{\bf u}}}^{n}\|^{2}\biggr{]}
Ck12.\displaystyle\leq Ck^{\frac{1}{2}}.

Finally, substituting (3.46)–(3.48) into (3.45) yields

(3.49) β2𝔼[km=0pm2]Ck12for 1M.\displaystyle\beta^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\mathcal{E}_{p}^{m}\|^{2}\biggr{]}\leq Ck^{\frac{1}{2}}\qquad\mbox{for }1\leq\ell\leq M.

The proof is complete. ∎

Remark 3.2.

It is interesting to point out that the above proof uses the technique from the (non-splitting) mixed method error analysis although Chorin scheme is a splitting scheme.

We conclude this subsection by stating two stability estimates for (𝐮~m,pm)(\tilde{{\bf u}}^{m},p^{m}) in high norms as immediate corollaries of the above error estimates, they will be used in the next section in deriving error estimates for a fully discrete finite element Chorin method. We note that these stability estimates improve those given in Lemma 3.1 and may not be obtained directly without using the above error estimates.

Corollary 3.1.

Under the assumptions of Theorem 3.1, there exists a positive constant CC which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f} such that

(3.50) 𝔼[km=0Mkn=0mpn2]C,\displaystyle\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\biggl{\|}k\sum_{n=0}^{m}\nabla p^{n}\biggr{\|}^{2}\bigg{]}\leq C,
(3.51) max0mM𝔼[kn=0m𝐮~n2]+𝔼[km=0Mkn=0mΔ𝐮~n2]C.\displaystyle\max_{0\leq m\leq M}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{n=0}^{m}\nabla\tilde{{\bf u}}^{n}\biggr{\|}^{2}\biggr{]}+\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n}\biggr{\|}^{2}\bigg{]}\leq C.
Proof.

(3.50) follows immediately from estimates (2.10) and (3.40) and (3.51)1\eqref{eq3.55}_{1} follows straightforwardly from the discrete Jensen inequality and (3.5a)3\eqref{eq3.5}_{3}. It remains to prove (3.51)2\eqref{eq3.55}_{2}. To that end, testing (3.19) by kn=0mΔ𝐮~n+1-k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}, we obtain

(3.52) kn=0mΔ𝐮~n+12\displaystyle\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{\|}^{2} =(𝐮~m+1𝐮0,kn=0mΔ𝐮~n+1)\displaystyle=\Bigl{(}\tilde{{\bf u}}^{m+1}-{\bf u}_{0},k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\Bigr{)}
+(kn=0mpn,kn=0mΔ𝐮~n+1)\displaystyle\quad+\biggl{(}k\sum_{n=0}^{m}\nabla p^{n},k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{)}
(n=0mtntn+1𝐁(𝐮~n)𝑑W(s),kn=0mΔ𝐮~n+1)\displaystyle\quad-\biggl{(}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}{\bf B}(\tilde{{\bf u}}^{n})\,dW(s),k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{)}
𝐮~m+1𝐮02+14kn=0mΔ𝐮~n+12\displaystyle\leq\|\tilde{{\bf u}}^{m+1}-{\bf u}_{0}\|^{2}+\frac{1}{4}\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{\|}^{2}
+kn=0mpn2+14kn=0mΔ𝐮~n+12\displaystyle\quad+\biggl{\|}k\sum_{n=0}^{m}\nabla p^{n}\biggr{\|}^{2}+\frac{1}{4}\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{\|}^{2}
+n=0mtntn+1𝐁(𝐮~n)𝑑W(s)2+14kn=0mΔ𝐮~n+12.\displaystyle\quad+\biggl{\|}\sum_{n=0}^{m}\int_{t_{n}}^{t_{n+1}}{\bf B}(\tilde{{\bf u}}^{n})\,dW(s)\biggr{\|}^{2}+\frac{1}{4}\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n+1}\biggr{\|}^{2}.

After an rearrangement, applying operators km=0Mk\sum_{m=0}^{M} and 𝔼[]\mathbb{E}[\cdot] to (3.52) we obtain

(3.53) 14𝔼[km=0Mkn=0mΔ𝐮~n2]\displaystyle\frac{1}{4}\mathbb{E}\biggl{[}k\sum_{m=0}^{M}\biggl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n}\biggr{\|}^{2}\biggr{]} 𝔼[km=0M(𝐮~m𝐮02+kn=0mpn2)]\displaystyle\leq\mathbb{E}\biggl{[}k\sum_{m=0}^{M}\biggl{(}\|\tilde{{\bf u}}^{m}-{\bf u}^{0}\|^{2}+\biggl{\|}k\sum_{n=0}^{m}\nabla p^{n}\biggr{\|}^{2}\biggr{)}\biggr{]}
+km=0M𝔼[n=0m1tntn+1𝐁(𝐮~n)𝑑W(s)2].\displaystyle\quad+k\sum_{m=0}^{M}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m-1}\int_{t_{n}}^{t_{n+1}}{\bf B}(\tilde{{\bf u}}^{n})\,dW(s)\biggr{\|}^{2}\biggr{]}.

The first two terms can be bounded by using (3.5a) and (3.50). The third term can be controlled by using the Itô isometry and (3.5a) as follows:

km=0M𝔼[n=0m1tntn+1𝐁(𝐮~n)𝑑W(s)2]\displaystyle k\sum_{m=0}^{M}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m-1}\int_{t_{n}}^{t_{n+1}}{\bf B}(\tilde{{\bf u}}^{n})\,dW(s)\biggr{\|}^{2}\biggr{]} =km=0M𝔼[n=0m1tntn+1𝐁(𝐮~n)2𝑑s]\displaystyle=k\sum_{m=0}^{M}\mathbb{E}\biggl{[}\sum_{n=0}^{m-1}\int_{t_{n}}^{t_{n+1}}\|{\bf B}(\tilde{{\bf u}}^{n})\|^{2}\,ds\biggr{]}
Ckm=0M𝔼[kn=0m1𝐮~n2ds]C.\displaystyle\leq Ck\sum_{m=0}^{M}\mathbb{E}\biggl{[}k\sum_{n=0}^{m-1}\|\tilde{{\bf u}}^{n}\|^{2}\,ds\biggr{]}\leq C.

The proof is complete. ∎

3.2. A modified Chorin projection scheme

In this subsection, we consider a modification of Algorithm 1 which was already pointed out in [8] but did not analyze there. The modification is to perform a Helmholtz decomposition of 𝐁(𝐮~m){\bf B}(\tilde{{\bf u}}^{m}) at each time step which allows us to eliminate the curl-free part in Step 1 of Algorithm 1, this then results in a divergent-free Helmholtz projected noise. The goal of this subsection is to present a complete convergence analysis for the modified Chorin scheme which includes deriving stronger error estimates for both velocity and pressure approximations than those for the standard Chorin scheme. We note that this Helmholtz decomposition enhancing technique was also used in [13] to improve the standard mixed finite element method for (1.1).

3.2.1. Formulation of the modified Chorin scheme

For ease of the presentation, we assume W(t)W(t) is a real-valued Wiener process and independent of the spatial variable. The case of more general W(t)W(t) can be dealt with similarly. The modified Chorin scheme is given as follows.

Algorithm 2

Set 𝐮~0=𝐮0=𝐮0\tilde{{\bf u}}^{0}={\bf u}^{0}={\bf u}_{0}. For m=0,1,,M1m=0,1,\cdots,M-1, do the following five steps.

Step 1: Given 𝐮~mL2(Ω,𝐇per1(D))\tilde{{\bf u}}^{m}\in L^{2}(\Omega,{\bf H}^{1}_{per}(D)), find ξmL2(Ω,Hper1(D)/)\xi^{m}\in L^{2}(\Omega,H^{1}_{per}(D)/\mathbb{R}) such that \mathbb{P}-a.s.

(3.54) Δξm=div 𝐁(𝐮~m)in D.\displaystyle\Delta\xi^{m}=\mbox{div }{\bf B}(\tilde{{\bf u}}^{m})\qquad\mbox{in }D.

Step 2: Set 𝜼𝐮~m=𝐁(𝐮~m)ξm\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}}={\bf B}(\tilde{{\bf u}}^{m})-\nabla\xi^{m}. Given 𝐮mL2(Ω,){\bf u}^{m}\in L^{2}(\Omega,\mathbb{H}) and 𝐮~mL2(Ω,𝐇per1(D))\tilde{{\bf u}}^{m}\in L^{2}(\Omega,{\bf H}_{per}^{1}(D)), find 𝐮~m+1L2(Ω,𝐇per1(D))\tilde{{\bf u}}^{m+1}\in L^{2}(\Omega,{\bf H}_{per}^{1}(D)) such that \mathbb{P}-a.s.

(3.55) 𝐮~m+1kΔ𝐮~m+1\displaystyle\tilde{{\bf u}}^{m+1}-k\Delta\tilde{{\bf u}}^{m+1} =𝐮m+k𝐟m+1+𝜼𝐮~mΔWm+1in D.\displaystyle={\bf u}^{m}+k{\bf f}^{m+1}+\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}}\Delta W_{m+1}\qquad\mbox{in }D.

Step 3: Find rm+1L2(Ω,Hper1(D)/)r^{m+1}\in L^{2}(\Omega,H^{1}_{per}(D)/\mathbb{R}) such that \mathbb{P}-a.s.

(3.56a) Δrm+1\displaystyle-\Delta r^{m+1} =1kdiv 𝐮~m+1\displaystyle=-\frac{1}{k}\mbox{\rm div\,}\tilde{{\bf u}}^{m+1} in D.\displaystyle\qquad\mbox{in }D.

Step 4: Define 𝐮m+1L2(Ω,){\bf u}^{m+1}\in L^{2}(\Omega,\mathbb{H}) as

(3.57) 𝐮m+1\displaystyle{\bf u}^{m+1} =𝐮~m+1krm+1in D.\displaystyle=\tilde{{\bf u}}^{m+1}-k\nabla r^{m+1}\qquad\mbox{in }D.

Step 5: Define the pressure approximation pm+1p^{m+1} as

(3.58) pm+1=rm+1+1kξmΔWm+1in D.\displaystyle p^{m+1}=r^{m+1}+\frac{1}{k}\xi^{m}\Delta W_{m+1}\qquad\mbox{in }D.
Remark 3.3.

It follows from (2.2b) and (3.54) that the Helmholtz projection 𝛈𝐮~m\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}} can be bounded in terms of 𝐮~m\tilde{{\bf u}}^{m} as follows:

(3.59) 𝜼𝐮~mL2\displaystyle\|\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}}\|_{L^{2}} 𝐁(𝐮~m)L2+ξ𝐮~mL22𝐁(𝐮~m)L2C𝐮~mL2\displaystyle\leq\|{\bf B}(\tilde{{\bf u}}^{m})\|_{L^{2}}+\|\nabla\xi^{m}_{\tilde{{\bf u}}}\|_{L^{2}}\leq 2\|{\bf B}(\tilde{{\bf u}}^{m})\|_{L^{2}}\leq C\|\tilde{{\bf u}}^{m}\|_{L^{2}}

3.2.2. Stability estimates for the modified Chorin scheme

In this subsection we first state some stability estimates for Algorithm 2. We then recall the Euler-Maruyama scheme for (1.1) and its stability and error estimates from [13], which will be utilized as a tool in the stability and error analysis of the modified Chorin scheme in the next subsection.

Lemma 3.2.

The discrete processes {(𝐮~m,rm)}m=0M\{(\tilde{{\bf u}}^{m},r^{m})\}_{m=0}^{M} defined by Algorithm 2 satisfy

(3.60a) max0mM𝔼[𝐮~m2]+𝔼[m=1M𝐮~m𝐮~m12]+𝔼[km=0M𝐮~m2]\displaystyle\max_{0\leq m\leq M}\mathbb{E}[\|\tilde{{\bf u}}^{m}\|^{2}]+\mathbb{E}\bigg{[}\sum_{m=1}^{M}\|\tilde{{\bf u}}^{m}-\tilde{{\bf u}}^{m-1}\|^{2}\bigg{]}+\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|\nabla\tilde{{\bf u}}^{m}\|^{2}\bigg{]} C,\displaystyle\leq C,
(3.60b) 𝔼[k2m=0Mrm2]\displaystyle\mathbb{E}\bigg{[}k^{2}\sum_{m=0}^{M}\|\nabla r^{m}\|^{2}\bigg{]} C,\displaystyle\leq C,

where CC is a positive constant which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f}.

Since the proof of this lemma follows the same lines as those of Lemma 3.1. We omit the proof to save space.

Next, we recall the Helmholtz enhanced Euler-Maruyama scheme for (1.1) which was proposed and analyzed in [13]. This scheme will be used as an auxiliary scheme in our error estimates for the velocity and pressure approximations of Algorithm 2 in the next subsection. The Euler-Maruyama scheme reads as

(3.61a) (𝐯m+1𝐯m)kΔ𝐯m+1+kqm+1\displaystyle({{\bf v}}^{m+1}-{\bf v}^{m})-k\Delta{\bf v}^{m+1}+k\nabla q^{m+1} =k𝐟m+1+𝜼𝐯mΔWm+1\displaystyle=k{\bf f}^{m+1}+\boldsymbol{\eta}_{{\bf v}}^{m}\Delta W_{m+1} inD,\displaystyle\qquad\mbox{in}\,D,
(3.61b) div 𝐯m+1\displaystyle\mbox{\rm div\,}{\bf v}^{m+1} =0\displaystyle=0 inD,\displaystyle\qquad\mbox{in}\,D,

where 𝜼𝐯m=𝐁(𝐯m)ξ𝐯m\boldsymbol{\eta}_{{\bf v}}^{m}={\bf B}({\bf v}^{m})-\nabla\xi^{m}_{{\bf v}} denotes the Helmholtz projection of 𝐁(𝐯m){\bf B}({\bf v}^{m}).

It was proved in [13] that the following stability and error estimates hold for the solution of the above Euler-Maruyama scheme.

Lemma 3.3.

The discrete processes {(𝐯m,qm)}m=0M\{({\bf v}^{m},q^{m})\}_{m=0}^{M} defined by (3.61) satisfy

(3.62a) max0mM𝔼[𝐯m2]+𝔼[m=1M𝐯m𝐯m12]+𝔼[km=0M𝐯m2]\displaystyle\max_{0\leq m\leq M}\mathbb{E}\bigl{[}\|{\bf v}^{m}\|^{2}\bigr{]}+\mathbb{E}\Bigl{[}\sum^{M}_{m=1}\|{\bf v}^{m}-{\bf v}^{m-1}\|^{2}\Bigr{]}+\mathbb{E}\Bigl{[}k\sum^{M}_{m=0}\|\nabla{\bf v}^{m}\|^{2}\Bigr{]} C,\displaystyle\leq C,
(3.62b) max0mM𝔼[𝐯m2]+𝔼[m=1M((𝐯m𝐯m1)2+k𝐀𝐯m2)]\displaystyle\max_{0\leq m\leq M}\mathbb{E}\Bigl{[}\|\nabla{\bf v}^{m}\|^{2}\Bigr{]}+\mathbb{E}\Bigl{[}\sum^{M}_{m=1}\Bigl{(}\|\nabla({\bf v}^{m}-{\bf v}^{m-1})\|^{2}+k\|{\bf A}{\bf v}^{m}\|^{2}\Bigr{)}\Bigr{]} C,\displaystyle\leq C,
(3.62c) 𝔼[km=0Mqm2]\displaystyle\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\|\nabla q^{m}\|^{2}\Bigr{]} C,\displaystyle\leq C,

where CC is a positive constant which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f}.

Lemma 3.4.

There hold the the following error estimates for the discrete processes {(𝐯m,qm)}m=0M\{({\bf v}^{m},q^{m})\}_{m=0}^{M}:

(3.63a) max0mM(𝔼[𝐮(tm)𝐯m2])12\displaystyle\max_{0\leq m\leq M}\bigl{(}\mathbb{E}\big{[}\|{\bf u}(t_{m})-{\bf v}^{m}\|^{2}\big{]}\bigr{)}^{\frac{1}{2}}\hskip 108.405pt
+(𝔼[km=0M(𝐮(tm)𝐯m)2])12\displaystyle+\bigg{(}\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|\nabla({\bf u}(t_{m})-{\bf v}^{m})\|^{2}\bigg{]}\bigg{)}^{\frac{1}{2}} Ck,\displaystyle\leq C\sqrt{k},
(3.63b) (𝔼[R(t)km=0qm2])12\displaystyle\bigg{(}\mathbb{E}\bigg{[}\bigg{\|}R(t_{\ell})-k\sum_{m=0}^{\ell}q^{m}\bigg{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}} Ck\displaystyle\leq C\sqrt{k}

for 0M0\leq\ell\leq M. Where CC is a positive constant which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f}.

Remark 3.4.

We note that the reason for the estimate (3.62c) to hold is because the Helmholtz projected noise 𝛈𝐯m\boldsymbol{\eta}_{{\bf v}}^{m} is divergent-free. Otherwise, the following weaker estimate can only be proved (cf. [12]):

𝔼[km=0Mqm2]Ck.\displaystyle\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\|\nabla q^{m}\|^{2}\Bigr{]}\leq\frac{C}{k}.

3.2.3. Error estimates for the modified Chorin scheme

The goal of this subsection is to derive error estimates for both the velocity and pressure approximations generated by Algorithm 2. The anticipated error estimates are stronger than those for the standard Chorin scheme proved in the previous subsection. We note that our error estimate for the velocity approximation recovers the same estimate obtained in [8, Theorem 3.1] although the analysis given here is a lot simpler. On the other hand, the error estimate for the pressure approximation is apparently new. The main idea of the proofs of this subsection is to use the Euler-Maruyama scheme analyzed in [13] as an auxiliary scheme which bridges the exact solution of (1.1) and the discrete solution of Algorithm 2.

The follow theorem gives an error estimate in strong norms for the velocity approximation.

Theorem 3.3.

Let {(𝐮~m,pm)}m=0M\{(\tilde{{\bf u}}^{m},p^{m})\}_{m=0}^{M} be the solution of Algorithm 2 and {(𝐮(t),\{({\bf u}(t), P(t)); 0tT}P(t));\,0\leq t\leq T\} be the solution of (1.1). Then there holds the following estimate:

(3.64) max0mM(𝔼[𝐮(tm)\displaystyle\max_{0\leq m\leq M}\bigg{(}\mathbb{E}\bigg{[}\|{\bf u}(t_{m}) 𝐮~m2])12\displaystyle-\tilde{{\bf u}}^{m}\|^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}
+(𝔼[km=0M(𝐮(tm)𝐮~m)2])12Ck,\displaystyle+\bigg{(}\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|\nabla({\bf u}(t_{m})-\tilde{{\bf u}}^{m})\|^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k},

where CC is a positive constant which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f}.

Proof.

Let 𝐞𝐮~m=𝐯m𝐮~m{\bf e}^{m}_{\tilde{{\bf u}}}={\bf v}^{m}-\tilde{{\bf u}}^{m}, erm=qmrme_{r}^{m}=q^{m}-r^{m}. Then, 𝐞𝐮~mL2(Ω,𝐇per1(D)){\bf e}_{\tilde{{\bf u}}}^{m}\in L^{2}(\Omega,{\bf H}_{per}^{1}(D)) and ermL2(Ω,Hper1(D)/)e_{r}^{m}\in L^{2}(\Omega,H_{per}^{1}(D)/\mathbb{R}). Subtracting (3.4) from (3.61) yields

(3.65a) (𝐞𝐮~m+1𝐞𝐮~m)kΔ𝐞𝐮~m+1+kerm\displaystyle({\bf e}^{m+1}_{\tilde{{\bf u}}}-{\bf e}^{m}_{\tilde{{\bf u}}})-k\Delta{\bf e}^{m+1}_{\tilde{{\bf u}}}+k\nabla e^{m}_{r} =k((qm+1qm))\displaystyle=-k(\nabla(q^{m+1}-q^{m}))
+(𝜼𝐯m𝜼𝐮~m)ΔWm+1\displaystyle\qquad+(\boldsymbol{\eta}^{m}_{{\bf v}}-\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}})\Delta W_{m+1} onD,\displaystyle\qquad\mbox{on}\,D,
(3.65b) div 𝐞𝐮~m+1kΔerm+1\displaystyle\mbox{\rm div\,}{\bf e}^{m+1}_{\tilde{{\bf u}}}-k\Delta e_{r}^{m+1} =kΔqm+1\displaystyle=-k\Delta q^{m+1} onD,\displaystyle\qquad\mbox{on}\,D,
(3.65c) 𝐧erm+1\displaystyle\partial_{\bf n}e_{r}^{m+1} =𝐧qm+1\displaystyle=\partial_{\bf n}q^{m+1} onD.\displaystyle\qquad\mbox{on}\,\partial D.

Testing (3.65a) with 𝐞𝐮~m+1{\bf e}^{m+1}_{\tilde{{\bf u}}} and integrating by parts, we obtain

(3.66) (𝐞𝐮~m+1𝐞𝐮~m,\displaystyle({\bf e}^{m+1}_{\tilde{{\bf u}}}-{\bf e}^{m}_{\tilde{{\bf u}}}, 𝐞𝐮~m+1)+k𝐞m+1𝐮~2+k(erm,𝐞𝐮~m+1)\displaystyle\,{\bf e}^{m+1}_{\tilde{{\bf u}}})+k\|\nabla{\bf e}^{m+1}_{\tilde{{\bf u}}}\|^{2}+k(\nabla e_{r}^{m},{\bf e}^{m+1}_{\tilde{{\bf u}}})
=k((qm+1qm),𝐞𝐮~m+1)+((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐞𝐮~m+1).\displaystyle=-k(\nabla(q^{m+1}-q^{m}),{\bf e}^{m+1}_{\tilde{{\bf u}}})+((\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}})\Delta W_{m+1},{\bf e}^{m+1}_{\tilde{{\bf u}}}).

Using the algebraic identity 2ab=(a2b2)+(ab)22ab=(a^{2}-b^{2})+(a-b)^{2} in the first term gives

(3.67) 12(𝐞𝐮~m+1L22\displaystyle\frac{1}{2}\bigl{(}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|_{L^{2}}^{2} 𝐞𝐮~mL22)+12𝐞𝐮~m+1𝐞𝐮~mL22\displaystyle-\|{\bf e}_{\tilde{{\bf u}}}^{m}\|_{L^{2}}^{2}\bigr{)}+\frac{1}{2}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|_{L^{2}}^{2}
+k𝐞𝐮~m+1L22+k(erm,𝐞𝐮~m+1)\displaystyle\qquad+k\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|_{L^{2}}^{2}+k(\nabla e_{r}^{m},{\bf e}_{\tilde{{\bf u}}}^{m+1})
=k((qm+1qm),𝐞𝐮~m+1)+((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐞𝐮~m+1).\displaystyle=-k(\nabla(q^{m+1}-q^{m}),{\bf e}_{\tilde{{\bf u}}}^{m+1})+((\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}})\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}).

Next, we derive a reformulation for each of the last term on the left-hand side and the first term on the right of (3.67) with a help of (3.65b). Testing (3.65b) by any qL2(Ω,Hper1(D)/)q\in L^{2}(\Omega,H_{per}^{1}(D)/\mathbb{R}) and using (3.65c), we obtain

(3.68a) (𝐞𝐮~m+1,q)k(erm+1,q)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla q\bigr{)}-k\bigl{(}\nabla e_{r}^{m+1},\nabla q\bigr{)} =k(qm+1,q)\displaystyle=-k\bigl{(}\nabla q^{m+1},\nabla q\bigr{)}
(3.68b) (𝐞𝐮~m+1𝐞𝐮~m,q)k((erm+1erm),q)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},\nabla q\bigr{)}-k\bigl{(}\nabla(e_{r}^{m+1}-e_{r}^{m}),\nabla q\bigr{)} =k((qm+1qm),q).\displaystyle=-k\bigl{(}\nabla(q^{m+1}-q^{m}),\nabla q\bigr{)}.

Setting q=ermq=e_{r}^{m} in (3.68a), we obtain

(3.69) (𝐞𝐮~m+1,erm)=k(erm+1,erm)k(qm+1,erm),\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla e_{r}^{m}\bigr{)}=k\bigl{(}\nabla e_{r}^{m+1},\nabla e_{r}^{m}\bigr{)}-k\bigl{(}\nabla q^{m+1},\nabla e_{r}^{m}\bigr{)},
=kerm+12k((erm+1erm),erm+1)k(qm+1,erm).\displaystyle\hskip 43.36243pt=k\|\nabla e_{r}^{m+1}\|^{2}-k\bigl{(}\nabla(e_{r}^{m+1}-e_{r}^{m}),\nabla e_{r}^{m+1}\bigr{)}-k\bigl{(}\nabla q^{m+1},\nabla e_{r}^{m}\bigr{)}.

and choosing q=erm+1q=e_{r}^{m+1} in (3.68b), we have

(3.70) k((erm+1erm),erm+1)\displaystyle k\bigl{(}\nabla(e_{r}^{m+1}-e_{r}^{m}),\nabla e_{r}^{m+1}\bigr{)} =(𝐞𝐮~m+1𝐞𝐮~m,erm+1)\displaystyle=\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},\nabla e_{r}^{m+1}\bigr{)}
+k((qm+1qm),erm+1).\displaystyle\qquad+k\bigl{(}\nabla(q^{m+1}-q^{m}),\nabla e_{r}^{m+1}\bigr{)}.

Substituting (3.70) to (3.69) yields

(3.71) (𝐞𝐮~m+1,erm)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla e_{r}^{m}\bigr{)} =kerm+12(𝐞𝐮~m+1𝐞𝐮~m,erm+1)\displaystyle=k\|\nabla e_{r}^{m+1}\|^{2}-\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},\nabla e_{r}^{m+1}\bigr{)}
k((qm+1qm),erm+1)k(qm+1,erm).\displaystyle\quad-k\bigl{(}\nabla(q^{m+1}-q^{m}),\nabla e_{r}^{m+1}\bigr{)}-k\bigl{(}\nabla q^{m+1},\nabla e_{r}^{m}\bigr{)}.

Alternatively, setting q=qm+1qmq=q^{m+1}-q^{m} in (3.68a), we obtain

(3.72) (𝐞𝐮~m+1,(qm+1qm))\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla(q^{m+1}-q^{m})\bigr{)} =k(erm+1,(qm+1qm))\displaystyle=k\bigl{(}\nabla e_{r}^{m+1},\nabla(q^{m+1}-q^{m})\bigr{)}
k(qm+1,(qm+1qm)).\displaystyle\qquad-k\bigl{(}\nabla q^{m+1},\nabla(q^{m+1}-q^{m})\bigr{)}.

Substituting (3.71) and (3.72) into (3.67) and rearranging terms, we get

(3.73) 12(𝐞𝐮~m+1L22𝐞𝐮~m2)+12𝐞𝐮~m+1𝐞𝐮~m2+k𝐞𝐮~m+12+k2erm+12\displaystyle\frac{1}{2}\bigl{(}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|_{L^{2}}^{2}-\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\bigr{)}+\frac{1}{2}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}+k\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}+k^{2}\|\nabla e_{r}^{m+1}\|^{2}
=k(𝐞𝐮~m+1𝐞𝐮~m,erm+1)+k2(qm+1,erm)\displaystyle\qquad=k\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},\nabla e_{r}^{m+1}\bigr{)}+k^{2}\bigl{(}\nabla q^{m+1},\nabla e_{r}^{m}\bigr{)}
+k2(qm+1,(qm+1qm))+((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐞𝐮~m+1).\displaystyle\qquad\qquad+k^{2}\bigl{(}\nabla q^{m+1},\nabla(q^{m+1}-q^{m})\bigr{)}+\bigl{(}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}\bigr{)}.

Finally, we bound each term on the right-hand side of (3.73). By Young’s inequality, for δ1,δ2,δ3>0\delta_{1},\delta_{2},\delta_{3}>0 we obtain

(3.74a) k(𝐞𝐮~m+1𝐞𝐮~m,erm+1)\displaystyle k\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},\nabla e_{r}^{m+1}\bigr{)} 14δ1𝐞𝐮~m+1𝐞𝐮~m2+δ1k2erm+12.\displaystyle\leq\frac{1}{4\delta_{1}}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}+\delta_{1}k^{2}\|\nabla e_{r}^{m+1}\|^{2}.
(3.74b) k2(qm+1,erm)\displaystyle k^{2}\bigl{(}\nabla q^{m+1},\nabla e_{r}^{m}\bigr{)} 14δ2k2qm+12+δ2k2erm2.\displaystyle\leq\frac{1}{4\delta_{2}}k^{2}\|\nabla q^{m+1}\|^{2}+\delta_{2}k^{2}\|\nabla e_{r}^{m}\|^{2}.
(3.74c) k2(qm+1,(qm+1qm))\displaystyle k^{2}\bigl{(}\nabla q^{m+1},\nabla(q^{m+1}-q^{m})\bigr{)} k2qm+12+14k2(qm+1qm)2.\displaystyle\leq k^{2}\|\nabla q^{m+1}\|^{2}+\frac{1}{4}k^{2}\|\nabla(q^{m+1}-q^{m})\|^{2}.

Rewriting

(3.75) ((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐞𝐮~m+1)\displaystyle\bigl{(}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}\bigr{)} =((𝜼𝐯m𝜼𝐮~m)Δm+1W,𝐞𝐮~m+1𝐞𝐮~m)\displaystyle=\bigl{(}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta_{m+1}W,{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}
+((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐞𝐮~m).\displaystyle\qquad\qquad+\bigl{(}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}.

Since the expectation of the second term on the right-hand side of (3.75) vanishes due to the martingale property of Itô’s integral, we only need to estimate the first term. Again, rewriting

(3.76) ((𝜼𝐯m𝜼𝐮~m)Δm+1W,𝐞𝐮~m+1𝐞𝐮~m)=((ξ𝐯mξ𝐮~m)Δm+1W,𝐞𝐮~m+1𝐞𝐮~m)\displaystyle\bigl{(}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta_{m+1}W,{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}=-\bigl{(}\nabla(\xi_{{\bf v}}^{m}-\xi_{\tilde{{\bf u}}}^{m})\Delta_{m+1}W,{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}
+((𝐁(𝐯m)𝐁(𝐮~m))ΔWm+1,𝐞𝐮~m+1𝐞𝐮~m).\displaystyle\hskip 130.08621pt+\bigl{(}({\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m}))\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}.

Then we have

(3.77a) 𝔼[((ξ𝐯m\displaystyle\mathbb{E}\big{[}\bigl{(}\nabla(\xi_{{\bf v}}^{m} ξ𝐮~m)ΔWm+1,𝐞𝐮~m+1𝐞𝐮~m)]\displaystyle-\xi_{\tilde{{\bf u}}}^{m})\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\bigr{)}\big{]}
2k𝔼[(ξ𝐯mξ𝐮~m)2]+18𝔼[𝐞𝐮~m+1𝐞𝐮~m2]\displaystyle\leq 2k\mathbb{E}\big{[}\|\nabla(\xi_{{\bf v}}^{m}-\xi_{\tilde{{\bf u}}}^{m})\|^{2}\big{]}+\frac{1}{8}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}
2k𝔼[(𝐁(𝐯m)𝐁(𝐮~m))2]+18𝔼[𝐞𝐮~m+1𝐞𝐮~m2]\displaystyle\leq 2k\mathbb{E}\big{[}\|({\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m}))\|^{2}\big{]}+\frac{1}{8}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}
Ck𝔼[𝐞𝐮~m2]+18𝔼[𝐞𝐮~m+1𝐞𝐮~m2],\displaystyle\leq Ck\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}+\frac{1}{8}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]},
(3.77b) 𝔼[((𝐁(𝐯m)\displaystyle\mathbb{E}\big{[}\bigl{(}({\bf B}({\bf v}^{m}) 𝐁(𝐮~m))ΔWm+1,𝐞𝐮~m+1)]\displaystyle-{\bf B}(\tilde{{\bf u}}^{m}))\Delta W_{m+1},{\bf e}_{\tilde{{\bf u}}}^{m+1}\bigr{)}\big{]}
14δ3k𝔼[(𝐁(𝐯m)𝐁(𝐮~m))L22]+δ3𝔼[𝐞𝐮~m+1𝐞𝐮~m2]\displaystyle\leq\frac{1}{4\delta_{3}}k\mathbb{E}\big{[}\|\bigl{(}{\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m})\bigr{)}\|_{L^{2}}^{2}\big{]}+\delta_{3}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}
Ck4δ3𝔼[𝐞𝐮~m2]+δ3𝔼[𝐞𝐮~m+1𝐞𝐮~m2].\displaystyle\leq\frac{Ck}{4\delta_{3}}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}+\delta_{3}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}.

Now, substituting (3.74) and (3.77) into (3.73) and taking expectation on both sides, we obtain

(3.78) 12𝔼[𝐞𝐮~m+12𝐞𝐮~m2]+(3814δ1δ3)𝔼[𝐞𝐮~m+1𝐞𝐮~m2\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}-\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\bigr{]}+\bigg{(}\frac{3}{8}-\frac{1}{4\delta_{1}}-\delta_{3}\bigg{)}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}
+k𝔼[𝐞𝐮~m+12]+k2(1δ1δ2)𝔼[erm+12]\displaystyle\qquad+k\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigr{]}+k^{2}\bigl{(}1-\delta_{1}-\delta_{2}\bigr{)}\mathbb{E}\bigl{[}\|\nabla e_{r}^{m+1}\|^{2}\bigr{]}
=12k2𝔼[qm+12]+12k2𝔼[(qm+1qm)2]+14δ3Ck𝔼[𝐞𝐮~m2].\displaystyle\quad=\frac{1}{2}k^{2}\mathbb{E}\bigl{[}\|\nabla q^{m+1}\|^{2}\bigr{]}+\frac{1}{2}k^{2}\mathbb{E}\bigl{[}\|\nabla(q^{m+1}-q^{m})\|^{2}\bigr{]}+\frac{1}{4\delta_{3}}Ck\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\bigr{]}.

Choosing δ1=1720,δ2=110,δ3=116\delta_{1}=\frac{17}{20},\delta_{2}=\frac{1}{10},\delta_{3}=\frac{1}{16}, taking the summation for 0mM10\leq m\leq\ell\leq M-1, and using (3.62c) and the discrete Gronwall’s inequality, we get

(3.79) 12𝔼[𝐞𝐮~+12]\displaystyle\frac{1}{2}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{\ell+1}\|^{2}\bigr{]} +5272𝔼[m=0𝐞𝐮~m+1𝐞𝐮~m2]+𝔼[km=0𝐞𝐮~m+12]\displaystyle+\frac{5}{272}\mathbb{E}\bigg{[}\sum_{m=0}^{\ell}\|{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\bigg{]}+\mathbb{E}\bigg{[}k\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|^{2}\bigg{]}
+120𝔼[k2m=0erm+12]exp(4CT)Ck.\displaystyle+\frac{1}{20}\mathbb{E}\bigg{[}k^{2}\sum_{m=0}^{\ell}\|\nabla e_{r}^{m+1}\|^{2}\bigg{]}\leq\exp\bigl{(}4CT\bigr{)}C\,k.

which and (3.63a) infer the desired estimate. The proof is complete. ∎

An immediate corollary of the above error estimate is the following stronger stability estimates for {(𝐮~m,rm)}\{(\tilde{{\bf u}}^{m},r^{m})\}, which may not be obtainable directly and will play an important role in the error analysis of fully discrete counterpart of Algorithm 2 in the next section.

Corollary 3.2.

There exists C>0C>0 which depends on DT,𝐮0D_{T},{\bf u}_{0} and 𝐟{\bf f} such that

(3.80a) max0mM𝔼[𝐮~m2]+𝔼[km=0MΔ𝐮~m2]\displaystyle\max_{0\leq m\leq M}\mathbb{E}[\|\nabla\tilde{{\bf u}}^{m}\|^{2}]+\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|\Delta\tilde{{\bf u}}^{m}\|^{2}\bigg{]} C,\displaystyle\leq C,
(3.80b) 𝔼[km=0Mrm2]\displaystyle\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|\nabla r^{m}\|^{2}\bigg{]} C.\displaystyle\leq C.
Proof.

The estimate (3.80b) follows immediately from (3.79) and the triangular inequality. To prove (3.80a), testing (3.4a) by Δ𝐮~m+1-\Delta\tilde{{\bf u}}^{m+1}, we obtain

(3.81) 12𝔼[𝐮~m+12𝐮~m2]\displaystyle\frac{1}{2}\mathbb{E}[\|\nabla\tilde{{\bf u}}^{m+1}\|^{2}-\|\nabla\tilde{{\bf u}}^{m}\|^{2}] +14𝔼[(𝐮~m+1𝐮~m)2]+14k𝔼[Δ𝐮~m+12]\displaystyle+\frac{1}{4}\mathbb{E}[\|\nabla(\tilde{{\bf u}}^{m+1}-\tilde{{\bf u}}^{m})\|^{2}]+\frac{1}{4}k\mathbb{E}[\|\Delta\tilde{{\bf u}}^{m+1}\|^{2}]
k𝔼[rm2]+k𝔼[𝐟m+12]+k𝔼[𝜼𝐮~m2].\displaystyle\leq k\mathbb{E}[\|\nabla r^{m}\|^{2}]+k\mathbb{E}\big{[}\|{\bf f}^{m+1}\|^{2}\big{]}+k\mathbb{E}[\|\nabla\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}\|^{2}].

Here we have used the periodic boundary condition for 𝜼𝐮~m\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m} to kill the boundary term which arises from integration by parts.

Finally, applying the operator m=0\sum_{m=0}^{\ell} to (3.81) and using the discrete Gronwall inequality and (3.80a)3, we obtain the desired estimate. The proof is complete. ∎

Similarly, the following estimate holds for {𝐮m}\{{\bf u}^{m}\}.

Corollary 3.3.

There exists C>0C>0 which depends on DT,𝐮0D_{T},\mathbf{u}_{0} and 𝐟\mathbf{f} such that

(3.82) max0mM(𝔼[𝐮(tm)𝐮m2])12+(𝔼[km=0M𝐮(tm)𝐮m2])12Ck.\displaystyle\max_{0\leq m\leq M}\big{(}\mathbb{E}[\|{\bf u}(t_{m})-{\bf u}^{m}\|^{2}]\big{)}^{\frac{1}{2}}+\bigg{(}\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\|{\bf u}(t_{m})-{\bf u}^{m}\|^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k}.

The proof of (3.82) readily follows from (3.57) and Theorem (3.3) as well as the estimate (3.80b).

Next, we derive error estimates for the pressure approximations rmr^{m} and pmp^{m} generated by Algorithm 2. First, we state the following lemma.

Lemma 3.5.

Let {rm}m=0M\{r^{m}\}_{m=0}^{M} be generated by Algorithm 2. Then, there exists a constant C>0C>0 depending on DT,𝐮0,𝐟D_{T},{\bf u}_{0},{\bf f} and β\beta such that for 0M0\leq\ell\leq M

(3.83) (𝔼[km=1(rmrm1)2])12Ck.\displaystyle\bigg{(}\mathbb{E}\bigg{[}\bigg{\|}k\sum_{m=1}^{\ell}(r^{m}-r^{m-1})\bigg{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k}.
Proof.

The idea of the proof is to utilize the inf-sup condition (3.41). Testing (3.57) by any 𝐯L2(Ω;𝐇per1(D)){\bf v}\in L^{2}(\Omega;{\bf H}^{1}_{per}(D)), we obtain

k(rm+1,div 𝐯)\displaystyle k\bigl{(}r^{m+1},\mbox{\rm div\,}{\bf v}\bigr{)} =(𝐮m+1𝐮~m+1,𝐯),\displaystyle=\bigl{(}{\bf u}^{m+1}-\tilde{{\bf u}}^{m+1},{\bf v}\bigr{)},
k(rm,div 𝐯)\displaystyle k\bigl{(}r^{m},\mbox{\rm div\,}{\bf v}\bigr{)} =(𝐮m𝐮~m,𝐯).\displaystyle=\bigl{(}{\bf u}^{m}-\tilde{{\bf u}}^{m},{\bf v}\bigr{)}.

Then, subtracting the above equations yields

(3.84) k(rm+1rm,div 𝐯)=((𝐮m+1𝐮m)(𝐮~m+1𝐮~m),𝐯).\displaystyle k\bigl{(}r^{m+1}-r^{m},\mbox{\rm div\,}{\bf v}\bigr{)}=\bigl{(}({\bf u}^{m+1}-{\bf u}^{m})-(\tilde{{\bf u}}^{m+1}-\tilde{{\bf u}}^{m}),{\bf v}\bigr{)}.

Applying the summation operator m=0\sum_{m=0}^{\ell} for 0M10\leq\ell\leq M-1 to (3.84), we get

(3.85) (km=0(rm+1rm),div 𝐯)\displaystyle\bigg{(}k\sum_{m=0}^{\ell}(r^{m+1}-r^{m}),\mbox{\rm div\,}{\bf v}\bigg{)} =((𝐮+1𝐮0)(𝐮~+1𝐮~0),𝐯)\displaystyle=\bigl{(}({\bf u}^{\ell+1}-{\bf u}^{0})-(\tilde{{\bf u}}^{\ell+1}-\tilde{{\bf u}}^{0}),{\bf v}\bigr{)}
=(𝐞𝐮+1𝐞𝐮~+1,𝐯)\displaystyle=\bigl{(}{\bf e}_{{\bf u}}^{\ell+1}-{\bf e}_{\tilde{{\bf u}}}^{\ell+1},{\bf v}\bigr{)}
C(𝐞𝐮+1+𝐞𝐮~+1)𝐯,\displaystyle\leq C\bigl{(}\|{\bf e}_{{\bf u}}^{\ell+1}\|+\|{\bf e}_{\tilde{{\bf u}}}^{\ell+1}\|\bigr{)}\|\nabla{\bf v}\|,

where 𝐞𝐮m{\bf e}_{{\bf u}}^{m} and 𝐞𝐮~m{\bf e}_{\tilde{{\bf u}}}^{m} are the same as defined in the preceding subsection and we have used the fact that 𝐮0𝐮~0=0{\bf u}^{0}-\tilde{{\bf u}}^{0}=0.

Finally, by using the inf-sup condition (3.41) and then taking the expectation we get

β2𝔼[km=0(rm+1rm)2]\displaystyle\beta^{2}\mathbb{E}\bigg{[}\Bigl{\|}k\sum_{m=0}^{\ell}(r^{m+1}-r^{m})\Bigr{\|}^{2}\bigg{]} C(𝔼[𝐞𝐮+12]+𝔼[𝐞𝐮~+12]),\displaystyle\leq C\Bigl{(}\mathbb{E}\bigl{[}\|{\bf e}_{{\bf u}}^{\ell+1}\|^{2}\bigr{]}+\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}}^{\ell+1}\|^{2}\bigr{]}\Bigr{)},

which and the estimates for 𝐞𝐮+1{\bf e}_{{\bf u}}^{\ell+1} and 𝐞𝐮~+1{\bf e}_{\tilde{{\bf u}}}^{\ell+1} infer the desired estimate (3.83). The proof is complete. ∎

We then are ready to state the following error estimate result for rmr^{m}.

Theorem 3.4.

Let {rm}m=0M\{r^{m}\}_{m=0}^{M} be generated by Algorithm 2 and R(t)R(t) be defined in (2.6). Then there exists a constant C>0C>0 depending on DT,𝐮0,𝐟D_{T},{\bf u}_{0},{\bf f} and β\beta such hat for 0M0\leq\ell\leq M

(3.86) (𝔼[R(t)km=0rm2])12Ck.\displaystyle\bigg{(}\mathbb{E}\bigg{[}\bigg{\|}R(t_{\ell})-k\sum_{m=0}^{\ell}r^{m}\bigg{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k}.
Proof.

Subtracting (3.61) from (3.4a) and then testing the resulting equation by 𝐯L2(Ω;𝐇per1(D)){\bf v}\in L^{2}(\Omega;{\bf H}^{1}_{per}(D)), we obtain

(3.87) (𝐞𝐮~m+1𝐞𝐮~m,𝐯)+k(𝐞𝐮~m+1,𝐯)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{m+1}-{\bf e}_{\tilde{{\bf u}}}^{m},{\bf v}\bigr{)}+k\bigl{(}\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla{\bf v}\bigr{)} k(qm+1rm,div 𝐯)\displaystyle-k\bigl{(}q^{m+1}-r^{m},\mbox{\rm div\,}{\bf v}\bigr{)}
=((𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐯).\displaystyle=\bigl{(}\big{(}\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}\big{)}\Delta W_{m+1},{\bf v}\bigr{)}.

Applying the summation operator m=0\sum_{m=0}^{\ell} to (3.87) for 0M10\leq\ell\leq M-1 yields

(3.88) (km=0(qm+1rm),div 𝐯)\displaystyle\bigg{(}k\sum_{m=0}^{\ell}(q^{m+1}-r^{m}),\mbox{\rm div\,}{\bf v}\bigg{)} =(𝐞𝐮~+1𝐞𝐮~0,𝐯)+(km=0𝐞𝐮~m+1,𝐯)\displaystyle=\bigl{(}{\bf e}_{\tilde{{\bf u}}}^{\ell+1}-{\bf e}_{\tilde{{\bf u}}}^{0},{\bf v}\bigr{)}+\bigg{(}k\sum_{m=0}^{\ell}\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1},\nabla{\bf v}\bigg{)}
(m=0(𝜼𝐯m𝜼𝐮~m)ΔWm+1,𝐯)\displaystyle\qquad-\bigg{(}\sum_{m=0}^{\ell}\big{(}\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}\big{)}\Delta W_{m+1},{\bf v}\bigg{)}
=𝙸+𝙸𝙸+𝙸𝙸𝙸.\displaystyle={\tt I}+{\tt II}+{\tt III}.

Now, we bound each term on the right-hand side of (3.88) as follows. By Schwarz and Poincaré inequalities, we get

|𝙸|+|𝙸𝙸|\displaystyle|{\tt I}|+|{\tt II}| C𝐞𝐮~+1𝐯+km=0𝐞𝐮~m+1𝐯,\displaystyle\leq C\|{\bf e}_{\tilde{{\bf u}}}^{\ell+1}\|\|\nabla{\bf v}\|+k\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}}^{m+1}\|\cdot\|\nabla{\bf v}\|,
|𝙸𝙸𝙸|\displaystyle|{\tt III}| m=0(𝜼𝐯m𝜼𝐮~m)ΔWm+1𝐯\displaystyle\leq\big{\|}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{{\bf v}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m})\Delta W_{m+1}\big{\|}\cdot\|\nabla{\bf v}\|
m=0(𝐁(𝐯m)𝐁(𝐮~m))ΔWm+1𝐯\displaystyle\leq\big{\|}\sum_{m=0}^{\ell}({\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m}))\Delta W_{m+1}\|\cdot\|\nabla{\bf v}\|
+m=0((ξ𝐯mξ𝐮~m))ΔWm+1𝐯\displaystyle\qquad+\big{\|}\sum_{m=0}^{\ell}(\nabla(\xi_{{\bf v}}^{m}-\xi_{\tilde{{\bf u}}}^{m}))\Delta W_{m+1}\big{\|}\cdot\|\nabla{\bf v}\|
=(𝙸𝙸𝙸𝟷+𝙸𝙸𝙸𝟸)𝐯.\displaystyle=\bigl{(}{\tt III_{1}+III_{2}}\bigr{)}\,\|\nabla{\bf v}\|.

By Itô isometry, the assumptions on 𝐁{\bf B} and Theorem 3.3, we obtain

𝔼[𝙸𝙸𝙸𝟷𝟸]\displaystyle\mathbb{E}[{\tt III_{1}^{2}}] =km=0𝔼[𝐁(𝐯m)𝐁(𝐮~m)2]Ckm=0𝔼[𝐞𝐮~m2]Ck.\displaystyle=k\sum_{m=0}^{\ell}\mathbb{E}\big{[}\|{\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m})\|^{2}\big{]}\leq Ck\sum_{m=0}^{\ell}\mathbb{E}\big{[}\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\big{]}\leq Ck.
𝔼[𝙸𝙸𝙸𝟸𝟸]\displaystyle\mathbb{E}[{\tt III_{2}^{2}}] =km=0𝔼[(ξ𝐯mξ𝐮~m)2].\displaystyle=k\sum_{m=0}^{\ell}\mathbb{E}\big{[}\|\nabla(\xi_{{\bf v}}^{m}-\xi_{\tilde{{\bf u}}}^{m})\|^{2}\big{]}.

Moreover, using the Poisson equation as defined in Step 1 of Algorithm 1, we have

(ξ𝐯mξ𝐮~m)𝐁(𝐯m)𝐁(𝐮~m).\|\nabla(\xi_{{\bf v}}^{m}-\xi_{\tilde{{\bf u}}}^{m})\|\leq\|{\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m})\|.

Thus,

𝔼[𝙸𝙸𝙸22]\displaystyle\mathbb{E}[{\tt III}_{2}^{2}] km=0𝔼[𝐁(𝐯m)𝐁(𝐮~m)2]km=0𝔼[𝐞𝐮~m2]Ck.\displaystyle\leq k\sum_{m=0}^{\ell}\mathbb{E}[\|{\bf B}({\bf v}^{m})-{\bf B}(\tilde{{\bf u}}^{m})\|^{2}]\leq k\sum_{m=0}^{\ell}\mathbb{E}[\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}]\leq Ck.

Finally, it follows from the inf-sup condition (3.41), Theorem 3.3, (3.88), and all above inequalities that

(3.89) (𝔼[km=0(qm+1rm)2])12Ck.\displaystyle\bigg{(}\mathbb{E}\bigg{[}\bigg{\|}k\sum_{m=0}^{\ell}(q^{m+1}-r^{m})\bigg{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k}.

The proof is completed by applying the triangular inequality, Lemma 3.5 and (3.63b). ∎

Corollary 3.4.

Let {pm}m=0M\{p^{m}\}_{m=0}^{M} be generated by Algorithm 2. Then, there exists a constant C>0C>0 which depends on DT,𝐮0,𝐟D_{T},{\bf u}_{0},{\bf f} and β\beta such that for 0M0\leq\ell\leq M

(3.90) (𝔼[P(t)km=0pm2])12Ck.\displaystyle\bigg{(}\mathbb{E}\bigg{[}\bigg{\|}P(t_{\ell})-k\sum_{m=0}^{\ell}p^{m}\bigg{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\sqrt{k}.
Proof.

The proof is similar to that of Theorem 3.4. Indeed, we have

(3.91) P(tm)km=0pm\displaystyle\bigg{\|}P(t_{m})-k\sum_{m=0}^{\ell}p^{m}\bigg{\|} R(tm)km=0rm\displaystyle\leq\bigg{\|}R(t_{m})-k\sum_{m=0}^{\ell}r^{m}\bigg{\|}
+m=0tmtm+1ξ(s)𝑑W(s)m=0ξmΔWm+1\displaystyle\qquad+\bigg{\|}\sum_{m=0}^{\ell}\int_{t_{m}}^{t_{m+1}}\xi(s)\,dW(s)-\sum_{m=0}^{\ell}\xi^{m}\Delta W_{m+1}\bigg{\|}
=𝙸+𝙸𝙸.\displaystyle={\tt I}+{\tt II}.

Clearly, the first term “𝙸{\tt I}” can be bounded by using Theorem 3.4. To bound the second term “𝙸𝙸{\tt II}”, we first have

(3.92) 𝙸𝙸\displaystyle{\tt II} m=0(ξ(tm)ξm)ΔWm+1+m=0tmtm+1(ξ(s)ξ(tm))𝑑W(s)\displaystyle\leq\bigg{\|}\sum_{m=0}^{\ell}(\xi(t_{m})-\xi^{m})\Delta W_{m+1}\bigg{\|}+\bigg{\|}\sum_{m=0}^{\ell}\int_{t_{m}}^{t_{m+1}}(\xi(s)-\xi(t_{m}))\,dW(s)\bigg{\|}
=𝙸𝙸1+𝙸𝙸2.\displaystyle={\tt II}_{1}+{\tt II}_{2}.

Then by Itô isometry we get

(3.93) 𝔼[𝙸𝙸1]2\displaystyle\mathbb{E}[{\tt II}_{1}]^{2} =𝔼[km=0ξ(tm)ξm2]C𝔼[km=0𝐁(𝐮(tm))𝐁(𝐮~m)2]\displaystyle=\mathbb{E}\bigg{[}k\sum^{\ell}_{m=0}\|\xi(t_{m})-\xi^{m}\|^{2}\bigg{]}\leq C\mathbb{E}\bigg{[}k\sum^{\ell}_{m=0}\|{\bf B}({\bf u}(t_{m}))-{\bf B}(\tilde{{\bf u}}^{m})\|^{2}\bigg{]}
C𝔼[km=0𝐞𝐮~m2]Ck.\displaystyle\leq C\mathbb{E}\bigg{[}k\sum^{\ell}_{m=0}\|{\bf e}_{\tilde{{\bf u}}}^{m}\|^{2}\bigg{]}\leq Ck.

Similarly, by the Hölder continuity (2.3) we have

(3.94) 𝔼[𝙸𝙸2]2\displaystyle\mathbb{E}[{\tt II}_{2}]^{2} =m=0𝔼[tmtm+1ξ(s)ξ(tm)2𝑑s]\displaystyle=\sum_{m=0}^{\ell}\mathbb{E}\bigg{[}\int_{t_{m}}^{t_{m+1}}\|\xi(s)-\xi(t_{m})\|^{2}\,ds\bigg{]}
m=0𝔼[tmtm+1𝐮(s)𝐮(tm)2𝑑s]Ck,\displaystyle\leq\sum_{m=0}^{\ell}\mathbb{E}\bigg{[}\int_{t_{m}}^{t_{m+1}}\|{\bf u}(s)-{\bf u}(t_{m})\|^{2}\,ds\bigg{]}\leq Ck,

Finally, substituting (3.92)–(3.94) into (3.91) and using Theorem 3.4 give us the desired estimate. The proof is complete. ∎

4. Fully discrete finite element methods

In this section, we formulate and analyze finite element spatial discretization for Algorithm 1 and 2. To the end, let 𝒯h\mathcal{T}_{h} be a quasi-uniform triangulation of the polygonal (d=2d=2) or polyhedral (d=3d=3) bounded domain DD. We introduce the following two basic Lagrangian finite element spaces:

(4.1) Vh\displaystyle V_{h} ={ϕC(D¯);ϕ|KP(K)K𝒯h},\displaystyle=\{\phi\in C(\overline{D});\quad\phi_{|_{K}}\in P_{\ell}(K)\quad\forall K\in\mathcal{T}_{h}\},
(4.2) Xh\displaystyle X_{h} ={ϕC(D¯);ϕ|KP(K)K𝒯h},\displaystyle=\{\phi\in C(\overline{D});\quad\phi_{|_{K}}\in P_{\ell}(K)\quad\forall K\in\mathcal{T}_{h}\},

where P(K)P_{\ell}(K) (1)(\ell\geq 1) denotes the set of polynomials of degree less than or equal to \ell over the element K𝒯hK\in\mathcal{T}_{h}. The finite element spaces to be used to formulate our finite element methods are defined as follows:

(4.3) 𝐇h=[VhHper1(D)]d,Lh=VhLper2(D)/,Sh=XhLper2(D)/.\displaystyle{\bf H}_{h}=[V_{h}\cap H^{1}_{per}(D)]^{d},\quad L_{h}=V_{h}\cap L^{2}_{per}(D)/\mathbb{R},\quad S_{h}=X_{h}\cap L^{2}_{per}(D)/\mathbb{R}.

In addition, we introduce spaces

(4.4) 𝕍h=L2(Ω,𝐇h),𝕎h=L2(Ω,Lh).\displaystyle\mathbb{V}_{h}=L^{2}(\Omega,{\bf H}_{h}),\qquad\mathbb{W}_{h}=L^{2}(\Omega,L_{h}).

Recall that the L2L^{2}-projection 𝒫h0:[Lper2(D)]d𝐇h\mathcal{P}^{0}_{h}:[L^{2}_{per}(D)]^{d}\rightarrow{\bf H}_{h} is defined by

(4.5) (ϕ𝒫h0ϕ,ξ)=0ξ𝐇h\displaystyle(\phi-\mathcal{P}_{h}^{0}\phi,\,\xi)=0\qquad\forall\xi\in{\bf H}_{h}

and the H1H^{1}-projection 𝒫h1:Hper1(D)/Lh\mathcal{P}_{h}^{1}:H^{1}_{per}(D)/\mathbb{R}\rightarrow L_{h} is defined by

(4.6) (((χ𝒫h1χ),η)=0ηLh.\displaystyle\bigl{(}(\nabla(\chi-\mathcal{P}_{h}^{1}\chi),\nabla\eta\bigr{)}=0\qquad\forall\eta\in L_{h}.

It is well known [6] that 𝒫h0\mathcal{P}_{h}^{0} and 𝒫h1\mathcal{P}_{h}^{1} satisfy following estimates:

(4.7) ϕ𝒫h0ϕ+h(ϕ𝒫h0ϕ)\displaystyle\|\phi-\mathcal{P}_{h}^{0}\phi\|+h\|\nabla(\phi-\mathcal{P}_{h}^{0}\phi)\| Ch2ϕH2ϕ𝐇per2(D),\displaystyle\leq C\,h^{2}\|\phi\|_{H^{2}}\qquad\forall\phi\in{\bf H}^{2}_{per}(D),
(4.8) χ𝒫h1χ+h(χ𝒫h1χ)\displaystyle\|\chi-\mathcal{P}_{h}^{1}\chi\|+h\|\nabla(\chi-\mathcal{P}_{h}^{1}\chi)\| Ch2χH2χHper1/Hper2(D).\displaystyle\leq C\,h^{2}\|\chi\|_{H^{2}}\qquad\forall\chi\in H^{1}_{per}/\mathbb{R}\cap H^{2}_{per}(D).

For the clarity we only consider P1P_{1}-finite element space in this section (i.e., =1\ell=1), the results of this section can be easily extended to high order finite elements.

4.1. Finite element methods for the standard Chorin scheme

Approximating the velocity space and pressure space respectively by the finite element spaces 𝐇h{\bf H}_{h} and LhL_{h} in Algorithm 1, we then obtain the fully discrete finite element version of the standard Chorin scheme given below as Algorithm 3. We also note that a similar algorithm was proposed in [8].

Algorithm 3

Let n0n\geq 0. Set 𝐮~h0=𝒫h0𝐮0\tilde{{\bf u}}_{h}^{0}=\mathcal{P}_{h}^{0}{\bf u}_{0}. For n=0,1,2,n=0,1,2,\cdots do the following steps:

Step 1: Given 𝐮hnL2(Ω,𝐇h){\bf u}^{n}_{h}\in L^{2}(\Omega,{\bf H}_{h}) and 𝐮~hnL2(Ω,𝐇h)\tilde{{\bf u}}^{n}_{h}\in L^{2}(\Omega,{\bf H}_{h}), find 𝐮~hn+1L2(Ω,𝐇h)\tilde{{\bf u}}^{n+1}_{h}\in L^{2}(\Omega,{\bf H}_{h}) such that \mathbb{P}-a.s.

(4.9) (𝐮~hn+1,\displaystyle\bigl{(}\tilde{{\bf u}}^{n+1}_{h}, 𝐯h)+k(𝐮~hn+1,𝐯h)\displaystyle{\bf v}_{h}\bigr{)}\,+\,k\bigl{(}\nabla\tilde{{\bf u}}^{n+1}_{h},\nabla{\bf v}_{h}\bigr{)}
=(𝐮hn,𝐯h)+k(𝐟n+1,𝐯h)+(𝐁(𝐮~hn)ΔWn+1,𝐯h)𝐯h𝐇h.\displaystyle=\bigl{(}{\bf u}_{h}^{n},{\bf v}_{h}\bigr{)}\,+\,k\bigl{(}{\bf f}^{n+1},{\bf v}_{h}\bigr{)}\,+\,\bigl{(}{\bf B}(\tilde{{\bf u}}^{n}_{h})\Delta W_{n+1},\nabla{\bf v}_{h}\bigr{)}\qquad\forall{\bf v}_{h}\in{\bf H}_{h}.

Step 2: Find phn+1L2(Ω,Lh)p^{n+1}_{h}\in L^{2}(\Omega,L_{h}) such that \mathbb{P}-a.s.

(4.10) (phn+1,ϕh)=1k(𝐮~hn+1,ϕh)ϕhLh.\displaystyle\bigl{(}\nabla p_{h}^{n+1},\nabla\phi_{h}\bigr{)}=\frac{1}{k}\bigl{(}\tilde{{\bf u}}_{h}^{n+1},\nabla\phi_{h}\bigr{)}\qquad\forall\phi_{h}\in L_{h}.

Step 3: Define 𝐮hn+1L2(Ω,𝐇h){\bf u}^{n+1}_{h}\in L^{2}(\Omega,{\bf H}_{h}) by

(4.11) 𝐮hn+1=𝐮~hn+1kphn+1.\displaystyle{\bf u}_{h}^{n+1}=\tilde{{\bf u}}_{h}^{n+1}-k\nabla p_{h}^{n+1}.

As mentioned in Section 1, eliminating 𝐮n{\bf u}^{n} in (4.9) using (4.10), we obtain

(4.12a) (𝐮~hn+1\displaystyle(\tilde{{\bf u}}^{n+1}_{h} 𝐮~hn,𝐯h)+k(𝐮~hn+1,𝐯h)\displaystyle-\tilde{{\bf u}}^{n}_{h},{\bf v}_{h})\,+\,k(\nabla\tilde{{\bf u}}^{n+1}_{h},\nabla{\bf v}_{h})
+k(phn,𝐯h)=k(𝐟n+1,𝐯h)+(𝐁(𝐮~hn)ΔWn+1,𝐯h)𝐯h𝐇h,\displaystyle+k(\nabla p^{n}_{h},{\bf v}_{h})=k\big{(}{\bf f}^{n+1},{\bf v}_{h}\big{)}+({\bf B}(\tilde{{\bf u}}_{h}^{n})\Delta W_{n+1},{\bf v}_{h})\qquad\forall{\bf v}_{h}\in{\bf H}_{h},
(4.12b) (𝐮~hn+1,ϕh)=k(phn+1,ϕh)ϕhLh.\displaystyle(\tilde{{\bf u}}_{h}^{n+1},\nabla\phi_{h})=k(\nabla p_{h}^{n+1},\nabla\phi_{h})\qquad\forall\phi_{h}\in L_{h}.

Next, we state the stability estimates for {(𝐮~hn,phn)}n=0M\{(\tilde{{\bf u}}_{h}^{n},p_{h}^{n})\}_{n=0}^{M} in the following lemma, which will be used in the fully discrete error analysis later. Since its proof follows from the same lines of that for Lemma 3.1, we omit it to save space.

Lemma 4.1.

Let {(𝐮~hn,phn)}n=0M\{(\tilde{{\bf u}}_{h}^{n},p_{h}^{n})\}_{n=0}^{M} be generated by Algorithm 3, then there holds

(4.13a) max0nM𝔼[𝐮~hn2]+𝔼[n=1M𝐮~hn𝐮~hn12]+𝔼[kn=0M𝐮~hn2]\displaystyle\max_{0\leq n\leq M}\mathbb{E}[\|\tilde{{\bf u}}_{h}^{n}\|^{2}]+\mathbb{E}\bigg{[}\sum_{n=1}^{M}\|\tilde{{\bf u}}_{h}^{n}-\tilde{{\bf u}}_{h}^{n-1}\|^{2}\bigg{]}+\mathbb{E}\bigg{[}k\sum_{n=0}^{M}\|\nabla\tilde{{\bf u}}_{h}^{n}\|^{2}\bigg{]} C,\displaystyle\leq C,
(4.13b) 𝔼[kn=0Mphn2]\displaystyle\mathbb{E}\bigg{[}k\sum_{n=0}^{M}\|\nabla p_{h}^{n}\|^{2}\bigg{]} Ck,\displaystyle\leq\frac{C}{k},

where CC is a positive constant depending only on DT,u0D_{T},u_{0}, and 𝐟{\bf f}.

The following theorem provides an error estimate in a strong norm for the finite element solution of Algorithm 3.

Theorem 4.1.

Let {(𝐮~m,pm)}m=0M\{\big{(}\tilde{{\bf u}}^{m},p^{m}\big{)}\}_{m=0}^{M} and {(𝐮~hm,phm)}m=0M\{(\tilde{{\bf u}}_{h}^{m},p_{h}^{m})\}_{m=0}^{M} be generated respectively by Algorithm 1 and Algorithm 3. Then under the assumptions of Lemmas 3.1, 4.1 and Corollary 3.1 there holds

(4.14) (𝔼[km=0M𝐮~m𝐮~hm2])12\displaystyle\bigg{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\bigl{\|}\tilde{{\bf u}}^{m}-\tilde{{\bf u}}_{h}^{m}\bigr{\|}^{2}\Bigr{]}\bigg{)}^{\frac{1}{2}}
+max0M(𝔼[kn=1(𝐮~n𝐮~hn)2])12C(k14+hk12),\displaystyle\qquad\qquad+\max_{0\leq\ell\leq M}\bigg{(}\mathbb{E}\Bigl{[}\Bigr{\|}k\sum_{n=1}^{\ell}\nabla(\tilde{{\bf u}}^{n}-\tilde{{\bf u}}_{h}^{n})\Bigr{\|}^{2}\Bigr{]}\bigg{)}^{\frac{1}{2}}\leq C\Bigl{(}k^{\frac{1}{4}}\,+\,hk^{-\frac{1}{2}}\Bigr{)},

where CC(DT,𝐮0,𝐟)C\equiv C(D_{T},{\bf u}_{0},{\bf f}) is a positive constant.

Proof.

The proof is conceptually similar to that of Theorem 3.1. Setting 𝐞𝐮~hm=:𝐮~m𝐮~hm{\bf e}_{\tilde{{\bf u}}_{h}}^{m}=:\tilde{{\bf u}}^{m}-\tilde{{\bf u}}_{h}^{m} and εphm=:pmphm\varepsilon_{p_{h}}^{m}=:p^{m}-p^{m}_{h}. Without loss of the generality, we assume 𝐞𝐮~h0=0{\bf e}_{\tilde{{\bf u}}_{h}}^{0}=0 and εph0=0\varepsilon_{p_{h}}^{0}=0 because they are of high order accuracy, hence are negligible.

First, applying the summation operator n=0m\sum_{n=0}^{m} to (4.12a), we obtain

(4.15) (𝐮~hm+1,𝐯h)\displaystyle\bigl{(}\tilde{{\bf u}}_{h}^{m+1},{\bf v}_{h}\bigr{)} +k(n=0m+1𝐮~hn,𝐯h)+k(n=0mphn,𝐯h)\displaystyle+k\Bigl{(}\sum_{n=0}^{m+1}\nabla\tilde{{\bf u}}_{h}^{n},\nabla{\bf v}_{h}\Bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla p^{n}_{h},{\bf v}_{h}\Bigr{)}
=(𝐮h0,𝐯h)+(n=0m𝐁(𝐮~hn)ΔWn+1,𝐯h)𝐯h𝐇h.\displaystyle=\bigl{(}{\bf u}^{0}_{h},{\bf v}_{h}\bigr{)}+\Bigl{(}\sum_{n=0}^{m}{\bf B}(\tilde{{\bf u}}_{h}^{n})\Delta W_{n+1},{\bf v}_{h}\Bigr{)}\qquad\forall{\bf v}_{h}\in{\bf H}_{h}.

Subtracting (3.19) from (4.15) yields the following error equations:

(4.16) (𝐞𝐮~hm+1,𝐯h)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},{\bf v}_{h}\bigr{)} +k(n=0m+1𝐞𝐮~hn,𝐯h)+k(n=0mεphn,𝐯h)\displaystyle+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla{\bf v}_{h}\Bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},{\bf v}_{h}\Bigr{)}
=(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝐯h)𝐯h𝐇h,\displaystyle=\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},{\bf v}_{h}\Bigr{)}\qquad\forall{\bf v}_{h}\in{\bf H}_{h},
(4.17) (𝐞𝐮~hm+1,ϕh)\displaystyle\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\nabla\phi_{h}\bigr{)} =k(εphm+1,ϕh)ϕhLh.\displaystyle=k\bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\nabla\phi_{h}\bigr{)}\qquad\forall\phi_{h}\in L_{h}.

Choosing 𝐯h=𝒫h0𝐞𝐮~hm+1=𝐞𝐮~hm+1𝜽m+1{\bf v}_{h}=\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}={\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}-\boldsymbol{\theta}^{m+1}; 𝜽m=𝐮~hm𝒫h0𝐮~hm\boldsymbol{\theta}^{m}=\tilde{{\bf u}}_{h}^{m}-\mathcal{P}_{h}^{0}\tilde{{\bf u}}_{h}^{m}, then (4.16) becomes

(4.18) 𝐞𝐮~hm+12\displaystyle\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2} +k(n=0m+1𝐞𝐮~hn,𝐞𝐮~hm+1)+k(n=0mεphn,𝒫h0𝐞𝐮~hm+1)\displaystyle+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}
=(𝐞𝐮~hm+1,𝜽m+1)+k(n=0m+1𝐞𝐮~hn,𝜽m+1)\displaystyle=\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\boldsymbol{\theta}^{m+1}\bigr{)}+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla\boldsymbol{\theta}^{m+1}\Bigr{)}
+(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝒫h0𝐞𝐮~hm+1).\displaystyle\qquad+\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}.

Setting ϕh=n=0m𝒫h1εphn=n=0mεphnn=0mξn\phi_{h}=\sum_{n=0}^{m}\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n}=\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n}-\sum_{n=0}^{m}\xi^{n} in (4.17), where ξn=pn𝒫h1pn\xi^{n}=p^{n}-\mathcal{P}_{h}^{1}p^{n}, we obtain

(4.19) (𝐞𝐮~hm+1,n=0m𝒫h1εphn)=k(εphm+1,n=0m𝒫h1εphn)\displaystyle\Bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\sum_{n=0}^{m}\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n}\Bigr{)}=k\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m}\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n}\Bigr{)}

In addition, by using the properties of 𝒫h0\mathcal{P}_{h}^{0}- and 𝒫h1\mathcal{P}_{h}^{1}-projection we have

(4.20) k(n=0mεphn,𝒫h0𝐞𝐮~hm+1)=k(n=0mεphn,𝐞𝐮~hm+1)k(n=0mεphn,𝜽m+1)\displaystyle k\biggl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\biggr{)}=k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}
=k(n=0m𝒫h1εphn,𝐞𝐮~hm+1)+k(n=0mξn,𝐞𝐮~hm+1)k(n=0mεphn,𝜽m+1)\displaystyle\quad=k\Bigl{(}\sum_{n=0}^{m}\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla\xi^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}
=k2(εphm+1,n=0m𝒫h1εphn)+k(n=0mξn,𝐞𝐮~hm+1)k(n=0mεphn,𝜽m+1)\displaystyle\quad=k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m}\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n}\Bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla\xi^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}
=k2(εphm+1,n=0m+1𝒫h1εphn)k2(εphm+1,𝒫h1εphm+1)\displaystyle\quad=k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{n}\Bigr{)}-k^{2}\bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\bigr{)}
+k(n=0mξn,𝐞𝐮~hm+1)k(n=0mεphn,𝜽m+1)\displaystyle\qquad+k\Bigl{(}\sum_{n=0}^{m}\nabla\xi^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}
=k2(εphm+1,n=0m+1εphn)k2(εphm+1,n=0m+1ξn)\displaystyle\quad=k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{)}-k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}
k2(εphm+1,𝒫h1εphm+1)+k(n=0mξn,𝐞𝐮~hm+1)k(n=0mεphn,𝜽m+1).\displaystyle\qquad-k^{2}\bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\bigr{)}+k\Bigl{(}\sum_{n=0}^{m}\nabla\xi^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}.

Moreover, by using the orthogonality property of 𝒫h1\mathcal{P}_{h}^{1}, we have

(4.21) k2(εphm+1,𝒫h1εphm+1)\displaystyle-k^{2}\bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\bigr{)} =k2((εphm+1𝒫h1εph),𝒫h1εphm+1)\displaystyle=-k^{2}\bigl{(}\nabla(\varepsilon_{p_{h}}^{m+1}-\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}),\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\bigr{)}
+k2𝒫h1εphm+12=k2𝒫h1εphm+12,\displaystyle\qquad+k^{2}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}=k^{2}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2},

which helps to reduce (4.20) into

(4.22) k(n=0mεphn,𝒫h0𝐞𝐮~hm+1)\displaystyle k\biggl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\biggr{)} =k2(εphm+1,n=0m+1εphn)k2(εphm+1,n=0m+1ξn)\displaystyle=k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{)}-k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}
+k2𝒫h1εphm+12+k(n=0mξn,𝐞𝐮~hm+1)\displaystyle\quad+k^{2}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}+k\Bigl{(}\sum_{n=0}^{m}\nabla\xi^{n},{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}
k(n=0mεphn,𝜽m+1).\displaystyle\quad-k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}.

Substituting (4.22) into (4.18) and rearranging terms yield

(4.23) 𝐞𝐮~hm+12+k(n=0m+1𝐞𝐮~hn,𝐞𝐮~hm+1)+k2(εphm+1,n=0m+1εphn)\displaystyle\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}+k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{)}
+k2𝒫h1εphm+12\displaystyle\qquad+k^{2}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}
=k2(εphm+1,n=0m+1ξn)+k(div 𝐞𝐮~hm+1,n=0mξn)\displaystyle\quad=k^{2}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}+k\Bigl{(}\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\sum_{n=0}^{m}\xi^{n}\Bigr{)}
+k(n=0mεphn,𝜽m+1)+(𝐞𝐮~hm+1,𝜽m+1)+k(n=0m+1𝐞𝐮~hn,𝜽m+1)\displaystyle\qquad+k\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}+\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\boldsymbol{\theta}^{m+1}\bigr{)}+k\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla\boldsymbol{\theta}^{m+1}\Bigr{)}
+(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝒫h0𝐞𝐮~hm+1).\displaystyle\qquad+\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}.

Next, we use the identity 2a(ab)=a2b2+(ab)22a(a-b)=a^{2}-b^{2}+(a-b)^{2} to create telescoping sums on the left side of (4.23) followed by taking the expectation to get

(4.24) 𝔼[𝐞𝐮~hm+12]+k2𝔼[n=0m+1𝐞𝐮~hn2n=0m𝐞𝐮~hn2+𝐞𝐮~hm+12]\displaystyle\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+\frac{k}{2}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\biggr{\|}^{2}-\biggl{\|}\sum_{n=0}^{m}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\biggr{\|}^{2}+\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\biggr{]}
+k22𝔼[n=0m+1εphn2n=0mεphn2+εphm+12]\displaystyle\qquad+\frac{k^{2}}{2}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m+1}\nabla\varepsilon_{p_{h}}^{n}\biggr{\|}^{2}-\bigg{\|}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n}\biggr{\|}^{2}+\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}\biggr{]}
+k2𝔼[𝒫h1εphm+12]\displaystyle\qquad+k^{2}\mathbb{E}\bigl{[}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}\bigr{]}
=k2𝔼[(εphm+1,n=0m+1ξn)]+k𝔼[(div 𝐞𝐮~hm+1,n=0mξn)]\displaystyle\quad=k^{2}\mathbb{E}\Bigl{[}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}\Bigr{]}+k\mathbb{E}\Bigl{[}\Bigl{(}\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\sum_{n=0}^{m}\xi^{n}\Bigr{)}\Bigr{]}
+k𝔼[(n=0mεphn,𝜽m+1)]+𝔼[(𝐞𝐮~hm+1,𝜽m+1)]\displaystyle\qquad+k\mathbb{E}\Bigl{[}\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}\Bigr{]}+\mathbb{E}\Bigl{[}\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\boldsymbol{\theta}^{m+1}\bigr{)}\Bigr{]}
+k𝔼[(n=0m+1𝐞𝐮~hn,𝜽m+1)]\displaystyle\qquad+k\mathbb{E}\Bigl{[}\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla\boldsymbol{\theta}^{m+1}\Bigr{)}\Bigr{]}
+𝔼[(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝒫h0𝐞𝐮~hm+1)].\displaystyle\qquad+\mathbb{E}\Bigl{[}\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}\Bigr{]}.

Now, applying km=0k\sum_{m=0}^{\ell} for 0<M0\leq\ell<M to (4.24) we obtain

(4.25) km=0𝔼[𝐞𝐮~hm+12]+12𝔼[kn=0+1𝐞𝐮~hn2]+k22m=0𝔼[𝐞𝐮~hm+12]\displaystyle k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+\frac{1}{2}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{k^{2}}{2}\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\Bigr{]}
+k2𝔼[kn=0+1εphn2]+k32m=0𝔼[εphm+12+2𝒫h1εphm+12]\displaystyle\qquad+\frac{k}{2}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{k^{3}}{2}\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}+2\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}\Bigr{]}
=𝔼[k3m=0(εphm+1,n=0m+1ξn)]+𝔼[k2m=0(div 𝐞𝐮~hm+1,n=0mξn)]\displaystyle\quad=\mathbb{E}\Bigl{[}k^{3}\sum_{m=0}^{\ell}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}\Bigr{]}+\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\sum_{n=0}^{m}\xi^{n}\Bigr{)}\Bigr{]}
+𝔼[k2m=0(n=0mεphn,𝜽m+1)]+𝔼[km=0(𝐞𝐮~hm+1,𝜽m+1)]\displaystyle\qquad+\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}\Bigr{]}+\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\boldsymbol{\theta}^{m+1}\bigr{)}\Bigr{]}
+𝔼[k2m=0(n=0m+1𝐞𝐮~hn,𝜽m+1)]\displaystyle\qquad+\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla\boldsymbol{\theta}^{m+1}\Bigr{)}\Bigr{]}
+km=0𝔼[(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝒫h0𝐞𝐮~hm+1)]\displaystyle\qquad+k\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}\Bigr{]}
:=𝙸++𝚅𝙸.\displaystyle\quad:={\tt I+\cdots+VI}.

Next, we bound the right-hand side of (4.25) as follows. By using the discrete Hölder inequality we get

(4.26) 𝙸\displaystyle{\tt I} =k𝔼[km=0(εphm+1,kn=0m+1ξn)]\displaystyle=k\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{(}\nabla\varepsilon_{p_{h}}^{m+1},k\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{)}\Bigr{]}
k𝔼[km=0εphm+1kn=0m+1ξn]\displaystyle\leq k\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla\varepsilon_{p_{h}}^{m+1}\|\biggl{\|}k\sum_{n=0}^{m+1}\nabla\xi^{n}\biggr{\|}\biggr{]}
k𝔼[(km=0εphm+12)12(km=0kn=0m+1ξn2)12]\displaystyle\leq k\mathbb{E}\Bigl{[}\Bigl{(}k\sum_{m=0}^{\ell}\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}\Bigr{)}^{\frac{1}{2}}\Bigl{(}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{\|}^{2}\Bigr{)}^{\frac{1}{2}}\Bigr{]}
k(𝔼[km=0εphm+12])12(𝔼[km=0kn=0m+1ξn2])12.\displaystyle\leq k\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}\Bigr{]}\Bigr{)}^{\frac{1}{2}}\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{\|}^{2}\Bigr{]}\Bigr{)}^{\frac{1}{2}}.

In addition, by using the stability estimates from (3.5b) and (4.13b), we have

𝔼[km=0εphm+12]Ck,\displaystyle\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}\Bigr{]}\leq\frac{C}{k},

Similarly, using (4.8) and the stability estimate from (3.50) we get

𝔼[km=0kn=0m+1ξn2]\displaystyle\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m+1}\nabla\xi^{n}\Bigr{\|}^{2}\biggr{]} =𝔼[km=0(kn=0m+1pn𝒫h1(kn=0m+1pn))2]\displaystyle=\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}\nabla\Bigl{(}k\sum_{n=0}^{m+1}p^{n}-\mathcal{P}_{h}^{1}\Bigl{(}k\sum_{n=0}^{m+1}p^{n}\Bigr{)}\Bigr{)}\Bigr{\|}^{2}\biggr{]}
C𝔼[km=0kn=0m+1pn2]C.\displaystyle\leq C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m+1}\nabla p^{n}\Bigr{\|}^{2}\biggr{]}\leq C.

Therefore, 𝙸Ck12{\tt I}\leq Ck^{\frac{1}{2}}.

Next, using the fact that n=0mξnChn=0mpn\Bigl{\|}\sum_{n=0}^{m}\xi^{n}\Bigr{\|}\leq Ch\Bigl{\|}\sum_{n=0}^{m}\nabla p^{n}\Bigr{\|} we have

(4.27) 𝙸𝙸\displaystyle{\tt II} =𝔼[k2m=0(div 𝐞𝐮~hm+1,n=0mξn)]\displaystyle=\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\mbox{\rm div\,}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\sum_{n=0}^{m}\xi^{n}\Bigr{)}\Bigr{]}
k28m=0𝔼[𝐞𝐮~hm+12]+C𝔼[k2m=0n=0mξn2]\displaystyle\leq\frac{k^{2}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+C\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\xi^{n}\Bigr{\|}^{2}\Bigr{]}
k28m=0𝔼[𝐞𝐮~hm+12]+Ch2k𝔼[km=0kn=0mpn2]\displaystyle\leq\frac{k^{2}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+\frac{Ch^{2}}{k}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m}\nabla p^{n}\Bigr{\|}^{2}\Bigr{]}
k28m=0𝔼[𝐞𝐮~hm+12]+Ch2k,\displaystyle\leq\frac{k^{2}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+\frac{Ch^{2}}{k},

where (3.50) was used to obtain the last inequality. Expectedly, the first term will be absorbed to the left-hand side of (4.25) later.

Next, using summation by parts we obtain

(4.28) 𝙸𝙸𝙸\displaystyle{\tt III} =𝔼[k2m=0(n=0mεphn,𝜽m+1)]\displaystyle=\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n},\boldsymbol{\theta}^{m+1}\Bigr{)}\Bigr{]}
=𝔼[k2(n=0+1𝜽n,n=0+1εphn)]𝔼[k2m=0(n=0m𝜽n,εphm)]\displaystyle=\mathbb{E}\Bigl{[}k^{2}\Bigl{(}\sum_{n=0}^{\ell+1}\boldsymbol{\theta}^{n},\sum_{n=0}^{\ell+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{)}\Bigr{]}-\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m}\boldsymbol{\theta}^{n},\nabla\varepsilon_{p_{h}}^{m}\Bigr{)}\Bigr{]}
Ck𝔼[n=0+1𝜽n2]+k38𝔼[n=0+1εphn2]\displaystyle\leq Ck\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{\ell+1}\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{k^{3}}{8}\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{\ell+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}
+C𝔼[km=0n=0m𝜽n2]+k38m=0𝔼[εphm2].\displaystyle\qquad+C\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{k^{3}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla\varepsilon_{p_{h}}^{m}\|^{2}\bigr{]}.

In addition, we can use (4.7) and (3.51) to control the first and third terms on the right side of (4.28) as follows:

(4.29) Ck𝔼[n=0+1𝜽n2]+C𝔼[km=0n=0m𝜽n2]\displaystyle Ck\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{\ell+1}\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}+C\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}
Ch4k𝔼[n=0+1Δ𝐮~n2]+Ch4𝔼[km=0n=0mΔ𝐮~n2]Ch4k2.\displaystyle\qquad\leq Ch^{4}k\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{\ell+1}\Delta\tilde{{\bf u}}^{n}\Bigr{\|}^{2}\Bigr{]}+Ch^{4}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n}\Bigr{\|}^{2}\Bigr{]}\leq\frac{Ch^{4}}{k^{2}}.

Therefore,

(4.30) 𝙸𝙸𝙸Ch4k2+k38𝔼[n=0+1εphn2]+k38m=0𝔼[εphm2],\displaystyle{\tt III}\leq\frac{Ch^{4}}{k^{2}}+\frac{k^{3}}{8}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{\ell+1}\nabla\varepsilon_{p_{h}}^{n}\biggr{\|}^{2}\biggr{]}+\frac{k^{3}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla\varepsilon_{p_{h}}^{m}\|^{2}\bigr{]},

Again, expectedly, the last two terms on the right-hand side of (4.30) will be absorbed to the left side of (4.25) later.

It follows from (4.7) and (3.5c) that

(4.31) 𝙸𝚅\displaystyle{\tt IV} =𝔼[km=0(𝐞𝐮~hm+1,𝜽m+1)]\displaystyle=\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},\boldsymbol{\theta}^{m+1}\bigr{)}\biggr{]}
km=0𝔼[𝜽m+12]+14km=0𝔼[𝐞𝐮~hm+12]\displaystyle\leq k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\boldsymbol{\theta}^{m+1}\|^{2}\bigr{]}+\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}
Ch4𝔼[km=0Δ𝐮~m+12]+14km=0𝔼[𝐞𝐮~hm+12]\displaystyle\leq Ch^{4}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\|\Delta\tilde{{\bf u}}^{m+1}\|^{2}\Bigr{]}+\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}
Ch4k+14km=0𝔼[𝐞𝐮~hm+12].\displaystyle\leq\frac{Ch^{4}}{k}+\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}.

To estimate term V, we approach similarly as done for term III. Namely, fist we use the summation by parts and then use (4.7) and (3.50).

(4.32) 𝚅\displaystyle{\tt V} =𝔼[k2m=0(n=0m+1𝐞𝐮~hn,𝜽m+1)]\displaystyle=\mathbb{E}\biggl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla\boldsymbol{\theta}^{m+1}\biggr{)}\Bigr{]}
=𝔼[k2(n=0+1𝐞𝐮~hn,n=0+1𝜽n)]𝔼[k2m=0(n=0m𝜽n,𝐞𝐮~hm+1)]\displaystyle=\mathbb{E}\biggl{[}k^{2}\Bigl{(}\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\sum_{n=0}^{\ell+1}\nabla\boldsymbol{\theta}^{n}\Bigr{)}\biggr{]}-\mathbb{E}\biggl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{(}\sum_{n=0}^{m}\nabla\boldsymbol{\theta}^{n},\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}\biggr{]}
18𝔼[kn=0+1𝐞𝐮~hn2]+C𝔼[kn=0+1𝜽n2]\displaystyle\leq\frac{1}{8}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+C\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}
+18𝔼[k2m=0𝐞𝐮~hm+12]+C𝔼[k2m=0n=0m𝜽n2]\displaystyle\quad+\frac{1}{8}\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\Bigr{]}+C\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\Bigl{\|}\sum_{n=0}^{m}\nabla\boldsymbol{\theta}^{n}\Bigr{\|}^{2}\Bigr{]}
18𝔼[kn=0+1𝐞𝐮~hn2]+18𝔼[k2m=0𝐞𝐮~hm+12]\displaystyle\leq\frac{1}{8}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{1}{8}\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\Bigr{]}
+Ch2𝔼[kn=0mΔ𝐮~n2]+Ch2k𝔼[km=0kn=0mΔ𝐮~n2]\displaystyle\quad+Ch^{2}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{Ch^{2}}{k}\mathbb{E}\Bigl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m}\Delta\tilde{{\bf u}}^{n}\Bigr{\|}^{2}\Bigr{]}
18𝔼[kn=0+1𝐞𝐮~hn2]+18𝔼[k2m=0𝐞𝐮~hm+12]+Ch2k.\displaystyle\leq\frac{1}{8}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{1}{8}\mathbb{E}\Bigl{[}k^{2}\sum_{m=0}^{\ell}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\Bigr{]}+\frac{Ch^{2}}{k}.

We use the Itô isometry to handle the noise term as follows:

(4.33) 𝚅𝙸\displaystyle{\tt VI} =km=0𝔼[(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝒫h0𝐞𝐮~hm+1)]\displaystyle=k\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},\mathcal{P}_{h}^{0}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\Bigr{)}\Bigr{]}
14km=0𝔼[𝐞𝐮~hm+12]+km=0𝔼[n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+12]\displaystyle\leq\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+k\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1}\Bigr{\|}^{2}\Bigr{]}
=14km=0𝔼[𝐞𝐮~hm+12]+km=0𝔼[kn=0m𝐁(𝐮~n)𝐁(𝐮~hn)2]\displaystyle=\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+k\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}k\sum_{n=0}^{m}\|{\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n})\|^{2}\Bigr{]}
14km=0𝔼[𝐞𝐮~hm+12]+Ckm=0kn=0m𝔼[𝐞𝐮~hn2].\displaystyle\leq\frac{1}{4}k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+Ck\sum_{m=0}^{\ell}k\sum_{n=0}^{m}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\|^{2}\bigr{]}.

Finally, substituting the above estimates for terms I, II, III, IV, V, VI into (4.25) yields the following inequality for X=km=0𝔼[𝐞𝐮~hm2]X^{\ell}=k\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m}\|^{2}\bigr{]}:

(4.34) 12X+1+38𝔼[kn=0+1𝐞𝐮~hn2]+k24m=0𝔼[𝐞𝐮~hm+12]\displaystyle\frac{1}{2}X^{\ell+1}+\frac{3}{8}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{k^{2}}{4}\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\|\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\Bigr{]}
+3k8𝔼[kn=0+1εphn2]+3k38m=0𝔼[εphm+12]\displaystyle\qquad+\frac{3k}{8}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{\ell+1}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}+\frac{3k^{3}}{8}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla\varepsilon_{p_{h}}^{m+1}\|^{2}\bigr{]}
+k3m=0𝔼[𝒫h1εphm+12]\displaystyle\qquad+k^{3}\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|\nabla\mathcal{P}_{h}^{1}\varepsilon_{p_{h}}^{m+1}\|^{2}\bigr{]}
Ck12+Ch2k+Ckm=0Xm.\displaystyle\leq Ck^{\frac{1}{2}}+\frac{Ch^{2}}{k}+Ck\sum_{m=0}^{\ell}X^{m}.

The desired error estimate (4.14) then follows from an application of the discrete Gronwall inequality to (4.34). The proof is complete. ∎

Next, we state an error estimate result for the pressure approximation generated by Algorithm 3 in a time-averaged fashion. Recall that an important advantage of Chorin-type schemes is to allow the use of a pair of independent finite element spaces which are not required to satisfy a discrete inf-sup condition, a price for this advantage is to make error estimates for the pressure approximations become more complicated even in the deterministic case. The idea for circumventing the difficulty is to utilize the following perturbed inf-sup inequality (cf. [16]): there exists δ>0\delta>0 independent of h>0h>0, such that

(4.35) 1δ2qh2sup𝐯h𝐇h|(qh,div𝐯h)|2𝐯h2+h2qh2qhSh,\frac{1}{\delta^{2}}\|q_{h}\|^{2}\leq\sup_{{\bf v}_{h}\in{{\bf H}}_{h}}\frac{|(q_{h},{\rm div}\,{\bf v}_{h})|^{2}}{\|\nabla{\bf v}_{h}\|^{2}}+h^{2}\|\nabla q_{h}\|^{2}\qquad\forall\,q_{h}\in S_{h}\,,

which was also used in [13] to derive an error estimate for a pressure-stabilization scheme for (1.1).

Theorem 4.2.

Under the assumptions of Theorem 4.1, there exists a positive constant CC(DT,𝐮0,𝐟,δ)C\equiv C(D_{T},{\bf u}_{0},{\bf f},\delta) such that

(4.36) (𝔼[km=0Mkm=1m(pnphn)2])12C(k14+hk12),\displaystyle\bigg{(}\mathbb{E}\bigg{[}k\sum_{m=0}^{M}\Bigl{\|}k\sum_{m=1}^{m}\bigl{(}p^{n}-p_{h}^{n}\bigr{)}\Bigr{\|}^{2}\bigg{]}\bigg{)}^{\frac{1}{2}}\leq C\Bigl{(}k^{\frac{1}{4}}\,+\,hk^{-\frac{1}{2}}\Bigr{)},
Proof.

We reuse all the notations from the proof of Theorem 4.1. First, from the error equations (4.16) we have

(4.37) (kn=0mεphn,div 𝐯h)\displaystyle\Bigl{(}k\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n},\mbox{\rm div\,}{\bf v}_{h}\Bigr{)} =(𝐞𝐮~hm+1,𝐯h)+(kn=0m+1𝐞𝐮~hn,𝐯h)\displaystyle=\bigl{(}{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1},{\bf v}_{h}\bigr{)}+\Bigl{(}k\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n},\nabla{\bf v}_{h}\Bigr{)}
(n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1,𝐯h)𝐯h𝐇h.\displaystyle\qquad-\Bigl{(}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1},{\bf v}_{h}\Bigr{)}\,\,\forall{\bf v}_{h}\in{\bf H}_{h}.

Using the Schwarz inequality on the right-hand side of (4.37) yields

(4.38) |(kn=0mεphn,div 𝐯h)|\displaystyle\Bigl{|}\Bigl{(}k\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n},\mbox{\rm div\,}{\bf v}_{h}\Bigr{)}\Bigr{|} =C𝐞𝐮~hm+1𝐯h+kn=0m+1𝐞𝐮~hn𝐯h\displaystyle=C\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|\|\nabla{\bf v}_{h}\|+\Bigl{\|}k\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}\|\nabla{\bf v}_{h}\|
+Cn=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+1𝐯h.\displaystyle\quad+C\Bigl{\|}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1}\Bigr{\|}\|\nabla{\bf v}_{h}\|.

Next, using (4.35) we conclude that

(4.39) 1δ2kn=0mεphn2h2kn=0mεphn2\displaystyle\frac{1}{\delta^{2}}\Bigl{\|}k\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}-h^{2}\Bigl{\|}k\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2} sup𝐯h𝐇h|(kn=0mεphn,div 𝐯h)|2𝐯h2\displaystyle\leq\sup_{{\bf v}_{h}\in{{\bf H}}_{h}}\frac{\Bigl{|}\Bigl{(}k\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n},\mbox{\rm div\,}{\bf v}_{h}\Bigr{)}\Bigr{|}^{2}}{\|\nabla{\bf v}_{h}\|^{2}}
C𝐞𝐮~hm+12+Ckn=0m+1𝐞𝐮~hn2\displaystyle\leq C\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}+C\Bigl{\|}k\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}
+Cn=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+12.\displaystyle+C\Bigl{\|}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1}\Bigr{\|}^{2}.

Then, applying operators km=0k\sum_{m=0}^{\ell} and 𝔼[]\mathbb{E}[\cdot] on both sides we obtain

(4.40) 1δ2𝔼[km=0kn=0mεphn2]\displaystyle\frac{1}{\delta^{2}}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m}\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\biggr{]} h2𝔼[km=0kn=0mεphn2]\displaystyle\leq h^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\biggr{]}
+Ckm=0𝔼[𝐞𝐮~hm+12]\displaystyle\quad+Ck\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}
+Ckm=0𝔼[kn=0m+1𝐞𝐮~hn2]\displaystyle\quad+Ck\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\Bigr{\|}^{2}\Bigr{]}
+Ckm=0𝔼[n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+12]\displaystyle\quad+Ck\sum_{m=0}^{\ell}\mathbb{E}\Bigl{[}\Bigl{\|}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1}\Bigr{\|}^{2}\Bigr{]}
:=𝙸+𝙸𝙸+𝙸𝙸𝙸+𝙸𝚅.\displaystyle\quad:={\tt I+II+III+IV}.

We now bound each terms on the right-hand side of (4.40). By using the discrete Jensen inequality and the stability estimates from (3.5b) and(4.13b) we get

(4.41) 𝙸\displaystyle{\tt I} =h2𝔼[km=0kn=0mεphn2]h2𝔼[km=0kn=0mεphn2]Ch2k.\displaystyle=h^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\Bigl{\|}k\sum_{n=0}^{m}\nabla\varepsilon_{p_{h}}^{n}\Bigr{\|}^{2}\biggr{]}\leq h^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}k\sum_{n=0}^{m}\|\nabla\varepsilon_{p_{h}}^{n}\|^{2}\biggr{]}\leq\frac{Ch^{2}}{k}.

Using Theorem 4.1, terms II and III can be bounded as follows:

(4.42) 𝙸𝙸+𝙸𝙸𝙸\displaystyle{\tt II+III} =Ckm=0𝔼[𝐞𝐮~hm+12]+Ckm=0𝔼[kn=0m+1𝐞𝐮~hn2]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\bigl{[}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{m+1}\|^{2}\bigr{]}+Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{n=0}^{m+1}\nabla{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\biggr{\|}^{2}\biggr{]}
C(k12+Ch2k).\displaystyle\leq C\Bigl{(}k^{\frac{1}{2}}+\frac{Ch^{2}}{k}\Bigr{)}.

Finally, using Itô’s isometry and Theorem 4.1 we have

(4.43) 𝙸𝚅\displaystyle{\tt IV} =Ckm=0𝔼[n=0m(𝐁(𝐮~n)𝐁(𝐮~hn))ΔWn+12]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}\biggl{\|}\sum_{n=0}^{m}({\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n}))\Delta W_{n+1}\biggr{\|}^{2}\biggr{]}
=Ckm=0𝔼[kn=0m𝐁(𝐮~n)𝐁(𝐮~hn)2]\displaystyle=Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}k\sum_{n=0}^{m}\|{\bf B}(\tilde{{\bf u}}^{n})-{\bf B}(\tilde{{\bf u}}_{h}^{n})\|^{2}\biggr{]}
Ckm=0𝔼[kn=0m𝐞𝐮~hn2]\displaystyle\leq Ck\sum_{m=0}^{\ell}\mathbb{E}\biggl{[}k\sum_{n=0}^{m}\|{\bf e}_{\tilde{{\bf u}}_{h}}^{n}\|^{2}\biggr{]}
C(k12+Ch2k).\displaystyle\leq C\Bigl{(}k^{\frac{1}{2}}+\frac{Ch^{2}}{k}\Bigr{)}.

The proof is complete after combining the above estimates. ∎

We are now ready to state the following global error estimate theorem for Algorithm 3 which is a main result of this paper.

Theorem 4.3.

Under the assumptions of Theorems 3.1, 3.2 and Theorems 4.1 and 4.2, there hold the following error estimates:

(4.44) (𝔼[km=0M𝐮(tm)𝐮~hm2])12\displaystyle\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\|{\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h}\|^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}} +max0M(𝔼[km=0(𝐮(tm)𝐮~hm)2])12\displaystyle+\max_{0\leq\ell\leq M}\biggl{(}\mathbb{E}\Bigl{[}\Bigl{\|}k\sum_{m=0}^{\ell}\nabla({\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h})\Bigr{\|}^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}
C(k14+hk12),\displaystyle\qquad\leq C\Bigl{(}k^{\frac{1}{4}}+hk^{-\frac{1}{2}}\Bigr{)},
(4.45) (𝔼[km=0MP(tm)kn=0mphn\displaystyle\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\Bigl{\|}P(t_{m})-k\sum_{n=0}^{m}p^{n}_{h} 2])12C(k14+hk12),\displaystyle\Bigr{\|}^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}\leq C\Bigl{(}k^{\frac{1}{4}}+hk^{-\frac{1}{2}}\Bigr{)},

where CC(DT,𝐮0,𝐟,β,δ)C\equiv C(D_{T},{\bf u}_{0},{\bf f},\beta,\delta) is positive constant independent of kk and hh.

4.2. Finite element methods for the modified Chorin scheme

In this subsection, we first formulate a finite element spatial discretization for Algorithm 2 and then present a complete convergence analysis by deriving error estimates which are stronger than those obtained above for the standard Chorin scheme.

Algorithm 4

Let m0m\geq 0. Set 𝐮~h0=𝒫h0𝐮0\tilde{{\bf u}}_{h}^{0}=\mathcal{P}_{h}^{0}{\bf u}_{0}. For m=0,1,2,m=0,1,2,\cdots do the following steps:

Step 1: For given 𝐮~hmL2(Ω,𝐇h)\tilde{{\bf u}}^{m}_{h}\in L^{2}(\Omega,{\bf H}_{h}), find ξhmL2(Ω,Sh)\xi^{m}_{h}\in L^{2}(\Omega,S_{h}) by solving the following Poisson problem: for \mathbb{P}-a.s.

(4.46) (ξhm,ϕh)=(𝐁(𝐮~hm),ϕh)ϕhSh.\displaystyle\big{(}\nabla\xi^{m}_{h},\nabla\phi_{h}\big{)}=\big{(}{\bf B}(\tilde{{\bf u}}^{m}_{h}),\nabla\phi_{h}\big{)}\qquad\forall\phi_{h}\in S_{h}.

Step 2: Set   𝜼𝐮~hm=𝐁(𝐮~hm)ξhm\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}_{h}}={\bf B}(\tilde{{\bf u}}^{m}_{h})-\nabla\xi^{m}_{h}. For given 𝐮hmL2(Ω,𝐇h){\bf u}^{m}_{h}\in L^{2}(\Omega,{\bf H}_{h}) and 𝐮~hmL2(Ω,𝐇h)\tilde{{\bf u}}^{m}_{h}\in L^{2}(\Omega,{\bf H}_{h}), find 𝐮~hm+1L2(Ω,𝐇h)\tilde{{\bf u}}^{m+1}_{h}\in L^{2}(\Omega,{\bf H}_{h}) by solving the following problem: for \mathbb{P}-a.s.

(4.47) (𝐮~hm+1,𝐯h)\displaystyle\big{(}\tilde{{\bf u}}^{m+1}_{h},{\bf v}_{h}\big{)} +k(𝐮~hm+1,𝐯h)\displaystyle+k\big{(}\nabla\tilde{{\bf u}}^{m+1}_{h},\nabla{\bf v}_{h}\big{)}
=(𝐮hm,𝐯h)+k(𝐟m+1,𝐯h)+(𝜼𝐮~hmΔWm+1,𝐯h)𝐯h𝐇h.\displaystyle=\big{(}{\bf u}^{m}_{h},{\bf v}_{h}\big{)}+k\big{(}{\bf f}^{m+1},{\bf v}_{h}\big{)}+\big{(}\boldsymbol{\eta}^{m}_{\tilde{{\bf u}}_{h}}\Delta W_{m+1},{\bf v}_{h}\big{)}\qquad\forall\,{\bf v}_{h}\in{\bf H}_{h}.

Step 3: Find rhm+1L2(Ω,Lh)r^{m+1}_{h}\in L^{2}(\Omega,L_{h}) by solving the following Poisson problem: for \mathbb{P}-a.s.

(4.48) (rhm+1,ϕh)\displaystyle\big{(}\nabla r^{m+1}_{h},\nabla\phi_{h}\big{)} =1k(𝐮~hm+1,ϕh)ϕhLh.\displaystyle=\frac{1}{k}\big{(}\tilde{{\bf u}}^{m+1}_{h},\,\nabla\phi_{h}\big{)}\qquad\forall\phi_{h}\in L_{h}.

Step 4: Define 𝐮hm+1L2(Ω,𝐇h){\bf u}^{m+1}_{h}\in L^{2}(\Omega,{\bf H}_{h}) by

(4.49) 𝐮hm+1=𝐮~hm+1krhm+1.\displaystyle{\bf u}^{m+1}_{h}=\tilde{{\bf u}}^{m+1}_{h}-k\nabla r^{m+1}_{h}.

Step 5: Define phm+1L2(Ω,Lh)p^{m+1}_{h}\in L^{2}(\Omega,L_{h}) by

(4.50) phm+1=rhm+1+1kξhmΔWm+1.\displaystyle p^{m+1}_{h}=r^{m+1}_{h}+\frac{1}{k}\xi^{m}_{h}\Delta W_{m+1}.

Since each step involves a coercive problem, hence, Algorithm 4 is well defined. The next theorem establishes a convergence rate for the finite element approximation of the velocity field. Since the proof follows the same lines as those in the proof of Theorem 4.1, we omit it to save space.

Theorem 4.4.

Let {𝐮~m}m=0M\{\tilde{{\bf u}}^{m}\}_{m=0}^{M} and {𝐮~hm}m=0M\{\tilde{{\bf u}}_{h}^{m}\}_{m=0}^{M} be generated respectively by Algorithm 2 and 4. Then, there exists a constant CC(DT,𝐮0,𝐟)>0C\equiv C(D_{T},{\bf u}_{0},{\bf f})>0 such that

(4.51) max1mM(𝔼[𝐮~m𝐮~hm2])12\displaystyle\max_{1\leq m\leq M}\biggl{(}\mathbb{E}\bigl{[}\|\tilde{{\bf u}}^{m}-\tilde{{\bf u}}_{h}^{m}\|^{2}\bigl{]}\biggr{)}^{\frac{1}{2}} +(𝔼[km=1M(𝐮~m𝐮~hm)2])12\displaystyle+\biggl{(}\mathbb{E}\biggl{[}k\sum_{m=1}^{M}\|\nabla(\tilde{{\bf u}}^{m}-\tilde{{\bf u}}_{h}^{m})\|^{2}\biggr{]}\biggr{)}^{\frac{1}{2}}
C(k+h+h2k12).\displaystyle\leq C\biggl{(}\sqrt{k}+h+h^{2}k^{-\frac{1}{2}}\biggr{)}.

In the next theorem, we establish an error estimate for the pressure approximation of the modified Chorin finite element method given by Algorithm 4.

Theorem 4.5.

Let {rm}m=1M\{r^{m}\}_{m=1}^{M} and {rhm}m=1M\{r_{h}^{m}\}_{m=1}^{M} be generated respectively by Algorithm 2 and 4. Then, there exists a constant CC(DT,𝐮0,𝐟,δ)>0C\equiv C(D_{T},{\bf u}_{0},{\bf f},\delta)>0 such that

(4.52) (𝔼[km=1M(rmrhm)2])12C(k+h+h2k12).\displaystyle\biggl{(}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{m=1}^{M}(r^{m}-r^{m}_{h})\biggr{\|}^{2}\biggr{]}\biggr{)}^{\frac{1}{2}}\leq C\biggl{(}\sqrt{k}+h+h^{2}k^{-\frac{1}{2}}\biggr{)}.
Proof.

Let 𝜺𝐮~m=𝐮~m𝐮~hm\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m}=\tilde{{\bf u}}^{m}-\tilde{{\bf u}}_{h}^{m} and εrm=rmrhm\varepsilon_{r}^{m}=r^{m}-r^{m}_{h}. It is easy to check that (𝜺𝐮~m,εrm)(\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m},\varepsilon_{r}^{m}) satisfies the following error equation:

(4.53) (𝜺𝐮~m+1𝜺𝐮~m,𝐯h)+k(𝜺𝐮~m+1,\displaystyle\bigl{(}\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1}-\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m},{\bf v}_{h}\bigr{)}+k\bigl{(}\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1}, 𝐯h)+k(εrm,𝐯h)\displaystyle\,\nabla{\bf v}_{h}\bigr{)}+k\bigl{(}\nabla\varepsilon_{r}^{m},{\bf v}_{h}\bigr{)}
=((𝜼𝐮~m𝜼𝐮~hm)ΔWm+1,𝐯h)𝐯h𝐇h.\displaystyle=\bigl{(}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1},{\bf v}_{h}\bigr{)}\qquad\forall\,{\bf v}_{h}\in{\bf H}_{h}.

Applying the summation operator m=0(0M1)\sum_{m=0}^{\ell}\,(0\leq\ell\leq M-1) to (4.53) yields

(4.54) (𝜺𝐮~+1𝜺𝐮~0,𝐯h)\displaystyle\bigl{(}\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{\ell+1}-\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{0},{\bf v}_{h}\bigr{)} +(km=0𝜺𝐮~m+1,𝐯h)+(km=0εrm,𝐯h)\displaystyle+\biggl{(}k\sum_{m=0}^{\ell}\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1},\nabla{\bf v}_{h}\biggr{)}+\bigl{(}k\sum_{m=0}^{\ell}\nabla\varepsilon_{r}^{m},{\bf v}_{h}\bigr{)}
=(m=0(𝜼𝐮~m𝜼𝐮~hm)ΔWm+1,𝐯h)𝐯h𝐇h.\displaystyle=\biggl{(}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1},{\bf v}_{h}\biggr{)}\qquad\forall\,{\bf v}_{h}\in{\bf H}_{h}.

Thus,

(4.55) (km=0εrm,div 𝐯h)\displaystyle\biggl{(}k\sum_{m=0}^{\ell}\varepsilon_{r}^{m},\mbox{\rm div\,}{\bf v}_{h}\biggr{)} =(𝜺𝐮~+1𝜺𝐮~0,𝐯h)+(km=0𝜺𝐮~m+1,𝐯h)\displaystyle=\bigl{(}\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{\ell+1}-\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{0},{\bf v}_{h}\bigr{)}+\biggl{(}k\sum_{m=0}^{\ell}\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1},\nabla{\bf v}_{h}\biggr{)}
(m=0(𝜼𝐮~m𝜼𝐮~hm)ΔWm+1,𝐯h)=:𝙸+𝙸𝙸+𝙸𝙸𝙸.\displaystyle\quad-\biggl{(}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1},{\bf v}_{h}\biggr{)}=:{\tt I}+{\tt II}+{\tt III}.

Using Poincaré and Schwarz inequalities, the three terms on the right-hand side of (4.55) can be bounded as follows:

(4.56) |𝙸|\displaystyle|{\tt I}| C(𝜺𝐮~+1+𝜺𝐮~0)𝐯h,\displaystyle\leq C\bigl{(}\|\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{\ell+1}\|+\|\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{0}\|\bigr{)}\|\nabla{\bf v}_{h}\|,
(4.57) |𝙸𝙸|\displaystyle|{\tt II}| km=0𝜺𝐮~m+1𝐯h,\displaystyle\leq\biggl{\|}k\sum_{m=0}^{\ell}\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1}\biggr{\|}\|\nabla{\bf v}_{h}\|,
(4.58) |𝙸𝙸𝙸|\displaystyle|{\tt III}| Cm=0(𝜼𝐮~m𝜼𝐮~hm)ΔWm+1𝐯h.\displaystyle\leq C\bigg{\|}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1}\bigg{\|}\bigl{\|}\nabla{\bf v}_{h}\bigr{\|}.

Using (4.35) and (4.56)–(4.58) in (4.55), we obtain

(4.59) 1δ2km=0εrm2h2km=0εrm2\displaystyle\frac{1}{\delta^{2}}\biggl{\|}k\sum_{m=0}^{\ell}\varepsilon_{r}^{m}\biggr{\|}^{2}-h^{2}\biggl{\|}k\sum_{m=0}^{\ell}\nabla\varepsilon_{r}^{m}\biggr{\|}^{2} C𝜺𝐮~+12+km=0𝜺𝐮~m+12\displaystyle\leq C\bigl{\|}\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{\ell+1}\bigr{\|}^{2}+k\sum_{m=0}^{\ell}\bigl{\|}\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1}\bigr{\|}^{2}
+Cm=0(𝜼𝐮~m𝜼𝐮~hm)ΔWm+12.\displaystyle\quad+C\biggl{\|}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1}\biggr{\|}^{2}.

Using itô isometry, the last term above can be bounded as

(4.60) 𝔼[m=0(𝜼𝐮~m𝜼𝐮~hm)Δm+1W2]\displaystyle\mathbb{E}\biggl{[}\biggl{\|}\sum_{m=0}^{\ell}(\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m})\Delta_{m+1}W\biggr{\|}^{2}\biggr{]} =𝔼[km=0𝜼𝐮~m𝜼𝐮~hm2]\displaystyle=\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\bigl{\|}\boldsymbol{\eta}_{\tilde{{\bf u}}}^{m}-\boldsymbol{\eta}_{\tilde{{\bf u}}_{h}}^{m}\bigr{\|}^{2}\biggr{]}
C𝔼[km=0𝜺𝐮~m2]+Ch2.\displaystyle\leq C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m}\|^{2}\biggr{]}+Ch^{2}.

Substituting (4.60) into (4.59) yields

(4.61) 1δ2𝔼[km=0εrm2]\displaystyle\frac{1}{\delta^{2}}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{m=0}^{\ell}\varepsilon_{r}^{m}\biggl{\|}^{2}\biggr{]} C𝔼[𝜺𝐮~+12]+𝔼[km=0𝜺𝐮~m+12]\displaystyle\leq C\mathbb{E}\bigl{[}\|\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{\ell+1}\|^{2}\bigr{]}+\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m+1}\|^{2}\biggr{]}
+C𝔼[km=0𝜺𝐮~m2]+h2𝔼[km=0εrm2]+Ch2.\displaystyle\quad+C\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\boldsymbol{\varepsilon}_{\tilde{{\bf u}}}^{m}\|^{2}\biggr{]}+h^{2}\mathbb{E}\biggl{[}k\sum_{m=0}^{\ell}\|\nabla\varepsilon_{r}^{m}\|^{2}\biggr{]}+Ch^{2}.

Finally, the desired estimate (4.52) follows from Theorem 4.4 and Step 3 of Algorithm 2 and 4. The proof is complete. ∎

Corollary 4.1.

Let {pm}m=1M\{p^{m}\}_{m=1}^{M} and {phm}m=1M\{p_{h}^{m}\}_{m=1}^{M} be generated respectively by Algorithm 1 and 2. Then, there exists a positive constant CC(DT,𝐮0,𝐟,δ)C\equiv C(D_{T},{\bf u}_{0},{\bf f},\delta) such that

(4.62) (𝔼[km=1M(pmphm)2])12C(k+h+h2k12).\displaystyle\biggl{(}\mathbb{E}\biggl{[}\biggl{\|}k\sum_{m=1}^{M}(p^{m}-p^{m}_{h})\biggr{\|}^{2}\biggr{]}\biggr{)}^{\frac{1}{2}}\leq C\biggl{(}\sqrt{k}+h+h^{2}k^{-\frac{1}{2}}\biggr{)}.
Proof.

Since the proof follows the same lines as those of the proof for Corollary 3.4, we only highlight the main steps. By definition of {pm}\{p^{m}\} and {phm}\{p^{m}_{h}\}, we have

(4.63) km=1M(pmphm)\displaystyle\biggl{\|}k\sum_{m=1}^{M}(p^{m}-p^{m}_{h})\biggr{\|} km=1M(rmrhm)+m=1M(ξ𝐮~mξ𝐮~hm)ΔWm+1\displaystyle\leq\biggl{\|}k\sum_{m=1}^{M}(r^{m}-r^{m}_{h})\biggr{\|}+\biggl{\|}\sum_{m=1}^{M}(\xi^{m}_{\tilde{{\bf u}}}-\xi_{\tilde{{\bf u}}_{h}}^{m})\Delta W_{m+1}\biggr{\|}
=:𝙸+𝙸𝙸.\displaystyle=:{\tt I}+{\tt II}.

Term 𝙸{\tt I} can be easily bounded by using Theorem 4.5. To bound 𝙸𝙸{\tt II}, by Itô isometry we get

𝔼[𝙸𝙸]2=𝔼[km=1Mξmξhm2].\displaystyle\mathbb{E}[{\tt II}]^{2}=\mathbb{E}\biggl{[}k\sum_{m=1}^{M}\|\xi^{m}-\xi_{h}^{m}\|^{2}\biggr{]}.

In addition, by (3.54) and (4.46), we get

(4.64) (ξ𝐮~mξ𝐮~hm)2\displaystyle\bigl{\|}\nabla(\xi_{\tilde{{\bf u}}}^{m}-\xi_{\tilde{{\bf u}}_{h}}^{m})\bigr{\|}^{2} C𝐁(𝐮~m)𝐁(𝐮~hm)2+Ch2ξ𝐮~mH2\displaystyle\leq C\bigl{\|}{\bf B}(\tilde{{\bf u}}^{m})-{\bf B}(\tilde{{\bf u}}^{m}_{h})\bigr{\|}^{2}+Ch^{2}\bigl{\|}\xi_{\tilde{{\bf u}}}^{m}\bigr{\|}_{H^{2}}
C𝐁(𝐮~m)𝐁(𝐮~hm)2+Ch2div (𝐁(𝐮~m))2\displaystyle\leq C\bigl{\|}{\bf B}(\tilde{{\bf u}}^{m})-{\bf B}(\tilde{{\bf u}}^{m}_{h})\bigr{\|}^{2}+Ch^{2}\bigl{\|}\mbox{\rm div\,}({\bf B}(\tilde{{\bf u}}^{m}))\bigr{\|}^{2}
C𝐁(𝐮~m)𝐁(𝐮~hm)2+Ch2(𝐁(𝐮~m))2\displaystyle\leq C\bigl{\|}{\bf B}(\tilde{{\bf u}}^{m})-{\bf B}(\tilde{{\bf u}}^{m}_{h})\bigr{\|}^{2}+Ch^{2}\bigl{\|}\nabla({\bf B}(\tilde{{\bf u}}^{m}))\bigr{\|}^{2}
C𝐁(𝐮~m)𝐁(𝐮~hm)2+Ch2𝐮~m2.\displaystyle\leq C\bigl{\|}{\bf B}(\tilde{{\bf u}}^{m})-{\bf B}(\tilde{{\bf u}}^{m}_{h})\bigr{\|}^{2}+Ch^{2}\bigl{\|}\nabla\tilde{{\bf u}}^{m}\bigr{\|}^{2}.

Finally, by Poincaré inequality, Lipschitz continuity of 𝐁{\bf B}, Theorem 4.4 and (4.64), we obtain

𝔼[𝙸𝙸]2=𝔼[km=1Mξmξhm2]C(k+h2+h4k).\displaystyle\mathbb{E}[{\tt II}]^{2}=\mathbb{E}\biggl{[}k\sum_{m=1}^{M}\|\xi^{m}-\xi_{h}^{m}\|^{2}\biggr{]}\leq C\biggl{(}k+h^{2}+\frac{h^{4}}{k}\biggr{)}.

The proof is complete. ∎

We conclude this section by stating the following global error estimate theorem for Algorithm 4, which is another main result of this paper.

Theorem 4.6.

Let (𝐮,P)({\bf u},P) be the solution of (1.1) and {(𝐮~hm,rhm,phm)}m=1M\{({\tilde{{\bf u}}}^{m}_{h},r_{h}^{m},p_{h}^{m})\}_{m=1}^{M} be the solution of Algorithm 4. Then, there exists a constant CC(DT,𝐮0,𝐟,β,δ)>0C\equiv C(D_{T},{\bf u}_{0},{\bf f},\beta,\delta)>0 such that

max1mM(𝔼[𝐮(tm)𝐮~hm2])12\displaystyle\max_{1\leq m\leq M}\Bigl{(}\mathbb{E}\bigl{[}\bigl{\|}{\bf u}(t_{m})-\tilde{{\bf u}}^{m}_{h}\bigr{\|}^{2}\,\bigl{]}\Bigr{)}^{\frac{1}{2}} +(𝔼[km=1M(𝐮(tn)𝐮~hn)2])12\displaystyle+\Bigl{(}\mathbb{E}\Bigl{[}k\sum_{m=1}^{M}\bigl{\|}\nabla({\bf u}(t_{n})-\tilde{{\bf u}}^{n}_{h})\bigr{\|}^{2}\,\Bigr{]}\Bigr{)}^{\frac{1}{2}}
C(k+h+h2k12),\displaystyle\quad\leq C\biggl{(}\sqrt{k}+h+h^{2}k^{-\frac{1}{2}}\biggr{)}\,,
(𝔼[R(tm)kn=1mrhn2])12\displaystyle\Bigg{(}\mathbb{E}\bigg{[}\bigg{\|}R(t_{m})-k\sum^{m}_{n=1}r^{n}_{h}\bigg{\|}^{2}\,\bigg{]}\Bigg{)}^{\frac{1}{2}} +(𝔼[P(tm)kn=1mphn2])12\displaystyle+\Bigg{(}\mathbb{E}\bigg{[}\bigg{\|}P(t_{m})-k\sum^{m}_{n=1}p^{n}_{h}\bigg{\|}^{2}\,\bigg{]}\Bigg{)}^{\frac{1}{2}}
C(k+h+h2k12).\displaystyle\quad\leq C\,\biggl{(}\sqrt{k}+h+h^{2}k^{-\frac{1}{2}}\biggr{)}\,.
Remark 4.1.

The above error estimates are of the same nature as those obtained in [12] for the standard Euler-Maruyama mixed finite element method. On the other hand, the error estimates obtained in [13] for the Helmholtz enhanced Euler-Maruyama mixed finite element method do not have the growth term h2k12h^{2}k^{-\frac{1}{2}}. Hence, in the case of general multiplication noise, the Helmholtz enhanced Euler-Maruyama mixed finite element method performs better than the Helmholtz enhanced Chorin finite element method in terms of rates of convergence.

5. Numerical experiments

In this section, we present two 2D numerical tests to guage the performance of the proposed numerical methods/algorithms. The first test is to verify the convergent rates proved in Theorem 4.3 for Algorithm 3 while the second test is to validate the convergent rates proved in Theorem 4.6. In both tests the computational domain is chosen as D=(0,1)×(0,1)D=(0,1)\times(0,1), the P1P1P_{1}-P_{1} equal-order pair of finite element spaces are used for spatial discretization, the constant source function 𝐟=(1,1)\mathbf{f}=(1,1) is applied, the terminal time is T=1T=1, the fine time and space mesh sizes k0=14096k_{0}=\frac{1}{4096} and h=150h=\frac{1}{50} are used to compute the numerical true solution, and the number of realizations is set as Np=500N_{p}=500 for the first test and Np=800N_{p}=800 for the second one. Moreover, to evaluate errors in strong norms, we use the following numerical integration formulas: for any 0mM0\leq m\leq M

𝓔𝐮m\displaystyle\boldsymbol{\mathcal{E}}_{{\bf u}}^{m} :=(𝔼[𝐮(tm)𝐮hm(k)2])12(1Np=1Np𝐮hm(k0,ω)𝐮hm(k,ω)2)12,\displaystyle:=\biggl{(}\mathbb{E}\Bigl{[}\|{\bf u}(t_{m})-{\bf u}_{h}^{m}(k)\|^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}\approx\biggl{(}\frac{1}{N_{p}}\sum_{\ell=1}^{N_{p}}\bigl{\|}{\bf u}_{h}^{m}(k_{0},\omega_{\ell})-{\bf u}_{h}^{m}(k,\omega_{\ell})\bigr{\|}^{2}\biggr{)}^{\frac{1}{2}}\,,
𝓔𝐮,avM\displaystyle\boldsymbol{\mathcal{E}}_{{\bf u},av}^{M} :=(𝔼[km=0M𝐮(tm)𝐮hm(k)2])12\displaystyle:=\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\|{\bf u}(t_{m})-{\bf u}_{h}^{m}(k)\|^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}
(1Np=1Np(km=0M𝐮hm(k0,ω)𝐮hm(k,ω)2))12,\displaystyle\approx\biggl{(}\frac{1}{N_{p}}\sum_{\ell=1}^{N_{p}}\Bigl{(}k\sum_{m=0}^{M}\bigl{\|}{\bf u}_{h}^{m}(k_{0},\omega_{\ell})-{\bf u}_{h}^{m}(k,\omega_{\ell})\bigr{\|}^{2}\Bigr{)}\biggr{)}^{\frac{1}{2}}\,,
p,avM\displaystyle{\mathcal{E}}_{p,av}^{M} :=(𝔼[km=0MP(tm)kn=0mphn(k)2])12\displaystyle:=\biggl{(}\mathbb{E}\Bigl{[}k\sum_{m=0}^{M}\Big{\|}P(t_{m})-k\sum_{n=0}^{m}p_{h}^{n}(k)\Big{\|}^{2}\Bigr{]}\biggr{)}^{\frac{1}{2}}
(1Np=1Np(km=0Mk0n=1tmk0phn(k0,ω)kn=1tmkphn(k,ω)2))12.\displaystyle\,\,\approx\biggl{(}\frac{1}{N_{p}}\sum_{\ell=1}^{N_{p}}\Bigl{(}k\sum_{m=0}^{M}\Bigl{\|}k_{0}\sum_{n=1}^{\frac{t_{m}}{k_{0}}}p_{h}^{n}(k_{0},\omega_{\ell})-k\sum_{n=1}^{\frac{t_{m}}{k}}p_{h}^{n}(k,\omega_{\ell})\Bigr{\|}^{2}\Bigr{)}\biggr{)}^{\frac{1}{2}}\,.

Test 1. In this test, the nonlinear multiplicative noise function 𝐁{\bf B} is chosen as 𝐁(𝐮)=10((u12+1)12,(u22+1)12){\bf B}({\bf u})=10\bigl{(}(u_{1}^{2}+1)^{\frac{1}{2}},(u_{2}^{2}+1)^{\frac{1}{2}}\bigr{)} and the initial value 𝐮0=(0,0){\bf u}_{0}=(0,0)^{\top}. Moreover, we choose J\mathbb{R}^{J}-valued Wiener process WW with increments satisfying

(5.1) ΔWm+1=WJ(tm+1,𝐱)WJ(tm,𝐱)=k0j,=0Jλj,ej,(𝐱)βj,m,\displaystyle\Delta W_{m+1}=W^{J}(t_{m+1},\mathbf{x})-W^{J}(t_{m},\mathbf{x})=k_{0}\sum_{j,\ell=0}^{J}\sqrt{\lambda_{j,\ell}}\,e_{j,\ell}(\mathbf{x})\beta_{j,\ell}^{m},

where 𝐱=(x1,x2)D\mathbf{x}=(x_{1},x_{2})\in D, βj,mN(0,1)\beta_{j,\ell}^{m}\sim N(0,1) and {ej,(𝐱)}j,\{e_{j,\ell}(\mathbf{x})\}_{j,\ell} are orthonormal functions defined by ej,(𝐱)=gj,gj,1e_{j,\ell}(\mathbf{x})=g_{j,\ell}\|g_{j,\ell}\|^{-1} with

(5.2) gj,(x1,x2)=sin(jπx1)sin(πx2)\displaystyle g_{j,\ell}(x_{1},x_{2})=\sin(j\pi x_{1})\sin(\ell\pi x_{2})

and λj,=1(j+)2gj,\lambda_{j,\ell}=\frac{1}{(j+\ell)^{2}}\|g_{j,\ell}\|. In this test, we set J=2J=2, ν=1\nu=1.

Figure 1 displays the convergence rates of the time discretization produced by Algorithm 3 (and Algorithm 1) using different time step size kk. The left figure shows the convergence rate O(k14)O(k^{\frac{1}{4}}) in the 𝓔𝐮,avM\boldsymbol{\mathcal{E}}_{{\bf u},av}^{M}-norm for the velocity approximation, while the right graph shows the same convergence rate in the 𝓔p,avM\boldsymbol{\mathcal{E}}^{M}_{p,av}-norm for the pressure approximation, both match the theoretical rates proved in our theoretical error estimates.

Refer to caption Refer to caption

Figure 1. Convergence rates of the time discretization for the velocity (left) and pressure (right) approximations by Algorithm 3 in the 𝓔𝐮,avM\boldsymbol{\mathcal{E}}^{M}_{{\bf u},av} norm and 𝓔p,avM\boldsymbol{\mathcal{E}}^{M}_{p,av} norm respectively.

Next, we want to verify that the dependence of the error bounds on the factor k12k^{-\frac{1}{2}} is valid. To the end, we fix h=120h=\frac{1}{20} and use again different time step size kk. The numerical results in Figure 2 shows that the errors for both the velocity and pressure approximations increase as the time step size decreases, which proves that the error bounds are indeed proportional to some negative power of kk.

Refer to caption Refer to caption

Figure 2. Errors the velocity approximation (left) in 𝓔𝐮,avM\boldsymbol{\mathcal{E}}^{M}_{{\bf u},av} norm and the pressure approximation (right) in p,avM\mathcal{E}_{p,av}^{M} norm by Algorithm 3.

To verify the sharpness of the error bounds on the factor k12k^{-\frac{1}{2}}, we implement Algorithm 3 using different pairs (k,h)(k,h), which satisfy the relation hkh\approx k, and display the numerical results in Figure 3. We observe 14\frac{1}{4} order convergence rate for both the velocity and pressure approximations as predicted in Theorem 4.3.

Refer to caption Refer to caption

Figure 3. Convergence rates in the 𝓔𝐮,avM\boldsymbol{\mathcal{E}}^{M}_{{\bf u},av} norm for the velocity (left) approximation and the p,avM\mathcal{E}^{M}_{p,av} norm for the pressure (right) approximation by Algorithm 3 under the mesh condition hkh\approx k.

Test 2. We use the same test problem as in Test 1 to validate the theoretical error estimates for our modified Chorin scheme given by Algorithm 4. However, the nonlinear multiplicative noise functions is chosen as 𝐁(𝐮)=((u12+1)12,(u22+1)12){\bf B}({\bf u})=\bigl{(}(u_{1}^{2}+1)^{\frac{1}{2}},(u_{2}^{2}+1)^{\frac{1}{2}}\bigr{)}. It should be noted that a similar numerical experiment was done in [8]. However, only the velocity approximation was analyzed and tested, no convergent rate for the pressure approximation was proved or tested in [8]. Here we want to emphasize the optimal convergence rate for the pressure approximation in the time-averaged norm.

Figure 4 displays the 12\frac{1}{2} order convergence rate in time for both the velocity and pressure approximations by Algorithm 4 as predicted by Theorem 4.6. We note that the velocity error is measured in the strong norm and the pressure error is measured in a time-averaged norm.

Refer to caption
Refer to caption
Figure 4. Convergence rates of the time discretization for the velocity in strong norm (left) and pressure in time-averaged norm (right) by Algorithm 4.

Similar to Test 1, we want to test whether the dependence of the error bounds on the factor k12k^{-\frac{1}{2}} is valid and sharp. To the end, we use the same strategy as we did in Test 1, namely, we fix mesh size h=120h=\frac{1}{20} and decrease time step size kk. As expected, we observe that the errors blow up as shown in Figure 5.

Refer to caption
Refer to caption
Figure 5. Errors for the velocity approximation in strong norm (left) and pressure approximation in time-averaged norm (right) by Algorithm 4.

Finally, Figure 6 shows the 12\frac{1}{2} order convergence rate for both the velocity and pressure approximations by Algorithm 4 when the time step size kk and the space mesh size hh satisfy the balancing condition hkh\approx\sqrt{k}, which verifies the sharpness of the dependence of the error bounds on on the factor k12k^{-\frac{1}{2}} as predicted by Theorem 4.6.

Refer to caption
Refer to caption
Figure 6. Convergence rates for the velocity approximation in strong norm (left) and pressure approximation in time-averaged norm (right) under the mesh condition hkh\approx\sqrt{k}.

References

  • [1] A. Bensoussan, Stochastic Navier-Stokes equations, Acta Appl. Math., 38, 267–304 (1995).
  • [2] A. Bensoussan and R. Temam, Equations stochastiques du type Navier-Stokes, J. Funct. Anal. 13, 195–222 (1973).
  • [3] H. Bessaih, Z. Brzeźniak, and A. Millet, Splitting up method for the 2D stochastic Navier-Stokes equations, Stoch. PDE: Anal. Comp., 2:433–470 (2014).
  • [4] H. Bessaih and A. Millet, On strong L2L^{2} convergence of time numerical schemes for the stochastic 2D Navier-Stokes equations, arXiv:1801.03548 [math.PR], to appear in IMA J. Numer. Anal.
  • [5] D. Breit and A. Dodgson, Convergence rates for the numerical approximation of the 2D stochastic Navier-Stokes equations, arXiv:1906.11778v2 [math.NA] (2020).
  • [6] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods (Third Edition), Springer-Verlag, New York, 2008.
  • [7] Z. Brzeźniak, E. Carelli, and A. Prohl, Finite element based discretizations of the incompressible Navier-Stokes equations with multiplicative random forcing, IMA J. Numer. Anal., 33, 771–824 (2013).
  • [8] E. Carelli, E. Hausenblas, and A. Prohl, Time-splitting methods to solve the stochastic incompressible Stokes equations, SIAM J. Numer. Anal., 50(6):2917–2939 (2012).
  • [9] E. Carelli and A. Prohl, Rates of convergence for discretizations of the stochastic incompressible Navier-Stokes equations, SIAM J. Numer. Anal., 50(5):2467–2496 (2012).
  • [10] P.-L. Chow, Stochastic Partial Differential Equations, Chapman and Hall/CRC, 2007.
  • [11] G. Da Prato and J. Zabczyk, Stochastic Equations in Infinite Dimensions, Cambridge University Press, Cambridge, UK, 1992.
  • [12] X. Feng and H. Qiu, Analysis of fully discrete mixed finite element methods for time-dependent stochastic Stokes equations with multiplicative noise, J. Scient. Comp., https://doi.org/10.1007/s10915-021-01546-4 (2021).
  • [13] X. Feng, A. Prohl, and L. Vo, Optimally convergent mixed finite element methods for the stochastic stokes equations, IMA J. Numer. Anal., 41(3):2280–2310 (2021).
  • [14] V. Girault and P.-A. Raviart, Finite Element Methods for Navier-Stokes Equations, Springer, New York, 1986.
  • [15] J. L.  Guermond, P. Minev, and J.  Shen, An overview of projection methods for incompressible flows, Comput. Methods Appl. Mech. Engrg., 195:6011–-6045 (2006).
  • [16] T. J .R. Hughes, L. P. Franca, and M. Balestra, A new finite element formulation for computational fluid mechanics: V. Circumventing the Babuska-Brezzi condition: A stable Petrov-Galerkin formulation of the Stokes problem accomodating equal order interpolation, Comp. Meth  Appl. Mech. Eng., 59:85–99 (1986).
  • [17] M. Hairer, and J. C. Mattingly, Ergodicity of the 2D Navier-Stokes equations with degenerate stochastic forcing, Ann. of Math., 164:993–1032 (2006).
  • [18] J.A. Langa, J. Real, and J. Simon, Existence and Regularity of the Pressure for the Stochastic Navier-Stokes Equations, Appl. Math. Optim., 48:195–210 (2003).
  • [19] A. Prohl, Projection and Quasi-Compressibility Methods for Solving the Incompressible Navier–Stokes Equations, Advances in Numerical Mathematics, B.G. Teubner, Stuttgart, 1997.
  • [20] R. Rannacher, On Chorin’s projection method for the incompressible Navier-Stokes Equations.The Navier-Stokes Equations II — Theory and Numerical Methods. Lecture Notes in Mathematics, vol 1530, Springer, Berlin, Heidelberg,, 1992.
  • [21] R. Temam, Navier-Stokes Equations. Theory and Numerical Analysis, 2nd ed., AMS Chelsea Publishing, Providence, RI, 2001.