Spectral and Structural Analysis of High Precision ®nite Di Erence Matrices For Elliptic Operators
Spectral and Structural Analysis of High Precision ®nite Di Erence Matrices For Elliptic Operators
Spectral and Structural Analysis of High Precision ®nite Di Erence Matrices For Elliptic Operators
www.elsevier.com/locate/laa
Abstract
In this paper we study the structural properties of matrices coming from high-pre-
cision Finite Dierence (FD) formulae, when discretizing elliptic (or semielliptic) dif-
ferential operators L
a; u of the form
k
d dk
ÿk a
x u
x :
dxk dxk
Strong relationships with Toeplitz structures and Linear Positive Operators (LPO) are
highlighted. These results allow one to give a detailed analysis of the eigenvalues lo-
calisation/distribution of the arising matrices. The obtained spectral analysis is then
used to de®ne optimal Toeplitz preconditioners in a very compact and natural way and,
in addition, to prove Szeg o-like and Widom-like ergodic theorems for the spectra of the
related preconditioned matrices. A wide numerical experimentation, con®rming the
theoretical results, is also reported. Ó 1999 Elsevier Science Inc. All rights reserved.
AMS classi®cation: 65F10; 15A12; 15A18; 65N22
Keywords: Finite dierences; Elliptic operators; Toeplitz and Vandermonde matrices; Linear
positive operators; Ergodic theorems; Matrix algebras; Preconditioning
*
Corresponding author. Department of Energetica ``S. Stecco'', University of Florence, Via
Lombroso 6/17, 50134 Florence, Italy; e-mail: serra@mail.dm.unipi.it
1
E-mail: Cristina.Tablino.Possio@mater.unimi.it
0024-3795/99/$ ± see front matter Ó 1999 Elsevier Science Inc. All rights reserved.
PII: S 0 0 2 4 - 3 7 9 5 ( 9 9 ) 0 0 0 2 2 - 1
86 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
1. Introduction
This paper concerns the main properties of matrices coming from a large
class of Finite Dierence (FD) discretizations of template problems of the form
8 k k
>
<
ÿk d a
x d u
x f
x on X 0; 1;
dxk dxk
1
>
:
Dirichlet B:C: on oX:
We have focused our attention on 1D problems even though the same analysis
can be also carried out in a very similar way in the multidimensional case as
well (see Section 7).
Recently this matter was considered by Tilli [44] and Serra Capizzano [30],
by looking at the problem from two dierent points of view. The Tilli approach
is in the style of the Szeg o results [17] and mainly concentrates on the as-
ymptotic estimates and ergodic theorems regarding eigen/singular values of a
large class of matrices that the author calls ``Locally Toeplitz'' matrices, which
contain Toeplitz structures and FD matrices as proper subsets.
The Serra Capizzano approach concerns with an in depth analysis (not
only asymptotic, but for any ®xed dimension) of the localization of the
spectra of FD matrices, Toeplitz-based preconditioners and especially of the
related preconditioned matrices. In eect, the second approach gives infor-
mation on the ``relative distribution'' of preconditioned matrices, while the
®rst applies to the ``real distribution'' of the original matrices (FD matrices,
Toeplitz, etc.).
Due to applicative interest, here we follow and generalize this second point
of view by showing relationships with the theory of linear positive operators
(LPOs) [20]. In fact, such relationships are very welcome when considering
recent results [38] which indicate that many dierent problems from applied
and numerical mathematics can be treated in a uniform way when LPOs are
encountered.
Therefore, for any ®xed k, we look for high-precision FD formulae by
formalising the problem in terms of the solution of special Vandermonde-like
ÿ1
systems (see also [23]). In such a way, by setting the mesh size h
n 1 ,
the discrete approximation of the template problem (1) gives rise to an n n
linear system whose coecient matrix is constructed by using two sets of
values: the solution of a Vandermonde-like linear system and a proper set of
equispaced samplings of the function a
x. When a
x 1, the resulting
matrices enjoy the Toeplitz structure; otherwise these matrices are simply
banded.
In addition, when a
x > 0 and when the ``weak formulation'' of the
problem (1) is considered, it is well known that the resulting bilinear functional
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 87
is self-adjoint and coercive [8]. We would like to ®nd again these important
structural properties in the discrete approximation.
In the case of an abstract Faedo±Galerkin approach (Finite Elements, etc.)
the symmetry and the coerciveness are directly inherited from the resulting
®nite dimensional bilinear forms, so that the obtained matrices are symmetric
and positive de®nite.
In the FD case the situation is quite dierent. In general the matrices dis-
cretizing the template problem are not symmetric. Moreover, by imposing
symmetry, the structure arising in the case a
x 1 is symmetric and Toeplitz,
but its positive de®niteness is actually not guaranteed (we can exhibit FD
formulae approximating the operator ÿd2 =dx2 , which are related to Toeplitz
matrices with ``big'' negative eigenvalues (see Section 3.1)).
This fact is connected to an interesting representation theorem proved in
k
Section 3: if a symmetric FD formula approximates the operator
ÿ d2k =dx2k
with a precision order m P 1, then the resulting Toeplitz matrix sequence admits
a generating function g
x such that g
x ÿ x2k xm2k (for the notion of see
De®nition 3.1) in any small neighbourhood of x 0. Therefore g
x must be
nonnegative (not identically zero) in a neighbourhood of zero, but not neces-
sarily everywhere.
Since we are interested in computation and since we want to use the pow-
erful preconditioned conjugate gradient (PCG) methods or multigrid methods,
we would like to deal with symmetric and positive de®nite matrices discretizing
problems (1). More speci®cally, we follow a natural approach for the discret-
ization of the operator appearing in (1): since it can be looked at as the
composition of two derivatives of order k, we leave the operator in ``divergence
form'' and we discretize the inner and outer derivatives separately. In fact, if
the operator in (1) is represented as
Xk j
k k d d2kÿj
ÿ a
x u
x
j0
j dxj dx2kÿj
is linear and normally positive in the sense that it maps nonnegative (posi-
tive) functions in nonnegative (positive) de®nite matrices.
· For any a and nonnegative b belonging to C0; 1, we ®nd
a a
r inf 6 k
Pn
a; b 6 sup R;
Y b Y b
holds true, where g
x a
x=b
x over fx : b
x > 0g and g
x equals zero
where the function b
x vanishes. Notice that g
x is the functional counter-
part of the matrices fA n
b; m; kAn
a; m; kgn obtained by pseudo Moore±
Penrose inversion [24,26], since g
x can be also de®ned as b a where
b
x 1=b
x over the set where b
x is positive, and is zero where b
x
is zero (a sort of pseudo-inversion of a nonnegative function).
The ®rst two statements are important for establishing the exact rank of the
matrices An
a; m; k, which is especially meaningful when a
x has zeros. The
third statement not only suggests natural preconditioners, but is a useful tool
as well for establishing a precise upperbound for the spectral condition num-
bers (with respect to the Euclidean vector norm) of these matrices.
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 89
We want to stress again that this analysis is very similar to the one followed
in [10,29,31,33,34] in a Toeplitz context (compare Theorem 3.1 [10], Theorem
2.4 [29], Theorem 2.1 [33] and Theorem 3.1 [34]), so revealing itself to be an
important con®rmation of the unifying theory devised in [38].
Lastly, the fourth and the ®fth statements are concerned with the asymptotic
distribution of the eigenvalues of the preconditioned matrices.
More in detail, if a
x is strictly positive (i.e. the problem is strictly elliptic),
then a simple choice of the preconditioner is An
1; m; k. The related PCG
method is optimal (see Theorem 4.4) in the sense that the expected number of
iterations for reaching the solution within a preassigned accuracy is bounded
by a constant not depending on the dimension n: of course this is true if the
chosen accuracy does not depend on the dimension n.
Furthermore, the ergodic relation (2) is equivalent to write that the eigen-
values of Pn
a; b distribute as the above de®ned function g
x: this fact implies
that the number of iterations predicted by the Axelsson±Lindskog bounds [2] is
very tight and precise.
Finally, the matrix An
1; m; k is a Toeplitz one, whose generating function is
explicitly known and, in particular, is nonnegative with a zero at x 0.
Therefore very fast methods can be applied to these Toeplitz structures as:
· multigrid methods requiring O
n
2q ÿ 1 ops and O
log n parallel steps
with O
2nq processors in the parallel PRAM model [27] of computation
[12,13];
· a recursive displacement-rank based technique requiring O
n log
2q ÿ 1
log
n
2q ÿ 1 log2
2q ÿ 1 ops and O
log n parallel steps with O
2nq pro-
cessors [5],
where n is the matrix dimension and 2q ÿ 1 the matrix bandwidth.
Clearly, in order to obtain the total computational cost, the quoted costs
have to be multiplied by the number of iterations, which is constant with
respect to n, and added to the cost of few matrix-vector multiplications (recall
the PCG algorithm). The ®nal cost is of O
n log
2q ÿ 1 ops and
O
log n
2q ÿ 1 parallel steps with O
2nq processors in the PRAM model of
computation.
In conclusion, we have reduced the asymptotic cost of these band systems to
the cost of the band-Toeplitz systems for which the recent literature provides
very sophisticated algorithms [12,13,5].
These simple but rather powerful results can be strongly generalized with
regard to two main directions.
· When considering a 2 L1 , we have to modify the de®nition of a
~xi as a
``mean value'' in order to deal with this irregular case (compare the sugges-
tions given in [30,38,36]).
· When considering p-dimensional elliptic problems on square regions (see
Section 7 for a speci®c example), we have to carefully extend the analysis giv-
en in this paper through tensor arguments [41,36].
90 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
k k
is an approximation of
d =dx
u
xjxxr . This means that we have to choose
cj for which /k 1 and /t 0 for t 6 s ÿ 1, t 6 k with s P k 1. The precision
order of the FD formula associated with the coecient vector c 2 Rq is, at
least, m s ÿ k.
In such a way we obtain a sequence of linear systems of the form
Ds Vs c ek1 ; s k 1; . . . ; q;
3
5
Proof. Case k even and q 2m 1 odd. It is evident that the symmetric vector c
with cÿj cj ; j 1; . . . ; m satis®es all the even index equations of the system
(3) (see the matrix Vq in (4)). By imposing these conditions, the following re-
duced system is obtained
2 32 3
1 2 ... 2 c0
60 76 c 7
6 76 1 7
Dm1 6
6 ..
76 . 7 e k
76 . 7
21;
6
4. 2Xm 54 . 5
0 cm
where
Dm1 diag
0!ÿ1 ;
2!ÿ1 ; . . . ;
2m!ÿ1 ;
Xm Ym diag y1 ; . . . ; ym
and
2 3
1 ... 1
6 y1 ... ym 7
6 7
Ym 6
6 .. .. .. 7
7
4 . . . 5
y1mÿ1 . . . ymmÿ1
If f is real-valued, then we notice that the quoted de®nition implies that aÿk
ak so that all the matrices of the sequence fTn
f gn are Hermitian.
A deeper relation between the spectral structure of the class fTn
f gn and the
symbol f is well described in the following fundamental results.
Theorem 3.2 [17,45]. Let f and Tn
f stand as in Theorem 3.1. Let k
X be the
generic eigenvalue of the matrix X and let mf and Mf be the essential in®mum
and supremum of f, respectively, i.e. the inf f and sup f up to within zero
measure sets [28]. Then for any n 2 N the following cases occur:
· mf < k
Tn
f < Mf if mf < Mf or
· k
Tn
f M if mf Mf M.
Corollary 3.1 [47]. Let f and Tn
f stand as in Theorem 3.1. Let k
n
i be the
eigenvalues of Tn
f ordered in a nondecreasing way. Then, for any choice of
a < b, the following asymptotic formula
1 1
lim #fi : k
n
i 2
a; bg mfx 2 ÿp; p : f
x 2
a; bg
8
n!1 n 2p
holds true, provided that mfx 2 ÿp; p : f
x ag mfx 2 ÿp; p :
f
x bg 0, where mfg denotes the usual Lebesgue measure on the real
line.
De®nition 3.1. Let f and g be the two Lebesgue integrable functions. We write
f g over the interval I if f and g are nonnegative over I and there exist two
positive constants c1 and c2 so that c1 g 6 f 6 c2 g almost everywhere in I. In
addition, we write f g over the interval I if either f g or ÿ f g
or f ÿg or ÿ f ÿg over I.
Proof. Let us consider a vector c such that x2 ÿ pc
x x2m ; m P 1 holds true.
From a direct calculation of the Taylor expansion
! !
Xm
1 Xm
x2 ÿ pc
x ÿ c0 2 cj 1 cj j2 x2
j1
2 j1
! !
X m
ÿ X
i m
ÿ cj j2i x2i Rm
x;
i2
2i! j1
In the same way a similar statement can be proved for the discretization of
k
the operator
ÿ d2k =dx2k .
X
m
c0 2 cj cos
jx:
j1
In other words, the relation x2k ÿ pc
x x2km is a dierent way of writing
and looking at the Vandermonde-like linear systems considered in Section 2, so
that a more precise FD formula is simply a more precise local expansion of the
spectral function x2k in a neighbourhood of x 0.
Therefore, in the language of numerical analysis of dierential equations,
the condition pc
x x2k is the consistency condition, while the relation x2k ÿ
pc
x x2km establishes that a consistent FD formula has a precision order
equal to m.
Notice that Corollary 3.2 tells us that a condition number growing as n2k is
necessary for the consistency condition and therefore is inherent in every good
FD discretization scheme.
Finally, in the light of these remarks, we can say that the FD discretization is
a ``local approximation'' of the continuous operators in the sense that the
discrete eigenvalue function pc
x tends toward the continuous eigenvalue
function x2k . Here the convergence is not intended as a global convergence in a
given functional topology but it is a convergence in the sense of the asymptotic
Taylor expansion in a neighbourhood of x 0: this ``locality'' in the approx-
imation has a spectral interpretation that we will highlight in the next section.
Let us suppose that c 2 Rq1 and d 2 Rq2 are two coecient vectors associ-
ated with two FD formulae discretizing the operator dk =dxk with k P 1 and
precision order m1 and m2 , respectively. Note that we consider both q1 and q2
odd when k is even and vice versa. In order to calculate the quantity zr obtained
as a discrete approximation of
k
k d dk
ÿ a
x k u
x ;
dxk dx
xxr
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 99
we make use of the two quoted formulae, applying the former to obtain the
discretization of the inner operator and the latter to obtain the discretization of
the outer operator. More precisely, by calling v
x dk =dxk
u
x, we have
Xq1
cj
v
xr u
xrjÿs1 O
hm1
j1
h k
so that
!
ÿ X X
k q2 q1
zr 2k di a
xriÿs2 cj u
xrijÿs2 ÿs1 ;
10
h i1 j1
where
mi 1 if qi 2mi 1;
si
mi 1=2 if qi 2mi ;
De®nition 3.2. The symbol An
a; c; d; k denotes the matrix discretizing the
problem (1) through the FD formula coecient vector c for the inner derivative
and FD formula coecient vector d for the outer derivative and according to
Eq. (11) for the rth row, where the factor h2k is neglected because it is included
in the right hand side of the associated linear system.
When d equals c and
c; c is the maximal precision conjugated pair, then the
matrix An
a; c; c; k is denoted by the shorter symbol An
a; m; k with m bq=2c
(or simply An
a when it is clear from the context).
100 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
Let c 2 Rq1 , d 2 Rq2 and let Gn f~xi g be the set of the equispaced samplings
of the coecient function a
x with respect to the FD formula coecient vector
d and let a^i
x be the piecewise linear and continuous function so that
1 if x ~xi ;
a^i
x
0 otherwise on Gn :
Theorem 3.5. For every dimension n, the matrix An
a; c; d; k can be expressed as
X
a
~xi An
^
ai ; c; d; k;
i
and #Gn n q2 ÿ 1.
Proof. As a consequence of (11) the rth row of the matrix An
a; c; d; k is given
by
k T
ÿ 0; . . . ; 0;
ar d T ; 0; . . . ; 0;
where the number of the zero entries at the beginning of the row is r ÿ a, with a
absolute constant depending on the chosen FD formulae (with the convention
that negative values are assumed to be zero).
Due to the linearity of the componentwise product, we can consider directly
the matrices An
^ ai ; c; d; k.
Case q2 odd (k even): In this case
ar j a
xrj ; j ÿbq2 =2c; . . . ; bq2 =2c.
So, if a
x a^i
x, it is evident that at most the q2 rows given by
i ÿ bq2 =2c 6 r 6 i bq2 =2c
1 6 r 6 n are not identically zero and, on each
row, the unique nonvanishing entry is selected in a backward manner. More
precisely, the vector ar d is equal to dt et , with t i ÿ r dq2 =2e, so that
ar dT T 0; . . . ; 0; dt c1 ; . . . ; dt cq1 ; 0; . . . ; 0 where the number of beginning
zero entries is equal to t ÿ 1.
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 101
To sum up, the number of initial zero entries of the rth-row of the matrix
An
^
ai ; c; d; k is equal to r ÿ a t ÿ 1 and is always the same since for increasing
r values, decreasing t values are obtained. Therefore, we have
2 3
0 0 0 gi ÿ bq22 c ÿ 1 rows;
An
^ai ; c; d; k
ÿk 4 0 Di 0 5 gq2 rows;
0 0 0
where
2 3
dq2
6 7
Di 4 ... 5 c1 ; . . . ; cq1 2 Rq2 q1 ;
d1
and where the
1; 1 entry of the rectangular block Di is located at the position
i ÿ bq2 =2c; i bq2 =2c ÿ a 1 with respect to the matrix An
^
ai ; c; d; k. Notice
that in the case of a nonpositive position index i ÿ bq2 =2c and/or i bq2 =2c ÿ
a 1 we have to consider only the submatrix having positive entry indices with
regard to those of An
^ ai ; c; d; k.
Case q2 even (k odd): In this case,
ar j a
xrj1=2 , j ÿq2 =2; . . . ;
q2 =2 ÿ 1, so that, if a
x a^i
x, the at most q2 not identically zero rows are
given by i ÿ q2 =2 1 6 r 6 i q2 =2. In the same manner as in the q2 odd case,
we obtain
2 3
0 0 0 gi ÿ q22 rows;
ai ; c; d; k
ÿ 4 0 Di 0 5 gq2 rows;
k
An
^
0 0 0
where the
1; 1 entry of the rectangular block Di is located at the position
i ÿ q2 =2 1; i q2 =2 ÿ a 1 with respect to the matrix An
^
ai ; c; d; k.
is symmetric and nonnegative de®nite. Here the vector ci denotes a vector of Rn
containing in ``its middle'' the vector c suitably shifted in accordance with i.
Proof. By taking into account Theorem 3.5, the proof is trivial. In fact, since a
is equal to b
2q ÿ 1n2c 1, Di 2 Rqq is a diagonal block and since
k
rev
d
ÿ c, it directly follows that the dyad
ÿk Qn
c; rev
d; i
ÿ2k Qn
c; c; i
12
102 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
has rank one and is symmetric and nonnegative de®nite by construction since
for each y 2 Rn nf0g; y
y1 ; x; y2 we have
2
yT Qn
c; c; iy xT ccT x jjxT cjj2 P 0:
By means of Corollary 3.3, we can state the following theorem which pro-
vides a link between the number of the zeros of the nonnegative function a
x
k
and the rank of the matrix An
a; c;
ÿ rev
c; k.
Theorem 3.6. Let Gn f~xi g be the set of equispaced samplings of the nonneg-
ative coecient function a
x as considered in the representation Theorem 3.5 and
let In
a fi : a
~xi > 0g. Suppose that the n q ÿ 1 vectors fci : i 1; . . . ;
n q ÿ 1g considered in Corollary 3.3 strongly generate Rn in the sense that each
subset fcik : 1 6 i1 < i2 < < in 6 n q ÿ 1g is a basis for Rn . Then
rank
An
a; c;
ÿk rev
c; k minfn; #
In
ag:
Finally, in the case a
x 1, we can give a global description of the matrices
An
a; c; d; k and, in particular, a spectral characterization of the matrices
An
a; c;
ÿk rev
c; k.
where z eix and with the assumption that qi 2mi 1 (the case qi 2mi can
be treated in a very similar way). Therefore, An
1; c; d; k is a Toeplitz matrix
with generating function
mX
1 m2
Here we want to show that An
An
; m; k is a LPO for any ®xed n, m and
k, that is a linear operator mapping nonnegative functions a
x into nonneg-
ative de®nite n n matrices. Moreover, we prove that this operator is also
normally positive, in the sense that it maps strictly positive functions into
positive de®nite symmetric matrices.
Let us denote by L
X ; Y the class of the linear operators from the real
vector space X to the real vector space Y and let us denote by LP
X ; Y the
104 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
Corollary 4.1. For any n; k and m, the operator An
: C0; 1 ! S nn is linear
and positive. In addition, the operator is normally positive, i.e. if a
x is a strictly
positive function then An
a will be a positive de®nite matrix.
Proof. By invoking Theorem 4.1 when a
x is nonnegative, we ®nd that An
a is
a nonnegative linear combination of the nonnegative de®nite dyads Qn
c; c; i.
So, the nonnegative de®niteness of the matrix An
a directly follows.
To prove the normality property use has been made of the Rayleigh quotient
of the considered matrix An
a. For any x 2 Rn
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 105
X
xT An
ax a
~xi xT Qn
c; c; ix
i
X
P amin xT Qn
c; c; ix;
i
We stress that the statement of Corollary 4.1 is still valid and with un-
changed proof for all the matrices An
a; c; d; k with d
ÿk rev
c as a con-
sequence of Corollary 3.3 and Proposition 3.3. Moreover the assumption of
continuity of the symbol a can be easily removed.
Let ker
X be the null space of the matrix X, that is the space of the
eigenvectors of X associated with the zero eigenvalue.
for any x.
cÿ1 T T ÿ1 T
2 x Ax 6 x Bx 6 c1 x Ax;
We prove that the nonnegative functions a
x partition the matrices An
a
into equivalence classes in the following theorem. Even if this result is a direct
consequence of the representation Theorem 4.1, we make use in the proof of
the positivity of the operator An
which is, in this speci®c FD context, a
consequence of Theorem 4.1.
Theorem 4.2. Let a and b be two equivalent functions. Then fAn
agn and
fAn
bgn are equivalent
and from Corollary 4.1, it ensues that for any x 2 Rn the following inequality
holds
c1 xT An
bx 6 xT An
ax 6 c2 xT An
bx:
The following result is of noteworthy and practical interest, since it shows
that the functions a
x also partition the matrices An
a with regard to the
spectral condition numbers.
Theorem 4.3. Let a and b be two equivalent functions and let An
b be positive
de®nite. Then
· k2
An
a k2
An
b;
2
· if a
x > 0 then k2
An
a k2
Tn
jpc j ;
2k
· if a
x P 0 then k2
An
a P cn ; with c > 0 and for n large enough.
Proof. From Theorem 4.2 it follows that for any x 2 Rn the following in-
equalities hold
c1 xT An
bx 6 xT An
ax 6 c2 xT An
bx
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 107
and therefore the matrix An
a is positive de®nite as well. The rest of the proof
arises from Proposition 3.3 and Theorem 3.3, while the technique is the same as
the one introduced in Theorem 2.2 of [33] in a Toeplitz context.
We notice that the given de®nition coincides with the one suggested by
Axelsson and Lindskog [3] when each An is invertible. In general fBn gn are
optimal preconditioners for fAn gn if and only if the two sequences of matrices
are equivalent. More precisely, the preconditioner must work only in the
subspace where An is invertible. In the orthogonal subspace where An is the null
matrix the problem of ®nding a solution to the related linear system is ill-posed
and therefore it is convenient that the matrix Bn acts as the null matrix in the
same subspace. All these ideas and properties are summarized in the preceding
de®nition of equivalent sequence of matrices.
The following result gives the theoretical support for our preconditioning
techniques.
Theorem 4.4. Let a and b be two equivalent functions. Then, for any n, the
matrix
A
n
bAn
a
has the same null space as An
a and An
b while the other nonzero eigenvalues
belong to the closed set c1 ; c2 ; ci being the equivalence constants in De®nition
4.1. Therefore, according to De®nition 4.3 , we can say that the matrices fAn
bgn
are optimal preconditioners for the matrices fAn
agn .
Proof. From Theorem 4.2 we know that fAn
agn and fAn
bgn are equivalent,
so that Lemma 4.1 assures that the null spaces of the symmetric nonnegative
de®nite matrices An
a and An
b coincide and they coincide also with the null
spaces of A
n
a and An
b as a consequence of the classical Schur decompo-
sition. In addition, since ker
An
a fy 2 Rn : y An
ax; x 2 Rn g? , with a
contradiction argument, it is easily proved that they also coincide with the null
space K of the matrix A n
bAn
a.
108 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
and
uT An
bu uT2 An
bu2 6 0;
A
n
bAn
au ku:
1=2
First we multiply by
An
b and we obtain
1=2 1=2
A
n
b An
au k
An
b u:
The latter result can be generalized in the case where the symbols a and b are
not equivalent. However, since we are interested in the applicative problem of
the preconditioning, we stress that this extension has a theoretical interest only.
In fact, Theorem 4.4 tells us that fAn
bgn are optimal preconditioners for
fAn
agn if a and b are equivalent. Theorem 4.5 tells us that the property of
equivalence of a and b is not only sucient but also necessary.
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 109
Theorem 4.5. Let a and b be functions de®ned on 0; 1 with b taking nonneg-
ative values. Then all the eigenvalues of
A
n
bAn
a
belong to the set c1 ; c2 [ f0g, where c1 inf Y a
x=b
x; c2 supY a
x=b
x
with Y fx 2 0; 1 : ja
xj jb
xj > 0g.
Therefore, the only way to guarantee that the matrices fAn
bgn are optimal
preconditioners for the matrices fAn
agn is for a and b to be equivalent. In the
case where a is not equivalent to b we have c1 6 0 or c2 1 (see also Corollary
4.3).
In order to prove the density and the distribution results (in the spirit of the
Szego Theorem 3.1) we need a preparatory lemma regarding the asymptotic
inertia of the matrices An
a.
Notice that in Theorem 4.2 and Lemma 4.1, we linked the inertia of
the matrices fAn
agn and fAn
bgn when a and b belonged to the same
equivalence class. Here we are interested in linking the asymptotical in-
ertia of fAn
agn to the distribution of the sign of the functional coe-
cient a
x.
We will make use of the following de®nitions and notations.
De®nition 4.4. Fix s < t and a, b two continuous functions with b taking
nonnegative values. Then
N
s; t; a; n #fi : ki
An
a 2
s; tg;
N
s; t;
a; b; n #fi : ki
A
n
bAn
a 2
s; tg;
A
a fx 2 0; 1 : a
x > 0g;
Aÿ
a fx 2 0; 1 : a
x < 0g;
A0
a fx 2 0; 1 : a
x 0g:
Finally, by aÿ
a, a
a and a0
a we denote the Lebesgue measure mfg of the
sets Aÿ
a, A
a and A0
a, respectively.
110 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
hold true when [Hp1] the continuous function a
x has a ®nite number of sign
changes.
Proof. Let In
a fi : a
~xi > 0g, Inÿ
a fi : a
~xi < 0g and In0
a
fi : a
~xi 0g where Gn f~xi g is the set of the equispaced samplings of the
function a
x de®ned in Section 4.1 and where the number of ~xi lying outside
the interval 0; 1 needed by the FD formula is a constant with respect to n, i.e.
O
1.
Under the assumption of continuity of the functional coecient a
x, it
follows that:
#
In
a na
a O
1;
#
Inÿ
a naÿ
a O
1;
16
#
In0
a na0
a O
1:
Now, by taking into account the dyadic decomposition of the matrix An
a
given in Theorem 4.1, it is evident that each dyad Qn
c; c; i is made up of a
weighted sum of at most q2 terms of the form es eTt where ek is the kth element
of the canonical basis of Rn and s; t belong to a proper neighbourhood of the
index i; more precisely it holds that
( Pm
cs ct eis eTit if q 2m 1;
Qn
c; c; i Ps;tÿm
ÿ1 T
Pm T
s;tÿm cs ct eis1 eit1 s;t1 cs ct eis eit if q 2m;
where the de®nition of the set W
a is necessary to give evidence of canonical
vectors belonging to those adjacent dyads Qn
c; c; i for which a change in the
sign of the functional coecient a
x occurs.
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 111
and that
#W
a O
1;
#
I^
a #
I
a ÿ O
1 ;
n n
we ®nd that Rn is the direct sum of the four orthogonal subspaces S
a,
Sÿ
a, S0
a and W
a. Notice that the presence of boundary dyads (i.e.
those dyads whose representation requires less than q2 canonical vectors) have
no in¯uence on the considered partitioning of Rn .
It is crucial to observe that for any choice of the vectors x 2 S
a, y 2
S
a and z 2 S0
a with kxk2 kyk2 kzk2 1, we ®nd
ÿ
where
xT An
^
ax > 0
since, from Corollary 4.1, we know that the linear operator An
An
; m; k is
also normally positive. The other two inequalities follow in a similar manner.
Now, by virtue of the relation An
az 0 and by virtue of the third relation
displayed in (16), we ®nd that
dim
Ker
An
a P #
I^n0
a na0
a O
1:
we deduce that
0 dim
S
a \
Zÿ
a Z0
a P #
I^n
a
n ÿ N ÿ n
#
I^
a ÿ N ;
n
#
I^ÿ
a ÿ N ÿ ;
n
and ®nally
#
I^n
a 6 N 6 #
I^n
a dim
W
a;
#
I^ÿ
a 6 N ÿ 6 #
I^ÿ
a dim
W
a;
n n
By recalling that dim
W
a O
1 and the relations linking the sizes of the
sets I^n
a and In
a with 2 f; ÿ; 0g, the claimed thesis directly follows.
Remark. The assumption [Hp1] of the preceding lemma concerning the ®nite
number of sign changes of a
x can be expressed more explicitly: in fact it is
equivalent to the assumption that the set of the zeros of a
x is made up by a ®nite
collection of isolated zeros and by a ®nite collection of closed disjoint intervals.
hold true when the continuous function a
x ful®lls the following assumption.
[Hp2] there exists a ®nite number of points fs1 ; . . . ; sp g such that for any positive
the function a
x has a ®nite number of sign changes in the new de®nition set
X 0; 1n [pj1
sj ÿ ; sj .
Proof. It is enough to repeat the same proof as in Lemma 4.2 by discarding the
indices i so that ~xi belongs to [pj1
sj ÿ ; sj .
Theorem 4.6. Let a and b be two equivalent functions and let s 2 R so that
a ÿ zb ful®lls assumption [Hp2] for z 2 f0; sg. Let
s; 1 be the generic half-line
and let us denote by Fs the characteristic function of
s; 1. Then, by denoting by
k
n
i the eigenvalues of the matrices fPn
a; bgn , we ®nd that
1X n Z
lim Fs k
n
i Fs
a=b dx mfA0
bgFs
0;
18
n!1 n A
b
i1
where X is symmetric and strictly positive de®nite. Since ker
Pn
a; b
ker
An
b we can also write that
2 3
ÿs
6 .
Un
a; b ÿ sI U 6
.. 0 77U H
4 ÿs 5
0 Hs
So, it follows that Hs and X Hs X have the same inertia, by virtue of the Syl-
vester inertia law, owing to the invertibility and the symmetry of X. In other
words we can say that the matrices W
s and Un
a; b ÿ sI have the same in-
ertia in the subspace orthogonal to Z0
b, where Z0
b is the null space of
An
b. Therefore, recalling that Pn
a; b and Un
a; b have the same spectra, we
we ®nd that
N
s; 1;
a; b; n #fi : ki
Pn
a; b > sg
#fi : ki
Un
a; b ÿ sI > 0g
#fi : ki
Un
a; b 0gFs
0 #fi : ki
Hs > 0g
dim
ker
Pn
a; bFs
0 #fi : ki
X Hs X > 0g:
Now, through Lemmas 4.2 and 4.3 and by comparison with the splitting of
W
s, we have
N
s; 1;
a; b; n
n mfA0
bg O
erFs
0 #fi : ki
X Hs X > 0g
n mfA0
bg O
erFs
0 #fi : ki
W
s > 0g;
Notice that, when a ÿ sb has only a ®nite number of isolated zeros and F, a
and b are Lipschitz continuous, then the quoted result tells us that the error
1 X n Z
n
F ki ÿ F
g dx
n i1 A
b
Proof. Call S the space of the linear combinations of functions of the form Fs
with s 2 H . Observe that the topological closure with respect to the in®nity
norm of the space S contains all the continuous functions with bounded sup-
port. Now the general statement given in (20) is a consequence of Theorem 4.6
and of the linearity with respect to F of both sides of Eq. (20).
Theorem 4.8. Let a and b be two continuous functions with b taking nonnegative
values and a0
b mfA0
bg 0. Then the eigenvalues k
n i of the matrices
fPn
a; bgn satisfy the following ergodic relation, i.e., for any continuous function
F with bounded support, it holds
1X n Z
lim F k
n
i F
a=b dx mfA0
bgF
0:
21
n!1 n A
b
i1
Proof. By Proposition 5.2 in [44] (see also Theorem 3.3 in [37]), it follows that
2
the sequence fAn
agn is Locally Toeplitz with regard to the pair
a; jpc j .
Therefore by Theorem 3.6 in [44] (Eq. (43)) we deduce that fAn
agn distributes
2
as a
x jpc
yj i.e. for any continuous function F with bounded support, it
holds
Z
1X n
1 2
lim F k
n
i
A n
a F
a
xjpc
yj dx dy:
n!1 n 2p 0;1ÿp;p
i1
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 117
2
Moreover by the assumption and since jpc
yj is a nonzero polynomial
we have mf
x; y 2 0; 1 ÿp; p : b
xjpc
yj2 0g 0 (i.e. b
xjpc
yj2 is
``sparsely vanishing'' [11]).
Finally An
is a Linear Positive Operator for any n. Therefore the as-
sumptions of Theorem 2.6 in [35] are ful®lled and the claimed thesis fol-
lows.
Remark. Theorem 4.7 does not follow from Theorem 4.8 since there exist
equivalent continuous functions a and b with A0
b having positive measure so
that one of the assumptions of Theorem 4.8 is violated. Conversely, Theorem
4.8 does not follow from Theorem 4.7. In fact there exist continuous functions
b with a0
b mfA0
bg 0 such that the set A0
b does not ful®ll condition
[Hp2]: take b
x inf y 2 K jx ÿ yj with K being the classical Cantor set of 0; 1 so
that A0
b K.
Corollary 4.2. Let a and b be two continuous functions satisfying the assump-
tions of Theorem 4.7 or of Theorem 4.8. Then the topological closure Z of the S set
of the eigenvalues of all the matrices fPn
a; bgn is contained in c1 ; c2 f0g
and the set Z contains the usual range of the function a=b. Moreover, the
eigenvectors related to the null eigenvalue belong, for any dimension, to the null
space of An
a.
When the two functions are not equivalent then the matrices fAn
bgn are
not good preconditioners since a ``lot'' of eigenvalues (not related to null spaces
of fAn
agn ) of the preconditioned matrices cluster at zero and/or at in®nity
and this is at odds with De®nition 4.3.
a
x 0g is strictly contained in the set fx 2 0; 1 : b
x 0g and there exists
an interval J1 0; 1 so that b
x 0 over J1 and a
x is strictly positive on J1 .
0
By referring to the notations used in Lemma 4.2, we consider the set In;J 1
b
0 ^ 0 ^0 0
fj : b
~xj 0; ~xj 2 J1 g In
b and the set In;J1
b In
b \ In;J1
b conse-
quently de®ned. Therefore, since
I^n;J
0
1
0
b In;J 1
b; #
I^n;J
0
1
0
b #
In;J 1
b ÿ O
1;
we deduce that the subspace S0J1
b of the vectors of the canonical basis, whose
indices belong to I^n;J
0
1
b, is contained in the null space of A
n
b, while
zT An az > 0
with z 2 S0J1
b, kzk2 1. Now, in light of Lemma 4.2, the dimension of the
subspace S0J1
b is equal to n mfJ1 g O
1 and the theorem is proved.
Therefore, if the equivalence condition does not hold, then the precondi-
tioned matrix is characterized by a linear quantity of very small and/or very big
eigenvalues. This fact, in light of the Axelsson±Linskog analysis [3], leads to a
substantial deterioration of the performances of the related PCG method.
In the previous section we have proved that optimal preconditioners for the
matrices fAn
agn are matrices of the same kind fAn
bgn , provided that b is a
function equivalent to a (according to De®nition 4.1). The computational in-
terest of such a result crucially depends on the ``cheap solution'' property of a
generic linear system whose coecient matrix is An
b.
Therefore, in the case of a strictly elliptic problem, i.e. a
x strictly positive,
it is evident that a preconditioner as the matrix Dm;k An
1; m; k An
1
(b 1) is a good choice for two reasons:
· fDm;k gn are optimal preconditioners in the Axelsson±Lindskog sense [3],
· Dm;k is a band-Toeplitz matrix, for any size n.
Notice that the optimality of the preconditioner Dm;k is a direct consequence
of Theorem 4.4. Moreover, we appreciate the role of this preconditioner since
the original coecient matrix An
a has a condition number which grows as-
ymptotically at least as n2k (recall Theorem 4.3).
A further improvement of the preconditioner Dm;k is based on the use of the
main diagonal of An
a. More speci®cally, we de®ne the preconditioner matrix
P as D1=2 1=2
n
aDm;k Dn
a where Dn
a is the suitable scaled main diagonal of
An
a.
The exceptional clustering properties of the preconditioned matrices asso-
ciated with P have been analysed in [30] for some very special choices of the
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 119
value m. In [39] a general analysis of the eigenvalue clustering of the matrix set
fP ÿ1 An
agn is performed. In Section 8, a wide numerical experimentation is
reported con®rming the goodness and the eectiveness of this approach.
Finally, we want to consider the case where a
x is a nonnegative function
and has zeros. In this case the simple preconditioner Dm;k does not assure the
optimality of the related PCG method. In eect, by using Corollary 4.3, we can
state a negative result: more precisely, for any > 0, 9 n and there exists a
positive constant c for which, for any n P n , at least dc ne eigenvalues of
Dÿ1
m;k An
a belong to 0; . Therefore, we must expect that the performances of
the related PCG methods are strongly spoiled (see all the tables in Section 8
when a
x has zeros).
On the other hand, the performances of the PCG method based on the
improved preconditioner P are still good and an in depth theoretical expla-
nation of this phenomenon is presented in [39].
However, Theorem 4.4 also furnishes further indication for dealing with the
case of a
x with zeros: supposing that a
x has isolated zeros and b
x is
chosen so that
a
0 < c1 6 6 c2 ;
b
then the matrices fAn
bgn are still optimal preconditioners according to the
optimality de®nition considered by Axelsson and Lindskog [3].
For numerical evidence of the associated PCG method see the last Table 5 in
Section 8, where a
x has a complicated expression but is asymptotic to x and
therefore b
x x can be chosen.
However, in this context it is interesting to give numerical evidence of the
spectral results found in Section 4. In Fig. 1 the plot of the eigenvalues of
An
a and An
b with a
x x
1 x2 = exp
x and b
x x is reported. No-
tice, for instance, that the plot of the eigenvalues of An
b is not close to a
sampling of the function b
x x. Nevertheless, this is not a surprising result
since, in the light of the Tilli analysis [44], we know that the eigenvalues of
2
fAn
bgn distribute as b
x jpc
yj over the rectangle 0; 1 ÿp; p.
In Fig. 2 the plot of the eigenvalues of the preconditioned matrix Pn
a; b
Aÿ1
n
bAn
a is reported. Notice that, in light of the ergodic results and since b
vanishes in a negligible subset of 0; 1, this plot must distribute as a=b.
To give numerical evidence of this, the comparison between these eigen-
values and the values
a=b
xi on a uniform n-dimension grid fxi g of 0; 1 is
reported in Fig. 3 where the absolute error j
a=b
xi ÿ ki
Pn
a; bj is plotted
with respect to the dierent cases of k and m values. These numerical results
perfectly con®rm the theoretical expectations of Theorems 4.7 and 4.8 and
Corollary 4.2.
6. Operation counts
Fig. 3. Absolute error ja
x=b
x ÿ ki
Pn
a; bj vs x (a
x x
1 x2 = exp
x, b
x x, n 300).
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 121
Setting k 2, q 3, we ®nd
1 1 p
n1 O 1=2 ; n2 O 1=4 O
n1 :
122 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
Therefore, in the second case for 10ÿ8 instead of solving a system of order
10 000, we must solve a system of order 100: this dramatic reduction of the
algebraic problem size is obtained at the cost of a small growth of the band-
width of the involved systems.
However, the bandwidth of An
a; m 1; k equals the bandwidth of
An
a; m; k plus 4 and, in the considered example, the cost of each iteration is
strongly reduced since it moves from c 104 log
5 to c 102 log
9, where c
is a suitable positive constant.
7. Further generalizations
These simple, but rather powerful, results can be strongly generalized with
regard to two main directions.
· When considering a 2 L1 , we have to modify the de®nition of a
~xi as a
``mean value'' in order to deal with this irregular case (compare the sugges-
tions given in [30,38,36]).
· When considering p-dimensional elliptic problems on square regions [36,41],
the only delicate step is the careful use of tensor arguments.
Notice that in the case where a
x is continuous the standard and the new
de®nition of the quantity a
~xi do not coincide but are asymptotically equiv-
alent. In particular the following result holds true.
Notice that the dyadic decomposition Theorem is still valid and if the
function a
x is (essentially) nonnegative, owing to the use of the Lebesgue
integral in Eq. (22), the matrix An
a can still be looked at as a Linear
Positive Operator from the linear space of the Lebesgue integrable functions
(with the ``essential'' ordering) to the linear space of the real symmetric n n
matrices.
Since these two properties are the key ones in the theory discussed in the
previous sections, we can conclude that all the theorems developed in Sections
3 and 4 can be generalized to this case (for a similar extension refer to Section 6
in [36]).
Here we are little interested in a general and detailed analysis of the mul-
tidimensional case due to the related cumbersome notations. Therefore in or-
der to show how the proposed ideas can be extended and used in this context,
we focus our attention on a two dimensional problem (p 2) only. More
precisely, let us consider the weighted Bilaplacian problem
o2 o2 o2 o2
a
x; y 2 u
x; y 2 a
x; y 2 u
x; y f
x; y
ox2 ox oy oy
2
over X 0; 1 with the homogeneous boundary conditions. We discretize
both the operators
o2 o2
a
x; y 2 u
x; y ; z 2 fx; yg;
oz2 oz
· the eigenvalues of Dÿ1mx ;my ;k An
a; mx ; my ; k belong to the interval c1 ; c2 with
c1 infa and c2 supa.
Therefore we can conclude that all the theorems developed in Sections 3 and
4 can be generalized to this case (for a similar, but more detailed extension, see
[30] in the case k 1, and [41]).
Concerning the computational cost we point out that the displacement
rank technique developed in [5] is no longer ``optimal'' in the multilevel
Toeplitz case. The same remark also holds for the classical band-solvers [16].
Therefore, in the multilevel case the only known optimal iterative solvers are
those based on the multigrid methods [13,43] or on mixed methods
(PCG + multigrid) [29]: with the latter multigrid-type choices, the computa-
tional cost is linear as the size of the linear system and linear as the band-
widths of the coecient matrix.
8. Numerical experiments
Table 1
Number of PCG steps in the case of An
a; m; k with k 1; m 1 and n 300
a
x Dn
a D1;1 P
1x 300 11 3
300 11 3
exp
x 300 14 3
300 14 4
x 300 107 7
300 106 7
x2 300 433 4
300 433 4
x ÿ 0:52 150 212 4
150 430 7
jx ÿ 0:5j 0:5 150 11 4
300 11 4
jx ÿ 0:5j 150 70 8
150 102 11
p
1 x if x P 0 300 11 4
1 if x < 0 300 11 4
sin2
7x 1 300 12 8
300 12 8
exp
x if x 6 2=3 300 11 5
2 ÿ x if x > 2=3 300 11 6
x
1 x2 = exp
x 300 93 7
300 92 7
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 125
Table 2
Number of PCG steps in the case of An
a; m; k with k 1; m 2 and n 300
a
x Dn
a D2;1 P
1x 323 11 3
325 11 3
exp
x 323 14 3
324 14 4
x 326 107 7
326 107 7
x2 327 434 4
326 433 4
x ÿ 0:52 158 213 6
267 431 7
jx ÿ 0:5j 0:5 158 11 4
324 11 4
jx ÿ 0:5j 158 70 8
324 102 11
p
1 x if x P 0 343 12 4
1 if x < 0 345 11 4
sin2
7x 1 323 12 8
325 12 8
exp
x if x 6 2=3 343 11 5
2 ÿ x if x > 2=3 345 11 5
x
1 x2 = exp
x 326 93 7
326 93 7
function a
x, ranging from the ``easy'' case given by k 1 and a strictly
positive coecient a
x ex to the ``dicult'' one given by k 2 and non-
negative coecient functions with zeros and/or discontinuity in the function
and/or in their derivatives. Note that in the last case, this choice of the func-
tional coecient greatly increases the ill-conditioning of the resulting sequence
of matrices fAn
a; m; kgn .
We do not make explicit comparison to circulant preconditioners [7,18,22]
because, as shown in [14,15], s and band-Toeplitz based preconditioners per-
form much better than circulant ones in the case of Dirichlet boundary value
problems. It is also true that circulant preconditioners are very eective in the
case of y-periodic linear elliptic problems [22] with strictly positive smooth
coecient functions.
This complementary behaviour of the two techniques is not surprising: in
fact, while the discretized y-periodic linear elliptic problems have a ``circulant
pattern'', the discretized Dirichlet BVPs show a symmetric pattern which is
naturally closer to the s and symmetric Toeplitz structure.
We performed our tests at dierent grid sizes and we noticed no signi®cant
change since, as proved in the previous sections, our PCG method shows a
126 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
Table 3
Number of PCG steps in the case of An
a; m; k with k 1; m 3 and n 300
a
x Dn
a D3;1 P
1x 343 11 3
345 11 3
exp
x 343 15 3
345 15 4
x 346 107 7
346 107 7
x2 347 434 4
346 433 4
x ÿ 0:52 167 213 6
264 431 7
jx ÿ 0:5j 0:5 167 11 4
344 11 4
jx ÿ 0:5j 167 70 8
344 102 11
p
1 x if x P 0 343 12 4
1 if x < 0 345 11 4
sin2
7x 1 343 12 8
345 12 8
exp
x if x 6 2=3 343 11 6
2 ÿ x if x > 2=3 345 11 6
x
1 x2 = exp
x 346 93 7
346 93 7
Table 4
Number of PCG steps in the case of An
a; m; k with k 2; m 2 and n 300
a
x Dn
a D2;2 P
1x ÿ 13 4
ÿ 13 4
exp
x ÿ 17 4
ÿ 17 4
x ÿ 124 11
ÿ 124 11
x2 ÿ 468 11
ÿ 467 11
x ÿ 0:52 ÿ 227 21
ÿ 448 27
jx ÿ 0:5j 0:5 ÿ 14 7
ÿ 14 8
jx ÿ 0:5j ÿ 82 23
ÿ 109 27
p
1 x if x P 0 ÿ 14 6
1 if x < 0 ÿ 14 6
sin2
7x 1 ÿ 14 13
ÿ 14 13
exp
x if x 6 2=3 ÿ 14 15
2 ÿ x if x > 2=3 ÿ 14 15
x
1 x2 = exp
x ÿ 108 11
ÿ 108 11
accuracy e 10ÿ7 is reported both in the case
fi 1 for every i 1; . . . ; n and
f vector of random numbers uniformly distributed on 0; 1.
The same tests are reported in Table 5 with respect to the case of Pn
a; b
Aÿ1 2
n
b; m; kAn
a; m; k where a
x x
1 x = exp
x and b
x x. Lastly,
Table 6 is devoted to the multidimensional example of the weighted bilaplacian
problem reported in Section 7.2.
Table 5
Number of PCG steps in the case of preconditioning of An
a; m; k; a
x x
1 x2 = exp
x
with An
b; m; k; b
x x (n 300)
m; k An
x; m; k
1,1 8
7
2,1 8
7
3,1 8
7
2,2 9
9
128
Table 6
Number of PCG steps in the case of An
a; mx ; my ; k with k 2; mx my 2 (n 100; 400; 900)
a
x; y Dn
a D2;2 P Dn
a D2;2 P Dn
a D2;2 P
1xy 56 13 3 206 14 4 448 14 4
69 13 3 252 14 4 543 14 4
exp
x y 55 19 3 205 23 3 448 24 4
57 19 3 209 22 3 449 24 4
xy 58 23 5 212 35 5 462 44 5
76 24 5 291 35 5 625 43 5
x y2 58 41 5 214 96 6 470 149 6
76 47 5 293 104 6 629 162 6
x ÿ 0:52
y ÿ 0:52 15 15 7 67 47 9 146 76 10
63 40 9 234 87 11 509 14 8
jx ÿ 0:5j jy ÿ 0:5j 0:5 15 10 5 67 13 6 147 14 7
65 11 7 238 13 7 509 14 8
jx ÿ 0:5j jy ÿ 0:5j 15 14 7 67 25 9 146 33 11
65 21 9 243 33 12 525 39 14
p
1 x y if x y P 0 56 9 4 206 10 4 446 11 4
1 if x y < 0 70 9 4 264 10 4 575 11 4
sin2
7
x y 1 57 10 11 207 11 14 456 12 14
80 10 12 304 11 15 660 12 15
exp
x y if x y 6 2=3 57 23 8 213 35 10 469 43 12
2 ÿ x if x y > 2=3 83 24 8 307 36 12 658 43 14
x y
1
x y2 58 21 5 212 31 5 462 38 5
= exp
x y
79 22 5 292 31 5 622 37 5
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 129
9. Concluding remarks
References
[1] F. Avram, On bilinear forms on Gaussian random variables and Toeplitz matrices, Probab.
Theory Related Fields 79 (1988) 37±45.
[2] O. Axelsson, V. Barker, Finite Element Solution of Boundary Value Problems, Theory and
Computation, Academic press, New York, 1984.
[3] O. Axelsson, G. Lindskog, On the rate of convergence of the preconditioned conjugate
gradient method, Numer. Math. 48 (1986) 499±523.
[4] D. Bini, Matrix structure in parallel matrix computation, Calcolo 25 (1988) 37±51.
[5] D. Bini, B. Meini, Eective methods for solving banded Toeplitz systems, invited lecture:
SIAM annual meeting ± Stanford, CA, July, 1997, SIAM J. Matrix Anal. Appl., in press.
[6] A. Bottcher, S. Grudsky, On the condition numbers of large semi-de®nite Toeplitz matrices,
Linear Algebra Appl. 279 (1998) 285±301.
[7] R.H. Chan, T.F. Chan, Circulant preconditioner for elliptic problems, J. Numer. Linear
Algebra Appl. 1 (1992) 77±101.
[8] P. Ciarlet, The Finite Element Method for Elliptic Problems, North-Holland, Amsterdam,
1978.
[9] P. Davis, Circulant Matrices, Wiley, New York, 1979.
[10] F. Di Benedetto, G. Fiorentino, S. Serra, C.G. preconditioning for Toeplitz matrices, Comput.
Math. Appl. 25 (1993) 35±45.
[11] F. Di Benedetto, S. Serra Capizzano, A unifying approach to matrix algebra preconditioning,
Numer. Math., 82 (1) (1999) 57±90.
[12] G. Fiorentino, S. Serra, Multigrid methods for Toeplitz matrices, Calcolo 28 (1991) 283±305.
[13] G. Fiorentino, S. Serra, Multigrid methods for symmetric positive de®nite block Toeplitz
matrices with nonnegative generating functions, SIAM J. Sci. Comput. 17 (1996) 1068±1081.
[14] G. Fiorentino, S. Serra, Tau preconditioners for (high order) elliptic problems, in: Vassilevski
(Ed.), Proceedings of the Second IMACS conference on Iterative Methods in Linear Algebra,
Blagoevgrad (Bulgaria), June 1995, pp. 241±252.
[15] G. Fiorentino, S. Serra, Fast parallel solvers for elliptic problems, Comput. Math. Appl. 32
(1996) 61±68.
[16] G. Golub, C. Van Loan, Matrix Computations, Johns Hopkins University Press, Baltimore,
1983.
[17] U. Grenander, G. Szeg o, Toeplitz Forms and Their Applications, 2nd ed., Chelsea, New York,
1984.
130 S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131
[18] T. Huckle, Fast transforms for tridiagonal linear equations, BIT 34 (1994) 99±112.
[19] M. Kac, W. Murdoch, G. Szeg o, On the extreme eigenvalues of certain Hermitian forms,
J. Rat. Mech. Anal. 13 (1953) 767±800.
[20] P.P. Korovkin, Linear Operators and Approximation Theory (English translation), Hindustan
Publishing, Delhi, 1960.
[21] T.K. Ku, C.C.J. Kuo, On the spectrum of a family of preconditioned Toeplitz matrices, SIAM
J. Sci. Stat. Comput. 13 (1992) 948±966.
[22] I. Lirkov, S. Margenov, P. Vassilevsky, Circulant block factorization for elliptic problems,
Computing 53 (1994) 59±74.
[23] N. Macon, A. Spitzbart, Inverses of Vandermonde matrices, Amer. Math. Monthly. 65 (1958)
95±100.
[24] E.H. Moore, General Analysis. Part I. Amer. Phil. Soc., Philadelphia, 1935.
[25] S.V. Parter, On the distribution of singular values of Toeplitz matrices, Linear Algebra Appl.
80 (1986) 115±130.
[26] R. Penrose, A generalized inverse for matrices, Proc. Cambridge Phil. Soc. 51 (1955) 406±413.
[27] M.J. Quinn, Parallel Computing: Theory and Practice, McGraw-Hill, New York, 1994.
[28] W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1985.
[29] S. Serra, Preconditioning strategies for asymtotically ill-conditioned block Toeplitz systems,
BIT 34 (1994) 579±594.
[30] S. Serra, The rate of convergence of Toeplitz based PCG methods for second order nonlinear
boundary value problems, Numer. Math. 81 (3) (1999) 461±495.
[31] S. Serra, Preconditioning strategies for Hermitian Toeplitz systems with nonde®nite generating
functions, SIAM J. Matrix Anal. Appl. 17 (4) (1996) 1007±1019.
[32] S. Serra, The extension of the concept of generating function to a class of preconditioned
Toeplitz matrices, Linear Algebra Appl. 267 (1997) 139±161.
[33] S. Serra, On the extreme eigenvalues of Hermitian (block) Toeplitz matrices, Linear Algebra
Appl. 270 (1998) 109±129.
[34] S. Serra, Asymptotic results on the spectra of block Toeplitz preconditioned matrices, SIAM J.
Matrix Anal. Appl. 20 (1) (1998) 31±44.
[35] S. Serra Capizzano, An ergodic theorem for classes of preconditioned matrices, Linear
Algebra Appl. 282 (1998) 161±183.
[36] S. Serra Capizzano, Spectral behavior of matrix-sequences and discretized boundary value
problems, A preliminary version in TR nr. 31, LAN, Dept. of Mathematics, Univ. of Calabria,
1998.
[37] S. Serra Capizzano, C. Tablino Possio, Superlinear preconditioning for optimal precondi-
tioners of collocation linear systems, submitted.
[38] S. Serra Capizzano, Some theorems on linear positive operators and functionals and their
applications, TR nr. 26, LAN, Dept. of Mathematics, Univ. of Calabria, 1997.
[39] S. Serra Capizzano, C. Tablino Possio, High-precision Finite Dierence schemes and Toeplitz
based preconditioners for Elliptic Problems, submitted.
[40] S. Serra Capizzano, C. Tablino Possio, Spectral and structural analysis of high precision ®nite
dierences matrices discretizing elliptic operators, TR nr. 28, LAN, Dept. of Mathematics,
Univ. of Calabria, 1997.
[41] S. Serra Capizzano, C. Tablino Possio, Preconditioning strategies for 2D Finite Dierence
matrix sequences, manuscript, 1999.
[42] S. Serra Capizzano, Locally X matrices, spectral distributions, preconditioning and applica-
tions, TR nr. 29, LAN, Dept. of Mathematics, Univ. of Calabria, 1998, SIAM J. Matrix Anal.
Appl., to appear.
[43] H. Sun, X. Jin, Q. Chang, Multigrid scheme for ill-conditioned block Toeplitz linear systems,
TR nr. 12, Dept. of Mathematics, Chinese Univ. of Hong Kong, 1998.
S. Serra Capizzano, C. Tablino Possio / Linear Algebra Appl. 293 (1999) 85±131 131
[44] P. Tilli, Locally Toeplitz matrices: Spectral theory and applications, Linear Algebra Appl. 278
(1998) 91±120.
[45] E.E. Tyrtyshnikov, A unifying approach to some old and new theorems on distribution and
clustering, Linear Algebra Appl. 232 (1996) 1±43.
[46] E.E. Tyrtyshnikov, N.L. Zamarashkin, Spectra of multilevel Toeplitz matrices: Advanced
theory via simple matrix relationships, Linear Algebra Appl. 270 (1998) 15±27.
[47] H. Widom, Toeplitz matrices, in: I. Hirshman Jr. (Ed.), Studies in Real and Complex Analysis,
Math. Ass. Amer., 1965.
[48] H. Widom, On the singular values of Toeplitz matrices, Zeit. Anal. Anw. 8 (1989) 221±229.