Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Apr 9.
Published in final edited form as: Appl Anal. 2019 Mar 25;100(1):116–134. doi: 10.1080/00036811.2019.1594201

Normal mode analysis of 3D incompressible viscous fluid flow models

Guoping Zhang 1, Mingchao Cai 1
PMCID: PMC8034259  NIHMSID: NIHMS1020022  PMID: 33840816

Abstract

In this paper, we study the normal mode solutions of 3D incompressible viscous fluid flow models. The obtained theoretical results are then applied to analyze several time-stepping schemes for the numerical solutions of the 3D incompressible fluid flow models.

Keywords: Normal mode analysis, Navier-Stokes equation, incompressible flow, boundary condition, splitting method, semi-discretization

1. Introduction

Many types of flow motions can be described by using various forms of incompressible viscous flow models [1]. In particular, the incompressible Navier–Stokes equations (NSEs) have been widely used to model many important applications [2] such as ocean currents [1], water flow in a pipe [3], water pollution, and blood flow in vessels. The NSEs are useful because they describe the physics of many phenomena of scientific and engineering interest. The NSEs for a viscous, incompressible flow read as

vt+p=vv+v2v+f(xΩ); (1)
v=0(xΩ). (2)

Here, v(x,t) is the velocity field at (x,t);p(x,t) is the pressure, v is the kinematic viscosity, f(x,t) is an external force and Ω is a bounded domain with piecewise smooth boundary ∂Ω. The derivation of the NSEs can be found in many references such as [49]. However, the theoretical analysis of NSEs in the three-dimensional case is notoriously difficult [4, 5, 912]. Many researchers resort to numerical investigation of incompressible viscous flow models [6, 1319]. Among various numerical algorithms for NSEs, the nonlinear term is usually treated explicitly while the reaction-diffusion term is solved implicitly. The pressure term can be viewed as a Lagrange multiplier or can be solved by using a deduced pressure Poisson equation [13, 15, 18]. Nevertheless, various numerical algorithms may lead to instability, inaccuracy, and artificial numerical boundary layers. In the literature, a powerful tool called normal mode analysis is frequently used to analyze the stability and the errors of various numerical methods. Such an analysis provides a convenient framework for the analysis of numerical methods to solve incompressible flow problems. In the normal mode analysis, we first linearize the nonlinear model (if the original model is nonlinear), then we find all possible eigenfunctions (so-called normal modes) of the eigenvalue problem associated with the linearized model. The general operator theory shows that all normal modes form a basis of a properly selected function space for the solutions so that the solution of the linearized model is a superposition of the normal modes. Therefore, we only need to test the stability of each normal mode solution in order to test it for the arbitrary solution of the problem. Our focus is the 3D NSEs because the increase in dimension will lead to essential difficulties on both theoretical aspect and numerical aspect. Correspondingly, the analysis is more difficult. In this paper, we start with the analysis of a quasi-one dimensional Stokes model. We derive the normal mode solution in a detailed way. Then, we investigate the stability and accuracy of several time-stepping schemes for the solution of incompressible viscous flow models.

The rest of the paper is organized as follows: in Section 2 we derive the normal mode solutions to a one-dimensional linearized model of NSEs. In Section 3, we perform the normal mode analysis to the Backward Euler time differencing method and Crank-Nicolson time differencing method. We study the stability and accuracy of solutions of a splitting method in the Section 4. Concluding remarks are given in Section 5.

2. Normal mode solutions

The linearized model of NSEs without the external force is the following three-dimensional Stokes equations :

vt+p=v2v(xΩ); (3)
v=0(xΩ). (4)

A one-dimensional reduced linear model that embodies the essential features of the incompressibility and viscous terms of the Navier-Stokes equations can be obtained by considering the solution of the form

v=(u(x,t),v(x,t),w(x,t))ei(ky+lz),p=p(x,t)ei(ky+lz)

for the three-dimensional Stokes equations with homogeneous Dirichlet boundary conditions along x-direction:

v(x,t)=0(ifx=1andx=1). (5)

To proceed normal mode analysis, as stated as above, we assume that the solution (v,p) is periodic along y and z directions with k and l being the corresponding wave numbers. Denoting D = ∂/∂x, then equations satisfied by (u, v, w, p) are

ut=Dp+v[D2(k2+l2)]u; (6)
vt=ikp+v[D2(k2+l2)]v; (7)
wt=ilp+v[D2(k2+l2)]w; (8)
Du+ikv+ilw=0. (9)

Here and there after, without losing generality, we set the domain

Ω=[1,1]×[π,π]×[π,π]={(x,y,z)|1x1;πyπ;πzπ}.

The homogeneous Dirichlet boundary conditions then become

u(±1,t)=v(±1,t)=w(±1,t)=0, (10)

and the wave numbers k and l become integers.

We notice that (u, v, w, p) = (0,0,0, p(t)) is always a solution of the problem (3)(5) for any single variable function p(t). This type of solutions are called trivial solutions. In this paper, we are only interested in nontrivial solutions of the problem (3)(5).

If the nontrivial solution of the boundary-value problem (6)(10) is in separable form:

u(x,t)=u^(x)U(t),v(x,t)=v^(x)V(t),w(x,t)=w^(x)W(t),p(x,t)=p^(x)P(t),

then we have the following proposition, which actually states that the time frequency is a negative number σ and ensures the solution is stable.

Proposition 2.1:

Without losing generality, we assume that

U(0)=V(0)=W(0)=P(0)=1.

Then, there exists a real constant σ < 0 such that

U(t)=V(t)=W(t)=P(t)=eσt.

Proof:

We break the proof into two steps: in first step we prove the existence of the real constant σ and then we prove that σ must be negative for a nontrivial solution in the second step.

After substituting the separate forms of u(x, t), v(x, t), w(x, t) and p(x, t) into (6)(10), we obtain

u^dUdt=(Dp^)P+v[D2(k2+l2)]u^U; (11)
v^dVdt=ikp^P+v[D2(k2+l2)]v^V; (12)
w^dWdt=ilp^P+v[D2(k2+l2)]w^W; (13)
(Du^)U+ikv^V+ilw^W=0. (14)

Applying the operator D on Equation (11), we have

(Du^)dUdt=(D2p^)P+v[D2(k2+l2)](Du^)U. (15)

Multiplying ik to Equation (12) and il to Equation (13) respectively, we obtain

(ikv^)dVdt=k2p^P+v[D2(k2+l2)](ikv^)V; (16)
(ilw^)dWdt=l2p^P+v[D2(k2+l2)](ilw^)W. (17)

Differentiating Equation (14) with respect to the variable t, we get

(Du^)dUdt+ikv^dVdt+ilw^dWdt=0. (18)

We denote

r=k2+l2. (19)

Applying (14) and (18) on the sum of (15), (16) and (17), we obtain

[(D2r2)p^]P(t)=0,

which implies

(D2r2)p^=0. (20)

Thus, we have

p^(x)=c1sinhrx+c2coshrx. (21)

Applying (D2r2) on (11), we get

(D2r2)u^dUdt=v(D2r2)2u^U,

which implies

dUdtU=v(D2r2)2u^(D2r2)u^=σ1. (22)

Here, σ1 must be a constant since the left-hand side of the above equation is a function depending on t only while the right-hand side of the above equation is a function depending on x only. Combining with U(0) = 1, we obtain

U(t)=eσ1t.

Applying (D2r2) on (12), we get

(D2r2)v^dVdt=v(D2r2)2v^V,

which implies that

dVdtV=v(D2r2)2v^(D2r2)v^=σ2

with σ2 being a constant. Combining with V(0) = 1, we obtain

V(t)=eσ2t.

Similarly, we obtain

dWdtW=v(D2r2)2w^(D2r2)w^=σ3

and

W(t)=eσ3t. (23)

Substituting V and W into (12) and (13) respectively, we obtain

e(σ2σ3)t=klσ3w^v(D2r2)w^σ2v^v(D2r2)v^,

which implies that σ2 = σ3 = σ since the right hand of the above equation is independent of t. Similarly, after multiplying ik on (11) and applying D on (12), we obtain

e(σ1σ2)t=σ2Dv^v(D2r2)Dv^ik(σ1u^v(D2r2)u^),

which implies that σ2 = σ1 = σ since the right hand of the above equation is independent of t.

Therefore, we complete the proof for the existence of the real constant σ.

Now we show that σ < 0 in the nontrivial solution of the form

u(x,t)=u^(x)eσt,v(x,t)=v^(x)eσt,w(x,t)=w^(x)eσt,p(x,t)=p^(x)eσt (24)

to the boundary-value problem (6)(10). This kind of solutions are called normal modes. Substituting (24) into (6)(10), we have

σu^=Dp^+v[D2r2]u^; (25)
σv^=ikp^+v[D2r2]v^; (26)
σw^=ilp^+v[D2r2]w^; (27)
Du^+ikv^+ilw^=0. (28)

The homogeneous Dirichlet boundary conditions become

u^(±1)=v^(±1)=w^(±1)=0. (29)

From (28) and (29), we obtain

Du^(±1)=0. (30)

It is easy to show that if r=k2+l2=0, then there are only trivial solutions to the problem (3)(5). Therefore, we can assume that r ≥ 1 from now on.

Applying (ikD) on (26) and (27), then adding the resulting equations, we obtain

σD2u^=r2Dp^v[D2r2]D2u^, (31)

which leads to the following equation after combining with (25):

v[D2r2]2u^=σ[D2r2]u^. (32)

We can rewrite (32) as

(D2r2)(D2r2σ/v)u^=0.

The corresponding characteristic equation is

(m2r2)(m2r2σ/v)=0.

Defining

μ=σ/vr2,

then we can show that σ must be negative by proving the following stronger statement.

Claim: r2+σ/v<0, and therefore μ=(σ/v+r2)>0andσ=v(r2+μ2)<0.

Proof:

We assert that r2+σ/v<0 and therefore μ > 0. We use the argument by contradiction to prove this statement. To proceed, we discuss the following two cases. ■

Case (1): r2+σ/v>0.

Let μ˜=r2+σ/ν.Ifσ0 then the general solution of (32) is

u^=c1cosh(rx)+c2sinh(rx)+c3cosh(μ˜x)+c4sinh(μ˜x),Du^=r(c1sinh(rx)+c2cosh(rx))+μ˜(c3sinh(μ˜x)+c4cosh(μ˜x)).

The boundary conditions give use that

u^(±1)=0,Du^(±1)=0.

This implies that

(coshrsinhrcoshrsinhr)(c1c2)=(coshμ˜sinhμ˜coshμ˜sinhμ˜)(c3c4) (33)

and

(rsinhrrcoshrrsinhrrcoshr)(c1c2)=(μ˜sinhμ˜μ˜coshμ˜μ˜sinhμ˜μ˜coshμ˜)(c3c4). (34)

More clearly, from Equation (33), we have

c1=coshμ˜coshrc3,c2=sinhμ˜sinhrc4,

from Equation (34), we see that

c1=μ˜sinhμ˜rsinhrc3,c2=μ˜coshμ˜rcoshrc4.

If c3 ≠ 0, then μ˜tanhμ˜ = r tanh r.

If c4 ≠ 0, then tanh μ˜/μ˜ = tanh r/r.

Since x tanh x is strictly increasing on the interval [0, ∞) and tanh x/x is strictly decreasing on [0, ∞), there must hold μ˜=r, which implies that σ = 0. This is a contradiction!

If σ = 0, then the general solution of (32) is

u^=c1sinhrx+c2coshrx+x(c3sinhrx+c4coshrx).

By a similar argument we can derive a contradiction. Therefore, in Case (1), there is only a trivial solution c1 = c2 = c3 = c4 = 0 for (32).

Case (2): r2+σ/v = 0.

The general solution of (32) is

u^=c1sinhrx+c2coshrx+c3x+c4.

By a similar argument, we can prove that there is only trivial solution c1 = c2 = c3 = c4 = 0 for (32) for Case (2). Thus the Claim has been proved. Therefore the proof of Proposition 2.1 is completed.

From Proposition 2.1, we only need to consider the following ODE with μ > 0.

v(D2r2)2u^=σ(D2r2)u^, (35)

and u^ satisfies the boundary conditions

u^(±1)=0,Du^(±1)=0. (36)

We notice that σ is actually an eigenvalue and the nonzero function u^ is the associated eigenfunction to the boundary value problem (35)(36).

The general solution of (35) is

u^=c1cosh(rx)+c2sinh(rx)+c3cos(μx)+c4sin(μx).

Therefore,

Du^=c1rsinh(rx)+c2rcosh(rx)c3μsin(μx)+c4μcos(μx).

Applying the boundary conditions (36), we obtain

(coshrsinhrcoshrsinhr)(c1c2)=(cosμsinμcosμsinμ)(c3c4)

and

(rsinhrrcoshrrsinhrrcoshr)(c1c2)=(μsinμμcosμμsinμμcosμ)(c3c4).

We only discuss the nontrivial solutions. Notice that if c4 = 0 and c3 ≠ 0, then −μ tan μ = r tanh r; if c4 ≠ 0 and c3 = 0, then μ cot μ = r coth r; while c4 ≠ 0, c3 ≠ 0 implies a contradiction. Thus, we have the following conclusion.

Proposition 2.2:

The boundary value problem (35)(36) has only the following two type of nontrivial solutions for r ≥ 1.

Case 1 (even solution):

μ=r2σ/v satisfies

μtanμ=rtanhr

and the corresponding eigenvalue σ=v(μ2+r2)<0 and the associated eigenfunction is a constant multiple of

u^=cosμcoshrxcoshrcosμx.

Correspondingly,

v^=ik[cosμrsinhrx+μcoshrr2sinμx],w^=il[cosμrsinhrx+μcoshrr2sinμx],p^=σcosμrsinhrx.

Case 2 (oddsolution):

μ=r2σ/v satisfies

μcotμ=rcothr,

and the corresponding eigenvalue σ = −v(μ2 + r2) < 0 and the associated eigenfunction is a constant multiple of

u^=sinμsinhrxsinhrsinμx.

Correspondingly,

v^=ik[sinμrcoshrxμsinhrr2cosμx],w^=il[sinμrcoshrxμsinhrr2cosμx],p^=σsinμrcoshrx.

Remark 2.3:

(1) For given r=k2+l21, by a simple analysis we know that on each interval ((π/2)+(k1)π,(π/2)+kπ),k=1,2,, there is a unique μk such that

μktanμk=rtanhr,

and the eigenvalue σk=v(μk2+r2)<0, the associated eigenfunction u^ is denoted by uk.

(2) For given r ≥ 1, on each interval (kπ,(k+1)π),k=1,2,, there is a unique μ˜k such that

μ˜kcotμ˜k=rcothr,

and the eigenvalue σ˜k=v(μ˜k2+r2)<0, the associated eigenfunction u^ is denoted by u˜k.

(3) If we define the function space

W02[1,1]={uH2[1,1]:u(±1)=Du(±1)=0},

then the general operator theory shows that the set of all eigenfunctions {uk,u˜k:k=1,2,3,} forms a basis of W02[1,1] (see [6]).

3. Semidiscrete implicit coupled methods

In this section, we are seeking normal mode solutions to some semidiscrete equations which are analogues to the normal mode solutions of the continuous model problem in Section 2. The corresponding normal modes are of the form

(un(x),vn(x),wn(x),pn(x))=Kn(u˜(x),v˜(x),w˜(x),p˜(x)), (37)

where K is the amplification factor and n denotes the n-th time step. The semidiscrete approximation is stable if |K|1+O(Δt) for all normal modes and is unstable otherwise [6]. The accuracy of the particular semidiscrete methods can be studied by computing the exponential growth rate σ˜ defined by

Kn=exp(σ˜nΔt),σ˜=lnKΔt. (38)

The error σ˜σ, where σ is given by (35)(36), measures the time discretization error.

3.1. Backward Euler time differencing

If the system (6)(10) is approximated by the backward Euler time differencing scheme, denoting r=k2+l2, then

un+1unΔt=Dpn+1+v[D2r2]un+1, (39)
un+1vnΔt=ikpn+1+v[D2r2]vn+1, (40)
wn+1wnΔt=ilpn+1+v[D2r2]wn+1, (41)
Dun+1+ikvn+1+ilwn+1=0, (42)
un+1(±1)=vn+1(±1)=wn+1(±1)=0, (43)

which imply

Dun+1(±1)=0. (44)

Applying D on (39), we have

Dun+1DunΔt=D2pn+1+v[D2r2](Dun+1). (45)

Multiplying ik to (40), we see that

ikun+1ikvnΔt=k2pn+1+v[D2r2](ikvn+1). (46)

Multiplying il to (41), we have

ilwn+1ilwnΔt=l2pn+1+v[D2r2](ilwn+1). (47)

Adding (45)(47) together and applying (42), we obtain

(D2r2)pn+1=0.

Applying (D2r2) on (37), we obtain

(D2r2)(un+1un)Δt=D(D2r2)pn+1+v(D2r2)2un+1, (48)
(D2r2)(un+1un)Δt=v(D2r2)2un+1, (49)
un+1(±1)=Dun+1(±1)=0. (50)

Substituting un=Knu˜,un+1=Kn+1u˜ into (49), we obtain

K1KΔt(D2r2)u˜=v(D2r2)2u˜,u˜(±1)=Du˜(±1)=0. (51)

Let

σ=K1KΔtK=11σΔt.

We have

σ(D2r2)u˜=v(D2r2)2u˜,

which is the same as (35). The time dependance of (un, vn, wn, pn) is proportional to Kn = exp (σ˜nΔt), where

σ˜=lnKΔt=ln(1σΔt)Δt.

Since

ln(1x)=n=0xn+1n+1

for small Δt, we have

ln(1σΔt)=n=0(σΔt)n+1n+1=σΔt+12σ2Δt2+13σ3Δt3+.

Thus

σ˜=σ+12σ2Δt+13σ3Δt2+,

which implies that

σ˜σ=O(Δt).

So the exponential growth rate σ˜ for the Backward Euler scheme is in error of the order O(Δt). Noting that σ < 0, we have K < 1, and therefore this numerical scheme is unconditionally stable. Thus, the normal analysis demonstrates the stability property of this implicit semidiscrete scheme. Moreover, this scheme is convergent if Δt → 0.

3.2. Crank-Nicolson time differencing

If the system (6)(10) is approximated by Crank-Nicolson scheme, then we have

un+1unΔt=Dpn+12+12v[D2r2](un+1+un), (52)
vn+1vnΔt=ikpn+12+12v[D2r2](vn+1+vn), (53)
wn+1wnΔt=ilpn+12+12v[D2r2](wn+1+wn), (54)
Dun+1+ikvn+1+ilwn+1=0, (55)

with

un+1(±1)=vn+1(±1)=wn+1(±1)=0. (56)

We will show that this numerical scheme is both unconditionally stable and accurate to the order of O(Δt2). Indeed, we can use a similar method to derive that

(D2r2)pn+12=0.

Applying (D2r2) on (52),we obtain

(D2r2)(un+1un)Δt=12v(D2r2)2(un+1+un). (57)

Let

un=Knu˜,vn=Knv˜,wn=Knw˜,

substituting them into (57), we have

(K1)Δt(D2r2)u˜=(K+1)v2(D2r2)2u˜.

Let

σ=2(K1)Δt(K+1),

we have

σ(D2r2)u˜=v(D2r2)2u˜,

which is the same as (35) where we have σ < 0. Let

σ=2(K1)Δt(K+1)K=2+σΔt2σΔt=1+12σΔt112σΔt<1.

Setting

Kn=exp(σ˜nΔt),

we have

σ˜=lnKΔt=ln(1+12σΔt)ln(112σΔt)Δt,ln(1+12σΔt)=n=0(1)n(12σΔt)n+1n+1,ln(112σΔt)=n=0(12σΔt)n+1n+1,σ˜=1Δtn=0[(1)n+1](12σΔt)n+1n+1=2Δtm=0σ2m+1Δt2m+1(2m+1)22m+1=m=0σ2m+1Δt2m(2m+1)22m=σ+σ312(Δt)2+σ580(Δt)4+....=σ+O(Δt2)

which implies

σ˜σ=O(Δt2).

As σ < 0, we see that σ˜<0and|K|<1. Thus, theexponentialgrowthratea for the Crank-Nicolson scheme is in error of the order O(Δt2) and this scheme is unconditionally stable.

4. Splitting method

In this section, we consider the velocity-pressure splitting with the normal velocity boundary conditions [15, 18]. The time-differencing scheme involves the following two split time steps.

The first step involves solution of the inviscid equation:

u*unΔt=Dpn+1, (58)
v*vnΔt=ikpn+1, (59)
w*wnΔt=ilpn+1, (60)
Du*+ikv*+ilw*=0, (61)
u*(±1)=v*(±1)=w*(±1)=0. (62)

The second step involves the solution of the viscous equation

un+1u*Δt=v(D2r2)un+1, (63)
un+1v*Δt=v(D2r2)vn+1, (64)
wn+1w*Δt=v(D2r2)wn+1, (65)
un+1(±1)=vn+1(±1)=wn+1(±1)=0. (66)

Remark: In the above splitting algorithm, (58)(66), un, vn, wn do not satisfy the incompressibility constraint, although the intermediate variables u*, v*, w* do.

4.1. The computation of ũ and ũ*

To analyze the splitting method, we firstly derive the results for the normal modes of the above system. The general conclusion then follow from the completeness ofthe normal modes.

Let

(un,vn,wn,pn)=Kn(u˜,v˜,w˜,p˜) (67)

and

(u*,v*,w*)=Kn(u˜*,v˜*,w˜*). (68)

Substituting (67) and (68) into (58)(66) gives

u˜*u˜=(KΔt)Dp˜, (69)
v˜*v˜=ik(KΔt)p˜, (70)
w˜*w˜=il(KΔt)p˜, (71)
Du˜*+ikv˜*+ilw˜*=0, (72)
u˜*(±1)=v˜*(±1)=w˜*(±1)=0, (73)
(D2r21vΔt)u˜=1KvΔtu˜*, (74)
(D2r21vΔt)u˜=1KvΔtv˜*, (75)
(D2r21vΔt)u˜=1KvΔtw˜*, (76)
u˜(±1)=v˜(±1)=w˜(±1)=0. (77)

Applying D on (69), multiplying ik on (70) and multiplying il on (71), then adding them together, we have

(KΔt)(D2r2)p˜=Du˜+ikv˜+ilw˜. (78)

Applying D on (74), multiplying ik on (75) and multiplying il on (76), then adding them together, we see that

(D2r21vΔt)(Du˜+ikv˜+ilw˜)=0. (79)

(78) and (79) imply that

(D2r2)(D2r21vΔt)p˜=0. (80)

Applying (D2r2)(D2r21vΔt) on (69) gives

(D2r2)(D2r21vΔt)u˜*=(D2r2)(D2r21vΔt)u˜=(D2r2)KvΔtu˜*,(D2r2)2u˜*=K1KvΔt(D2r2)u˜*. (81)

(72) and (73) imply

Du˜*(±1)=u˜*(±1)=0. (82)

The characteristic equation of (81)

(m2r2)2=K1KνΔt(m2r2)m=±r,±r2K1KvΔt.

Let

μ˜=(r2K1KvΔt)1/2>0.

The general solution of Equation (81) can be expressed as

u˜*=c1coshrx+c2sinhrx+c3cosμ˜x+c4sinμ˜x,Du˜*=c1rsinhrx+c2rcoshrxc3μ˜sinμ˜x+c4μ˜cosμ˜x,u˜*(1)=0c1coshr+c2sinhr+c3cosμ˜+c4sinμ˜=0,u˜*(1)=0c1coshrc2sinhr+c3cosμ˜c4sinμ˜=0,Du˜*(1)=0c1rsinhr+c2rcoshrc3μ˜sinμ˜+c4μ˜cosμ˜=0,Du˜*(1)=0c1rsinhr+c2rcoshr+c3μ˜sinμ˜+c4μ˜cosμ˜=0,

which imply that

c1=cosμ˜coshrc3,c2=sinμ˜sinhrc4,c1=μ˜sinμ˜rsinhrc3,c2=μ˜cosμ˜rcoshrc4.

By a similar analysis to that in Section 2, we obtain the boundary value problem (81)(84) has only the following two types of solutions.

(i) Even solution: The eigenvalue μ˜ satisfies

μ˜tanμ˜=rtanhr,

and the corresponding eigenfunction is a constant multiple of

u˜*=cosμ˜coshrxcoshrcosμ˜x.

(ii) Odd solution: The eigenvalue μ˜ satisfies

μ˜cotμ˜=rcothr,

and the corresponding eigenfunction is a constant multiple of

u˜*=sinμ˜sinhrxsinhrsinμ˜x.

Let

σ=K1KΔt,

then (81) becomes

v(D2r2)2u˜*=σ(D2r2)u˜*,

which is exactly the same as (35). Since

μ˜=(r2σv)1/2=μ,

we see that

K=11σΔt=eσ˜Δt.

The leading behavior σ of the growth rate σ˜ agrees with that of the exact solution:

σ˜=lnKΔt=ln(1σΔt)Δt=σ+12σ2Δt+13σ3Δt3+, (83)

which implies that

σ˜σ=O(Δt).

Let

λ=(r2+1vΔt)1/2>0.

Equation (74) becomes

(D2λ2)u˜=1KvΔtu˜*,
u˜(±1)=0. (84)

We consider two cases:

Case (i):

u˜*=cosμcoshrxcoshrcosμx

is even.

Noting that r ≠ λ and r ≠ μ, we see that the general solution of (84) is

u˜=c1coshλx+c2sinhλx+u˜p.

Here,

u˜p=Acoshrx+Bcosμx,Du˜p=ArsinhrxBμsinμx,D2u˜p=Ar2coshrxBμ2cosμx,(D2λ2)u˜p=A(r2λ2)coshrxB(μ2+λ2)cosμx,=cosμKvΔtcoshrx+coshrKvΔtcosμx, (85)

which implies that

A=cosμKvΔt(λ2r2),B=coshrKvΔt(λ2+μ2),u˜p=cosμcoshrxKvΔt(λ2r2)coshrcosμxKvΔt(λ2+μ2).

Noting that

K=11σΔtK(1σΔt)=1.

Thus,

u˜=c1coshλx+c2sinhλx+cosμcoshrxKvΔt(λ2r2)coshrcosμxKvΔt(λ2+μ2).

From

λ2r2=1vΔt,λ2+μ2=1vΔtσv,

we obtain

u˜=c1coshλx+c2sinhλx+(1σΔt)cosμcoshrxcoshrcosμx.

Applying the boundary conditions:

u˜(1)=0c1coshλ+c2sinhλσΔtcosμcoshr=0,u˜(1)=0c1coshλc2sinhλσΔtcosμcoshr=0,

we have

c1=σΔtcosμcoshrcoshλ,c2=0.

Thus,

u˜=σΔtcosμ(coshrcoshλxcoshλcoshrx)+(cosμcoshrxcoshrcosμx)

is also even.

Case (ii):

ũ* = sin μ sinh rx − sinh r sin μx is odd.

u˜=c1coshλx+c2sinhλx+u˜p,

where

u˜p=Asinhrx+Bsinμx,Du˜p=Arcoshrx+Bμcosμx,D2u˜p=Ar2sinhrxBμ2sinμx,(D2λ2)u˜p=1KvΔtu˜*=sinμKvΔtsinhrx+sinhrKvΔtsinμx,A(r2λ2)sinhrxB(μ2+λ2)sinμx=sinμKvΔtsinhrx+sinhrKvΔtsinμxA=sinμK,B=sinhr.u˜=c1coshλx+c2sinhλx+sinμsinhrxKsinhrsinμx.

Applying the boundary condition,

u˜(1)=0c1coshλ+c2sinhλ=sinμsinhrK+sinμsinhr,u˜(1)=0c1coshλc2sinhλ=sinμsinhrKsinhrsinμ.

We have

c1=0,c2=σΔtsinμsinhrsinhλ.

Thus,

u˜=σΔtsinμ(sinhrsinhλxsinhλsinhrx)+(sinμsinhrxsinhrsinμx)

is also odd.

Since ũ* is the exact normal mode solution, we can compute the error u˜u˜* as follows.

Case (i): ũ* = cos μ cosh rx − cosh r cos μx,

u˜=σΔtcosμ(coshrcoshλxcoshλcoshrx)+(cosμcoshrxcoshrcosμx),

which implies that

u˜u˜*=σΔtcosμ(coshrcoshλxcoshλcoshrx).

Obviously, we have the error estimate:

u˜u˜*=O(Δt).

The error is uniformly bounded with respect to x in [−1,1] as Δt0 because

λ=(r2+1vΔt)12asΔt0,|x|<1coshλxcoshλ0asλ.

Case (ii):ũ* = sin μ sinh rx − sinh r sin μx,

u˜=σΔtsinμ(sinhrsinhλxsinhλsinhrx)+(sinμsinhrxsinhrsinμx),

which implies that

u˜u˜*=O(Δt).

The error is uniformly bounded with respect to x in [1,1]asΔt 0 because

λ=(r2+1vΔt)12asΔt0,|x|<1sinhλxsinhλ0asλ.

4.2. The Computation of p˜

Case (i): ũ* is even.

From (69),

Dp˜=u˜u˜*KΔt=σKcosμcoshλ[coshrcoshλxcoshλcoshrx]p˜=σKcosμcoshλ[coshrλsinhλxcoshλrsinhrx]+c

where c is a constant. From (81):

(D2r2)(D2λ2)p˜=0.p˜=c1coshrx+c2sinhrx+c3coshλx+c4sinhλx.

Thus, we see that c = 0 and

p˜=σcosμcoshλ[coshrλsinhλxcoshλrsinhrx](1σΔt).

Notice that

p^=σcosμrsinhrx,p˜p^=σcosμcoshλcoshrλsinhλxΔtσ2cosμcoshλ[coshrλsinhλxcoshλrsinhrx].

Moreover,

λ=(r2+1vΔt)1/2=O(Δt(1/2))1λ=O(Δt1/2)

Because sinh λ x/cosh λ is uniformly bounded on [−1,1] as Δt → 0, we have

p˜p^=O(Δt1/2).

Case (ii): ũ* is odd.

Dp˜=u˜u˜*KΔt=σK[sinhrsinhλxsinhλsinhrx]sinμsinhλ.

From (80),

p˜=c1coshrx+c2sinhrx+c3coshλx+c4sinhλx.

Thus,

p˜=σsinμsinhλ[sinhrλcoshλxsinhλrcoshrx](1σΔt),p^=σsinμrcoshrx,p˜p^=σsinμsinhrλcoshλsinhλΔtσ2sinμsinhλ[sinhrλcoshλxsinhλrcoshrx].

We see that

p˜p^=O(Δt1/2).

Similarly, one can compute v˜,v˜*,w˜,w˜*. As the derivations are similar, we omit the details.

5. Concluding remarks

In this work, we derive the normal mode solution of a 3D linearized incompressible fluid flow model. Then, we apply the results to analyze some implicit time stepping schemes including the Backward Euler and the Crank-Nicolson schemes, as well as the splitting method. By using the normal mode analysis, we rigorously prove that both the Backward Euler scheme and Crank-Nicolson schemes for the 3D Stokes equations are unconditionally stable; The time errors of the Backward Euler scheme and the Crank-Nicolson scheme are of the order O(Δt)andO(Δt2), respectively. Moreover, based on the normal mode analysis, we give the estimates of error orders of each variable and the intermediate variables for the splitting method.

Acknowledgments

Funding

This author’s work is supported in part by the NIH BUILD grant (ASCEND pilot project) through UL1GM118973, NSF HBCU-UP Research Initiation Award through HRD-1700328 andNSF HBCU-UP Excellence in Research Award through DMS-1831950.

Footnotes

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • [1].Lions JL, Temam R, Wang S. Mathematical theory for the coupled atmosphere-ocean models (CAO III). J Math Pures Appl. 1995;74(9):105–163. [Google Scholar]
  • [2].Constantin P, Foias C, Manley O, et al. Determining modes and fractal dimension of turbulent flow. J Fluid Mech. 1985;150:427–440. [Google Scholar]
  • [3].Babin AV. The attractor of a Navier-Stokes system in an unbounded channel-like domain. J Dyn Differ Equ. 1992;4:555–584. [Google Scholar]
  • [4].Constantin P, Foias C. Navier-Stokes equations. Chicago: University of Chicago Press; 1988. [Google Scholar]
  • [5].Foias C, Manley O, Rosa R, et al. Navier-Stokes equations and turbulence. Cambridge: Cambridge University Press; 2001. [Google Scholar]
  • [6].Orszag SA, Israeli M, Deville MO Boundary conditions for incompressible flows. J SciComput. 1986;1(1):75–111. [Google Scholar]
  • [7].Rhebergen S, Cockburn B, van der Vegt J. A space time discontinuous Galerkin method for the incompressible Navier Stokes equations. J Comput Phys. 2013;233:339–358. [Google Scholar]
  • [8].Reynolds O On the dynamical theory of incompressible viscous fluids and the determination of the criterion. Philos Trans R Soc Lond Ser A. 1895;186:123–164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].Temam R Navier-Stokes equations: theory and numerical analysis. 1st ed. Providence (RI): American Mathematical Society; published in 1979. [Google Scholar]
  • [10].Fefferman C Existence and smoothness of the Navier-Stokes equation. The Millennium Prize Probl. 2006;57:67. [Google Scholar]
  • [11].Hoff D Global well-posedness of the Cauchy problem Stokes equations of nonisentropic flow with discontinuous initial data. J Differ Equ. 1992;95:33–74. [Google Scholar]
  • [12].Nan Z, Zheng X. Existence and uniqueness of solutions for Navier-Stokes equations with hyper-dissipation in a large space. J Differ Equ. 2016;261(6):3670–3703. [Google Scholar]
  • [13].Brown D, Cortez R, Minion M. Accurate projection methods for the incompressible Navier Stokes equations. J Comput Phys. 2001;168(2):464–99. April 10 [Google Scholar]
  • [14].Cahouet J, Chabard J-P. Some fast 3D finite element solvers for the generalized Stokes problem. Int J Numer Methods Fluids. 1988;8(8):869–895. [Google Scholar]
  • [15].Chorin AJ. Numerical solution of the Navier-Stokes equations. Math Comput. 1968;22:745–762. [Google Scholar]
  • [16].Dubois T, Jauberteau F, Temam R. Incremental unknowns, multilevel methods and the numerical simulation of turbulence. Comput Methods Appl Mech Eng. 1998;159:123–189. [Google Scholar]
  • [17].Girault V, Raviart P. Finite element methods for Navier-Stokes equations: theory and algorithms. New York Tokyo: Springer-Verlag Berlin Heidelberg; 1986. [Google Scholar]
  • [18].Kim J, Moin P. Application of a fractional-step method to incompressible Navier Stokes equations. J Comput Phys. 1985;59:308–323. [Google Scholar]
  • [19].Taylor C, Hood P. A numerical solution of the Navier-Stokes equations using the finite element technique. Comput Fluids. 1973;1(1):73–100. [Google Scholar]

RESOURCES