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 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 denotes the time mesh size. On the other hand, an optimal convergence rate 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 , where denotes the
time step size, which explains the deteriorating performance of the standard Chorin scheme
when 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.
†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)
(1.1b)
(1.1c)
where represents a period of the periodic domain in , and stand for respectively the velocity field and the pressure of the fluid,
is an operator-valued random field, denotes an -valued
-Wiener process, and is a body force function (see Section 2 for their precise definitions). Here we seek periodic-in-space solutions with period , that is,
and
almost surely
and for any and , where
denotes the canonical basis of .
The system (1.1a) is a stochastic perturbation of the deterministic Stokes system by introducing
a multiplicative noise force term
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 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
is nonlinear in the velocity . 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., 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 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.,
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)
(1.2b)
(1.2c)
where
denotes the normal derivative of and is the time step size.
One of advantages of the above Chorin scheme is that the spatial approximation
spaces for and 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 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 and then rewrite (1.2a) as
(1.3)
where . Our modified Chorin scheme consists of
(1.3), (1.2b)–(1.2c) and the Helmholtz decomposition
. Since is divergence-free,
it turns out that the finite element approximation of the modified Chorin scheme
has better convergence properties. Notice that can be recovered from
via the simple algebraic relation .
The main contributions of this paper are summarized below.
•
We proved the following error estimates in strong norms
for the Chorin- finite element method (see Algorithm 3) for problem (1.1)
with general multiplicative noises:
where are the solution to
problem (1.1) while are the discrete solution of Algorithm 3, see Sections 2 and 4 for their precise definitions.
•
We proposed a modified Chorin- finite element method (see Algorithm 4) and
proved the following error estimates in strong norms for problem (1.1)
with general multiplicative noises:
where is the solution to problem (1.1) and is defined
as the time-average of the pseudo pressure while 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 , which
explains the deteriorating performance of the standard (and modified) Chorin scheme
when and the mesh size 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 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 as well as recall the definition of solutions to (1.1). We then state and prove a Hölder continuity property for the pressure 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 )
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 denote the subspace of whose -valued functions have zero trace on , and denote the standard -inner product, with induced norm . We also denote and as the Lebesgue and Sobolev spaces of the functions that are periodic and have vanishing mean, respectively.
Let be a filtered probability space with the probability measure , the
-algebra and the continuous filtration . For a random variable
defined on ,
denotes the expected value of .
For a vector space with norm , and , we define the Bochner space
, where
.
We also define
We recall from [14] that the (orthogonal) Helmholtz projection
is defined
by for every ,
where is a unique tuple such that
and solves the following Poisson problem
with the homogeneous Neumann boundary condition:
(2.1)
We also define the Stokes operator .
Throughout this paper we assume that is a
Lipschitz continuous mapping and has linear growth, that is,
there exists a constant such that for all
(2.2a)
(2.2b)
Since we shall not explicitly track the dependence of all constants on ,
for ease of the presentation, unless it is stated otherwise, we shall set in the rest of
the paper and assume that . In addition,
we shall use to denote a generic positive constant
which may depend on , the datum functions and , and the domain
but is independent of the mesh parameter and .
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 , let be an -valued Wiener process on it.
Suppose and .
An -adapted stochastic process is called
a variational solution of (1.1) if ,
and satisfies -a.s. for all
(2.3)
We cite the following Hölder continuity estimates for the variational solution whose proofs
can be found in [8, 12].
Lemma 2.1.
Suppose and . Then there exists a constant , such that the variational solution to problem (1.1) satisfies
for
(2.4a)
(2.4b)
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 for (1.1),
its associated pressure is subtle to define. In that regard we quote the following
theorem from [13].
Theorem 2.1.
Let be the variational solution of (1.1). There exists a unique adapted process
such that satisfies -a.s. for all
(2.5a)
(2.5b)
System (2.5) is a mixed formulation for the stochastic Stokes
system (1.1), where the (time-averaged) pressure is defined.
The distributional derivative , which was
shown to belong to ,
was defined as the pressure. Below, we also define another time-averaged “pressure”
(2.6)
using the Helmholtz decomposition , where
such that
(2.7)
Then, it is easy to check that
(2.8)
The process will also be approximated
in our numerical methods.
Next, we establish some stability estimates for the velocity and the pressure in the following lemma.
Lemma 2.2.
Suppose that . Let solve (2.5). Then there exists a constant such that
(2.9)
(2.10)
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)
Then, the desired estimate from immediately after taking supreme over all and using (2.9).
∎
We finish this section by establishing the following Hölder continuity result for , which is used for the error analysis in sections.
Lemma 2.3.
Suppose that , and . Then, there holds
(2.12)
where the constant depends on , and .
Proof.
From its definition of , we get
(2.13)
Therefore, for any , we have
(2.14)
Thus,
(2.15)
The first term I can be controlled by using (2.5a). For II and III, by Schwarz inequality, we have
(2.16)
In addition, using Itô isometry and (2.2b), we obtain
(2.17)
In summary, we have
(2.18)
Finally, the desired estimate follows from the assumptions on and .
∎
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 be a (large) positive integer and be the time step size.
Set for , then forms a
uniform mesh for . The standard Chorin projection scheme is
given as follows (cf. [8, 21, 14]):
Algorithm 1
Let . For , do the following three steps.
Step 1: Given and , find such that -a.s.
(3.1)
Step 2: Find such that -a.s.
(3.2)
Step 3: Define by
(3.3)
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
satisfies the following system:
(3.4a)
(3.4b)
where .
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.
Applying the summation operator to (3.13) for any yields
(3.14)
Finally, by the discrete Gronwall’s inequality we obtain
which implies the desired estimate (3.5a) and (3.5b).
It remains to show (3.5c). To the end, testing (3.4a) by we get
(3.15)
Proceeding with similar arguments as used above yields
(3.16)
Then by the discrete Gronwall inequality we get
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 in this subsection.
First, we state the following error estimate result for the velocity.
Theorem 3.1.
Let be generated by
Algorithm 1, then there exists a positive constant which depends
on and such that
(3.17)
Proof.
Let and .
Obviously, and . In addition, from (2.5a), we have
Next, squaring both sides of (3.44) followed by applying operators and , we obtain
(3.45)
We now bound each term above as follows. Using Theorem 3.1 we get
(3.46)
By the Hölder continuities given in (2.4b) and (2.12), we have
(3.47)
By using the Itô isometry and Theorem 3.1 and (2.4a), we conclude that
(3.48)
Finally, substituting (3.46)–(3.48) into (3.45) yields
(3.49)
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 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 which
depends on and such that
(3.50)
(3.51)
Proof.
(3.50) follows immediately from estimates (2.10) and (3.40) and follows straightforwardly from the discrete Jensen inequality and . It remains to prove . To that end, testing (3.19) by , we obtain
(3.52)
After an rearrangement, applying operators and to (3.52) we obtain
(3.53)
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:
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 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 is
a real-valued Wiener process and independent of the spatial variable. The case of more general can be dealt with similarly.
The modified Chorin scheme is given as follows.
Algorithm 2
Set . For , do
the following five steps.
Step 1: Given , find such that -a.s.
(3.54)
Step 2: Set . Given and , find such that -a.s.
(3.55)
Step 3: Find such that -a.s.
(3.56a)
Step 4: Define as
(3.57)
Step 5: Define the pressure approximation as
(3.58)
Remark 3.3.
It follows from (2.2b) and (3.54) that the Helmholtz projection can be bounded in terms of as follows:
(3.59)
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
defined by Algorithm 2 satisfy
(3.60a)
(3.60b)
where is a positive constant which depends on and .
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)
(3.61b)
where denotes the Helmholtz
projection of .
It was proved in [13] that the following stability and error estimates hold for
the solution of the above Euler-Maruyama scheme.
where is a positive constant which depends on and .
Lemma 3.4.
There hold the the following error estimates for the discrete processes :
(3.63a)
(3.63b)
for . Where is a positive constant which depends on and .
Remark 3.4.
We note that the reason for the estimate (3.62c) to hold is
because the Helmholtz projected noise is divergent-free.
Otherwise, the following weaker estimate can only be proved (cf. [12]):
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 be the solution of Algorithm 2
and be the solution of (1.1).
Then there holds the following estimate:
(3.64)
where is a positive constant which depends on and .
Proof.
Let , . Then, and . Subtracting (3.4) from (3.61) yields
(3.65a)
(3.65b)
(3.65c)
Testing (3.65a) with and integrating by parts,
we obtain
(3.66)
Using the algebraic identity in the first term gives
(3.67)
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 and using (3.65c), we obtain
Substituting (3.71) and (3.72) into (3.67) and rearranging terms, we get
(3.73)
Finally, we bound each term on the right-hand side of (3.73). By Young’s inequality, for we obtain
(3.74a)
(3.74b)
(3.74c)
Rewriting
(3.75)
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)
Then we have
(3.77a)
(3.77b)
Now, substituting (3.74) and (3.77) into (3.73) and taking expectation on both sides, we obtain
(3.78)
Choosing , taking the summation for , and using (3.62c) and the discrete Gronwall’s inequality, we get
(3.79)
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 , 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 which depends on and such that
(3.80a)
(3.80b)
Proof.
The estimate (3.80b) follows immediately from (3.79) and the triangular inequality. To prove (3.80a), testing (3.4a) by , we obtain
(3.81)
Here we have used the periodic boundary condition for to kill the boundary term which arises from integration by parts.
Finally, applying the operator 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 .
Corollary 3.3.
There exists which depends on and
such that
(3.82)
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
and generated by Algorithm 2. First, we state the following lemma.
Lemma 3.5.
Let be generated by Algorithm 2. Then, there exists a constant
depending on and such that for
(3.83)
Proof.
The idea of the proof is to utilize the inf-sup condition (3.41).
Testing (3.57) by any , we obtain
Then, subtracting the above equations yields
(3.84)
Applying the summation operator for to (3.84), we get
(3.85)
where and are the same as defined in the preceding subsection and we have used the fact that .
Finally, by using the inf-sup condition (3.41) and then taking the expectation we get
which and the estimates for and infer the desired estimate (3.83). The proof is complete. ∎
We then are ready to state the following error estimate result for .
Theorem 3.4.
Let be generated by Algorithm 2 and be defined in (2.6).
Then there exists a constant depending on and such hat
for
(3.86)
Proof.
Subtracting (3.61) from (3.4a) and then testing the resulting equation by , we obtain
(3.87)
Applying the summation operator to (3.87) for
yields
(3.88)
Now, we bound each term on the right-hand side of (3.88) as follows.
By Schwarz and Poincaré inequalities, we get
By Itô isometry, the assumptions on and Theorem 3.3, we obtain
Moreover, using the Poisson equation as defined in Step 1 of Algorithm 1, we have
Thus,
Finally, it follows from the inf-sup condition (3.41), Theorem 3.3, (3.88), and all above inequalities that
(3.89)
The proof is completed by applying the triangular inequality, Lemma 3.5 and (3.63b).
∎
Corollary 3.4.
Let be generated by Algorithm 2. Then, there exists a constant
which depends on and such that for
(3.90)
Proof.
The proof is similar to that of Theorem 3.4. Indeed, we have
(3.91)
Clearly, the first term “” can be bounded by using Theorem 3.4. To bound the second term “”,
we first have
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 be a quasi-uniform triangulation of the polygonal () or polyhedral () bounded domain . We introduce
the following two basic Lagrangian finite element spaces:
(4.1)
(4.2)
where denotes the set of polynomials of degree less than or equal to over the element . The finite element spaces to be used to
formulate our finite element methods are defined as follows:
(4.3)
In addition, we introduce spaces
(4.4)
Recall that the -projection is defined by
(4.5)
and the -projection is
defined by
(4.6)
It is well known [6] that and satisfy following estimates:
(4.7)
(4.8)
For the clarity we only
consider -finite element space in this section (i.e., ),
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 and 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 . Set . For do the following steps:
Step 1: Given and , find such that -a.s.
(4.9)
Step 2: Find such that -a.s.
(4.10)
Step 3: Define by
(4.11)
As mentioned in Section 1, eliminating in (4.9) using (4.10), we obtain
(4.12a)
(4.12b)
Next, we state the stability estimates for 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 be generated by Algorithm 3, then there holds
(4.13a)
(4.13b)
where is a positive constant depending only on , and .
The following theorem provides an error estimate in a strong norm for the finite element solution of Algorithm 3.
Theorem 4.1.
Let and 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)
where is a positive constant.
Proof.
The proof is conceptually similar to that of Theorem 3.1.
Setting and . Without loss of the generality, we assume and because they are of high order accuracy, hence are negligible.
First, applying the summation operator to (4.12a), we obtain
(4.15)
Subtracting (3.19) from (4.15) yields the following error equations:
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)
We use the Itô isometry to handle the noise term as follows:
(4.33)
Finally, substituting the above estimates for terms I, II, III, IV, V, VI into (4.25) yields the following inequality for :
(4.34)
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 independent of ,
such that
(4.35)
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 such that
(4.36)
Proof.
We reuse all the notations from the proof of Theorem 4.1. First, from the error equations (4.16) we have
(4.37)
Using the Schwarz inequality on the right-hand side of (4.37) yields
Then, applying operators and on both sides we obtain
(4.40)
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)
Using Theorem 4.1, terms II and III can be bounded as follows:
(4.42)
Finally, using Itô’s isometry and Theorem 4.1 we have
(4.43)
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)
(4.45)
where is positive constant independent of and .
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 . Set . For do the following steps:
Step 1: For given , find by solving the
following Poisson problem: for -a.s.
(4.46)
Step 2: Set . For given and , find by solving the following problem: for -a.s.
(4.47)
Step 3: Find by solving the following Poisson problem: for -a.s.
(4.48)
Step 4: Define by
(4.49)
Step 5: Define by
(4.50)
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 and be generated respectively by Algorithm 2 and 4.
Then, there exists a constant such that
(4.51)
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 and be generated respectively by Algorithm 2 and 4. Then, there exists a constant such that
(4.52)
Proof.
Let and . It is easy to check that satisfies the following error equation:
Finally, by Poincaré inequality, Lipschitz continuity of , Theorem 4.4 and (4.64), we obtain
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 be the solution of (1.1)
and be the solution of Algorithm 4.
Then,
there exists a constant such that
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 .
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 , the equal-order pair of finite element spaces are used for spatial discretization, the constant source function is applied,
the terminal time is , the fine time and space mesh sizes
and are used to compute the numerical true solution, and the number of realizations is set as for the first test and for the second one.
Moreover, to evaluate errors in strong norms, we use the following numerical
integration formulas: for any
Test 1. In this test, the nonlinear multiplicative noise function is chosen as and the initial value . Moreover, we choose -valued Wiener process with increments satisfying
(5.1)
where , and are orthonormal functions defined by with
(5.2)
and . In this test, we set , .
Figure 1 displays the convergence rates of the time discretization produced by
Algorithm 3 (and Algorithm 1) using different time step size . The left figure shows the convergence rate in the -norm for the velocity approximation, while the right graph shows the same convergence rate in the -norm for the pressure approximation, both match the theoretical rates proved in our theoretical error estimates.
Figure 1. Convergence rates of the time discretization for the velocity (left) and pressure (right) approximations by Algorithm 3 in the norm and norm respectively.
Next, we want to verify that the dependence of the error bounds on the factor is valid. To the end, we fix and use again different time step size . 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 .
Figure 2. Errors the velocity approximation (left) in norm and the pressure approximation (right) in norm by Algorithm 3.
To verify the sharpness of the error bounds on the factor , we implement Algorithm 3 using different pairs , which satisfy the relation , and display the numerical results in Figure 3. We observe order convergence rate for both the velocity and pressure approximations as predicted in Theorem 4.3.
Figure 3. Convergence rates in the norm for the velocity (left) approximation and the norm for the pressure (right) approximation by Algorithm 3 under the mesh condition .
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 .
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 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.
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 is valid and sharp. To the end, we use the same strategy as we did in Test 1,
namely, we fix mesh size and decrease time step size . As expected, we observe
that the errors blow up as shown in Figure 5.
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 order convergence rate for both the velocity and pressure approximations by Algorithm 4 when the time step size and the space mesh size satisfy the balancing condition , which verifies the sharpness of the dependence
of the error bounds on on the factor as predicted by Theorem 4.6.
Figure 6. Convergence rates for the velocity approximation in strong norm (left) and pressure approximation in time-averaged norm (right) under the mesh condition .
[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 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).
[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.