Next Article in Journal
A Posteriori Error Estimators for the Quasi-Newtonian Stokes Problem with a General Boundary Condition
Next Article in Special Issue
A New Entropy Stable Finite Difference Scheme for Hyperbolic Systems of Conservation Laws
Previous Article in Journal
Stability and Neimark–Sacker Bifurcation of a Delay Difference Equation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unconditional Superconvergence Error Estimates of Semi-Implicit Low-Order Conforming Mixed Finite Element Method for Time-Dependent Navier–Stokes Equations

School of Mathematics, Zhengzhou University of Aeronautics, Zhengzhou 450046, China
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(8), 1945; https://doi.org/10.3390/math11081945
Submission received: 24 March 2023 / Revised: 15 April 2023 / Accepted: 19 April 2023 / Published: 20 April 2023
(This article belongs to the Special Issue Advanced Computational Methods for Fluid Dynamics and Applications)

Abstract

:
In this paper, the unconditional superconvergence error analysis of the semi-implicit Euler scheme with low-order conforming mixed finite element discretization is investigated for time-dependent Navier–Stokes equations. In terms of the high-accuracy error estimates of the low-order finite element pair on the rectangular mesh and the unconditional boundedness of the numerical solution in L -norm, the superclose error estimates for velocity in H 1 -norm and pressure in L 2 -norm are derived firstly by dealing with the trilinear term carefully and skillfully. Then, the global superconvergence results are obtained with the aid of the interpolation post-processing technique. Finally, some numerical experiments are carried out to support the theoretical findings.

1. Introduction

In this paper, we consider the following two-dimensional time-dependent incompressible Navier–Stokes equations:
u t Δ u + ( u · ) u + p = f , ( x , t ) Ω × ( 0 , T ] ,
· u = 0 , ( x , t ) Ω × ( 0 , T ] ,
u ( x , t ) = 0 , ( x , t ) Ω × ( 0 , T ] ,
u ( x , 0 ) = u 0 ( x ) , x Ω , t = 0 ,
where Ω R 2 is a rectangular domain with boundary Ω , u = ( u 1 , u 2 ) = ( u 1 ( x ) , u 2 ( x ) ) represents the velocity vector, p = p ( x ) denotes the pressure, and f = ( f 1 ( x ) , f 2 ( x ) ) is the prescribed body force, x = ( x , y ) . Moreover, u 0 is the initial velocity and T > 0 is the final time.
Navier–Stokes equations are a classical incompressible fluid model and have been widely applied in the mathematical physics and the computational fluid dynamics fields [1,2,3]. It is an unrealistic thing to find the the exact solutions of the Navier–Stokes equations due to their nonlinear and incompressible properties. Therefore, numerous works have been devoted to the development of efficient numerical approximations for Navier–Stokes Equations (1)–(4), including finite difference methods [4,5,6,7], Galerkin finite element methods [8,9,10,11,12,13,14,15,16,17,18,19,20,21] and other methods [22,23,24,25,26]. In particular, a new second-order accurate finite difference scheme for the incompressible Navier–Stokes equations was discussed in [4] by the primitive variable formulation. Based on the vorticity stream-function formulation and a fast Poisson solver defined on a general domain using the immersed interface method, a fast finite difference method was proposed and studied in [5] for the incompressible Navier–Stokes equations. A new fully discrete finite element nonlinear Galerkin method was established in [8] for the Navier–Stokes equations with two-grid finite element discretization in the spatial direction and Euler explicit scheme with variable time step size in the temporal direction. The boundedness, convergence and stability condition of the presented method were discussed under certain time-step constraints dependent on the coarse grid parameter. A Lagrange–Galerkin mixed finite element approximation was presented for the Navier–Stokes equations in [13], and optimal error estimates were obtained with the mesh restriction τ = O ( h σ ) , where σ > ( d 1 ) / 2 , and τ , h and d denote the time step size, the mesh size and the dimension of the domain, respectively. In [14], in terms of the special properties of a low-order nonconforming mixed finite element pair on the rectangular mesh, the superconvergence error estimates were derived for Navier–Stokes equations with the time-step constraints τ = O ( h 1 + α ) , α > 0 due to the inverse inequality used in the error analysis. In [18], in terms of the error splitting technique developed in [27,28], the unconditional stability and convergence of a typical modified characteristics finite element method were studied for the time-dependent Navier–Stokes equations by introducing an iterated characteristic time-discrete system. Optimal error estimates were obtained under the boundedness of the numerical solution in W 1 , -norm. Subsequently, the unconditionally optimal error estimates were derived in [20] under the boundedness of the numerical solution in L -norm, which is weaker than that in [18].
It should be pointed out that optimal error estimates were obtained in [20] due to roughly handling trilinear terms. Moreover, the superconvergence error estimates were derived in [14] with a certain time-step restriction. To the best of our knowledge, there are few contributions on the unconditionally superclose and superconvergence error estimates for problems (1)–(4). The purpose of this paper is to consider the unconditionally superclose and superconvergence error estimates for incompressible time-dependent Navier–Stokes Equations (1)–(4). It should be pointed out that the difficulties come from the trilinear term (i.e., the convection term) ( u · ) u in the superclose and superconvergence error analysis. Therefore, we should deal with them carefully and skillfully.
In the present work, we focus on a low-order conforming finite element approximation, which is called the bilinear-constant scheme [4,29,30], for problems (1)–(4). The key to our analysis is to employ the special properties (high-accuracy error estimation; see Lemmas 1–2 below) on the rectangular mesh and to treat the convection term ( u · ) u rigorously and skillfully. The superclose error estimates are obtained firstly for the velocity in H 1 -norm and for the pressure in L 2 -norm. Then, in terms of an efficient interpolation post-processing approach, the global superconvergence results are derived for the velocity in the H 1 -norm and for the pressure in the L 2 -norm. In addition, some numerical experiments are presented and tested.
The remainder of this paper is organized as follows. In Section 2, some preliminaries and lemmas are introduced, and the semi-implicit Euler fully discrete finite element approximation of the problem (1)–(4) is presented. In Section 3, the detailed superclose error estimates are studied with the help of the special properties of the bilinear-constant finite element pair combined with skillfully dealing with the trilinear term. Then, the global superconvergence error estimates are established by the interpolation post-processing technique. In Section 4, some numerical results are provided to verify the theoretical findings.

2. Preliminaries

We denote by W m , p ( Ω ) the Sobolev spaces with the norm · m , p and semi-norm | · | m , p defined by
u m , p = | β | k Ω | D β u | p d x d y 1 p , 1 p < , | β | k e s s s u p Ω | D β u | , p = ,
where D β = | β | x β 1 y β 2 for the multi-index β = ( β 1 , β 2 ) , β 1 0 , β 2 0 and | β | = β 1 + β 2 . For p = 2 , we let H m ( Ω ) denote W m , 2 ( Ω ) . We omit the subscript when p = 2 and write · 0 , 2 and | · | 0 , 2 as · 0 and | · | 0 for simplicity. Moreover, we use ( · , · ) to represent the L 2 inner product, i.e.,
( u , v ) = Ω u v d x d y .
For the mathematical setting of the problem (1)–(4), we introduce the following Sobolev spaces V and M [31], i.e.,
V = ( H 0 1 ( Ω ) ) 2 , P = L 0 2 ( Ω ) = { q L 2 ( Ω ) : Ω q d x d y = 0 } .
Moreover, for any Banach space X and I = [ 0 , T ] , let L p ( I ; X ) be the space of all measurable function f : I X with the norm
f L p ( I ; X ) = ( 0 T f X p d t ) 1 p , 1 p , ess sup t I f X , p = .
Let T h = { e } be a uniform rectangular mesh over Ω with mesh size h. For a given element e T h , its four vertices are denoted by a i ( x i , y i ) , i = 1 , 2 , 3 , 4 in the counterclockwise order (see Figure 1 left). For the velocity, we choose V h as the general bilinear finite element space. For the pressure, we assume that the subdivision T h is obtained from T 2 h = { τ } by dividing each element of T 2 h into four small congruent rectangles. Let P h consist of piecewise constant functions with respect to T h such that the local basis functions for P h on a 2 × 2 -patch of τ (see Figure 1 right) are indicated in Figure 2. Then, the finite element space for pressure is defined by P h L 0 2 ( Ω ) . In the following discussion, we always assume that τ = i = 1 4 e i T 2 h with e i T h ( 1 i 4 ) (see Figure 1 right). Thus, the finite element approximation spaces V h and M h for the bilinear-constant scheme are described by ([1,29,30])
V h = { v ( C ( Ω ¯ ) ) 2 : v | e ( Q 11 ( e ) ) 2 , v | Ω = 0 , e T h } , M h = { p L 0 2 ( Ω ) : p | τ = i = 1 3 λ i τ φ i τ , τ T 2 h λ 1 τ = 0 , τ T 2 h } ,
where Q 11 denotes the space of all polynomials of degree 1 with respect to each of the two variables, x and y. It is shown in [1,29,30] that the bilinear-constant scheme satisfies the Babuška–Brezzi condition, i.e.,
sup 0 v h X h ( q h , · v h ) v h 1 β q h 0 , q h M h ,
where β > 0 is a constant, independent of h.
For the velocity, we use the Lagrange nodal interpolation operator I h : ( C ( Ω ¯ ) ) 2 V h as the corresponding interpolation operator. For the pressure, we first introduce the local L 2 -projection J h p of p by
J h p = 1 | e | e p d x 1 d x 2 , e T h ,
and then define the operator J h with respect to e ˜ by
J h p = J h p 1 4 α e ˜ , i = 1 , 4 , J h p + 1 4 α e ˜ , i = 2 , 3 ,
where α e ˜ = p 1 e ˜ p 2 e ˜ p 3 e ˜ + p 4 e ˜ with the notations of p i e ˜ = 1 | e i | e i p d x 1 d x 2 , ( i = 1 , 2 , 3 , 4 ) . We can check that
J h p | e ˜ = 1 4 i = 1 4 p i e ˜ φ 1 e ˜ + ( p 1 e ˜ p 2 e ˜ + p 3 e ˜ p 4 e ˜ ) φ 2 e ˜ + ( p 1 e ˜ + p 2 e ˜ p 3 e ˜ p 4 e ˜ ) φ 3 e ˜ ,
which implies that J h p M h for p L 0 2 ( Ω ) .
Lemma 1
([29]). Suppose that u ( H 3 ( Ω ) ) 2 and p H 2 ( Ω ) , then there hold
( ( u I h u ) , v h ) C h 2 u 3 v h 0 , v h V h ,
( p J h p , · v h ) C h 2 p 2 v h 0 , v h V h ,
( · ( u I h u h ) , q h ) C h 2 u 3 q h 0 , q h P h .
Lemma 2.
Supposing that u H 3 ( K ) on element e, we have
( ( u I h u ) x , v ) e + ( ( u I h u ) y , v ) e C h e 2 u 3 , e v 0 , e , v P h .
Furthermore, there holds for u H 3 ( Ω )
( ( u I h u ) x , v ) + ( ( u I h u ) y , v ) C h 2 u 3 v 0 v P h .
Proof. 
In order to obtain a second-order accuracy estimate, we adopt the high-accuracy integral technique developed in [32]. We introduce the following error function on an element e (see Figure 3):
F ( y ) = 1 2 [ ( y y e ) 2 h y 2 ] ,
where ( x e , y e ) denotes the barycentric coordinate of element e. One can check that F ( y ) = 1 , F ( y ) | l 1 , l 3 = 0 . Note that v is constant on an element e, and we have
( ( u I h u ) x , v ) e = e ( u I h u ) x v d x d y = v e ( u I h u ) x d x d y = v e ( u I h u ) x F ( y ) d x d y = v l 3 l 1 ( u I h u ) x F ( y ) d x e ( u I h u ) x y F ( y ) d x d y = l 3 l 1 ( u I h u ) x y F ( y ) d x e ( u I h u ) x y y F ( y ) d x d y = e F ( y ) u x y y v d x d y C h e 2 u 3 , e v 0 , e ,
where we have that ( u I h u ) ( a i ) = 0 , ( i = 1 , , 4 ) and F ( y ) | l 1 , l 3 are constants in the above estimate.
In the same way, we also have
( ( u I h u ) x , v ) e C h e 2 u 3 , e v 0 , e .
Therefore, by adding (14) and (15), the desired result (11) is derived. Furthermore, summing up (11) with respect to e T h gives the result (12). □
We present the discrete Gronwall’s inequality, which is useful in the following error analysis.
Lemma 3
([33,34]). Let τ, B and a k , b k , c k , γ k , for integers k 0 , be non-negative numbers such that
a n + τ k = 0 n b k τ k = 0 n γ k a k + τ k = 0 n c k + B , for n 0 ,
suppose that τ γ k < 1 , for all k, and set σ k = ( 1 τ γ k ) 1 . Then
a n + τ k = 0 n b k exp τ k = 0 n γ k σ k τ k = 0 n c k + B , for n 0 .
The Young inequality will be frequently used in the following analysis, and we present it here.
a b a 2 2 ϵ + ϵ b 2 2 , f o r ϵ > 0 .
The weak formulation of (1)–(4) is as follows: find ( u , p ) : [ 0 , T ] ( H 0 1 ( Ω ) ) 2 × L 0 2 ( Ω ) , such that
( u t , v ) + ( u , v ) + ( ( u · ) u , v ) ( p , · v ) = ( f , v ) , v ( H 0 1 ( Ω ) ) 2 ,
( · u , q ) = 0 , q L 0 2 ( Ω ) .
Moreover, in order to give the fully discrete scheme, let 0 = t 0 < t 1 < < t N = T be a given uniform partition of the time interval with time step τ = T / N and t n = n τ , n = 0 , 1 , , N . For a smooth function w defined on [ 0 , T ] , denote
w n = w ( t n ) , D τ w n = ( w n w n 1 ) τ .
Then, the semi-implicit backward Euler low-order conforming mixed finite element scheme (bilinear-constant scheme) is as follows: for given ( u h n 1 , p h n 1 ) V h × M h , find ( u h n , p h n ) V h × M h such that
( D τ u h n , v h ) + ( u h n , v h ) + ( ( u h n 1 · ) u h n , v h ) ( p h n , · v h ) = ( f n , v h ) , v h V h ,
( · u h n , q h ) = 0 , q h M h ,
with the initial approximation u h 0 = I h u 0 = I h u ( 0 ) .
The unconditionally optimal error estimate for scheme (18)–(19) is shown in [20] by using the error splitting technique with the low-order conforming mixed finite element method, specifically, by introducing the following time-discrete system [18,20]:
D τ U n Δ U n + ( U n 1 · ) U n + P n = f n ,
· U n = 0 ,
U n = 0 ,
U 0 = u 0 .
Then, the error between the exact solution and the numerical solution comprises two parts, where one is the temporal error and the other is the spatial error. Therefore, the numerical solution in the L -norm can be bounded without any time-step restrictions. Here, we present a lemma to state the τ -independent boundedness of the numerical solution.
Lemma 4
([20]). Suppose that ( u n , p n ) is the solution of (1)–(4) with suitable regularity and ( U n , P n ) and ( u h n , p n ) are the solutions of (20)–(23) and (18)–(19). Then, there exist positive constants τ 0 and h 0 such that when τ τ 0 and h h 0 , we have the τ-independent boundedness of numerical solutions
u h n 0 , K , n = 0 , 1 , , N ,
where K is a constant, which is not dependent on h, τ and n.

3. Superclose and Superconvergence Error Estimates

In this section, we firstly present the superclose error estimates for the velocity and pressure variables. Then, we give the superconvergence result in terms of the superclose error estimates as well as the interpolation technique.
Theorem 1.
Suppose that ( u , p ) and ( u h n , p h n ) are the solutions of (16)–(19) and u L ( ( H 3 ( Ω ) ) 2 ( H 0 1 ( Ω ) ) 2 ) , u t L ( ( H 3 ( Ω ) ) 2 ) , u t t L 2 ( ( L 2 ( Ω ) ) 2 ) and p L ( H 2 ( Ω ) ) . Then, we have the following superclose estimates
( I h u n u h n ) 0 C ( h 2 + τ ) , τ k = 1 n J h p n p h n 0 C ( h 2 + τ ) .
Furthermore, we have the following optimal error estimates:
u n u h n 0 C ( h 2 + τ ) , ( u n u h n ) 0 C ( h + τ ) , τ k = 1 n p n p h n 0 C ( h + τ ) .
Proof. 
For simplicity, we split the errors u u h and p p h as follows:
u n u h n = u n I h u n + I h u n u h n : = ρ n + θ n , p n p h n = p n J h p n + J h p n p h n : = ξ n + η n .
At t = t n , from (16)–(17), we have
( D τ u n , v h ) + ( u n , v h ) + ( ( u n 1 · ) u n , v h ) ( · v h , p n ) = ( f n , v h ) + ( D τ u n u t n , v h ) + ( ( ( u n 1 u n ) · ) u n , v h ) , v h V h ,
( · u n , q h ) = 0 , q h M h .
Subtracting (18)–(19) from (27)–(28) gives the following error equations:
( D τ θ n , v h ) + ( θ n , v h ) ( · v h , η n ) = ( D τ ρ n , v h ) ( ρ n , v h ) + ( · v h , ξ n ) ( ( u n 1 · ) u n ( u h n 1 · ) u h n , v h ) + ( D τ u n u t n , v h ) + ( ( ( u n 1 u n ) · ) u n , v h ) , v h V h ,
( · θ n , q h ) = ( · ρ n , q h ) , q h M h .
Alternatively, we have
( · D τ θ n , q h ) = ( · D τ ρ n , q h ) , q h M h .
Taking v h = D τ θ n in (29) and q h = η n in (31) yields that
D τ θ n 0 2 + 1 2 τ ( θ n 0 2 θ n 1 0 2 + ( θ n θ n 1 ) 0 2 ) = ( D τ ρ n , D τ θ n ) ( ρ n , D τ θ n ) + ( ξ n , · D τ θ n ) ( · D τ ρ n , η n ) ( ( u n 1 · ) u n ( u h n 1 · ) u h n , D τ θ n ) + ( D τ u n u t n , D τ θ n ) + ( ( ( u n 1 u n ) · ) u n , D τ θ n ) : = k = 1 7 E k .
Now, we start to estimate E k , ( k = 1 , , 7 ) term by term. By the Cauchy–Schwarz inequality and interpolation theory, we have
E 1 = ( D τ ρ n , D τ θ n ) C h 2 u t L ( ( H 2 ( Ω ) ) 2 ) D τ θ n 0 C h 2 D τ θ n 0 C h 4 + ϵ D τ θ n 0 2 .
Applying the following equality
( a n , D τ b n ) = D τ ( a n , b n ) ( D τ a n , b n 1 ) ,
we have
E 2 = ( ρ n , D τ θ n ) = D τ ( ρ n , θ n ) + ( D τ ρ n , θ n 1 ) D τ ( ρ n , θ n ) + C h 2 u t L ( H 3 ( Ω ) ) 2 θ n 1 0 D τ ( ρ n , θ n ) + C h 4 + C θ n 1 0 2 ,
where we used (8) of Lemma 1 in the above estimate.
With an application of (34) again, it follows that
E 3 = ( ξ n , · D τ θ n ) = D τ ( ξ n , · θ n ) ( D τ ξ n , · θ n 1 ) D τ ( ξ n , · θ n ) + C h 2 p t L ( H 2 ( Ω ) ) θ n 1 0 D τ ( ξ n , · θ n ) + C h 4 + C θ n 1 0 2 ,
where we used (9) of Lemma 1 in the above estimate.
According to (10) of Lemma 1, E 4 can be bounded by
E 4 = ( · D τ ρ n , η n ) C h 2 u t L ( H 3 ( Ω ) ) 2 η n 0 C h 2 η n 0 .
To bound E 5 , we split ( u n 1 · ) u n ( u h n 1 · ) u h n as follows:
( u n 1 · ) u n ( u h n 1 · ) u h n = ( ( u n 1 u h n 1 ) · ) u n + ( u h n 1 · ) ( u n u h n ) = ( ( u n 1 I h u n 1 ) · ) u n + ( ( I h u n 1 u h n 1 ) · ) u n + ( u h n 1 · ) ( u n I h u n ) + ( u h n 1 · ) ( I h u n u h n ) = ( ( u n 1 I h u n 1 ) · ) u n + ( ( I h u n 1 u h n 1 ) · ) u n + ( ( u h n 1 I h u n 1 ) · ) ( u n I h u n ) + ( ( I h u n 1 u n 1 ) · ) ( u n I h u n ) + ( u n 1 · ) ( u n I h u n ) + ( u h n 1 · ) ( I h u n u h n ) : = k = 1 6 F k .
In terms of the Cauchy–Schwarz inequality and interpolation theory, we have
( F 1 , D τ θ n ) ( F 2 , D τ θ n ) ( ρ n 1 0 + θ n 1 0 ) u n 0 , D τ θ n 0 C ( h 2 + θ n 1 0 ) D τ θ n 0 C h 4 + C θ n 1 0 2 + 2 ϵ D τ θ n 0 2 .
Using inverse inequality and interpolation theory, we have
( F 3 , D τ θ n ) ( F 4 , D τ θ n ) ( θ n 1 0 + ρ n 1 0 ) ρ n 0 D τ θ n 0 , C ( θ n 1 0 + h 2 u n 1 2 ) ( C h u n 2 ) ( C h 1 D τ θ n 0 ) C ( h 2 + θ n 1 0 ) D τ θ n 0 C h 4 + C θ n 1 0 2 + 2 ϵ D τ θ n 0 2 .
With the aid of (35), one can check that
( F 5 , D τ θ n ) = ( ( u n 1 · ) ρ n , D τ θ n ) = D τ ( ( u n 1 · ) ρ n , θ n ) + ( D τ ( ( u n 1 · ) ρ n ) , θ n 1 ) = D τ ( ( u n 1 · ) ρ n , θ n ) + ( ( D τ u n 1 · ) ρ n , θ n 1 ) + ( ( u n 2 · ) D τ ρ n , θ n 1 ) .
In order to obtain a high-accuracy error estimate, i.e., second-order accuracy, we introduce the local L 2 -projection defined as follows:
P e u = 1 | e | e u d x d y ,
where e is an element of the partition of T h and | e | is the measure of e. Then, we have
P e u 0 , e u 0 , e , u P e u 0 , e C h e u 1 , e , u P e u 0 , , e C h e u 1 , , e .
Therefore, we have
( ( D τ u n 1 · ) ρ n , θ n 1 ) = e T h ( ( D τ u n 1 · ) ρ n , θ n 1 ) e = e T h ( ( ( D τ u n 1 P e D τ u n 1 ) · ) ) ρ n , θ n 1 ) e + e T h ( ( P e D τ u n 1 · ) ) ρ n , θ n 1 P e θ n 1 ) e + e T h ( ( P e D τ u n 1 · ) ) ρ n , P e θ n 1 ) e e T h D τ u n 1 P e D τ u n 1 0 , , e ρ n 0 , e θ n 1 0 , e + C e T h ρ n 0 , e θ n 1 P e θ n 1 0 , e + C h 2 e T h u n 3 , e P e θ n 1 0 , e C h 2 e T h u n 2 , e θ n 1 0 , e + C h 2 e T h u n 2 , e θ n 1 1 , e + C h 2 e T h u n 3 , e θ n 1 0 , e C h 2 u n 3 θ n 1 1 C h 2 u n 3 θ n 1 0 C h 4 + C θ n 1 0 2 ,
where we used Lemma 2 in the above estimate.
In the same way, we also have
( ( u n 2 · ) D τ ρ n , θ n 1 ) C h 2 θ n 1 0 C h 4 + C θ n 1 0 2 .
Hence, there holds
( F 5 , D τ θ n ) = D τ ( ( u n 1 · ) ρ n , θ n ) + C h 4 + C θ n 1 0 2 .
By (24) of Lemma 4, the numerical solution u h n is bounded unconditionally in L -norm, and we have
( F 6 , D τ θ n ) u h n 0 , θ n 0 D τ θ n 0 C θ n 0 D τ θ n 0 C θ n 0 2 + ϵ D τ θ n 0 2 .
Based on the above estimates ( F k , D τ θ n ) , ( k = 1 , , 6 ) , E 5 reduces to
E 5 C h 4 + C ( θ n 0 2 + θ n 1 0 2 ) + 5 ϵ D τ θ n 0 2 D τ ( ( u n 1 · ) ρ n , θ n ) .
Moreover, by Taylor expansion, we have
E 6 + E 7 C τ u t t L ( L 2 ( Ω ) ) D τ θ n 0 + C τ u t L ( L 2 ( Ω ) ) u L ( L ( Ω ) ) D τ θ n 0 C τ D τ θ n 0 C τ 2 + ϵ D τ θ n 0 2 .
With the estimates E k , ( k = 1 , , 6 ) , we obtain
D τ θ n 0 2 + 1 2 τ ( θ n 0 2 θ n 1 0 2 ) C ( h 4 + τ 2 ) + C ( θ n 0 2 + θ n 1 0 2 ) + 7 ϵ D τ θ n 0 2 + C h 2 η n 0 D τ ( ρ n , θ n ) + D τ ( ξ n , · θ n ) D τ ( ( u n 1 · ) ρ n , θ n ) .
On the other hand, from (29), it follows that
( · v h , η n ) = ( D τ θ n , v h ) + ( D τ ρ n , v h ) + ( θ n , v h ) + ( ρ n , v h ) ( · v h , ξ n ) + ( ( u n 1 · ) u n ( u h n 1 · ) u h n , v h ) ( D τ u n u t n , v h ) ( ( ( u n 1 u n ) · ) u n , v h ) , v h V h .
Then, one can check that by the Cauchy–Schwarz inequality and interpolation theory,
( D τ θ n , v h ) + ( D τ ρ n , v h ) ( D τ θ n 0 + D τ ρ n 0 ) v h 0 C ( h 2 + D τ θ n 0 ) v h 0
and
( θ n , v h ) θ n 0 v h 0 .
By Lemma 1, we have
( ρ n , v h ) ( · v h , ξ n ) C h 2 v h 0 .
Using a similar estimate process as E 5 , there holds
( ( u n 1 · ) u n ( u h n 1 · ) u h n , v h ) C ( h 2 + θ n 0 + θ n 1 0 ) v h 0 .
As an application of the Taylor expansion, we have
( D τ u n u t n , v h ) ( ( ( u n 1 u n ) · ) u n , v h ) C τ v h 0 .
Therefore, we conclude that
( · v h , η n ) C ( h 2 + τ + θ n 0 + θ n 1 0 + D τ θ n 0 ) v h 0 .
Then, thanks to the discrete LBB condition ( β is a positive constant independent of mesh-size h), we have
β η n 0 sup 0 v h V h ( · v h , η n ) v h 0 C ( h 2 + τ + θ n 0 + θ n 1 0 + D τ θ n 0 ) .
Substituting (58) into (50) and using the Young inequality, (50) reduces to
D τ θ n 0 2 + 1 2 τ ( θ n 0 2 θ n 1 0 2 ) C ( h 4 + τ 2 ) + C ( θ n 0 2 + θ n 1 0 2 ) + 8 ϵ D τ θ n 0 2 D τ ( ρ n , θ n ) + D τ ( ξ n , · θ n ) D τ ( ( u n 1 · ) ρ n , θ n ) .
Choosing ϵ = 1 16 in (59) and summing up the resulting inequality, we have
τ 2 k = 1 n D τ θ k 0 2 + 1 2 θ n 0 2 C ( h 4 + τ 2 ) + C τ k = 1 n θ k 0 2 ( ρ n , θ n ) + ( ξ n , · θ n ) ( ( u n 1 · ) ρ n , θ n ) , C ( h 4 + τ 2 ) + C τ k = 1 n θ k 0 2 + C h 2 u n 3 θ n 0 + C h 2 p n 3 θ n 0 + C h 2 u n 3 θ n 0 C ( h 4 + τ 2 ) + C τ k = 1 n θ k 0 2 + 1 4 θ n 0 2 ,
where we used θ 0 = 0 , Lemma 1 and the same estimate process (44) for ( ( u n 1 · ) ρ n , θ n ) .
Hence, from (60), there holds
τ k = 1 n D τ θ k 0 2 + θ n 0 2 C ( h 4 + τ 2 ) + C τ k = 1 n θ k 0 2 .
Then, an application of the Gronwall lemma (see Lemma 3) yields
τ k = 1 n D τ θ k 0 2 + θ n 0 2 C ( h 4 + τ 2 ) .
Putting (62) into (58) gives that
τ k = 1 n η 0 C ( h 2 + τ ) + τ k = 1 n D τ θ n 0 C ( h 2 + τ ) + C τ k = 1 n D τ θ k 0 2 1 2 C ( h 2 + τ ) .
The desired results are obtained, and the proof is complete. □
With the aid of the superclose error estimate in Theorem 1, we adopt the interpolation post-processing approach to improve the accuracy of the numerical solution ( u h n , p h n ) on the whole domain. Let I 2 h be the piecewise biquadratic nodal interpolation operator for the velocity associated with T 2 h . Moreover, for the pressure, the postprocessing operator J 2 h is defined as
J 2 h p Q 11 ( e ˜ ) , 1 | e i | e i ( J 2 h p p ) d x 1 x 2 = 0 , i = 1 , 2 , 3 , 4 .
The following properties are shown in [29]:
I 2 h I h = I 2 h , I 2 h v 1 C v 1 , v V h , I 2 h u u 1 C h 2 u 3 , J 2 h J h = J 2 h , J 2 h q 0 C q 0 , q P h , J 2 h p p 0 C h 2 p 2 .
Moreover, we also have from [29] that for p H 2 ( Ω ) ,
p J 2 h J h p 0 C h 2 p 2 .
Now, we state the global superonvergence result in the following theorem:
Theorem 2.
Under the condition of Theorem 1, there holds
u n I 2 h u h n 1 + p n J 2 h p h n 0 C ( h 2 + τ ) .
Proof. 
By (64) and (65), we have
u n I 2 h u h n 1 + p n J 2 h p h n 0 u n I 2 h I h u n 1 + I 2 h I h u n I 2 h u h n 1 + p n J 2 h J h p n 0 + J 2 h J h p n J 2 h p h n 0 C h 2 u n 3 + C I h u n u h n 1 + C h 2 p n 2 + C J h p n p h n 0 C ( h 2 + τ ) .
The proof is complete. □

4. Numerical Experiment

In this section, we present some numerical results to confirm the correctness of the theoretical analysis. The software we used is MATLAB 2018a with a 3.20 GHz Intel Core i5-6500 CPU processor and 8 GB memory.
Example 1.
Let Ω = ( 0 , 1 ) 2 , and divide Ω into m × n uniform rectangles. Moreover, the function f and the initial and boundary conditions are chosen corresponding to the exact solution [35]:
u 1 = e t ( x 1 4 2 x 1 3 + x 1 2 ) ( 4 x 2 3 6 x 2 2 + 2 x 2 ) , u 2 = e t ( x 2 4 2 x 2 3 + x 2 2 ) ( 4 x 1 3 6 x 1 2 + 2 x 1 ) , p = 10 e t ( 2 x 1 1 ) ( 2 x 2 1 ) .
We set the final time T = 1.0 in the computation. In order to confirm the error estimates in Theorems 1 and 2, we present the numerical errors in Table 1, Table 2, Table 3, Table 4, Table 5 and Table 6 at t = 0.1 , 0.6 , 1.0 , respectively. Obviously, it can be seen that the numerical results are in agreement with the theoretical analysis, i.e., the errors u n u h n 0 , u n u h n 1 , I h u n u h n 1 and u n I 2 h u h n 1 for the velocity u are of order O ( h 2 ) , O ( h ) , O ( h 2 ) and O ( h 2 ) , respectively. In addition, the errors p n p h n 0 , J h p n p h n 0 and p n J 2 h p h n 0 for the pressure p are of orders O ( h ) , O ( h 2 ) and O ( h 2 ) , respectively. At the same time, we also give the graphics of the exact solutions ( u , p ) and finite element solutions ( u h , p h ) at t = 1.0 on mesh 32 × 32 (see Figure 4, Figure 5, Figure 6 and Figure 7), respectively. It can be seen that the numerical results are also in good agreement with the theoretical analysis.
Example 2.
Let Ω = ( 0 , 1 ) 2 and divide Ω into m × n uniform rectangles. Moreover, the function f and the initial and boundary conditions are chosen corresponding to the exact solution [14]:
u 1 = e 2 t sin 2 ( π x ) sin ( π y ) cos ( π y ) , u 2 = e 2 t sin ( π x ) cos ( π x ) sin 2 ( π y ) , p = e 4 t sin ( π x ) sin ( π y ) .
We set the final time T = 10 in the computation. In order to confirm the error estimates in Theorems 1 and 2, we present the numerical errors in Table 7 and Table 8 at t = 10 . Obviously, it can be seen that numerical results are in agreement with the theoretical analysis.

5. Conclusions

In this paper, a low-order conforming mixed finite element method is investigated for time-dependent Navier–Stokes equations with the semi-implicit Euler scheme. With the help of the error splitting technique and the high-accuracy error estimates of the element pair on the rectangular mesh, the numerical solution in L i n f t y -norm is obtained without any time-step restrictions. Furthermore, the unconditionally superclose error estimates are derived by treating the trilinear term rigorously and skillfully. Moreover, the global superconvergence results are acquired in terms of the interpolation post-processing approach. It should be pointed out there are many interesting topics to study in future works, such as high-order time (e.g., Crank–Nicolson scheme and BDF2 scheme), low-order nonconforming mixed finite element methods, and fast computing for discretized linear system. We will study these topics in the future.

Author Contributions

X.M.; writing—original draft preparation; H.Y.; writing—review and editing, methodology, software. All authors have read and agreed to the published version of the manuscript.

Funding

H.Y. is supported by the National Natural Science Foundation of China (No. 12101568).

Data Availability Statement

The authors will supply the relevant data in response to reasonable requests.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Girault, V.; Raviart, P.A. Finite Element Methods for Navier-Stokes Equations; Theory and Algorithms; Springer: Berlin/Heidelberg, Germany, 1986. [Google Scholar]
  2. Temam, R. Navier-Stokes Equations, Theory and Numerical Analysis; American Mathematical Society: North-Holland, The Netherlands; Amsterdam, The Netherlands; New York, NY, USA; Oxford, UK, 1984. [Google Scholar]
  3. Giraldo, F.X.; Restelli, M.; Laeuter, M. Semi-implicit formulations of the Navier-Stokes equations: Application to nonhydrostatic atmospheric modeling. SIAM J. Sci. Comput. 2010, 32, 3394–3425. [Google Scholar] [CrossRef]
  4. Strikwerda, J.C. Finite difference methods for the Stokes and Navier-Stokes equations. SIAM J. Sci. Stat. Comput. 1984, 50, 56–68. [Google Scholar] [CrossRef]
  5. Wang, C.; Liu, J.G. Analysis of finite difference schemes for unsteady Navier-Stokes equations in vorticity formulation. Numer. Math. 2002, 91, 543–576. [Google Scholar] [CrossRef]
  6. Li, Z.L.; Wang, C. A fast finite difference method for solving Navier-Stokes equations on irregular domains. Commun. Math. Sci. 2003, 1, 181–197. [Google Scholar] [CrossRef]
  7. Hou, T.Y.; Wetton, B.T.R. Convergence of a finite difference scheme for the Navier-Stokes equations using vorticity boundary conditions. SIAM J. Numer. Anal. 1992, 29, 615–639. [Google Scholar] [CrossRef]
  8. He, Y.N.; Li, K.T. Convergence and stability of finite element nonlinear Galerkin method for the Navier-Stokes equations. Numer. Math. 1998, 79, 77–106. [Google Scholar] [CrossRef]
  9. He, Y.N.; Sun, W. Sability and convergence of the Crank-Nicolson/Adams-Bashforth scheme for the time-dependent Navier-Stokes equations. SIAM J. Numer. Anal. 2007, 45, 837–869. [Google Scholar] [CrossRef]
  10. Li, J.; He, Y.N.; Chen, Z.X. A new stabilized finite element method for the transient Navier-Stokes equations. Comput. Methods Appl. Mech. Eng. 2007, 197, 22–35. [Google Scholar] [CrossRef]
  11. He, Y.N.; Li, J. A stabilized finite element method based on local polynomial pressure projection for the stationary Navier-Stokes equation. Appl. Numer. Math. 2008, 58, 1503–1514. [Google Scholar] [CrossRef]
  12. He, Y.N. A fully discrete stablized finite element method for the time-dependent Navier- Stokes problem. IMA J. Numer. Anal. 2003, 23, 665–691. [Google Scholar] [CrossRef]
  13. Suli, E. Convergence and nonlinear stability of the Lagrange-Galerkin method for the Navier-Stokes equations. Numer. Math. 1988, 53, 459–483. [Google Scholar] [CrossRef]
  14. Yang, H.J.; Shi, D.Y.; Liu, Q. Superconvergence analysis of low order nonconforming mixed finite element methods for time-dependent Navier-Stokes equations. J. Comput. Math. 2021, 39, 63–80. [Google Scholar] [CrossRef]
  15. Han, H.D. A finite element approximation of Navier-Stokes equations using nonconforming elements. J. Comput. Math. 1984, 2, 77–88. [Google Scholar]
  16. Xu, X.J. On the accuracy of nonconforming quadrilateral Q1 element approximation for the Navier-Stokes problem. SIAMJ. Numer. Anal. 2000, 38, 17–39. [Google Scholar] [CrossRef]
  17. Shi, D.Y.; Ren, J.C.; Hao, X.B. A new second order nonconforming mixed finite element scheme for the stationary Stokes and Navier-Stokes equations. Appl. Math. Comput. 2009, 207, 462–477. [Google Scholar] [CrossRef]
  18. Si, Z.Y.; Wang, J.L.; Sun, W.W. Unconditional stability and error estimates of modified characteristics FEMs for the Navier-Stokes equations. Numer. Math. 2016, 134, 139–161. [Google Scholar] [CrossRef]
  19. Li, M.H.; Shi, D.Y.; Li, Z.Z.; Chen, H.R. Two-level mixed finite element methods for the Navier-Stokes equations with damping. J. Math. Anal. Appl. 2019, 470, 292–307. [Google Scholar] [CrossRef]
  20. Yang, H.J.; Shi, D.Y. Unconditionally optimal error estimates of the bilinear-constant scheme for time-dependent Navier-Stokes equations. J. Comput. Math. 2022, 40, 127–146. [Google Scholar]
  21. Li, M.H.; Xiao, L.C.; Li, Z.Z. The supercloseness property of the Stoke projection for the transient Navier-Stokes equations and global superconvergence analysis. J. Appl. Math. Comput. 2023, 69, 2041–2057. [Google Scholar] [CrossRef]
  22. Shen, J. On error estimates of projection methods for Navier-Stokes equations: First-order schemes. SIAM J. Numer. Anal. 1992, 29, 57–77. [Google Scholar] [CrossRef]
  23. Shen, J. On error estimates of the projection, methods for the Navier-Stokes equations: Second-order schemes. Math. Comput. 1996, 65, 1039–1066. [Google Scholar] [CrossRef]
  24. Huang, P.Z.; Zhang, T.; Ma, X.L. Superconvergence by L2-projection for a stabilized finite volume method for the stationary Navier-Stokes equations. Comput. Math. Appl. 2011, 62, 4249–4257. [Google Scholar] [CrossRef]
  25. Huang, P.Z.; Feng, X.L. Error estimates for two-level penalty finite volume method for the stationary Navier-Stokes equations. Math. Methods Appl. Sci. 2013, 36, 1918–1928. [Google Scholar] [CrossRef]
  26. Liu, X.; Li, J.; Chen, Z.X. A weak Galerkin finite element method for the Navier–Stokes equations. J. Comput. Appl. Math. 2018, 333, 442–457. [Google Scholar] [CrossRef]
  27. Li, B.Y.; Sun, W.W. Error analysis of linearized semi-implicit Galerkin finite element methods for nonlinear parabolic equations. Int. J. Numer. Anal. Model. 2013, 10, 622–633. [Google Scholar]
  28. Li, B.Y.; Sun, W.W. Unconditional convergence and optimal error estimates of a Galerkin-mixed FEM for incompressible miscible flow in porous media. SIAM J. Numer. Anal. 2013, 51, 1959–1977. [Google Scholar] [CrossRef]
  29. Pan, J. Global superconvergence for the bilinear-constant scheme for the Stokes problem. SIAM J. Numer. Anal. 1997, 36, 2424–2430. [Google Scholar] [CrossRef]
  30. Liu, H.P.; Yan, N.N. Global superconvergence for optimal control problems governed by Stokes equations. Int. J. Numer. Anal. Model. 2006, 3, 283–302. [Google Scholar]
  31. Adams, R.A.; Fournier, J.J.F. Sobolev Spaces; Academic Press: Cambridge, MA, USA, 2003. [Google Scholar]
  32. Lin, Q.; Lin, J.F. Finite Element Methods: Accuracy and Improvement; Science Press: Beijing, China, 2006. [Google Scholar]
  33. Heywood, J.G.; Rannacher, R. Finite element approximation of the nonstationary Navier-Stokes problem IV: Error analysis of second-order time-discretization. SIAM J. Numer. Anal. 1990, 27, 353–384. [Google Scholar] [CrossRef]
  34. Gao, H.D. Optimal error analysis of Galerkin FEMs for nonlinear Joule heating equations. J. Sci. Comput. 2014, 58, 627–647. [Google Scholar] [CrossRef]
  35. Xu, C.; Shi, D.Y.; Liao, X. Low order nonconforming mixed finite element method for nonstationary incompressible Navier-Stokes equations. Appl. Math. Mech. Engl. Ed. 2016, 37, 1095–1112. [Google Scholar] [CrossRef]
Figure 1. The element e (left) and e ˜ (right).
Figure 1. The element e (left) and e ˜ (right).
Mathematics 11 01945 g001
Figure 2. Local basis functions of P h .
Figure 2. Local basis functions of P h .
Mathematics 11 01945 g002
Figure 3. The element e ˜ .
Figure 3. The element e ˜ .
Mathematics 11 01945 g003
Figure 4. The graphics of u 1 n and u 1 h n on mesh 32 × 32 .
Figure 4. The graphics of u 1 n and u 1 h n on mesh 32 × 32 .
Mathematics 11 01945 g004
Figure 5. The graphics of u 2 n and u 2 h n on mesh 32 × 32 .
Figure 5. The graphics of u 2 n and u 2 h n on mesh 32 × 32 .
Mathematics 11 01945 g005
Figure 6. The graphics of the vector fields on mesh 32 × 32 .
Figure 6. The graphics of the vector fields on mesh 32 × 32 .
Mathematics 11 01945 g006
Figure 7. The graphics of p n and p h n on mesh 32 × 32 .
Figure 7. The graphics of p n and p h n on mesh 32 × 32 .
Mathematics 11 01945 g007
Table 1. The numerical errors and convergence orders at t = 0.1 of u .
Table 1. The numerical errors and convergence orders at t = 0.1 of u .
m × n 4 × 4 8 × 8 16 × 16 32 × 32
u n u h n 0 6.9832 × 10 4 1.7443 × 10 4 4.0642 × 10 5 1.0779 × 10 5
Order/2.00122.10161.9147
u n u h n 1 2.7565 × 10 2 1.3929 × 10 2 6.9778 × 10 3 3.4900 × 10 3
Order/0.984690.997300.99953
I h u n u h n 1 5.1545 × 10 3 1.7229 × 10 3 4.9365 × 10 4 1.1691 × 10 4
Order/1.58101.80332.0780
u n I 2 h u h n 1 1.6546 × 10 2 4.4691 × 10 3 1.1290 × 10 3 2.7815 × 10 4
Order/1.88841.98492.0211
Table 2. The numerical errors and convergence orders at t = 0.1 of p.
Table 2. The numerical errors and convergence orders at t = 0.1 of p.
m × n 8 × 8 16 × 16 32 × 32 64 × 64
p p h 0 1.1969×10 + 0 5.4990×10 1 2.6876×10 1 1.3356×10 1
Order/1.12211.03291.0088
J h p p h 0 5.7535×10 1 1.4260×10 1 3.6042×10 2 8.9144×10 3
Order/2.01251.98422.0154
p J 2 h p h 0 7.6225×10 1 1.8945×10 1 4.7655×10 2 1.1841×10 2
Order/2.00841.99112.0089
Table 3. The numerical errors and convergence orders at t = 0.6 of u .
Table 3. The numerical errors and convergence orders at t = 0.6 of u .
m × n 4 × 4 8 × 8 16 × 16 32 × 32
u n u h n 0 2.4270×10 4 6.3566×10 5 1.4898×10 5 3.9554×10 6
Order/1.93292.09311.9132
u n u h n 1 1.6737×10 8.4489×10 3 4.2323×10 3 2.1168×10 3
Order/0.986230.997340.99953
I h u n u h n 1 3.5044e×10 3 1.0579×10 3 3.0068×10 4 7.1119×10 5
Order/1.72791.81492.0799
u n I 2 h u h n 1 1.0152×10 2 2.7162×10 3 6.8536×10 4 1.6879×10 4
Order/1.90221.98672.0216
Table 4. The numerical errors and convergence orders at t = 0.6 of p.
Table 4. The numerical errors and convergence orders at t = 0.6 of p.
m × n 8 × 8 16 × 16 32 × 32 64 × 64
p p h 0 7.2597×10 1 3.3353×10 1 1.6301×10 1 8.1008×10 2
Order/1.12211.03291.0088
J h p p h 0 3.4897×10 1 8.6491×10 2 2.1860×10 2 5.4069×10 3
Order/2.01251.98422.0154
p J 2 h p h 0 4.6233×10 1 1.1491×10 1 2.8905×10 2 7.1817×10 3
Order/2.00841.99112.0089
Table 5. The numerical errors and convergence orders at t = 1.0 of u .
Table 5. The numerical errors and convergence orders at t = 1.0 of u .
m × n 4 × 4 8 × 8 16 × 16 32 × 32
u n u h n 0 1.2510×10 4 3.2199×10 5 8.0816×10 6 2.0220×10 6
Order/1.95801.99431.9989
u n u h n 1 1.1202×10 2 5.6624×10 3 2.8366×10 3 1.4189×10 3
Order/0.984300.997240.99938
I h u n u h n 1 1.7780×10 3 6.0911×10 4 1.6246×10 4 4.1252×10 5
Order/1.54551.90661.9775
u n I 2 h u h n 1 6.6492×10 3 1.7801×10 3 4.4318×10 4 1.1058×10 4
Order/1.90122.00602.0029
Table 6. The numerical errors and convergence orders at t = 1.0 of p.
Table 6. The numerical errors and convergence orders at t = 1.0 of p.
m × n 8 × 8 16 × 16 32 × 32 64 × 64
p p h 0 4.8472×10 1 2.2345×10 1 1.0923×10 1 5.4300×10 2
Order/1.11721.03251.0084
J h p p h 0 2.2993×10 1 5.7482×10 2 1.4370×10 2 3.5926×10 3
Order/2.00002.00002.0000
p J 2 h p h 0 3.0657×10 1 7.6642×10 2 1.9160×10 3 4.7901×10 3
Order/2.00002.00002.0000
Table 7. The numerical errors and convergence orders at t = 10 of u .
Table 7. The numerical errors and convergence orders at t = 10 of u .
m × n 4 × 4 8 × 8 16 × 16 32 × 32
u n u h n 0 3.9072×10 18 9.6945×10 19 2.4129×10 19 6.0248×10 20
Order/2.01092.00642.0018
u n u h n 1 2.0623×10 9 1.0366×10 9 5.1886×10 10 2.5950×10 10
Order/0.992410.998390.99959
I h u n u h n 1 3.0875×10 10 1.0500×10 10 2.8119×10 11 7.1483×10 12
Order/1.55601.90081.9759
u n I 2 h u h n 1 1.2562×10 9 4.2567×10 10 1.0848×10 10 2.7240×10 11
Order/1.56121.97231.9937
Table 8. The numerical errors and convergence orders at t = 10 of p.
Table 8. The numerical errors and convergence orders at t = 10 of p.
m × n 8 × 8 16 × 16 32 × 32 64 × 64
p p h 0 2.4684×10 10 7.9753×10 11 2.0617×10 11 5.1913×10 12
Order/1.63001.95171.9896
J h p p h 0 2.4684×10 10 7.9753×10 11 2.0617×10 11 5.1913×10 12
Order/1.63001.95171.9896
p J 2 h p h 0 2.8503×10 10 8.7022×10 11 2.1086×10 11 5.2215×10 12
Order/1.71172.04512.0138
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Meng, X.; Yang, H. Unconditional Superconvergence Error Estimates of Semi-Implicit Low-Order Conforming Mixed Finite Element Method for Time-Dependent Navier–Stokes Equations. Mathematics 2023, 11, 1945. https://doi.org/10.3390/math11081945

AMA Style

Meng X, Yang H. Unconditional Superconvergence Error Estimates of Semi-Implicit Low-Order Conforming Mixed Finite Element Method for Time-Dependent Navier–Stokes Equations. Mathematics. 2023; 11(8):1945. https://doi.org/10.3390/math11081945

Chicago/Turabian Style

Meng, Xiaoling, and Huaijun Yang. 2023. "Unconditional Superconvergence Error Estimates of Semi-Implicit Low-Order Conforming Mixed Finite Element Method for Time-Dependent Navier–Stokes Equations" Mathematics 11, no. 8: 1945. https://doi.org/10.3390/math11081945

APA Style

Meng, X., & Yang, H. (2023). Unconditional Superconvergence Error Estimates of Semi-Implicit Low-Order Conforming Mixed Finite Element Method for Time-Dependent Navier–Stokes Equations. Mathematics, 11(8), 1945. https://doi.org/10.3390/math11081945

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop