The Least-Squares Finite Element Method Theory and Applications in Computational Fluid Dynamics and Electromagnetics
The Least-Squares Finite Element Method Theory and Applications in Computational Fluid Dynamics and Electromagnetics
The Least-Squares Finite Element Method Theory and Applications in Computational Fluid Dynamics and Electromagnetics
Editorial Board
J.-J. Chattot, San Francisco, CA, USA
C. A. J. Fletcher, Sydney, Australia
R. Glowinski, Toulouse, France
W. Hillebrandt, Garching, Germany
M. Holt, Berkeley, CA, USA
Y. Hussaini, Hampton, VA, USA
H. B. Keller, Pasadena, CA, USA
J. Killeen, Livermore, CA, USA
D. I. Meiron, Pasadena, CA, USA
M. L. Norman, Urbana, IL, USA
S. A. Orszag, Princeton, NJ, USA
K. G. Roesner, Darmstadt, Germany
V. V. Rusanov, Moscow, Russia
The least-Squares
Finite Element Method
Theory and Applications
in Computational Fluid Dynamics
and Electromagnetics
, Springer
Dr. Bo-nan Jiang
Institute for Computational Methods in Propulsion
NASA Lewis Research Center
Cleveland, OH 44135, USA
ISSN 0172-5726
ISBN 978-3-642-08367-9
Library of Congress Cataioging-in-Publication Data. Jiang, Bo-nan, 1940- The least-squares finite
element method: theory and appIications in computational fluid dynamics and electromagnetics /
Bo-nan Jiang. p. cm. - (Scientific computation, ISSN 0172-5726) Includes bibliographicai references
and index.
ISBN 978-3-642-08367-9 ISBN 978-3-662-03740-9 (eBook)
DOI 10.1007/978-3-662-03740-9
1. Fluid mechanics-Mathematics. 2. Electro-
megnetics-Mathematics. 3. Finite element method. 4. Least squares. 5. Differential equaions, Par-
tial-Numerical solutions. 1. Tide. II. Series. QC151.J53 1998 532'.051'01515353-dc21 97-51980
This work is subject to copyright. AII rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broad-
casting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this
publication or parts thereof is permitted only under the provisions of the German Copyright Law of
September 9, 1965, in its current version, and permission for use must always be obtained from
Springer-Verlag. Violations are liable for prosecution under the German Copyright Law.
© Springer-Verlag Berlin Heidelberg 1998
Originally pubIisbed by Springer-Verlag Berlin Heidelberg New York in 1998
Softcover reprint of the hardcover 1 st edition 1998
The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant pro-
tective laws and regulations and therefore free for general use.
Typesetting: Data conversion by Satztechnik Katharina Steingraeber, Heidelberg
Cover design: design & production GmbH, Heidelberg
SPIN 10559938 55/3144 - 5432 10- Printed on acid-free paper
Preface
Friedrichs inequalities related to grad, div and curl operators. Of course, the
analysis can also be based on the famous Lax-Milgram theorem. However, the
proof of the Lax-Milgram theorem needs knowledge of functional analysis and
its application is limited to elliptic problems. Therefore, in this book we em-
phasize the approach based on the bounded inverse theorem. This approach
gives a clear picture of the least-squares method without specialized math-
ematical knowledge beyond calculus and elementary differential equations.
Following this principle, this book establishes an almost self-sufficient and
reasonably rigorous mathematical framework for the least-squares method,
while the mathematics has been kept as simple as possible.
This book is written mainly for both engineers and physicists as well as
researchers. Most parts of the book cover diverse applications of the LSFEM.
It is presumed that the reader has basic knowledge about the finite el-
ement method for second-order elliptic partial differential equations. The
introduction of the finite element method is reduced to a minimum in order
to keep the book short. For those readers who are not familiar with finite
elements, many excellent introductory books are available. For example, the
readers may consult Zienkiewicz and Morgan (1983) or Becker et al. (1983).
The book contains fifteen chapters and four appendices.
Part I (Chaps. 1-3) gives some basic ideas about the least-squares method.
Chapter 1 provides an overview of the LSFEM.
Through simple one-dimensional examples, Chap. 2 explains why the
standard Galerkin method or the central difference method fails and why
the LSFEM doesn't need upwinding and is perfectly suitable for convective
transport problems described by first-order derivatives.
Chapter 3 compares the LSFEM with the mixed Galerkin method for
first-order elliptic systems and shows why the LSFEM can accommodate
equal-order elements.
Part II (Chaps. 4-6) introduces some theoretical aspects of the LSFEM.
Chapter 4 provides the mathematical foundation and general formulation
of the LSFEM based on a first-order system of partial differential equations.
Chapter 5 deals with the div-curl system which is fundamental for study-
ing the incompressible Navier-Stokes equations in fluid dynamics and the
Maxwell equations in electromagnetics. This chapter shows that the three-
dimensional div-curl system is not overdetermined and the LSFEM is the best
one for seeking its solution. This chapter introduces the div-curl method and
the least-squares method for deriving equivalent second-order equations and
their boundary conditions.
Chapter 6 deals with diffusion problems which are usually described by
a second-order scalar elliptic equation. A second-order equation can be de-
composed into a grad-div system or a div-curl-grad system. Both the mixed
Galerkin method and the conventional least-squares method are based on
the grad-div system. This chapter shows that the LSFEM based on the div-
curl-grad system has significant advantages: accommodation of equal-order
Preface VII
The book reflects recent developments of the LSFEM. Most of the results
in this book have been obtained by the author and his collaborators, and
many of them are not published elsewhere.
1. Introduction.............................................. 3
1.1 Why Finite Elements? .................................. 3
1.2 Why Least-Squares? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
5. Div-Curl System .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 81
5.1 Basic Theorems. .. . . .. . . .. . . .. . . . . . . . .. . . .. . . . . . . . . . . .. 81
5.2 Determinacy and Ellipticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 86
5.3 The Div-Curl Method .................................. 88
5.4 The Least-Squares Method .............................. 90
5.5 The Euler-Lagrange Equation. . . . . . . . . . . . . . . . . . . . . . . . . . .. 91
5.6 The Friedrichs Second Div-Curl Inequality ................ 93
5.7 Concluding Remarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 95
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
In this chapter we briefly overview the basic ideas and features of the least-
squares finite element method. In the following chapters we will elaborate on
the view points discussed in this chapter.
t
a "weighted residual" ,
where Vi and Vi are "suitably chosen" test functions, and T denotes the
transpose. In the conventional Galerkin method, the choice is
1.2 Why Least-Squares? 5
Vi = Vi = <l?i. (1.4)
For equations with self-adjoint and positive definite operators, the Galer-
kin formulation results in the same system of equations as in the Rayleigh-
Ritz formulation. However, the Galerkin method appears more general, be-
cause it is applicable for non-self-adjoint equations such as arise in fluid me-
chanics. This generality of the Galerkin finite element method provided, in the
early 1970s, a strong impetus for the utilization of the method in fluid dynam-
ics. It was thought that the significant advantages gained in solid mechanics
and diffusion-type problems by using the Rayleigh-Ritz method would again
be open to exploitation in the area of fluid flow by using the Galerkin method.
In reality, this proved to be too optimistic, especially for modeling convection
dominated flow problems. Convection operators are of first order, and thus
non-self-adjoint; as a result, the Galerkin method does not exhibit the best
approximation property.
In practice, solutions to convection dominated transport problems by the
Galerkin method are often corrupted by spurious oscillations or "wiggles".
These can only be removed by severe mesh refinements which clearly under-
mine the practical utility of the method.
The classic Galerkin method also behaves poorly for high-speed com-
pressible flow and shallow water wave problems. In these cases, the flows are
governed by nonlinear first-order hyperbolic equations; their solutions may
be discontinuous, that is, shocks often develop even if the initial flow field is
smooth.
Notorious difficulties arise even for the solution of elliptic problems by
the Galerkin mixed method. One example is incompressible irrotational flow
problems governed by Laplace or Poisson equations of the potential. In com-
mon practice, the primal variable, i.e., the potential is solved by the Rayleigh-
Ritz method, then a posteriori numerical differentiation is required to obtain
the dual variables, i.e. the velocity components which are more important.
The computed velocity is not continuous across element boundaries, and in
general its accuracy is one order lower than that of the potential. The Galerkin
mixed method based on the first-order equations was devised in the hope of
obtaining better accuracy for both primal and dual variables. Here the term
'mixed' means that both primal variables and dual variables are approxi-
mated as fundamental unknowns. Unfortunately, the Galerkin mixed method
brings perhaps more troubles than benefits, at least for second-order ellip-
tic problems. First, the original simple minimization problem is turned into
a difficult saddle-point problem, and one must use different elements to in-
terpolate primal variables and dual variables, and often lower-order elements
must be employed for dual variables to satisfy the celebrated Ladyzhenskaya-
Babuska-Brezzi (LBB) condition (see, e.g., Babuska 1971, Brezzi 1974, and
Oden and Carey 1983). As a consequence, better accuracy for dual variables
may not be achieved. Second, the Rayleigh-Ritz method leads to symmetric
positive-definite matrices, while the Galerkin mixed method produces non-
6 1. Introduction
positive-definite matrices which have been hard to solve for large-scale prob-
lems.
Another well-known example is viscous flow problems governed by the
incompressible Navier-Stokes equations. Here one encounters the same insta-
bilities when the Galerkin mixed method is used for the solution of velocity
and pressure.
In computational electromagnetics (CEM), the most popular finite ele-
ment methods for static and eddy currents problems are based on the use of
potentials that involve difficulties related to the appropriate gauging method
and the loss of accuracy and inter-element continuity of the calculated field
intensity by numerical differentiation. It is commonly believed that the di-
vergence equations in the Maxwell equations are redundant. Therefore, con-
ventional time-domain methods are based only on the two curl equations,
and frequency methods are based only on the curl-curl equations. Due to the
appearance of spurious modes which violate the divergence-free equations,
the direct solution of the field intensity vectors in the first-order Maxwell
equations by using the node-based Galerkin-type procedure has not been
successful.
It is not an exaggeration to say that during the passed three decades much
of finite element research in fluid mechanics was devoted to modification
and improvement of the Galerkin method. The achievement of mathematical
analysis ofthe Galerkin method is remarkable; however, the application of the
Galerkin method in computational fluid dynamics (CFD) and computational
electromagnetics (CEM) has not produced entirely satisfactory results. This
is a major reason why the finite difference and finite volume methods are
much in vogue in CFD and time-domain CEM.
It is the author's opinion that for non-self-adjoint systems, such as arise
in fluid mechanics and electromagnetics, the application of the least-squares
finite element method (LSFEM) is the right direction to go.
The LSFEM is based on the minimization of the residuals in a least-
squares sense. The method seeks the minimizer of the following functional
(1.5)
within the constraint of a given boundary condition (1.1b). That is, the least-
squares solution is calculated from the following variational statement:
(1.6)
(4) Optimality. In many cases it can be rigorously proven that the LSFEM
solution is the best approximation, that is, the error of the LSFEM solution
has the same order as the interpolation error. In many areas of engineering
and applied science, it is as important to evaluate the accuracy of an ap-
proximate solution as it is to obtain that solution. The least-squares method
meets the need for posterior error analysis by supplying an error indicator in
the form of the residuals that are minimized by the procedure. This very re-
liable error indicator can be used for adaptive refinement to achieve optimal
solutions. No other approximate method can supply this information without
additional calculation.
(5) Concurrent simulation of multiple physics. Since the LSFEM is a uni-
fied method for the approximate solution of differential equations governing
diverse physical phenomena, a one-algorithm, one-code and one-pass simu-
lation tool for concurrent analysis of different disciplines, such as conjugate
heat transfer and solid-fluid interaction, can be developed.
(6) General-purpose coding. As mentioned above, the LSFEM is formulated in
a very general setting. Therefore, the LSFEM can be programmed systemat-
ically, such that for a new application, one need only add simple subroutines
to supply the coefficients, the load vector, and the boundary conditions for
the first-order system. This can drastically reduce the time, cost and pro-
gramming errors in code development.
The least-squares method also has a great potential for the theoretical
study of the mathematical properties of partial differential equations includ-
ing uniqness and stability of solutions and the permissibility of boundary
conditions, as well as the derivation of equivalent high-order versions of equa-
tions.
The mathematical study of the incompressible Navier-Stokes equations
has been mainly confined to the standard case in which velocity compo-
nents are prescribed on the boundary, and traditionally conducted via the
Galerkin method which leads to a difficult saddle-point problem. On the
other hand, computational researchers often choose non-standard boundary
conditions based on physical intuition. Although the Navier-Stokes equations
were established more than one hundred years ago, there has been a lack of
systematic study about their permissible boundary conditions. Recently, by
considering the div-curl structure of the incompressible Stokes and Navier-
Stokes equations in the velocity-pressure-vorticity formulation and by us-
ing the least-squares method, for the first time, mathematically permissible
boundary conditions have been systematically and rigorously derived and
analyzed.
The Maxwell equations in electromagnetics were established by James
Clerk Maxwell in 1873. The original first-order full Maxwell equations, which
consist of two curl equations (the Faraday and Ampere laws) and two di-
vergence equations (the Gauss laws), reflect general and independent laws
of physics. These laws can not be deduced from each other. They govern all
1.2 Why Least-Squares? 9
world problems. For these reasons, we believe that the LSFEM will play an
important role in computational engineering and science in the twenty first
century.
2. First-Order Scalar Equation
in One Dimension
f(x) = {2X/1,
1,
if °: ;
x ::; 1/2;
if 1/2::; x::; 1
(2.7)
Proof: Since
lu(xW = 11 x
1· u'(t)dtI2 :::; 1 ·l
x
1dt
x
lu'(t)1 2dt :::; l .1 1
lu'(xWdx.
we have
lu(x)1 ~ l x
lU' f + u!'ldt ~ IlflllllUlll ~ Clluill.
By changing f(x) we can prove (2.9) if x E [0, l/2J. o
Proof' Assume that f(x) i= 0, then there exists a point Xo in n such that
f(xo) i= 0 (assuming that f(xo) > 0 without loss of generality). From con-
tinuity we know that f(x) > 0 in a neighborhood Kp = {Ix - xol < p} for
some p > O. Let
r/1(x) = {p2 - (x - xo)2, if x E ~P;
0, otherwIse.
Obviously r/1 E GO(n) and r/1 > 0 in Kp, therefore
At first, let us try to use the classic Galerkin method for finding the approxi-
mate solution of the problem (2.1). The Galerkin method is a member of the
class of weighted residual methods. The starting point for a weighted residual
method is to assume an approximate solution written as
x E n= [0,1]' (2.10)
where the basis functions r/11, ... , r/1n are known. That is, the approximate
(trial) solution is expressed as a linear combination of the basis functions. Of
course, u(x) should satisfy the boundary condition u(O) = O. In one spatial
dimension, the simple basis functions might be polynomials or trigonometric
functions, e.g.,
2.3 The Classic Galerkin Method - Global Approximation 15
or j = 1,2, ....
The coefficients aj are unknowns to be determined. If the approximate solu-
tion (2.10) is substituted into (2.1), it will not generally be identically zero.
Thus we can write
R(x) = u'(x) - f(x), (2.11)
where R(x) is referred to as the equation residual.
In the method of weighted residual, the coefficients aj are determined by
requiring that the integral of the weighted residual over the computational
domain is zero, i.e.,
l v(x)R(x)dx = 0. (2.12)
If the integral in (2.12) is zero for any weight (test) function v(x) E COUl),
due to the basic lemma of variational principles, the residual R will be
identical to zero, i.e., the differential equation (2.1) is satisfied by the so-
lution u(x). However, in practice, it is impossible or unnecessary to test
the residual R by all CO functions in [0,1]' since the approximate solu-
tion u(x) itself has only n degrees of freedom, and n test functions would
be enough. Now the question is how to choose the test functions. Differ-
ent choices for the test functions v(x) in (2.12) give rise to different meth-
ods. In the Galerkin method, the test functions are chosen from the same
family as the approximating (trial) functions. If the basis functions form
a complete set (on [0,1] a complete set of trigonometric functions would
be 1, sin(1l'x), cos (1l'x) , sin(21l'x), cos(21l'x) , ... , sin(n1l'x), cos(n1l'x)) , (2.12) indi-
cates that the residual is orthogonal to every member of a complete set.
Consequently, one hopes that as n tends to infinity the approximate solution
will converge to the exact solution.
Unfortunately, in solving first-order differential equations, the classic
Galerkin method does not behave as expected. Let us choose the following
simple problem as an example:
u' = 21l'cos( 1l'x) in [0,1]' (2.13a)
u=o at x = 0. (2.13b)
Obviously,
uexact = 2sin(1l'x).
The one-point boundary value problem (2.13) can now be given in terms
°
of the following Galerkin formulation: Find u E V = {u E H1(0,1) : u =
at x = O} such that
and thus
a = 2,
which is the correct solution. However, the Petrov-Galerkin method itself
cannot answer the question: How to choose an appropriate space of test
functions for general problems? Therefore, the Petrov-Galerkin method is
not always a satisfactory method.
or
(u ' , Vi) = (j, Vi) Vv E V. (2.21 )
Obviously, the least-squares method expressed by (2.21) can be interpreted
as a Petrov-Galerkin method in which Vi is chosen as a test function in stead
of v.
We shall now use the least-squares method to solve the first-order differ-
ential equation (2.13). As in Sect. 2.3, we select the trial function
u = asin(1I'x). (2.22)
Thus the test function is
Vi = 1I'cos(1I'x). (2.23)
To determine a we substitute (2.22) and (2.23) into the least-squares formu-
lation (2.21) and obtain
1/J~j)(X), x E ej;
I};j(x) = { 1/JiJ+l) (x), x E ej+1; 1 :::; j :::; n - 1
0, otherwise;
x E en;
otherwise;
are the so-called basis functions. I};j(x) is a piecewise continuous linear func-
tion with a value unity at node Xj and null at other nodes, as illustrated in
Fig. 2.2. Therefore, Uh(Xj) = Uj. Especially, if we take Uo = 0, then (2.26)
satisfies the boundary condition (2.1b), and other parameters (nodal values)
Ul, U2, ... , Un can be arbitrary. All functions Uh(X) constitute the space of trial
functions Vh .
I____ ~----~----~----~----~--~~ }{
1
Xj+l
E'(x) = l x
E"(t)dt.
Since Ihu is linear, E = u - Ihu implies E" = u" - (Ihu)" = u" within ej.
Applying the Schwarz inequality (2.4), we find
r/
IE'(x)1
E(x) = l x
X;-l
E'(t)dt
r/
and use (2.28) to obtain
X;-l
; E"(t)2dt 2
(2.29)
Squaring (2.28) and (2.29), integrating, summing over all elements, and
taking square roots, we arrive at the estimates of the error:
lI(u-Ihu)'lIo :::; hlul2' (2.30a)
lIu-Ihullo :::; h21u12. (2.30b)
Finally since
lIu -lhull~ = lIu -lhull~ + lI(u -lhu),II~,
we find easily that
lIu -lh u ll1 :::; V2h lu l2' (2.31)
Note that the bounds (2.30) and (2.31) are satisfactory in the sense that
the powers of h are the best possible. To obtain an improved result, by ex-
panding the error E(x) in a sine series, one can verify that (Strang and Fix
1973, p.45)
II(u -lhu)'llo :::; 1I'-1hluI2, (2.32)
lIu -lhullo :::; 1I'- 2h2IuI2' (2.33)
Since
n
Uh(X) = L <pj(x)Uj, (2.35)
j=l
and (2.34) must be valid by taking Vh = <Pi(X), i = 1,2, ... , n as the test
functions, we can write (2.34) as
n
L(<Pi' <pj)Uj = (<Pi, I) i = 1,2, ... ,n, (2.36)
j=l
Ki,i-l
,
= (<Pi, <Pi-I) =
lX, (x - Xi-I) -1
h . · -;:;:dx =
1
-2'
Xi-l t ,
and
K
nn
= ('" "") =
'l'n, 'l'n
lxn
Xn-l
(x - Xn-l) . ~d = ~
h
n
h X 2·
n
l x,
Xi-l
(X - Xi-l ) .
hi
fdx +
Xi
xi + 1 (
Xi+l
hi+l
- X) • fdx
1 1
2 •• + -f·h·+
-f·h· 2" l ,
and
U3 - Ul - f
2h - 2,
U4 - U2 - f
2h - 3·
The above is, incidentally, identical to the standard central difference approx-
imation obtained by putting
, Ui+l - Ui-l
U ~ 2h
The matrix is obviously non-symmetric with zeros on the diagonal, and has
odd-even decoupling which leads to oscillatory solutions. This trouble persists
even after the mesh is refined. In order to overcome this bad approximation
problem, finite difference practitioners use one-sided or upwind differencing
Ui+l - Ui-l h Ui+l - 2Ui + Ui-l
2h - '2 h2
which is equivalent to adding some numerical dissipation into the central
difference scheme. That is, they solve, instead of (2.1), the following second-
order equation by the central difference method
h
u' - -u"
2 = f.
2.7 The Least-Squares Finite Element Method 23
For the first-order differential equation (2.1) the least-squares finite element
method is formulated according to (2.21) as follows: Find Uh E Vh such that
Since
Uh(X) = ¢>1(X)Ul + ... + ¢>n(x)un , (2.40)
(2.39) can be written as
n
~)¢>~, ¢>j)Uj = (¢>~, 1) i = 1,2, ... ,n, (2.41 )
j=1
or in matrix form as
KU=F.
l
By using one-point Simpson quadrature we obtain for i = 1,2, ... , n - 1
l X
1
Xi-l hi
•
-fdx+
Xi
+ -1
xi 1
-fdx
hHI
(fi-l + fi) 1 h (fi + fHl) 1 h
2 hi i - 2 hHI HI
(fHl - fi-l)
2
and
l Xi-l
Xi
1
hr1
-dx+ l X
Xi
i+ 1 1
--dx
hf+!
1
= -+- i=1,2, ... ,n-1,
hi hi+!
i = 2, ... ,n -1,
Xi
+1 1
~dx
HI
1
= --,;:-
,+1
i = 2, ... ,n -1,
and
24 2. First-Order Scalar Equation in One Dimension
Knn = (¢~,¢~) = l xn
Xn-l
1 1
h 2 dx = h'
n n
we have
n n
U TKU = (2: Ui¢~' 2: Uj¢j) = (u~, u~) ~ 0
i=l j=l
o 0 0 0 0 -1 2 -1
o 0 0 0 0 0 -1 1 Un
12-10
iJ-il
1 14-12
(2.42)
2h
-In-1 - In
Obviously, the left-hand side of (2.42) can be interpreted as a standard
central difference for u" and the right-hand side as a central difference for 1'.
We clearly see again that the least-squares finite element method for the first-
order differential equation leads to a symmetric and positive-definite system
of linear algebraic equations.
We shall now study the error of the least-squares finite element solution for
the model problem (2.1). Let U be the solution of
w = 0 at x = 0,
w' =0 at x = 1,
where e is the error of the least-squares finite element solution of (2.47). From
Sect. 2.5 we know that the finite element interpolant Ihw satisfies
II(w -lhw)'lla ::; Chllw"lla = Chllella. (2.53)
Using integration by parts and the fact that e = 0 at x = 0 and w' = 0 at x =
1,
(e,e) = -(e,w") = (e',w') = (e',w'-llhW'),
where the last equality follows from (2.48), since Ihw E Vh so that (e', Ihw')
= o. Applying the interpolation estimate (2.53) to wand using also (2.50),
we find
Ile116::; Ile'llall(w - llhW)'lla ::; Chlul2 h ll e lla.
Dividing by Ilelia we finally obtain
Ilelia ::; Ch 21u12. (2.54)
The result (2.54) indicates that the least-squares finite element method for
first-order equations has an optimal rate of convergence in the sense that the
error measured in L2 norm is of the same order of h as the finite element
interpolation.
28 2. First-Order Scalar Equation in One Dimension
[
u'(x) = 1- exp -~ ( 1)]-11~exp (I-X) --E- XE[O,lJ, (2.61a)
u(O) = 0, (2.61b)
with 0 < E < < 1. The exact solution of this problem is given by
3.0 ,-----,r----,----,.---r---,---,------,----,-----.-----,
- - Exact
Go - -0 LSFEM
2.0
......... Central FD /
- - Upwinding FD /
- - Galerkin FEM /
1.0 / ~
/ ~
We have compared in this chapter the least-squares method with the Galerkin
method for the solution of first-order scalar differential equations.
The Galerkin method is identical to the common central finite difference
approximation. The algebraic equations generated by the Galerkin method
are obviously non-symmetric, non-positive-definite and odd-even decoupled,
and the solution is thus purely oscillatory and bears almost no relation to
the underlying problem.
The least-squares method for first-order scalar differential equations is
formally equivalent to the Galerkin method for corresponding second-order
problems, and hence has an optimal rate of convergence. The least-squares
approach yields a symmetric and positive-definite system that has signifi-
cant computational advantages. That "the least-squares method needs extra
boundary conditions and produces an ill-conditioned matrix" is a misconcep-
tion. We have shown that the least-squares method does not need any ad-
ditional boundary condition and that the condition number of the algebraic
equations resulting from the least-squares method based on the first-order
equations has an order of h- 2 which is the same as in the classic Galerkin
method for second-order equations.
For purely convective problems a common technique in finite difference
methods is upwinding which introduces excessive dissipation and reduces the
accuracy of the approximate solution. Upwinding and free parameters turn
out to be unnecessary when the least-squares method is employed.
3. First-Order System in One Dimension
In fact, it can be shown (see, e.g., Carey and Oden 1983a, p.106) that the
solution pair {u, p} is a saddle-point of the functional (3.6).
Let us now study first the stability of the problem (3.5). Here, stability
means that the solution (u, p) depends continuously on the force term I,
that is, if the force varies a little, so does the solution. A natural stability
inequality for (3.5) would be the following: There is a constant C such that
if {u,p} E H x S satisfies (3.5), then
IIul11 + Ilpllo :=:; Clllll-1, (3.7)
where
(f,v)
11111-1 = O"/-vEH
sup -II-II (3.8)
v 1
in which "sup" denotes the least upper bound.
To be able to conclude (3.7), we need the Ladyzhenskaya-Babuska-Brezzi
(LBB) condition (Babuska 1971, Brezzi 1974): Given any q E S, there exists
a constant 'Y > 0 such that
34 3. First-Order System in One Dimension
(q, v')
sup-11-11-
O;lvEH v 1
~ 'Yllqllo, (3.9)
and v(O) = v(l) = 0 due to the fact that q satisfies the constraint fol q(e)de =
0, that is, v belongs to H. Obviously,
We have
IIvll~ = IIvll6 + IIv'lIo = I vll6 + IIqll6 :::; 211q1l6·
3.3 The Mixed Galerkin Method 35
Therefore,
(3.17)
That is, the LBB condition (3.10) indeed holds and 'Y = 1/../2.
To further understand the features of the mixed Galerkin method, we
consider a discrete solution of (3.5). A natural idea to obtain a discrete ana-
logue of (3.5) is now to replace Hand S by finite element subspace Hh and
Sh which satisfy the discrete LBB condition:
(3.18)
This gives the following mixed finite element method: Find Uh E Hh and
Ph E Sh such that
(Ph,qh) - (U~,qh) = 0 Vqh E Sh, (3.19a)
Introducing the basis functions {(Pl, ... , <Pn} and {'l/Jl, ... , 'l/Jm} for Shand
Hh , respectively, the discrete problem (3.19) can be written in matrix form:
(3.20)
where
... a 1n
. ° 0
.
:
)
,
... ann
We shall show that if the rank condition (3.21) holds, then (3.20) has a unique
solution. Here we need only to prove that, under the condition (3.21), the
solution of the homogeneous version of (3.20) is null. To do this we assume
that {P, U} is the solution of the corresponding homogeneous equation of
(3.20), then BT P = O. Multiplying the upper part in (3.20) by pT yields
pT AP _ pT BU = O.
Therefore,
pTAP=O.
Since A is positive-definite, we have P = 0, and thus BU = O. Due to
condition (3.21), U = O.
In the above we have proved that the rank condition (3.21) is a sufficient
and necessary condition for (3.20) to have a unique solution.
We now investigate the relationship between the rank condition (3.21)
and the following condition:
pT BU > O. (3.22)
If condition (3.21) holds, then the only solution of BU = 0 is U = O. If
U i= 0, then BU i= O. Therefore, there exists aPE m,n, such that (3.22)
is valid. Conversely, if (3.21) does not hold, i.e., n < m, then the equation
BU = 0 has a non-zero solution U. In such a case, for all P E m,n, we have
pT BU = 0, and thus (3.22) is not valid. Therefore (3.22) and (3.21) are
equivalent.
It is clear that if condition (3.21) does not hold, then pT BU = (qh, v~) =
0, i.e., the discrete LBB condition cannot be realized. If the rank condition
(3.21) holds, then pT BU = (qh, v~) > 0 which does not mean that the
discrete LBB condition (3.18) is verified. In other words the rank condition
(3.21) is only a necessary condition for the discrete LBB condition (3.18) to
hold. The rank condition (3.21) guarantees only the existence of the discrete
solution, but does not guarantee the stability of the solution. As a comment
we note that the parameter 'Y appearing in the condition (3.18) in general is
related to the mesh size h. Once the parameter 'Y in (3.18) is proved to be
unrelated to h, the error in the finite element solution will depend only on
the ability of the chosen finite element subspaces to approximate the exact
solution. However, for particular choices of Sh and H h , it is not an easy matter
to verify whether the LBB condition holds or not. There are different ways
in which arbitrarily chosen finite element spaces may fail to satisfy the LBB
condition. In particular, mixed methods with equal-order finite elements are,
in general, unstable in that the stability parameter 'Y tends to zero as h -+ O.
Now we investigate some particular finite element pairs of Sh and Hh . In
all cases let the interval [0,1] be divided into N cells (N > 3). An example of
instability is the following seemingly natural choice. For the v approximations,
we choose Hh = set of piecewise continuous quadratic functions. For discrete
3.4 The Least-Squares Finite Element Method 37
Au = (~1 ~1) ~: + (~ ~) u,
where X = {q E Hl(O, In
and H = {v E Hl(O,I) : v(O) = 0, v(l) = O}.
A necessary condition that u = {p, u} E X x H be a minimizer of the
functional I is that its first variation vanishes at u for all admissible v =
{q, v} E X x H. This leads to the least-squares variational formulation: Find
u = {p,u} E X x H such that
- 11
t-+O t
Vv = {q,v} E X X H. (3.26)
Let
B (u, v) = (Au, Av) = (p - u', q - v') + (p', q'), (3.27)
L(v) = (f,Av) = - (f,q'), (3.28)
then (3.26) can be written as
(Au,Av) = (f,Av) Vv = {q,v} E X x H, (3.29a)
or
B (u, v) = L(v) Vv = {q,v} E X x H. (3.29b)
We shall now analyze the least-squares method (3.29). The reader will fully
understand these analyses after studying the bounded inverse theorem in
Chap. 4. The analysis can also be based on the famous Lax-Milgram theorem
(see Appendix D). Formulation (3.29a) is convenient for the analysis based
on the bounded inverse theorem, while (3.29b) is popular for the application
of the Lax-Milgram theorem. For both approaches the essential issue in the
analyses is the same: the boundedness of the operator A or the coerciveness
of the bilinear form B (u, v). In the following we use both notations.
Apparently, B(u, v) is symmetric. It is straightforward to verify that
I(Au, Av)1 = IB(u, v)1 ~ Cllulll ·lIvliI, (3.30)
where lIull~ = IIpll~ + lIull~· Therefore A or B(·,·) is bounded on X x H.
Now let us prove that A is bounded below or B(v, v) is coercive: There
exists a constant a > 0 such that
IIAvll~ = B(v, v) ~ allvll~ Vv E X x H. (3.31)
We note that
B(v, v) = IIq - v'lI~ + IIq'II~· (3.32)
Consequently,
B(v, v) ~ IIq'II~, (3.33)
In the model problem (3.1) only the boundary conditions on u itself are
given. When the problem (3.1) is reduced to the first-order system (3.4)
by introducing the flux p, no boundary condition on P is required. This is
an important observation and is contrary to a common misunderstanding
that says "whenever a new variable is introduced to reduce a higher order
differential equation to a lower order system, a boundary condition on this
new variable is needed."
For the same problem with other boundary conditions such as:
(1) u(O) = 0, p(1) = OJ
(2) u(O) = 0, p(O) = O.
We can prove the boundedness (below) ofthe A following the same procedure
as above, see Jiang and Chang (1990).
The symmetry of the bilinear form (Au, Av) and the boundedness (be-
low) of A guarantee that the matrix for the least-squares finite element
method is symmetric and positive-definite. This is an important advantage
of the least-squares method.
Theorem 3.1 Assume f is smooth enough and the finite element interpola-
tion error estimates hold (see Sect. 4.6.3); that is,
lip -lIhPlll ~ CphllplHl, Ilu -lIhulll ~ Cuhmlulm+l, (3.49)
then
(3.50)
where f. and m denote the orders of polynomials for P and u, respectively;
r = min(f., m) and Cp , Cu and C are the constants which do not depend on
the mesh size h.
We may utilize the Aubin-Nitsche trick (see Sect. 2.9) to obtain the op-
timal L2-estimates of the error:
(3.51)
We remark that for the least-squares method the choice of interpolations
for P and u is not subject to any restriction as long as f. ~ 1, m ~ 1. In
particular, the equal-order finite elements are permissible. For example, if
the linear finite element is chosen, the accuracy of the least-squares solution
is of O(h2) for both P and u. However, inspection ofthe equation (3.4a) shows
that in order to make the residual of this equation equal to zero throughout,
one may choose the interpolation for u as one order higher than that for p,
such as a pair of quadratic (u) and linear (p) elements. As seen in Sect. 3.3,
this reasonable combination is not allowed in the mixed Galerkin method.
8
Problem 1
-- --
Pl'Obiem 2
8
Probtem 1
-- 8
Problem 3
/-
6 m=2 6 msl
g:
...g:
.~
...
I
I
4 4
2 2
Fig. 3.1. Computed convergence rates for one-dimensional problems (From Jiang
and Chang 1988)
Table 3.1. Computed convergence rates for model problems (From Jiang and
Chang 1988)
Fundamentals of LSFEM
4. Basis of LSFEM
the arcs are greater than zero (see, e.g., Oden and Reddy 1976, p.59 or Krizek
and Neittaanmiik 1990, p.4). A domain with a piecewise smooth boundary
represents a sufficiently wide class of domains necessary for most practical
purposes. Simple domains such as a ball, cube, torus, triangle, polygon, poly-
hedron, ... , etc., certainly have a piecewise smooth boundary.
As usual, Lp(D), 1 :::; p :::; 00 denotes the space of of functions u defined
on D whose absolute value have pth powers which are Lebesque-integrable
on D. The norm on Lp(D) is given by
The Lebesque space Lp(D) is a Banach spaces. For the purpose of this book
the spaces L1 (D) and L2 (D) are of particular importance.
A Banach space U is called a Hilbert space, if there exists an inner product
(., .) on U such that Ilullu = (u, u)1/2 for all u E U. We note that L1(D) is
not a Hilbert space.
L 2 (D) is a Hilbert space, since L 2 (D) is the space of square-integrable
functions defined over D and equipped with the inner product
For the function space H1({}) we may also introduce the semi-norm
lull= ( ~IIl}u
~ ~ 112) 1/2 . (4.4)
i=1 UXi °
Also we denote by H-I({}) the dual space consisting of bounded linear
functionals on HJ({}), i.e., u E H- I ({}) implies that (u, v) < 00 for all
v E HJ({}). A norm for H-I({}) is given by
(u,v)
lIull-1 = sup -I-I ' (4.5)
O~VEHJ(!J) v I
lIull~ = L Ilujll~·
j=1
(u,v) = l u·vdD.
50 4. Basis of LSFEM
Theorem 4.1 (The Bounded Inverse Theorem) The sufficient and necessary
condition for a linear operator A : U ~ V to have a continuous inverse
operator is that A is bounded below, Le., there exists a positive constant a
such that
VuEU. (4.11)
Proof: We follow the proof given by Chen (1982). Let us consider the two
dimensional case. The proof for the three dimensional case is similar. Assume
that u = 0 outside of ti, where ti denotes the closure of n.We use a circle
with diameter D to surround the domain n.
Let the boundary r l be located
as shown in Fig. 4.1 (this can always be done by rotating the coordinates),
such that the set of y coordinates of n contains a ~ y ~ (3.
We choose a point A on rl. The ordinate of A is y.
Through A we draw a straight line segment AB which is parallel to the
x axis. We have
lu(x,Y)1 2 = (i B
:~d~f ~ D i (:~)2d~.B
lu(x, y)12({3 - a) In
~ 2D (:~f dxdy + 2D({3 - a) i : (~~f dy.
Finally integrating the above inequality with respect to x and y yields the
Friedrichs inequality (4.14). 0
(4.18)
Proof: For simplicity let us prove its validity for two-dimensional rectangular
domains, see Fig. 4.2.
(O.b) 1---------------.
< 2{ ( r 2
8u(x, yd dX) 2
iXl ax
+ (lY2au~2,Y)dY)2}.
Yl Y
For the right hand side by using the Schwarz inequality we have
U2(X2' Y2) 2U(X2, Y2)U(XI. Yl) + U2(XI, Yl)
~ 2a la(aU~~YI)fdx+2b lb(aU~;,Y))2dY.
Integrating both sides of the above equation with respect to Xl, YI, X2, Y2
yields
56 4. Basis of LSFEM
diameter of all circles inscribed in the triangles. We shall assume that the
subdivision satisfies the standard regularity condition, Le., there is a positive
constant f3 independent of h such that
(4.21 )
This condition means that the triangles K E Th are not allowed to be ar-
bitrarily thin, or equivalently, the angles of the triangles K are not allowed
to be arbitrarily small; the constant f3 is a measure of the smallest possible
angle in any K E Th for any triangulation Th.
Consider a triangle in the mesh (Fig. 4.4). The nodes of this -triangle are
the vertices Ai(i = 1,2,3). Now we choose the linear interpolant
(4.22)
To evaluate the three constants ao, al and a2, we must provide three values
of u, x and y at each of the three nodes, i.e.,
4.6 Finite Element Spaces 59
Solving for the constants (aO,al,a2) and substituting them into (4.22) gives
Ihu(x, y) = '1/11 (x, Y)Ul + 'l/J2(X, Y)U2 + 'l/J3(X, Y)U3, (4.23)
where
'l/Jl (x, y) = D
1 det C :')
x
~ X2
X3 Y3
, (4.24a)
1 det
'l/J2(X,y) = D C
~
Xl
x Y"1 ) ,
x3 Y3
(4.24b)
'l/J3(X, y) = D1 det C Xl
~ x2
X ~
"1 ) , (4.24c)
in which
'l/J2 = area(A1AA3) ,
area(A1A2A3)
'l/J3 = area(A1A2A) .
area(A1A2 A3)
We see that
3
L'l/Ji(X,y) = 1. (4.25)
i=l
we follow the analysis presented by Ying (1988). This analysis has three steps.
First, the nodal value of U(Ai) is expressed by the Taylor expansion of u at any
point A(x, y). Next, the relation between the point-wise error u(A) - Ihu(A)
and the remainders of the Taylor expansion is established. Finally, the L2
and HI norms of the error are derived.
(1) We assume that u E H2(n). For any point A(x, y) within the element,
we have
U(Ai) = u(A) + 11 ! U(tXi + (1- t)x, tYi + (1- t)y)dt, i = 1,2,3. (4.26)
Let
~i = tXi + (1 - t)x, "'i = tYi + (1 - t)y,
then (4.27) can be written as
. 8u(A) 8u(A)
u(A) = u(A) + ---a;-(xi - x) + a:y(Yi - y) + Ri(A), (4.28)
in which
(4.29)
8 (A)
Ihu(A) = L {u(A) + T
3
i=1
(xi -
X
x)
(4.30)
4.6 Finite Element Spaces 61
Ihl(A) = L3 {
u(A)
8u(A)
+ -ax-(xi - x)
8u(A)
+ -8-(Yi -
}
y) 1/Ji(A).
i=l Y
The interpolant of a linear function is equal to itself, therefore
Ihl(A) = l(A) = u(A),
and
L {8u(A)
3
-ax-(xi - x)
8u(A)
+ -8-(Yi -
}
y) 1/Ji(A) = 0,
i=l Y
which together with (4.30) gives
3
Ihu(A) - u(A) = L Ri(A)1/Ji(A). (4.31 )
i=l
Equation (4.31) is the expression for the point-wise interpolation error Ihu-
u.
Next, we shall give the derivatives of the point-wise error. Differentiation
of (4.31) with respect to x leads to
Now we show that the last term in (4.32) is equal to zero. We differentiate
(4.28) with respect to x to obtain
82u(A) ( . _)
8x2 x, X +
8 2u(A) ( . _)
8x8y y, Y +
8Ri(A) _
8x - ,
°
therefore from the above equation we obtain
t 8~~A)
i=l
1/Ji(A) = - t {82;;~)
i=l
(Xi- X)+ 8;:~~) (Yi-Y) }1/Ji(A).(4.33)
Define a linear function A such that
A(A) = 0,
8A(A) 8 2u(A) 8A(A)
-ax-=- 8x2 ' 8y
62 4. Basis of LSFEM
2
Ih)"(A) = - L {a2u(A)
3 a u(A) }
ax2 (Xi - X) + axay (Yi - y) 'l/Ji(A). (4.34)
i=1
But
Ih)"(A) = )"(A) = 0,
which implies the right-hand sides of (4.34) and (4.33) equal zero, and hence
L
3
a~(A) 'l/Ji(A) = o.
i=1 ax
Therefore from (4.32) we obtain
3
a[lhu(A) - u(A)] = '"' ~(A) a'I/Ji(A) . (4.35)
ax L..J ax
i=1
Similarly,
=L
3
a[lhu(A) - u(A)] Ri(A) a'I/Ji(A) . (4.36)
ay i=1 ay
(3) Using (4.31) we can estimate lIu(A) -lhu(A)lIo,n•. To this end, we use
(4.31) and l'l/Ji(A) I ~ 1 to obtain
~ 3 L
3
i=1
1 R~(A)dxdy.
n.
Recall that
IXi -xl ~ h, IYi -yl ~ h,
which together with (4.29) give
+ 2 a2u(ei,Tli)( . _ )( . _ )
aeiaTli X, X y, Y
+ a2U~~lTli)(Yi_y)2]dtr
< Ch 4 {1 1
(1- t)(1 ~~ll + 2Ia~i2;TlJ + I~~ll)dt r·
4.6 Finite Element Spaces 63
Therefore
82u 12)
+ 1 8T/; dxdydt.
(4.37)
in which the constant J3 from (4.21) is absorbed into C. In fact, we have proved
the following theorem by summing (4.38) and (4.39) over all elements:
The estimates (4.40) and (4.41) are typical examples of estimates for the
interpolation error u - IIhu with piecewise linear functions. If we work with
piecewise polynomials of order r ~ 1 on the triangulation Th satisfying (4.21),
and u is smooth enough, we have the following estimates:
liu - IIhUlio,{l ::; Chr+1lulr+l,{l, (4.42)
(4.43)
For a proof of the above estimates, see, e.g., Ciarlet (1991) and Oden and
Carey (1983).
(~ aen) au
~Ak8 at + ... =/
k=l Xk <"n
in which, except for the first term, all terms depend only on the value of u
and its internal derivatives on the surface S, and thus are known. If au/aen
can be determined, then the solution can be extended off S. This depends
on whether the matrix E;=l Akaen/aXk is of full rank or not. If it is of full
rank, then the external derivative au/aen can be determined. Noting that
en = <p, if
n
det(2: Ak
k=l
a:a )k
= 0 (4.46)
66 4. Basis of LSFEM
then l is called the characteristic direction, and (4.47) is called the character-
istic equation. The surface S which satisfies (4.47) is called the characteristic
surface. Since (8cp/8xl, ... , 8cp/8x n ) represents the normal direction of the
surface cp = 0, the characteristic surface is a surface such that at every point
the normal is in the characteristic direction.
The classification of the system (4.45) depends on the solution of (4.47):
(1) If a!, ... , an-l are arbitrarily given real numbers,
°
there exist m real roots an satisfying (4.47), the system is hyperbolic;
(2) If det(L:Z=1 Akak) :/= for all non-zero real a!, ... , an, it is elliptic.
For large systems some roots may be complex and some may be real; this
gives a mixed system.
In the elliptic case all roots are complex; since complex roots occur as
pairs, the number of equations in an elliptic system must be even. In other
words a system with an odd number of equations cannot be elliptic.
We remark that the classification discussed here is in the ordinary sense.
A system which is not elliptic in the sense discussed here may still be elliptic
in the Agmon-Douglis-Nirenberg theory (Agmon et al. 1964).
Au = (4.49)
i=l
(4.53)
(4.57)
where (Ub U2, ... , um)j are the nodal values at the jth node, and h denotes
the mesh parameter.
Introducing the finite element approximation defined in (4.57) into the
variational statement (4.56), we have the linear algebraic equations
KU=F, (4.58)
where the U is the global vector of nodal values. The global matrix K is
assembled from the element matrices
(4.59)
in which ile C il is the domain of the eth element, and T denotes the
transpose, and the vector F is assembled from the element vectors
(4.60)
(4.61)
We remark that the boundary conditions (4.48b) can also be included into
the quadratic functional (4.55). In this approach, no boundary conditions are
imposed on the subspace V. This is another advantage of the least-squares
method.
4.9 The Euler-Lagrange Equation 69
For a given particular problem (4.48) we often can identify the subspace
V and its norm, see examples in Sect. 8.2.2. For a general problem (4.48)
without further information about the operator A, we do not know all details
of the space V and its associated norm lIuliv. But fortunately, if the operator
equation (4.48) is well-posed, we can at least know some thing about the
relations between IIAulio and lIullo.
(4.66a)
Moreover, if the solution u E H1(f1), then there exists a positive constant
M1 such that
(4.66b)
Theorem 4.7 (Errors of LSFEM for General Problems) If the linear first-
order differential system (4.48) is well-posed, and its solution is smooth
enough, then the following error estimates hold for the approximate solu-
tion by LSFEM:
IIR(uh)lIo ~ Clhrlulr+l, (4.72)
(4.73)
The proof of (4.77) follows from similar steps in the proof of Theorem
4.6. Using the Aubin-Nitsche trick as in Sect. 2.7.3 we can further obtain the
error estimate (4.78).
This theorem implies that for strictly elliptic problems the rate of conver-
gence of the least-squares method with equal-order finite elements is optimal
for all variables, since the rate predicted is as high as possible for a given
order of interpolation.
It should be noted that for strictly elliptic problems we do not assume
in advance that the solution uniquely exists. The coerciveness or the bound-
edness below in the H1 norm guarantees the existence and uniqueness of
the solution. For many problems in mathematical physics, the existence and
uniqueness of the solutions are often proved by using the Galerkin method.
The least-squares technique introduced in this section not only has practical
value in numerical solution, but can also be used for investigating the exis-
tence, uniqueness and stability of the solutions for some difficult problems,
such as the solvability of the incompressible Navier-Stokes equations with
non-standard boundary conditions (see Chap. 8).
is of great importance for the success of LSFEM, and from this knowledge to
establish a principle for choosing a proper order of Gaussian quadrature.
We begin by studying the least-squares method for linear algebraic equa-
tions.
knm Um in
or simply
Ku=/, (4.80)
where K, u and / denote the corresponding matrices in system (4.79). We
note that in system (4.79) the number of equations n may not be equal to
it)
the number of unknowns m. We define the augmented matrix as
k12
kim
k22 k2m 12
(4.81)
knm in
and the rank of a matrix as the order of the largest square array within that
matrix, formed by deleting certain rows and columns, whose determinant
does not vanish. Obviously, if n 2: m, the possible maximum rank of the
augmented matrix (4.81) is m + 1.
Depending on the properties of both the coefficient matrix K and the
augmented matrix, system (4.79) may have a unique solution, an infinite
number of solutions, or no solution at all. We shall investigate the conditions
under which the least-squares method can be used to find the solution to
system (4.79).
The least-squares method amounts to minimizing the following summa-
tion of the weighted squared residuals:
I(u) = Wi(kuUi + k12U2 + ... + kimum - id 2
+ W2(k2iUi + k22U2 + ... + k2mum - 12)2
+
+ Wn(kniUi + kn2U2 + ... + knmum - in)2, (4.82)
where (Wi> 0, i = 1, n) are the weighting factors. The minimizer of I is the
solution of the normal equation:
KTWKu = KTW/, (4.83)
74 4. Basis of LSFEM
This system has four equations and two unknowns. Moreover, the rank of
(113)
214
102
226
is three. Hence system (4.86) is overdetermined and has no solution. If (Wi =
1, i = 1, n) ~e chosen, the normal equation corresponding to (4.86) is
2) (;
2 1 0
2
~)
2
(Xl) = (1 2 1 2)
X2 1 1 0 2
(!)
2'
6
or
(4.87)
The least-squares solution to system (4.86), i.e., the solution to the normal
equation (4.87) is (1.545,1.364).
We emphasize that the satisfaction of the rank condition is important
for classifying a system as overdetermined. The rank condition guarantees
that m + 1 linearly independent equations can be found in the system. In
example (4.86), the first three equations are linearly independent, i.e., the
third equation cannot be obtained by a linear combination of the first two
equations.
If n > m, but the rank of the augmented matrix equals m, then the system
is determined. For example, the rank of the augmented matrix of
Xl +X2 = 3,
2Xl + X2 4,
4Xl + 2X2 = 8,
2Xl + 2X2 = 6 (4.88)
is 2 = m, therefore (4.88) is determined. The rank of the augmented matrix
of
Xl +X2 3,
2Xl + 2X2 = 6,
3Xl + 3X2 9,
4Xl + 4X2 12 (4.89)
is 1 < m = 2, therefore system (4.89) is underdetermined.
In summary, the least-squares method can be used only for the solution
of determined and overdetermined systems of linear algebraic equations for
which n ~ m and rankaug = m or m + 1.
76 4. Basis of LSFEM
U1)
R(x,y) = E(
Nn
j=l
a:
a'lj;· a'lj;·
A1 + TA2 +'Ij;jAo)
Y
( U2
:
.
-f· (4.91)
Urn j
{ ~(a'lj;,a: + a:
~ Al a'lj;' A.+"';Ao) .:.(:~) - J} ~ 0. (4.92)
J ({c.1)c)
The unknown nodal values can be found by solving the resulting system of
linear algebraic equations (4.92). In order to obtain a determined system, we
should choose a proper number of interior collocation points in each element
such that the total number of equations is equal to the total number of un-
known nodal values. Since the resulting system of algebraic equations (4.92)
by the collocation method is sparse but not symmetric, it is difficult to find
its solution.
Instead of solving the residual equations (4.92) directly, we may use the
least-squares method discussed in the above section to obtain the solution.
4.11 Implementation of LSFEM 77
We now collocate at Neal interior points in each element and minimize the
summation of weighted, squared residuals:
ideas and maintain simplicity, we often further assume that the domain n is
convex or simply connected, although these restrictions are not necessary in
many cases.
In this book, we use the symbols V, V" Vx and Ll to denote the gradi-
ent, divergence, curl and Laplacian operators, respectively. We also use the
following notations:
(u,v}r = l uvdr,
(u,v}r = l u· vdr.
When there is no confusion, we will often omit the measure r from these
inner products.
Lemma 5.1 Let n be a bounded open subset of IR3 with a piecewise smooth
boundary r. Then every function u of [Hl (n)p with n x u = 0 on r satisfies
l
lul~ + (~l + ~Ju, udr = IIV· ull~ + IIV x ull~, (5.1)
Proof: In this case we still have (5.2). Since n· u = 0, we may assume that
u = UT on r.
By virtue of the triple scalar product
(V x u) . (n x u) = n· (u x (V x u))
and using (A.5), the boundary integral terms in (5.2) can be written as
Theorem 5.1 Let rJ be a bounded and convex open subset of R3. Then
every function u of [HI(rJ)j3 with n· u = 0 on and n x u = 0 on r 2n
satisfies:
(5.4)
84 5. Div-Curl System
Theorem 5.2 (The Div-Curl Theorem) Suppose that n is a (1) bounded and
convex, or (2) bounded and simply connected subset of JR3. If u E [Hl{n)j3
satisfies
'V·u=O inn,
'Vxu=O inn,
n·u =0 on Fl ,
n xu= 0 on F2 ,
then
u == 0 in n.
Remark. We distinguish between case (1) and case (2) only for convenience.
In fact, case (2) includes case (1), because a convex domain must be simply
connected.
Proof: Case (1). Theorem 5.2 indicates that the homogeneous div-curl system
with homogeneous boundary conditions has only a trivial solution. In other
words the div-curl system under the given boundary conditions has a unique
solution. Therefore, by virtue of Theorem 4.6, there exists a positive constant
a such that
5.1 Basic Theorems 85
This theorem indicates the fact that for the function space
H = {[H 1 (D)]3: n· u = 0 on n,n x u = 0 on r 2}
lIull~ and (IIV . ull~ + IIV x ull~) are equivalent norms for an appropriate
domain .0.
The proof of Theorem 5.3 can also be based on the use of contradiction
arguments together with Theorem 5.1, see e.g., Saranen (1982) and the refer-
ences therein. In the two-dimensional case, a direct proof is available (Krizek
and Neittaanmaki 1984a or 1990).
9= 0 on r 1 # 0 (or on r 2 # 0),
then
9 == 0 in .0.
i i
and from the Stokes theorem we have a contradiction:
The proof of Theorem 5.5 can also be found in Pironneau (1989, p.53).
n·u=O on r I , (5.8c)
nxu=O on r 2, (5.8d)
where the given vector function (.oj E [L2(il)j3 cannot be arbitrary; it must
satisfy the following solvability conditions:
\7·w=O in il, (5.9a)
n·w=O (5.9b)
l n·wdr=O. (5.9c)
If r2 is empty, then the given scalar function p E L2(il) must satisfy the
solvability condition:
fnPdil = o. (5.9d)
n·u=O (5.10c)
5.2 Determinacy and Ellipticity 87
(5.lOd)
nxu=O (5.10e)
Notice that we impose 79 = 0 on n, and do not specify any boundary condi-
tion for the dummy variable 79 on r 2 •
By virtue of Theorem 5.2, the vector equation (5.lOa) is equivalent to the
following equations and boundary conditions:
\1 x (\179 + \1 x u - w) =0 in il, (5.lIa)
\1 . (\179 + \1 x u - w) =0 in il, (5.lIb)
n x (\179 + \1 x u - w) = 0 on rl. (5.lIc)
n· (\179 + \1 x u - w) = 0 onn. (5.lId)
Taking into account the solvability conditions (5.9a) and (5.9b), the boundary
condition (5.lOe) and Theorem 5.5, from (5.lIb), (5.10d) and (5.lId) we have
Ll79=0 in il, (5.12a)
79 = 0 on n, (5.12b)
079 = 0 (5.12c)
on
From (5.12) we know that 79 == 0 in il. That is, the introduction of 79 into
(5.8) does not change anything, and thus system (5.10) with four equations
and four unknowns is indeed equivalent to system (5.8).
We cannot classify system (5.8), since the classification introduced in Sect.
4.7 requires that the coefficient matrices of a first-order system be square. But
now we can classify system (5.10). In Cartesian coordinates the equations in
system (5.10) are given as
079
-
ow OV
ox+oy- - oz
- = Wx ,
079
-
OU ow
oy +oz- -ox
- = wy ,
-
079
+---
ov OU
wz ,
OZ ox oy
ou ov ow
ox + oy + oz p. (5.13)
We may write system (5.13) in standard matrix form:
ou OU OU
Ai 0-; + A2 0; + A3 0-; + AoY = /, (5.14)
in which
88 5. Div-Curl System
A,~ 0 0 0
0 -1
1 0
0 0
D. A, ~ (~1 10) oo
o
0 1
0 0
1 0 0
'
-1 0 0 0
A3~G D·~~G D·
0 0 0 0
0 0 0 0
0 1 0 0
f~ (~). ~~G)·
The characteristic polynomial associated with system (5.13) is
= (e + 772 + (2)2 #= 0
for all nonzero real triplets (e, 77, (), system (5.10) is thus elliptic and properly
determined.
The first-order elliptic system (5.10) has four equations in four unknowns,
so two boundary conditions on each boundary are needed to make system
(5.10) well-posed. Here {} = 0 and n· u = 0 serve as two boundary conditions
on Ft; while n x u = 0 implies that two tangential components of u are zero
on r2 •
Since system (5.8) is equivalent to system (5.10), and system (5.10) is
elliptic and properly determined, so is system (5.8).
Remark. In fact, the solvability conditions (5.9a), (5.9b) can be obtained
by applying the div-curl method (see the next section) to (5.8a).
n . (\7 X u - w) = 0 (5.15d)
n·u=O (5.19c)
n x (V x u) = n x w (5.19d)
n xu = 0 on r2, (5.1ge)
\7 . U = P on r2· (5.19f)
90 5. Div-Curl System
Theorem 5.6 The solution of (5.8) or (5.22) uniquely exists and satisfies:
Ilulh ~ C(llwllo + Ilpllo). (5.23)
Theorem 5.7 The LSFEM based on (5.22) has an optimal rate of conver-
gence and an optimal satisfaction of the divergence equation:
Ilu - uhllo ~ C1hr+1l1 ullr+b (5.24a)
Taking into account the vector identity (5.18) and that v satisfies n . v =0
on nand n x v = 0 on r2, we obtain from (5.26)
(-Llu - \7 x w + \7 p, v)
-(n x (\7 x u - w),v)r1 + ((\7. u - p),n· v)r2 = 0 (5.27)
for all admissible v E V, hence we have the Euler-Lagrange equation and
boundary conditions:
Llu = - \7 x w + \7 p in il, (5.28a)
n·u=O on rl , (5.28b)
n x (\7 x u) = n xw on rl , (5.28c)
nxu=O r2 ,
on (5.28d)
\7·u=p on r2 . (5.28e)
(5.8); (2) the least-squares variational formulation (5.22); (3) the uncoupled
Poisson equations (5.28); and (4) the Galerkin formulation (5.27).
It turns out that the least-squares method (5.22) for the div-curl equa-
tions (5.8) corresponds to using the Galerkin method (5.27) to solve system
(5.28) which consists of three independent second-order Poisson equations
(5.28a) and three coupled boundary conditions on each boundary, where the
original first-order equations (5.28c) and (5.28e) serve as the natural bound-
ary conditions, and (5.28b) and (5.28d) as the essential boundary conditions.
The least-squares method (5.22) is the simplest approach among these
equivalent methods, because (1) the sources wand p need only to be square
integrable; (2) it does not need any additional boundary conditions; the trial
function u and the test function v need to satisfy only the original essential
boundary conditions; (3) the corresponding finite element method has an
optimal rate of convergence and leads to a symmetric positive-definite matrix.
These are the reasons why we strongly recommend the least-squares method.
Proof: If we can actually find q and 'I/J, this theorem is proved. By virtue of
Theorem 5.2, (5.29) is equivalent to the following equations and boundary
condition:
'\1 . ('\1 q + '\1 x 'I/J - u) = 0 in !1, (5.30a)
'\1 x ('\1q + '\1 x 'I/J - u) =0 in !1, (5.30b)
n· ('\1q + '\1 x 'I/J - u) = 0 on r. (5.30c)
Taking into account '\1 . '\1 x 'I/J = 0 and '\1 x '\1q = 0, system (5.30) can be
written as follows:
Llq = '\1 . u in !1, (5.31a)
'\1 x ('\1 x 'I/J) = '\1 x u in !1, (5.31b)
94 5. Div-Curl System
n . (V q + V X 1/J) = n . u on r. (5.31c)
To obtain q we may solve the following Poisson equation with the Neumann
boundary condition:
Llq = V . u in n, (5.32a)
n . Vq = n . u on r, (5.32b)
where the boundary condition (5.32b) is additionally supplied. Although q is
not unique, i.e., an arbitrary constant can be added into q, Vq is uniquely
determined.
Now 1/J should satisfy
V . (V x 1/J) = 0 in n, (5.33a)
V x (V x 1/J) =V x u in n, (5.33b)
n . (V x 1/J) = 0 on r. (5.33c)
V x 1/J in system (5.33) may be considered as an unknown vector that can be
uniquely determined by the least-squares method described in Sect. 5.4.
Finally, we can solve the following div-curl system to obtain 1/J:
V·1/J=O inn, (5.34a)
V x 1/J = given in n, (5.34b)
nx1/J=O onr. (5.34c)
Therefore the validation of the decomposition (5.29) is proved.
Using Appendix (B.4) and (5.33c) we find
(Vq, V x 1/J) = (n· (V x 1/J),q}r = 0,
that is, V q and V x 1/J are orthogonal. o
We remark that Theorem 5.9 is valid also for bounded and multiply con-
nected domains (Girault and Raviart 1986).
In this chapter we study and compare different finite element methods for
the solution of two- and three-dimensional diffusion-type problems usually
governed by second-order elliptic partial differential equations. In the mixed
Galerkin method, the second-order scalar equation is decomposed into a first-
order grad-div system by introducing additional variables (the fluxes). The
conventional least-squares finite element method is also based on the same
grad-div system. We shall present theoretical analysis and numerical results
to show that this simple procedure of reduction destroys ellipticity and thus
the conventional LSFEM is not optimal, that is, the rate of convergence
for the fluxes is one order lower than optimal. In order to have an optimal
LSFEM, the div-curl-grad system should be employed.
S = {v E [Hl({})td : n x v = 0 on F},
98 6. Div-Curl-Grad System
W = {v E [L 2 (D)t d },
The most commonly used method for probleql. (6.1) is the Rayleigh-Ritz
method. However, as indicated in Sect. 3.2 for one-dimensional problems, a
posteriori numerical differentiation is required to obtain the dual variables
(flux for heat transfer; velocity for fluid flows; or electric field intensity for
electrostatics) which are often of most interest. In general, the accuracy of so-
computed dual variables is one order lower than that of the primal variable.
Moreover, the computed dual variables are not continuous across the element
boundary.
6.2 The Mixed Galerkin Method 99
c/>=o on r. (6.3c)
Now multiplying (6.3a) by v E W and integrating, multiplying (6.3b) by
'l/J E H and integrating by parts, we are led to the mixed Galerkin variational
statement: Find a pair {c/>, u} E H x W such that
(u, v) - (\lc/>, v) = 0 'Vv E W, (6.4a)
sup (
Oi-t/>EH
r\lc/>. udn)(Ic/>h)-l ~ 'Yllullo
In
'Vu E W, (6.5)
where the constant 'Y > o. The LBB condition precludes the application of
simple equal-order finite elements. It can be proved that the finite element
spaces Hh and Wh satisfy the discrete LBB condition (6.5), and if the solution
(c/>, u) of (6.3) belongs to Hr+1(n) x [Hr(il)]nd, we have the following error
estimate (see, e.g., Roberts and Thomas 1991, p.578):
(6.6)
The estimate (6.6) shows that in this mixed method, the accuracy of the flux
u is always one order lower than that for the primal variable c/>.
Inspecting equation (6.4) shows that the matrix associated with the mixed
method is non-positive-definite. This makes the use of iterative methods to
solve large-scale problems relatively difficult.
100 6. Div-Curl-Grad System
(f,AV) = (-/,V·v).
The corresponding finite element problem is then to find Uh = {<Ph, Uh} E
Hh x M h , such that
(6.9)
where
Theorem 6.1 Assume that f(a;) E L 2 (il), the solution (¢, u) of (6.3) belongs
to Hr+1(il) x [Hr+l(il)]nd, and the finite element interpolation estimates
(6.2a), (6.2b) and (6.2f) hold. Then for the approximate solution associated
with (6.9), we have the error estimate:
(6.21)
(6.22c), and (6.22e) can be derived by (6.22d). We shall show that the inclu-
sion of these relations is important.
We note that the calculation of ¢ and u in (6.22) can be decoupled. We
may solve the div-curl system of (6.22a) and (6.22b) with the boundary
condition (6.22e) to obtain u, and then (if necessary) to find ¢ by using
numerical integration or by solving (6.22c) and (6.22d) in terms of LSFEM.
For more general problems such as
-\1. \1¢ + ¢ = f(x) in il,
8¢ =0
8n on ,
r
this decoupling is impossible (Jiang and Povinelli 1993).
e D' C1 0 ~1 ).
in which
0 0
At~ -
1
0 0
0
1 A2 =
-1
0
0 0
0 0
0 0 0 o 0 1
Ao = CO o
-1
o -1
0
0
oo0
o 0
o 0
0)
' f~cn, Y~m·
Since
0
~ det ( ~q ~ ) ~ (e + q')' '" 0
rJ
~ 0
det(At{ + A,q) 0 ~ -rJ
0 rJ ~
6.4 The Optimal LSFEM 105
for all nonzero real pairs (~, TJ), system (6.24) and thus the two-dimensional
version of system (6.22) is determined and elliptic, as contended. The elliptic
first-order system (6.24) has four equations involving four unknowns, so two
boundary conditions (6.24e) and (6.24f) are required.
Remark. In this chapter we analyze only the boundary condition (6.lb). If
equation (6.la) is supplemented by the boundary condition
a¢
an = o. on ,
r
then the following boundary conditions
'19 = 0 on r, (6.26a)
n·u=O on r, (6.26b)
should be specified for (6.23a)-(6.23c).
\1'19 + \1 x u = 0 in n, (6.27b)
\1xX+\1¢-u=O in n, (6.27c)
\1·X=O in n, (6.27d)
¢=O on r, (6.27e)
n·x=O on r, (6.27f)
nxu=O on r, (6.27g)
where u = (u,v,w)j while '19 and X = {XbX2,X3} are dummy variables. In
system (6.27) there are eight equations and four algebraic boundary condi-
tions involving eight unknowns.
In the following we use the div-curl method discussed in Sect. 5.3 to
show that the dummy variables are identical to zero. For simplicity, some
details have been omitted. By applying the divergence operation to (6.27b)
and considering \1 . \1 x u = 0 we have
Ll'19 = 0 in n.
By applying the n· operation to (6.27b) on the boundary and considering
(6.27g) and Theorem 5.5, we have
n . \1'19 = 0 on r.
106 6. Div-Cud-Grad System
1 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0
0 0 -1 0 0 0 0 0
0 1 0 0 0 0 0 0
AI= 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 -1
0 0 0 0 0 0 1 0
0 0 0 0 0 1 0 0
0 1 0 0 0 0 0 0
0 0 1 0 0 0 0 0
0 0 0 0 1 0 0 0
-1 0 0 0 0 0 0 0
A2= 0 0 1
,
0 0 0 0 0
0 0 0 1 0 0 0 0
0 0 0 0 0 -1 0 0
0 0 0 0 0 0 1 0
0 0 1 0 0 0 0 0
0 -1 0 0 0 0 0 0
1 0 0 0 0 0 0 0
0 0 0 0 1 0 0 0
A3= -1 0
,
0 0 0 0 0 0
0 0 0 0 0 1 0 0
0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 1
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
Ao= ,
-1 0 0 0 0 0 0 0
0 -1 0 0 0 0 0 0
0 0 -1 0 0 0 0 0
0 0 0 0 0 0 0 0
-1 u
0 v
0 w
0 ¢
/= 0
!!= 79
0 Xl
0 X2
0 X3
108 6. Div-Curl-Grad System
Since
e TJ ( 0 0 0 0 0
0 -( TJ 0 e 0 0 0
( 0 -e 0 TJ 0 0 0
det(Ale + A 2TJ + A 3() = det
-TJ e 0 0 ( 0 0 0
0 0 0 e 0 0 -( TJ
0 0 0 TJ 0 ( 0 -e
0 0 0 ( 0 -TJ e 0
0 0 0 0 0 e TJ (
= -(e + TJ2 + (2)4 # 0
for all nonzero real triplets (e, TJ, (), systems (6.32), (6.27) and thus system
(6.22) is properly determined and elliptic. The elliptic first-order system
(6.27) has eight unknowns, so that four boundary conditions are needed.
Considering that two algebraic conditions are included in (6.27g), this re-
quirement is satisfied in (6.27).
Remark. If equation (6.1a) is supplemented by the Neumann boundary con-
dition:
8¢ = 0
8n
on r,
then for three-dimensional problems the following boundary conditions
{} = 0 on r, (6.34a)
n x X =0 r,
on (6.34b)
on r, (6.34c)
should be given to (6.27a)-(6.27d). If the original second-order problem (6.1)
:!
has a mixed boundary condition:
¢ + f3 = 0 on r,
then (6.34c) should be replaced by
¢ + f3n . u =0 on r.
6.4.3 Error Analysis
Theorem 6.2 Assume that /(x) E L 2([}), the solution (c/J,u) E Hr+I([}) x
[Hr+1([})]nd and the finite element interpolation estimates (6.2a)-(6.2d)
hold. Then for the approximate solution associated with (6.37), we have the
error estimate:
(6.41)
110 6. Div-Cud-Grad System
This theorem implies that the rate of convergence (in the H1 norm) of
LSFEM based on the first-order div-curl-grad system (6.22) with equal-order
finite elements is optimal for all variables. Optimal L2 convergence can be
obtained by using the Aubin-Nitsche trick. The optimality is attributed to
the fact that the optimal LSFEM controls the errors in the full H1 norm.
As discussed in Sect. 4.11, the LSFEM with numerical quadrature is equiv-
alent to a weighted collocation least-squares method. We may use this idea
to choose an appropriate number of Gaussian points. The conventional LS-
FEM with linear (01) elements and one-point quadrature will lead to a sin-
gular global matrix, because it corresponds to solving an underdetermined
algebraic system. In contrast, the optimal LSFEM with reduced integration
works very well; although the computed nodal values of u may have some
oscillations, the values of u at Gaussian points are very smooth. The use of
reduced integration is important in practice, because it significantly reduces
the computing work when the matrix-free element-by-element conjugate gra-
dient method is employed for the solution of the resulting system of algebraic
equations.
4.0
3.5 _--4.
_--1111
II
•
___I ~
u-optlmal
~-optlmal
& - -.... u-c:onventlonal
-c:onventlonal
3.0
......,
j2.5
I
2.0
1.5
1.0
0.5 0.7 0.9 1.1 1.3 1.5
-Iog(h)
av_au_ o (7.2b)
ax By - ,
with the boundary conditions
v = sin (x) on y = 0,
u = eYcos(1) on x = 1, (7.3)
v = esin(x) on y = 1 ,
u == eY on x = O.
Combining the equations in (7.2), we see that this example corresponds to
solving the Laplace equation for the potential ¢ in the unit square. For ex-
ample, this problem describes incompressible potential flow in the square
domain with the normal derivative a¢/an prescribed on the boundary. The
analytic solution to the model problem is
u = eYcos(x), v = eYsin(x). (7.4)
This is now used with the approximate solution to calculate the norm of the
error in a sequence of mesh refinement studies.
The domain is divided into uniform bilinear elements. In the computation
we are faced with the problem of evaluating an integral of the form (4.59)
for the element matrices. The integrand consists of the squares of the first-
order derivatives of the shape functions. Since bilinear elements are used,
the integrand is a polynomial of degree 2, and hence the integral can be
evaluated exactly with 2 x 2 Gaussian quadrature. However, as explained in
Sect. 4.11.2, the LSFEM with Gaussian quadrature is equivalent to the least-
squares finite element collocation method, and hence here one-point Gaussian
quadrature should be employed. From this example, we understand that in
LSFEM computations, in general, reduced integration should be used instead
of full (exact) integration. For comparison purposes we have tested both one-
point and 2 x 2 Gaussian quadrature in the computation.
The L2 norm of the error is plotted against the mesh size h in a log-log
plot in Fig. 7.1. We see that the error is O(h2), which is consistent with the
optimal rate predicted by the estimate (5.24a) with this choice of elements.
We remark that, although the estimate (5.24a) is derived for the case with
homogeneous boundary conditions, it is not very difficult to show its validity
for inhomogeneous boundary conditions.
Both full integration with 2 x 2 Gaussian quadrature and the reduced
integration with one-point Gaussian quadrature produced the same rates, as
shown in Fig. 7.1. The reason is as follows: first, the coefficients in system
(7.2) are constants; second, we used uniform square elements which are not
distorted and thus the Jacobian of the coordinate transformation in (4.94)
is a constant; and more important, there are no zero-order terms in system
(7.2), i.e., system (7.2) is of homogeneous first-order type. The consequence
is that the equation can be satisfied everywhere in an element, and hence
an increase in the order of quadrature does not mean the inclusion of more
7.1 Incompressible Irrotational Flow 117
3.5
2.D~~_L-~_L-~_L-_L_L-_L~
Fig. 7.1. Experimental rates of convergence of LSFEM for the model problem:
6. 2 x 2 Gaussian quadrature; x one-point Gaussian quadrature
1.0
0.5
a.
()
I
0.0
-0.5
-1.0
0.0 0.2 0.4 0.6 0.8 1.0
X
Fig. 7.3. Pressure coefficient distribution for NACA0012 airfoil at a = 2°
2.0
1.0
a.
()
I
0.0
(7.6a)
Since the flow is also irrotational, the velocity components are related by
8p = _p~(q2)/a2. (7.8a)
8x 8x 2
Similarly we may obtain
8p 8 (q2) 2
8y = -P8y "2 /a , (7.8b)
8p = _p!....(q2)/a 2. (7.8c)
8z 8z 2
Substituting the above three equations into (7.6a), we obtain
7.2 Subsonic Compressible Irrotational Flow 121
(7.10b)
where
0;2 = 1 + ,~1 (Moo 2 _ if.2), (7.lOc)
in which the free stream Mach number Moo = qoo/a oo .
It is also appropriate here to introduce some useful local dimensionless
parameters for the fluid flow. Using a~ = ,Pool Poo and 15 = pi Poo and (7.9b),
we obtain the dimensionless local density
, - 1 2 2 ] 1/(1-1)
15= [ l+-(Moo -if.) (7.11)
2
By virtue of (7.10c), the local Mach number can be expressed as
Now let us apply the LSFEM to the solution of plane steady compressible
flow problems. The fundamental dimensionless equations for two-dimensional
steady subsonic irrotational flows are
8v_8u=O in n
H, (7.14b)
8x 8y
where a is the local sound speed normalized by the sound speed of the free
stream, and for convenience the dimensionless bar notation is suppressed.
Usually, on the solid surface the normal component of the velocity is zero,
that is,
n·u=O on r. (7.15)
The uniform condition u = (Uoocoso:, Uoosino:) is prescribed on the outer
boundary.
The quasi-linear first-order partial differential equations (7.14) may be
linearized, for example, by using the successive substitution method. Let
A = (1 _ U22 )
B = (1 _
V2) = _ uv2 C (7.16)
a ' a2 ' a •
Coefficients A, B, C are regarded as known quantities and are determined by
the velocity field solved in the preceding iteration step. Then system (7.14)
can be rewritten as
8u C(8u 8v - 0
A -+
8x
- +8V)
-
8y 8x
+B-
8y
- in il, (7.17a)
8v_8u=O in n
H. (7.17b)
8x 8y
Now we write system (7.17) in standard matrix form:
8u· 8u
A l8x + A2 8y + Aou = f in il, (7.18)
in which
Al = (~ ~),
Ao=(~ ~),
7.2 Subsonic Compressible Irrotational Flow 123
Since
AB1/Ji,x1/Jj,y
+AC(1/Ji,x1/Jj,y + 1/Ji,y1/Jj,x) +AC1/Ji,x1/Jj,x + BC1/Ji,y1/Jj,y
+(C2 + 1)1/Ji,y1/Jj,y +(C2 - 1)1/Ji,y1/Jj,x
In. AB1/Ji,y1/Jj,x
dxdy.
(::)=c(:), (7.20a)
= (7.21)
Kil Ki2 Kii Kin Ui Fi
U i = (:) / Fi = (~:)i
are the velocity components and the components of the 'force' vector at the
ith node respectively. For the homogeneous equations (7.14), the components
of the 'force' vectors in system (7.21) are zero. In general, they are not zero.
For nodes on the inclined solid surface, the normal and the tangential
velocity components are taken as the unknowns. By considering (7.20b), we
can rewrite system (7.21) as
Kll K12 KliC Kin Ui Fi
K2i K22 K 2i C K 2n U2 F2
U":l (7.22)
Kil Ki2 KiiC Kin Fi
U: = (un) .
Ut i
U":l =
CK il CK i2 CKiiC CKin CF i
7.2.3 Examples
In the following two numerical examples, the flow field obtained by solving
the corresponding incompressible problem was taken as an initial field. At
all nodal points of the field, the solution was considered convergent if the
relative variation of two successive values of the dimensionless density p =
[1 + ~(M! - (u 2 + v2))]1/b- 1) is less than a given small quantity f..
Flow Past Circular Cylinder. The flow condition at seven times the ra-
dius of the cylinder was considered undisturbed from the free stream. One
quarter of the circular cylinder was divided into 4 x 5 eight-node quadrilateral
elements. Reduced integration with 2 x 2 Gaussian quadrature was used. f.
was taken as 0.0009. The solutions at various Moo including the critical Mach
number 0.42 depicted in Fig. 7.5 were obtained in 2 '" 3 linearizations.
1.2
1.0 y
0.8
M 0.6
&~.
0.4
0.2
18 36 S4 72 90
0'
Fig. 7.5. Mach number on the surface of the circular cylinder at upstream Mach
number of 0.42, 0 LSFEM, - results from Imai (1941) (from Jiang and Cai 1980)
Flow Past NACA0012 Airfoil. The flow around the symmetrical NACA
0012 airfoil with zero angle of attack was calculated. The upper half region of
the airfoil was divided into 5 x 12 eight-node quadrilateral elements and f. =
0.0001 was taken as the convergence criterion. In the computation reduced
integration with 2 x 2 Gaussian quadrature was used. The solution at the
critical Mach number 0.72 was obtained after six linearizations and is shown
in Fig. 7.6.
The LSFEM has also been tested for lifting cases. Figure 7.7 and Figure
7.8 show the pressure distribution for the subsonic flow on the NACA0012
airfoil at 3.5 0 incidence and Moo = 0.5, and 100 incidence and Moo = 0.5,
126 7. Inviscid Irrotational Flows
Il~---------------
LI
1.11
0.7
NG.6
0.5
0.4
G.3
G.2
0.1
0.°0 U
' "
G.2 G.3 ~ 0.5 G.6
I
V
I
U
I
U U
¥jc
Fig. 7.6. Mach number on the surface of the NACA0012 airfoil surface at 00
incidence and upstream Mach number of 0.72, 0 LSFEM, - results from Lock
(1970) (from Jiang and Cai 1980)
2.0
1.0
a.
()
I
0.0
-1.0
-2.0
0.0 0.2 0.4 0.6 0.8 1.0
X
Fig. 7.7. Pressure distribution on the NACA0012 airfoil surface at 3.50 incidence
and upstream Mach number of 0.5
7.3 Concluding Remarks 127
respectively. In these computations, the flow domain and mesh, the number
of integration points, the boundary conditions and the Kutta condition are
the same as those used for the incompressible case. Converged solutions are
obtained after four linearizations.
6.0
4.0
Co
()
I
2.0
0.0 : : : : : : : : : :
:
-2.0
0.0 0.2 0.4 0.6 0.8 1.0
X
Fig. 7.8. Pressure distribution on the NACA0012 airfoil surface at 10° incidence
and upstream Mach number of 0.3
Fletcher (1979) first used a LSFEM based on the compressible Euler equa-
tions to solve subsonic flow problems. The feature of his formulation is the
designation of groups of variables rather than single variables.
8. Incompressible Viscous Flows
The variational statement (8.2) can be obtained directly from the station-
ary condition for the Lagrangian
1
L(u,p) = "2a(u, u) - (f, u) - (V· u,p). (8.3)
For the Stokes equations, Hughes and his colleagues put additional least-
squares terms into the mixed Galerkin functional (8.3):
132 8. Incompressible Viscous Flows
1
LM/L(U,P) = "2a(u, u) - (I, u) - (\7 . u,p)
where i'J denotes element interiors, i' denotes element interfaces, H is the
jump operator. a and (3 are nondimensional stability parameters, such that
for small parameters this new functional (8.6) has the desired coerciveness
over the finite element space H h X Sh in the sense that as Ilpllo --t 00 and
Ilull! --t 00, LM/L(U,P) = +00. This coerciveness implies unique solvability
of the problem without imposition of the LBB condition. Because it involves
least-squares terms, this method is named the mixed Galerkin/least-squares
method (Franca et al. 1989). This method is attractive because one obtains
convergence for arbitrary combinations of velocity and pressure interpola-
tions. However, when it is extended to solve the Navier-Stokes equations,
the resulting matrix is nonsymmetric and thus difficult to solve by iterative
methods for large-scale problems.
We shall show that system (8.7) is really properly determined and elliptic
by using the same technique as discussed in Sect. 5.2. To proceed we may
consider the following first-order system for the Stokes equations:
\7p+ \7 x w = f in n, (8.8a)
\7 . w =. ° in n, (8.8b)
-w + \7 ¢ + \7 x u = 0 in n, (8.8c)
\7 . U = ° in n. (8.8d)
° °
Here we have introduced a dummy variable ¢ in (8.8c) , which satisfies the
r.
boundary condition ¢ = on Substituting (8.8c) into (8.8b) yields iJ.¢ = 0,
n.
thus ¢ == in It means that the introduction of ¢ does not change anything.
However, now there are eight unknowns and eight equations in (8.8). System
(8.8), and hence system (8.7) is indeed properly determined.
We note that in some cases specification of the boundary condition for
the dummy variable ¢ is unnecessary, and \7¢ == 0 can also be guaranteed,
see Sect. 8.2.2.
Now let us classify system (8.8). In Cartesian coordinates, system (8.8) is
given as
op oW z ow y - f
ox + oy - OZ - x,
op + oW x _ oW z - f
oy oz ox - y,
OWx oW y oW z _ 0
ox + oy + OZ - , (8.9)
o¢ ow OV
-w x +-+----0
ox oy oz - ,
o¢ ou ow
-w
y
+ -oy + -OZ - -ox = 0,
o¢ ov ou
-W z +- +- - - = 0,
oz ox oy
OU ov ow_ O
ox + oy + OZ - .
134 8. Incompressible Viscous Flows
0 0 0 0 0 0 1 0
0 0 0 0 0 0 0 1
0 0 0 0 -1 0 0 0
0 0 0 0 0 1 0 0
A2= 0 0
,
0 0 1 0 0 0
0 0 0 1 0 0 0 0
-1 0 0 0 0 0 0 0
0 1 0 0 0 0 0 0
0 0 0 0 0 -1 0 0
0 0 0 0 1 0 0 0
0 0 0 0 0 0 0 1
0 0 0 0 0 0 1 0
A3= ,
0 -1 0 0 0 0 0 0
1 0 0 0 0 0 0 0
0 0 0 1 0 0 0 0
0 0 1 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
A= ,
0 0 0 0 -1 0 0 0
0 0 0 0 0 -1 0 0
0 0 0 0 0 0 -1 0
0 0 0 0 0 0 0 0
8.2 The Stokes Equations in the u - p - w Formulation 135
fx u
fy v
fz w
o ¢
[= o :!!=
Wx
o Wy
o Wz
o P
The characteristic polynomial associated with system (8.10) is
0 0 0 0 0 -( 'T} ~
0 0 0 0 ( 0 -~ 'T}
0 0 0 0 -'T} ~ 0 (
0 0 0 0 ~ 'T} ( 0
det(Al~ + A 2'T} + A 3() = det
0 -( 0 0 0 0
'T} ~
( 0 -~ 'T} 0 0 0 0
-'T} ~ 0 ( 0 0 0 0
~ 'T} ( 0 0 0 0 0
= (e + 'T}2 + (2)4 i= 0
for all nonzero triplet (~, 'T}, (). System (8.10) or (8.8) and thus system (8.7)
is indeed elliptic. Ellipticity can be easily understood in the following way:
(8.8a) and (8.8b) constitute a div-curl system for vorticity, and (8.8c) and
(8.8d) constitute a div-curl system for velocity. These two div-curl systems
are coupled through the vorticity in (8.8c) and form the Stokes equations. In
other words, the principal part of the Stokes operator consists of two identical
elliptic div-curl operators.
The classification here is based on an ordinary method (see Sect. 4.7), so
(8.7b) is needed to guarantee the ellipticity.
must be the permissible boundary conditions for system (B.1). The reverse
is also true. Consequently, it is not possible that some boundary conditions
that have been shown to be legitimate for system (B.7) might not be so for
system (B.1), and vice versa.
As will be pointed out in Sect. B.4, the nonlinear convective term has no
effect on the classification of the Navier-Stokes equations, hence the boundary
conditions for the Stokes equations are valid for the Navier-Stokes equations.
So we need. only analyze the Stokes equations (B.7) or (B.B). Since system
(B.B) is of first order, the boundary conditions do not involve derivatives of
unknowns. In other words, there are only algebraic boundary conditions for
the solution of first-order partial differential equations. This fact precludes
the mathematical legitimacy of specifying the derivative of pressure on the
boundary for system (B.7) or (B.1).
For convenience we rewrite the Stokes equations (B.B) as the following two
coupled generalized div-curl systems:
V'p+ V' x w = f in n, (B.lla)
V'·u=O in n. (B.12b)
If the vorticity w in (B.12a) is known, the structure of these two generalized
div-curl systems would be identical. In Chap. 5 we have thoroughly investi-
gated this type of div-curl system. In order to solve (B.ll) to obtain p and
w, on the boundary we should specify
(p and n· w) or nxw,
where n x w is counted as two conditions.
When w is known, to obtain cP and u, we may solve (B.12) with the
boundary conditions specified as
(cP and n· u) or nx u.
From the above considerations we can immediately list four permissible
combinations of boundary conditions ((1)-(4) in Table B.1) for the Stokes
equations. In Table B.1 for three-dimensional problems, we don't explicitly
include the boundary condition cP = 0 on r.If we understand that cP = 0 al-
ways goes with the condition of n·u and include this condition, then the total
number of boundary conditions is four for three-dimensional problems. Since
there are eight equations and eight unknowns in the first-order elliptic system
(B.B), we indeed need four boundary conditions on each fixed boundary. In
(5) and (6) of Table B.1 we list other possible choices which are not obtained
from the above arguments but still satisfy the requirement of four conditions
8.2 The Stokes Equations in the u - p- w Formulation 137
details can be found in Jiang et al. (1994b), and thus are omitted here. Case
(5) needs special attention.
(1) Symmetric Plane. The boundary conditions Un = 0, Wri = 0, W r 2 =
o (n . u = 0, n X w = 0) on r may be used for symmetric planes. The
inhomogeneous version may be used for inflow boundaries where the normal
component of the velocity and two tangential components of the vorticity
are prescribed. These conditions correspond to those in the velocity-pressure
formulation, in which the normal velocity and the tangential 'Stresses are
given. For example, let us consider a piece of boundary with n = (1,0,0).
We have that
u=o,
aU ow
wy=---=O,
az ax
av au
wz=---=O'
ax ay
From u = 0 on this boundary we know that
au_ O au=O
az - , ay .
Therefore,
ow -0 av = o.
ax - , ax
This implies that
au ow
az + -ax = 0,
'Yxz = -
av au
'Yxy = ax + Oy = O.
That is, the tangential strains and thus the tangential stresses are equal to
zero.
In order to guarantee the uniqueness of the solution of pressure p, we
require that the pressure has a zero mean over a as indicated in (8.2d), i.e.,
!npdx=O. (8.13)
follow from the bounded inverse theorem (Theorem 4.1) in a standard man-
ner. Consequently, the corresponding finite element method has an optimal
rate of convergence for all unknowns by virtue of Theorem 4.8.
Now we examine the coerciveness of IIAUllo. Let us expand the first term
in (8.14). Since n x w = 0 on r, using Green's formula (B.6), we have
(V'p, V' x w) = (V'p, n x w) = 0, (8.15)
and thus
lIV'p + V' x wll5 = lIV'pll5 + IIV' x wll5 + 2(V'p, V' x w)
= lIV'pll5 + IIV' x w1l5· (8.16)
By virtue of (8.16) we have
IIAUII5 = lIV'pll5 + IIV' x wll5 + IIV' .wll5 + IIw - V' x ull5 + IIV' .uIl5·(8.17)
From (8.17) we have
IIAUII5 ~ lIV'pll5 = Ipli· (8.18)
Since p satisfies the zero mean constraint (8.13), from the Poincare inequality
(C.1) or (4.18) we have
Clpli ~ IIplli. (8.19)
The combination of (8.18) and (8.19) yields
CIIAUII5 ~ IIplli· (8.20)
From (8.17) we also have
IIAUII5 ~ IIV' x wll5 + IIV' . w1l5, (8.21 )
or
C(IIAUII5)! ~ IIwllo. (8.26)
From (8.22) we have
140 8. Incompressible Viscous Flows
oe/> = 0 on r. (8.36b)
on
Equations (8.36a) and (8.36b) imply that e/> is a constant or V'e/> == 0 in n.
Therefore, four conditions in the present case automatically guarantee that
the dummy variable e/> can be eliminated in (8.8c).
The boundary conditions in this case correspond to those in the velocity-
pressure formulation, in which the tangential velocity components and the
normal stress are prescribed. To show this let us consider, for example, the
surface with n = (1,0,0). Since v = w = 0, we have
ov = 0, ow = o.
oy oz
Hence from the continuity of velocity we know that
ou =0.
ox
Therefore,
ou
Ux = P + 211 ox = 0,
that is, the normal stress is equal to zero.
By using p = 0 on rand (B.4) we can show the validity of (8.16). Due to
the fact that n x u = 0 and Wn = 0, the Friedrichs first div-curl inequality is
142 8. Incompressible Viscous Flows
valid. Therefore, the coerciveness of IIAUI15 can be proved in the same way
as in case (1).
If we really specify that the dummy variable ¢J be zero on r in advance,
°
then only three boundary conditions are needed, and Wn = can be imposed
in a weak sense. In this case, the least-squares method minimizes the following
functional:
J(U) = lIV'p+ V' x wll~ + IIV'· wll~ + IIw - V' x ull~ + IIV'· ull~
+ lin . WIl~/2,r' (8.37)
and coerciveness can still be proved. For details see Jiang et al. (1994b).
0, WTI = 0, W T 2 = °
(4) Uniform Outflow. The boundary conditions are UTI = 0, U T 2 =
(n x u = 0, n x w = 0) on r. For the same reason
°
as explained in the previous case, ¢J == is guaranteed even if no boundary
condition for ¢J is given. The coerciveness of AU can be proved by just
following the steps in case (1).
(5) Wall. The boundary conditions
0, n x u = 0) on r
Un = 0, UTI = 0, U T 2 = °
(n . u =
are the so called no slip conditions on the wall. This is
a standard case of the permissible velocity boundary conditions. Bochev and
Gunzburger (1994) pointed out that in this case the full HI coerciveness of
AU does not hold. To confirm this, they consider a two-dimensional situation:
u = 0, p = sin(nx)eny and W = W z = -cos(nx)eny • Since curl W + V'p = 0
(see the definition of curl in Sect. 6.4.1),
IIAUII~ = IlwlI~ = 0(1).
However, IIwII~ = 0(n 2 ). Thus it is not true that
CIIAUII5 ~ IIwlI~·
Therefore, we shall try to find a proper norm to bound IIAUlio from below.
It is well known that the Stokes equations (8.1) have a unique solution under
the standard velocity boundary condition. This can be proved by using the
Galerkin method, see e.g., Girault and Raviart (1986). The first-order version
of the Stokes equations (8.7) is equivalent to (8.1), hence (8.7) must have a
unique solution in the subspace of [L 2 (n)J3 x L 2 (n) x [L 2 (n)J3 under the
velocity boundary conditions. By virtue of Theorem 4.6, we have
CIIAUII~ ~ Ilull~ + IIpll~ + Ilwll~, (8.38)
and particularly
CIIAUII~ ~ Ilwll~· (8.39)
On the other hand, we have the definition of IIAUllo:
IIAUII~ = IIV'p + V' x wll5 + IIV' . wll5 + IIw - V' x ull5 + IIV' . uIl5,(8.14)
and therefore
IIAUII~ ~ IIw - V' x ull~, (8.40)
8.2 The Stokes Equations in the u - p - w Formulation 143
°p(6)(p Pressure
=
=
0,
and Vorticity. In this case p 0, = 0,Wn =0, W7'l
0, n·w 0, n x w 0) are given on r. Using the boundary conditions
Wn = °one can solve (8.11) to obtain p and However, cannot
= =
w.
=
u
W7'2 =
au av
w+---=o in il, (8.46d)
ay ax
where w denotes the z-component of w, and il is a bounded domain in lR?
with piecewise smooth boundary r. The boundary conditions listed in Table
8.1 should be supplemented to complete the definition of the boundary-value
problem. They may be written as:
144 8. Incompressible Viscous Flows
0
where
0 0 1 0
A,~ (~ 0
0
-1
1
0
0
~1 ).
A2~ 0
0
0
0
1
0 D·
Ao=
Coo
o 0)
000
0000 ' l~ (~). Y~(D·
000 1
The LSFEM formulation for equal-order interpolations is given in Sect.
4.8. For the Stokes equations, the element matrices and vectors are computed
from the following submatrices:
p,u u,v
Fig. 8.la,b. Boundary conditions for the Stokes equations (from Jiang 1998)
146 8. Incompressible Viscous Flows
in n. (8.49b)
Here all variables are nondimensionalized, and Re denotes the Reynolds num-
ber, defined as
Re= UL,
11
3.0 3.0
~ ~
Ci
~ ~
2.0 2.0
--velocity --velocity
pressu re
I>--EI pressure
I>--EI
-vorticity -vorticity
1.0 1.0
0.0 1.0 2.0 0.0 1.0 2.0
-Log(h) -Log(h)
slope=2
(l'~~1
3.0 3.0
~
Ci
~
2.0 2.0
--velocity
pressure
I>--EI
-vorticity
1.0 L -_ _ ~_---' __ ~ _ _--'
1.0
0.0 1.0 2.0 0.0 1.0 2.0
-Log(h) -Log(h)
Fig. 8.2a-d. Computed convergence rate for the Stokes equations (from Jiang
1998)
148 8. Incompressible Viscous Flows
bL-________________ ~
, .
, , " , .. " " ' 1
.......
""
\\" '~/"
...... ~,~""
, , ' , ... - - ... '" "'11111
\ \ \ \ , , ' , ... - - , ""'1/1111
: . : : . ':'
c~ ____ ~~ ____ ____
~ ~
Fig. 8.3. Computed results for the cavity Stokes flow: (a) Streamline (b) Pressure
(c) Vorticity (d) Velocity (From Jiang and Chang 1990)
'V·w=O in D, (8.50b)
w-'Vxu=O in D, (8.50c)
'V . u = 0 in D. (8.50d)
required by the Navier-Stokes equations are exactly the same as for the Stokes
equations. Therefore, the boundary conditions listed in Table 8.1 can be used
for the Navier-Stokes equations (8.50) without any modification.
Before application of LSFEM, the convective term U· V'u in the momen-
tum equation (8.50a) should be linearized by using successive substitution or
Newton's method. In the successive substitution method, (8.50a) is linearized
as
. 1
Uo . V'u + V'p + Re V' x w = f in n. (8.51 )
Here the superscript "0" indicates that the value of the corresponding variable
is known from the previous calculation step. In Newton's method, (8.50a) is
linearized as
1
Uo . V'u + u . V'uo + V'p + Re V' x w = f + Uo . V'uo in n. (8.52)
The general first-order Navier-Stokes equations (8.50) take the following form
for two-dimensional problems in rectangular coordinates:
8u+8v =0 in n, (8.53a)
8x 8y
8u 8u 8p 18w
u Bx +v By + Bx + Re By = Ix in n, (8.53b)
8v 8v 8p 18w
u 8x +v 8y + 8y - Re 8x = Iy in n, (8.53c)
8u 8v
w+---=O
8y 8x
in n. (8.53d)
8u+8v=0 in n, (8.54a)
8x 8y
8u 8u 8uo 8uo 8p 1 8w
Uo -+vo -+u - + v - + - + - -
8x 8y 8x 8y 8x Re 8y
8uo 8uo
= Ix +uo 8x +vo 8y in n, (8.54b)
150 8. Incompressible Viscous Flows
av av avo avo ap 1 aw
uo -+vo -+u - + v - + - - - -
ax ay ax ay ay Re ax
avo avo
= fy + Uo ax + Vo ay in fl, (8.54c)
au av
w+- - - =0 in Q. (8.54d)
ay ax
We can write (8.54) in standard matrix form (8.47) of the first-order system,
(f 1)
in which
_~, )
0 0 1 0
A 1=
(1 0 1
Uo 0
-1 0 0
Re
, 0 0
A2~
Vo 1 o '
0 0 0
0 0
~}
~ ~ 0
ay
Ao= ( ~
0
~
ax ay 0
0 0 0
[= (f+U~+V~)
x ax
f y +u ax
ay
0
ay ,
0 ~+v 0 ~
0
0
~~ (D (8.55)
y u x v
1.0000 1.0000 1.0000 0.0000
0.9775 0.4884 0.9725 -0.5669
0.9700 0.4959 0.9675 -0.5777
0.9625 0.4996 0.9625 -0.5603
0.9525 0.4949 0.9600 -0.5463
0.8500 0.3494 0.9525 -0.5037
0.7300 0.2016 0.9450 -0.4752
0.6200 0.0868 0.9050 -0.4277
0.4500 -0.0727 0.8050 -0.3090
0.2800 -0.2312 0.5000 0.0083
0.1700 -0.3356 0.2500 0.2571
0.1000 -0.3969 0.1500 0.3645
0.0800 -0.4198 0.1000 0.4190
0.0700 -0.4363 0.0700 0.4567
0.0625 -0.4485 0.0600 0.4628
0.0525 -0.4554 0.0500 0.4567
0.0400 -0.4286 0.0400 0.4329
0.0300 -0.3704 0.0200 0.3278
0.0000 0.0000 0.0000 0.0000
The profile of the horizontal velocity along the vertical center line and the
profile of the vertical velocity along the horizontal centerline of the driven-
cavity for Re = 10 000 are illustrated in Fig. 8.4a and b, respectively. Overall,
the profile compares well with that given by Ghia et al. (1982), except in the
boundary layers. In our results, the boundary layer phenomena are more pro-
nounced. Table 8.2 lists the numerical values of selected nodes corresponding
152 8. Incompressible Viscous Flows
1.0...-----r----J"""---,
0.4 - - pre... nt
0.8 --pAMol
• Ghio et al
• Chia Itt 01
0.2
0.8 ~
} 0.0
f
0.4 ~
~.2
~.4
~.5
0.0 O.~ 1.0 0.0 0.2 0.4 0 .6 0.8 \ .0
(a) (b)
(e) (d)
Fig. 8.4. Numerical results for two-dimensional cavity flow at Re = 10 000: (a) Hor-
izontal velocity profile, (b) Vertical velocity profile, (c) Streamlines, (d) Vorticity
contours (From Jiang et al. 1994a)
to the velocity profiles shown in Fig. 8.4a and b. The computed streamlines
and vorticity contours are given in Fig. 8.4c and d.
au au ap law w
u az + v ar + az + Re (ar + -;) = fz in fl, (8.56b)
8.3 The Navier-Stokes Equations in the u - p - w Formulation 153
8v 8v 8p l8w
u 8z +v 8r + 8r - Re 8z = fr
in n, (8.56c)
8u 8v
w+---=O in n, (8.56d)
8r 8z
where z denote the axial coordinate, r the radial coordinate, u the velocity
component in the axial direction, and v the velocity in the radial direction.
For example, when (8.56) is linearized by using simple substitution, the
coefficient matrices in the standard first-order system form of (8.47) are
1o 0 0) 0 .l.
Re
vo 1 0 '
o 0 0
1r 0
o 0 1 1
o )
ReT (8.57)
o 0 o '
o 0 1
u8w
- + v8w
- + w8w
- +8p
- + 1- (8w
-- y
-8Wx)
- =
fz in n, (8.58c)
8x 8y 8z 8z Re 8x 8y
8wx + 8w y + 8w z = 0 (8.58d)
8x 8y 8z in n,
8w 8v
Wx - -
8y
+ -8z = 0 in n, (8.58e)
8u 8w
Wy - 8z + 8x = 0 in n, (8.58f)
8v 8u
Wz - - +- = 0 in n, (8.58g)
8x 8y
8u+8v+8w =0 in n. (8.58h)
8x 8y 8z
154 8. Incompressible Viscous Flows
In order to use the LSFEM given in Sect. 4.8, system (8.58) should be
linearized, for example by Newton's method, and written in standard matrix
form (8.10) in which:
Uo 0 0 0 0 0 1
1
0 Uo 0 0 0 -Re 0
1
0 0 Uo 0 Re 0 0
0 0 0 1 0 0 0
A1 = 0 0
0 0 0 0 0
0 0 -1 0 0 0 0
0 1 0 0 0 0 0
1 0 0 0 0 0 0
1
Vo 0 0 0 0 Re 0
0 Vo 0 0 0 0 1
1
0 0 Vo -Re 0 0 0
0 0 0 0 1 0 0
A2 = 0 0 1 0 0 0 0
0 0 0 0 0 0 0
-1 0 0 0 0 0 0
0 1 0 0 0 0 0
1
Wo 0 0 0 -Re 0 0
1
0 Wo 0 Re 0 0 0
0 0 Wo 0 0 0 1
0 0 0 0 0 1 0
A3=
0 -1 0 0 0 0 0
1 0 0 0 0 0 0
0 0 0 0 0 0 0
0 0 1 0 0 0 0
~ ~ ~ 0 0 0 0
8x 8y 8z
~ ~ ~ 0 0 0 0
8x 8y 8z
~ ~ ~ 0 0 0 0
8x 8y 8z
A= 0 0 0 0 0 0 0
0 0 0 -1 0 0 0
0 0 0 0 -1 0 0
0 0 0 0 0 -1 0
0 0 0 0 0 0 0
8.3 The Navier-Stokes Equations in the 'U - P - w Formulation 155
f x +u0 ~+v
8x 0 ~+w
y 0 ~z u
f y +u0 ~+v
8x ~+w
0 8y ~
0 8z v
f z +u0 ~+v 0 ~+w 0 ~
w
t=
x y 8z
0 Y.= Wx (8.59)
0 Wy
o· Wz
0 P
Three-Dimensional Driven-Cavity Flow. The least-squares solution of
the cubic driven-cavity problem was obtained by Jiang et al. (1994a) using
formulation (8.59) and also by Tang et al. (1995) using the time marching
method. Almost all previous researchers used time-marching methods, and
claimed that steady-state solutions had been obtained for Re = 2000. Also
no Taylor-Gortler-like vortices were reported for Re < 2000 by previous
researchers.
Koseff and Street (1984a,b,c) conducted systematic experiments of flow
in a driven-cavity. Their experiments reveal that the flow field is highly un-
steady and possesses significant secondary motions (end-wall corner eddies
and Taylor-Gortler-like vortices). They mention that the flow is not stable
at Re = 2000 and show results for Re = 3200. However, a very important cri-
terion, the Reynolds number at which the flow becomes unstable, had never
been reported.
In a study conducted by Jiang et al. (1994a), the computation is carried
out up to Re = 3200. The boundary conditions are as follows: u = 1, v =
w = 0 are specified on the top driven surface (y = 1.0), and u = v = w = 0
on all solid walls (Fig. 8.5). At the center of the bottom (x = 0.5, y =
0.0, z = 0.5), p = 0 is specified. The mesh is based on 50 x 50 x 50 uniform
(i)
(g) (h)
Fig. 8.6. Numerical results for three-dimensional cavity flow at Re = 1000: (a)
Flow pattern, (b) Vorticity(wx ) contours, (c) Pressure contour at x=0.5; (d) Flow
pattern, (e) Vorticity(wx ) contours , (f) Pressure contour at y=0.5; (g) Flow pat-
tern, (h) Vorticity(wx ) contours, (i) Pressure contour at z=0.5 (From Jiang et al.
1994a)
8.3 The Navier-Stokes Equations in the u - p - w Formulation 157
1.0
0.8 • Re-100
• Re-400
Re=1000
0.6
:0-
0.4-
0.2
0.0
-0.5 0.0 0.5 1.0
u-velocity
tom shown in Fig. 8.8 become more clearer and stronger. The L2 norm of the
residual remained around 6.0 x 10- 5 • The maximum difference of the velocity
magnitudes between the results of two sequential Newton's steps is around
9.0 x 10-3 and couldn't be further reduced. Newton's method employed here
may be considered as an implicit time-marching method with a very large
time step. The fact that the method cannot lead to a convergent solution may
imply that at Re = 1000 the flow in the cavity is not stable. The appearance
of the Taylor-Gortler-like vortices is consistent with this instability.
It appears that the results with LSFEM are highly accurate and also
predict instability at a lower Reynolds number than shown by previous works
in this field.
Three-Dimensional Backward-Facing Step Flow. The LSFEM based
on (8.59) is further tested by solving the backward-facing step flow problem
(Jiang et al. 1995). It is well known that for backward-facing step flows,
the reattachment length obtained by two-dimensional calculations cannot
8.3 The Navier- Stokes Equations in the u - p - w Formulation 159
Fig. 8.8. Velocity vector at x = 0.5 for three-dimensional cavity flow at Re = 1000
(From Jiang et al. 1994a)
mid-plane
\
o 5 10 15 20
Fig. B.10. Nonuniform mesh (xy plane and xz plane) (From Jiang et al. 1995)
26
(a) 7/W.,f).8974
14 20 24 26
(b) :jWJ.J.76
26
(e) 7/W=(J.0 (mid-plane)
Fig. 8.11. Velocity vectors for Re=800 (From Jiang et a1. 1995)
x/s-()
Fig. 8.12. Pressure contours for Re=800 (From Jiang et al. 1995)
162 8. Incompressible Viscous Flows
y
:0.4
_ _= _-_
--_---0.2
--:-:---=------------;
.~:::;:-=-=-=-:--0
-0.2 - ___":.'
0.2=
-0_
-c._
~ -3;;.:.·2.6 ... ·2.,;:;.......;- . - ' . -. . \~ . 1.8 1
Z=6.5 (b) 9.0
in the x, y and z direction individually, while the largest element has a size
of 3.0, 0.06938 and 1.68367 steps respectively.
Figure 8.10 shows the nonuniform mesh which has more elements close to
the sidewall, the floor and the roof in the test channel. The matrix-free Jacobi
preconditioned conjugate gradient method is employed for the solution of the
resulting algebraic equations. One-point Gaussian quadrature is used for the
evaluation of element matrices. The solution of the Stokes equations is taken
as an initial guess for the case of Re = 100, and the 'converged' solution
is then used as an initial guess for higher Reynolds numbers. 'Converged'
solutions are those whose L2 norm of the residuals less than 1.0 x 10- 4 .
The problem is solved using about 5M words of memory on a CRAY-YMP.
8.3 The Navier- Stokes Equations in the u - p - w Formulation 163
-~ ~-
-~.I
Y
;;,;,j ~
;.:
::::; -
l
~ ~
~ ~ ~ ~
\
~
Z=6.S 9.0
(a)
-=0.1 Y
-- -~- -- -
- .,..--
Jj JJ
~ ~ ;:::
~ ,
II
~
~
-~
Z=6.S 9.0
(b)
- =0.1 Y
,1
-
- -
::::::
- - - -
t
:::::: ~
:;:::
~ ~
l
~
~ ::::
;
l
Z=6.S
~ §.
~ ~
~
~
~
9.0
(e)
Fig. 8.14. Velocity vectors for Re=800 at x/S=(a) 6.25, (b) 9.0, (c) 12.25 (From
Jiang et al. 1995)
~:~ {-----;.;,=O.O--------------------------
0.8 ~ ~ .. ~ .. .:. .:. .:. ..
U 6.25
0.7 O--o--()---o--o--o-........
0.6 9.0
0.5
0.4 ~-C...~
•. _-i--+~.5I::~~Q:s::5~~I-o-«:r-CJ
0.3 /
18.1S3
0.2
0.1
Re=648 y=7.5 mm
0.0
~.1+-----~----------------~----~~~
0.0 0.2 0.4 7/W 0.6 0.8 1.0
1.0
Rc=648 y=2.35 mm
0.9
0.8
U
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
~.1
0.0 0.2 0.4 7/W 0.6 0.8 1.0
Fig. 8.15. Spanwise distributions of u-velocity profiles for Re=648 (From Jiang et
al. 1995)
8.3 The Navier-Stokes Equations in the u - p - w Formulation 165
16 a a I
a
a II Re::389
[
14
x a
• Re::500 _
x x )C
A Re=6OO
0 x
C) 0 0 0 (
0 0
12 Re=648 -
A A 6. 6.
• A ~ )C
Re::700
~
a
10 • • • • • •• I~ 0 Re::800 -
8 .. .a
.. - '" -- G
.. III III ~~
I
~
6
~~
4
0.0 0.2 0.4 'Z/W 0.6 0.8 1.0
Fig. 8.1 6. Spanwise locations of reattachment line for various Reynolds numbers
(From Jiang et al. 1995)
,!=-l8J, I -
15
l!! 14
.
;c I
.!f 13
! 11
1! 11
10
a
9
a
8
7
I
6 ~.
200 300 400 soo 600 700 800
Re
Fig. 8.17. Comparison of experimental and numerical results for primary reattach-
ment length up to Re=800 (From Jiang et al. 1995)
layer, then drops and resumes two-dimensional flow from z/W = 0.5 to the
mid-plane (z/W = 0.0).
Figure 8.16 demonstrates the three-dimensional effects by depicting the
spanwise location of the reattachment length for the primary separated-flow
region. Next to the sidewall, the reattachment length increases mainly due
to the interaction of viscous stress between the channel floor and sidewall.
Figure 8.17 shows the computed reattachment length of the primary recir-
culation zone, and compares well with the experimental results obtained by
Armalyet al. (1982). In the Re range for which most simulations fail to pre-
dict the reattachment length because of the three-dimensional phenomenon
(Re = 450 '" 800), the calculated results by LSFEM agree very well with
experimental results up to Re = 800.
As the Reynolds number increases, an additional separated-flow region
occurs near the channel ceiling. Figure 8.18 illustrates the spanwise detach-
ment (X4/S) and reattachment (X5/S) lines of this second eddy. The present
results show that as the Reynolds number increases, this upper-wall eddy
propagates toward the mid-plane and its length decreases in the spanwise
direction. For example, as the Reynolds number increases from 600 to 800, it
ranges from 17.5 to 22 step heights at z/W = 0.97708 and increases from 0.2
to 9.35 step heights at z/W = 0.76. The upper-wall eddy at the mid-plane is
not observed.
From these comprehensive numerical results, it is seen that the LSFEM
is a remarkably robust method even for three-dimensional flows at large
Reynolds numbers.
8.4 The Navier-Stokes Equations in the u - b - w Formulation 167
2S
20
rs
IS
10
5
.,
a
0.5 0.6 0.7 7/W 0.8 0.9 1.0
Fig. 8.18. Spanwise distributions of detachment and reattachment length for var-
ious Reynolds numbers (From Jiang et al. 1995)
in a, (8.61b)
in a, (8.61c)
168 8. Incompressible Viscous Flows
ov + -au = 0 in D, (B.61g)
ox oy
Wz - -
au + ov + ow =0 in n. (B.61h)
ox oy oz
For two-dimensional problems, let w = wz , we have
ob + -
-vw + -ox 1 ow
- = Ix in D, (B.62a)
Re oy
ob 1 -
uw + -
ow = Iy in D,
oy - -
Re ox
(B.62b)
ov au
w--+-=O inD, (B.62c)
ox oy
au + ov = 0 in n. (B.62d)
ox oy
We note that the first-order velocity-total pressure-vorticity formulation
is almost-linear. In this formulation, a good portion of the nonlinearity has
been absorbed into the Bernoulli pressure; the nonlinear term u x w is of zero
order, and thus is not related to any derivatives; while the rest of the terms
constitutes the linear Stokes equations. Therefore, the non-linear term has
no effect on the classification of the system, and it does not matter how large
the Reynolds number is, the whole system is elliptic. For this reason, the
incompressible Navier-Stokes equations have the same boundary conditions
as those listed in Table B.1 for the Stokes equations.
The velocity-total pressure-vorticity formulation is also suitable for LS-
FEM. It is found that for tested problems the velocity-total pressure-
vorticity formulation leads to slightly slow convergence (Jiang et' al. 1994a).
aU AU 00' xx 00'
aU) op ( --+ xy 00' xz ) = f x in a,(8.63 a)
p ( u -+v -+w -oz +-- --+ --
ax oy ax ax oy oz
OW
p ( u -+v -+W -
ow op (OO'xz
OW) +-- --+ oO'yz -- = f z ina,(8. 63c)
--+oO'zz)
ax oy oz oz ox oy oz
ou (8.63d)
O'xx-2 ",ox= O ina,
OV (8.63e)
O'yy - 2", oy = 0 in a,
au = 0 in a, (8.63f)
0' zz - 2", oz
in a, (8.63g)
in a, (8.63h)
( OV OW) in a, (8.63i)
O'yz - '" oz + oy = 0
au + ov + ow = 0 in a, (8.63j)
ax oy oz
where p is the density, '" is the viscosity.
In (8.63) there are ten equations, that is, three momentum equations,
ng ten
six constit utive equations and one incompressibility equation, involvi
due to
unknown variables (u, v, W,p, O'xx, O'yy, O'zz, O'xy, O'xz, O'yz). However,
ically inde-
incompressibility, three normal stress O'xx, O'yy, O'zz are not algebra
penden t, that is
O'zz = -(O'xx + O'yy).
fact only
Therefore, in the velocity-pressure-stress formulation there are in
un-
nine (odd) indepe ndent differential equations in nine (odd) indepe ndent
der sys-
known variables. As discussed before in Sect. 4.7, this kind of first-or
y sense, and it is also difficult to specify
tem cannot be elliptic in the ordinar
ons. Numer ical experim ents con-
an approp riate number of bounda ry conditi
l Stokes equatio ns with this for-
ducted by the author on the two-dimensiona
the pressur e and the stresses
mulation reveal that the rate of convergence for
ation has
is one order lower than that of the velocity. Moreover, this formul
tations to obtain
too many unknowns. It is not profitable to do more compu
ian fluid
directly the stress components which are not import ant in Newton
ry,
flows. In case someone is interested in the shear stresses on the bounda
170 8. Incompressible Viscous Flows
the shear stresses can be algebraically expressed by the vorticity. For these
reasons it is not recommended that the velocity-pressure-stress formulation
be used for Newtonian fluid flows.
+(1 - ()) ( un. V'un + V'pn + ~ V' X wn) = I n+1 in n x (0, T]'(B.65a)
in n x (0, Tj, (8.65b)
in n x (0, Tj, (8.65c)
in nx (0, T). (B.65d)
Usually, two choices can be made: () = 1/2 corresponds to the Crank-Nicolson
scheme and provides second-order accuracy O(Llt 2 ), and is used for true
8.6 Time-Dependent Problems in the 1.1. - P - w Formulation 171
v n+
_ _1 _ vn
_ _ +() (u no
+1 v +1
_n_ +vn +1 ov
__
n+1 + _____
opn+l 1 ow n_+1 )
Llt 8x 8y 8y Re 8x
8v n n av n 8pn f n+1 in n, (B.66c)
+(1 - ()) ( 1.1.n -+v
8x ay ay
- -_
- + - - -1-awn) Re ax Y
aun+1 n
8v +1
wn +1 + __ -- - =0 in n. (B.66d)
. 8y 8x
The convective term in (B.66) can be linearized by using the simple substi-
tution method or Newton's method as explained in Sect. B.3. When a small
time-step is used, one Newton's linearization is often enough at each time-
step. Then, for example, u n+18un+1 /8x in (B.66b) can be approximated as
172 8. Incompressible Viscous Flows
System (8.66), after linearization, can be written in standard form (8.47) for
urI
first-order systems in which
A - ( +0°ax
1
.It
8u n
0
08u n
8y
o 0)
0
1!= 0- 08v n
8x
.l+08v n
.It 8y o 0 '
0 0 o 1
A1 = (O!.0
0
0
0
Oun
-1
0
1
0
0
o
0) ,A
-~ie 2 =
C rOv n
~ Ov n
1
0
0
0
0
1
0
o)
O.l
,(8.67)
+
-(1 _ O)(u n 8u n
8x
+vn 8u n
8y
+ ~O
8x
+ .l8w
Re 7fY
+ O(u n n
) 8u n
8x
+vn 8u n
8y
»)
( -(1 _ O)(u n 8v n +vn 8v n +~ _ .l8w + O(u n n
) 8v n +vn 8v n ) •
8x 8y 8y Re 8x 8x 8y
o
If more than one linearization is needed at each time-step, the formulation
will be similar to that given in Sect. 10.3.
The first-order system is then discretized in space following the standard
LSFEM procedure in Sect. 4.8. The matrix-free Jacobi preconditioned conju-
gate gradient method (see Chap. 15) is recommended for solving the resulting
linear algebraic equations. The advantage of the conjugate gradient method
is that the time step can be very small so that the largest element CFL num-
ber is close to unity. In this way, time accuracy is achieved without heavy
expense of computer time. This is attributed to the fact that when the time
step is small, the solution at the previous time step serves as a very good
initial guess for the solution at the current time-step, thus the conjugate gra-
dient method needs fewer iterations than when a larger time step is used.
Therefore, although the time-marching LSFEM is an implicit method which
has the advantage of unconditional stability, when combined with the con-
jugate gradient method, its implementation is similar to an explicit method,
and thus is relatively inexpensive.
Vortex Shedding Behind Circular Cylinder. A circular cylinder is im-
mersed in an infinite and uniform flow field with u = 1, v = o. When Re 2: 40,
a periodic oscillatory flow occurs behind the cylinder. This is the well-known
von Karman vortex street problem. The interest in this problem is due to its
periodic flow pattern and the unknown outflow boundary conditions.
8.6 Time-Dependent Problems in the u - p - w Formulation 173
Fig. 8.19. Mesh grids for vortex shedding behind a circular cylinder (From Wu
and Jiang 1995)
St = 0.186
> 0.0
-0.8
500.0 550.0 600.0 650.0 700.0
time step
Fig. 8.20. Time history of velocity component v at point P (From Wu and Jiang
1995)
(a.)
(b)
Fig. 8.21. Vortex shedding by a circular cylinder at Re = 200 at t = 150: (a)
Pressure contours, (b) Streamlines (From Wu and Jiang 1995)
D
St = UT'
8.7 Fluid-Thermal Coupling 175
where D is the diameter of the cylinder and U the characteristic velocity. The
computed Strouhal number falls within the range of experimental data given
by Roshko (1953) and numerical results obtained by using other methods, see
e.g., Engleman and Jamnia (1990). Figure 8.21 shows the instantaneous pres-
sure contours and streamlines after the stable periodical solution is reached.
More details about the transient LSFEM solutions for this and other
problems can also be found in Tang (1994).
in which
Reynolds number
Re = UL t'o.J inertial force,
v viscous force
Pe = UL advection of heat .
Peclet number
t'o.J
conduction of heat
K,
in which
v
Prandtl number Pr= -,
K,
is a property of the fluid, not of the particular flow. The Rayleigh number
indicates the relative importance of the buoyancy force.
In order to apply LSFEM to obtain the approximate solution of (8.72),
we need to introduce the vorticity wand the heat flux q = (qx,qy,qz) as
independent variables and transform system (8.72) into the following first-
order 1.1. - P - w - () - q system:
81.1. g
at + u· Y'u + Y'p + PrY' x w + PrRaO jgj = 0 in n x (0, TJ, (8.73a)
n x (O,Tj,
in (8.73b)
and d = 1 is the thickness of the vessel, see Fig. 8.23. The gravitational
force is in the negative z direction. The vertical walls are adiabatic. The
bottom wall is heated isothermally to a higher temperature Bh , and the top
wall is maintained at a lower temperature Be. The fluid starts motion when
the Rayleigh number Ra reaches the critical value. This is termed Rayleigh-
Benard convection.
For fluid between infinite parallel plates, the critical Rayleigh number
Rae = 1708. Theoretical (Davis 1967 and Catton 1972) and experimental
(Stork and Muller 1972) results show that the critical value in small vessels
8.7 Fluid-Thermal Coupling 179
(a) (b)
(c) (d)
Fig. 8.23. Velocity vectors on the plane z = 0.5 at t = (a) 25, (b) 45, (c) 60 and
(d) Steady state (From Tang and Tsang 1997)
is higher than in large one. In general, the critical Rayleigh number increases
when the aspect ratio of the container decreases, because sidewalls have a
stablizing effect on the onset of convection due to viscous drag. Moreover, the
roll structures and thermal fields are decisively affected by the aspect ratio of
the vessel, thermal properties of the sidewalls, and the physical properties of
the fluid. The most interesting phenomenon of Rayleigh-Benard convection
is the pattern selection process. Due to the nonlinearity of the governing
equations, several solutions are possible for the same set of parameters. The
mechanisms on how different solutions to the problem evolve and compete
and the processes by which a particular flow configuration undergoes change
are broadly known as pattern selection. A numerical study using LSFEM that
180 8. Incompressible Viscous Flows
(b)
(c) (d)
Fig. 8.24. Temperature contours on the plane z = 0.5 at t = (a) 25, (b) 45, (c)
60 and (d) Steady state (From Tang and Tsang 1997)
1
q + Pe '\If) = 0 in n x (0, Tj, (8.74f)
u 0 on r,
° on= x O,a x ,
°1 on
on
z
= y O,ay,
=0,
° zz = 0, on
° z = 1, on =0,
° z = 1, on
qx
° z=
= on
qy
°
= on 1.
The numerical technique is based on an implicit, fully coupled, and time
accurate method, which consists of the Crank-Nicolson scheme for time inte-
gration, Newton's method for the convective terms with extensive lineariza-
tion steps, and the LSFEM for space discretization. A matrix;-free Jacobi
preconditioned conjugate gradient method (see Chap. 15) is implemented to
solve the resulting symmetric positive definite linear system of equations.
For the solution illustrated in Fig. 8.23, Ra = 8 x 103 , Pr = 0.71 (air), the
geometric aspect ratio is ax : ay : a z = 5 : 5 : 1, and the mesh consists of
60 x 60 x 14 trilinear elements. One-point Gaussian quadrature is used for
the computation.
The velocity vectors and temperature contours on the plane z = 0.5 are
shown in Fig. 8.23 and Fig. 8.24, respectively. Figure 8.25 shows the results
on the plane y=0.5. When steady-state is reached, three vortex tubes, i.e.,
182 8. Incompressible Viscous Flows
Fig. 8.25. Velocity vectors and temperature contours on the plane y = 0.5 at t =
(a) 25, (b) 45, (c) 60 and (d) Steady state (From Tang and Tsang 1997)
8.7 Fluid-Thermal Coupling 183
six rolls in a vertical cross section, ring around the vertical center line of
the vessel can be clearly observed. The flow structure is referred to as a
multicellular flow pattern. More results for different cases can be found in
Tang and Tsang (1997).
must be used. In Table 8.4 the properties of the silicon oil used in their
experiments are listed. According to these data, the Prandtl number of the
silicon oil Pr = v / /'i, is about 1000, and the capillary number Cc = Po/'i,V hd
is about one thousandth. Note that d is the depth of the fluid layer and is
in the order of millimeter. Since the flow motion is thermally driven, the
Prandtl number is a measure of the sluggishness of the fluid: higher Prandtl
number implies slower motion. On the other hand, in the absence of gravity,
the capillary number Cc is a measure of the surface deflection. And smaller
Cc implies higher surface tension that corresponds to a non-deflecting free
surface.
The boundary condition on the bottom and the side walls of the container
is the no-slip condition:
U=V =W =0. (8.75d)
In addition, the bottom is heated to a constant temperature (h, and the side
walls are insulated:
() = ()1 on the bottom, (8.75e)
n . V () = ° on the sidewalls. (8.75f)
On the top free surface, the Marangoni boundary conditions are applied:
8.7 Fluid-Thermal Coupling 185
au aB
POIl- = -'1-, (8.75g)
az ax
av a(}
POIl- = -'1-, (8.75h)
az ay
w = o. (8.75i)
The Marangoni conditions represent the relationship between the flow shear
stress and the tangential surface tension force across the free surface. Any
inhomogeneity of the surface tension due to temperature variations creates
a shear force on the free surface and therefore results in fluid motion. These
Marangoni conditions are the driving force for Marangoni-Benard convection.
The heat loss on the free surface is modeled by the usual heat transfer
condition:
(8.75j)
where
Ma= ,ed (8.80)
p/l/'i,
(8.82)
(a) (b)
(c) (d)
Fig. 8.26. The computed patterns in the square container: (a) Two-cell, (b) Three-
cell, (c) Four-cell, (d) Five-cell (From Yu et al. 1996c)
188 8. Incompressible Viscous Flows
1
q+~'VO=O in [} x (0, T), (8.88f)
PrJGr
'Vxq=O in [} x (0, T], (8.88g)
ac
at + U· 'Vc + 'V. J = ° in [} x (0, T], (8.88h)
1
J + --='Vc = 0 in [} x (O,Tj, (8.88i)
ScJGr
'V x J =0 in [} x (0, Tj, (8.88j)
1.0 rt::~;::::============~
0.8
0.8
0.4
0.2
0.8 0.8
0.8 0.8
0 .• 0 .•
0.2 0.2
0.0
0.0 0.2 0.4 0.8 0.8 1.0 0.0 0.2 0.. 0.8 0.8 1.0
(c) (d)
Fif' 8.27. Augmented flow of double-diffusive convection in a square cavity at Gr =
10 and X = -1.0: (a) Velocity vectors, (b) Stream functions, (c) Temperature
contours, (d) Concentration contours (From Tang and Tsang 1994)
8.8 The Second-Order u - w Formulation 191
The LSFEM solution of the coupled system for heated side-walls and
diffusion in a square cavity is given by Tang and Tsang (1994). The cavity
has insulated and impermeable top and bottom walls but with heated and
solute coated vertical walls. The fluid motion driven by temperature and
solute gradients consists of two flow structures: augmenting and opposing
flows. Each flow structure has two cases, Gr c > 0 and Grc < o. In their
study, the solute is the heavier component. Therefore, the buoyancy ratio
X is always negative, Le., Grc < o. Here we only discuss augmented flows.
Results for opposing flows can be found in Tang and Tsang (1994).
Augmented flow is the flow driven vertically up by high temperature near
the left wall and down by high concentration near the right wall. The fluid
flow is always in the clockwise direction. The boundary conditions are: u =
v = qy = J y = 0 everywhere; () = 1, c = 0 at the left wall and () = 0, c = 1
at the right wall. The reference pressure p = 0 is specified at the lower
left corner. A non-uniform mesh of 50 x 50 bilinear elements is used. Since
the steady-state solution is sought, the system is discretized in time by the
backward-Euler differencing. A large time step Llt = 1.0 is applied. One-point
Gaussian quadrature is used in the computation. The results for the case of
Gr = 106 , X = -1.0, Pr = 0.7 and Sc = 0.6 are given in Fig. 8.27. These
results compare well with Beghein et al. (1992) and Lin et al. (1990).
We will also use the div-curl method introduced in Sect. 5.3 to rigorously
derive the popular second-order velocity-vorticity formulation, the pressure
Poisson equation and their boundary conditions. We believe that many con-
troversies over the permissibility of the boundary conditions for derived equa-
tions can be resolved definitively via this div-curl method.
+(Vp+Vxw-f, VXT)
+(V·w, V'T)
+(w - V x u, T)
- (w - V x u, V xv)
+(V·u, V·v)=O V(v,q,T)EH o. (B.90)
where Ho = {[Hl(!1)j1 : (v, q, T) satisfies the corresponding homogeneous
boundary conditions on r}. By using Green's formulae (B.1), (B.3) and (B.5),
from (B.90) we have
+(w - V x u, T)
-(Vx(w-Vxu), v)+(nx(w-Vxu), v)
From (8.91) after simplification by using the vector identity (A.4), we obtain
the following Euler-Lagrange equations and boundary conditions:
Llp = \l . f in il, (8.92a)
p = given or n· (\lp + \l x w - f) = 0 on r, (8.92b)
(1) n . u, n x w given
n . (\l p + \l x w - f) = 0,
\l·w = 0,
n x (w - \l x u) O.
(2) p, n . u, n . w given
n x (\lp + \l x w - f) 0,
n x (w - \l x u) 0.
194 8. Incompressible Viscous Flows
(3) p, n X u, n . w given
n x (V'p + V' x w - f) = 0,
V'·u = O.
(4) n x u, n x w given
n· (V'p + V' x w - f) = 0,
V'·w = 0,
V'·u = O.
(5) U given
V'p+ V' x w - f = 0,
V'·w = O.
(6) p, n· w, n x w given
n x (w - V' x u) = 0,
V'·u = o.
We emphasize again that the least-squares finite element method based on
the first-order velocity-pressure-vorticity formulation (8.7) does not need any
additional boundary conditions. However, if another discretization scheme
such as the finite difference method is used to solve the second-order velocity-
pressure-vorticity formulation (8.92a), (8.92c) and (8.92f), the additional
natural boundary conditions should be included.
We notice that in the second-order velocity-pressure-vorticity formula-
tion, the solution of the pressure Poisson equation (8.92a) is coupled with
the solution of the velocity and the vorticity through the boundary condi-
tions. The significant advantages of the present second-order formulation are
that (1) it guarantees the satisfaction of the continuity of velocity and the
solenoidality constraint on the vorticity without explicitly including these
two divergence-free conditions; (2) it is suitable not only for the standard
boundary conditions but also for non-standard boundary conditions; (3) the
differential operator is self-adjoint (symmetrical).
The given boundary data Ur should satisfy the global mass conservation:
n. (8U
8t
-uxw+'Vb+~'VxW-f) =0
Re
on r x (0, T], (8.94c)
8w 1
at - 'V x (u x w) - ReLlw = 'V x f in n x (O,T], (8.95a)
in n x (O,TJ, (8.95g)
196 8. Incompressible Viscous Flows
U = Ur on r x (O,T]. (8.95h)
Since (8.95a) implies that
o(V . w) _ ~Ll(V . w) =
at Re
° in n x (O,T], (8.96a)
if we specify that
V .w = ° on r x (0, T], (8.96b)
then V·w == °
in n, that is, the divergence-free property of the vorticity vector
is guaranteed. Therefore we can replace the divergence equation (8.95f) by
the boundary condition (8.96b).
As explained in Sect. 5.5, the divergence equation (8.95g) can be elim-
inated, since (8.95d) and (8.95f) and the boundary conditions (8.95e) and
(8.95h) guarantee the divergence-free of the velocity.
Finally we obtain the second-order velocity-vorticity equations and the
pressure Poisson equation for the Navier-Stokes equations:
ow 1
at - V X (u x w) - ReLlw =V x f in n x (O,T], (8.97a)
n x (w - V X u) = 0 on r x (O,Tj, (8.97b)
Llu + V x w = 0 in n, (8.97d)
U = Ur on r x (O,Tj, (8.97e)
n . (Vb - U x w + -V
Re
1
x w- f) = ° on r x (O,T]. (8.97g)
From (8.97) it is clear that the calculation of the velocity and vorticity
is decoupled from that of the pressure; (8.97a) and (8.97d) are termed the
velocity-vorticity formulation. For simplicity, in this section we have studied
only the case with the velocity boundary conditions specified on the whole
boundary. For cases with other boundary conditions the derivation of these
additional Neumann type of boundary conditions is similar. It is worth not-
ing that if the pressure is prescribed on a part of boundary, the additional
boundary condition on that part of the boundary for the vorticity equation
(8.97a) is more complicated than (8.97b).
In the finite difference framework one may solve the vorticity and the
velocity in a sequential manner: using the velocity field known from the pre-
vious time step one may solve (8.97a) with boundary conditions (8.97b) and
(8.97c) to obtain the vorticity, then by solving the Poisson equations (8.97d)
update the velocity. When convergent solutions for velocity and vorticity are
8.9 Concluding Remarks 197
obtained, one can solve the pressure Poisson equation (8.97f) with the Neu-
mann boundary condition (8.97g) to obtain pressure.
Obviously, the LSFEM based on the first-order velocity-pressure-vorticity
formulation is much more simple, since (1) no Neumann-type boundary con-
ditions (8.97b), (8.97c), and (8.97g) are involved; and (2) neither upwinding
nor staggered grids are needed; (3) the pressure boundary condition can be
handled easily.
K-(v, q, r) = IIV'q + V' x rllg + IIV' . rllg + IIr - V' x vll~ + IIV' . vll~. (8.98)
Note that this functional involves products of second derivatives of the ve-
locity, and thus the corresponding method is not practical due to the need of
C 1 elements for the approximation of the velocity. In order to overcome this
difficulty, by considering the inverse inequality (2.56) for finite element shape
functions they suggest that the norm II . 111 in (8.98) might be simulated by
h- 1 11 ·110 where h is an appropriate measure of the grid size. Therefore, the
following simple weighted functional is recommended by them:
8.9 Concluding Remarks 199
Their numerical experiments support this idea. Their formulation was ap-
plied to boundary control problems (Bochev and Bedivan 1997). The same
weighted functional (without the \l . T term) was also proposed by Harbord
and Gellert (1991). However, the weighted functional (8.99) perhaps has more
theoretical value than practical value, because (1) for real world problems,
the mesh is often nonuniform, then the question is how to choose the mesh
size h; (2) for practical problems, the boundary conditions are not always
the pure standard velocity condition, they can be mixed with anyone listed
in Table 8.1; (3) more important, in Sect. 4.11 we point out that the least-
squares finite element method with Gaussian quadrature is equivalent to the
least-squares finite element collocation method. In order to obtain a good
approximate solution, the number of Gaussian points must be selected such
that one solves a determined or a slightly overdetermined collocated residual
equations. As discussed in Sect. 4.11, for a determined set of linear alge-
braic equations, weighting has no effect on the least-squares solution. That
is, in practical computations we must do our best to make the weighting have
minimal effect.
Chang (1990) and Chang et al. (1995) also proposed a least-squares
method for the Stokes equations based on the velocity-velocity derivatives-
pressure formulation. This formulation achieves optimal rate of convergence
in the HI norm for the velocity boundary condition case. However, it has too
many variables.
As pointed out originally by Lighthill (1986), velocity and vorticity are
the most suitable variables from the viewpoint of a fluid dynamic description
of incompressible viscous flows. Therefore, the first-order velocity-pressure-
vorticity formulation is preferred for the least-squares method.
9. Convective Thansport
dy
-ds = /32, (9.3b)
dz
-ds = /33. (9.3c)
These curves are the characteristics, since their normal direction is (a1' a2, (3).
By virtue of the chain-rule
d au dx au dy au dz
ds (u(:z:(s» = ax ds + ay ds + az ds
au au au
/31 ax + /32 ay + /33 az = (3. Vu,
i.e., along each characteristic the partial differential equation (9.1) is reduced
to an ordinary differential equation:
d
ds u(:z:(s» + u(:z:(s» = o. (9.4)
901 Steady-State Problems 203
(u, v)out = r
Jrout
uvn {3ds,
0
where
rout = {(x, y) E r : n(x, y) {3
0 ~ O}o
We recall Green's formula
(u,8, v) = (u,v) - (u,v,8), (905)
and thus
1 2
(u,8,u) = 2"l u lr o (9.6)
Further we also define the function space
S = {u E Hl([}) : u = 0 on lin},
and the corresponding finite element subspace Sh, Le. Sh is the space of
continuous piecewise polynomial functions of order ro From the finite element
interpolation theory (Theorem 405) we have: Given a function u E Hr+l([}),
there exists an interpolant Ihu E Sh such that
lIu -lhullo ::; Chr+lllullr+I. (907a)
Now let us analyze the following classic Galerkin method for problem (9.1):
Find Uh E Sh such that
(9.9)
Since the exact solution satisfies (9.1), we also have
(9.10)
By subtracting (9.9) from (9.10) we obtain the following orthogonality for
the error e = U - Uh:
(9.11)
Let Ihu E Sh be the interpolant of U satisfying (9.7) and write P = U - Ihu
and "I = Ihu - Uh so that e = P + "I. Then, by using (9.5), (9.6), (9.11) and
(9.8) we have
1
21el} + IIel1 2 = (efJ + e, e) = (efJ + e, p) + (efJ + e, "I)
= (efJ'p) + (e,p) = -(e,pfJ) + (e,p) + (e,p)
212212 212
~ IlpfJllo + 411ello + IPlr + 41elr + Ilpllo + 411ello. (9.12)
By moving all terms related to e into the left-hand side, the inequality (9.12)
becomes
1212 22 2
41elr + 2"e"o ::; IIp/311o + IPlr + IIpllo· (9.13)
Gathering (9.20) and (9.21), moving the terms related to e into the left-hand
side, we obtain the error estimate for the SUPG method:
problem (cf. (9.16)). The major weakness of the SUPG method in practical
calculation is the non-symmetry of the resulting algebraic equations that
makes the solution of large-scale problems difficult.
In recent years, Hughes and coworkers have been promoting the Galerkin/
Least-Squares method (Hughes et al. 1989). In this method, which is more
general than the SUPG method and applicable to a wide variety of other
problem types, squares of residuals are added to the classic Galerkin method.
These terms yield improved stability properties with respect to the classic
Galerkin method without degrading the accuracy. For problem (9.1) the for-
mulation of the Galerkin/Least-Squares method is: Find Uh E Sh such that
(Uh,{3 + Uh, Vh) + h(Uh,{3 + Uh, Vh,{3 + Vh)
= (j, Vh) + h(j, Vh,{3 + Vh) VVh E Sh. (9.25)
It is easy to show that the mathematical properties (9.23) and (9.24) are still
valid for the Galerkin/Least-Squares method.
u=o (9.32b)
where n = {(x,y) E 1R?: 0 < x < 1,0 < y < I} is the unit square, and
rin = {(x, y) E r : x = 0 or y = O}, in which r is the boundary. This
problem has a smooth exact solution u = sin(x)sin(y).
A uniform meshes with n x n bilinear elements is employed. At first, one-
point Gaussian quadrature was used for calculating element matrices. As
pointed out in Sect. 4.11, in this case, the LSFEM is equivalent to the least-
squares finite element collocation method with one collocation point at the
center of each element. It is easy to check that in such a collocation method,
the number of discretized algebraic equations is equal to n 2 ; while the number
of unknowns is also equal to Nnode - Nbc = (n + 1)2 - 2 x n - 1 = n 2 • In
other words, in such a case we solve a determined system. The numerical
result for the convergence rate is shown in Fig. 9.1. The optimal rate, i.e.
lIelio '" O(h2), is observed. The numerical rate of convergence with the 2 x 2
Gaussian rule is also included in Fig. 9.1. In this case, the LSFEM solves an
overdetermined system. The convergence rate is around O(h1.75), which is
suboptimal.
We also repeated the numerical tests with specified extra boundary con-
ditions on the outflow boundary rout = {(x, y) E r : x = 1 or y = I}. In this
case, the least-squares method with 2 x 2 Gaussian rule gives the optimal
rate of convergence Ilello '" O(h2) (Fig. 9.1). .
From these numerical results we may conclude that the LSFEM for the
hyperbolic equation has an optimal or near optimal rate of convergence de-
pending on the number of Gaussian points in the calculation of element ma-
trices. More Gaussian points yield slightly less optimal results, because the
least-squares method is not able to make the residual of each equation in the
overdetermined system equal to zero.
208 9. Convective Transport
....... x IX I
............ I 2X2 EX
6.5
-- 0 2X2
~ 5.5
r.q
"-
t.!>
C
'-'l
I 4.5
3.5
1.5 2.5
-[De(H)
Fig. 9.1. Computed convergence rate for the pure convection problem (from Jiang
1993)
9.2.1 Introduction
U = g on Iin, (9.33b)
where g is the specified data on the inflow boundary Fin. The characteristics of
problem (9.33) are the streamlines defined by (9.3), and the analytic solution
of problem (9.33) is constant along a characteristic. The value of this constant
is equal to the given value of g at the intersection of this characteristic and the
inflow boundary. Thus the solution will be discontinuous with a jump across
a characteristic curve for two-dimensional cases or a characteristic surface for
three-dimensional cases, if the boundary data g is discontinuous.
Usually one uses finite difference or finite element methods based on a
me~ which may not be adapted to fit the characteristics of the particular
problem. In such a case, if the exact solution has a jump discontinuity (con-
tact discontinuity) across a characteristic, all conventional finite difference
and finite element methods will produce approximate solutions which either
oscillate or smear out a sharp front. Finding accurate approximations of the
discontinuous solutions of hyperbolic equations has been a persistent difficult
task in modern numerical mathematics and computational physics.
9.2 Contact Discontinuity 209
This trouble comes from the fact that all existing conventional methods
are based on discretization of the following equation:
au + f32-
f31-
au + f33-
au = 0 in n (9.34)
ax ay az
in the Cartesian coordinates instead of (9.33a) in the streamline coordinates.
If the boundary data 9 is discontinuous, then (9.33a) and (9.34) are not equiv-
alent mathematically. Equation (9.33a) admits a solution that is discontinu-
ous in the direction a: = (Q b Q2, Q3) which is orthogonal to the streamline
direction (3, because in (9.33a) there is no derivative term with respect to
the a: direction. However, if the solution is discontinuous, (9.34) and its dis-
cretized version do not hold, since at least one of the derivatives with respect
to the coordinate direction approaches infinity across the jump. Common
methods do not account about this trouble, and force (9.34), which looses
meaning in the "shock" cells or elements where the front is located, to be
satisfied everywhere, and this leads to an oscillatory and diffused solution
around the contact front.
If we can identify the "shock" elements, and permit the equations in
the "shock" elements not to be satisfied while requiring that the equations
in remaining elements be satisfied exactly, the solution will then not have
oscillations. This idea can be realized by minimizing the L1 norm of the
residuals of the overdetermined linear algebraic equations.
The L1 idea can be explained as followings. In tp.e usual L2 (least-squares)
curve fitting approach, the L2 procedure does its best in the sense of least-
squares of the residual to make the curve pass by all of the data. If the data
are smooth, the L2 fitting leads to a very good approximation. However, if the
data contain abrupt changes, the L2 procedure will produce an oscillatory and
diffuse curve around sharp changes. In such a case, the L2 fitting makes the
use of individual datum equally important. The tendency of L1 fitting is to
give up the outliers in the data and to require the remaining data be satisfied
exactly. Therefore, L1 fitting is the choice for discontinuous functions. The
same thing happens in the L2 and L1 solutions of discretized hyperbolic
equations. The L1 approach translates into a capacity to totally give up the
equations in the "shock" elements, in which the discretized scheme is not
valid, while making the residuals in the smooth elements be zero exactly.
k nrn Urn In
210 9. Convective Transport
or simply
kiu = Ii, i=I, ... ,n, (9.36)
where the number of equations n is greater than the number of unknowns
m, and the rank of the augmented matrix (see (4.81)) equals to m + 1. The
L1 method is a special case of the general Lp (1 ~ p < 00) methods that
minimize the following summation of the pth powered absolute value of the
residuals:
Ip(u) = W11kuU1 + k12U2 + ... + k1mum - /tIP
+ W21k21U1 + k22U2 + ... + k2mum - hiP
+
+ Wnlkn1U1 + kn2U2 + ... + knmum - InI P, (9.37)
where (Wi> 0, i = 1, ... , n) are the weighting factors. We have already stud-
ied the case p = 2 in Sect. 4.11. Since 12(u) is a continuously differentiable
function of u, the minimizer of 12(u) can be found easily by using the cal-
culus, and the L2 minimizer is the solution of the normal equation (4.83).
However, the situation in the case p = 1 becomes difficult. In fact, it is not
a simple task to find a L1 minimizer. Since the residuals are linear functions
of u, It (u) is a piecewise linear function of u, and its derivatives with re-
spect to u are piecewise constant. In other words, the L1 norm is not globally
differentiable. Therefore the calculus is useless for finding the L1 minimizer,
except for the problems with only one unknown. It is for this reason that
people rely on linear programming to deal with L1 problems.
In the rest of this section we review some basic knowledge about the
L1 solution to the linear algebraic equations. More details can be found in
Bloomfield and Steiger (1983).
Let us first consider the simplest overdetermined system with only one
unknown:
k 1u = It,
k2u = 12,
= ... ,
knu = In, (9.38)
where we assume that the rank of the augmented matrix is 2, and ki (i =
1, ... , n) is not equal to zero. The L1 solution problem is to find a minimizer
of the L1 distance function
n n
11(u) =L Ikiu - Iii = L
IRi(U)I· (9.39)
i=1 i=1
Now imagine that /i/ki ~ 1i+1/ki+1, which can always be arranged by
renumbering the data. If we restrict u to the interval (fq/kq, I q+1/kq+1) ,
11 (u) in (9.39) becomes
9.2 Contact Discontinuity 211
q n
It(u) = L Ikil(u - fi/ki) - L Ikil(U -/i/ki ). (9.40)
i=1 i=q+l
Differentiation reveals that
Lemma 9.1 I 1 (u) in (9.39) has a minimizer u = fi/ki for some i = 1, ... , n,
say i = q. Thus, the Ll solution to (9.38) satisfies exactly at least one equa-
tion, i.e., the residual Rq(u) is zero.
11
u
Fig. 9.2. Graph of h(u)
to
u = 1,
u = 2. (9.41 )
The next is an important theorem on the L1 solution to system (9.35).
Prool: Assume that the initial u satisfies exactly some equations in (9.35). Let
Zu = {i : Ii = kiu} and suppose that the number of independent equations
in Zu is equal to s > O. Because s < m, we can choose v -# 0 E lRm such
that kiv = 0, Vi E Zu, and kiv -# 0 for some i f/. Zu' We then consider
u(t) = u+tv t E lR.
Algorithm 9.2 For each distinct subset J = {i1, ... , jm} of {1, ... , n} of size
m, when possible solve kiu = Ii, i E J for u, U is the u that minimizes 11.
At this stage we should not worry about the efficiency of the 11 method.
We appreciate its important and unique feature: For an overdetermined sys-
tem of linear equations the 11 solution exactly satisfies some equations and
permits the remaining equations not satisfied.
In this section we shall combine the 11 concept with finite element inter-
polation to obtain highly accurate and non-oscillatory solutions for contact
discontinuities. The 11 method must be based on an overdetermined sys-
tem. Fortunately, it is trivial to have an overdetermined system in the fi-
nite element context. As discussed in Sect. 4.11, the LSFEM with numerical
quadrature is equivalent to the weighted least-squares finite element collo-
cation method, in which the equation residuals are first collocated at the
interior points in each element, then the algebraic system is approximately
solved by the weighted least-squares method. The Gaussian points for calcu-
lating the element matrices in LSFEM correspond to the collocation points
in the collocation method. If the order of the Gaussian quadrature or the
number of quadrature points is appropriately chosen, the LSFEM amounts
to solving an overdetermined system. For the same reason, the 11 finite el-
ement method with the Gaussian quadrature is equivalent to the 11 finite
element collocation method.
Let us consider problem (9.33). The 11 method is based on solving a
weighted 11 problem: Find the minimizer Uh of
Nelem NCauss
11 (uh) = L (L
j=11=1
Wl w dRdlJ(€z,7]I)I).,
J
(9.43)
in which
(9.44)
and RI stands for the residual at each Gaussian point, NGauss denotes the
number of Gaussian points, WI is the weight of the Gaussian quadrature,
WI is an additional weighting factor, IJI is the determinant of the Jacobian
matrix, and (€I, 7]1) is the local coordinates of the Gaussian points. As usual,
Uh can be expressed as
Nnode
Uh(€, 7]) = L
1/Jm(€,7])U m, (9.45)
m=1
where "Nnode" is the number of nodes in an element, 1/Jm denotes the shape
functions, and Um is the nodal values. In order to make problem (9.43) mean-
ingful, the number of Gaussian points NGauss should be chosen such that
Nelem x NGauss is greater than the number of the total unknown nodal val-
ues.
214 9. Convective Transport
where n = {(x, y) E IR2 : 0 < x < 1, 0 < y < I} is the unit square with the
boundary r. The inflow boundary conditions are
U =2 on n = {(x, y) E r: x = O}, (9.46b)
U = 1 on r 2 = {(x, y) E r :x > 0 and y = O}. (9.46c)
Equation (9.46) represents uniform flow along straight lines inclined at an
angle of ,8 with respect to the x-axis. The jump discontinuity occurs along
the line y = xtan(,8). In this case, the analytic solution is
U = 2 on and above the line y = xtan(,8),
U = 1 below the line y = xtan(,8).
Consider a uniform n x n mesh generated by linear triangular elements,
such as that shown in Fig. 9.3. We use the finite element collocation method
with one Gaussian point in each triangle. In this approach there are 2n2
elements for n x n grids, so we have 2n 2 equations. Since there are (n + 1)2
nodal values and (2n + 1) boundary conditions, the number of unknowns is
(n + 1)2 - (2n + 1) = n 2 • That is, the number of equations is two times
of the number of unknowns. Therefore, the Ll method amounts to solving
an overdetermined system. It does not make sense to take more collocation
points in each element, because in a linear triangular element aUh/aX and
aUh/ay are constant, and thus the residuals at different points in an element
are the same. Now the Lp norm of the residuals is defined as
Nelem
Ip = L WjIRjIP, (9.47)
j=1
where the Gaussian weights and the Jacobians have been suppressed, since
they are constant and thus have no effect. The general Lp method is based on
minimizing the total Lp norm of the equation residuals in the whole domain.
Here we must emphasize the words "total" and ''whole'', since the correct
solution or the minimizer of the total Lp norm may not minimize the Lp
norm for each element or for a local subdomain.
For two typical linear triangles shown in Fig. 9.4, the residuals of (9.46a)
can be expressed as
Rupper = Une - Unw + tan (,8) (Unw - Usw ), (9.48a)
Rlower = Use - Usw + tan(,8) (Une - Use), (9.48b)
in which the uniform mesh size h has no effect and thus has been suppressed.
9.2 Contact Discontinuity 215
6
3Q---------~---------Q9
2~--------~---------e8
The unknown is Une = U*. By using (9.48) the L1 norm of this problem can
be written as
11(U*) = IU* - UL + "I(UL - UL)I + IUR - UL + "I(U* - UR)I, (9.49)
where "I = tan(35°).
Unw~--------------..QUne
, ,
,,
,
,,
upper
, ,,
triangle ,,,
,,
,,,
,
,,
: 75' Fig. 9.4. Two typical linear tri-
,
angular elements
Usw t'J-------.....'--"'-o4) Use
A
15.0
\
- - L1
\\ --- L2
10.0 \
\
\
,so
5.0
where WI denotes the weight set, which depends on the information from the
previous step, f is a very small positive number, for example f = 1.0E-20,
to prevent an overflow in computation, and all other notations have been
defined in (9.43) and (9.44).
The IRLSFEM would begin with the initial weight WI = 1. This first
step is nothing but the least-squares method introduced in Sect. 9.1.3. The
result then determines a new set of weights by (9.54). In the second iteration,
the residual IRzI is larger in the "shock" elements. Thus the weight WI for
the "shock" elements is smaller, and their influence becomes less important.
This procedure is repeated until IIUhcurrent - uhprevlou.11 is small. Our numer-
ical experiments reveal that the difference between the residuals of "shock"
elements and those of their neighboring elements in the first least-squares
solution is not significant enough. We put the sixth power in (9.54) in or-
der to additionally increase the importance of "smooth" elements and reduce
the contamination of "shock" elements, and also accelerate the convergence.
This is reasonable, since we want to eliminate completely equations which
9.2 Contact Discontinuity 219
will have nonzero residuals from the system. This trick is applicable because
the non-weighted least-squares method is good enough to locate the "shock"
elements. That is, in the results of LSFEM, the absolute value of the residuals
in the "shock" elements is always greater than that in other elements.
We may use another simple and reliable "shock" indicator - the variation
of nodal values in each element - instead of the residual. The variation is
defined as
Nnode
v= L IUm -Um - 11 U o = UNnode. (9.55)
m=1
Therefore, the following weight is suggested:
1015 if IVI~revious < 10- 7 ;
W/= { 1
otherwise. (9.56)
I VI:reV!OUB
Here some measures have been taken to prevent overflow. The advantage of
using the variation as a "shock" indicator is as follows: Once the jump in
the boundary data is given, we may know the exact values of the variation
in the "shock" elements in advance. There are only a few possible values,
which depend on the type of finite element and are independent of the shape
and size of the particular element, and have no relation with the location of
quadrature points.
The implementation of this reweighted least-squares method is relatively
straightforward. If a least-squares finite element code is already available, it
needs only a few additional lines of FORTRAN statements.
Table 9.1. Nodal values of IRLSFEM solution for constant convection problem
(50 triangles) (from Jiang 1993)
Fig. 9.6. Contours of IRLSFEM solution for constant convection problem (50
triangles) (from Jiang 1993)
which means that the four discretized equations at four Gaussian points are
independent in this case. (If the flow is inclined at an angle of 45 0 or 135 0 with
respect to the x-axis, we have three independent equations, because the lo-
cation of Gaussian points is symmetric.) All together we have 4n 2 equations
involving n 2 unknown nodal values. Therefore, we deal with an overdeter-
mined system.
The contours of the least-squares solution on a mesh with 40 x 40 bilinear
elements are illustrated in Fig. 9.7. This approximate solution is reasonably
good, although the discontinuity is smeared out. Taking this initial least-
squares solution, and after 8 steps of processing, we obtained a clean, non-
diffusive solution illustrated in Fig. 9.8.
Fig. 9.7. Contours of LSFEM solution for constant convection problem (40 x 40
bilinear elements) (from Jiang 1993)
Fig. 9.B. Contours of IRLSFEM solution for constant convection problem (40 x 40
bilinear elements) (from Jiang 1993)
Fig. 9.9. Contours of LSFEM solution for circular convection problem (100 x 100
bilinear elements) (from Jiang 1993)
Fig. 9.10. Contours of IRLSFEM solution for circular convection problem (100 x
100 bilinear elements) (from Jiang 1993)
9. Convective Transport
I-
~ -'- I I I I I I I I
!boO 0.2 0.4- 0.6 O. 1.0
X
Fig. 9.11. Boundary distribution of u for circular convection problem (from Jiang
1993)
used for problems with purely smooth solutions such as problem (9.32). In
the smooth case the L1 method will take elements with larger gradients as
"shock" elements and completely ignore them, and will then yield a wrong
discontinuous solution. For problems with smooth data, we should use the
ordinary LSFEM.
If the given data 9 in problem (9.33) are varying along the boundary
I1n with large gradients and a very small jump, any method will have dif-
ficulty. This is related to the numerical measure of discontinuity which is
different from the mathematical definition of discontinuity. That is, we must
distinguish discontinuity from large gradients. The distinction can be realized
easily by using fine meshes such that the difference between the nodal values
at any two neighboring nodes on I1n is less than the jump and by specifying
an appropriate level for a particular problem. Only if the variation indicator
of an element defined in (9.55) exceeds this level, should the contribution of
this element be eliminated.
The choice ofthe sixth power in (9.54) or (9.56) is based on our experience.
In fact, the method is not sensitive to the number of the power. The 3rd,
4th, 5th, 7th ... powers, all of them, work well. The only difference is that the
number of iterations to convergence is different. Here the essential issue is
how to completely eliminate the equations in the "shock" elements, and it
doesn't matter what means are employed.
Lowrie and Roe (1994) tried to analyze the L1 method and the itera-
tively reweighted least-squares method, and claimed that they had "proved"
a "distressing theorem" which says that correct discontinuos solutions cannot
9.3 Transient Problems 225
minimize any weighted residuals and the minimizer of any weighted residuals
cannot be a correct solution. Clearly, their statement is incorrect, because it
is against the residual minimization theorem (Theorem 4.2) which is rigor-
ously proved in a very general situation. In particular, their "mathematical
analysis" is erroneous. They mistakenly chose a segment of solution to inves-
tigate, and found that the correct solution does not minimize the Ll norm of
a sub domain, and hence made their conclusion. As pointed out in Sect. 9.2.3,
the minimizer of the global Ll norm may not minimize the local Ll norm of
a subdomain. Moreover, the Ll norm is the summation of the absolute values
of the element residuals, and thus is a piecewise linear function; therefore as
emphasized in Sect. 9.2.2, to find a Ll minimizer is not a simple task that can
be performed just by using easy calculus which they adopted. Unfortunately,
in their calculation of the derivative of IRil with respect to the unknown, the
sign of the absolute value was completely ignored.
u = Uo in!l at t = 0, (9.58c)
where un is for u{x, t n ), Llt = tn+l - tn, and the subscript "t' denotes the
order of temporal derivative at tn. Equation (9.58a) permits restatement of
the successive derivatives with respect to time in (9.60) as
Ut = -{3 . \1u = -u/3, (9.61a)
Utt = -{3 . \1Ut = {3 . \1{{3 . \1u) = u/3{3, (9.61b)
Uttt = Ut/3{3' (9.61c)
Substituting (9.61) into the Taylor series (9.60) and rearranging some
terms yields the following semidiscrete equation:
(9.62b)
Vv E V. (9.63)
For simplicity of exploration we consider a scalar, one-dimensional version
of problem. In the case of a uniform mesh of linear elements, (9.63) takes the
following form at an interior node j:
(9.64)
9.3 Transient Problems 227
or
(9.72)
From (9.7), (9.72) and (9.69) we obtain
lIell~ + 02Llt2I1e,8l1~ + OLltlel~ ~ lip + OLltp,8)II~
~ C(h 2 (r+1) + 02Llt2h2r)lIcPlI~+1' (9.73)
where C is a general constant independent of hand cP, and may have a
different value when it appears in different places. If the time-step Llt = O(h),
we can then use (9.73) to establish the optimal error estimates for each time-
step solution:
Ilello ~ Ch r +1llcPllr+b (9.74a)
The case for linear elements over a uniform mesh is important and readily
amenable to analysis. Combining element contributions for two linear ele-
ments adjacent to node j, and arranging terms to correspond to the order in
(9.76), we obtain the nodal equation
[1 + G-02c2)82](Uj+1_ Uj)
(9.80)
A standard von Neumann stability analysis can now be applied to the
recurrence relation (9.80) for LSFEM formulation. Consider a solution to
(9.78) of the form
(9.81)
where k is the wave number (wavelength>. = 27r/k) of mode eFi'kx so that
u(x, t + L1t) = u(x, t)e-Fi'k{jilt, (9.82)
where G = e-Fi'k{jilt is the complex amplification factor for the exact solu-
tion. The nodal values of the finite element solution can be written as
(9.83)
If IZ21 > IZ11 or equivalently IZ212_IZl12 > 0, then IGLSI < 1 and the scheme
is stable. Using the above expressions and simplifying,
1
for-<O<l.
2- -
Hence, the scheme is stable if ~ ~ 0 ~ 1 at all CFL (Courant-Friedrichs-
Lewy) number.
In the limit as kh --t 0 we have the asymptotic expansion for GLS :
GLS = 1 - Hckh - Oc2(kh)2 + H(Pc3(kh)3 + O((kh)4).
For 0 = 1/2, GLS = e-v:::rckh + O((kh)3) so the scheme is second-order
accurate in the time domain, as anticipated.
The amplification factor IGLS I determines the growth or decay of modes in
the solution. The control of oscillation is also important, and this is influenced
by both the relative amplification and the phase error of the modes. The phase
error can be expressed as arg(GLs)/(-ckh) in which
-1 [Im(G Ls )]
arg(GLs) = tan Re(GLs ) '
The numerical properties of the scheme are shown in Fig. 9.12, given the polar
diagrams of the phase and amplitude errors versus kh for several values of
the CFL number c.
i
\\
\,
\\
\
0.50
\
u
0.25
0.00+----l--~~:::f_---'t7----
-0.25+---..,......--....,..----,---.,-----.
0.00 0.20 0.40 0.60 0.80 1.00
Fig. 9.13. Propagation of a sine wave, comparison of LSFEM (0 = 0.5) with the
Taylor-Galerkin method, initial condition at t = 0 (left) and solutions at t = 0.6:
-, LS, c = 0.75; - - -, TG, c = 0.75; -. -, LS, c = 1.25 (from Carey and Jiang
1988)
1.00.,-------d-/
i
\I
0_75 \
\
\
i
i
\
I
0.00 L........................................i.>.,.<=-------
V
-0.25+---..,......--....,----,-----.---......,
0.00 0.20 0.40 0.60 0.80 1.00
Fig. 9.14. Propagation of a front, comparison of LSFEM (0 = 0.5) with the Taylor-
Galerkin method, initial condition at t = 0 (left) and solutions at t = 0.5: -, LS,
c = 0.75; - - -, TG, c = 0.75 (from Carey and Jiang 1988)
234 9. Convective Transport
(31(X,y) = wy,
(32 (X, y) = -wx. (9.86)
The flow domain is a square. The lower-left and the upper-right coordinates
of the domain are (-16, -16) and (16,16), respectively. The peak of the hill
is initially located at (-8,0), and the initial distribution of the concentration
is as follows:
u(x y) = { ~(1 + cos~d) d < 4, (9.87)
'0 d~ 4.
where d is the distance from the peak. The angular velocity w is adjusted so
that one revolution takes 300 time-steps. This corresponds to a CFL number
of 0.16755 at the peak of the rotating cone.
The Crank-Nicolson scheme for time discretization together with LSFEM
has been applied to this problem by Burrell et al. (1995). Figure 9.15 shows
the computed distribution of the concentration by using 32 x 32 nine-node
quadratical (Q2) elements after 1,2 and 4 and 6 revolutions. The quality of the
solution can be recognized by noting that the peak maximum is maintained
at about 1.0 throughout the course of six revolutions.
Rotating Puff in a Parabolic Angular Velocity Flow Field. This prob-
lem was originally proposed by Odman and Russell (1993) for further testing
numerical methods, and solved by Burrell et al. (1995). In this problem, the
initial puff concentration is located at (-8,0) and is given by
(9.88)
Both x and y range from -16 to 16. The flow field is given as follows:
x2 +y2)
(31 (x, y) = wyJx 2 + y2 ( 1 - W2 ' (9.89a)
(9.89b)
The angular velocity is set to 0.01 so that one full revolution takes 240
time-steps. This corresponds to a CFL number of 0.2094 at the peak of the
puff. The parameter W is chosen as 14.3. As compared with the results of
Odman and Russell (1993), the initial puff distribution in (9.88) has much
steeper gradients than that in the rotating cone problem, and the analytical
solution under the conditions (9.88) and (9.89) also has steeper gradients
around the puff after one revolution.
The analytical solution which can be obtained by the method of charac-
teristics is shown in Fig. 9.16 at different stages of a revolution. Notice that
after one revolution, the concentration distribution is much distorted from its
initial Gaussian distribution, and parts of distribution have steep gradients.
Figure 9.17 shows the results of LSFEM using the Crank-Nicolson scheme
and 32 x 32 nine-node quadratic elements. The results compare well with the
analytic solution in Fig. 9.16.
9.3 Transient Problems 235
c
c
1.IXI
LOO
0 .'711
QAI5
0 .<4
CU7
•
0~7
Fig. 9 .1 5. Concentration distribution of the rotating cone after (a) one, (b) two ,
(c) four and (d) six revolutions by using the Crank-Nicolson LSFEM and 32 x 32
quadratic elements (from Burrell et al. 1995)
236 9. Convective Transport
(a) (b)
(c) (d)
Fig. 9.16. Analytical solution for the rotating puff in a parabolic angular velocity
flow after (a) 0.25, (b) 0.5, (c) 0.75 and (d) 1.0 revolution (from Burrell et al. 1995)
9.3 Transient Problems 237
(a) (b)
(c) (d)
Fig. 9.17. Concentration distribution of the rotating puff computed by the Crank-
Nicolson LSFEM and 32 x 32 quadratic elements after (a) 0.25, (b) 0.5, (c) 0.75
and (d) 1.0 revolution (from Burrell et al. 1995)
238 9. Convective Transport
Using this specified velocity field in the least-squares finite element analysis,
the finite-element system follows directly from the introduction of the finite
element expansion for u into (9.67) with () = 1. The unit square domain is
discretized to a uniform 40x40 grid of bilinear elements with the initial data
u(x, y, 0) = 0 and the injection value u(O, 0, t) = 1 for t ~ O. The equi-spaced
solution contours for concentration u(x, y, t) are shown in Fig. 9.18 at t = 0.05
and 0.3 and the fixed time-step Llt = 0.0025. We note that there is numerical
dissipation of the propagating front and no oscillation .
•
Fig. 9.18. Concentration contours for the water-oil displacement problem: (a) at
t = 0.05, (b) at t = 0.30 by using the backward-Euler LSFEM and 40 x 40 bilinear
elements, .jt = 0.005 (from Jiang and Povinelli 1990)
9.4 Concluding Remarks 239
The LSFEM described in Sect. 4.8 can be used for the solution of both
steady-state and time-dependent convective transport problems. The LS-
FEM naturally and properly accounts for the role of the characteristics in
multi-dimensional hyperbolic equations. The LSFEM automaticly acts in a
manner similar to upwinding and requires no free parameters. The method
has nearly optimal or optimal rate of convergence. The backward-Euler and
the Crank-Nicolson LSFEM are unconditionally stable. Although the gov-
erning hyperbolic equations are non-selfadjoint, the method still produces
symmetric and positive definite matrix systems, and hence iterative methods
such as the conjugate gradient method can be employed. In particular, for
transient problems smaller time-steps are allowed for higher accuracy without
the penalty of increased computing time.
For pure convection problems with contact discontinuities, conventional
finite difference and finite element solutions in general exhibit large spurious
oscillations even far from the jumps and often are not close to the exact so-
lution. New procedures based on the L1 concept and the iteration of LSFEM
for the solution of contact discontinuities have been developed. The overde-
termined algebraic system is inherently obtained by choosing an appropriate
number of Gaussian points in the formation of element matrices. Through L1
minization or reweighting in the L2 method, the contamination of "shock"
elements is eliminated. These finite element methods capture discontinuities
in bands of elements that are only one element wide on both coarse and fine
meshes. The solution of this method has neither smearing nor oscillation, and
has superior accuracy. The iterative reweighted LSFEM is simple, robust and
efficient.
The LSFEM is general, very simple and easy to formulate and program.
This lays down the foundation of the unified LSFEM for complex phenomena
such as (1) convection-diffusion with small or vanishing diffusivity where
convection dominates the fluid flow; (2) transonic flows in which the flow
in subsonic regions is governed by elliptic equations and in supersonic bulbs
by hyperbolic equations; and (3) two-fluid flows where the location of the
interface is governed by pure convection of the so called color function. In
all these cases, the LSFEM has a single unified formulation. However, other
methods need different schemes in different regions or separate schemes for
different equations.
The idea of LSFEM for convection problems was first appeared in Nguyen
and Reynen (1984). They presented a space-time LSFEM for the advection-
diffusion problems. Their numerical results show that the use of upwinding
techniques or the Taylor-Galerkin approach in finite elements is not necessary
when the least-squares weak formulation is extended into the time domain
using standard shape functions. Donea and Quartapelle (1992) compared
the space-time LSFEM with the Crank-Nicolson LSFEM described in Sect.
240 9. Convective Transport
9.3.2, and found out that the space-time approach is less accurate and more
dissipative.
The LSFEM with the backward-Euler and Crank-Nicolson time dis-
cretization was first published in Carey and Jiang (1988). This work was
originally based on stimulating discussions between Srinivas Cheri and Jiang.
Later, Li (1990) proposed a characteristic LSFEM. An accurate method for
treating transient convection problems was introduced by Park and Liggett
(1990, 1991) by combining the Taylor-Galerkin and least-squares concepts.
Due to the presence of second-order spatial derivatives, elements of C 1 type,
such as the Hermite cubic element must be used in this Taylor LSFEM. This
method represents a valuable step towards simulations using relatively coarse
meshes.
Chen (1992) provided an error analysis of the Crank-Nicolson LSFEM
and her version of the method for a general linear first-order hyperbolic sys-
tems, namely, the Friedrichs system (Friedrichs 1958). The optimal error es-
timate given in Sect. 9.3.2 is a simplified version of her results. The analysis
for steady-state problems in Sect. 9.1.3 was given in Jiang (1991, 1993).
The radical Ll procedure for non-oscillatory discontinuous solutions was
first proposed by Lavery (1988, 1989) for one-dimensional problems in the
finite difference context. Because standard finite differencing leads to deter-
mined linear algebraic systems, in order to obtain an overdetermined system
Lavery relies on non-traditional tricks, such as gradually adding a small vis-
cous term to the one-dimensional Burger equation. However, it is very difficult
to extend these tricks to two-dimensional problems.
The Ll method based on overdetermined systems generated by numeri-
cal quadrature in the finite element context for multi-dimensional problems
including an inexpensive solution method - the iterative reweighted LSFEM,
was proposed by Jiang (1991, 1993). This Ll approach is much more easy
and general.
10. Incompressible Inviscid Rotational Flows
where u(x, t) is the velocity vector, p(x, t) is the pressure divided by the
constant density of the fluid.
Equation (IO.lb) allows the pressure to be correct up to any constant. To
determine the pressure uniquely we require that the mean pressure over il is
zero, namely,
!npdX=O. (IO.lc)
Vx
au
(at - u xw + Vb - f) = 0 in {} x (0, T), (10.4a)
au
V . (at - u x w + Vb - f) = 0 in {} x (0, T), (1O.4b)
au
n· ( - - u x w + Vb - f) = 0 on r x (O,T). (1O.4c)
at
By noting that the conservative force f has a potential V, namely, f = - VV,
and using the vector identity (A.6), (10.4a) can be written as
244 10. Incompressible Inviscid Rotational Flows
ow + (u· V)w -
- (w· V)u = 0 in n X (O,T]. (10.5)
at
Equations (lO.4b) and (lO.4c) are the Poisson equation for the total pressure
with its Neumann boundary condition:
Llb = V . (u x w) +V .f in n x (0, T], (10.6a)
ob og
an =n·(uxw)+n·f- at onrx (O,T]. (1O.6b)
It is clear from the above derivation that the computation of the pressure
can be separated from that of the velocity and the vorticity.
Now we can write the governing equations of incompressible inviscid ro-
tational flows in the following form:
(1O.7c)
the solution for the vorticity with an optimal rate of convergence provided
that the time step is properly chosen.
By contrast in three dimensions, due to the existence of the term (w·
V')u in (1O.7a), vorticity is no longer conserved along the trajectories of the
particles. Moreover, this term is responsible for the difficulties which arise
in the construction of a global solution of the Euler equation. Since w is
associated with the gradient of the velocity, we may conjecture that (w· V')u
is of the order w 2 and ow / f)t w 2 • Because the ordinary differential equation
I'J
aX = X2, (1O.9a)
f)t
ow
-at + (u· V')w - (w· V')u = 0 in n x (O,Tj, (10.12b)
in n x (O,Tj, (1O.l2c)
246 10. Incompressible Inviscid Rotational Flows
(1O.13)
in n x {O,Tj, (10.16b)
10.3 The Least-Squares Finite Element Method 247
8v 8v 8v 8p
at + u 8x + V 8y + 8y = fy in il x (0, TJ, (10.16c)
8u 8v
---+w=O in il x (O,Tj, (1O.16d)
8y 8x
8w 8w 8w
-+u-+v-=O in il x (O,Tj. (10.16e)
8t 8x 8y
In Sect. 10.1 we point out that one could use the splitting method to con-
struct the solution. Of course, one may solve all five equations in (10.16)
simultaneously at each time-step. Numerical tests conducted by Wu et al.
(1994), and Wu and Jiang (1995b) show that the simultaneous solution of
(10.16a)-(10.16e) generally results in fast convergence and good accuracy.
When the vorticity field has discontinuities such that the spatial derivative
of the vorticity approaches infinity across the discontinuities, we simply do
not include (10.16e) in the computation.
Before applying LSFEM, the equations should be discretized in the time
domain. As usual, the general () form can be utilized:
8un+! 8v n+!
-8- + -8- = 0 in il, (1O.17a)
x y
un+1 ___
_-..,. un + () (un+!
8u _n_
+1 + V n+! 8u n_
_ +! 8 pn+1 )
+ __
..1t 8x 8y 8x
()) ( U n - n
8u+ 8un+8-pn ) fn+l
+(1 - vn - = in il, (10.17b)
8x 8y 8x x
vn+1 _ v n + () (un+!
____ 8v _n_
+! + vn+! 8v n_
_ +1 8 pn +!)
+ __
Llt 8x 8y 8y
()) ( u n - n
8v+ 8v n+8-pn ) _ f n+1
+(1 - vn - - Y in il, (lO.17c)
8x 8y 8y
in il, (1O.17d)
in il, (1O.17e)
where the superscript 'n' denotes the previous time-step and 'n + l' the
current time-step.
Usually, two choices of (), () = 1/2 or () = 1, can be made in the computa-
tion. () = 1/2 represents the Crank-Nicolson scheme, provides second-order
248 10. Incompressible Inviscid Rotational Flows
where k stands for the iteration count of the linearization and (u n+1 ) [0] = un.
System (10.17), after linearization, can be written in standard matrix
form as:
Al
au + A2 au + Ao!! = 1
f'l- f'l- (10.19)
vx vy -
or'
in which
0 0
(~[~I
at +0 08u
..1... 08u 0
8x 8y
(
..1... + 08v
Ikl
!!= Ao = 08v 0
8x at 8y
0 0 0
08u
8x
08u
8y 0
1.
o o
It r
n+l n+l
!
0 0 1 0
o~L
Ou 0 1 Ov 0 0
( 1
Ou 0 1
A,~ ~ -1
A2~ Ov
0 0 0
0 0 0 0
o
f Xn + l + un
at (1 _ O)(u 8u
_
8x
+V 8u
8y
+ !!1!.)n
8x
+ O(u 8u
8x
+V 8u)n+l
8y [k]
_
I = r+
y
l + vatn _ (1 - O)(u 8x
8v +V 8v
8y
+ !!1!.)n
8y
+ O(u 8v
8x
+v 8v)n+l
8y [k]
o
wn _ 8w + v 8w)n + O(u 8w + v 8w)n+l
at (1 _ O)(u ax 8y 8x 8y [k]
The first-order system (10.19) is then discretized in space following the stan-
dard LSFEM procedures given in Sect. 4.8.
10.4 Numerical Results of LSFEM 249
(a)
(b)
(c)
Fig. 10.1. Solution of the standing vortex problem: (a) Contours of vorticity, (b)
Absolute value of velocity, (c) Pressure, left-analytical solutions; right-numerical
solutions (from Wu and Jiang 1995b)
10.4 Numerical Results of LSFEM 251
Solutions of the pressure, velocity and vorticity after 600 time steps that
corresponds to t = 30 when the fluid in r < 2 has traveled 24 cycles are shown
in Fig. 10.1. The numerical solutions and the exact solutions compare very
well particularly for the velocity and pressure. Since in this case the vorticity
is discontinuous, it is reasonable that the vorticity solution is smeared.
From Sect. 10.2 we know that the kinetic energy should be conserved
in this flow. If significant numerical viscosity exists in a scheme, the kinetic
energy of the computed fluid flow will not be conserved. To compare the
numerical dissipation in the LSFEM and the SUPG method (Tezduyar et al.
1992c), Fig. 10.2 illustrates the history of the kinetic energy as the percentage
of the original. Clearly, the LSFEM maintains very good accuracy for this
test case all the time, and has almost no dissipation. Even after a very long
period of time t = 30, 99.8% of the kinetic energy is still retained.
104.0 r--,--.,--:---;--;----;-..,.---.-,--...,.----:--.,
"or' .
84.0 '----'----'--'---'---'----'---'----'--~-'---'-~
o 100 200 300 400 500 600
time step
Fig. 10.2. Time history of kinetic energy, - LSFEM, ... Tezduyar et al. (1992c)
(from Wu and Jiang 1995b)
¢ = p or w,
so that the vortex can accurately propagate out of the domain without being
artificially reflected back into the computational domain to contaminate the
rest of the flow field. This property at the outflow boundary is important in
many practical applications.
The computed contours of pressure and absolute value of the velocity at
different times are shown in Fig. 10.3. The distribution of the pressure and
the vertical velocity at the center line at different times is illustrated in Fig.
lOA. As demonstrated in these figures, the solution of LSFEM is of high
accuracy and has no visible smearing.
'----__110 I
L---IIL---_._I
'-----_IIL-----.J
.I
(a) (b)
Fig. 10.3. Vortex propagation: the computed solutions at different times: (a) Pres-
sure, (b) Absolute value of the velocity (from Wu and Jiang 1995b)
Shear Flow Past Solid Bodies. We first consider steady-state shear flow
past a circular cylinder. The center of this circular cylinder is placed at the
origin of the rectangular coordinates. The cylinder is exposed to uniform
shear flow with the velocity expressed by
u = Uo +woY,
v = 0,
where Uo and Wo are the given constants. The analytical solution to this
problem (Batchelor 1970) is
10.4 Numerical Results of LSFEM 253
/\
I "\
0.5
I \
I \
&> I \
·u
.,
.Q 0.0
\ I
>
\ I
-0.5 \ I
\ I \ I
-1.0
\I ',;
-1.5
0.0 1.0 2.0 3.0 4.0
0.0
... -----r---------,
\!
~ / "
\ ( ,
\ I '.
f \ I \
.," \ I
\
o.~ -0.5
\
\ (
I
\
\ (
~ : \ I
\ ...../ \./
Fig. 10.4. Vortex propagation: solutions at the center line (from Wu and Jiang
1995b)
1 2 2 Y 4 (x 2 - y2)
"2WOY + uoY - uoa r2 + woa 4r 4 '
81/J
u =
8y
where 1/J is the stream function, a is the radius of the circular cylinder and
r = (x 2 + y2)1/2.
The computational domain is a circular ring. The distance between the
outer boundary and the surface is 12 times the diameter of the cylinder.
254 10. Incompressible Inviscid Rotational Flows
The finite element mesh consists of 4302 bilinear elements with 4318 nodes.
Figure 10.5a shows a part of the mesh near the cylinder. The velocity and
the vorticity of uniform shear flow are prescribed on the inflow boundary;
the pressure (Poo = 0) is specified on the outflow boundary. The undisturbed
uniform flow is taken as an initial solution. A converged solution is obtained
after six Newton's linearizations.
In Fig. 10.5b the calculated pressure distribution on the cylinder surface is
compared with the analytical distribution. In Fig. lO.5c and d, the computed
streamlines in the vicinity of the cylinder are compared with the analytical
streamlines. The LSFEM results show excellent agreement with the analytical
solutions.
5 .0 .---~--.-~---.--~--.--~--,
00
-5.0
-10.0
-15.0 L-~_-,--~_-,--~_-,--~-.I
-1.0 -0.5 0.0 0.5 1.0
(a) (b)
(c) (d)
Fig. 10.5. Shear flow past a circular cylinder: (a) Finite element mesh, (b) Pressure
distribution on the surface (. LSFEM, - analytical), (c) Analytical streamlines,
(d) Computed streamlines (from Wu and Jiang 1995b)
10.4 Numerical Results of LSFEM 255
We then consider the same type of flow past a NACA0012 airfoil. The
outer boundary is 20 chords away from the airfoil. A mesh with 3521 bilinear
elements and 3640 nodes is used. There are 170 nodes on the surface of
the airfoil. The mesh near the airfoil is shown in Fig. 1O.6a. The computed
distribution of the pressure coefficient C p = (p - Poo)/O.5 on the surface is
shown in Fig. 1O.6b.
(a)
0.5 r----~------...,...---_r- _ __,
0.0
Q.
()
-0.5
problem looks simple at first sight, it turns out to be quite challenging compu-
tationally. It was demonstrated by Pulliam (1989), to the surprise of many,
that calculations done by almost all Euler codes available at the time for
flows over cylinders and ellipses at angle of attack produced lifting solutions,
and the solutions suffered from a general lack of consistency across mesh re-
finements, artificial viscosity parameters and algorithm types (e.g., central
differences, upwind differences, implicit methods, multigrid schemes, direct
solvers). Moreover, different converged solutions with lift varying from large
positive to large negative values can be obtained, even within the framework
of a single code and a single grid with varying algorithm parameters. Hafez
and Brucker (1991) attribute the above trouble to contamination caused by
artificial viscosity and vorticity due to discretization errors, and hence pro-
posed a finite difference scheme on staggered grids with averaging of variables.
Winterstein and Hafez (1993) suggested a way to avoid such difficulty by ad-
justing the location of the trailing edge stagnation point to produce zero lift
solutions. This approach is effective, but may be cumbersome to implement.
(a) (b)
Fig. 10.7. Uniform flow (0 = 5°) past an ellipse: (a) Finite element mesh, (b)
Computed pressure contours (from Wu and Jiang 1995b)
1.0
0.5
0.0
0.
()
I
• LSFEM
-0.5
- - analytical
-1.0
Newton's linearizations. A part of the mesh with 238 nodes on the surface
and the computed contours of the pressure are illustrated in Fig. 10.7a and b,
respectively. Figure 10.8 demonstrates that the computed pressure coefficient
on the surface is compared very well with the analytical solution. It is clear
that the calculated flow is nonlifting and attached to the solid surface. We
note that in the LSFEM no special treatment is needed.
In this chapter, low-speed compressible viscous flows are of interest. This sub-
ject has been neglected in finite element simulations, but has been intensively
studied using finite difference and finite volume methods.
We shall show that by employing the velocity, pressure, vorticity, temper-
ature, and heat flux as independent variables, the governing equations of this
type of fluid flows can be written as an almost-linear first-order system. In
this system, the principle part is the same div-curl-grad operator as that for
the incompressible flow-heat coupling problems studied in Sect. 8.7. There-
fore, the mathematical properties, such as the ellipticity for steady-state cases
and the permissible boundary conditions of this system are exactly the same
as that for incompressible viscous flows. Application of LSFEM based on this
system to compressible buoyant flows are provided. The presentation of this
chapter mainly follows the works of Yu et al. (1995a, 1996a,b).
11.1 Introduction
Low-speed flows with significant temperature variations are compressible due
to the density variation induced by heat addition. For example, significant
heat addition occurs in combustion related flow fields; inside chemical vapor
deposition reactors, strong heat radiation also results in significant density
variation. Although the flow speed is low, one must employ the compressible
flow equations to simulate such flows.
It is well known that the simple explicit time-marching method, which
can readily handle high-speed compressible flows for both steady-state and
time-dependent problems, fails miserably when applied to low-speed or in-
compressible flows. This is due to the fact that for stability, the time step
must be chosen inversely proportional to the largest eigenvalue of the system
which is approximately the speed of sound for slow flows. Thus, this method is
very time consuming. For incompressible flows, the speed of sound approaches
infinity, thus simple explicit methods do not work at all. Another equivalent
explanation of this difficulty is that the incompressible Navier-Stokes equa-
tions do not contain any time derivative of the pressure, hence the pressure
variation propagates infinitely fast. To overcome this difficulty, Chorin (1968)
260 11. Low-Speed Compressible Viscous Flows
added an artificial time derivative of the pressure into the continuity equa-
tion together with a multiplicative parameter {3. The free parameter {3 is then
chosen so as to reach the steady state rapidly. This idea can be extented by
adding the pressure time derivative to the momentum equations and by in-
troducing a second free parameter a. With these artificial terms the transient
nature of the system is changed, and the steady state can be reached quickly
by time-marching methods. Based on an analogy with the conjugate gradient
method such a method is called the preconditioned method, since the object is
to reduce the condition number of the matrix. Recently, noteworthy progress
in simulating low-speed compressible flows has been achieved by using this
preconditioning technique in the context of finite difference and finite volume
methods, see e.g., Merkle and Choi (1987, 1988), Choi and Merkle (1993),
Turkel (1987, 1993) and van Leer et al. (1991). However, the convergence of
this preconditioned method is dependent on the choice of the free parame-
ters. The optimal parameters depend on the dimensionalization, boundary
conditions and particularly the inflow conditions of the problem. In addition,
in the preconditioned equation set, one must specify the pressure boundary
condition which usually is derived based on the boundary layer assumption.
For recirculating flows, however, this approximate boundary condition poses
a significant error.
Since for small Mach numbers the compressible equations basically have
the same mathematical properties as the incompressible equations, the mixed
Galerkin finite element method discussed in Sect. 8.1.1 is suitable for low-
speed flows, as demonstrated by Einset and Jensen (1992) for flows inside
a chemical vapor deposition furnace. It has been mentioned many times in
this book that in the setting of the mixed Galerkin finite element method,
only certain combinations of the approximation functions for velocity and
pressure are acceptable for stable solutions. i.e., the LBB condition must be
satisfied. While the LBB condition is cherished for its mathematical elegance,
the condition renders no easy verification. For three-dimensional calculations,
few combinations of shape functions are acceptable. In addition, the final
coefficient matrix is nonsymmetric, and hence the inversion of such matrices
is not trivial.
All the above mentioned difficulties associated with conventional methods
have promoted the development of LSFEM for low-speed flow problems.
ap ap ap (au
-+u-+v-+p -+-av) =0 in n x (0, T], (ll.la)
at ax ay ax ay
au au au ap a [au 2 (au av)]
p-+pu-+PV-+-=J..l-
at ax ay ax ax 2ax- - -3 -+-
ax ay
a (au av)
+J..l- n x (0, T],
ay -+-
ay ax in (11.1b)
av +pu-
av av ap a (au av)
p-
at ax +pv- ax -ay +-
ay + -ay =J..l- ax
a [av -
+J..l- -2 (au
- + -av)] - pg n x (0, T],
ay 2-ay 3 ax ay in (11.1c)
(11.1e)
where p is the density, u and v are the velocity components in the respective
directions, () is the temperature, and tP is the viscous dissipation. Note that
the coordinate system is chosen so that gravity acts in the negative y direc-
tion. Equation (ll.la) is the continuity equation; (11.1b) and (ll.lc) are the
momentum equations; and (ll.ld) is the energy equation.
In order to reduce (11.1) to a first-order system, the following new vari-
ables are introduced:
'0- au av (ll.2a)
- ax + ay'
av au
W=---, (11.2b)
ax ay
a()
qx = -k ax' (11.2c)
a()
qy = -k ay' (1l.2d)
262 11. Low-Speed Compressible Viscous Flows
au
p at
au au ap
+ pu ax + pv ay + ax = f.J, 3 ax
(4 a'D aw)
- ay , (11.3b)
av
p at
av av
+ pu ax + pv ay + ay = f.J,
ap (43 a'D
ay + ax
aw)
- pg, (11.3c)
aqy _ aqx =
ax ay
° (11.4)
should be added to the above first-order equations to make the system have
good mathematical properties.
The above governing equations (11.2)-(11.4) are closed by the equation
of state
p = pR(), (11.5)
where R is the gas constant.
To nondimensionalize the equations (11.2)-(11.5), the following dimen-
sionless variables are introduced:
- p u v
p=-, ii= Uoo ' v=U'
Poo 00
- 'DL _ wL -I p'
'D=-, w=-, =--rJ2'
Uoo Uoo P
Poo 00
x
X= L'
- qy
qy = PooVoo
TT ()
00
Cp '
where Poo, Uoo , ()oo, and L are reference values of density, velocity, tem-
perature, and length scale. Note that special care is taken to deal with the
11.2 Two-Dimensional Case 263
8u
p at
8u
+ pu 8x + pv 8y +
8u 8p' 1
8x = Re
(48V
'3 8x -
8w)
8y , (11.6b)
80 80 80 ( 1)M2 (8 pl pl
(-y - 1)M2
P at + PU-8x + pv 8y - "{ - -at- + u-8p' 8 )
8-y = .;....:....--::Re~-
8-x + v-
1 80
qy = -Pe 8y' (11.6h)
8qy _ 8qx _ 0
(11.6i)
8x 8y-'
Note that the dimensionless bar notation has been suppressed in the above
equation system for convenience. The dimensionless numbers in these equa-
tions are defined as
M- Uoo Ra= 2€gL3
-~' /la '
264 11. Low-Speed Compressible Viscous Flows
v
Pr= -,
a
Re = Jlli1Fr
Pr '
Pe = RePr,
k C
a = PooCp ' 'Y = C: '
where M is the Mach number, Fr the Froude number, Ra the Rayleigh num-
ber, Pr the Prandtl number, Re the Reynolds number, Pe the Peclet number,
a the thermal diffusivity, and 'Y the ratio of specific heats. The temperature
difference parameter € is defined as
(h - ()e
€ = ()h + ()e' (11.7)
where ()h and ()e are the specified hot and cold temperatures in the thermal
convection problem, respectively.
V_8u 8v
- 8x + 8y' (l1.8e)
8v 8u
W=--- (l1.8f)
8x 8y'
1 8()
qx = -Pe 8x' (l1.8g)
11.2 Two-Dimensional Case 265
1 {)()
qy = -Pe {)y' (11.8h)
{)qy _ {)qx _ 0
{)x {)y - ,
(11.8i)
1 = p(). (1l.8j)
(11.9)
Substituting (11.9) into the continuity equation (11.8a), choosing the tem-
perature as a primitive variable, and replacing the density by the reciprocal
of temperature lead to
{)() {)()
u {)x + v 8y = ()V. (11.10)
(11.11)
The left hand side of (11.10) is the material derivative of temperature, which
can be substituted into the energy equation (11.8d), and we obtain
8qx8qy 'D - 0 (11.12)
8x+8y+ - .
The above equation directly correlates the compressibility effects with the
heat fluxes for low-speed flows. The nonlinear convective terms of the en-
ergy equation (1l.8d) in terms of the temperature now become an algebraic
expression in V.
266 11. Low-Speed Compressible Viscous Flows
(11.13)
au au ap' _ ab _ vw qx Pe ( 2 2)
Pu ax + Pv ay + ax - ax () + 2()2 u + v ,
(11.14a)
av av ap' ab uw qyPe 2 2
Pu ax + Pv ay + ay = ay + 7i + 2()2 (u +v ). (11.14b)
av au
---=W,
ax ay
(11.15b)
aw
ay +
aB
ax
= Re [vw _ qx Pe (
() 2()2 u
2+ 2)]
V ,
(11.15c)
aw aB '
--+-=-Re [uw
-+-qyPe
( u 2 +v)
2 ]
- -Re
-, (11.15d)
ax ay () 2()2 2fFr()
a()
ax = -Peqx, (11.15e)
a()
ay = -Peqy, (11.15f)
aqx aqy Pe ( )
ax + ay = +7) uqx +vqy , (11.15g)
(11.16a)
rot u = w, (11.16b)
(11.16d)
rot q = 0, (ll.16e)
"VO = -Peq, (11.16f)
where the notations ofrot and curl have been defined in Sect. 6.4.1; the right
hand side vector / in (11.16c) is defined as
~ _ q.. ~e(u2 +v 2) )
/-Re
-
( 8 28
_:!!!!! _ qy~e(u2 +v2) _ _ 1_ .
8 28 2eFr8
Note that all right hand sides in (11.16) are algebraic, and thus they have
nothing to do with the classification of this equation system.
The principle part of system (11.16) consists of a Stokes operator (11.16a)-
(11.16c) for the velocity vector, the scalar B and the vorticity, and a div-
curl-grad operator (11.16d)-(11.16f) for the temperature and heat fluxes. As
such, we arrive at an almost-linear first-order system with eight equations
involving seven unknowns: u, v, B, w, 0, qx and qy. The inconsistency be-
tween the number of the unknowns and the number of equations results in an
"overdetermined" problem. However, as emphasized many times in this book,
this "overdetermined" problem is a notion borrowed from linear algebra. For
partial differential equations, this interpretation leads to a misconception.
By introducing a dummy variable K, into the div-curl-grad system (11.16d)-
(11.16f) as explained in Sect. 6.4.1, it is easy to show that the system is elliptic
and properly determined (see Yu et al. 1995a and 1996a,b for details).
In this first-order elliptic system augmented by a dummy variable there
are eight unknowns governed by eight equations, thus on each boundary
four boundary conditions are required. To facilitate the discussion of bound-
ary conditions, the system of equations is divided into two groups: the flow
equations (11.16a)-(11.16c), and the heat equations (11.16d)-(11.16f). Ac-
cordingly, by virtue of the knowledge given in Sects. 6.6 and 8.2.2, we can
list some (not all) permissible boundary conditions for each group in Table
11.1.
In Table 11.1, n denotes the outward normal vector for the boundary;
n . q and n x q are the normal and tangential components of q. Without
losing generality, the symmetry condition is assumed to be with respect to
the x axis. Note that for certain combinations, a null boundary condition for
the dummy variable f<i, is invoked in order to make the system (augmented by
the dummy variable) well posed. It should be emphasized however that this
is only for the purpose of discussion, because the dummy variable is not used
in the numerical computation. Therefore, these null boundary conditions are
put in parentheses.
In the conventional second-order velocity-vorticity method used by Ern
and Smooke (1993), there have been endeavors to derive wall conditions for
the vorticity (see Sect. 8.8.2). For the preconditioning technique, one always
needs to impose a wall condition for pressure. Here, we show that since only
two boundary conditions are required from the flow equations, no wall bound-
ary condition for vorticity or pressure is needed because the no-slip condition
(u = v = 0) is enough to make the system well posed. Similarly, no boundary
condition for vorticity is needed for specified inflow or outflow conditions.
For the heat equations, we list only two types of boundary conditions: (1)
the specified temperature condition; and (2) the specified heat flux condition.
A linear combination of the two is also permissible (see Remark in Sect. 6.4.2).
For the first case, it is convenient to impose an additional condition for heat
flux tangential to the wall. This tangential heat flux condition can be deduced
from the specified wall temperatures. For the second case, we activate a null
boundary condition for f<i, in addition to the specified normal heat flux. Note
again that f<i, is a dummy variable, and the adoption of this pseudo boundary
condition poses no theoretical difficulty.
(11.17)
where Uj,j = 1"",7 are the unknown variables. These nonlinear terms are
linearized by Newton's method in the following fashion:
(11.20)
To proceed, the governing equations are cast into the following operator form
ready for finite-element discretization:
L8:!! = 1, (11.21)
where the linear operator L is defined as
L = Al :x + A2 ~ + A~, (11.22)
In each iteration step the standard linear first-order system (11.21) can be
solved by LSFEM, see Sect. 4.8.
the Prandtl number Pr = 0.71 (air), the Proude number and the aspect ratio
are unity. The variable B is specified to be zero at the center of the bottom.
For the grid refinement study, three meshes (32 x 32, 64 x 64 and 128 x 128)
were used. In all three meshes, grid lines were clustered near all four walls
to resolve high gradients of the flow properties. In all calculations, nine-node
quadratic elements are used, and the integration is performed using 2 x 2
Gaussian quadrature.
In the computation there are two loops of iterations: the outer loop up-
dates the coefficient matrices and source terms (see (11.21)) by Newton's
method; while the inner loop solves the variable increment O!! using the Ja-
cobi preconditioned conjugate gradient (JCG) method (see Chap. 15).
Figure 11.1 shows typical convergence numbers for the inner iterations by
the JCG method. The case shown here is Ra = 106 with 128 x 128 mesh.
Figure 11.1 shows that the error drops to lower than the sixth decimal digit
in about 300 JCG iterations. For these calculations, the specified criteria for
stopping the inner iterations is that error ~ 10- 7 or 300 iterations, whichever
is satisfied first. The error in the JCG method is defined as the marching dis-
tance of the CG method divided by the absolute value of the global unknown
vector.
-4.0
g
w
0' -5.0
..-
!
-6.0
Figure 11.2 shows typical convergence behavior of the outer loop by New-
ton's method for the same case. In about six iterations, the magnitude of
O!! reduces by about seven orders of magnitude. Since the criterion for the
inner iteration by the JCG method is set at 10- 7 , the outer iterations level
11.2 Two-Dimensional Case 271
1.0
G---e ContInuity
~ X-Momentum
off at approximately the same order of magnitude. Figure 11.2 also shows
quadratic convergence, which is a typical characteristic of Newton's method.
Figure 11.3 shows a comparison of the present result with previous data
reported by Chenoweth et al. (1986). The x-axis is the Rayleigh number on
a logarithmic scale, and the y-axis is the Nusselt number. Flow with large
Rayleigh number results in enhanced heat transfer by natural convection,
and therefore a larger Nusselt number is obtained. For all twelve calculations
with four Rayleigh numbers and three meshes, favorable agreements were
obtained. Only slight improvements were found for solutions using the fine
mesh.
Figure 11.4 shows a comparison of the predicted heat flux on the cold wall
for the case of Ra = 106 by using three different meshes. A slight discrepancy
can be observed for the solution between the 32 x 32 mesh and the other two
meshes. For 64 x 64 and 128 x 128 meshes, the solution is essentially same.
In Fig. 11.5, streamlines of simulated flow fields for different Rayleigh
numbers are presented. The streamlines are obtained by plotting constant
contours of the stream function 'lj;. The distribution of 'lj; is obtained by solving
the following equations by LSFEM:
o'lj; u
ox 8'
272 11. Low-Speed Compressible Viscous Flows
20.0 ~----r---~-"'---~---'------'
o 32 X 32
15.0 0 64 X 64
o 128 X 128
- - Chenoweth et al.
i 10.0
5.0
0.0
0.0 2.0 4.0 6.0 8.0
Log 10 (Ra)
Fig. 11.3. Comparison of the calculated Nusselt numbers with Chenoweth and
Paolucci's correlation (from Yu et al. 1996a)
1.0
0.8
0.6
>-
0.4 32 X 32
64 X 64
128 X 128
0.2
0.0
0.000 0.010 0.020 0.030
RIght Heat Rule
(c) Ra = 105
Fig. 11.5a-d. Streamlines of 2D buoyancy-driven gas flows (from Yu et al. 1996a)
274 11. Low-Speed Compressible Viscous Flows
: ~"----..::~~~~=:':~
___ ...........................
/./.;'----
L/,~
,
.......... " ...... _____ ----~ ~\
... _- --......
,\,~_,_._.
,. - ...... -- - :- " -' ~.. ~\ ~- .' .'.,.,""1
_' •••
'., '. __ - .... .. , I. ' • • • ,
\, ...... - ................................ ,.
,------- ..... - - - -: :
n .. . ,__....
---~
:~.:::.:;.::..- f~
~~~~~g ~
o o
in fl, (1l.24b)
where
w = \1 x u, (11.25b)
q = -k\1B, (11.25c)
\1 x q = 0, (11.26f)
\1. W = 0, (1 1. 26g)
11.3 Three-Dimensional Case 277
(11.26h)
Note that in the above system the dimensionless bar notation has been sup-
pressed for convenience. The dimensionless body force in the momentum
equation is f = (O, -pj{2fFr), 0). Also note that the irrotational condition
(11.26f) for the heat flux and the solenoidal condition (11.26g) for the vortic-
ity are added to make the system have better mathematical properties, see
Chaps. 6 and 8.
For low-speed flows (M ~ 0.3), the pressure derivative terms, the viscous
dissipation terms, and the buoyancy term in the energy equation (11.26c)
become negligible. In addition, according to the equation of state (11.26h),
the density and temperature become reciprocals of each other and we have
(11.9). Substituting (11.9) into the mass conservation equation (11.26a) leads
to
(u· \l)O = OV. (11.27)
Here the density is replaced by the reciprocal of temperature. As a result,
the dilatation V can be expressed by the following algebraic equation:
OV = -Pe(u· q). (11.28)
The left hand side of (11.28) is the material derivative of temperature, which
can be substituted into the energy equation (11.26c) to obtain
(11.29)
Similarly, in order to transform the nonlinear convective terms of the
momentum equations into algebraic expression, the total pressure b is intro-
duced:
(11.32a)
\l x u = w, (11.32b)
\l·w = 0, (11.32c)
278 11. Low-Speed Compressible Viscous Flows
Vxq = 0, (11.32f)
VB = -Peq. (11.32g)
Here all left hand sides are linear first-order derivatives; and all right hand
sides are algebraic, and thus they have nothing to do with the classification
of this system.
This set of equations consists of a div-curl system (11.32a) and (11.32b)
for the velocity, a generalized div-curl system (11.32c) and (11.32d) for the
scalar B and the vorticity, and a div-curl-grad system (11.32e)-(11.32g) for
the temperature and heat flux. Although this set has fifteen equations with
eleven unknowns, it is elliptic and determined (see Chaps. 6 and 8).
For the purpose of discussing the boundary conditions, we divide the sys-
tem of equations into two groups: the flow equations (11.32a)-(11.32d) and
the heat equations (11.32e)-(11.32g). Any combination of the permissible
boundary conditions for the flow equations listed in Table 8.1 and the per-
missible boundary conditions for the heat equations discussed in Sect. 6.4.2 is
permissible for low-speed equations in the div-curl-grad formulation (11.32).
Some of these combinations are listed in Table 11.2.
this study, four Rayleigh numbers, Ra = 103 , 104 , 105 , and 5 x 105 , are
considered with a temperature difference parameter E = 0.6. For all four
cases, the Prandtl number Pr = 0.71 (air), the Froude number and the aspect
ratio are unity. For the grid refinement study, three meshes (16 x 16 x 8,
32 x 32 x 16 and 64 x 64 x 32) were used. Fluid flows are assumed symmetric
with respect to the central x - y plane, and half of the grid nodes are used
in the z direction. In all three meshes, grid lines were clustered near the
walls to resolve high gradients of the flow properties. In all calculations,
trilinear elements with one-point Gaussian quadrature are used. In addition,
the calculation is carried out by using the solution of coarse mesh as initial
guess for the fine mesh solution.
insulated wall
y u=v=w=qy=O
hot cold
6=6h 6=6c
u=v=w=qy=qz=O u=v=w=qy=qz=O
I I
I I
I I ____
r--
~x
I
I /
1//
r----- go
/ --~
/
z
/ symmetric plane
w=c.ox=c.oy=qz=O
In the computation there are two loops of iterations: the outer loop up-
dates the coefficient matrices and source terms by Newton's method; while
the inner loop solves the variable increment 15!! by using the JCG method.
Typically, it takes about 10 to 15 global iterations to reduce 15!! to about
eight to ten orders of magnitude. The limitation of the global convergence is
due to the prescribed criterion for the inner iterations. Usually, two to three
hundred inner iterations are used for each global iteration.
Figure 11.8 shows the distribution of the u velocity at the central vertical
line for Ra = 105 using three meshes, respectively. Little difference is observed
between solutions for the 32 x 32 x 16 mesh and the 64 x 64 x 32 mesh.
In another test, the error of overall heat transfer obtained by integrating
qx over the hot and cold conducting surfaces is evaluated. Due to the conser-
vation of energy, the integrated heat fluxes on the cold and hot walls should
280 11. Low-Speed Compressible Viscous Flows
o16 X 16 X 8
0.8 32 X 32 X 16
- - - - 64 X 64 X 32
0.6
>-
0.4
0.2
0.0
-0.20 -0.10 0.00 0.10 0.20
U velocity
Fig. 11.8. Velocity u distribution along the vertical centerline at Ra = 105 (from
Yu et al. 1996b)
be equal. The error tabulated in Table 11.3 is the difference between the two
integrated heat fluxes normalized by their average.
Table 11.3. Errors of overall heat transfer (%) (from Yu et al. 1996b)
Ra Mesh 17x17x9 32x32x17 64x 64 x32
103 - 2.040 1.029 0.206
104 3.164 1.143 0.732
105 2.410 1.510 1.175
5x105 3.326 1.203 0.910
(a) Ra = 103
The computation with Ra higher than a half million was also conducted.
No steady-state solution could be obtained; this means that the flow may
become unstable for Ra higher than a half million.
By introducing the vorticity and the heat flux as additional variables, the
governing equations for low-speed compressible viscous flows can be written
as an almost-linear first-order system. The principle part of this system is
exactly the same as that for incompressible flow-heat coupling problems. It
consists of div-curl-grad operators. Consequently, the steady-state governing
equation system for low-speed flows is elliptic, and the required boundary
conditions become verifiable.
In contrast to conventional approaches, the LSFEM based on the quasi-
linear or almost linear first-order system evades the difficulty of the lack
of a timEHierivative of pressure in the continuity equation, and no special
treatments, such as preconditioning, staggered grids, or operator-splitting
. are needed. Moreover, due to the merit of LSFEM, the coefficient matrix
is symmetric positive definite; its solution can be conducted by an element-
by-element Jacobi preconditioned conjugate gradient method. The precondi-
tioning matrix in LSFEM is automatically chosen by the preconditioned CG
method itself after discretization, instead of the free choice of only two pa-
rameters to precondition the time derivative terms in the partial differential
equations. Numerical solutions with good resolution for two and three di-
mensional compressible buoyant flows at various Rayleigh numbers have been
obtained by LSFEM. No previously reported results showed such resolution.
This illustrates the merits of LSFEM: (1) no artificial boundary condition for
pressure is used; (2) no added artificial damping is employed; and (3) first-
order derivative variables such as vorticity and heat fluxes are discretized to
the same order of accuracy as the velocity, pressure, and temperature.
12. Two-Fluid Flows
12.1 Introduction
Many engineering problems involve multi-fluid/interfacial flows. Injection
molding, metal casting, crystal growth and spray atomization are some sig-
nificant examples. At the interface of different fluids, surface tension exists as
a result of uneven molecular forces. The interface behaves in a way similar to
a thinly stretched membrane. The prediction of the evolution of the interface
and the treatment of the interface conditions have been a challenging task
for numerical simulations.
Most existing numerical methods for multi-fluid/free-surface flows fall into
two categories: (1) those which use a fixed grid; and (2) those which allow
the grid to deform in time so that it remains surface-intrinsic.
In the first category, the computational grid is fixed throughout the cal-
culation. An additional variable is used to identify the interface (front). Ex-
286 12. Two-Fluid Flows
amples of such methods are the marker and cell (MAC) method proposed
by Harlow and Welch (1965), the volume of fluid (VOF) method by Hirt
and Nichols (1981), and the level set method by Zhu and Sethian (1992).
The MAC method uses massless marker particles which travel with the fluid
to trace the fluids and the interface. The VOF method modifies the MAC
method by replacing the discrete marker particles with a continuous field
variable - the color function or the level set function. This function assigns
a unique constant (color) to each fluid. At fluid interfaces this color function
has a sharp gradient. Numerical methods in this category are sometimes re-
ferred to as "front capturing" methods. Such methods possess great flexibility
in handling large deformations and topological changes, as demonstrated by
Daly (1967) and Harlow and Shannon (1967). The most difficult task with
the front capturing approach is to accurately identify the interface and to
impose the interface condition; these are exemplified by the elaborate work
of Daly (1969). This difficulty can be alleviated by using a continuum surface
force (CSF) model proposed by Brackbill et al. (1992). The CSF model in-
terprets surface tension as a continuous, three-dimensional body force across
an interface, rather than as a boundary condition on the interface. The com-
puter implementation of the CSF model is therefore relatively simple com-
pared with other approaches. The combination of the VOF method (or the
level set method) and the CSF model has been used by a number of authors
to simulate multi-phase phenomena involving surface tension and complex
topological changes, see e.g., Richards et al. (1994), Lafaurie et al. (1994),
Sussman et at. (1994), and Chang et al. (1996). A detailed description of the
CSF model will be given in the next section.
For methods in the second category, see e.g., Pritt and Boris (1979), Fyfe
et al. (1988), Jue and Ramaswamy (1992), Tezduya et al. (1992a,b), which are
referred to as "front tracking" methods, imposing the interface condition is
easy compared to the first, because the interface always coincides with mesh
sides. However, it requires frequent updating of the computational mesh,
which can be a complex and time-consuming procedure. In particular, it
encounters severe difficulty when the flow experiences severe distortions and
complex topological changes.
Another approach which can be regarded as a combination of the above
is the front-tracking method introduced by Unverdi and Tryggvason (1992).
This approach uses a fixed, structured grid to represent the flow field. A
separate, unstructured grid is used to represent the interface. The interface
is explicitly tracked and kept at a constant thickness of the order of the
mesh size. This ensures that the interface will not be smeared by numerical
diffusion. Much success has been achieved in solving a variety of two-fluid
flow problems using this approach (Nobari and Tryggvason 1994a,b). The
difficulty with this approach is the handling of complex topological changes.
In this chapter, numerical solutions to problems involving two immiscible
fluids are sought. A large number of such problems deal with the interaction
12.2 Continuum Surface Force Model 287
between a liquid and gas or air, which are often simply treated as free-surface
problems. In the free-surface formulation the flow equations are solved only
for the liquid, and zero traction is assumed on the interface. In this chapter
such problems are treated as true two-fluid cases. The VOF approach and
a modified version of the CSF model proposed by Jacqmin (1995) are used
in the simulation. The LSFEM is used to discretize the governing system of
equations.
which acts as a source term in the momentum equations. In the above for-
mulation, [CJ denotes the jump of C across the interface; the curvature K. is
calculated from
288 12. Two-Fluid Flows
f ,.-
- aTij
aXj' (12.7)
in which the stress tensor Tij is related to the color function C by:
Tij _~ _1_ (ac ac _ fJ .. ac ac)
[C] IVCj aXi aXj '3 aXk aXk
a( 1 acac
- [C] IVCI aXi aXj - fJij Ivcl ,
) (12.8)
where fJij is the Kronecker delta. This formulation for the body force was first
derived by Jacqmin (1995). The advantage of using the formulation given by
(12.7) and (12.8) instead of the original CSF model (12.3) is that (12.7)
and (12.8) do not require the explicit calculation of the normalized gradient
term in the right hand side of (12.5), whose definition is not clear when the
denominator IVCj approaches zero. In (12.7) and (12.8), both T and fare
well defined in the whole domain, and naturally vanish when IVCj becomes
zero. An additional advantage of using the stress tensor T is that it can be
regarded as part of the momentum flux. In many numerical procedures T can
be used directly, and there is no need to calculate f.
12.3 The First-Order Governing Equations 289
au
P at + P (u . V') u + V'p + J.t (V' x w)
- (V'J.t. V'u + V'J.t. (V'U)T) = f in a, (12.9b)
w-V'xu=O in a, (12.9c)
V'·w=O ina, (12.9d)
in which P is the density, u is the velocity vector, p is the pressure, w is
the vorticity vector, J.t is the dynamic viscosity, and f is the body force,
which generally consists of the surface tension effect given by (12.7) and the
gravitational force. The terms in the last pair of brackets on the left hand
side represent the effect of non-uniform viscosity.
The fluids are identified by different values of the color function C, which
is convected by the flow field:
ac
at + (u . V')C = O. (12.10)
Fluid properties such as the density and the viscosity are assumed to be
distributed in the same manner as C, i.e.,
P = PI + CP22 -- PI ( )
C C - C1 ,
1
(12.11a)
The governing equations (12.9) and (12.10) are first discretized in time.
The backward-Euler difference scheme is used:
in a, (12.12a)
in a, (12.12b)
in a, (12.12c)
290 12. Two-Fluid Flows
in fl, (12.12d)
in fl, (12.12e)
where the superscript 'n' denotes the previous time-step and 'n + l' denotes
the current time-step.
The above system of equations (12.12) is further linearized. To ensure
time-accuracy for time-dependent problems, linearization iterations are per-
formed within each time-step (typically three iterations are used). For steady
state problems only one linearization is performed in each time-step. Let k
denote the number of linerization, the linearization is carried out as follows:
\7 . u~r = 0, in fl, (12.13a)
u n +1
n+! [kJ
P[k-IJ Llt
+ P[k-lJ
n+1 (n+1
u[k-IJ'
n)
Y u[kJ
n+1 + nY P[kJ
n+1 + n+! \7
Il[k-IJ
n+1
x w[kJ
n+1 n+1 un
+I[k-IJ + P[k-IJ ..1t in fl, (12.13b)
n
y,w[kJ
n+l
= 0 . n
lllu, (12.13d)
e n+1 en
~
..1t
+ (u n[kJ+1 . \7) en+! _
[kJ -
[k-IJ
..1t in fl, (12.13e)
8ur
8r
+ !u + ! 8ue + 8u
r r r 8(} 8z
z = 0 in n, (12.14a)
8u r u~ 8u 1 8u 8u 8p
p - - p-
at r
+ pur -8rr + PUe-- r
r 8(}
+ pu z -8zr +-
8r
+J.L (!r 8w
8(}
z _ 8we) _ 2 8 J1. 8u r _ 8J1.
8z 8r 8r r 8(}
! (!r 8u80r + 8ue
8r
_ ue)
r
+J1. (8w r _ 8W z )
8z 8r
_ 8J1.
8r
(!r 8u8(}r + 8ue
8r
_ u e ) _ ~ 8J1. (! 8 u e + u r )
r r 8(} r 8(} r
0' ( 1 1 OC oC )
TIJIJ = - [CJ IV'CI r2 of) of) - IV'CI , (12.17b)
0' ( 1 OC oC ) (12.17c)
Tzz = - [CJ 1V'C1 oz oz - 1V'C1 ,
0' ( 1 1 oC OC) (12.17d)
Tr() = T()r = - [CJ 1V'C1 ~ or of) ,
1 1 oC OC)
IV'CI ~ of) a; ,
0' (
T()z = Tz() = - [C] (12.17e)
(12.18a)
(12.18b)
12.4 Numerical Examples 293
When the flow is axisymmetric, all the O-derivatives become zero. More-
over, if the flow is non-swirling, u() becomes zero. For such cases the governing
equations are further simplified.
Broken Dam. The broken dam problem has been used by many as a test
case for simulating free-surface problems. Experimental data for this case are
available (Martin and Moyce 1952). Here the problem is solved as a two-fluid
problem involving both water and air. Zero surface tension and slippery walls
(left and bottom sides) are assumed. On the top and right sides zero pressure
is imposed. In addition, if inflow is detected on the top and right sides, the
density of air is imposed as a boundary condition. The computational domain
is 2 units high and 6 units long. Initially, water occupies a 1 x 1 area at the
bottom left corner. The computational mesh consists of 120 x 40 uniform
linear quadrilateral elements. The time-step is Llt = 0.05. The nondimen-
sionalized gravitational acceleration, g, is taken to be unity. The viscosity is
set at 3.05 x 10- 5 for water which is the same choice as in Nakayama and
Mori (1996) and 3.05 x 10- 8 for air. The densities for water and air are 1 and
0.001, respectively.
Figure 12.1 shows free surface profiles (a contour line of density at p =
0.5) and the pressure contours at various times. The calculated water front
location and water column height are compared with the experimental data
in Fig. 12.2. It can be seen that the calculated results are in good agreement
with the experimental data given by Martin and Moyce (1952) and with the
calculation results in Nakayama and Mori (1996).
Two-Liquid Interface. This example is taken from Tezduyar et al. (1992b).
In this problem, a two-dimensional 0.8 x 0.6 box is fully occupied by two
liquids of equal volume. The density of the fluids are 1.0 and 2.0, respectively.
The gravitational acceleration is 0.294. The lighter liquid is placed on top of
the heavier one. Initially the interface is a straight line, with a slope of 0.25.
The dynamical viscosity is 0.001 (the same for both liquids). Zero surface
tension is assumed. The computational mesh consists of 21 x 42 uniform
bilinear elements. A time-step of Llt = 0.25 is used throughout the calculation
which corresponds to a CFL number close to 1 when the velocity reaches its
maximum at approximately t = 6. Slippery wall conditions are used on all
four sides. Figure 12.3 shows the interface profile and velocity field at different
times. Figure 12.4 shows the time history of the interface location on the left
and right hand side walls (relative to and normalized by the average height).
The results agree well with those presented in Tezduyar et al. (1992b).
294 12. Two-Fluid Flows
b I
K I
b: I ~
C I
I -.... I
[ I
k :::. I
[ I
Fig. 12.1. A broken dam, free surface profile and pressure contours, from top to
bottom: t = 0.5, 1.0, 1.5,2.0,2.5,3.0 (from Wu et al. 1996)
12.4 Numerical Examples 295
•.0 0 .8
2.0 0 .•
02~--~--~~--~
2.0 3.0 0 0.0 1 .0 2 .0 3.0
Fig. 12.2. A broken dam, front location and water column height, solid line -
calculated results; dots - experimental data (from Wu et al. 1996)
Fig. 12.3. A two-liquid interface problem, interface profile and velocity field, t =
0, I, 2, 3, 4, 5, 6, 7, 8, 9,10,11,12,13,14, IS, 16, 17, 18, 19, respectively (from Wu et al.
1996)
296 12. Two-Fluid Flows
0.4
-0.4
0.0 5.0 10.0 15.0 20.0 25.0
Fig. 12.4. A two-liquid interface problem, time history of the interface location on
the left and right hand side walls (from Wu et al. 1996)
0 0 0 0
0 0 0 0
0 0 0 0
0 0 0 0
Fig. 12.5. An oscillating bubble, interface profile for the two-dimensional case,
t = 0,1,2,3,4,5,6,7,8,9,10,11,12,13,14,40, respectively (from Wu et al. 1996)
00 00
000 0
0000
0000
Fig. 12.6. An oscillating bubble, interface profile for the axisymmetric case, t =
0,1,2,3,4,5,6,7,8,9,10,11,12,13,14,40, respectively (from Wu et al. 1996)
298 12. Two-Fluid Flows
2.0 p;;----------.-------_--------,
1.5
)(
1.0
0.5~------~-------~------~
0.0 5.0 10.0 15.0
2.0,----------.-------_--------,
1.8
)( 1.6
1.4
Fig. 12.7. An oscillating bubble, time history of the interface location on the x-
axis. upper: the two-dimensional case; lower: the axisymmetric case (from Wu et
al. 1996)
12.4 Numerical Examples 299
Fig. 12.8. Coalescence of two initially stationary drops, t = 0.0, 1.0, 1.5,2.0,2.5,3.0,
respectively (from Wu et al. 1996)
300 12. Two-Fluid Flows
1{---=--0
'"
'"
....
-~
2.0
4.0
Fig. 12.9. Flow in a pressure swirl atomizer, the problem definition (from Wu et
al. 1996)
Fig. 12.10. Flow in a pressure swirl atomizer, the computational mesh and the
interface profile (from Wu et al. 1996)
302 12. Two-Fluid Flows
In this chapter a numerical procedure based on a CSF model and the LSFEM
is presented for two-fluid flow problems. Numerical tests carried out on a
number of two-dimensional planar and axisymmetric flows indicate that this
approach is capable of simulating such flow phenomena. The present approach
has the advantage in the ability to handle complex topological changes such as
the breakup of liquid jet. Another advantage is the ability to handle complex
geometrical configurations in practical engineering environments, since the
LSFEM is based on totally unstructured grids. In the presented examples
the interface is typically spread over 3 '" 5 grids. A change of the interface
thickness is also observed in the calculations. The analysis by Haj-Hariri and
Borhan (1994) indicates that the migration velocity is strongly affected by the
smearing of interface. Thus, to ensure good accuracy of the simulation a fine
grid would have to be used. Improvement of the accuracy of the numerical
scheme is expected by introducing higher order discretization schemes in
both time and space. The tests conducted by Yu et al. (1995b) indicate
that discontinuity can be modeled within 3 grids, even when the interface
undergoes very large deformation due to a highly vortical flow field.
13. High-Speed Compressible Flows
°
boundary data. Depending on the problem, the boundary data may be given
at x = 1, or specified at both x = and x = 1.
304 13. High-Speed Compressible Flows
(13.2)
(13.3)
where k stands for the iteration count, a~r = a( u~r) and u~t = un. In
the following discussion, for neatness and convenience we just use (13.2) to
write the related formulations.
We can now apply the general formulation of LSFEM given in Sect. 4.8
to (13.2) to find the solution un+! at each time-level.
We proceed to discern why this method can capture shocks. For this
reason we deduce the variational statement. The basic least-squares method
for (13.2) amounts to minimizing the L2 norm of the residual R for admissible
un+! in (13.2), i.e., minimizing the objective functional:
r1 1 r
J(u n+1) = Jo R 2 dx = Jo (u n+1 - un
8u n+1
+ ..1tan ----a;-)
2
dx. (13.4)
1.0
\
0.8
0.6
0.4
0.2
::;) 0.0
-0.2
-0.4
-0.6
-0.8
-1.0
1\
0.0 0.2 0.4 0.6 0.8 1.0
X
1= 10 [R2 + p(~:r]dx
1
(13.8)
Vv E Ho, (13.9)
( Oa n ) 2 OV OUn+!
{3 1 + Llt ox ox ----a;-'
This term corresponds to adding the stabilization functional
{3la
l
(1 + Llt~~)2 (O~:+! fdx
to the functional in (13.4). Integration by parts reveals that the corresponding
numerical viscosity term in the associated Euler-Lagrange equation is
\~
0.8 0.8
0.6 0.6
:;)
0.4 0.4
0.2 0.2
0.0 0.0
-0.2 '-_'--_'----......."""""---'~_l -0.2
0.0 0 .2 0 .4 0 .6 0 .8 1.0 0.0 0 ,2 0.4 0 .6 0.8 1 .0
X X
(C) Non-conservative L2-scheme (d) Non-conservative H1-scheme
two-point Gaussian quadrature
two-polnt Gaussian quadrature
\
1 . 0+----'!"'""~ 1,0
0 ,8 0 ,8
0.6
:;)
0.6
0,4
0.4
0.2
0.2
0.0 ....._ _ _.,
0.0
-0.2 L..................J'----.........J~.........J~---'_ ___!
0.0 0 .2 0.4 0.6 0.8 1.0 -0.2
0.0 0.2 0 ,4 0.6 0 ,8 1.0
X X
Fig. 13.2. Solution to the Burger equation from initial slant step with 100 uni-
form linear elements: (a) ..1t = 0.02,t = 0.2,k = 0; (b) ..1t = O.OI,t = O.4,fJ =
0.00004,k = 1; (c) Llt = O.OI,t = 0.2 , k = 1; (d) Llt = O.OOl,t = 0.4,,8 =
0.000002, k = 1
F n +1
(13.13)
Applying the linearization procedure of (13.13) to the flux derivative aF/ax,
(13.12) becomes
aa n aLl aFn 1
(1 + (}Llt ax )Llq + (}Llta n axq = -Llt ax + (() - 2)O(Llt2). (13.14)
When () = 1, this scheme corresponds to the backward-Euler differencing
scheme and is of first-order accuracy O(Llt) in the sense of the Taylor-series
expansion; for () = 1/2 one obtains the Crank-Nicolson scheme with a second-
order accuracy O(Llt)2.
The corresponding least-squares variational statement is: Find Llu E Ho
such that
Applying the idea introduced in Sect. 13.1.2, we may also develop a conserva-
tive approximate H1-residual scheme which can be obtained just by adding
the following term into the right-hand side of (13.15 ):
r
1 ( aan ) 2 av aLlq
Jo {3 1 + 2(}Llt ax ax ax dx.
in which
Ao = ( 1 + (}Llt ax
oan )
o '
f = ( -Llt!!..I.::..)
o ax .
The general formulation of LSFEM given in Sect. 4.8 can be directly applied
to solving (13.16).
To obtain time-accurate results, at each time-level we may perform two
or three iterations as discussed in Sect. 13.1.1.
310 13. High-Speed Compressible Flows
Figure 13.3 shows the computed solutions to the same Burger equation
tested in the previous section with the conservative HI scheme. As expected,
the first-order scheme (0 = 1) produces a diffused shock front, while the
second-order scheme (0 = 0.5) can provide sharp resolution of shocks but
with some oscillations.
1.0
0.8
, 1.0
0.8
0.6 0.6
:J :J
0.4 0.4
0.2 0.2
0.0 0.0
-0.2 -0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
X X
Fig. 13.3. Application of the conservative HI scheme to the Burger equation with
100 uniform linear elements, Llt = 0.016, t = 0.4, k = 2, one-point Gaussian quadra-
ture: (a) first-order (B = 1), (3 = 0.00003; (b) second-order (B = 0.5), (3 = 0.00003
+ LltAn1 aX
au n + 1 = un + Llt8.
un+! (13.18)
u_(pa)
- pau '
Al = ( 2
c -u
a 2 8 = (pc~~~)' (13.19)
where p is the density, u is the velocity, the speed of sound c is normalized
to unit and the cross-sectional area of the nozzle is
a= 1a (x - 2.5)2 a ::; x ::; 5. (13.20)
. + 12.5 '
Numerical results for subsonic inflow and outflow with interior supersonic
flow regime terminating in a shock, using the non-conservative HI scheme,
are shown in Fig. 13.4. The computations are performed on 80 uniform linear
elements with a time-step Llt = 0.5 and f3 = 0.0001. The steady-state solution
is attained after 80 time steps.
Shock Tube Problem. The one-dimensional Euler equations for compress-
ible flow can be expressed as a system in the form (13.17) with
Al = ( a uPO)
u
p-I , 8=0, (13.21 )
a 'Y u
where p is the density, u is the velocity, p is the pressure and 'Y = 1.4 is the
ratio of specific heats. For the shock tube problem the initial data are
U=
1.0)
( 0.0 for x ::; 0.5,
0.125)
U= ( 0.0 for x > 0.5. (13.22)
1.0 0.1
Here the initial density and pressure difference between two sections of the
tube is maintained by a diaphragm which is destroyed at t = O. The objective
is to compute accurate approximations to p, u and p at subsequent times as
the shock and contact discontinuities propagate along the tube. The problem
is solved here using the non-conservative L2 scheme (13.5) with a uniform
mesh of 100 linear elements, a fixed time step Llt = 0.005 and one-point
Gaussian quadrature. The solution profiles for density, velocity and pressure
at t = 0.14 are shown in Fig. 13.5. There is moderate smearing of the 'fronts'.
The solution is neither oscillatory nor unstable.
312 13. High-Speed Compressible Flows
...
0
0
on
-
>-
~o
HO
O.
0-
-J
W
>
0
on
0
0
0
...
0
-
on
>-
~O
HO
Ill':
Z
w
Q
0
on
0
0
0
Fig. 13.4. Computed solution for isothermal flow in a nozzle using the non-
conservative HI scheme and linear elements: h = 0.125, Llt = 0.5, t = 40, f3 = 0.0001
(from Jiang and Carey 1988)
13.2 One-Dimensional Flows 313
1.0
0.8
>-
I-
enz 0.6
w
Cl 0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
X
1.0
0.8
~ 0.6
C3
g
w 0.4
>
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
X
1.0
0.8
w
c: 0.6
::J
CI)
CI)
w
c: 0.4
a..
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
X
Fig. 13.5. Solutions to the shock tube problem with 100 uniform linear elements,
Llt = 0.005, t = 0.14, k = 1 using a non-conservative L2 scheme
314 13. High-Speed Compressible Flows
(13.23)
( o v 0:PO)
0
A1=
U 0 ' 0o p-l , V
o 'YP 0 u o
0 'YP v
in which p is the density, (u, v) are the Cartesian velocity components, p is
the pressure, and 'Y is the specific heat ratio. Boundary conditions and an
initial condition are needed to complete the description of the problem.
Introducing .tJ.U = Un+! - un as an unknown vector we linearize the
system (13.23) by neglecting higher-order terms. After time-discretization by
using the backward-Euler difference, system (13.23) becomes
A.tJ.U=! (13.24)
8Un An8Un)
_ - ( An1--+
! -
8x
28y
--'
!!..e.
8y
o
( 1
Llt
+ 8x
8u + 8v)
8y
'Y~
in which I is the identity matrix. Here we put .tJ.t in Ao. In this way, the
arithmetic operation count will be reduced. For steady-state problems, the
derivative terms in Ao can also be ignored.
Note that even for capturing stationary shocks, least-squares methods
must be based on time-marching algorithms such as formulation (13.24).
Because of a lack of the entropy contraint condition, the original compressible
Euler equations (13.23) are not well posed in the sense of admitting both
compression and expansion shocks. The direct application of least-squares
methods to steady-state Euler equations is suitable only for subsonic flows; for
13.3 Two-Dimensional Flows 315
supersonic flows an artificial viscosity term must be added (see examples given
by Laminie 1994). As discussed in Sect. 13.1, the least-squares procedure
for time-discretized equations inherently generates numerical dissipation to
exclude non-physical expansion shocks.
Grid Adaptation. For the compressible Euler equations, the least-squares
method is naturally diffusive, no explicit artificial viscosity is added to the
formulation. In general, the inherent numerical dissipation produced by the
least-squares method does not allow shocks to be sharply resolved, unless
fine grids are used. However, shock resolution can be significantly improved
by adapting the grid to the solution. Taghaddosi et al. (1997) have success-
fully obtained high quality shock solutions by combining the LSFEM with a
directional-adaptive remeshing method. In the following we briefly describe
the idea of their approach.
Their edge-based moving-node method (r-method) is originally developed
by Ait-Ali-Yahia et al.(1996). It uses second-order derivatives of a given flow
variable as an error indicator. In this approach, the number of both ele-
ments and nodes is fixed, connectivity is conserved, and the error is equally
distributed over element edges. This anisotropic refinement where more reso-
lution is introduced along those direction with rapidly changing flow variables
is particularly useful for directional flow phenomena, such as shock waves.
Let us first consider a one-dimensional problem, in which the solution
variable U is approximated by the linear interpolation Uh. If quadratic in-
terpolation Uh is available and the nodal error is zero, then the local error
of linear interpolation over an element can be estimated as the difference
between the quadratic and linear one:
, ~ d2uh
Ee = Uh - Uh = '2 (he -~) dx 2 ' (13.25)
e
where is the local element coordinate and he the element length. The overall
error in the element may be measured by the root-mean-square value of Ee:
e he yf120 e dx 2 e
An optimal mesh may be defined as one for which the error is equally dis-
tributed over elements, thus the following criteria should be satisfied by each
element:
h 21 d2uh
e dx2
Ie = C
'
(13.27)
whose value is calculated for all nodes throughout the domain. The value
of u at the middle point of each element edge can be obtained by linear
interpolation. Then the second derivative along any edge through its middle
point can be calculated by a transformation of coordinates.
Since the solution is obtained by using linear elements, the quadratic solu-
tion Uh inside each element is not available. Therefore, the second derivatives
have no representation inside an element. However, one may investigate the
solution Uh in the elements surrounding node i. This yields an estimate of
the second derivatives. For example,
f Uh A. ·do f Uh A.. dO
uh,xyli ~ Jn~ ;~;~ = - Jnil, ';~;~ , (13.29)
n, ' ni '
where oi represents the elements sharing node i, and cPi is a piecewise con-
tinuous linear function with a value unity at node i and null at other nodes.
To equally distribute the error over the entire domain, a moving-node
method based on a spring analogy is used (Gnoffo 1983). Here the mesh
is interpreted as a network of springs, where each edge is considered as a
fictitious spring with its stiffness representing the measure of the error, see
Fig. 13.6. The equilibrium of forces in this spring system will then determine
the movement of each node.
j ~ passive node
j
Fig. 13.6. Local spring network corresponding to node i
where ~ is the position vector, and the stiffness kij is defined as:
(13.31)
in which u~ is the second derivative along the edge (~i - ~j) at its mid-
dle point. Minimizing P in (13.30) with respect to ~i yields the following
equation:
13.3 Two-Dimensional Flows 317
1.0 )
u= ( 2.9 for upstream,
0.0
0.7143
1.7 )
U = ( 2.6193 for upper boundary.
-0.5063
1.5282
A shock emanates from the upper left corner; this shock is reflected at the
lower wall where v = 0, and the downstream boundary conditions remain
free for outflow. The initial data were prescribed as constant at values given
on the upper boundary and the specific heat ratio is 'Y = 1.4.
In the initial calculation, a uniform 60 x 20 mesh of bilinear elements was
used. In the first two adaptive cycles, the solution was obtained with a time
step Llt = 0.1. The numerical viscosity is reduced beginning at the third
cycle by reducing the time step to 0.05. The resulting algebraic equations are
solved by using the Jacobi preconditioned conjugate gradient method. Figure
13.8 shows the pressure contours and the grids after each adaptation. The
improvements in the shock resolution after adaptation is quite evident in Fig.
13.9.
318 13. High-Speed Compressible Flows
Moo = 2.9
110
Fig. 13.7. Shock reflection from a solid wall: the computation domain (from
Taghaddosi et al. 1997)
I~
3rd adaptation (llt = 0.05)
1:S:Z
5th adaptation (llt = 0.05)
Fig. 13.8. Evolution of the grid and the pressure contours during adaptation for
shock reflection problem (from Taghaddosi et al. 1997)
13.3 Two-Dimensional Flows 319
3 . 5.-------~--~----~--~--~--~,_--~
1 ............... ;._.. . . !
··· __·!-············--···1'·
/
O.5+----r---T--~----+----r---T--~r_--4J
o 0.5 1.5 2 2.5 3 3.5 4
x
Fig. 13.9. Pressure distribution at y = 0.5 before and after adaptation for shock
reflection problem (from Taghaddosi et al. 1997)
l~iE I
initial solution (~t = 0.3)
I~I
1st adaptation (Dot = 0.3)
I dig? I
2nd adaptation (~t = 0.1)
Fig. 13.10. Grids and Mach number contours before and after adaptation for
transonic channel flow (from Taghaddosi et al. 1997)
abapled LSi
1.5 i" -'--,-' Ni el. lJ.
Eidelman . •
1.25
ci
Z
.c
:;l
::2 0.75
0.5 -.
x
Fig. 13.11. Mach number distribution on the lower and upper walls for transonic
channel flow (from Taghaddosi et al. 1997)
13.3 Two-Dimensional Flows 321
where
pu2Pu+p ) G- pv
puv )
( - ( pv2 +p , (13.35b)
F = (p::-~)u ' (pe + p)v
in which e is the total energy, and for the case of perfect gas the equation of
state is
(13.35c)
Retaining the conservative variables as an intermediate step, we first con-
vert (13.35) (temporarily) into the following nonconservative form:
oq - oq - oq
-ot + A 1 -ox + A 2 -oy = 0, (13.36a)
with
1 0
_ 1;Y(u2 +0v 2) _ u 2 (3 - ,)u -;Yv
Al= ( 2
-uv V U
n'
0 1
- ( -uv
0 v u
A2 = (13.36c)
!;y(u2 + v2) _ v2 -;yu (3 - ,)v
(;y(u 2 + v 2) - ,e)v -;yuv € _ ;yv 2 ,v
in which
322 13. High-Speed Compressible Flows
1.1,-----:-----r----...,.....--------- .
~o adapt.
loa ..
4th
5th
0.9 .. ~.- ...
0.4 +---+---+---+---+---+--- -1
o 0.5 1.5 2 2.5 3
x
Fig. 13.13. Pressure distribution at y = 0.2 before and after adaptation for super-
sonic channel flow (from Taghaddosi et al. 1997)
I aA n aA n
A~ = .t1t + ax'" + ayY, (13.37b)
(13.39)
is satisfied, the increment Llq becomes zero, and the calculation is terminated.
Our numerical experiments on the shock wave reflection problem show that
there is no essential difference between the results of nonconservative and
conservative LSFEM.
Lefebvre et al. (1993) applied scheme (13.38) to problems of transonic and
supersonic flows on triangular meshes. In their approach, a local time step is
calculated in each element according to
h
Llt = CFL-,- , (13.40)
I'Imax
where h is the element size, Amax is the maximum eigenvalue of the present
hyperbolic system (see, e.g., Hirsch 1990), and CFL (Courant-Friedrichs-
Levy) number is a user specified constant. The time step is chosen so as
to obtain a uniform CFL number of 5 throughout the mesh. The resulting
algebraic equations is solved by the incomplete Choleski conjugate gradient
method.
Transonic Flow Past an Airfoil. The first example is transonic flow past
a NACA0012 airfoil with free stream Mach number of 0.95. A structured
O-type mesh of quadratic (P2) elements employing 160 nodes around the
surface of the airfoil and 49 nodes in the normal direction is used. A detail of
the mesh in the vicinity of the airfoil is shown in Fig 13.14. The pressure and
Mach number distribution are shown in Fig. 13.15. The oblique shocks are
reasonably well captured and compare favorably with highly accurate results
of this test problem obtained by Pullian and Barton (1985).
Fig. 13.14. Mesh of quadratic elements used for the solution of the transonic flow
past a NACA0012 airfoil, Nnode=7840, Nelem=3840 (from Lefebvre et al. 1993)
13.3 Two-Dimensional Flows 325
0 .& ,....-.,--r--r-,.-.--,--r-.,--,---,r-T"""
....
0.'
0 .2 .' ~ '-"' .
:
0 .0
a. - 0 .2 1-'-+--+--+-+-+-+- I-'--+--+----1I-'--t--.-
u f-;"-=c~
I -0.' r-t--+-t-t-+-t-r-',....-t--ll--t--t
- 0 .& F--+-+----1-+-+-+-~+-+-if-+-~
I .'
.... ....
.... .... ~-
1.2
.-,' I'- I- •
.:
1.0
0 .8
~~
~
0 .'
0 .'
0 .2
Fig. 13.15. Transonic flow past a NACA0012 airfoil, Moo = 0.95, (a) Pressure
coefficient distribution and contours, (b) Mach number distribution and contours
(from Lefebvre et al. 1993)
326 13. High-Speed Compressible Flows
(b) solution
..
Z.O
L r-=-tPl
.
It
/
.
Q. 1.0
\
,,
U 01
01 . \
\
,
Q.
\
00
'-..J ~ ... ..
-o! I.' _ ' .0 -0.' 0 ,0 0 _' 1.0 '-,
x
.,
' ,0 1
I
r
l"
) ,0 ~
, I
"
Z"
0
Z,O
J
I"
.
I.:'
. 1//
0,'
....
'\ .... .. 0
,,
~.o • 0
-1.0 0 ,0
•
Fig. 13.16. Supersonic flow past a cylinder, Moo = 3, (a) Adapted mesh, (b)
Pressure and Mach number solution (from Lefebvre et al. 1993)
13.4 Concluding Remarks 327
(13.42)
in which ile is the element area. The refinement strategy is then to refine all
these elements for which the indicator is larger than a certain specified value.
Each triangle, for which refinement is needed, is subdivided into four smaller
triangles, and to avoid the problem of "hanging nodes", transition elements
are introduced. After the refinement, the sides of the elements are replaced
by springs of unit stiffness and a few relaxation iterations are used to move
the nodes until nodal equilibrium of forces is achieved, as described in Sect.
13.3.!.
This refinement technique is applied to supersonic flow past a cylinder.
The free stream Mach number is 3, and due to the assumed symmetry,
only one half of the cylinder is considered. The initial mesh consists of 1028
quadratic (P2) elements and 2145 nodes. The adapted mesh and the steady-
state solutions are presented in Fig. 13.16.
We have demonstrated that the general LSFEM given in Sect. 4.8, when ap-
plied to semi-discretized compressible Euler equations obtained by various
implicit time-difference schemes, can capture shocks in high-speed compress-
ible flows. In contrast to other methods, the LSFEM does not use upwinding,
Riemann solvers, flux splitting or staggered grids as building blocks. There-
fore, the LSFEM is very simple. The implicit nature of LSFEM does not
cause problems with computational efficiency, because the conjugate gra-
dient method can be employed to solve the resulting positive and definite
matrix. The resolution of the computed results can be significantly improved
by using adaptive mesh refinement. The equation residual or the recovered
second-order derivative of a given flow variable can serve as a reliable error
indicator for refinement.
The time-marching LSFEM for high-speed compressible flows was orig-
inally proposed in Jiang and Carey (1988, 1990) and Jiang and Povinelli
(1990). The research group led by K. Morgan (Lefebvre et al. 1992, 1993) and
the group led by W.G. Habashi (Taghaddosi et al. 1997) have demonstrated
that by coupling the least-squares method with adaptive mesh refinement,
shocks can be captured with high resolution using relatively coarse grids.
328 13. High-Speed Compressible Flows
LSFEM in Electromagnetics
14. Electromagnetics
(14.2d)
for magnetic fields, where n is the unit vector normal to the interface, pointing
from medium (-) to medium (+), p~mp is the imposed surface charge density
and J~mp is the imposed surface electric current density.
334 14. Electromagnetics
The boundary conditions can be reduced to a special case when one of the
media, say medium (-), becomes a perfect conductor. Since a perfect conduc-
tor cannot sustain a field within it, the corresponding boundary conditions
for a perfectly conducting wall becomen
n x E=O onn, (14.3a)
n . (J.LH) = 0 on n, (14.3b)
where E and H are the fields exterior to the conductor. Similarly, for a
r
perfect magnetic wall 2 we have
(14.3c)
n· (eE) = 0 on r 2. (14.3d)
l n.KimPdr=O, (14.4c)
14.1.2 Determinacy
The original Maxwell equations (14.1), for the three-dimensional case, con-
sists of eight first-order equations but with only six unknown vector com-
ponents; and for two-dimensional transverse electric (TE) and transverse
magnetic (TM) cases, four equations involving only three unknowns. That
is, the number of equations is larger than the number of unknown functions.
By taking the divergence of the Faraday and Ampere laws one can see that
two divergence equations will be satisfied for all time if they are satisfied ini-
tially. For these two reasons, in the CEM community it has been traditionally
believed that the full first-order Maxwell equations are "overspecified", and
14.1 The First-Order Maxwell Equations 335
8(eE) .
Vx + V x H - ~ - erE = Jlmp in [2, (14.5b)
n xE =0 on r 1, (14.5e)
on r 1, (14.5f)
nxH=O (14.5g)
The introduction of the variables 'P and X is purely for showing that the
Maxwell equations are not "overspecified". They do not enter into computa-
tion in contrast to the method proposed by Assous et al. (1993).
We shall prove that 'P and X in (14.5) are identically zeros, Le., system
(14.5) is equivalent to system (14.1). In fact, by virtue of the div-curl theorem
(Theorem 5.2), (14.5a) and (14.5b) are equivalent to the following equations:
8~E) ~p]
V· [VX+VxH-~-erE-J =0 in [2, (14.7b)
336 14. Electromagnetics
8(IEE)
n· [V'X+V'xH-~-aE-J
imp] =0 (14.7d)
Taking into account the divergence-free condition (14.5d) and the solvability
condition (14.4a), from (14.6b) we find that
Llcp = 0 in [2. (14.8a)
Taking into account the boundary condition (14.5e), Theorem 5.5, the bound-
ary condition (14.5f) and the solvability condition (14.4b), from (14.6c) we
obtain
n·V'cp=O onn. (14.8b)
From (14.5h) we have
cp=O (14.8c)
System (14.8) implies that cp == 0 in [2. Similarly, we can show that X == 0 in
[2. Therefore, cp and X in (14.5) are really dummy variables, and thus system
(14.5) is an equivalent formulation of system (14.1).
The first-order system (14.5) has eight equations, eight unknowns, and
four boundary conditions on each part of the boundary, and thus it is properly
determined. This is valid for static, transient and time-harmonic cases.
In the static case, the time-derivative terms in (14.5a) and (14.5b) dis-
appear, and aE is included into the given current density. System (14.5)
becomes two independent div-curl systems for the electric field and the mag-
netic field, respectively. It is well known that each div-curl system is elliptic
(Sect. 5.2).
In the time-harmonic case, when the time factor e jwt is used and sup-
pressed, the time-derivative terms become the zero-order terms, and system
(14.5) becomes two coupled div-curl systems. The coupling is through the
zero-order terms. The principle part, namely, the first-order derivative terms
which classify the system, still have the div-curl structure, and thus the whole
system is elliptic.
In the transient case, the whole system (14.5) is hyperbolic. However, in
time-domain numerical methods, the time-derivative terms are discretized by
either explicit or implicit finite differences, hence the time-derivative terms
become the zero-order terms in the space domain. For each time-step, the
time-discretized system is still elliptic.
In summary, in all cases, system (14.5) is properly determined, and is el-
liptic in the space domain. Since system (14.1) is equivalent to system (14.5),
system (14.1) is indeed properly determined and also elliptic in the space
domain. Therefore, the divergence equations (14.1c), (14.1d) and the bound-
ary conditions (14.3b) and (14.3d) are not "redundant", and must always be
taken into account.
14.1 The First-Order Maxwell Equations 337
adies in CEM could be cured by using the edge element method. The edge
elements allow the normal component to jump across material interfaces and
do not yield conflicting conditions at sharp corners. However, (1) such an
approach can only be used in the simple divergence-free case for isotropic
media; (2) the vector basis functions for the general quadrilateral elements,
which are desired for modeling complex geometry with higher solution accu-
racy than triangles, are in fact not divergence-free (Jin 1993, p.240); (3) edge
elements violate the normal field continuity between adjacent elements in the
homogeneous material domain, this will cause a stability problem for numer-
ical modeling in plasma physics (Assous et al. 1993); (4) the accuracy of edge
elements is lower than that of the nodal elements for the same number of un-
knowns, or the computational cost of edge elements is much higher than that
of nodal elements for the same accuracy (Mur 1994 and Monk 1993); (5) the
edge element interpolation itself cannot solve singularity problems at sharp
corners, the singularity problems can be solved only by the use of singular
elements, high-order interpolations (p-version elements) or fine meshes; (6)
the edge element method also needs non-conventional meshing and display
programs which are not normally available.
The curl-curl equations are derived from the first-order Maxwell curl
equations by applying the curl operator. As pointed out above the curl equa-
tions alone cannot determine a unique solution. Moreover, the curl-curl equa-
tions obtained by simple differentiation without additional equations and
boundary conditions admit more solutions than do its progenitors.
In order to derive an equivalent higher-order system from a system of
vector partial differential equations, one should use the div-curl method that
is based on Theorem 5.2: if a vector is divergence-free and curl-free in a
domain, and its normal component or tangential components on the bound-
ary are zero, then this vector is identically zero. In other words, the curl
and the divergence operators must act together with appropriate boundary
conditions to ensure that there are no spurious solutions in the resulting
higher-order equations. In this section, this div-curl method introduced in
Chap. 5 is employed to derive the second-order system of time-dependent
Maxwell equations and its boundary conditions; and to show that the diver-
gence equations and additional boundary conditions must be appended to the
curl-curl equations, and the Helmholtz-type equations must be solved with
divergence conditions enforced on the corresponding part of boundary. We
shall establish the Galerkin formulation corresponding to the Helmholtz-type
equations. We shall see that this Galerkin formulation is of the same form
as the popular Galerkin/penalty method with the penalty parameter s = 1.
Or more precisely, the parameter s should be chosen such that the resulting
equation becomes equivalent to the Helmholtz-type equation in homogeneous
regions. We shall also give a simple least-squares look-alike method to obtain
a correct variational formulation which rigorously justifies choosing s = 1 in
the penalty method.
340 14. Electromagnetics
Assume that the field intensities and the sources in system (14.1) are suf-
ficiently smooth. By virtue of the div-curl theorem (Theorem 5.2), system
(14.1) is equivalent to
\1 x (\1 x E) 8[8~E)]
+ J.L-
8t
- - + aE = -\1 X
8t
.
K1mp
8Jim p
-J.L--at"" in il, (14.11a)
real reason that the numerical methods based on the curl-curl equations will
give rise to spurious solutions.
It is difficult to solve a second-order curl-curl equation (14.11a) mixed
with an explicit constraint of the first-order divergence equation (14.11b).
We should look for a simple way. By virtue of Theorem 5.4 and the vector
identity (A.4), system (14.11) can be reduced to
in il, (14.13b)
nxE=O (14.13c)
\1 . (cE) = pimp (14.13d)
n· (cE) = 0 (14.13e)
n x (\1 X E) = -n X Kimp (14.13f)
It can be shown that (14.13b) may be eliminated (Sect. 14.2.3). That is, the
divergence equation (14.13b) is implicitly satisfied by the Helmholtz-type
equation (14.13a) with the boundary conditions. Therefore, system (14.13)
can be further simplified as
a [a(cE)] . aJimp
-L1E + J.L at ---at + aE = -\7 X K1mp -11---at
nxE=O (14.14b)
\1. (cE) = pimp on rl , (14.14c)
n· (cE) = 0 on r 2, (14.14d)
n X (\1 X E) = -n X Kimp (14.14e)
Similarly, we can have
(14.15e)
Since we are dealing with a second-order problem in the time-domain,
in addition to initial conditions for E and H, we need initial conditions for
aE/at and aH/Ot which can be obtained by using (14.1a) and (14.1b).
We note that the divergence equations are required to be satisfied only on
a part of boundary. This was first observed by Boyse et al. (1992) for a special
case. We will rigorously prove this in Sect. 14.2.3 by using the least-squares
method. The Helmholtz-type equations (14.14a) and (14.15a) can be found in
most text books on electromagnetics. However, it seems that all these books
(except the papers by Boyse et al. 1992, Mayergoyz and D'Angelo 1993)
claim that the Helmholtz-type equation must be solved with the divergence
equation satisfied in the whole domain. The rigorous derivation using the
div-curl method shows that the Helmholtz-type equation can stand alone,
and the divergence equation should be satisfied only on a part of boundary.
The advantages of using the Helmholtz-type equation over the curl-curl
equation are obvious: one avoids the difficulty involving explicit satisfaction
of the divergence equations, instead one solves three decoupled second-order
equations with coupled boundary conditions. We also should remark here that
the Helmholtz-type equations (14.14a) and (14.15a) are valid only for homo-
geneous and isotropic material regions. For inhomogeneous or anisotropic
materials, we still can use the div-curl method. However, the correct second-
order equations do not have LlE and LlH terms.
One may elect to use, for example, the finite difference method, to solve
the Helmholtz-type equations (14.14) or (4.15). Usually, the finite difference
method is based on rectangular structured grids. In this case, for example,
the divergence equation (14.15e) can be simplified as the Neumann boundary
condition:
a
-Hn=O
an
For complex geometry it is not straightforward to implement the Neumann
boundary condition in the finite difference method. By using the finite ele-
ment method based on a variational principle, only the essential boundary
conditions should be explicitly imposed, all natural boundary conditions in-
cluding divergence conditions on the boundary will be satisfied automatically
by the variational process. In the following we derive the variational formu-
lation corresponding to (14.14).
By taking into account the vector identity (A.4), the Galerkin formulation
associated with (14.14) is: Find E E £ such that
344 14. Electromagnetics
+ (- ~ [~. E -
pimp]) pimp
E* + (~ . E - - g - ' n . E*) n
-g- ,
( ~
+Ji- at
[a(gE)
at + ,
aE] E*)
+ Ji- at '
(aJimp E*) = 0 (14.16)
for all E* E e.
For time-harmonic eigenvalue problems with a = 0, the variational for-
mulation takes the form
(~x E, ~ x E*) + (~. E, ~. E*) -w 2Ji-g(E,E*) = 0, (14.18)
where w is the angular frequency.
The formulations for the magnetic field are analogous: Find H E 1£ such
that
(imp
"·E=- inn, (14.21b)
c
n x E= 0 on n, (14.21c)
n· (cE) = 0 on r2 , (14.21d)
where H is assumed to be known and to satisfy (14.1b) and (14.1d) as well as
the boundary conditions (14.3b) and (14.3c), and the source terms satisfy the
solvability conditions (14.4a)-{14.4e). In other words, when the magnetic field
and the sources are given, the solution of (14.21) will give the corresponding
electric field.
Obviously, system (14.21) is a typical div-curl system that can be treated
easily by the least-squares method, see the details in Sect. 5.4. To proceed,
we define a quadratic functional:
I: e -+ JR.,
I{E) = II" x E + a(~~) + Kimpl12 + II"· E - pi:P r (14.22)
346 14. Electromagnetics
pimp
+ ( \1. E - -c-' \1 . E
*) = 0, (14.23)
where E* = aE E E.
By using Green's formula (B.5) and considering the fact that H satisfies
(14.1b) and (14.3c), the H related term in (14.23) can be written as
( 8(J.LH)
8t '
\1 x E*) = (8(\1 x H) E*)
J.L at ' +
(8(J.LH)
at ,n x
E*)
r
= ( J.L ~ [8(eE)
m at +0- E + Jimp] , E*) . (14.24)
(14.25)
which is exactly the same as (14.17). By using Green's formula, from (14.25)
we can obtain the Euler-Lagrange equation (14.14a) and the natural bound-
ary condition (14.14c) and (14.14e). That is, the correctness of (14.14) or
(14.15) is completely proved.
Now we understand that if the field intensities are sufficiently smooth,
then the variational formulation (14.25), the Helmholtz-type equation (14.14a)
with its boundary conditions, and the first-order system (14.21) are equiv-
alent to one another. However, the finite element method based on (14.25)
has advantages: (1) the sources are required only to be square integrable; (2)
the test and trial functions from standard finite element spaces are required
to satisfy only the essential boundary conditions (14.21c) and (14.21d); (3)
the divergence equation (14.21b) is automatically satisfied,
We remark that the procedure of deriving formulation (14.25) in this sec-
tion is not a true least-squares approach, because (1) we have assumed that
H is given and satisfies (14.1b), and hence H is not subject to the varia-
tion; (2) the true least-squares method always leads to a symmetric bilinear
form; here the a related term is not symmetric. Even so, this procedure
is mathematically justifiable. It is nothing but a rigorous method to estab-
lish the Galerkin variational formulation corresponding to the Helmholtz-
type equations (14.14a) and their boundary conditions. All derivations pro-
vided in this section have rigorously proved that the penalty parameter in
the Galerkin/penalty method should be chosen such that it generates the
Helmholtz-type equation in homogeneous isotropic material regions.
14.2 The Second-Order Maxwell Equations 347
l n.KimPdr=O, (14.27c)
The permittivity e and the permeability I' are real symmetric positive-
definite matrices, i.e., 3 x 3 tensors in three-dimensions. Their inverse matrices
exist and are denoted by e- l and 1'-1, respectively. We may choose their
maximum eigenvalues as their norms denoted by lei and 11'1, respectively.
We always can decompose the matrix I' such that
I' = p.!(p.!)T. (14.28)
We would like to use the least-squares method to derive the second-order
equations for the electric field and to establish the corresponding variational
formulation. Consider the following div-curl system for the electric field:
348 14. Electromagnetics
a(ILH) _Kimp
\1xE= in n, (14.29a)
at
\1. (eE) = 0 in n, (14.29b)
nxE=O on r 1 , (14.29c)
where E* = ~E E E.
Upon using a relation similar to (14.24), from (14.31) we finally obtain
the variational formulation:
11L1(1L- 1 \1 x E, \1 x E*) + ler 2(\1. (eE), \1. (eE*))
nxE=O on rl , (14.33b)
V'·(eE)=O on r l , (14.33c)
n· (eE) = 0 on r 2 , (14.33d)
n x (JL-IV' x E) = -n x (JL-IKimp) on r 2• (14.33e)
We remark that both (14.32) and (14.33) can be generalized to spatially
variable media.
Similarly, we can obtain the variational formulation for the magnetic field:
Find H E 1£ such that
lei (e-IV' x H, V' x H*) + IJLI- 2 (V' . (JLH) , V' . (JLH*))
+Iel (:t22 (JLH) , H*) = lei (e- l Jimp, V' x H*) -lei (gt Kimp, H*) (14.34)
for all H* E 1£, where 1£ = {H E [H I (il)]3 : n x H = 0 on r 2, n· (JLH) =
o on rd.
By using the variational formulation we can show that the magnetic field
satisfies the following equation and boundary conditions:
82
lelV' x (e-IV' x H) -IJLI- 2 JLV'[V" (eH)] + lel-2 (JLH)
8t
in il, (14.35a)
n· (eH) = 0 (14.35b)
(14.35c)
nxH=O (14.35d)
V' . (eH) = 0 (14.35e)
When the field quantities do not vary with time, we have a static field. In
this case, there is no interaction between electric and magnetic fields, and
therefore we can separate either an electrostatic case or a magnetostatic case.
350 14. Electromagnetics
n xE = 0 on re , (14.36c)
n·E= 0 on r a, (14.36d)
n x E+ =n x E- on l1nt, (14.36e)
on r int , (14.36f)
8¢ =0 on r a, (14.38c)
8n
where gimp is a given constant voltage. The equation (14.38) is the well-
known Poisson equation with Dirichlet and Neumann boundary conditions
The classic Galerkin finite element method can be employed to obtain the
approximate solution of ¢ with an optimal accuracy. However, as we have
pointed out many times, the calculated electric field intensity E by numerical
differentiation of ¢ is of lower-order accuracy and discontinuous across the
element boundary.
System (14.36) is a typical div-curl system. In Sect. 5.2 we have shown by in-
troducing a dummy variable {} that in the three-dimensional case this system
is properly determined and elliptic. Now let us apply the least-squares method
to solve the div-curl system (14.36). We construct the following quadratic
functional:
14.3 Electrostatic Fields 351
I : E ----+ JR,
By using Green's formulae (B.3) and (B.5), the equation (14.40) can be con-
verted to
(V x (V x E),E*) + (V x E, n x E*)
pimp
+ ( V · E - - , n·E
*) =0 (14.44)
e ro
for all admissible E* E E, hence we have the Euler-Lagrange equation and
boundary conditions:
pimp)
6E=V ( - in il, (14.45a)
e
352 14. Electromagnetics
n x E= 0, (14.45b)
6Ez = -
a (pimp) in il, (14.46c)
aZ -c-
aEn pimp
El = 0, E2 = 0, - - = - - on Fe, (14.46d)
an c
E
n = °'an =
aEl 0,
an = 0,
aE2
on
r
a (14.46e)
(14.47g)
(14.47h)
(14.47i)
As discussed in Chap. 6, the curl equation (14.47a), which seems redun-
dant, must be included to ensure an optimal rate of convergence for the field
intensity and good numerical properties for iterative solvers.
Microstrip Line. As an example, let us apply LSFEM to a two-dimensional
problem of the shielded microstrip transmission line illustrated in Fig. 14.1.
This problem is taken from Jin (1993, p.98). When the microstrip line oper-
ates at low frequency, its capacitance and inductance can be obtained from a
static analysis, which amounts to solving (14.47) by assuming a certain po-
tential, say 1 volt, on its strip and a different potential, say zero volts, on the
shielding conductor. Due to its symmetric geometry, the domain of analysis
is reduced to half by specifying Ex = 0 at the plane of symmetry. On the
conducting surfaces, we impose n x E = 0 as shown in Fig. 14.1. It is impor-
tant not to specify the field intensity at the sharp corners of the strip. The
finite element mesh consisting of 949 linear quadrilateral elements is shown
in Fig. 14.2. The nodes at the material interface are doubly numbered. This
is a small-scale problem, so a direct solver is employed to solve the resulting
algebraic equations. For each interface node we add the following term
Q[(q,+ - q,-)2 + (Et - E;)2 + (c+ E: - c- E;)2]
<I>=O,Ex=O
<1>=0
Ex=O Ey=O
'<1>= 1,Ex=0
C\l=O,Ex=O
Fig. 14.1. Shielded microstrip line
354 14. Electromagnetics
W-
J
A-\\
\\
\ \ \
~c--
\ \\'\.J....\..
\
~ -
f-
II
~
~
~
~H
1--'-- j--l-
I-
I-- H l-
I-- jl7'lli
H ~kk<
~ ;4:
.x Fig. 14.2. The finite element mesh
of a shielded microstrip line
I •• , ••••••••• •• '
,
, ,
. I
•
I
, ' ,
I
I
If
,
••
...... .
••••••
• , ......
•
.
.. .
I \ I
I
I
I •
I',
t
. .. ... . .
, . .
f ' ' ., ••••• • •
, I I •••••• • • , • •
• ' , I I I , , , • • • • '• • •
',',
11/~""""--------" " " "'" ..... .
/, ~...:.::.:: : : : : ::: : : : : :
- ............. ....... ........ ....: :':.: :. :. ~
(14.48c)
(14.48d)
(14.48e)
on r int , (14.48f)
where n
is a perfectly conducting electric wall or a symmetric plane, r 2 is
a perfect magnetic wall, and nnt is an interface between two media without
imposed surface electric current. The given electric current density should
satisfy the solvability conditions:
'\l. Jimp = 0 in fl, (14.49a)
System (14.48) is of first order, and thus is difficult to solve directly by the
classic Galerkin method. In order to convert (14.48) into a system of second-
order equations, usually a vector potential A is introduced and is defined
as
J.LH=V x A. (14.50)
In this way, the divergence equation (14.48b) is automatically satisfied. Sub-
stituting (14.50) into (14.48a) yields
vx (1 Vx A) = JimP. (14.51)
J=O
j.Fl
Hy=O H:r=O
J=10
j.FlOOO
-----_
.... -.- .................................
, ,
'"--_ ,
~~ \
""
....................... \ I
.. . . . , . . . "'" , \ \
!--............._ ...... , ..... " " , \ I
!---~, \ I
~
_
.. !::-..'~""'" \ \ , 1
~ !:-..
~~~~~;"\\\\\, I
..........., ........" ' - " " " " " \ \ \ \ I
.. ''''\\\\\ I
'"-...... ~ ~~~ ~ ~ ~~ X\ , t
--'\\\
.....
-' ,
, ,\ \
\
\
I
\ \
I
Fig. 14.6. The computed
H field of a bus-bar
£ n.KimPdr=O, (14.60c)
V'. Er =0 in Q, (14.61e)
V'. Ei = 0 in Q, (14.61£)
V'·Hr=O in Q, (14.61g)
V'. Hi =0 in Q. (14.61h)
Obviously, system (14.61) is elliptic, since its principle part consists of four
div-curl systems. For the solution of (14.61) the least-squares variational
formulation is: Find u = (E r , E i , H r, Hi) E 1/. such that
B(u,v) = L(v) Vv = (E;,E;,H;,H;) E 1/., (14.62)
where 1/. = {u E [Hl(Q)]3 x [Hl(QW x [Hl(QW x [Hl(Q)j3 : n x E =
n, n, r
o on n . H = 0 on n x H = 0 on 2 , n . E = 0 on 2 }, and B(·, .) is r
the bilinear form
B(u, v) = (V' x Er - WILH i , V' x E; - WILH;)
+ (V' x Ei +wILHr. V' x E; +wILH;)
+ (V' x Hr +weEi , V' x H; + weE;)
+ (V' x Hi - weEr, V' x H; - weE;)
+ (V'. E r , V' . E;) + (V' . E i , V' . E;)
+ (V'. H r , V'. H;) + (V'. Hi, V'. H;), (14.63a)
and L(·) is the linear form
L(v)
(14.63b)
· *E 8Hz 0
JWe x - 8y = in Q, (14.64b)
· *E
JWe y + 8H
8x
z
=
0 in Q, (14.64c)
Hz = const on r, (14.64e)
on r. (14.64f)
Condition (14.64e) is an inhomogeneous version corresponding to (14.59c),
and (14.64f) is a 2D version of (14.59d). We also note that the boundary
conditions (14.64e) and (14.64f) satisfy the boundary compatibility condition
on r, (14.65)
JWe
· *Ex+-
af) - -
ax
aH-z =
ay
0 in n, (14.66b)
· *E
JWe y+ af) aHz
ay + ax = 0 in n, (14.66c)
in n. (14.66d)
By applying the operation ajax to the equation (14.66b) and the operation
ajay to the equation (14.66c), and by adding the results together we obtain
the Laplace equation for f):
a2{} a2{} in n.
ax 2 + ay2 = 0 (14.67a)
By taking the operation n· to (14.66b) and (14.66c) and using the boundary
compatibility condition (14.65) we obtain
af)=O
an on
r. (14.67b)
From (14.67) we know that f) = const, that is, system (14.64) is completely
equivalent to the augmented system (14.66) with four equations involving
four unknowns. Since system (14.66) consists of two two-dimensional div-curl
systems, and thus is elliptic. Therefore, system (14.64) is not 'overdetermined'
but indeed properly determined and elliptic.
For numerical calculation, separating the real and imaginary parts in
(14.64a)-(14.64d) leads to
aHzr
-W(erEyi + eiEyr) + ---a;- = 0 in fl, (14.68c)
aHzi
w (erExr - eiExi ) - BY =0 in fl, (14.68f)
aHzi
w (erEyr - eiEyi ) + --a;- =0 in fl, (14.68g)
in which
0 0 1 0 0 0
0 0 0 0 0 0
1 0 0 0 0 0
0 1 0 0 0 0
(14.70a)
0 0 0 0 0 1
0 0 0 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 -1 0 0 0 0
-1 0 0 0 0 0
0 0 0 0 0 0
0 0 1 0 0 0
(14.70b)
0 0 0 0 -1 0
0 0 0 -1 0 0
0 0 0 0 0 0
0 0 0 0 0 1
14.5 Time-Harmonic Fields 363
0 0 0 -Wf1 0 0
0 -WEi 0 0 -WEr 0
0 0 -WEi 0 0 -WEr
0 0 0 0 0 0
Ao= Wf1 0 0 0 0 0
(14.70c)
0 WEr 0 0 -WEi 0
0 0 WEr 0 0 -WEi
0 0 0 0 0 0
0 Hzr
0 Exr
0 Eyr
/= 0
U=
Hzi
(14.70d)
0 Exi
0 Eyi
At the interface r int between two contiguous media (+) and (-) the
following general conditions should be satisfied:
n x E+ =n x E- on r int , (14.71a)
(14.71b)
n· (E*+ E+) = n· (E*- E-) (14.71c)
n· (f1+ H+) = n· (f1- H-) (14.71d)
For two-dimensional TE waves, the interface conditions (14.71a) and
(14.71c) become
nx E yr xr -- nx E-
+ - ny E+ yr - ny E-
xr' (14.72a)
(14.72b)
Fig. 14.7. (a) The split cylinder and the mesh, (b) Contours of constant magnetic
field intensity Hr , (c) Contours of constant magnetic field intensity Hi, (d) Vectors
of the computed electric field intensity E r , (e) Vectors of the computed electric
field intensity Ei (from Jiang et al. 1996)
366 14. Electromagnetics
Fig. 14.S. (a) The off-center cylinder, (b) Contours of constant magnetic field
intensity H r , (c) Contours of constant magnetic field intensity Hi, (d) Vectors of
the computed electric field intensity E r , (e) Vectors of the computed electric field
intensity Ei (from Jiang et al. 1996)
14.6 Transient Scattering Waves 367
_ aE aHy
ax +lI
,.. at =0
z (14.73b)
in il,
(14.76)
ay + ~",aE
_ aHz
atx = 0
in il, (14.77a)
(14.77b)
(14.77c)
(14.77d)
14.6.2 Time-Discretization
____
1 aEn+l
z _____
z
1 aEn +J.t Hn+1
Y
- HnY =0, (14.79b)
2 ax 2 ax Llt
1 ( aH n+1
2 Y + aHn+l)
x + 1 (aHnY + aHn)
x +c En+l
Z
- EnZ 0
ax ay 2" - ax 8y Llt =,
(14.79c)
(14.79d)
and
aHzn+l _ _
1 _::"-_ 1 aHn
__z + IE: En+1
x
- Enx = 0, (14.80a)
2 ay 2 ay Llt
(14.80b)
aEn+1 aEn+l
tf;-+-jy=0, (14.80d)
where the superscript n denotes the nth time-step. Equations (14.79) and
(14.80) can be rewritten in a compact matrix form as:
au au
A1ax + A2 ay + Aou = j, (14.81)
U=
14. Electromagnetics
(E~f'
Hx
Hy
, /=
( _l~+~&
2
1 (8H y
2 8x
8y
1!t§... + J.l&
2 8x
_ 8H" )
8y
L\t
L\t
+ e!b..
L\t
r '
0
A2~ U
1
(~~ ~t)
0 0
~).
0 0
A1 =
e,
1
0 "2
1 0
1
Ao~ ~l}
J.l L\t
0
o '
r
e L\t 0
0 0 0
and for TE waves:
l~+e~
(H'f' ( 2 ay L\t
+ S
_l.aH.
2 ax
e
L\t
U= Ex , /=
Ey
_12 (~
ax
- ~
ay
) + J.l&
L\t
'
0
n ~). (T D'
1
0 0
A1 = 0
0 A2~ 0
1
-"2
1 a
1
Ao~ ( ~,
e L\t
J.l L\t
0
0
el,o ) .
. 0 0 0
At this stage it becomes clear that once one has a general purpose L8-
FEM code based on the standard first-order system (14.81), the only required
programming work is to write a subroutine to supply the matrices A, Ai and
/, and the boundary conditions.
We use a Jacobi-preconditioned conjugate gradient method to solve the
resulting algebraic equation system. By using this method, the iterations can
be conducted in such a manner that there is no need to form the element
matrices and to assemble the global matrix K. This greatly reduces the re-
quirement for computer memory. We have found that five conjugate gradient
iterations are enough to guarantee the accuracy of the solution in our calcu-
lations. The use of such a small number of iterations can be attributed to the
fact that when the time-step is small, the solution in the previous time-step
serves as a very good initial guess for the next step, thus the conjugate gra-
dient method needs fewer iterations than the case when a larger time-step is
14,6 Transient Scattering Waves 371
, Ez,i cos a
Hy"=-----
J.LC
for the TM case; and
Hz,i = Ho sin k(x cos a + y sina - ct),
, Hz,i sina
EX,t = - - - - - ,
€C
, Hz,icos a
EY"= - - - -
€c
for the TE case, where c is the wave velocity, k is the wave number, a is the
angle of incidence.
Numerical results are presented in terms of the surface current (for TM
case) and the radar cross section (RCS). The calculation ofthese requires the
field variables in the frequency-domain, thus the Fourier transform is per-
formed to transfer the field variables from the time-domain to the frequency-
domain.
The surface current J 8 for the TM case is defined as
InxHtl
J =
8 IHil '
where Hi and Ht are the incident and total magnetic fields in the frequency-
domain, respectively.
The RCS for the TM wave, STM' is defined as:
· 27rp 1E:(p,
STM (0) = 11m E'
0) 12 ,
p--+oo ~
where E: and E! are the scattered and incident electric fields, respectively,
in the frequency-domain, p is the distance from the center of the scattering
object, and 0 is the angle of observation. Note that the calculation of STM
from the above expression requires the knowledge of E:
at an infinite distance
from the scatterer. This is not available, since the Maxwell equations are
372 14. Electromagnetics
· 2 1H;(p, 0) 12
STE (0) = 11m
p-+oo
trp H'~
Fig. 14.9. Scattering by a circular cylinder: problem definition (from Wu and Jiang
1996)
Fig. 14.10. Scattering by a circular cylinder: finite element mesh (from Wu and
Jiang 1996)
14.6 Transient Scattering Waves 373
ka= 1, TM case
10' , - - - - - - - - - - - - - - - - ,
~ 10'
a:
10'
0 30 60 90 120 150 180
angle
ka=5, TM case
10'
\
1-- anat(b~
- LSFEM
\
(j)
() 10°
a:
\;~.-
10·'
0 30 60 90 120 150 180
angle
ka=10, TM case
10'
1- analytical \
~
- LSFEM
(j)
() 10°
a:
".
1\;......--~
• ••M
~
10·'
0 30 60 90 120 150 180
angle
Fig. 14.11. ReS for the circular cylinder (TM mode) (from Wu and Jiang 1996)
374 14. Electromagnetics
ka=l, TE case
10'
'" ~ ...
(f)
()
a:
10°
"-.. _
...../
I-analyticall
- LSFEM
10"
0 30 60 90 120 150 180
angle
ka=5, TE case
10'
10°
(f)
() 10"
a:
10.2
1- analytical 1
- LSFEM
10"
0 30 60 90 120 150 180
angle
ka=10, TE case
10'
\ ~--"-
100
vrv~
(f)
() 10"
a:
10.2
•
I=~:I
10.3
0 30 60 90 120 150 180
angle
Fig. 14.12. ReS for the circular cylinder (TE mode) (from Wu and Jiang 1996)
14.6 Transient Scattering Waves 375
Fig. 14.13. Scattered fields for the circular cylinder (Ez, Hz, Hy contours and H
vectors), TM mode, ka = 5 (from Wu and Jiang 1996)
376 14. Electromagnetics
e
"'----L--I-_ _ _...:x
EZ.i~
I" 2a ~I
Fig. 14.14. Scattering by a square cylinder: problem definition (from Wu and Jiang
1996)
Fig. 14.15. Scattering by a square cylinder: finite element mesh (from Wu and
Jiang 1996)
14.6 Transient Scattering Waves 377
5~---------------,
.. LSFEM
-MOM
• Shankar
o~----~------~------~----~
o 45 90 135 180
angle (degree)
Fig. 14.16. Normalized surface current for the square cylinder (from Wu and Jiang
1996)
Fig. 14.17. Scattered Ez contours for the square cylinder (from Wu and Jiang
1996)
2a
NACA0012 Airfoil. The LSFEM is also tested for two cases of scattering
TM wave by a NACA0012 airfoil defined in Fig. 14.18. The angles of incidence
are 0 and 90 degrees; and the values of ka are ka = 10 and ka = 1011",
respectively. Figure 14.19 shows the computational mesh which consists of
1782 bilinear elements and 1912 nodes. There are 156 divisions on the surface
of the airfoil. Figures 14.20 and 14.21 show the computed RCS. The result
in the case of 90 degree incidence is compared with those by Shankar et al.
(1989) and Vinh et al. (1992). Generally, fair agreement is observed. Note
that we have used a very coarse mesh and a very small domain. Finally Fig.
14.22 shows the contours of Ez for the 0 degree incidence case.
Fig. 14.19. Scattering by a NACA0012 airfoil: finite element mesh (from Wu and
Jiang 1996)
(f)
()
a:
o 30 60 90 120
angle
Fig. 14.20. RCS for the NACA0012 airfoil (0 degree incidence, ka = 10) (from Wu
and Jiang 1996)
14.6 Transient Scattering Waves 379
105
- - - LSFEM
104 --.
--
Vinh (lax-Wendrotf)
Vinh (Crank-Nicolson)
.c .--_ .. _.- Shankar
0,
c:
Q) 103
(j)
>
ctl
102
~
(J)
0
a: 10'
10°
10-'
0 90 180 270 360
angle
Fig. 14.21. RCS for the NACA0012 airfoil (90 degree incidence, ka = 1071") (from
Wu and Jiang 1996)
Fig. 14.22. Scattered Ez contours for the NACA0012 airfoil (0 degree incidence,
ka = 10) (from Wu and Jiang 1996)
Fig. 14.23. Comparison of the solutions of. the full Maxwell equations (left) and
the two curl equations (right), TM case, ka=lO, E z , Hx, Hy respectively (top to
bottom) (From Wu and Jiang 1996)
conditions significant spurious patterns exist for Hx and Hy near the surface.
These spurious patterns are similar to those in Figs. 2-b and 2-c of Vinh
et al. (1992) . Further calculation shows that the group 2 solutions possess
very large divergence of H in the vicinity of the scatterer surface. For the
case ka = 1 the absolute value of V' . H reaches its maximum of 21.28 in
an element adjacent to the surface. On the other hand, the maximum value
of V' . H for the group 1 is 0.054, which occurs near the outer boundary.
Since the divergence-free condition for E is automatically satisfied, almost
identical Ez distributions are obtained with the two formulations.
We should remark here that though group 1 and group 2 solutions differ
significantly, very little difference was observed in the ReS calculated based
on the two solutions. This is due to the fact that the erroneous part of the
solution is static, and the static field does not radiate. However, the erro-
neous solutions of the field intensities caused by neglecting the divergence
conditions cannot be accepted, for example, in plasma computations and in
the computation of impedance in digital circuits and electronic packaging.
14.7 Conclusion Remarks 381
their approach, edge elements are placed at interfaces to ensure that all local
continuity conditions can be met, and nodal elements are used elsewhere.
In the past, Assous et al. (1993) realized the importance of divergence
equations in the time-domain Maxwell equations and correctly worked with
second-order wave equations. Paulsen and Lynch (1991) and Boyse et al.
(1992) first gave a correct variational formulation corresponding to the
Helmholz-type operator in homogeneous regions.
Part V
(15.4)
An
where U e is an 'expanded' vector with non-zero values only in those com-
ponents corresponding to nodes of element e. Only non-zero entries of each
element matrix and element vector enter in the sum in (15.4). Hence the
individual matrix products V; for the sum in (15.4) can be computed as
dense element matrix-vector products at the element level rather than in the
expanded form shown. That is, we may calculate
(15.5)
independently and concurrently, and accumulate the vectors Y e into appro-
priate positions of the global vector yn for (15.4).
15.2 Matrix-Free Algorithm 387
If several iterations are carried out, then the element matrices may be saved
in a compact form. In fact, it is not even imperative that we store the element
matrices in cases where storage is extremely limited, since we can perform
the matrix-vector multiplication without forming the element matrices.
To illustrate the idea of matrix-free algorithm, we consider the computa-
tion of (15.5), i.e., the product Ve of an element matrix Ke by an element
vector U e , where the element matrix is given by (4.59); that is, we wish to
evaluate the expression
(15.6)
U e = ((Ub U2, ... , umh, (Ub U2, ... , umh, ... , (UI, U2, ... , Um)Nn ) T, (15.8)
in which all notations have been defined in Sect. 4.8.
Now two algorithms can be chosen. Both of them avoid formation and
storage of the element matrices. In the first algorithm, one forms the matrix
A~j, and computes
In Sect. 2.10, we have shown that for LSFEM K,(K) = O(h- 2 ) and thus
here we would have n = O(h- 2 ), i.e., a very large number of iterations have
to be performed. However, by choosing the direction vectors differently, one
obtains a more efficient method - the CG method.
o ~j ~ n. (15.16c)
By using the K-orthogonality and the linear independence of the Dj vec-
tors, the Nth residual RN can be shown to be orthogonal to N linearly
independent vectors and hence is the null vector. Therefore, in the absence
of round-off errors the method produces the solution in no more than N
iterations.
From computational considerations, we can have better formulas for the
CG method. Using
D n = _Rn + (3n_l Dn -l,
KD n = ~(Rn+1 _
R n ),
Tn
and orthogonality properties (15.16), we have
_(Rn,Dn) = (Rn, R n ),
15.3 The Conjugate Gradient Method 391
(R n+1 , Rn+ 1)
f3n = (Rn, Rn) . (15.18)
U:=U+TD;
R=R+TB; 81:= (R,R);
IF 81 ~ f THEN STOP;
392 15. The Element-by-Element Conjugate-Gradient Method
f3 := 81/80; 80 := 81;
D:= -R+f3D;
GOTO IT;
It can be shown that to reduce the error IU n - ill to less than f times
the initial error IUO - ill, the required number of iteration necessary in the
CG method is roughly given by
1 1
n ~ 2v'K(K)log~, (15.20)
which should be compared with K(K) in the case of the steepest descent
method. Thus, for large K(K), the conjugate gradient method is much more
efficient than the steepest descent method. For a typical application of LS-
FEM, the required number of iterations is proportional to order h- 1 which
is greatly superior to h- 2 for the steepest descent method.
U:= U +TD;
R=R+TB;
B:= Q-1R; 81 := (R,B);
IF 81 ~ f. THEN STOP;
f3 := 81/80; 80:= 81
D:= -B+f3D;
GOTO IT;
The statement B := Q-l R should be interpreted as solving the system
Q B = R. If the main diagonal of K is chosen to construct a diagonal matrix
Q, we have the well-known Jacobi preconditioned CG method (JCG). The
best choice would be Q = K = WTW with WTW the Cholesky factor-
ization of K, in which case ~(W-T KW- 1) = 1. However, this results in
a method which is not different from a direct solver. Since K is a sparse
matrix, we may require W to have the same sparsity as that of K. This
leads to an approximate Cholesky factorization of K. For example, we may
allow the entry Wij of W to be non-zero only if the corresponding entry k ij
of K is non-zero. This is a common variant of the so-called incomplete fac-
torization which takes only O(N) operations with considerable reduction of
the condition number, that is, ~(W-TKW-l) = O(h- 1 ) (Axelsson 1994).
We should mention that, although various incomplete factorization pre-
conditioning techniques can significantly reduce the number of iterations,
their overall performance should be evaluated by considering the cost of ad-
ditional memory requirement and total computer time.
Most LSFEM results presented in this book are obtained by using the
JCG method, since the diagonal pre conditioner is the simplest. Only five
N-dimensional arrays are needed for B, U, R, D and Q in the JCG method.
The experiments conducted by Jiang (1986) and Carey and Jiang (1986) show
that, even for small-scale (several hundred unknowns) LSFEM computations,
the element-by-element JCG method performs better than the sparse direct
solvers in storage requirement and computer time, particularly for non-linear
problems.
394 15. The Element-by-Element Conjugate-Gradient Method
Table 15.1. Comparison of total CPU time and computer storage of LSFEM and
GFEM (from Tang and Tsang 1995)
Flow Method Solver Degree of Total CPU Stopping Memmory
Freedom Time(sec) Time (Mb)
LSFEM JCG 10404 1795 80 1.07
LDCF LINPACK 10404 11426 80 18.42
GFEM NAG 7702 5111 80 30.39
LSFEM JCG 18207 16404 130 1.57
TDCF LINPACK 18207 64982 130 55.02
GFEM NAG 10303 18022 130 53.31
LSFEM JCG 9282 1336 75 0.80
RBC LINPACK 9282 8756 100 15.06
GFEM NAG 5228 1791 100 14.50
LSFEM JCG 26010 14510 180 2.07
DDF LINPACK 26010 98397 179 111.52
GFEM NAG 12904 49574 180 82.86
step was 0.5 for LSFEM, and 0.25 for the GFEM, respectively (the Newton's
method in the GFEM does not converge for Llt = 0.5). The LSFEM-JCG
required 1.3 Mb memory, whereas the GFEM in FIDAP required about 11
Mb. The CPU time on a HP-720 workstation were 8756 seconds and 39257
seconds for LSFEM and GFEM, respectively.
The matrix-free LSFEM-JCG has been successfully implemented on the
parallel computer CRAY-T3D by J. Wu (1994). Table 15.2 shows the speedup
for a typical computation of two-dimensional incompressible viscous flow.
The LSFEM gives rise to symmetric, positive definite systems of linear al-
gebraic equations. This system of equations can be efficiently solved by the
matrix-free element-by-element preconditioned conjugate gradient method.
The matrix-free element-by-element CG method proves to be very attrac-
tive when solving large problems, because
(1) The need for assemblage, storage and factorization of the global matrix
is circumvented, and thus the total storage and computational costs are low;
(2) For direct methods, it is advisable that the elements and nodes be num-
bered judiciously to reduce the bandwidth for both storage and efficiency. In
some cases, this is difficult to accomplish and one must rely on a sophisti-
cated preprocessor. Moreover, once the mesh in a local area is changed or
modified, as in adaptive refinement procedures, the elements and nodes must
be renumbered if a good band structure is to be preserved. However, for the
element-by-element CG method, the mesh structure and the numbering of
nodes can be arbitrary, the amount of storage is independent of meshing and
node numbering, and depends only on the number and the type of elements
in the mesh;
(3) Since element contributions are computed independently, the element-by-
element CG method can be easily implemented in parallel on multi-processors
without any special partitioning of the domain, topology of the mesh, or node
numbering;
(4) An interpolation of the coarse-mesh solution can serve as a very good
initial guess for the fine-mesh solution.
396 15. The Element-by-Element Conjugate Gradient Method
A. Operations on Vectors
\7 . (qv) = q\7 . v + \7q. v, (A.24)
\7 x (qv) = q\7 x v + \7q x v, (A.25)
\7 . (u x v) = (\7 x u) . v - u· (\7 x v), (A.26)
\7 x \7 x v = \7(\7 . v) - Llv, (A.27)
1
v x (\7 x v) = 2\7(v 2 ) - (v· \7)v, (A.28)
\7 x (u x v) = (v· \7)u - v(\7 . u) - (u· \7)v + u(\7 . v). (A.29)
B. Green's Formula
Assume that u, v and q are smooth enough. Integrating (A.I) and using the
Gauss divergence theorem lead to
(\7. v, q) + (v, \7q) = (n· v, q). (B.I)
Substituting v = Vp into (B.I) yields
(Llp, q) + (Vp, Vq) = (n· Vp, q). (B.2)
Substituting q = V . u into (B.I) yields
(V·v, V·u)+(v, V(V·u)) =(n·v, V·u). (B.3)
Replacing v by V x v in (B.I) leads to
(V x v, Vq) = (n· (V x v), q). (B.4)
Integrating (A.3) and using the Gauss divergence theorem lead to
(\7 x u, v) - (u, V x v) = (n x u, v). (B.5)
Substituting u = \7 q into (B.5) yields
(\7 x v, \7q) = -(n x \7q, v) = (\7q, n x v). (B.6)
Replacing v by \7 x v in (B.5) yields
(V x u, \7 x v) - (u, V x V x v) = (n x u, \7 x v). (B.7)
398 Appendices
C. Poincare Inequality
Let {1 be a bounded domain with a piecewise C1 boundary r, then
Ilplll " c {IIVpll~ + (In pdx) '} Vp E H'(Il), (C.1)
D. Lax-Milgram Theorem
Agmon, S., Douglis, A., Nirenberg, L. (1964): Estimates near the boundary for so-
lutions of elliptic partial differential equations satisfying general boundary con-
ditions II. Comm. Pure. Appl. Math. 17, 35-92
Ait-Ali-Yahia, D., Habashi, W.G., Tam, A., Vallet, M.-G., Fortin, M. (1996): A
directionally-adaptive finite element method for high-speed flows. Int. J. Num.
Meth. Fluids 23, 673-690
Ambrosiano, J.L., Brandon, S.T., Lohner, R. (1994): Electromagnetics via the
Taylor-Galerkin finite element method on unstructured grids. J. Compo Phys.
110, 310-319
Anilkumar, A.V., Lee, C.P., Wang, T.G. (1991): Surface-tension-induced mixing
following coalescence of initially stationary drops. Phys. Fluids A 3, 2587-2591
Armaly, B.F., Durst, F., Pereira, J.C.F., Schonung, B. (1982): Experimental and
theoretical investigation of backward-facing step flow. J. Fluid Mech. 127, 473-
496
Assous, F., Degond, P., Heintze, E., Raviart, P.A., Seger, J. (1993): On a finite-
element method for solving the three-dimensional Maxwell equations. J. Compo
Phys. 109, 222-237
Axelsson, O. (1994): Itemtive Solution Methods (Cambridge University Press, Cam-
bridge, England)
Axelsson, 0., Barker, V.A. (1984): Finite Element Solutions of Boundary Value
Problems (Academic Press, Orlando, Fla.)
Aziz, A.K., Kellogg, R.B., Stephens, A.B. (1985): Least squares methods for elliptic
systems. Math. Compo 44, 53-70
BabuSka, I. (1971): Error bounds for finite element method. Numer. Math. 16,
322-333
Barragy, E., Carey, G.F. (1988): A parallel element-by-element solution scheme.
Int. J. Num. Meth. Engrg. 26, 2367-2382
Batchelor, G.K. (1970): An Introduction to Fluid Dynamics (Cambridge University
Press, London)
Becker, E.B., Carey, G.F., Oden J.T. (1983): Finite Elements: An Introduction,
Vol. I (Prentice-Hall, Englewood Cliffs, NJ)
Beghein, C., Haghighat, F., Allard, F. (1992): Numerical study of double-diffusive
natural convection in a square cavity. Int. J. Heat Mass Trans. 35, 833-846
Bentley, L.R., Pinder, G.F. (1992): Least-squares method for solving the mixed
form of the groundwater flow equations. Int. J. Num. Meth. Fluids 14, 729-751
Bestehorn, M. (1993): Phase and amplitude instabilities for Benard-Marangoni
convection in fluid layers with large aspect ratio. Physical Review E 48, 3622-
3634
Bloomfield, P., Steiger, W.L. (1983): Least Absolute Deviations, Theory, Applica-
tions, and Algorithms (Birkhauser, Boston)
400 References
Carey, G.F., aden, J.T. (1983b): Finite Elements: Computational Aspects, Vol. III
(Prentice-Hall, Englewood Cliffs, NJ)
Carey, G.F. (1986): Parallelism in finite element modeling. Commun. Appl. Num.
Methods 2,281-287
Carey, G.F., Jiang, B.N. (1986): Element-by-element linear and nonlinear solution
schemes. Commun. Appl. Num. Methods 2, 145-153
Carey, G.F., aden, J.T. (1986): Finite Elements: Fluid Dynamics, Vol. IV (Prentice-
Hall, Englewood Cliffs, NJ)
Carey, G.F., Barragy, E., Mclay, R, Sharma, M. (1988): Element-by-element vector
and parallel computations. Commun. Appl. Num. Methods 4, 299-307
Carey, G.F., Jiang, B.N. (1988): Least-squares finite elements for first-order hyper-
bolic systems. Int. J. Numer. Meth. Engrg. 26, 81-93
Cartton, I. (1972): The effect of insulating vertical walls on the onset of motion in
a fluid heated from below. Int. J. Heat Mass Trans. 15, 665-572
Chan, D.C., Bailey, D.H., Bjorstad, P.E., Gilbert, J.R, Mascagni, M.V., Schreiber,
RS., Simon, H.D., Torczon, V.J., Watson, L.T. (1995): "A parallel least-squares
finite element method for subsonic and supersonic flows", in Proc. the Sev-
enth SIAM Conference on Parallel processing for Scientific Computing (SIAM,
Philadelphia) 179-180
Chang, C.L., Gunzburger, M.D. (1987): A finite element method for first order
elliptic systems in three dimensions. Appl. Math. Comput. 23, 135-146
Chang, C.L. (1990): A mixed finite element method for Stokes problem:
acceleration-pressure formulation. Appl. Math. Comput. 36, 135-146
Chang, C.L., Jiang, B.N. (1990): An error analysis of least-squares finite element
method of velocity-pressure-vorticity formulation for Stokes problem. Comput.
Meth. Appl. Mech. Engrg. 84, 247-255
Chang, C.L. (1992): Finite element approximation for grad-div type systems in the
plane. SIAM J. Numer. Anal. 29,452-461
Chang, C.L., Yang, S.Y., Hsu, J.S. (1995): A least-squares finite element method
for incompressible flow in stress-velocity-pressure version. Comput. Meth. Appl.
Mech. Engrg. 128, 1-9
Chang, Y.C., Hou, T.Y., Merriman, B., Osher, S. (1996): A level set formulation of
Eulerian interface capturing methods for incompressible fluid flows. J. Comput.
Phys. 124, 449-462
Chattot, J.J., Guiu-roux, J., Lamine, J. (1982): Numerical solution of a first-order
conservation equation by a least square method. Int. J. Num. Meth. Fluids 2,
209-219
Chen, C.M. (1982): Analysis of the Finite Element Method with Acceleration of
Convergence (Chinese) (Hunan Science and Technology, Changsha, China)
Chen, T.F. (1986): On least-squares approximations to compressible flow problems.
Numerical Method for Partial Differential Equations 2, 207-228
Chen, T.F., Fix, G.J. (1986): Least-squares finite element simulation of transonic
flows. Appl. Numer. Math. 2, 399-408
Chen, T.F. (1992): Semidiscrete least squares methods for linear convection-
diffusion problem. Comput. Math. Applic. 24, 29-44
Chenoweth, D.R, Paolucci, S. (1986): Natural convection in an enclosed vertical air
layer with large horizontal temperature differences. J. Fluid Mech. 169, 173-210
Chew, W.C. (1990): Waves and Fields in Inhomogeneous Media (Van Nostrand
Reinhold, New York)
Choi, Y.-H., Merkle, C.L. (1993): The application of preconditioning in viscous
flows. J. Comput. Phys. 105,207-223
Chorin, A.J. (1967): A numerical method for solving incompressible viscous flow
problems. J. Comput. Phys. 2, 12-26
402 References
Christie, I., Griffiths, D.F., Mitchell, A.R., Zienkiewicz, O.C. (1976): Finite element
methods for second order differential equations with significant first derivatives.
Int. J. Numer. Meth. Engrg. 10, 1389-1376
Ciarlet, P.G. (1991): Basic error estimates for elliptic problems. in Handbook of
Numerical Analysis, Vol. II, Finite Element Methods (Part 1), ed. by P.G. Ciarlet,
J.L. Lions (North-Holland, Amsterdam) 23-351
Cloot, A., Lebon, G. (1984): A nonlinear stability analysis of the Benard-Marangoni
problem. J. Fluid Mech. 145,447-469
Corr, D.G., Davies, J.B. (1972): Computer analysis of fundamental and higher
order modes in single and coupled microstrip. IEEE Trans. Microwave Theory
and Tech. 20,669-678
Cox, C.L., Fix, G.J. (1984): On the accuracy of least squares methods in the pres-
ence of corner singularities. Comput. Math. Appl. 10, 463-476
Crowley, C.W., Silvester, P.P., Hurwitz, H. (1988): Covariant projection elements
for 3D vector field problems. IEEE Trans. Magn. 24, 397-400
Daly, B.J. (1967): Numerical study of two fluid Rayleigh-Taylor instability. Phys.
Fluids 10, 297-307
Daly, B.J. (1969): A technique for including surface tension effects in hydrodynamic
calculations. J. Comput. Phys. 4, 97-117
Davis, S.H. (1967): Convection in a box: linear theory. J. Fluid Mech. 30,465-478
Davis, S.H. (1969): Buoyancy-surface tension instability by the method of energy.
J. Fluid Mech. 39, 347-359
Deconinck, H., Powell, K.G., Roe, P.L., Struijs, R. (1991): Multi-dimensional
schemes for scalar advection. AIAA-91-1532-CP.
Dendy, J.E. (1974): Two methods of Galerkin type achieving optimal L2 accuracy
for first order hyperbolics. SIAM J. Numer. Anal. 11, 637-£53
Donea, J. (1984): A Taylor-Galerkin method for convective transport problems.
Int. J. Numer. Meth. Engrg. 20, 101-119
Donea, J., Quartapelle, L. (1992): An introduction to finite element methods for
transient advection problems. Int. J. Num. Meth. Engrg. 95, 169-203
Eason, E.D. (1976): A review ofleast-squares methods for solving partial differential
equations. Int. J. Numer. Meth. Engrg. 10, 1021-1046
Eidelman, S., Colella, P., Shreeve, R.P. (1984): Application of the Godunov method
and its second-order extension to cascade flow modeling. AlA A J. 22, 1609-1651
Einset, E.O., Jensen, K.F. (1992): A finite element solution of three-dimensional
mixed convection gas flows in horizontal channels using preconditioned iterative
matrix methods. Int. J. Num. Meth. Fluids 14, 817-841
Engleman, M.S., Jamnia, M. (1990): Transient flow past a circular cylinder: a bench-
mark solution. Int. J. Num. Meth. Fluids 11, 985-1000
Ern, A., Smooke, M.D. (1993): Vorticity-velocity formulation for three-dimensional
steady compressible flows. J. Comput. Phys. 105, 58-71
Farrar, A., Adams, A.T. (1976): Computation of propagation constants for the
fundamental and higher modes in microstrip. IEEE Trans. Microwave Theory
and Tech. 24, 456-460
Feistauer, M. (1993): Mathematical Methods in Fluid Dynamics (Longman Scientific
& Technical, Harlow, England)
Fix, G.J., Rose, M.E. (1985): A comparative study of finite element and finite
difference methods for Cauchy-Reimann-type equations. SIAM J. Num. Analysis
22, 250--260
Fix, G.J., Gunzburger, M.D. (1978): On least squares approximations to indefinite
problems of the mixed type. Int. J. Numer. Meth. Engrg. 12,453-469
Fix, G.J., Gunzburger, M.D., Nicolaides, R.A. (1979): On finite element methods
of the least squares type, Compo & Maths. with Appls. 5, 87-98
References 403
Fletcher, C.A.J. (1979): A primitive variable finite element formulation for inviscid
compressible flow. J. Comput. Phys. 33,301-312
Foresti, S. Hassanzadeh, S., Murakami, H., Sonnad, V. (1990): Parallel implemen-
tation of iterative solution techniques with rapid operation on shared memory
machines. IBM Kingston Research Report, KGN-217
Fox, RL., Stanton, E.L. (1968): Developments in structural analysis by direct en-
ergy minimization. AIAA J. 6, 1036-1042
Franca, L.P., Hughes, T.J.R, Loula, A.F.D., Miranda, I. (1989): "A new family of
stable elements for the Stokes problem based on a mixed Galerkin/least squares
finite element formulation" , in Proc. Seventh International Conference on Finite
Element Methods in Flow Problems, Finite Element Analysis in Fluids, ed. by
T.J.Chung, G.RKarrr (VAH press, Huntsville, Alabama) 1067-1074.
Fried, I. (1969a): More on generalized iterative methods in finite element analysis.
AIAA J. 7,565-567
Fried, I. (1969b): Gradient methods for finite element eigenproblems. AIAA J., 7,
739-741
Fried, I. (1970): A gradient computational procedure for the solution of large prob-
lems arising from the finite element discretization method. Int. J. Num. Meth.
Engrg. 2, 477-494
Friedrichs, K.O. (1958): Symmetric positive linear differential equations. Comm.
Pure Appl. Math. 11,333-418
Fritts, M.J., Boris, J.P. (1979): The Lagrangian solution of transient problems in
hydrodynamics using a triangular mesh. J. Comput. Phys. 31, 173-215
Fyfe, D.E., Oran, E.S., Fritts, M.J. (1988): Surface tension and viscosity with La-
grangian hydrodynamics on a triangular mesh. J. Comput. Phys. 76, 349-384
Ganguly, A.K., Spielman, B.E. (1977): Dispersion characteristics for arbitrarily
configured transmission media, IEEE 7rans. Microwave Theory and Tech. 25,
1138-1141
Ghia, V., Ghia, K.N., Shin, C.T. (1982): High-Re solutions for incompressible flow
using the Navier-Stokes equation and a multigrid method. J. Comput. Phys. 48,
387-411
Girault, V., Raviart, P.-A. (1986): Finite Element Methods for Navier-Stokes Equa-
tions (Springer-Verlag, Berlin)
Gnoffo, P.A. (1983): A finite-volume, adaptive grid algorithm applied to planetary
entry flow-fields. AIAA J. 21, 1249-1254
Golub, G.H., Loan, C.F.V. (1989): Matrix Computations, 2nd Ed. (The Johns Hop-
kins Vniversity Press, Baltimore)
Gresho, P.M., Chan, S.T. (1988): Semi-consistent mass matrix technique for solv-
ing the incompressible Navier-Stokes equations. Lawrence Livermore National
Laboratory, VCRL-99503.
Gresho, P.M. (1991a): Some current CFD issues relevant to the incompressible
Navier-Stokes equations. Comput. Meth. Appl. Mech. Engrg. 87, 201-252
Gresho, P.M. (1991b): Incompressible fluid dynamics: some fundamental formula-
tion issues. Annu. Rev. Fluid Mech. 23, 413-453
Gresho, P.M. (1992): Some interesting issues in incompressible fluid dynamics, both
in the continuum and in numerical simulation. Advances in Applied Mechanics
28,45-137
Gresho, P.M., Chan, S.T. (1995): An update on projection methods for transient
incompressible viscous flow. Lawrence Livermore National Laboratory, VCRL-
JC-121525
Gunzburger, M.D. (1989): Finite Element Methods for Viscous Incompressible
Flows (Academic Press, San Diego)
404 References
Hafez, M., Brucker, D. (1991): The effect of artificial vorticity on the discrete solu-
tion of Euler equations. AIAA paper 91-1553
Hageman, L.A., Young, D.M. (1981): Applied Itemtive Methods (Academic Press,
New York)
Haj-hariri, H., Shi, Q., Borhan, A. (1994): Effect of local property smearing on
global variables: implication for numerical simulations of multiphase flows. Phys.
Fluids 6, 2555-2257
Hara, M., Wada, T., Fukasawa, T., Kikuchi, F. (1983): Three dimensional analysis
of RF electromagnetic fields by finite element method. IEEE Trans. Magn. 19,
2417-2420
Harbord, R, Gellert, M. (1991): A simple least-squares method for FE analysis of
the Navier-Stokes problem. Comput. Mech. 8, 19-24
Harlow, F.H., Welch, J.E. (1965): Numerical calculations of time-dependent viscous
incompressible flow of fluid with free Surface. Phys. Fluids 8, 2182-2189
Harlow, F.H., Shannon, J.P. (1967): The splash of a liquid drop. J. Comput. Phys.
38, 3855-3866
Hayes, L.J., Devloo, P. (1985): "An element-by-element block iterative method for
large non-linear problems", in Innovative Methods for Nonlinear Behavior, ed.
by W.K. Liu et al. (Pineridge Press, Swansea)
Hayes, L.J., Devloo, P. (1986): A vectorized version of a sparse matrix-vector mul-
tiply. Int. J. Num. Meth. Engrg. 23, 1043-1056
Hayes, L.J. (1989): "Advances and trends in element-by-element techniques", in
State-of-the-Art Surveys on Computational Mechanics, ed. by A.K.Noor, J.T.
aden, 219-236
Hestenes, M.R, Stieffel, E.L. (1952): Methods of conjugate gradient for solving
linear systems. Nat. Bur. Std. J. Res. 49, 409-436
Hillion, P. (1997): Beware of Maxwell's divergence equations. J. Comput. Phys.
132, 154-155
Hirsch, C. (1990): Numerical Computation of Internal and External Flows, Vol. 2
(John Wiley, Chichester)
Rirt, C.W., Nichols, B.D. (1981): Volume of fluid (VOF) method for the dynamics
of free boundaries. J. Comput. Phys. 39, 201-225
Hughes, T.J.R., Levit, I., Winget, J. (1983a): An element-by-element implicit algo-
rithm for heat conduction. ASCE J. Engrg. Mech. Div. 109,576-583
Hughes, T.J.R, Levit, I., Winget, J. (1983b): An element-by-element solution algo-
rithm for problems of structural and solid mechanics. Compo Meth. Appl. Mech.
Engrg. 36, 241-254
Hughes, T.J.R (1987): Recent progress in the development and understanding of
SUPG methods with special reference to the compressible Euler and Navier-
Stokes equations. Int. J. Numer. Meth. Fluids 7, 1261-1275
Hughes, T.J.R, Ferencz, RM., Hallquist, J.O. (1987): Large-scale vectorized im-
plicit calculation in solid mechanics on a CRAY X-MP/48 utilizing EBE precon-
ditioned conjugate gradients. Compo Meth. Appl. Mech. Engrg. 61, 215-248
Hughes, T.J.R, Franca, L.P., Hulbert, G.M. (1989): A new finite element formula-
tion for computational fluid mechanics: VIII. The Galerkin/least-squares method
for advective-diffusive equations. Comput. Methods Appl. Mech. Engrg. 73, 173-
189
Ikohagi, T., Shin, B.R, Daiguji, H. (1992): Application of an implicit time-marching
scheme to a three-dimensional incompressible flow problem in curvilinear coor-
dinate systems. Computer Fluids 21, 163-175
Imai, I. (1941): On the flow of a compressible fluid past a circular cylinder, II. Proc.
Phys-Math. Soc. Japan 23, 180-193
References 405
Iwatsu, R., Ishii, K, Kawamura, T., Kuwahara, K, Hyun, J.M. (1989a): Simulation
of transition to turbulence in a cubic cavity. AIAA papaer 89-0040
Iwatsu, R., Ishii, K, Kawamura, T., Kuwahara, K, Hyun, J.M. (1989b): Numerical
simulation of three dimensional flow structure in a driven cavity. Fluid Dynamics
Research 5,173-189
Jacqmin, D. (1995): Three-dimensional computations of droplet collisions, coales-
cence, and droplet/wall interactions using a continuum surface-tension method.
AIAA-95-D883
Jiang, B.N., Chai, J.Z. (1980): Least squares finite element analysis of steady high
subsonic plane potential flows. Acta Mechanica Sinica, No.1, 90-93 (Chinese).
[English transl.: Foreign Technology Div., Air Force Systems Command, Wright-
Patterson AFB, OH., FTD-ID(RS)T-0708-81, 53-59]
Jiang, B.N., Carey, C.F. (1984): "Sub critical flow computation using an element-
by-element conjugate-gradient method" , in Proc. 5th Int. Symp. Finite Elements
and Flow Problems (University of Texas at Austin) 103-106
Jiang, B.N. (1986): Least-squares finite element methods with element-by-element
solution including adaptive refinement. Ph.D. dissertation (The University of
Texas at Austin)
Jiang, B.N., Carey, C.F. (1987): Adaptive refinement for least-squares finite ele-
ments with element-by-element conjugate gradient solution. Int. J. Numer. Meth.
Engrg. 24, 569-580
Jiang, B.N., Carey, C.F. (1988): A stable least-squares finite element method for
nonlinear hyperbolic problems. Int. J. Numer. Meth. Fluids 8, 933-942
Jiang, B.N., Chang, C.L. (1988): Least-squares finite elements for Stokes problem.
NASA TM 101308, ICOMP-88-1
Jiang, B.N., Chang, C.L. (1990): Least-squares finite elements for Stokes problem.
Comput. Meth. Appl. Mech. Engrg. 78, 297-311
Jiang, B.N., Carey, C.F. (1990): Least-squares finite element method for compress-
ible Euler equations. Int. J. for Num. Fluids 10, 557-568
Jiang, B.N., Povinelli, L.A. (1990): Least-squares finite element method for fluid dy-
namics. Comput. Meth. Appl. Mech. Engrg. 81, 13-37. First appeared as NASA
TM 102352, ICOMP-89-23
Jiang, B.N. (1991): The L1 finite element method for pure convection problems.
NASA TM 103773, ICOMP-91-03
Jiang, B.N. (1992): A least-squares finite element method for incompressible
Navier-Stokes problems. Inter. J. Numer. Meth. Fluids 14, 843-859. First ap-
peared as NASA TM 102385, ICOMP-89-28
Jiang, B.N. (1993): Non-oscillatory and non-diffusive solution of convection prob-
lems by the iteratively reweighted least-squares finite element method. J. Com-
put. Phys. 105, 108-121
Jiang, B.N., Povinelli, L.A. (1993): Optimal least-squares finite element method
for elliptic problems. Comput. Meth. Appl. Mech. Engrg. 102, 199-212. First
appeared as NASA TM 105382, ICOMP-91-29
Jiang, B.N., Lin, T.L., Povinelli, L.A. (1994a): Large-scale computation of incom-
pressible viscous flow by least-squares finite element method. Comput. Methods
Appl. Mech. Engrg. 114, 213-231. First appeared as NASA TM 105904, ICOMP-
93-06
Jiang, B.N., Loh, C.Y., Povinelli, L.A. (1994b): Theoretical study of the incompress-
ible Navier-Stokes equations by the least-squares method. NASA TM 106535,
ICOMP-94-04
Jiang, B.N., Sonnad, V. (1994): Least-squares solution of incompressible Navier-
Stokes equations with the p-version of finite elements. Comput. Mech. 15, 129-
136. First appeared as NASA TM 105203, ICOMP-91-14
406 References
Jiang, B.N., Hou, L.J., Lin, T.L., Povinelli, L.A. (1995): Least-squares finite ele-
ment solutions for three-dimensional backward-facing step flow. Inter. J. Comput.
Fluid Dynamics 4, 1-19. First appeared as NASA TM 106353, ICOMP-93-31
Jiang, B.N., Wu, J., Povinelli, L.A. (1996): The origin of spurious solutions in
computational electromagnetics. J. Comput. Phys. 125, 104-123. First appeared
as NASA TM 106921, ICOMP-95-8
Jiang, B.N. (1997): "The true origin of spurious solutions and their avoidance
by the least-squares finite element method", in Computational Electromagnet-
ics and Its Applications, ed. by Campbell, T.G., Nicolaides, RA., Salas, M.D.,
ICASE/LaRC Interdisciplinary Series in Science and Engineering, VoI.5, (Kluwer
Academic, Dordrecht, Netherlands), pp. 155-184
Jiang, B.N. (1998): On the least-squares method. Comput. Meth. AppI. Mech.
Engrg. 152, 239-257
Jin, J.M. (1993): The Finite Element Method in Electro-Magnetics (John Wiley and
Sons, New York)
Johnson, C. (1987): Numerical Solution of Partial Differential Equations by the
Finite Element Method (Cambridge University Press, Cambridge, England)
Johnson, C., Navert, U., Pitkaranta, J. (1984): Finite element methods for linear
hyperbolic problems. Comput. Meth. AppI. Mech. Engrg. 45, 285-312
Jue, T.C., Ramaswamy, B. (1992): "Cavity natural convection with a deformable
free surface", in Advances in Finite Element Analysis in Fluid Dynamics, ed.
by M. N. Dhaubhadel, et aI. (ASME, New York)
Kangro, U., Nicolaides, R (1997): Spurious fields in time-domain computations of
scattering problems. IEEE Trans. Antennas and Propagat. 45, 228-234
Kececioglu, I., Rubinsky, B. (1989): A mixed-variable continuously deforming finite
element method for parabolic evolution problems. Part II: The coupled problem
of phase-change in porous media. Int. J. Num. Meth. Engrg. 28,2609-2634
Kong, J.A. (1990): Electromagnetic Wave Theory (John Wiley and Sons, New York)
Koschmieder, E.L. (1967): On convection under an air surface. J. Fluid Mech. 30,
9-15
Koschmieder, E.L., Biggerstaff, M.1. (1986): Onset of surface-tension-driven Benard
convection. J. Fluid Mech. 161,49-64
Koschmieder, E.L., Prahl, S.A. (1990): Surface-tension-driven Benard convection
in small containers. J. Fluid Mech. 215, 571-583
Koschmieder, E.L., Switzer, D.W. (1992): The wavenumbers of supercritical
surface-tension-driven Benard convection. J. Fluid Mech. 240, 533-548
Koseff, J.R, Street, RL. (1984a): Visualization studies of a shear driven three-
dimensional recirculation flow. J. Fluids Eng. 106, 21-29
Koseff, J.R., Street, RL. (1984b): On end wall effects in a lid-driven cavity flow. J.
Fluids Eng. 106, 385-389
Koseff, J.R., Street, RL. (1984c): The lid-driven cavity flow: a synthesis of quali-
tative and quantitative observations. J. Fluids Eng. 106, 390--398
Kfizeki, M., Neittaanmaki, P. (1984a): On the validity of Friedrich's inequalities.
Math. Scand. 54, 17-26
Kfizeki, M., Neittaanmaki, P. (1984b): Finite element approximation for a div-rot
system with mixed boundary conditions in non-smooth plane domains. ApI. Mat.
29,272-285
Kfizeki, M., Neittaanmaki, P. (1990): Finite Element Approximation of Variational
Problems and Applications (Pitman Scientific and Technical, Hawlow, England)
Ku, H.C., Hirsh, R.S., Taylor, T.D. (1987) A pseudospectral method for solution of
the three dimensional incompressible Navier-Stokes equations. J. Comput. Phys.
10,439-462
References 407
Ku, H.C., Hirsh, RS., Taylor, T.D., Rosenberg, A.P. (1989): A pseudospectral
matrix element for solution of three-dimensional incompressible flows and its
parallel implementation. J. Comput. Phys. 83, 260-291
Kunz, K.S., Luebbers, RJ. (1993): The Finite Difference Time Domain Method for
Electro-Magnetics (CRC Press, Boca Raton)
Lafaurie, B., Nardone, C., Scardovelli, R, Zaleski, S., Zanetti, G. (1994): Modeling
merging and fragmentation in multiphase flows with SURFER J. Comput. Phys.
113, 134-147
Lager, I.E. (1996): Finite element modeling of static and stationary electric and
magnetic fields. Thesis (Delft University Press, Delft, Netherlands)
Lager, I.E., Mur, G. (1996): The finite element modeling of static and stationary
electric and magnetic field. IEEE Transactions on Magnetics 32, 631-634
Laminie, J. (1994): Some aspect of computational fluid dynamics: finite element
discretization, parallel implementation. Thesis (Universite de Paris-Sud)
Landau, L. D. and Lifshitz, E. M. (1959): Fluid Mechanics (Pergammon, New York)
Lavery, J.E. (1988): Nonoscillatory solution of the steady-state inviscid Burgers'
equation by mathematical programming. J. Comput. Phys. 79, 436-448
Lavery, J.E. (1989): Solution of steady-state one-dimensional conservation laws by
mathematical programming. SIAM J. Numer. Anal. 26, 1081-1089
Lee, J.F., Lee, R, Cangellaris, A. (1997): Time-domain finite element methods.
IEEE Transaction on Antennas and Propagat. 45, 430-441
Lee, RL., Madsen, N.K. (1990): A mixed finite element formulation for maxwell's
equations in the time domain. J. Comput. Phys. 88, 284-304
Lefebvre, D., Peraire, J., Morgan, K. (1992): Least squares finite element solution
of compressible and incompressible flows. Int. J. Num. Meth. Heat Trans. Fluid
Flow 2,99-113
Lefebvre, D., Peraire, J., Morgan, K. (1993): Finite element Least squares solution of
the Euler equations using linear and quadratic approximations. Inter. J. Comput.
Fluid Dynamics 1, 1-23
Li, C.W. (1990): Least-squares characteristics and finite elements for advection
dispersion simulation. Int. J. Num. Meth. Engrg. 29, 1343-1364
Lighthill, J. (1986): Informal Introduction to Theoretical Fluid Dynamics (Claren-
don Press, Oxford)
Lin, T.F., Huang, C.C., Chang, T.S. (1990): Transient binary mixture natural con-
vection in square enclosures, Int. J. Heat Mass Trans. 33, 287-299
Lions, J.L., Magenes, E. (1972): Non-homogeneous Boundary- Value Problems and
Applications, Vol. I (Springer-Verlag, New York)
Lock, RC. (1970): Test cases for numerical methods in two-dimensional transonic
flows. AGARD-R-575.
Long, P.E., Peper, D.W. (1981): An examination of some simple numerical schemes
for calculating scalar advection. J. Appl. Met. 20, 1146-1156
Lowrie, RB., Roe, P.L. (1994): On the numerical solution of conservation laws by
minimizing residuals. J. Comput. Phys. 113, 304-308
Lynn, P.P., Arya, S.K. (1973): Use of the least squares criterion in the finite element
formulation. Int. J. Num. Meth. Engrg. 6,75-88
Madsen, N.K., Ziolkowski, R (1988): Numerical solution of Maxwell's equations in
the time domain using irregular nonorthogonal grids. Wave Motion 10, 583-596
Martin, J.C., Moyce, W.J. (1952): An experimental study of the collapse of liquid
columns on a horizontal plane. Philos. Trans. R. Soc. Lond. A 244, 312-324
Mayergoyz, I.D., D'Angelo, J. (1993): A new point of view on the mathematical
structure of Maxwell's equations. IEEE Trans. Magn. 29, 1315-1320
Merkle, C.L., Choi, Y.H. (1987): Computation oflow-speed flow with heat addition.
AIAA J. 25,831-838
408 References
Merkle, C.L., Choi, Y.H. (1988): Computation of low-speed compressible flows with
time-marching procedures. Int. J. Num. Meth. Engrg. 25,293-311
Monk, P. (1993): "Finite element time domain methods for Maxwell equations",
in Second International Conference on Mathematical and Numerical Aspects of
Wave Propagation, ed. by I. Stakgold et al (SIAM, Philadelphia) 380-389
Morgan, K, Hassan, 0., Peraire, J. (1996): A time-domain unstructured grid ap-
proach to the simulation of electromagnetic scattering in piecewise homogeneous
media. Int. J. Num. Meth. Engrg. 134, 17-36
Morton, KW., Parrot, A.K (1980): Generalised Galerkin methods for first-order
hyperbolic equations. J. Comput. Phys. 36, 249-270
Mur, G. (1994): Edge elements, their advantages and their disadvantages. IEEE
Trans. Magn. 30, 3552-3557
Nakayama, T., Mori, M. (1996): An Eulerian finite element method for time-
dependent free surface problems in hydrodynamics. Int. J. Num. Meth. Fluids
22, 175-194
Nedelec, J. (1980): Mixed finite elements in R3. Numer. math. 35, 315-341
Neittaanmiiki, P., Sarannen, J. (1981a): Finite element approximation of vector
fields given by curl and divergence. Math. Methods Appl. Sci. 3, 328-335
Neittaanmiiki, P., Sarannen, J. (1981b): Finite element approximation of electro-
magnetic fields in the three dimensional space. Numer. Funct. Anal. Optim. 2,
487-506
Nelson, J.J., Chang, C.L. (1995): A mass conservative least-squares finite element
method for the Stokes problem. Commun. Appl. Num. Methods 11, 965-970
Nguyen, H., Reynen, J. (1984): A space-time least-squares finite element scheme for
advection-diffusion equations. Comput. Meth. Appl. Mech. Engrg. 42, 331-342
Nield, D. A. (1964): Surface tension and buoyancy effects in cellular convection. J.
Fluid Mech., 19, 341-352
Ni, R-H. (1982): A multiple grid scheme for solving the Euler equations. AIAA J.
20, 1565-1571
Noack, RW., Anderson, D.A. (1992): Time-domain solution of Maxwell equations
using a finite-volume formulation. AIAA Paper 92-0451
Nobari, M.R, Tryggvason, G. (1994a): Numerical simulation of drop collisions.
NASA TM 106751, ICOMP-94-23
Nobari, M.R, Tryggvason, G. (1994b): The flow induced by the coalescence of two
initially stationary drops. NASA TM 106752, ICOMP-94-24
aden, J.T., Reddy, J.N. (1976): An Introduction to the Mathematical Theory of
Finite Elements (John Wiley and Sons, New York)
aden, J.T., Carey, G.F. (1983): Finite Elements, Mathematical Aspects, Vol. IV
(Prentice-Hall, Englewood Cliffs, New Jersey)
aden, J.T., Jacquotte, O.-P. (1984): Stability of some mixed finite element methods
for Stokesian flows. Comput. Meth. Appl. Mech. Engrg. 43, 231-248
aden, J.T. (1991): "Finite elements: An introduction", in Handbook of Numerical
Analysis, Vol II, Finite Element Methods (Part 1), ed. by P.G. Ciarlet, J.L. Lions
(North-Holland, Amsterdam) 3-15
aden, J.T., Demkowicz, L.F. (1996): Applied Functional Analysis (CRC, Boca Ra-
ton)
adman, M.T., Russell, A.G. (1993): A nonlinear filtering algorithm for multi-
dimensional finite element pollutant advection schemes. Atmospheric Environ-
ment 27A, 793-799
Ortiz, M., Pinsky, P.M., Taloy, RL. (1983): Unconditionally stable element-by-
element algorithm for dynamic problems. Compo Meth. Appl. Mech. Engrg. 36,
223-239
References 409
Park, N.S., Liggett, J.A. (1990): Taylor-least squares finite element for two-
dimensional advection dominated advection-diffusion problems. Int. J. Numer.
Meth. Fluids 11, 21-28
Park, N.S., Liggett, J.A. (1991): Application of Taylor-least squares finite element
for three-dimensional advection-diffusion equation. Int. J. Num. Meth. Fluids
13,759-733
Paulsen, K.D., Lynch, D.R. (1991): Elimination of vector parasites in finite element
Maxwell solutions. IEEE Trans. Microwave Theory and Tech. 39, 395-404
Pearson, J.R. (1958): On convection cells induced by surface tension. J. Fluid Mech.
4,489-500
Pehlivanov, A.I., Carey, G.F., Lazarov, R.D., Shen, Y. (1993): Convergence ofleast-
squares finite elements for first-order ODE systems. Computing 51, 111-123
Pinchuk, A.R., Crowley, C.W., Silvester, P.P. (1988): Spurious solutions to vector
diffusion and wave field problems. IEEE Trans. Magn. 24, 158-161
Pironneau, O. (1989): Finite Element Methods for Fluids (John Wiley and Sons,
Chichester)
Prenter, P.M. (1975): Splines and Variational Methods (John Wiley and Sons,
Chichester)
Pulliam, T.H., Barton, J.T. (1985): Euler computation of AGARD working group
07, Airfoil test cases. AIAA Paper 85-0018
Pulliam, T.H. (1989): A computational challenge: Euler solution for ellipses. AIAA
paper 89-0469
Quartapelle, L. (1993): Numerical Solution of the Incompressible Navier-Stokes
Equations (Birkhauser Verlag, Basel)
Rahman, B.M.A., Davies, J.B. (1984): Penalty function improvement of waveguide
solution by finite elements. IEEE Trans. Microwave Theory and Tech. 32, 922-
928
Raviart, P.-A, Thoman, J.-M. (1977): Primal hybrid finite element methods for
second order elliptic equations. Math. Compo 31, 391-413
Reddy, J.N., Gartiing, D.K. (1994): The Finite Element Method in Heat Transfer
and Fluid Dynamics (CRC Press, Boca Raton)
Reddy, M.P., Reddy, J.N., Akay, H.U. (1992): Penalty finite element analysis of
incompressible flows using element by element solution algorithm. Compo Meth.
Appl. Mech. Engrg. 100, 169-205
Richards, J. R., Lenhoff, A. M., Beris, A. N. (1994): Dynamics breakup of liquid-
liquid jets. Phys. Fluids A 6, 2640--2655
Roberts, J.E., Thomas, J.-M. (1991): "Mixed and hybrid methods", in: Handbook
of Numerical Analysis, Vol. II. Finite Element Methods, ed. by P.G. Ciarlet, J.L.
Lions (North-Holland, Amsterdam) 523-639
Roshko, A. (1953): On the development of turbulent wakes from vortex streets.
NASA TN 2913
Saad, Y. (1986): GMRES: A generalized minimal residual algorithm for solving
nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 7, 856-869
Saranen, J. (1982): On an inequality of Friedrichs. Math. Scand. 51,310-322
Scanlon, J.W., Segal, L.A. (1967): Finite amplitude cellular convection induced by
surface tension. J. Fluid Mech. 30, 149-162
Schwieg, E., Bridges, W.B. (1984): Computer analysis of dielectric waveguides: a
finite difference method. IEEE Trans. Microwave Theory and Tech. 32, 531-541
Shakib, F., Hughes, T.J.R., Johan, Z. (1989): A multi-element group preconditioned
GMRES algorithm for nonsymmetric systems arising in finite element analysis.
Compo Meth. Appl. Mech. Engrg. 75, 415-456
Shang, J.S. (1995): A fractional-step method for solving 3-D time-domain Maxwell's
equations. J. Comput. Phys. 109-119
410 References
Tezduyar, T.E., Behr, M., Liou, J. (1992a): A new strategy for finite element com-
putations involving moving boundaries and interfaces - the deforming-spatial-
domain/space-time procedure: I. the concept and the preliminary numerical tests.
Compo Meth. Appl. Mech. Engrg. 94, 339-351
Tezduyar, T.E., Behr, M., Mittal, S., Liou, J. (1992b): A new strategy for finite ele-
ment computations involving moving boundaries and interfaces - the deforming-
spatial-domain/space-time procedure: II. computation of free-surface flows, two-
liquid flows, and flows with drifting cylinders. Compo Meth. Appl. Mech. Engrg.
94,353-371
Tezduyar, T.E., Mittal, S., Ray S.E., Shih, R. (1992c): Incompressible flow com-
putations with stablized bilinear and linear equal-order-interpolation velocity-
pressure elements. Comput. Meth. Appl. Mech. Engrg. 95, 221-242
Thess, A., Orszag, S.A. (1995): Surface-tension-driven Benard convection at infinite
Prandtl number. J. Fluid Mech. 283, 201-230
Turkel, E. (1987): Preconditioning methods for solving the incompressible and low
speed compressible equations. J. Comput. Phys. 72, 277-298
Turkel, E. (1993): Review of preconditioning methods for fluid dynamics. Appl.
Num. Math. 12, 257-284
Umashankar, KR., Taflove, A. (1993): Computational Electromagnetics (Artech
House, Boston)
Unverdi, S.O., Tryggvason, G. (1992): A front-tracking method for viscous, incom-
pressible, multi-fluid flows. J. Comput. Phys. 100, 25-37
Van Leer, B., Lee, W.T., Roe, P. (1991): Characteristic time stepping or local
preconditioning of the Euler equations. AIAA paper 91-1552
Vinh, H., Dwyer, H.A., van Dam, C.P. (1992): Finite-difference algorithms for the
time-domain Maxwell equations - a numerical approach to RCS analysis. AIAA
Paper 92-2989
Wahlbin, L.B. (1974): A dissipative Galerkin method applied to some quasilinear
hyperbolic equations. RAIRO Anal. Numer. 8, 109-117
Washizu, K (1975): Variational Methods in Elasticity and Plasticity, 2nd ed. (Perg-
amon Press, Oxford)
Wathen, A.J. (1989): An analysis of some element-by-element techniques. Comput.
Methods Appl. Mech. Engrg. 74,271-287
Wendland, W.L. (1979): Elliptic Systems in the Plane (Pitman, London)
Williams, P.T., Baker, A.J. (1997): Numerical simulations of laminar flow over a
3D backward-facing step. Int. J. Num. Meth. Fluids 24, 1159-1183
Winget, J.M., Hughes, T.J.R. (1985): Solution algorithms for nonlinear tran-
sient heat conduction analysis employing element-by-element iterative strategies.
Compo Meth. Appl. Mech. Engrg. 52, 711-815
Winterscheidt, D., Surana, KS. (1993): P-version least squares finite element for-
mulation for two dimensional incompressible fluid flow. Int. J. Numer. Meth.
Engrg. 36, 3629-3646
Winterstein, R., Hafez, M. (1993): Euler solution for blunt bodies using triangular
meshes: artificial viscosity forms and numerical boundary conditions. AIAA 93-
3333-cp
Wong, S.H., Cendes, Z.J. (1988): Combined finite element-model solution of three-
dimensional eddy current problems. IEEE Trans. Magn., 24, 2586-2687
Wong, S.H., Cendes, Z.J. (1989): Numerically stable finite element methods for
Galerkin solution of eddy current problems. IEEE Trans. Magn. 25, 3019-3021
Wu, J., Jiang, B.N., Yu, S.T., Liu, N.S. (1994): Time-accurate LSFEM for fluid flows
and electromagnetic scattering problems. Least-Squares Finite Element Methods
Workshop, Oct. 13-14, Cleveland.
412 References
Wu, J., Jiang, B.N. (1995a): "A least-squares finite element method for electromag-
netic scattering problems", in Proceedings of the Third US National Congress on
Computational Mechanics, ed. by J.N. Reddy (Texas A & M University, College
Station) 295
Wu, J., Jiang, B.N. (1995b): A time-accurate least-squares finite-element method
for incompressible inviscid and viscous flows. ICDMP report (unpublished)
Wu, J., Jiang, B.N. (1996): A least-squares finite element method for electromag-
netic scattering problems. NASA Contract Report 202313, ICDMP-96-12
Wu, J., Yu, S.T., Jiang, B.N. (1996): Simulation of two-fluid flows by the least-
squares finite-element method using a continuum surface tension model. NASA
Contract Report 2023124, ICDMP-96-13
Vee, K.S. (1966): Numerical solution of initial boundary value problems involving
Maxwell's equations in isotropic media. IEEE Trans. Antennas and Popagat. 14,
302-307
Ying, L.A. (1988): Lecture Notes on the Finite Element Method (Chinese) (Beijing
University Press, Beijing)
Yu, S.T., Jiang, B.N., Liu, N.S., Wu, J. (1995a): The least-squares finite-element
method for low-Mach-number compressible viscous flows. Int. J. Num. Meth.
Engrg. 38, 3591-3610
Yu, S.T., Jiang, B.N., Wu, J., Jacqmin, D. (1995b): "A unified approach for sim-
ulating free-surface flows by the least-squares finite element method", in Proc.
Sixth Int. Symp. on Computational Fluid Dynamics, A Collection of Technical
Papers, Vol. III, 1467-1472
Yu, S.T., Jiang, B.N., Wu, J., Liu, N.S. (1996a): A div-curl-grad formulation for
compressible buoyant flows solved by the least-squares finite-element method.
Compo Meth. Appl. Mech. Engrg. 137, 59-88
Yu, S.T., Jiang, B.N., Wu, J. (1996b): A div-curl-grad formulation of three-
dimensional low-Mach-number flows solved by the least-squares finite-element
method. Compo Meth. Appl. Mech. Engrg. (submitted)
Yu, S.T., Jiang, B.N., Wu, J., Duh, J.C. (1996c): Three-dimensional simulations
of Marangoni-Benard convection in small containers by the least-squares finite
element method. AIAA paper 96-0863
Zhu, J., Sethian, J.A. (1992): Projection methods coupled to level set interface
techniques. J. Comput. Phys. 102, 128-138
Zienkiewicz, D.C., Dwen, D.RJ., Lee, K.N. (1974): Least squares finite element for
elasto-static problems - use of reduced integration. Int. J. Numer. Meth. Engrg.
8,341-358
Zienkiewicz, D.C. (1975): "Why finite elements", in Finite Elements in Fluids, Vol.
1, ed. by R H. Gallagher et al. (John Wiley & Sons) 1-23
Zienkiewicz, D.C., Morgan, K (1983): Finite Elements and Approximation (John
Wiley & Sons, Chichester)
Zienkiewicz, D.C., Zhu, J.Z. (1987): A simple error estimator and adaptive proce-
dure for practical engineering analysis. Int. J. Num. Meth. Engrg. 24, 337-357
Zienkiewicz, D.C., Taylor, RL. (1988): The Finite Element Method, 4th ed., Vol. 1
(McGraw-Hill, London)
Zienkiewicz, D.C., Taylor, RL. (1991): The Finite Element Method, 4th ed., Vol. 2
(McGraw-Hill, London)
Zienkiewicz, D.C., Zhu, J.Z. (1992): The superconvergent patch recovery and a
posteriori error estimates. Part 1: the recovery technique. Int. J. Num. Meth.
Engrg. 33, 1331-1364
Index