More on convergence of
Chorin’s projection method for incompressible Navier-Stokes equations
Abstract
Kuroki and Soga [Numer. Math. 2020] proved that a version of Chorin’s fully discrete projection method, originally introduced by A. J. Chorin [Math. Comp. 1969], is unconditionally solvable and convergent within an arbitrary fixed time interval to a Leray-Hopf weak solution of the incompressible Navier-Stokes equations on a bounded domain with an arbitrary external force. This paper is a continuation of Kuroki-Soga’s work. We show time-global solvability and convergence of our scheme; -error estimates for the scheme in the class of smooth exact solutions; application of the scheme to the problem with a time-periodic external force to investigate time-periodic (Leray-Hopf weak) solutions, long-time behaviors, error estimates, etc.
Keywords: fully discrete projection method; incompressible Navier-Stokes equations; Leray-Hopf weak solution; time-periodic solution; error estimate
AMS subject classifications: 35Q30; 35D30; 65M06; 65M15
1 Introduction
We consider the incompressible Navier-Stokes equations on a bounded domain of
(1.5) | |||
where is the velocity, is the pressure, is a given external force, is an arbitrary positive number, is initial data and , , etc., stand for the partial (weak) derivatives of . Let and be arbitrarily taken as
Here, is the family of -functions : with a compact support; ; is the closure of with respect to the norm ; (resp. ) is the closure of with respect to the norm (resp. ).
A function belonging to is called a time global Leray-Hopf weak solution of (1.5), if satisfies (1.6) for each fixed .
This paper is a continuation of the work [5]. In [5], Kuroki-Soga proposed a version of Chorin’s fully discrete projection method applied to (1.5) and proved its convergence within an arbitrarily fixed time interval to a Leray-Hopf weak solution (up to a subsequence) by means of a new compactness argument (the standard Aubin-Lions-Simon approach fails). It seems that Chorin’s fully discrete projection method is no longer very popular in modern computational fluid dynamics because of its less accuracy, i.e., discretization of into a uniform mesh and the Dirichlet boundary condition cause a less accurate result. However, we believe that Chorin’s fully discrete projection method can be one of strong mathematical tools to analyze the Navier-Stokes equations including complicated issues such as free boundary problems, long time behaviors, time-periodic solutions, bifurcations, etc. Unlike Galerkin type methods, the projection method solves the equations more directly, which could be an advantage for better understandings. Motivated by such an opinion, we further develop mathematical analysis of Chorin’s fully discrete projection method beyond the convergence to a Leray-Hopf weak solution of the initial boundary value problem.
In Section 2, we first formulate a version of Chorin’s fully discrete projection method and recall the results in [5]. Note that [5] deals with the one-sided difference and the discrete Helmholtz-Hodge decomposition formulated by the zero Dirichlet boundary condition for both of the divergence-free part and potential part. Here, we deal with the central difference and the discrete Helmholtz-Hodge decomposition formulated by the zero Dirichlet boundary condition for the divergence-free part and the zero mean condition for the potential part. This modification in the discrete Helmholtz-Hodge decomposition is particularly important to obtain error estimates, since the exact pressure term does not necessarily satisfy the zero Dirichlet boundary condition. The new result of Section 2 is the time-global solvability of our discrete problem with a fixed discretization parameter, under the assumption that the -norm of the external force within is uniformly bounded for any . This result yields a sequence of step functions that is convergent locally in time to a time-global Leray-Hopf weak solution.
In Section 3, we demonstrate an error estimate for our scheme in the -class. In [1], Chorin showed an -error estimate of in the -class for problems with the periodic boundary conditions, where and are the mesh size for the space variables and time variable, respectively. In the case of the zero Dirichlet boundary condition, the issue is more complicated due to the gap between the exact boundary and the boundary of the grid space. Semi-discrete projection methods, i.e., discrete in time with the mesh size and continuous in space, are free from this complication and one can do a lot also in the class of strong solutions. In fact, Rannacher [8] gave an error estimate of for the Dirichlet problem. Since Chorin took the diffusive scaling condition in his fully discrete setting, the two results by Chorin and Rannacher seem to be “consistent”. We also refer to Shen [10] and the references therein for further investigation on semi-discrete projection methods. Although a fully discrete projection method applied to the Dirichlet problem is said to be less accurate, to the best of the authors’ knowledge, there is no rigorous error analysis. We will show an -error estimate of for a discrete solution and exact -solution under the scaling condition . Note that Chorin [1] and Temam [11] proved convergence of their schemes with the standard diffusive scaling condition, while Kuroki-Soga [5] gave scale-free results; The diffusive scaling does not yield such an error estimate in our formulation. We will see that our error bound and scaling condition arise from the discrete Helmholtz-Hodge decomposition, not from the discrete Navier-Stokes equations. Although the error estimates of does not sound very sharp, the proof provides a new idea to estimate a remainder term on the boundary arising from “summation by parts” in the discrete problem, which is reminiscent of the construction of the trace operator. This idea would provide further applications in analysis of finite difference methods.
In Section 4, we apply our scheme to investigation of the problem with a time-periodic external force. In this problem, one of the main issues is to find a time-periodic solution with the same time-period as that of the external force. After the first attempt by Serrin [9], many results have been obtained. We refer to Kyed [4] for a nice review of the literature on time-periodic solutions to the Navier-Stokes equations. To the best of the authors’ knowledge, there is no mathematical investigation of time-periodic solutions in terms of the fully discrete projection method. We find a discrete time-periodic solution as a fixed point of the time- map of the discrete Navier-Stokes equations and prove convergence to a time-periodic Leray-Hopf weak solution. We also investigate long-time behaviors of discrete solutions, assuming that there exists a “small” discrete solution in the -sense. We obtain exponential contraction of any other discrete solutions. Since the rate of contraction is independent of the size of the discretization parameters, we see that similar exponential contraction holds for exact Leray-Hopf weak solutions, where we do not assume any regularity except for the -bound of a solution. This idea would provide further applications in analysis of stability of a time-periodic solution, its bifurcation, etc. of the exact problem through the discrete problem. Furthermore, we prove that any discrete solution falls into the -neighborhood of an exact time-periodic solution, provided the exact solution is of the -class and “small”. These results can be seen as a version of the results by Serrin [9], Miyakawa-Teramoto [6] and Teramoto [12]. We refer also to Jauslin-Kreiss-Moser [2] and Nishida-Soga [7] for similar investigations on time-periodic entropy solutions of forced Burgers equations through finite difference methods. Finally, we point out Kagei-Nishida-Teramoto [3] for analysis on stability of stationary solutions of the incompressible Navier-Stokes equations via the corresponding artificial compressible system.
In Section 5, we briefly state results corresponding to Section 3 and 4 in the case of the periodic boundary conditions, where the diffusive scaling and the central difference play an essential role. Since the -estimates can be sharpened to be , we obtain an -estimate of through the inequality used by Chorin [1].
2 Construction of Leray-Hopf weak solution
We investigate a version of the scheme studied in [5] with the central difference, as well as the discrete Helmholtz-Hodge decomposition with the zero Dirichlet boundary condition for the divergence-free part and the zero mean condition for the potential part.
2.1 Calculus on grid
Let be the mesh size for the space variables and consider the grid
Let be the standard basis of . For , the boundary of is defined as . Let be a bounded connected open subset of with a Lipschitz boundary . For and , set
Define the discrete derivatives of a function as follows: For each ,
where these operations work under the condition that is extended to be outside , i.e., (resp. ) if (resp. ). For , set , . Define the discrete gradient of a function and the discrete divergence of a function as
We often use discrete versions of integration by parts, “summation by parts”, where we need careful treatments of reminder terms on the boundary. For this purpose, we introduce the following notation (see with Figure 1):
e.g., if there is a sequence of points of on a line parallel to as shown in Figure 1, we have , , , , , , , .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/f156f601-1c70-44ad-99e2-4c316e07f7a5/Figure-1.png)
Figure 1.
Lemma 2.1.
For functions and , we have
Proof.
We may carry out by the summation along sequences of grid points of on each line parallel to ( or ). By shifting to in the summation, we obtain
Since and are supposed to be outside , we obtain the assertion. ∎
We give two Poincaré type inequalities for functions on the grid.
Lemma 2.2 (Poincaré type inequality I).
For each function with , we have
where is a constant depending only on the diameter of .
Proof.
This is proved in [5]. ∎
The second one is for functions without the zero boundary condition, where the mean value of each function is involved in the inequality. This is essentially applied to the discrete Helmholtz-Hodge decomposition below, where we need to avoid the presence of values of the derivatives on (see Theorem 2.8). Since we formulate the inequality with the central difference, we must look at the -translation invariance in the grid : is divided into , each of which is invariant under the -translation, i.e., are the sets of grid points with index even, even, even, even, even, odd, even, odd, even, odd, even, even, even, odd, odd, odd, odd, even, odd, even, odd, odd, odd, odd, respectively. Introduce the following notation:
The reason why we introduce is to avoid the presence of values of the derivatives on in the Poincaré type inequality. For each , there exists such that ; Then, and there exists . Hence, we have .
(2.1) |
Similarly, for each , we have such that and such that . Hence, (2.1) implies
(2.2) |
We always assume that is small enough so that is connected, i.e., for any , we have such that for all and .
Lemma 2.3 (Poincaré type inequality II).
For each function , we have
where is a constant depending only on .
Proof.
See Appendix. ∎
Next two lemmas state Lipschitz interpolation of step functions, which is used to show convergence of a step function in .
Lemma 2.4.
For a function with and the step function derived from as
there exists a Lipschitz continuous function with a compact support such that
where is a constant independent of and .
Proof.
This is proved in [5]. ∎
Lemma 2.5.
For a function and the step function derived from as
there exists a Lipschitz continuous function
such that
where is a constant independent of and .
Proof.
See Appendix. ∎
2.2 Discrete Helmholtz-Hodge decomposition
Here is the discrete Helmholtz-Hodge decomposition.
Theorem 2.6.
For each function , there exist unique functions and such that
where does not necessarily need to banish on .
Proof.
We modify the proof of Theorem 2.2 of [5]. First, we note that any function with satisfies for each ,
(2.4) |
due to cancelation. We label each point of and as
Let be unknown functions to be determined. Introduce and as
where has zero in front of . Then, the equations on , on with the zero mean constraint of give a -system of linear equations, which is denoted by with a -matrix . Due to (2.4), we find eight trivial from . Therefore, can be deduced to be with a -matrix and . Note that is independent of , and that if .
Our assertion holds, if is invertible. To prove invertibility of , we show that if and only if . There is at least one satisfying . Then, we obtain at least one pair satisfying
(2.5) |
By Lemma 2.1, we obtain
Therefore, on and on . The latter equality implies that is constant on for each , and hence (2.5) implies . Thus, is invertible.
Suppose that there are two pairs and which satisfy the assertion. Then, we see that , yields the unique trivial solution of . Therefore, we conclude that and . ∎
Definition 2.7.
Define the discrete Helmholtz-Hodge decomposition operator for each function as
Theorem 2.8.
The following estimates hold for the decomposition :
where is the constant from the discrete Poincaré type inequality II. Furthermore, if on , we have
(2.6) |
Proof.
Note that the values of on are out of control in the decomposition, which requires the discrete Poincaré type inequalities not to contain those values; We will discuss an inequality corresponding to (2.6) without the condition on in Section 3.
2.3 Discrete Navier-Stokes equations
Let be the time-discretization parameter and let be such that . For initial data and the external force , introduce , , as
We define functions , and , in the following manner ( is already defined above):
(2.8) | |||||
(2.10) | |||||
(2.11) |
where (2.8)-(2.11) are recurrence equations in the implicit form and called the discrete Navier-Stokes equations.
For functions or , we define the discrete -inner product and norm as
The next theorem states unconditional solvability of the implicit equations (2.8)-(2.10).
Theorem 2.9.
Proof.
Although our proof is essentially the same as the proof of Theorem 3.1 of [5], we demonstrate some calculation. We label the elements of as . Introduce as
Then, (2.3)-(2.10) are re-written as the linear equations with a -matrix depending on .
We prove that the matrix is invertible if satisfies in . It is enough to check that has the unique solution . Let be a solution of . Then, there exists at least one function with such that
Then, we have
The above two summations are denoted by (i), (ii), respectively. With the zero boundary condition of , we see that
Shifting to in the last summation, we obtain
Similarly, we see that
Hence, the discrete divergence free constraint of implies
Thus, we conclude that and . ∎
Theorem 2.10.
Proof.
We may follow the proof of Theorem 4.1 of [5]. ∎
Theorem 2.10 implies convergence of the discrete solution to a Leray-Hopf weak solution (up to a subsequence). Set . For the solution of (2.8)-(2.11), define the step functions , as
(2.19) | |||||
(2.22) | |||||
(2.25) |
where and the notation is seen in Lemma 2.4. In the rest of our argument, the statement “there exists a sequence …” means “there exists a sequence with as …”.
Theorem 2.11.
There exists a sequence and a function for which the following weak convergence holds:
(2.26) | |||
(2.27) | |||
(2.28) |
Theorem 2.12.
Take under the condition , where is any constant. Then, the sequence , which satisfies (2.27), converges strongly to in as .
Proof.
We may follow the proofs of Lemma 6.1 and Theorem 6.2 of [5], where we slightly change in to be (note that we use the central difference for the discrete divergence). ∎
Theorem 2.13.
The limit function of and derived under the condition with is a Leray-Hopf weak solution of (1.5).
Proof.
We may follow the proof of Theorem 7.1 of [5]. ∎
2.4 Time-global solvability
We sharpen Theorem 2.10 by taking the dissipative effect of into account and prove time-global solvability of the discrete Navier-Stokes equations under the assumption that there exists a constant for which the external force satisfies
A typical example of such is time-periodic one, which will be discussed in Section 4.
Take with . Define the set of initial data as
and the constant as
where is the constant from the discrete Poincaré type inequality I. Note that depends only on the diameter of .
Theorem 2.14.
Proof.
It follows from the equalities for (i), (ii) in the proof of Theorem 2.9 and the discrete Poincaré type inequality I that the inner product of (2.3) and yields
Hence, we have
Therefore, we obtain
We see that
Thus, if satisfies with , we have . We, then, repeat the same estimate to obtain , and so on. ∎
Theorem 2.15.
Proof.
Standard Cantor’s diagonal argument yields the assertion. In fact, for each , Theorem 2.13 implies that there exists a sequence with as such that , converge to a Leray-Hopf weak solution defined in . Then, we may subtract a subsequence from such that , converge to a Leray-Hopf weak solution defined in . Repeating this process for and taking the sequence , we obtain our assertion. ∎
3 Error estimate in -class
We give an error estimate for our projection method, supposing that the external force is smooth and that the limit of Theorem 2.13 belongs to the -class with the pressure . Note that a Leray-Hopf weak solution is smooth within a certain time interval, provided initial data and are smooth enough. The argument below itself does not require smoothness of , and we proceed with the Lipschitz regularity of (we do not discuss if there is a special situation where a Lipschitz domain yields a -solution).
Difficulty here is that is not contained in ; Hence, the exact solution does not satisfy the zero boundary condition on ; The calculus on applied to leaves reminder terms coming from . Careful estimates of such reminder terms are necessary. For this purpose, we assume that our Lipschitz domain satisfies the following property:
Condition A. There exist a constant and a family of open subset of planes in such that
-
•
Each is contained in and has a normal vector such that for each ,
where is seen as a hight function between and ,
-
•
for all and ,
-
•
.
Note that Condition A is fulfilled if is smooth; being rectangular with orthogonal to , or fails to satisfy Condition A (the edge is left over in the last condition), but we may directly deal with such an through the reasoning in Subsection 3.1.
Our goal is to prove the next Theorem.
Theorem 3.1.
Suppose that Condition A holds. Suppose also that (1.5) with a smooth external force possesses the solution such that and . Then, the solution to the discrete problem under the scaling condition , with fixed constants satisfies
where is a constant independent of and .
Our strategy is the following: Let , satisfy (1.5) in the sense of classical solutions. For , define
For each , set
It follows from the Taylor expansion that
Hence, the exact Navier-Stokes equations imply that
Let be the solution of (2.8)-(2.11) with given by . Set
Then, we have for ,
(3.1) | |||
In order to have a recurrence inequality of the norm of with respect to from (3.1), we need the estimate
The term must be treated as small increment of error within even though it does not contain ; namely, we have to take out of this term with an appropriate scaling condition.
We remark that in the rest of this section, the discrete differential operators operate on and in (3.1) without the -extension outside , while operates on with the -extension outside .
We will demonstrate -estimates of (3.1) and (3), where we must take care of remainder terms of “summation by parts” coming from and . For this purpose, we prepare several lemmas below. Note that, if and have more regularity, in can be , which is essential in the problem with the periodic boundary conditions for a shaper error estimate (see Section 5 and [1]). In the Dirichlet problem, however, is not necessary in , because (3) gives lower order error.
3.1 Estimates on boundary
We show that is of at best in general. Then, we must take out of with a scaling condition in accordance with the other remainder terms. We will see that the appropriate scaling is , which implies . One can say that convergence rate of the fully discrete projection method with the Dirichlet boundary is governed by the estimate of in (3.1) and (3). Our argument requires several estimates on/near the boundary of , which is reminiscent of the construction of the trace operator on .
It is useful to observe that we have a constant such that
Let be the indicator function supported by . We sometimes use calculation like
(3.3) | |||
(3.4) |
We prepare several lemmas.
Lemma 3.2.
There exist constants depending only on for which each function satisfies the estimate
Proof.
There exists with the zero mean such that on . We will apply the Poincaré type inequality II to (that is why is involved). With discrete divergence free constraint of , we have
(3.5) | |||
Set
With the Poincaré type inequality II, we have
(3.6) | |||
We estimate the terms in of , where (3.3), (3.4) are not available because is not estimated in and (3.4) leaves . Take a smooth function such that
where is the one in Condition A. Now we use Condition A. Since () are open subsets of planes, we still have the statements of Condition A with instead of for some , where stands for the set . Define
For , we have the following estimate: Fix such that ; For each , we have
where . Note that for all because of instead of . Hence, we see that
For , we have
We have the same estimate for . In this way, we obtain
(3.8) |
with some constant . (3.5), (3.6) and (3.8) conclude the proof. ∎
Lemma 3.3.
For each function such that , and , there exists a constant independent of for which we have
where with the -extension of outside .
Proof.
Lemma 3.4.
There exists a constant depending only on such that for any function and we have
Proof.
We deal with the case of . Due to the same reasoning and notation as those of the proof of Lemma 3.2, we have for each ,
Hence we see that
The other cases are proved in the same way. ∎
3.2 Proof of Theorem 3.1
For each , we have with Lemma 3.3,
(3.9) |
We take the inner product of (3.1) and over : Observe that
With the discrete divergence free constraint of and for , we obtain
We estimate each term: By Lemma 3.4 and (3.9), we have
The other terms are also estimated in this way. Hence, we obtain
Since outside , we have with (3.9),
(3.11) | |||
where we took out of the inner product.
Observe that
Since outside , we have
where we note that , and
(3.12) | |||
By the discrete Poincaré type inequality I, where we note that and the discrete Poincaré type inequality I does not work for itself, we have
(3.13) | |||
Hence, we obtain
(3.14) | |||
Observe that
With (3.1), we have
We estimate each term: Since on and on , we have
Since and on , we have with (3.12);
Therefore, we obtain
(3.15) | |||
Finally, we have
(3.16) |
The estimates (3.2), (3.11), (3.14), (3.15) and (3.16) together with the scaling yield
(3.17) | |||
where to are some positive constants independent of and .
4 Problem with time-periodic external force
We investigate the Navier-Stokes equations with a time-periodic external force. Suppose that the external force is time-periodic with the period , i.e.,
Take with . Then, we may introduce the time- map
of the discrete Navier-Stokes equations. We find a fixed point of , which yields a time-periodic solution of the discrete Navier-Stokes equations, i.e., a solution of (2.8)-(2.11) such that
Then, we show that tend to a time-periodic Leray-Hopf weak solution with the period as , where is called a time-periodic Leray-Hopf weak solution of
(4.4) |
with the period , if
We also discuss long-time behaviors of the (discrete) Navier-Stokes equations, as well as an error estimate, assuming that there exists a smooth time-periodic solution of (4.4).
4.1 Time-global solvability and time-periodic solution
Define the set of initial data of the discrete Navier-Stokes equations as
and the constant as
where is the constant in the discrete Poincaré type inequality I. The next theorem is an immediate consequence of Theorem 2.14.
Theorem 4.1.
For any and fixed , the time- map maps to itself. For each , the discrete Navier-Stokes equations is solvable for all and the solution satisfies for all .
Theorem 4.2.
For any and fixed , the time- map possesses at least one fixed point, which yields a time-periodic solution of the discrete Navier-Stokes equations.
Proof.
Since is finite, we find a one to one onto mapping , i.e., we tag the points of as and define for each . Since , the Euclidian norm of is also bounded by for each fixed . It is clear that is a convex set, and hence, is a bounded convex subset of .
Since is obtained though finitely many basic arithmetic operations, is continuous with respect to . In fact, let and be solutions of the discrete Navier-Stokes equations with and set , ; It is enough to check that (), as in the sense of ; Since due to the property of , we have as ; Suppose that as for some ; The discrete Navier-Stokes equations implies (4.6) in the proof of Theorem 4.4 below; Taking the inner product of (4.6) and together with the calculation for (i) and (iii) in the proof, we obtain
where we note that is fixed and is bounded in the process of and ; By induction, we have our assertion.
Therefore, the map is continuous with respect to the Euclidian norm of . Brouwer’s fixed point theorem guarantees existence of a fixed point. ∎
4.2 Time-periodic Leray-Hopf weak solution
Let , be the solution of the discrete Navier-Stokes equations with initial data equal to a fixed point of . Define the step functions with and in the same way as (2.19) to (2.25). The argument on weak and strong convergence in [5] proves that weakly converge to some function and strongly convergence to in as (up to a subsequence). Let be periodically extended in time with the period , i.e., and for a.e. . Since are time-periodic with the period , it is clear that, for any fixed , weakly converge to and strongly convergence to in as (the same subsequence as the above). Furthermore, is a time-periodic Leray-Hopf weak solution of (4.4) with the period . By taking in Theorem 4.2, we find a time-periodic solution which tends to as . To sum up, we have
Theorem 4.3.
A time-periodic solution of the discrete Navier-Stokes equations tends to a time-periodic Leray-Hopf weak solution of (4.4) as (up to a subsequence). There exists a family of time-periodic (discrete and Leray-Hopf weak) solutions which tends to in the -norm as .
4.3 Stability of small solution
We prove that a “small” solution is exponentially stable. Suppose that there exists a small solution in the sense of , i.e., a solution , of the discrete Navier-Stokes equations (2.8)-(2.11) with , such that
(4.5) |
where is the constant from the Poincaré type inequality I. We remark that it is not clear when one can find such a solution (nevertheless, one could check with a computer). Let be an arbitrary solution of the discrete Navier-Stokes equations with ().
Theorem 4.4.
We have
Proof.
Set , . Observe that for ,
(4.6) | |||
Since is discrete-divergence-free, we have
Summation by part yields
Since is discrete-divergence-free, we obtain with (4.5),
The Poincaré type inequality I implies
Hence, with , we obtain
Suppose that for ,
Then, for any sufficiently small , we have with the Poincaré type inequality I,
which leads to
This is a contradiction. Therefore, we obtain
∎
Corollary 4.5.
Suppose that the discrete Navier-Stokes equations possess a solution satisfying (4.5). Then, a time-periodic solution found in Theorem 4.2 is necessarily unique and bounded by for all and . Furthermore, any other solutions of the discrete Navier-Stokes equations tend to the time-periodic solution as time goes to infinity.
Proof.
Suppose that there exist initial data and a sequence such that the solution of the discrete Navier-Stokes equations solved with satisfies (4.5) for each element of the sequence. Then, Theorem 4.3 and Corollary 4.5 imply that there exists a time-periodic Leray-Hopf weak solution that is bounded by in the -sense. Furthermore, since the decay estimate given in Theorem 4.4 is independent of the size of , we have a stability result on a “small” Leray-Hopf weak solution: Let be a convergent sequence of the step functions derived from the above , where tends to a time-global Leray-Hopf weak solution with initial data as . Let be any time-global Leray-Hopf weak solution with initial data that can be a limit of a sequence derived from the discrete Navier-Stokes equations solved with , where is the same sequence as the above. Note that we do not suppose any regularity of except for the -bound coming from (4.5).
Theorem 4.6.
We have
Proof.
Fix an arbitrary . Take arbitrary small and large . Since (resp., ) strongly converges to (resp., ) in as , we have with sufficiently small and Theorem 4.4,
where are such that , . Taking smaller if necessary, we have
Since and are arbitrary, we obtain our assertion. ∎
4.4 Error estimate for time-periodic solution in -class
Suppose that there exists a time-periodic (period ) solution of the exact Navier-Stokes equations that belongs to the -class and satisfies the smallness condition
(4.7) |
where is the constant from the Poincaré type inequality I. We take which satisfies
(4.8) |
where is some constant specified later. Let be any time-global solution of the discrete Navier-Stokes equations.
Theorem 4.7.
There exist constants and for which we have with each sufficiently small,
Remark. This theorem states that any solution of the discrete Navier-Stokes equations (including time-periodic one!) falls into the -neighborhood of the exact time-periodic solution as time goes to infinity. However, it does not claim that a discrete solution tends to a time-periodic state (we do not assume the existence of a discrete solution satisfying (4.5)) and hence we do not know about the contraction stated in Theorem 4.4.
Proof.
Set , and . Observe that we have for ,
(4.9) | |||
Following the estimate given in Section 3, we have (3.9), (3.2), (3.14), (3.15) and (3.16) (with instead of ) also for (4.9). We estimate by taking out of the inner product after “summation by parts”. For this purpose, observe that
We have
Since , we have
where comes from the values outside . The terms with are estimated as
Therefore, with (4.7) and , we obtain
This estimate together with (3.2), (3.14), (3.15) ( instead of ) and the scaling condition of lead to
(4.10) | |||
where are some positive constants independent of , , , and .
Lemma 4.8.
Suppose that
in (4.8) (the first inequality guarantees that ). Then, we have for each sufficiently small ,
(4.11) |
Proof.
5 Problem with periodic boundary conditions
We briefly summarize results on the problems in , i.e., the problems with the periodic boundary conditions. By taking with , one can formulate the discrete Navier-Stokes equations with the periodic boundary conditions in the same way as Section 2.
Since there is no boundary of , the Poincaré type inequality II is obtained in a simpler way (see [1]) and we are not bothered by the remaining terms from the boundary in the arguments corresponding to Section 3. Hence, we may optimize our error estimates by the central difference and the diffusive scaling . In fact, Lemma 3.3 is improved to be
Furthermore, Theorem 3.1 is improved to be
provided an exact solution belongs to the -class. Then, using the inequality
we obtain the -error estimate
The results in Section 4 are also improved with the diffusive scaling and with a -exact solution, where we need to argue with initial data with a common average over (the average of a solution is conserved both for the Navier-Stokes equations and discrete Navier-Stokes equations). In particular, Theorem 4.7 becomes
Then, we have an -estimate of to be for all sufficiently large . This implies that there exists a solution of the discrete Navier-Stokes equations that satisfies (4.5), provided there exists an exact smooth time-periodic solution that satisfies (4.7). Hence, we obtain
Theorem 5.1.
Suppose that there exists an exact time-periodic solution satisfying (4.7). Then, a time-periodic discrete solution , with the same average as is unique and asymptotically stable within initial data with the same average. The -error between and is for all .
Therefore, one can approximate a time-periodic discrete solution and exact one only by solving an initial value problem of the discrete Navier-Stokes equations for a long time.
Acknowledgement. The second author, Kohei Soga, is supported by JSPS Grant-in-aid for Young Scientists #18K13443.
Appendix Appendix
1. Proof of Lemma 2.3.
It is enough to prove that we have a constant depending only on for which
hold for each .
We first find such a constant with fixed : Suppose that there is no such constant . Then, for each , we have such that
We normalize as
Then, we see that
which implies that is bounded on . Furthermore, since belongs to for any , we have for ,
which implies that is bounded on
Hence, since is fixed, we have a subsequence of whose restriction on converges to some . We have
Since is connected, this is a contradiction.
We next prove that there exists such that for . Suppose that there is no such . Then, for each , we have and such that as and
We normalize as
Then, we see that
Set
Let be the step function defined as
Let be the Lipschitz interpolation of the step function derived from as Lemma 2.5. Then, we have
(1.1) |
where are some constant, which leads to
(1.2) | |||
(1.3) |
We extend to be a function of with the estimates
(1.4) |
where is a constant independent from . This is possible because is bounded and Lipschitz: Let be a family of open balls covering such that each is described as the graph of a Lipschitz map , where is a Cartesian coordinate pointing to in the original space spanned by in such a way that is uniformly away from for , and all () has a common Lipschitz constant (the coordinate depends on ); For all large enough, cover also , where consists of -squares orthogonal to , or ; Each is arbitrarily close to as ; We see that is described as the graph of a Lipschitz map ; We see also that has a common Lipschitz constant for all and ; Then, we may apply the standard extension argument for -functions to obtain (1.4). Since , are bounded in , we have subsequences, still denoted by the same symbol, which weakly converge to some , , respectively. For each , we have
which implies that with . On the other hand, the Rellich-Kondrachov theorem yields a subsequence of , still denoted by the same symbol, such that strongly converges to in as . For each , we have with (1.1),
Hence, we obtain a.e. in , which implies that is constant in . Since due to (1.2), we see that and in . This is a contradiction, since (1.3) implies
∎
2. Proof of Lemma 2.5 .
For each , define the following functions:
Then, we see that
It is clear that can be Lipschitz continuously connected with each other, yielding that satisfies the inequalities. ∎
References
- [1] A. J. Chorin, On the convergence of discrete approximations to the Navier-Stokes equations, Math. Comp. 23 (1969), pp. 341-353.
- [2] H. R. Jauslin, H. O. Kreiss and J. Moser, On the forced Burgers equation with periodic boundary conditions, Proc. Sympos. Pure Math. 65 (1999), pp. 133-153.
- [3] Y. Kagei, T. Nishida and Y. Teramoto, On the spectrum for the artificial compressible system, J. Differential Equations 264 (2018), No. 2, pp. 897-928.
- [4] M. Kyed, Time-Periodic Solutions to the Navier-Stokes Equations, Habilitationsschrift, TU-Darmstadt (2012).
- [5] H. Kuroki and K. Soga, On convergence of Chorin’s projection method to a Leray-Hopf weak solution, Numer. Math. (2020). https://doi.org/10.1007/s00211-020-01144-w
- [6] T. Miyakawa and Y. Teramoto, Existence and periodicity of weak solutions of the Navier-Stokes equations in a time dependent domain, Hiroshima Math. J. 12 (1982), pp. 513-528.
- [7] T. Nishida and K. Soga, Difference approximation to Aubry-Mather sets of the forced Burgers equation, Nonlinearity 25 (2012), 2401-2422.
- [8] R. Rannacher, On Chorin’s projection method for the incompressible Navier-Stokes equations, Lecture Notes in Mathematics 1530, Springer, Berlin-Heidelberg (1992), pp. 167-183.
- [9] J. Serrin, A note on the existence of periodic solutions of the Navier-Stokes equations, Arch. Rational Mech. Anal. 3 (1959), pp. 120-122.
- [10] J. Shen, On error estimates of the projection methods for the Navier-Stokes equations: Second-oder scheme, Math. Comp. 65 (1996), pp. 1039-1065.
- [11] R. Temam, Sur l’approximation de la solution des équations de Navier-Stokes par la méthode des pas fractionnaires. II. (French), Arch. Rational Mech. Anal. 33 (1969), pp. 377-385.
- [12] Y. Teramoto, On the stability of periodic solutions of the Navier-Stokes equation in a noncylindrical domain, Hiroshima Math. J. 13 (1983), pp. 607-625.