+ All Categories
Home > Documents > A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA...

A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA...

Date post: 26-Jun-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
25
A pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania Ben-Artzi a , Jean-Pierre Croisille b , Dalia Fishelov c,d, * , Shlomo Trachtenberg e a Institute of Mathematics, The Hebrew University, Jerusalem 91904, Israel b Department of Mathematics, University of Metz, Metz, France c Tel-Aviv Academic College of Engineering, 218 Bnei-Efraim St., Tel-Aviv 69107, Israel d School of Mathematical Sciences, Tel Aviv University, Ramat Aviv, Tel Aviv 69978, Israel e Department of Membrane and Ultrastructure Research, The Hebrew University-Hadassah Medical School, P.O. Box 12271, Jerusalem 91120, Israel Received 28 April 2004; received in revised form 10 November 2004; accepted 28 November 2004 Available online 21 December 2004 Abstract A pure-streamfunction formulation is introduced for the numerical simulation of the two-dimensional incompress- ible Navier–Stokes equations. The idea is to replace the vorticity in the vorticity-streamfunction evolution equation by the Laplacian of the streamfunction. The resulting formulation includes the streamfunction only, thus no inter-function relations need to be invoked. A compact numerical scheme, which interpolates streamfunction values as well as its first order derivatives, is presented and analyzed. A number of numerical experiments are presented, including driven and double driven cavities, where the Reynolds numbers are sufficiently large, leading to symmetry breaking of asymptotic solutions. Ó 2004 Elsevier Inc. All rights reserved. Keywords: Navier–Stokes equations; Streamfunction formulation; Vorticity; Numerical algorithm; Compact schemes; Driven cavity; Symmetry breaking; Asymptotic behavior 1. Introduction A new methodology for tracking vorticity dynamics was introduced in [3,12]. More specifically, we stud- ied the time evolution of the planar flow subject to the Navier–Stokes equations. It is the purpose of the 0021-9991/$ - see front matter Ó 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.jcp.2004.11.024 * Corresponding author. Tel.: +972 3 640 5397; fax: +972 3 640 9357. E-mail addresses: [email protected] (M. Ben-Artzi), [email protected] (J.-P. Croisille), [email protected] (D. Fishelov), [email protected] (S. Trachtenberg). Journal of Computational Physics 205 (2005) 640–664 www.elsevier.com/locate/jcp
Transcript
Page 1: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Journal of Computational Physics 205 (2005) 640–664

www.elsevier.com/locate/jcp

A pure-compact scheme for the streamfunction formulationof Navier–Stokes equations

Matania Ben-Artzi a, Jean-Pierre Croisille b,Dalia Fishelov c,d,*, Shlomo Trachtenberg e

a Institute of Mathematics, The Hebrew University, Jerusalem 91904, Israelb Department of Mathematics, University of Metz, Metz, France

c Tel-Aviv Academic College of Engineering, 218 Bnei-Efraim St., Tel-Aviv 69107, Israeld School of Mathematical Sciences, Tel Aviv University, Ramat Aviv, Tel Aviv 69978, Israel

e Department of Membrane and Ultrastructure Research, The Hebrew University-Hadassah Medical School, P.O. Box 12271,

Jerusalem 91120, Israel

Received 28 April 2004; received in revised form 10 November 2004; accepted 28 November 2004

Available online 21 December 2004

Abstract

A pure-streamfunction formulation is introduced for the numerical simulation of the two-dimensional incompress-

ible Navier–Stokes equations. The idea is to replace the vorticity in the vorticity-streamfunction evolution equation by

the Laplacian of the streamfunction. The resulting formulation includes the streamfunction only, thus no inter-function

relations need to be invoked. A compact numerical scheme, which interpolates streamfunction values as well as its first

order derivatives, is presented and analyzed. A number of numerical experiments are presented, including driven and

double driven cavities, where the Reynolds numbers are sufficiently large, leading to symmetry breaking of asymptotic

solutions.

� 2004 Elsevier Inc. All rights reserved.

Keywords: Navier–Stokes equations; Streamfunction formulation; Vorticity; Numerical algorithm; Compact schemes; Driven cavity;

Symmetry breaking; Asymptotic behavior

1. Introduction

A new methodology for tracking vorticity dynamics was introduced in [3,12]. More specifically, we stud-

ied the time evolution of the planar flow subject to the Navier–Stokes equations. It is the purpose of the

0021-9991/$ - see front matter � 2004 Elsevier Inc. All rights reserved.

doi:10.1016/j.jcp.2004.11.024

* Corresponding author. Tel.: +972 3 640 5397; fax: +972 3 640 9357.

E-mail addresses: [email protected] (M. Ben-Artzi), [email protected] (J.-P. Croisille), [email protected] (D.

Fishelov), [email protected] (S. Trachtenberg).

Page 2: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 641

present paper to upgrade this methodology by further reducing the role of vorticity and concentrating on

the streamfunction instead.

We recall the basic setup. Let X ˝ R2 be a bounded, simply connected domain with smooth boundary

oX. An incompressible, viscid flow in X is governed by the Navier–Stokes equations [21] (in its ‘‘vortic-

ity–velocity’’ formulation)

otnþ ðu � rÞn ¼ mDn in X; ð1:1Þ

r � u ¼ 0 in X: ð1:2Þ

The system (1.1) and (1.2) expresses the evolution of the vorticity n = oxv � oyu, where u = (u,v) is the veloc-

ity (and (x,y) are the coordinates in X). The coefficient m > 0 is the viscosity coefficient.The system (1.1) and (1.2) is supplemented by the initial data

n0ðx; yÞ ¼ nðx; y; tÞjt¼0; ðx; yÞ 2 X ð1:3Þ

and a boundary condition on oX. Indeed, as has been discussed in [3], this condition is the ‘‘source of

(numerical and theoretical) trouble’’, since it is normally expressed in terms of the velocity, rather than

the vorticity. In our presentation here we take the most common condition, the so-called ‘‘no-slip’’

condition,

uðx; y; tÞ ¼ 0 for ðx; yÞ 2 oX and all t P 0: ð1:4Þ

The difficulty of ‘‘translating’’ (1.4) to a boundary condition adequate for use in (1.1) and (1.2) is a major

topic of any numerical simulation. In this paper, we overcome this difficulty by transforming the system

(1.1) and (1.2) to the ‘‘pure streamfunction’’ version. It will also be clear how to replace the boundary con-

dition (1.4) by more general ones. In fact, most of our numerical examples in this paper are studied in this

more general case.The vorticity formulation (1.1) and (1.2) has been the starting point for a wide variety of methods

designed to solve numerically the Navier–Stokes equations. Here, we mention two of them, which are

of particular relevance to the present paper. In fact, each one of them can be regarded as a ‘‘family’’

of algorithms, which share some common basic structural hypotheses, yet differ considerably in their

technical details. The first is generally labeled as the ‘‘vortex method’’. It consists of a wide array of

algorithms, all based on the approximation of the vorticity field by a collection of ‘‘singular objects’’

(such as point vortices, vortex filaments, etc.). These objects (which are often mathematically ‘‘regular-

ized’’) are advected and diffused in a way which preserves the main physical features of the flow.Clearly, the generation of vorticity on the boundary is of crucial significance in this approach. We refer

the reader to the recent book [6] for a comprehensive treatment. The second method is usually referred

to as the ‘‘vorticity-streamfunction’’ method, and has gained increasing attention in recent years (see e.g.

[4,7,8,14,15,18,22]). In some sense the streamfunction-vorticity formulation is an evolution of vortex

methods. However, while the latter is a ‘‘particle method’’, which does not require a grid, the former

assumes a smooth distribution of vorticity laid out on a regular grid. The vorticity equation (1.1) is then

treated by temporal discretization. Once again, there are numerous ways of handling the spatial discret-

ization, such as spectral techniques, finite differences, finite volume or finite element algorithms. Thevelocity field is typically updated, subject to the incompressibility constraint (1.2), by means of the evo-

lution of the streamfunction (see Section 2 below for some mathematical background). Consider for

example, the recent paper [4], where the evolution of the vorticity is accomplished by a ‘‘fractional step

scheme’’. The first step (hyperbolic) takes care of the advection. The second step, which is labeled there

as a ‘‘Stokes flow step’’, is a ‘‘parabolic-elliptic’’ system, where the vorticity is diffused by a heat-type

equation, coupled to a Poisson equation which ties the vorticity to the streamfunction. Since the Poisson

equation allows for only one boundary condition on the streamfunction (say, of Dirichlet type), the

Page 3: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

642 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

second one (see Section 2 below for details) must be accommodated by accounting appropriately for the

vorticity boundary values. Thus, boundary conditions for the vorticity must be brought into play. This

approach should be compared with Gresho�s observation [11, pp. 428, 429], that ‘‘there are no boundary

conditions for the vorticity, and none is needed’’ hence ‘‘the elliptic equation for the streamfunction

cannot be viewed in isolation because the inevitable conclusion is that it carries too many boundaryconditions’’. This ‘‘overdeterminacy’’ problem was addressed in the previous paper [3]. The key idea

of ‘‘vorticity projection’’ was introduced; instead of solving the Poisson equation n = Dw (which, as al-

ready observed, cannot take care of the two conditions on w) one solves Dn = D2w. The two boundary

conditions are applied directly to w, and there is no need for boundary values of the vorticity. This idea

is carried one step further in this paper (compare [12]). The vorticity ‘‘disappears’’ altogether and only

the streamfunction and its gradient (i.e., velocity) are discretized in a ‘‘box-scheme’’ style. This means

that all discretized values are attached to the grid nodes. The gradient values (which are regarded inde-

pendently) are related to the function values via suitable compatibility conditions, preserving the overallaccuracy of the scheme. We are therefore justified in labeling the scheme presented here as a ‘‘pure

streamfunction’’ scheme, which follows closely the theoretical treatment of Eqs. (1.1) and (1.2).

The plan of the paper is as follows. In Section 2 we recall the classical construction of the streamfunction

and present the mathematical background needed for our treatment. A basic element in this study consists

of using the bilaplacian D2 as the ‘‘driving generator’’ of the evolution. In Section 3 we describe our numer-

ical scheme, where the spatial discretization of D2 plays a significant role. Briefly, we assign, at each node,

values for the streamfunction and its gradient, and use a compact (second-order) scheme for D2. It allows a

‘‘clean’’ representation of the boundary condition, restricted fully to the boundary points. Section 4 is de-voted to a detailed study of questions of stability and convergence in a suitable linearized model. In Section

5, we present detailed results of numerical experiments, including a driven and a doubly-driven cavity.

Here, we go beyond the mere inspection of the time evolution and study also aspects of asymptotic behavior

and ‘‘breakdown of symmetry’’ [3,17].

The present scheme, as well as some numerical results, have been presented by three of the authors at a

conference (see [9]).

2. Pure-streamfunction formulation

The streamfunction w(x,y,t) was already introduced by Lagrange (see [13]) as a prime object in the inves-

tigation of the two-dimensional incompressible flow. The incompressibility condition (1.2) entails the exis-

tence of a function w(x,y,t) such that, for any fixed t P 0,

uðx; tÞ ¼ r?w ¼ � owoy

;owox

� �: ð2:1Þ

It follows that n = Dw and Eq. (1.1) takes the form

otðDwÞ þ ðr?wÞ � rðDwÞ ¼ mD2w; in X: ð2:2Þ

Observe that the velocity field u is divergence-free due to (2.1). Furthermore, the boundary condition (1.4)now reads

rwðx; y; tÞ ¼ 0 for ðx; yÞ 2 oX; t P 0: ð2:3Þ

Since w is clearly only determined up to an additive constant, we can rewrite (2.3) as

wðx; y; tÞ ¼ owon

ðx; y; tÞ ¼ 0; ðx; yÞ 2 oX; t P 0; ð2:4Þ

Page 4: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 643

where oon is the outward normal derivative. Finally, the initial data (1.3) is now written in terms of w,

w0ðx; yÞ ¼ wðx; y; tÞjt¼0; ðx; yÞ 2 X: ð2:5Þ

For functions w which are sufficiently regular the boundary condition (2.4) is equivalent to

wðx; y; tÞ 2 H 20ðXÞ for any fixed t P 0; ð2:6Þ

where H2 is the Sobolev space of order 2, equipped with the norm

kwð�; �; tÞk2H2 ¼ZXw2 dx dy þ

ZXðDwÞ2 dx dy ð2:7Þ

(see [16]). The closed subspace H 20ðXÞ � H 2ðXÞ is defined as the closure of the subspace C1

0 of smooth com-

pactly-supported functions, with respect to the H2-norm.

The Sobolev space H4 is defined in exactly the same fashion, adding the integral of (D2w)2 to the right-

hand-side of (2.7). As is well-known, the operator D2 is a positive (self-adjoint) operator in H4 [16], whosedomain is H 4 \ H 2

0. It therefore gives rise to a contraction (analytic) semigroup which solves (uniquely) the

linear equation

otðDHÞ ¼ mD2H; Hð�; �; tÞ 2 H 20ðXÞ: ð2:8Þ

Observe the presence of DH in the left-hand-side of (2.8). It makes the equation more subtle than a simple

generalization of the heat equation. For example, a ‘‘formal division’’ by D might lead one to conclude that

the ‘‘spatial order’’ of the equation is two, hence (in analogy with the heat equation) only one boundary

condition is needed (i.e., H 2 H 10). This is in fact not the case, and a double condition (i.e., H 2

0, as in

(2.4) is needed. Even the definition of ‘‘eigenfunctions’’ for (2.8) (and their completeness) is not quite clear.

We refer to [12] for the one dimensional case (where X ˝ R is an interval).Comparing Eqs. (2.2) and (2.8) we see that the convective nonlinear term in (2.2) adds yet another

difficulty to the mathematical study of the equation. Furthermore, an important objective of this study

is the extension of the theory to ‘‘rough’’ initial data, namely, letting n0 = Dw0 be a singular function.

This is not only a ‘‘pure mathematical interest’’ but, on the contrary, represents the common physical

(and numerical) models of point vortices or vortex filaments. We refer the reader to [2] for a full treat-

ment of the mathematical aspects. We emphasize that our ‘‘pure streamfunction’’ approach in this paper

is very closely linked to the theoretical treatment. In what follows we indicate how the questions of

uniqueness and asymptotic decay are handled in this framework. The results of the two theorems arecertainly not new, but their proofs in terms of the streamfunction shed light on the usefulness of this

formulation. Furthermore, they are very close to the proofs in the discrete case. In particular, the esti-

mates used in the proof of Theorem 2.1 are analogous to the stability proof for the convergence of the

discrete scheme in Theorem 4.1.

Theorem 2.1 (Uniqueness). Let w; ~w 2 H 20ðXÞ be solutions of (2.2)–(2.4) having the same initial data. Let

u ¼ r?w; v ¼ r?~w be the corresponding velocity fields, and n ¼ Dw; g ¼ D~w the corresponding vorticities.

Then w � ~w.

Proof We consider Eq. (1.1) and the corresponding one for g, v. Taking their difference and multiplying by

w� ~w we get,

ZXðw� ~wÞotðn� gÞdx�

ZXnððu� vÞ � rÞðw� ~wÞdx�

ZXðn� gÞðv � rÞðw� ~wÞdx ¼ m

ZXðn� gÞ2 dx:

ð2:9Þ

Page 5: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

644 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

But clearly ðu� vÞ � rðw� ~wÞ � 0. The first term in the LHS of (2.9) can be rewritten as

ZXðw� ~wÞotðn� gÞdx ¼ � 1

2

d

dt

ZXjrðw� ~wÞj2 dx: ð2:10Þ

As for the third integral in the LHS of (2.9), we use Holder�s inequality, an interpolation inequality for the

L4 norm (see [21, Section 3.3.3]) and standard elliptic estimates to obtain

ZXðn� gÞðv � rÞðw� ~wÞdx

���� ���� 6 kn� gkL2ðXÞkvkL4ðXÞkrðw� ~wÞkL4ðXÞ

6 Ckn� gkL2ðXÞkvk1=2

L2ðXÞkrvk1=2L2ðXÞkrðw� ~wÞk1=2

L2ðXÞkDðw� ~wÞk1=2L2ðXÞ ð2:11Þ

(where C is a ‘‘generic’’ constant depending only on X). We now note that, by definition,

kvkL2ðXÞ 6 Ck~wkH10ðXÞ 6 C½k~wkL2ðXÞ þ kD~wkL2ðXÞ�;

krvkL2ðXÞ 6 C½k~wkL2ðXÞ þ kD~wkL2ðXÞ�;

hence (2.11) can be rewritten as,

ZXðn� gÞðv � rÞðw� ~wÞdx

���� ���� 6 C kn� gk3=2L2ðXÞkrðw� ~wÞk1=2

L2ðXÞ½k~wkL2ðXÞ þ kD~wkL2ðXÞ�n o

: ð2:12Þ

The RHS in (2.12) can be further estimated by

kn� gk3=2L2ðXÞkrðw� ~wÞk1=2

L2ðXÞ½k~wkL2ðXÞ þ kD~wkL2ðXÞ�

6 �kn� gk2L2ðXÞ þ64

81�3½k~wkL2ðXÞ þ kD~wkL2ðXÞ�

4krðw� ~wÞk2L2ðXÞ: ð2:13Þ

We take � ¼ m2C in this estimate and insert it in (2.9). In conjunction with (2.10), (2.12) we get,

d

dtkrðw� ~wÞk2L2ðXÞ 6 Ckrðw� ~wÞk2L2ðXÞ; ð2:14Þ

where C > 0 depends on ~w (in addition to m, X) but not on w. Since wðx; 0Þ ¼ ~wðx; 0Þ, the Gronwall inequal-

ity yields w � ~w. h

Turning to the asymptotic behavior of the solution to (2.2), we have the following.

Theorem 2.2 (Decay of solutions). Let w(x,t) be a solution to (2.2). Then there exists a positive constant k,

depending only on X, such that

kjrwðx; tÞjkL2ðXÞ 6 e�mktkjrwðx; 0ÞjkL2ðXÞ: ð2:15Þ

Proof Let us first note that

ZX

oDwot

w dx ¼ 1

2

d

dt

ZXwDw dx: ð2:16Þ

We may rewrite the integral in RHS of (2.16) as follows.

ZXwDw dx ¼

ZXwr � ðrwÞ ¼ �

ZXjrwj2 dx: ð2:17Þ

Combining Eqs. (2.16), (2.17) and (2.2), we find that

Page 6: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 645

� 1

2

o

ot

ZXjrwj2 dx ¼ �

ZXwðr?w � rÞDw dxþ m

ZXwD2w dx: ð2:18Þ

Using (2.1), the first term in the RHS of (2.18) may be rewritten as follows.

ZXwðu � rÞDw dx ¼ �

ZXðrw � uÞDw dx ¼ 0; ð2:19Þ

since $w Æ u ” 0.

Now, we treat the second term in the RHS of (2.18). Since D is a self-adjoint operator on H 20ðXÞ,

Z

XwD2w ¼

ZXðDwÞ2 dx: ð2:20Þ

Applying (2.19) and (2.20) to (2.18), we find that

1

2

o

ot

ZXjrwj2 dx ¼ �m

ZXðDwÞ2 dx:

Using the Poincare inequality

ZXðDwÞ2 dx P k

ZXjrwj2 dx;

where k is a positive constant depending on X, we conclude that

kjrwðx; tÞjkL2ðXÞ 6 e�mktkjrwðx; 0ÞjkL2ðXÞ: �

3. The numerical scheme

To simplify the exposition, assume that X is a rectangle [a,b] · [c,d]. We lay out a uniform grida = x0 < x1 < � � � < xN = b, c = y0 < y1 < � � � < yM = d. Assume that Dx = Dy = h. At each grid point (xi,yj)

we have three unknowns wij, pij, qij, where p = wx and q = wy.

The time discretization is obtained by a Crank–Nicolson scheme, which approximates (2.2). The latter is

applied at interior points 1 6 i 6 N � 1, 1 6 j 6M � 1. On the boundary i = 0,N or j = 0,M

w,p = wx,q = wy are determined by the boundary conditions (2.4). In order to do that we have to give

discrete expressions for the spatial operators which appear in (2.2).

3.1. Spatial discretization

3.1.1. The viscous term

Our scheme is based on Stephenson�s [20] scheme for the biharmonic equation

D2w ¼ f :

Later on, Altas et al. [1] and Kupferman [12] applied Stephenson�s scheme, using a multigrid solver. Ste-

phenson�s compact approximation for the biharmonic operator is the following:

ðD2hÞ

cwi;j ¼ 1h4f56wi;j � 16ðwiþ1;j þ wi;jþ1 þ wi�1;j þ wi;j�1Þ

þ2ðwiþ1;jþ1 þ wi�1;jþ1 þ wi�1;j�1 þ wiþ1;j�1Þþ6h½ðwxÞiþ1;j � ðwxÞi�1;j þ ðwyÞi;jþ1 � ðwyÞi;j�1�g

¼ fi;j:

8>>><>>>: ð3:1Þ

Page 7: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

646 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

Here, ðD2hÞ

cwi;j is the compact second-order approximation for D2w specified in caption of Fig. 1. We have

also to relate wx and wy to w. This is done via the following fourth-order compact schemes.

Fig. 1

polyno

hðwxÞi;j ¼3

4ðwiþ1;j � wi�1;jÞ �

h4

ðwxÞiþ1;j þ ðwxÞi�1;j

h i; ð3:2Þ

hðwyÞi;j ¼3

4ðwi;jþ1 � wi;j�1Þ �

h4

ðwyÞi;jþ1 þ ðwyÞi;j�1

h i: ð3:3Þ

Eqs. (3.1)–(3.3) form a second order compact scheme for D2w, involving values of w, wx and wy at (i,j) and

at its eight nearest neighbors (see Fig. 1). Thus, the scheme is compact. The approximation above is applied

at any interior point 1 6 i 6 N � 1, 1 6 j 6 M � 1. On the boundary i = 0,N or j = 0,M, w, wx, wy are deter-

mined from the boundary conditions (2.4).

3.1.2. The Laplacian of a discrete function

For any function g we define the discrete approximation to Dg by Dhg, where Dhg is

Dhg ¼ d2xg þ d2yg ð3:4Þ

and

d2xgi;j ¼giþ1;j � 2gi;j þ gi�1;j

h2; d2ygi;j ¼

gi;jþ1 � 2gi;j þ gi;j�1

h2:

3.1.3. The convective term

The convective term ($^w) Æ $(Dw) is approximated as follows:

ðr?wÞi;j ¼ �ðwyÞi;j; ðwxÞi;jh i

: ð3:5Þ

. Stephenson�s scheme for D2w = f: The finite difference operator ðD2hÞ

cw at point (i,j) is D2P(xi,yj) where w = P(x,y) is a

mial in P3 + Span(x4,x2y2,y4) defined by the 13 collocated values for w, wx, wy displayed.

Page 8: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 647

No further approximation is needed, since wx and wy are part of the unknowns in our discretization. Now,

rðDwÞi;j ¼ ððDwxÞi;j; ðDwyÞi;jÞ ¼ ððDhwxÞi;j; ðDhwyÞi;jÞ þOðh2; h2Þ: ð3:6Þ

Note that the above discretization is well defined for any interior point 1 6 i 6 N � 1, 1 6 j 6M � 1. The

resulting scheme has the following form.

3.2. The scheme

Combining (3.1) and (3.6) and the time discretization, we obtain the following scheme:

ðDhwi;jÞnþ1=2 � ðDhwi; jÞn

Dt=2¼ �½�ðwn

yÞi;j; ðwnxÞi;j� � ½ðDhw

nxÞi;j; ðDhw

nyÞi;j� þ

m2

ðD2hÞ

cwnþ1=2i;j þ ðD2

hÞcwn

i;j

h i; ð3:7Þ

ðDhwi;jÞnþ1�ðDhwi;jÞn

Dt¼�½�ðwnþ1=2

y �i;j; ½wnþ1=2x Þi;j� � ½ðDhw

nþ1=2x Þi;j;ðDhw

nþ1=2y Þi;j�þ

m2½ðD2

hÞcwnþ1

i;j þðD2hÞ

cwni;j�;

ð3:8Þ

where ðD2hÞ

cis defined in (3.1).

Remark that we apply the scheme above to all interior point, and on boundary points we impose the

boundary conditions by determining w, wx and wy from (2.4).

4. Stability and convergence in two dimensions

4.1. Stability of the predictor-corrector scheme in two dimensions

We consider the predictor-corrector scheme (3.7) and (3.8) applied to the linear model equation

Dwt ¼ aDwx þ bDwy þ mD2w: ð4:1Þ

This scheme reads

Dhwnþ1=2�Dhw

n

Dt=2 ¼ aDhwnx þ bDhw

ny þ m

2ðD2

hwn þ D2

hwnþ1=2Þ; ðaÞ

Dhwnþ1�Dhw

n

Dt ¼ aDhwnþ1=2x þ bDhw

nþ1=2y þ m

2ðD2

hwnþ1 þ D2

hwnÞ; ðbÞ

8<: ð4:2Þ

where Dhwn, D2

hwn, wn

x and wny are defined in (3.4) and (3.1)–(3.3), respectively. Denote

g ¼ Dth; l ¼ mDt

h2: ð4:3Þ

We have

Proposition 4.1 The difference scheme (4.2) is stable in the Von Neumann sense under the sufficient condition

maxðjaj; jbjÞg 6 min

ffiffiffi8

p

3

ffiffiffil

p;

ffiffiffi2

p

3

!: ð4:4Þ

Proof Let h = ah 2 [0,2p[, / = bh 2 [0,2p[ and wnjk ¼ cwnða; bÞeijheik/. We denote by g1(h,/) the amplification

factor of the predictor step (4.2)a, g2(h,/) the amplification factor after the two steps (4.2). The factor

g1(h,/) is

Page 9: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

648 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

g1ðh;/Þ ¼A1ðh;/Þ � B1ðh;/Þ þ iC1ðh;/Þ

A1ðh;/Þ þ B1ðh;/Þ; ð4:5Þ

with

A1ðh;/Þ ¼2� 2 cos h

h2þ 2� 2 cos/

h2;

B1ðh;/Þ ¼l4

62� 2 cos h

h21� cos h2þ cos h

þ 62� 2 cos/

h21� cos/2þ cos/

þ 22� 2 cos h

h2� 2 cos/

h

� �;

C1ðh;/Þ ¼g2

2� 2 cos h

h2þ 2� 2 cos/

h2

� �3a sinðhÞ2þ cos h

þ 3b sinð/Þ2þ cos/

� �:

The factor g2(h,/) is

g2ðh;/Þ ¼A1ðh;/Þ � 2B1ðh;/Þ þ 2iC1ðh;/Þg1ðh;/Þ

A1ðh;/Þ þ 2B1ðh;/Þ: ð4:6Þ

The stability condition suph,/jg2(h,/)j 6 1 is equivalent for each h,/ 2 [0,2p[ to

�4C1ðh;/ÞImðg1ðh;/ÞÞ½A1ðh;/Þ � 2B1ðh;/Þ� þ 4C21ðh;/Þjg1ðh;/Þj

26 8A1ðh;/ÞB1ðh;/Þ: ð4:7Þ

We restrict ourselves to the case where

suph;/2½0;2p½

jg1ðh;/Þj 6 1: ð4:8Þ

A sufficient condition for (4.8) to be satisfied is

maxðjaj; jbjÞg 6

ffiffiffi8

p

3

ffiffiffil

p:

Then (4.7) is satisfied under the sufficient condition

C21 1� A1 � 2B1

A1 þ B1

� �6 2A1B1:

The latter is equivalent to

1

4g2

3a sinðhÞ2þ cos h

þ 3b sinð/Þ2þ cos/

� �2� 3

A1

A1 þ B1

6 2: ð4:9Þ

A sufficient condition for (4.9) is

maxða2; b2Þg2 6 2

9;

which completes the proof. h

Observe that in the nonconvective case, a = b = 0, the scheme is unconditionally stable, as could

be expected. Thus, the presence of lower-order convective terms makes it necessary to limit the

timestep.

Remark Note that the restriction of the CFL number (4.4) by a formula of the type

maxðjaj; jbjÞg 6 Cffiffiffil

p ð4:10Þ

Page 10: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 649

pertains to a centered scheme for the convection-diffusion equation, with an implicit discretization of the

diffusive term and an explicit discretization of the convective term, even in the one-dimensional

situation.

4.2. Convergence of the spatially semi-discrete two-dimensional scheme

In the next theorem, we prove a rate of convergence of h2 for the time continuous version of scheme (3.7)

and (3.8), when applied to the linear Eq. (4.1) on [0,1] · [0,1] in the H 20 setting.

Define h = 1/N, xi = yi = ih, 0 6 i 6 N. We call L2h;0 the space of N · N arrays in x and y directions, for

which u0,j = uN,j = ui,0 = ui,N = 0. For u; v 2 L20;h, the scalar product is

ðu; vÞh ¼XN�1

i¼1

XN�1

j¼1

uijvijh2

and the norm

jujh ¼XN�1

i¼1

XN�1

j¼1

juijj2h2 !1=2

:

For ~w 2 L2h;0, the spatial discrete operators Dh

~w, D2h~w, ~wx and ~wy are defined in (3.4) and (3.1)–(3.3),

respectively.

In the investigation of the convergence properties of our scheme we use the exact Eq. (2.2) and a semi-

discrete analog of (3.7) and (3.8). Thus the discrete solution is represented by the grid functions~wi;jðtÞ; ~wx;i;jðtÞ; ~wy;i;jðtÞ which approximate the exact solution w(x,y,t), wx(x,y,t), wy(x,y,t) at (x,y) =

(ih, jh). We have ~w, ~wx,~wy 2 L2

h;0, so that the boundary values of ~w, ~wx,~wy vanish. Thus, the equation sat-

isfied by the discrete functions is

o

otDh

~w ¼ aDhð~wxÞ þ bDhð~wyÞ þ mD2hð~wÞ ð4:11Þ

subject to initial conditions

~wi;j ¼ w0ðih; jhÞ; ð~wx;~wyÞi;j ¼ w0;xðih; jhÞ;w0;yðih; jhÞ

� �: ð4:12Þ

For every discrete u 2 L2h;0 we set as usual

ðdþx uÞi;j ¼uiþ1;j � ui;j

h; ðdþy uÞi;j ¼

ui;jþ1 � ui;jh

ð4:13Þ

and

ðdxuÞi;j ¼uiþ1;j � ui�1;j

2h; ðdyuÞi;j ¼

ui;jþ1 � ui;j�1

2h: ð4:14Þ

Our convergence result is the following.

Theorem 4.1 Let the discrete solution ~w; ~wx;~wy satisfy (4.11), and the boundary conditions

~wi;j ¼ ~wx;i;j ¼ ~wy;i;j ¼ 0 for i 2 f0;Ng or j 2 f0;Ng:

Let w be the exact solution of (2.2) subject to the boundary conditions (2.4). Fix s > 0. Then, for 0 6 t 6 s,

there exists a constant C > 0, depending only on s and the initial data, such that

jejh þ jdþx ejh þ jdþy ejh 6 Ch2; ð4:15Þ

Page 11: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

650 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

where e ¼ ~w� w is the difference between the approximate and the exact solutions on the grid, the latter being

supposed sufficiently regular.

Proof The exact solution w satisfies

o

otDhw ¼ aDhðwxÞ þ bDhðwyÞ þ mD2

hðwÞ � T ; ð4:16Þ

where T = O(h2) is the truncation error. Observe that in (4.16), the values (wx)ij, (wy)ij are not the com-

ponents of the gradient of the given smooth solution w (at (i,j)) but are the values obtained from the

discrete values wij by use of (3.2) and (3.3). Subtracting (4.16) from (4.11) and denoting the errore ¼ ~w� w, we have

o

otDhe ¼ aDhðexÞ þ bDhðeyÞ þ mD2

heþ T : ð4:17Þ

The viscous term given in (3.1) is

mD2heij ¼

m

h4

n56eij � 16ðeiþ1;j þ ei�1;jÞ � 16ðei;jþ1 þ ei;j�1Þ þ 2ðeiþ1;jþ1 þ ei�1;jþ1Þ

þ2ðeiþ1;j�1 þ ei�1;j�1Þ þ 6h½ðexÞiþ1;j � ðexÞi�1;j þ ðeyÞi;jþ1 � ðeyÞi;j�1�o;

which may be rewritten as

mD2heij ¼

m

h4

n�8eij � 16ðeiþ1;j � 2eij þ ei�1;jÞ � 16ðei;jþ1 � 2eij þ ei;j�1Þ þ 2ðeiþ1;jþ1 � 2ei;jþ1 þ ei�1;jþ1Þ

þ2ðeiþ1;j�1 � 2ei;j�1 þ ei�1;j�1Þ þ 4ðei;jþ1 þ ei;j�1Þ þ 6h½ðexÞiþ1;j � ðexÞi�1;j þ ðeyÞi;jþ1 � ðeyÞi;j�1�o:

The latter may be simplified to

mD2heij ¼

m

h4�12h2d2xeij � 12h2d2yeij þ 2h4d2xd

2yeij þþ6h½ðexÞiþ1;j � ðexÞi�1;j þ ðeyÞi;jþ1 � ðeyÞi;j�1�

n o;

ð4:18Þ

or

mD2hei;j ¼ m d4xei;j þ d4yei;j þ 2d2xd

2yei;j

h i; ð4:19Þ

where

d4xui;j ¼12

h2dxðuxÞi;j � d2xui;jh i

; d4yui;j ¼12

h2dyðuyÞi;j � d2yui;jh i

; d2xd2yui;j ¼ ðdxyÞ2ui;j: ð4:20Þ

The rows (and the columns) of elements in L2h;0 are (N + 1)-vectors h = (h0, h1, . . ., hN) such that h0 = hN = 0.

We denote by l2h;0 the ((N � 1)-dimensional) space of such vectors. It will be convenient to refer to the

(N � 1)-dimensional part (h1, . . ., hN� 1) of vectors in l2h;0 and the operators acting on it (with the under-

standing that h0 = hN = 0). The scalar product and the norm in l2h;0 are

ðh; ~hÞh ¼ hXN�1

i¼1

hi ~hi; jhj2h ¼ hXN�1

i¼1

h2i ; ð4:21Þ

which is in agreement with the notation for L2h;0 (in the case of arrays).

The proof consists now of applying the energy method to (4.17). In order to do that, we need the

discrete analog of the following L2(X) scalar products denoted by (Æ,Æ), where it is assumed that

w 2 H 20ðXÞ.

Page 12: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 651

ðiÞ ðDw;wÞ ¼ �jrwj2;

ðiiÞ aDowox

þ bDowoy

;w

� �¼ 0;

ðiiiÞ ðD2w;wÞ ¼ jDwj2:

ð4:22Þ

The main difficulty is that the discrete gradient (ex, ey) in (4.17) is defined implicitly by (3.2) and (3.3), so

that a classical discrete integration by parts like (dxe,e)h = 0 for e 2 l2h;0 no longer holds. Also the bihar-

monic operator applied to e as defined by (3.1) involves the values of ex. These values have to be related

to the standard difference operators like d�x e in order to get an equivalent form to (4.22(iii)).

Let us introduce P be as the finite difference operator acting in l2h;0 by

ðPhÞi ¼ hi�1 þ 4hi þ hiþ1; 1 6 i 6 N � 1; h 2 l2h;0: ð4:23Þ

The operator P is positive symmetric (diagonally dominated), so that by the Cauchy–Schwartz inequality,

2jhj2h 6 ðPh; hÞh 6 6jhj2h; h 2 l2h;0: ð4:24Þ

Note that by (3.2), (ex)i,j is defined by

ðPexÞi;j ¼3

hðeiþ1;j � ei�1;jÞ; 1 6 i; j 6 N � 1: ð4:25Þ

In the sequel, we handle any grid function ui;j 2 L2h;0 as well as finite difference operators acting on them, as

(N � 1) · (N � 1) matrices. Denoting ejx ¼ ððexÞ1;j; . . . ; ðexÞN�1;jÞT, ej = (e1,j,. . .,eN� 1,j)

T, 1 6 j 6 N � 1 the

jth columns of the matrices ex, e, we can rewrite (4.25) as

Pejx ¼ 6dxej; 1 6 j 6 N � 1; ð4:26Þ

or simply in matrix form

Pex ¼ 6dxe; ð4:27Þ

and similarly

eyP ¼ 6dye: ð4:28Þ

Note that in (4.27) and (4.28), we refer to P as the symmetric positive definite matrix

P i;m ¼4; m ¼ i;

1; jm� ij ¼ 1;

0; jm� ij P 2:

8><>: ð4:29Þ

In addition, due to (dx u,u)h = 0 for u 2 l2h;0, we have

ðPejx; ejÞh ¼ 0; j ¼ 1; . . . ;N � 1: ð4:30Þ

Note also that multiplication on the left of a matrix A by P results in replacing its ith row Ai by

4Ai + Ai+1 + Ai� 1. Multiplication on the right has the same effect on the columns.

The matrix representing d2x is (see (3.4)) h�2(P � 6I), as multiplication on the left (of the matrix g), while

d2y is expressed by the same multiplication on the right.

Taking the scalar product of (4.17) with e yields

ðddtDhe; eÞh ¼ mðD2

he; eÞh ðIÞþaðDhex; eÞh þ bðDhey ; eÞh ðIIÞþðT ; eÞh ðIIIÞ

8><>: ð4:31Þ

(I), (II), (III) are respectively the diffusive, convective, and truncation terms.

Page 13: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

652 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

We first consider the diffusive term (I). The crucial step in the proof of the theorem is the derivation of a

suitable lower bound for ðD2he; eÞh, for e; ex; ey 2 L2h;0ðXÞ. In the continuous case, if / 2 H 2

0ðXÞ, an

integration by parts yields

ðD2/;/ÞL2ðXÞ ¼ kD/k2L2ðXÞ: ð4:32Þ

Our discrete analog is given by the following claim.

Claim: There exists a constant C P 0 independent of h, such that, for all grid functions u 2 L2h;0

ðD2hu; uÞh P C jdþx uxj

2h þ jdþy uy j

2h þ jdþy uxj

2h þ jdþx uy j

2h

h i; ð4:33Þ

where ux, uy are related to u as in (4.27) and (4.28).

Proof of the claim: Let us first observe that for all u; v 2 l2h;0

ðdxu; vÞh ¼ ðdþx u;PvÞh; ð4:34Þ

where P : l2h;0 ! l2h;0 is the averaging operator defined by

ðPvÞi ¼1

2ðvi þ viþ1Þ; 1 6 i 6 N � 1: ð4:35Þ

Indeed, to prove (4.34), we note that

ðdxu; vÞh ¼1

2ðdþx þ d�x Þu; v

h: ð4:36Þ

But since dþx ¼ d�x S, where (Sv)k = vk+1, 1 6 k 6 N�1 is the forward shift in the x direction, we have

ðdxu; vÞh ¼ �ðu; dxvÞh ¼ � 1

2d�x ðI þ SÞv; u

� �h

¼ 1

2ðI þ SÞv; dþx u

� �h

¼ ðPv; dþx uÞh: ð4:37Þ

For any u 2 L2h;0, we have now

ðD2hu; uÞh ¼ ðd4xu; uÞh þ ðd4yu; uÞh þ 2ðd2xd

2yu; uÞh

¼ 12

h2dxux � d2xu; u

hþ 12

h2dyuy � d2yu; u� �

hþ 2 d2xd

2yu; u

� �h:

Next we check that for any v 2 l2h;0

12

h2dxvx � d2xv; v

hP Cjdþx vxj

2h: ð4:38Þ

Noting (4.34) and �ðd2xv;wÞh ¼ ðdþx v; dþx wÞh for all v;w 2 l2h;0, we get

ðdxvx � d2xv; vÞh ¼ �ðdþx v;PvxÞh þ ðdþx v; dþx vÞh

¼ ðdþx v; dþx v�PvxÞh ¼ ðdþx v�Pvx; d

þx v�PvxÞh þ ðPvx; d

þx v�PvxÞh

P ðPvx; dþx v�PvxÞh:

Recall that P ¼ 6I þ h2d2x . Then, by (4.27)

vx; zþh2

6d2xz

� �h

¼ 1

6ðvx; PzÞh ¼

1

6ðPvx; zÞh ¼ ðdxv; zÞh ¼ ðdþx v;PzÞh; z 2 l2h;0: ð4:39Þ

Setting z = vx in (4.39), we have

Page 14: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 653

ðPvx; dþx v�PvxÞh ¼ vx; vx þ

h2

6d2xvx

� �h

� jPvxj2h ¼ jvxj2h �h2

6jdþx vxj

2

h � jPvxj2h: ð4:40Þ

Using finally that for all w 2 l2h;0

jwj2h � jPwj2h ¼h2

4jdþx wj

2h; ð4:41Þ

we deduce from (4.40) that

ðdxvx � d2xv; vÞh P ðPvx; dþx v�PvxÞh ¼

h2

4� h2

6

� �jdþx vxj

2h ¼

h2

12jdþx vxj

2h; ð4:42Þ

which is the desired result (4.38). Clearly, the same result holds for a bidimensional grid function v 2 L2h;0

(summation over all columns of the matrix v).Consider now the mixed term ðd2xd

2yu; uÞh ¼ jdþx d

þy uj

2h. We assert that

jdþx dþy ujh P

1

6jdþx uy jh: ð4:43Þ

To prove (4.43), we write first

dþx dþy ui;j ¼

dþy uiþ1;j � dþy ui;jh

: ð4:44Þ

Using dþy ui;j ¼ dyui;j þ h2d2yui;j and (4.28), we deduce

dþx dþy ui;j ¼

dyuiþ1;j � dyui;jh

þ 1

2d2yuiþ1;j � d2yui;jh i

¼ 1

hðuyÞiþ1;j � ðuyÞi;jh i

þ h6

d2yðuyÞiþ1;j � d2yðuyÞi;jh i

þ 1

2d2yuiþ1;j � d2yui;jh i

¼ dþx ðuyÞi;j þh2

6d2yd

þx ðuyÞi;j þ

h2d2yd

þx ui;j: ð4:45Þ

In addition, using the definition of d2y we have

jd2ydþx uy jh 6

4

h2jdþx uy jh; ð4:46Þ

and using d2y ¼ d�y dþy , we have again by definition

jd2ydþx ujh 6

2

hjdþy d

þx ujh: ð4:47Þ

Therefore, we deduce from (4.45)

jdþx dþy ujh P jdþx uy jh �

h2

6jd2yd

þx uy jh �

h2jd2yd

þx ujh P jdþx uy jh �

2

3jdþx uy jh � jdþx d

þy ujh:

This gives finally 2jdþx dþy ujh P 1

3jdþx uy jh which is (4.43). We proceed in the same way in proving the symmet-

ric estimate

jdþx dþy ujh P

1

6jdþy uxjh: ð4:48Þ

This concludes the proof of the claim (4.33). h

Page 15: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

654 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

The convective term (II) = a(Dhex,e)h + b(Dhey,e)h in (4.31) is

ðIIÞ ¼ aðd2xðexÞ; eÞh þ aðd2yðexÞ; eÞh þ bðd2xðeyÞ; eÞh þ bðd2yðeyÞ; eÞh: ð4:49Þ

Since we do not have a strict discrete equivalent of (4.22(ii)) we proceed as follows.The first term in the right-hand-side of (4.49) is

aðd2xex; eÞh ¼ �aðdþx ex; dþx eÞh; ð4:50Þ

so that

jaðd2xex; eÞhj 6 jaj ejdþx exj2

h þ1

4ejdþx ej

2

h

� �; ð4:51Þ

where e > 0 will be selected latter. Proceeding in the same way for the three other terms in (4.49), we find the

estimate of the convective term

jaðDhex;eÞh þ bðDhey ;eÞhj6maxðjaj; jbjÞ efjdþx exj2h þ jdþy ey j

2h þ jdþx ey j

2h þ jdþy exj

2hgþ

1

2efjdþx ej

2h þ jdþy ej

2hg

� �:

Finally the truncation term (III) in (4.31) is estimated by

jðIIIÞj ¼ jðT ; eÞhj 61

2jT j2h þ

1

2jej2h 6 C0 h4 þ jdþx ej

2

h þ jdþy ej2

h

h i: ð4:52Þ

Collecting now the terms (I), (II), (III) in (4.31), we find

1

2

d

dtjdþx ej

2h þ jdþy ej

2h

h i¼ � d

dtDhe; e

� �h

6 emaxðjaj; jbjÞ � Cm½ � jdþx exj2h þ jdþy ey j

2h þ jdþx ey j

2h þ jdþy exj

2h

h iþ 1

2emaxðjaj; jbjÞ þ C0

� �jdþx ej

2h þ jdþy ej

2h

h iþ C0h4: ð4:53Þ

Selecting now e sufficiently small in order that [e max(jaj;jbj)�Cm] 6 0, we obtain

d

dtjdþx ej

2h þ jdþy ej

2h

h i6 C00ðm; jaj; jbjÞ jdþx ej

2h þ jdþy ej

2h þ h4

h i; ð4:54Þ

which gives the conclusion of the theorem for jdþx ejh þ jdþx ejh by the Gronwall inequality. The estimate for

jejh now follows from the discrete Poincare inequality. h

Remark Note that the error estimate (4.15) is directly linked to the truncation error T = O(h2).

5. Numerical results

We present here several problems, towards which we applied our scheme. In the first test problem we

take an exact solution to (2.2) (see[5]).

wðx; y; tÞ ¼ �0:5e�2mt sin x sin y; 0 6 x; y 6 p: ð5:1Þ

We have picked m = 1. Table 1 summarizes the error, e, and the relative error, er, where

el2 ¼ kwcomp � wexactkl2 ;

Page 16: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Table 1

Streamfunction formulation: compact scheme for problem (1) – l2 error, relative l2 error and max error in the horizontal component of

the velocity field

Mesh 17 · 17 Rate 33 · 33 Rate 65 · 65

t = 1 el2 2.437 · 10�5 4.05 1.394 · 10�6 3.93 9.114 · 10�8

er 2.134 · 10�4 1.349 · 10�5 8.371 · 10�7

eu 2.797 · 10�5 4.00 1.749 · 10�6 4.00 1.093 · 10�7

t = 2 el2 3.180 · 10�6 3.78 2.322 · 10�7 4.07 1.319 · 10�8

er 2.232 · 10�4 1.334 · 10�5 8.736 · 10�7

eu 4.254 · 10�6 4.00 2.663 · 10�7 4.00 1.665 · 10�8

t = 3 el2 4.289 · 10�7 3.97 2.738 · 10�8 3.90 1.831 · 10�9

er 2.235 · 10�4 1.400 · 10�5 8.750 · 10�7

eu 6.199 · 10�7 4.00 3.882 · 10�8 4.00 1.677 · 10�9

t = 4 el2 7.864 · 10�8 4.00 4.925 · 10�9 4.27 1.831 · 10�9

er 2.224 · 10�4 1.400 · 10�5 8.750 · 10�7

eu 9.019 · 10�8 4.00 5.648 · 10�9 4.00 3.530 · 10�10

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 655

er ¼ e=kwexactkl2

and

eu ¼ max jucomp � uexactj:

Here, wcomp, ucomp and wexact, uexact are the computed and the exact streamfunction and x-component

of the velocity field, respectively. We represent results for different time-levels and number of meshpoints.

Similar results are shown for the solution w = e�t(x2 + y2)2 of the non-homogeneous problem

oðDwÞot

þ ðr?wÞ � rðDwÞ ¼ D2w� 16e�tðx2 þ y2 þ 4Þ

for 0 6 x,y 6 1. Table 2 displays the error, e, and the relative error, er, and eu (the same quantities as in

Table 1). We represent results for different time-levels and number of mesh points, and see that the conver-gence rate is around 2.

We turn now to the class of driven cavity problems, which has been used for benchmark test prob-

lems by many authors. In particular, we compare our results to the steady state results of Ghia, Ghia

and Shin [10]. First we show numerical results for m = 1/400. Here the domain is X = [0,1] · [0,1] and the

fluid is driven in the x-direction on the top section of the boundary (y = 1). Thus, u = 1, v = 0 for y = 1,

and u = v = 0 for x = 0, x = 1 and y = 0. In Table 3 we present computational quantities for different

meshes and time-levels. We show wX,t = maxXw(x,y,t), ð�x; �yÞ, where ð�x; �yÞ is the point where wX,t occurs,

and wX,t = minXw(x,y,t). The meshes are of 65 · 65, 81 · 81 and 97 · 97 points and the time levels aret = 10, 20, 40, 60. Note that the highest value of the streamfunction at the latest time step is 0.1136.

Here the maximum occurs at ð�x; �yÞ ¼ ð0:5521; 0:6042Þ, and the minimal value of the streamfunction is

�6.498 · 10�4. Note that wX,t in Table 3 has been stabilized at t = 40. The location ð�x; �yÞ of the primary

vortex remains constant from t = 20 and on. In [10] wX,t = 0.1139 occurs at (0.5547,0.6055), and the

minimal value of the streamfunction is �6.424 · 10�4. Fig. 2(a) displays streamfunction contours at

t = 60, using a 97 · 97 mesh. In Fig. 4(a), we present velocity components u(0.5,y) and v(x,0.5) (solid

lines) at T = 60 compared with values obtained in [10] (marked by �0�), for m = 1/400. Note that the

match between the results is excellent.

Page 17: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Table 2

Streamfunction formulation: compact scheme for w = e�t(x2 + y2)2 with RHS f = �16e�t(x2 + y2 + 4) on [0,1] · [0,1] – l2 error, relative

l2 error and max error in the horizontal component of velocity field

Mesh 17 · 17 Rate 33 · 33 Rate 65 · 65

t = 0.25 el2 6.202 · 10�5 1.986 1.564 · 10�5 2.003 3.903 · 10�6

er 8.176 · 10�5 1.895 · 10�5 4.535 · 10�6

eu 2.070 · 10�4 1.986 5.224 · 10�5 2.002 1.304 · 10�5

t = 0.5 el2 7.632 · 10�5 2.000 1.908 · 10�5 2.002 4.762 · 10�6

er 1.030 · 10�4 2.368 · 10�5 5.671 · 10�6

eu 2.572 · 10�4 2.000 6.431 · 10�5 2.003 1.605 · 10�5

t = 0.75 el2 7.896 · 10�5 2.007 1.964 · 10�5 2.002 4.904 · 10�6

er 1.091 · 10�4 2.498 · 10�5 5.984 · 10�6

eu 2.667 · 10�4 2.001 6.664 · 10�5 2.008 1.657 · 10�5

t = 1 el2 7.818 · 10�5 2.007 1.945 · 10�5 2.002 4.856 · 10�6

er 1.110 · 10�4 2.535 · 10�5 6.072 · 10�6

eu 2.643 · 10�4 2.007 6.576 · 10�5 2.002 1.642 · 10�5

Table 3

Streamfunction formulation: compact scheme for the driven cavity problem, Re = 400

Time Quantity 65 · 65 81 · 81 97 · 97

10 max w 0.1053 0.1057 0.1059

ð�x;�yÞ (0.5781,0.6250) (0.5750,0.6250) (0.5833,0.6354)

min w �4.786 · 10�4 �4.758 · 10�4 �4.749 · 10�4

20 max w 0.1124 0.1128 0.1130

ð�x;�yÞ (0.5625,0.6094) (0.5625,0.6125) (0.5521,0.6042)

min w �6.333 · 10�4 �6.371 · 10�4 �6.361 · 10�4

40 max w 0.1131 0.1134 0.1136

ð�x;�yÞ (0.5625,0.6094) (0.5500,0.6000) (0.5521,0.6042)

min w �6.513 · 10�4 �6.5148 · 10�4 �6.498 · 10�4

60 max w 0.1131 0.01134 0.1136

ð�x;�yÞ (0.5625,0.6094) (0.5500,0.6000) (0.5521,0.6042)

min w �6.514 · 10�4 �6.5155 · 10�4 �6.498 · 10�4

[10]�s results: max w = 0.1139 at (0.5547,0.6055), min w = �6.424 · 10�4.

656 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

In Table 4, we show the same flow quantities as in Table 3, but for m = 1/1000 at t = 20, 40, 60, 80.

The grids are of 65 · 65, 81 · 81 and 97 · 97 points. Note that with each of the meshes the flow quan-

tities tend to converge to a steady state as time progresses. At the latest time level on the finest grid the

maximal value of w is 0.1178, compared to 0.1179 in [10]. The maximum is obtained atð�x; �yÞ ¼ ð0:5312; 0:5625Þ, compared to (0.5313,0.5625) in [10]. The minimum value of the streamfunction

is �0.0017, same as in [10]. Fig. 2(b) displays streamfunction contours at t = 80, using a 97 · 97 mesh.

In Fig. 4(b), we present velocity components u(0.5,y) and v(x,0.5) (solid lines) at T = 80 compared with

values obtained in [10] (marked by �0�), for m = 1/1000. Note again the excellent match between the

results.

Results for m = 1/3200 on a 81 · 81 mesh and a 97 · 97 mesh are shown in Table 5. At the latest time

level on the finest grid the maximal value of w is 0.1174, compared to 0.1204 in [10]. The latter is

obtained at ð�x; �yÞ ¼ ð0:5208; 0:5417Þ, compared to (0.5165,0.5469) in [10]. The minimum value of the

Page 18: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Table 4

Streamfunction formulation: compact scheme for the driven cavity problem, Re = 1000

Time Quantity 65 · 65 81 · 81 97 · 97

20 max w 0.1129 0.1139 0.1143

ð�x; �yÞ (0.5469,0.5781) (0.5375,0.5750) (0.5417,0.5729)

min w �0.0015 �0.0015 �0.0015

40 max w 0.1160 0.1169 0.1175

ð�x; �yÞ (0.5312,0.5625) (0.5250,0.5625) (0.5312,0.5625)

min w �0.0017 �0.0017 �0.0017

60 max w 0.1160 0.1171 0.1177

ð�x; �yÞ (0.5312,0.5625) (0.5250,0.5625) (0.5312,0.5625)

min w �0.0017 �0.0017 �0.0017

80 max w 0.1160 0.1172 0.1178

ð�x; �yÞ (0.5312,0.5625) (0.5250,0.5625) (0.5312,0.5625)

min w �0.0017 �0.0017 �0.0017

[10]�s results: max w = 0.1179 at (0.5313,0.5625), min w = �0.0017.

0 0.1 0.2 0.3 0.4

Re=400(a) Re=10000.5 0.6 0.7 0.8 0.9 1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1 Streamfunction contours, Re=400, T=60, mesh 97X97

0.00055

0.00045

0.000350.000250.00015

0.000155e05 5e

05

0

0

0

00.

0113

6

0.01

136

0.011360.01136

0.01136

0.01

136

0.01136 0.01136 0.01136

0.02

272

0.02

272

0.02272

0.02272

0.02272

0.02272 0.02272 0.02272

0.03

408

0.034

08

0.03408

0.03408

0.03

408

0.03408 0.03408

0.04

544

0.04544

0.04544

0.04

544

0.04544 0.04544

0.05

68

0.0568

0.0568

0.0568

0.05680.0568

0.06

816

0.06816

0.06816

0.068160.06816

0.07

952

0.07952

0.07

952

0.07952

0.09

088

0.09088

0.09

088

0.09088

0.10

224

0.10224

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction contours, T=80, Re=1000, mesh 97X97

0.0013

0.000

9

0.0005

0.0005

1e04

1e041e040

0

0

0

0

0.01

18

0.01

18

0.01180.0118

0.0118

0.01

18

0.0118 0.0118 0.0118

0.02

36

0.02

36

0.02360.02360.0236

0.02

36

0.0236 0.0236

0.03

54

0.0354

0.03540.0354

0.03

54

0.0354 0.0354

0.04

72

0.04

72

0.04720.0472

0.04

72

0.0472 0.0472

0.05

9

0.0590.059

0.05

9

0.059 0.059

0.07

08

0.0708

0.0708

0.07

08

0.0708

0.08

26

0.08260.0826

0.08260.0826

0.09

44

0.0944

0.09

44

0.0944

0.10

62

0.1062

0.1062

(b)

Fig. 2. Driven cavity for Re = 400, 1000: streamfunction contours.

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 657

streamfunction is �0.0027, where [10] reports on �0.0031. Fig. 3(a) displays streamfunction contours at

t = 360, using a 97 · 97 mesh. In Fig. 5(a), we present velocity components u(0.5,y) and v(x,0.5) (solid

lines) at T = 360 compared with values obtained in [10] (marked by �0�), for m = 1/3200. There is an

excellent match.

Finally, in Table 6 we display results for m = 1/5000. At the latest time level on the finest grid the maximal

value of w is 0.1160, compared to 0.11897 in [10]. The location of the maximal value is

ð�x; �yÞ ¼ ð0:5104; 0:5417Þ, compared to (0.5117,0.5352) in [10]. The minimum value of the streamfunction

is �0.0029, where the value �0.0031 was found in [10]. Fig. 3(b) displays streamfunction contours att = 400, with a 97 · 97 mesh. In Fig. 5(b), we present velocity components u(0.5,y) and v(x,0.5) (solid lines)

at T = 400 compared with values obtained in [10] (marked by �0�), for m = 1/5000. Note the excellent match

in this case too.

Page 19: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

0 0.1 0.2 0.3 0.4 0.5 0.6

Re=3200 Re=5000

0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=3200, T=360, mesh 97X97

0.0023

0.0019

0.0015

0.0011

0.00

11

0.0007

0.0007

0.00

07

0.00

03

0.0003

0.000

3

0.0003

0.00

03

0

0

0

0

0

0

0

0

0

0.01

16

0.01

16

0.0116

0.0116

0.0116

0.01

16

0.01160.0116 0.0116

0.02

32

0.02

32

0.02320.0232

0.0232

0.02

32

0.0232 0.0232

0.0348

0.03

48

0.0348

0.0348

0.03

48

0.0348

0.0348 0.0348

0.04

64

0.04640.0464

0.04640.0464

0.0464

0.0464

0.05

8

0.0580.058

0.05

8

0.0580.058

0.06

96

0.0696

0.0696

0.06

96

0.0696

0.0696

0.08

12

0.0812

0.0812

0.0812 0.0812

0.09

28

0.0928

0.09

28

0.0928

0.1044

0.1044

0.1044

0.116

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=5000, T=400, mesh 97X97

0.0025

0.00210.0017

0.0013

0.00

13

0.0009

0.0009 0.0009

0.0009

0.00

05

0.0005

0.00

05

0. 0005

0.00

05

1e041e04

1e04

1e04

1 e 04

1e0

4

0

0

0

0

0

0

0 0

0

0.01

16

0.01

16

0.0116

0.0116

0.0116

0.01

16

0.01160.0116 0.0116

0.02

32

0.0232

0.0232

0.0232

0.02

32

0.0232

0.0232 0.0232

0.0348

0.03

48

0.0348

0.0348

0.03

48

0.0348

0.0348

0.0348

0.04

64

0.0464

0.0464

0.0464

0.04

64

0.04640.0464

0.05

8

0.0580.058

0.05

8

0.0580.058

0.06

96

0.06960.0696

0.06

96

0.06960.0696

0.08

12

0.0812

0.0812

0.08120.0812

0.09280.0928

0.0928 0.0928

0.10

44

0.1044

0.1044

(b)(a)

Fig. 3. Driven cavity for Re = 3200, 5000: streamfunction contours.

Table 5

Streamfunction formulation: compact scheme for the driven cavity problem, Re = 3200

Time Quantity 81 · 81 97 · 97

40 max w 0.1157 0.1145

ð�x; �yÞ (0.5125,0.5500) (0.5104,0.5417)

min w �0.0024 �0.0025

80 max w 0.1152 0.1154

ð�x; �yÞ (0.5125,0.5375) (0.5208,0.5417)

min w �0.0026 �0.0027

160 max w 0.1155 0.1169

ð�x; �yÞ (0.5125,0.5375) (0.5208,0.5417)

min w �0.0026 �0.0027

200 max w 0.1155 0.1172

ð�x; �yÞ (0.5125,0.5375) (0.5208,0.5417)

min w �0.0027 �0.0027

240 max w 0.1156 0.1173

ð�x; �yÞ (0.5125,0.5375) (0.5208,0.5417)

min w �0.0027 �0.0027

360 max w 0.1156 0.1174

ð�x; �yÞ (0.5125,0.5375) (0.5208,0.5417)

min w �0.0027 �0.0027

[10]�s results: max w = 0.1204 at (0.5165,0.5469), min w = �0.0031.

658 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

We also investigated the behavior of the flow for m = 1/7500 and m = 1/10000. Here, the initial flow was

taken from the results of m = 1/5000 at T = 400. For m = 1/7500 at T = 560 with a 97 · 97 mesh, the maximal

value of w is 0.1175, compared to 0.11997 in [10]. The location of the maximal value is

ð�x; �yÞ ¼ ð0:5104; 0:5312Þ, compared to (0.5117,0.5322) in [10]. The minimum value of the streamfunction

is �0.003, where the value �0.0033 was found in [10]. Fig. 6(a) displays streamfunction contours and

Page 20: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

–1 –0.8 –0.6 –0.4 –0.2 0

R=3200 R=5000(b)(a)0.2 0.4 0.6 0.8 1

–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Components, Re=3200, T=360, mesh 97X97

–1 –0.8 –0.6 –0.4 –0.2 0 0.2 0.4 0.6 0.8 1–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Componenets, Re=5000, T=400, mesh 97X97

Fig. 5. Driven cavity for Re = 3200, 5000: velocity components. [10]�s results are marked by �0�.

Re=400 Re=1000–1 –0.8 –0.6 –0.4 –0.2 0 0.2 0.4 0.6 0.8 1

–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Components, Re=400, T=60, mesh 97X97

–1 –0.8 –0.6 –0.4 –0.2 0 0.2 0.4 0.6 0.8 1–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Components, Re=1000, T=80, mesh 97X97

(b)(a)

Fig. 4. Driven cavity for Re = 400, 1000: velocity components. [10]�s results are marked by �0�.

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 659

Fig. 7(a) represents velocity components u(0.5,y) and v(x,0.5) (solid lines) compared with values obtained in

[10] (marked by �0�). The match is excellent.

For m = 1/10000 at T = 500 with a 97 · 97 mesh, the maximal value of w is 0.1190, compared to

0.1197 in [10]. The location of the maximal value is ð�x; �yÞ ¼ ð0:5104; 0:5312Þ, compared to

(0.5117,0.5333) in [10]. The minimum value of the streamfunction is �0.0033, where the value

�0.0034 was found in [10]. Fig. 6(b) displays streamfunction contours and Fig. 7(b) represents velocity

components u(0.5,y) and v(x,0.5) (solid lines) compared with values obtained in [10] (marked by �0�).Note again that the match between the computed u(0.5,y) and v(x,0.5) at T = 500 and [10]�s results

is excellent. However, a steady state has not been reached, as we can observe from Fig. 8(b), which

represents the maximal value of the streamfunction from T = 400 to T = 500, m = 1/10000. A similar

Page 21: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Table 6

Streamfunction formulation: compact scheme for the driven cavity problem, Re = 5000

Time Quantity 81 · 81 97 · 97

40 max w 0.0936 0.0983

ð�x; �yÞ (0.4875,0.6125) (0.5114,0.6146)

min w �0.0029 �0.0030

80 max w 0.1007 0.1010

ð�x; �yÞ (0.5000,0.5125) (0.5312,0.5312)

min w �0.0027 �0.0029

120 max w 0.1060 0.1068

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5417)

min w �0.0028 �0.0028

160 max w 0.1095 0.1105

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5312)

min w �0.0028 �0.0028

200 max w 0.1117 0.1127

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5312)

min w �0.0028 �0.0029

240 max w 0.1131 0.1141

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5417)

min w �0.0028 �0.0029

280 max w 0.1139 0.1150

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5417)

min w �0.0028 �0.0029

400 max w 0.1149 0.1160

ð�x; �yÞ (0.5125,0.5375) (0.5104,0.5417)

min w �0.0028 �0.0029

[10]�s results: max w = 0.11897 at (0.5117,0.5352), min w = �0.0031.

0 0.1 0.2 0.3 0.4 0.5

Re=7500(a) (b) Re=10000

0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=7500, T=560, mesh 97X97

0.0026

0.00

22

0.0018

0.0014

0.0014

0.0014

0.001

0.001

0.00

1

0.00

1

0.0006

0.0006

0.0006

0.000 6

0.00

06

0.0002

0.0002

0.0002

0.0002

0.00

02

0.00020.0002

0

0

0

0

0

0

0

0

0

0.01

175

0.01

175

0.01175

0.01175

0.01175

0.01

175

0.01175

0.01175 0.01175

0.02

35

0.0235

0.0235

0.0235

0.02

35

0.0235

0.0235 0.0235

0.03

525

0.035

25

0.03525

0.03525

0.03

525

0.035250.03525

0.04

7

0.047

0.047

0.047

0.04

7

0.0470.047

0.05

875

0.05875

0.05875

0.05875

0.05875

0.05875

0.07

05

0.07050.0705

0.07

05

0.07050.0705

0.08

225

0.08225

0.08225

0.08225

0.08225

0.094

0.094

0.094 0.094

0.10

575

0.10575

0.10575

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=10000, T=600, mesh 97X97

0.00

35

0.00310.0027

0.0027

0.0027

0.0023

0.0023

0.0023

0.0019

0.0019

0.0019

0.00

19

0.0015

0.0015

0.0015 0.00

15

0.00110.0011

0.0011

0.0011

0.0007

0.0007

0.0007

0.000

7

0.00 07

0.00

07

0.00

03

0.0003

0.0003

0.0003 0.00

03

0.0

003 0.0003

0

0

0

0

0

0

0

0

0

0

0.01198

0.01

198

0.01198

0.01198

0.01198

0.01

198

0.01198

0.01198 0.01198

0.02

396

0.02396

0.02396

0.02396

0.02

396

0.023960.02396

0.02396

0.03

594

0.03594

0.03594

0.03594

0.03

594

0.035940.03594

0.04

792

0.04792

0.04792

0.04792

0.047

92

0.047920.04792

0.05

99

0.05990.0599

0.05

99

0.05990.0599

0.07

188

0.07188

0.07188

0.07

188

0.07188

0.07188

0.08

386

0.08386

0.08386

0.08386 0.08386

0.09

584

0.09584

0.09584

0.09584

0.10

782

0.10782

0.10782

Fig. 6. Driven cavity for Re = 7500, 10000: streamfunction contours, (a) max w = 0.1175, [10]�s 0.11998; location is (0.5104,0.5312),

[10]�s (0.5117,0.5322); min w = �0.0030, [10]�s = �0.0033. (b): max w = 0.1190, [10]�s 0.1197; location is (0.5104,0.5312), [10]�s(0.5117,0.5333); min w = �0.0033, [10]�s = �0.0034.

660 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

Page 22: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

–1 –0.8 –0.6 –0.4 –0.2 0

Re=7500 Re=10000(b)(a)0.2 0.4 0.6 0.8 1

–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Components, Re=7500, T=560, mesh 97X97

–1 –0.8 –0.6 –0.4 –0.2 0 0.2 0.4 0.6 0.8 1–1

–0.8

–0.6

–0.4

–0.2

0

0.2

0.4

0.6

0.8

1Velocity Components, Re=10000, T=500, mesh 97X97

Fig. 7. Driven cavity for Re = 7500, 10000: velocity components. [10]�s results are marked by �0�.

Re=7500 Re=10000(b)(a)400 410 420 430 440 450 460 470 480 490 500

0.115

0.1155

0.116

0.1165

0.117

0.1175

0.118

0.1185

0.119Max Streamfunction, Re=7500, T=400 to 500, mesh 97X97

400 410 420 430 440 450 460 470 480 490 5000.115

0.116

0.117

0.118

0.119Max Streamfunction, Re=10000, T=400 to 500, mesh 97X97

Fig. 8. Driven cavity for Re = 7500, 10000: max streamfunction, T = 400–500.

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 661

plot – Fig. 8(a) – shows that for m = 1/7500 the same quantity grows monotonically toward a steady-

state, while for m = 1/10000 we observe that it grows non-monotonically. A similar phenomena was

observed for m = 1/8500, in agreement with [17] and [12]. Therefore, it seems that in [10] a steady state

solution was computed, however the solution of the time-dependent problem does not tend to a steady

state. A similar phenomenon was observed in [12,17,19]. It is commonly interpreted as an indication

that, while a steady state solution is computed in [10] for high Reynolds numbers, they are unstable

and experience a Hopf bifurcation into time-periodic solutions. The rigorous analysis proving this

bifurcation has not yet been performed.A similar problem, that we considered, has the same geometry X, but here the fluid is driven also in the

negative y-direction at the left-end of X. Thus, u = 1,v = 0 for y = 1, u = 0,v = �1 for x = 0, u = v = 0 for

Page 23: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

0 0.1 0.2 0.3 0.4

Re=400(a) (b) Re=1000

0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Velocity Components, Re=400, T=100, mesh 81X81

0.080718

0.080718

0.07

2221

0.07

2221

0.063725

0.063725

0.063725

0.055228 0.05

5228

0.055228

0.046731

0.046731

0.046731

0.04

6731

0.038235

0.038235

0.03

8235

0.038235

0.029738

0.029738

0.02

9738

0.029738

0.02

9738

0.021242

0.021242

0.021242

0.021242

0.02

1242

0.01

2745

0.01

2745

0.0127450.012745

0.012745

0.012745

0.00

4248

30.

0042

483

0.00424830.0042483

0.0042483

0.0042483

0.0042483

0.00

4248

3

0.0042483

0.0042483

0.0042483

0.0042483 0.0042483 0.0042483

0.01

2745

0.0127450.012745

0.012745

0.012745 0.012745 0.012745

0.02

1242

0.021242

0.021242

0.02

1242 0.021242 0.021242

0.02

9738

0.029738

0.029738

0.029738 0.029738

0.03

8235

0.038235

0.038235

0.038235

0.04

6731

0.0467310.

0467

31

0.046731

0.0552280.055228

0.055228

0.063725

0.063725

0.063725

0.072221

0.072221

0.080718

0.080718

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=1000, T=100, mesh 81X81

0.08394

0.08394

0.075103

0.075103

0.066266

0.06

6266

0.066266

0.057428

0.0574280.057428

0.0485910.048591

0.0485910.048591

0.0397540.039754

0.039754

0.039754

0.03

0916

0.030916

0.030916

0.030916

0.030916

0.0220790.022079

0.022

079

0.022079

0.022079

0.01

3242

0.0132420.013242

0.01

3242

0.013242

0.013242

0.00

4404

50.

0044

045

0044040.0044045

0.00

4404

5

0.0044045

0.0044045 0.00

4432

7

0.00443270.0044327

0.0044327

0044327

0.0044327 0.0044327

0.01

327

0.01327

0.01327

0.01327 0.01327 0.01327

0.02

2107

0.022107

0.022107

0.02

2107

0.022107

0.03

0945

0.030945

0.030945

0.030945 0.030945

0.03

9782

0.039782

0.03

9782

0.039782

0.048619

0.048619

0.048619 0.048619

0.05

7456

0.057456

0.057456

0.066294

0.066294

0.066294

0.075131

0.075131

0.083968

0.083968

Fig. 9. Double driven cavity for Re = 400, 1000: streamfunction contours.

Re=3200(a) (b) Re=5000

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=3200, T=100, mesh 97X97

0.063983

0.063983

0.055464

0.0554640.046946

0.046946

0.0469460.038427

0.038427

0.0384270.029909

0.029909

0.029909

0.02

9909

0.02139

0.02139

0.02139

0.02139

0.01

2872

0.012872

0.012872

0.012872

0.00

4353

30.

0043

533

0.0043533

0.0043533

0.0043533

0.0043533

0.0043533

0.0043533

0.00

4165

2

0.0041652

0.0041652

0.00416520.0041652

0.0041652 0.0041652 0.0041652

0.0041652

0.01

2684

0.012684

0.012684

0.012684

0.01

2684

0.012684 0.012684

0.02

1202

0.021202

0.021202

0.021202

0.021202 0.021202

0.02

9721

0.029721

0.029721

0.029721

0.029721 0.029721

0.03

8239

0.038239

0.038239

0.03

8239

0.038239 0.038239

0.0467580.046758

0.04

6758

0.046758 0.046758

0.05

5276

0.055276

0.055276

0.055276 0.055276

0.063

795

0.063795

0.06

3795

0.063795

0.07

2313

0.072313

0.0723130.072313

0.08

0832

0.080832

0.080832 0.080832

0.08935

0.089350.08935

0.097869

0.09

7869

0.10639

0.10639

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1Streamfunction Contours, Re=5000, T=100, mesh 97X97

0.039815

0.031162

0.031162

0.02

2509

0.02

2509

0.013856

0.013856

0.013856

0.00

5202

70.

0052

027

0.0052027

0.0052027

0.0052027

0.0052027

0.00

3450

3

0.00

3450

3

0.0034503

0.0034503

0.0034503

0.00

3450

3

0.0034503 0.0034503

0.01

2103

0.01

2103

0.012103

0.012103

0.012103

0.012103 0.012103 0.012103

0.02

0756

0.020756

0.020756

0.020756

0.02

0756

0.020756 0.020756

0.02

9409

0.029409

0.029409

0.029409

0.0294090.029409 0.029409

0.03

8062

0.038062

0.038062

0.038062

0.0380620.038062 0.038062

0.04

6716

0.0467160.046716

0.046716

0.046716 0.046716

0.05

5369

0.055369

0.055369

0.05

5369

0.0553690.055369

0.0640220.064022

0.06

4022

0.0640220.064022

0.07

2675

0.072675

0.072675

0.072675 0.072675

0.0813280.081328

0.08

1328

0.081328

0.08

9981

0.089981

0.0899810.089981

0.09

8634

0.098634

0.098634

0.10729

0.10729

0.10729

0.11594

0.11594

Fig. 10. Double driven cavity for Re = 3200, 5000: streamfunction contours.

662 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

x = 1 and y = 0. We picked m = 1/400, 1/1000, 1/3200, 1/5000 and started the flow impulsively from zero.Figs. 9(a), (b) and 10(a), (b) represent the streamfunction at t = 100 for the various viscosity coefficients,

respectively. Note that at m = 1/3200 we start to observe symmetry breaking with 81 · 81 and 97 · 97

meshes. Numerical results by Pan and Glowinski [17] indicate the same phenomena for 1/m between

4000 and 5000. In Fig. 11(a) and (b), the maximum of the streamfunction from T = 0 to T = 200 is dis-

played, for Re = 3200 and Re = 5000, respectively. A closer look to T = 400 shows that no steady state

is achieved for both cases. Fig. 12(a) and (b) shows the same quantities as in Fig. 11(a) and (b), for

T = 200 to T = 400.

Page 24: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

Re=3200(a) (b) Re=50000 20 40 60 80 100 120 140 160 180 200

0

0.05

0.1

0.15Max Streamfunction Double Driven, Re=3200, T=0200, mesh 97X97

0 20 40 60 80 100 120 140 160 180 2000

0.05

0.1

0.15Max Streamfunction Double driven, Re=5000, T=0200, mesh 97X97

Fig. 11. Double driven cavity for Re = 3200, 5000: max streamfunction, T = 0–200.

200 220 240 260 280 300

Re=3200(a) Re=5000(b)320 340 360 380 400

0.115

0.116

0.117

0.118

0.119

0.12Max Streamfunction DoubleDriven, Re=3200, T=200 to 400

200 220 240 260 280 300 320 340 360 380 400

0.11

Max Streamfunction DoubleDriven, Re=5000, T=200 to 400

Fig. 12. Double driven cavity for Re = 3200, 5000: max streamfunction, T = 200–400.

M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664 663

Acknowledgment

Supported in part by funds from the Israel Science Foundation (ST).

References

[1] I. Altas, J. Dym, M.M. Gupta, R.P. Manohar, Mutigrid solution of automatically generated high-order discretizations for the

biharmonic equation, SIAM J. Sci. Comput. 19 (1998) 1575–1585.

[2] M. Ben-Artzi, Vorticity dynamics in planar domains, in preparation.

[3] M. Ben-Artzi, D. Fishelov, S. Trachtenberg, Vorticity dynamics and numerical resolution of Navier–Stokes equations, Math.

Model. Numer. Anal. 35 (2) (2001) 313–330.

[4] D. Calhoun, A cartesian grid method for solving the two-dimensional streamfunction-vorticity equations in irregular regions, J.

Comp. Phys. 176 (2002) 231–275.

[5] A.J. Chorin, Numerical solution of the Navier–Stokes equations, Math. Comp. 22 (1968) 745–762.

[6] G.-H. Cottet, P.D. Koumoutsakos, Vortex Methods. Theory and Practice, Cambridge University Press, 2000.

[7] W. E, J.-G. Liu, Essentially compact schemes for unsteady viscous incompressible flows, J. Comp. Phys. 126 (1996) 122–138.

[8] W.E, J.-G. Liu, Vorticity boundary condition and related issues for finite difference scheme, J. Comp. Phys. 124 (1996) 368–382.

Page 25: A pure-compact scheme for the streamfunction …math.huji.ac.il/.../navier-stokes.JCP.pdfA pure-compact scheme for the streamfunction formulation of Navier–Stokes equations Matania

664 M. Ben-Artzi et al. / Journal of Computational Physics 205 (2005) 640–664

[9] D. Fishelov, M. Ben-Artzi, J.-P. Croisille, A compact scheme for the stream-function formulation of Navier–Stokes equations,

Notes on Computer Science (Springer-Verlag) 2667 (2003) 809–817.

[10] U. Ghia, K.N. Ghia, C.T. Shin, High-Re solutions for incompressible flow using the Navier–Stokes equations and a multigrid

method, J. Comp. Phys. 48 (1982) 387–411.

[11] P.M. Gresho, Incompressible fluid dynamics: some fundamental formulation issues, Annu. Rev. Fluid Mech. 23 (1991) 413–453.

[12] R. Kupferman, A central-difference scheme for a pure streamfunction formulation of incompressible viscous flow, SIAM J. Sci.

Comput. 23 (1) (2001) 1–18.

[13] H. Lamb, Hydrodynamics, 6th ed., Dover, 1932.

[14] Ming Li, Tao Tang, A compact fourth-order finite difference scheme for unsteady viscous incompressible flows, J. Sci. Comput. 16

(1) (2001) 29–45.

[15] Zhilin Li, Cheng Wang, A fast finite difference method for solving Navier–Stokes equations in irregular domains, Comm. Math.

Sci. 1 (1) (2003) 180–196.

[16] J.L. Lions, E. Magenes, Non-Homogeneous Boundary Value Problems and Applications I, Springer-Verlag, Berlin, 1972.

[17] T.W. Pan, R. Glowinski, A projection/wave-like equation method for the numerical simulation of incompressible viscous fluid

flow modeled by the Navier–Stokes equations, Comput. Fluid Dynamics 9 (2000).

[18] L. Quartapelle, Numerical Solution of the Incompressible Navier–Stokes Equations, Birkhauser Verlag, 1993.

[19] J. Shen, Hopf bifurcation of the unsteady regularized driven cavity, J. Comp. Phys. 95 (1991) 228–245.

[20] J.W. Stephenson, Single cell discretizations of order two and four for biharmonic problems, J. Comp. Phys. 55 (1984) 65–80.

[21] R. Temam, Navier–Stokes Equations, AMS Edition, 2001.

[22] T.E. Tezduyar, J. Liou, D.K. Ganjoo, M. Behr, Solution techniques for the vorticity-streamfunction formulation of the two-

dimensional unsteady incompressible flows, Int. J. Numer. Meth. Fluids 11 (1990) 515–539.


Recommended