Model Reduction For Large-Scale Linear Dynamical Systems: Abstract. The Optimal
Model Reduction For Large-Scale Linear Dynamical Systems: Abstract. The Optimal
c 2008 Society for Industrial and Applied Mathematics
Vol. 30, No. 2, pp. 609–638
Abstract. The optimal H2 model reduction problem is of great importance in the area of dy-
namical systems and simulation. In the literature, two independent frameworks have evolved focusing
either on solution of Lyapunov equations on the one hand or interpolation of transfer functions on the
other, without any apparent connection between the two approaches. In this paper, we develop a new
unifying framework for the optimal H2 approximation problem using best approximation properties
in the underlying Hilbert space. This new framework leads to a new set of local optimality condi-
tions taking the form of a structured orthogonality condition. We show that the existing Lyapunov-
and interpolation-based conditions are each equivalent to our conditions and so are equivalent to
each other. Also, we provide a new elementary proof of the interpolation-based condition that clar-
ifies the importance of the mirror images of the reduced system poles. Based on the interpolation
framework, we describe an iteratively corrected rational Krylov algorithm for H2 model reduction.
The formulation is based on finding a reduced order model that satisfies interpolation-based first-
order necessary conditions for H2 optimality and results in a method that is numerically effective
and suited for large-scale problems. We illustrate the performance of the method with a variety of
numerical experiments and comparisons with existing methods.
AMS subject classifications. 34C20, 41A05, 49K15, 49M05, 93A15, 93C05, 93C15
DOI. 10.1137/060666123
The work of these authors was supported in part by the NSF through grants DMS-050597 and
DMS-0513542, and by the AFOSR through grant FA9550-05-1-0449.
‡ Department of Electrical and Computer Engineering, Rice University, Houston, TX (aca@ece.
rice.edu). The work of this author was supported in part by the NSF through grants CCR-0306503
and ACI-0325081.
609
multiple input/multiple output (MIMO) case in section 3.2.1, but will confine our
analysis and examples to the SISO case.)
G(s) is the transfer function of the system: if u (s) and y(s) denote the Laplace
transforms of the input and output u(t) and y(t), respectively, then y(s) = G(s) u(s).
With a standard abuse of notation, we will denote both the system and its transfer
function by G. The “dimension of G” is taken to be the dimension of the underlying
state space, dim G = n in this case. It will always be assumed that the system, G, is
stable, that is, that the eigenvalues of A have strictly negative real parts.
The model reduction process will yield another system,
ẋr (t) = Ar xr (t) + br u(t)
(1.2) Gr : or Gr (s) = cTr (sI − Ar )−1 br ,
yr (t) = cTr xr (t)
2π −∞
We
seek a reduced order dynamical system, Gr , such that
G − Gr H2 , the “H2 error,” is as small as possible;
(i)
(ii)
critical system properties for G (such as stability) exist also in Gr ; and
the computation of Gr (i.e., the computation of Ar , br , and cr ) is both
(iii)
efficient and numerically stable.
The problem of finding reduced order models that yield a small H2 error has
been the object of many investigations; see, for instance, [6, 37, 34, 9, 21, 26, 22, 36,
25, 13] and the references therein. Finding a global minimizer of G − Gr H2 is a
hard task, so the goal in making G − Gr H2 “as small as possible” becomes, as for
many optimization problems, identification of reduced order models, Gr , that satisfy
first-order necessary conditions for local optimality. There is a wide variety of such
conditions that may be derived, yet their interconnections are generally unclear. Most
methods that can identify reduced order models satisfying such first-order necessary
conditions will require dense matrix operations, typically the solution of a sequence
of matrix Lyapunov equations, a task which becomes computationally intractable
rapidly as the dimension increases. Such methods are unsuitable even for medium-
scale problems. In section 2, we review the moment matching problem for model
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
reduction, its connection with rational Krylov methods (which are very useful for
large-scale problems), and basic features of the H2 norm and inner product.
We offer in section 3 what appears to be a new set of first-order necessary condi-
tions for local optimality of a reduced order model comprising in effect a structured
orthogonality condition. We also show its equivalence with two other H2 optimality
conditions that have been previously known (thus showing them all to be equivalent).
An iterative algorithm that is designed to force optimality with respect to a set of
conditions that is computationally tractable is described in section 4. The proposed
method also forces optimality with respect to the other equivalent conditions as well.
It is based on computationally effective use of rational Krylov subspaces and so is
suitable for systems whose dimension n is of the order of many thousands of state
variables. Numerical examples are presented in section 5.
2. Background.
2.1. Model reduction by moment matching. Given the system (1.1), reduc-
tion by moment matching consists in finding a system (1.2) so that Gr (s) interpolates
the values of G(s), and perhaps also derivative values as well, at selected interpolation
points (also called shifts) σk in the complex plane. For our purposes, simple Hermite
interpolation suffices, so our problem is to find Ar , br , and cr so that
or, equivalently,
cT (σk I − A)−1 b = cTr (σk I − Ar )−1 br and cT (σk I − A)−2 b = cTr (σk I − Ar )−2 br
(2.2) Ar = (WrT Vr )−1 WrT AVr , br = (WrT Vr )−1 WrT b, and cTr = cT Vr .
Evidently the choice of Vr and Wr determines the quality of the reduced order model.
Rational interpolation by projection was first proposed by Skelton et al. in [11, 38,
39]. Grimme [17] showed how one can obtain the required projection using the rational
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
Krylov method of Ruhe [33]. Krylov-based methods are able to match moments
without ever computing them explicitly. This is important since the computation of
moments is in general ill-conditioned. This is a fundamental motivation behind the
Krylov-based methods [12].
In Lemma 2.1 and Corollary 2.2 below, we present new short proofs of rational
interpolation by Krylov projection that are substantially simpler than those found in
the original works [17, 11, 38, 39].
Lemma 2.1. Suppose σ ∈ C is not an eigenvalue of either A or Ar .
Proof. Define Nr (z) = Vr (zI − Ar )−1 (WrT Vr )−1 WrT (zI − A) and N r (z) =
−1
(zI−A)Nr (z)(zI−A) . Both Nr (z) and Nr (z) are analytic matrix-valued functions
in a neighborhood of z = σ. One may directly verify that N2r (z) = Nr (z) and
N r (z) and that Vr = Ran Nr (z) = Ker (I − Nr (z)) and W ⊥ = Ker N
2 (z) = N r (z) =
r
r
Ran I − N r (z) for all z in a neighborhood of σ. Then
T
G(z) − Gr (z) = (zI − AT )−1 c r (z) (zI − A) I − Nr (z) (zI − A)−1 b.
I−N
G(σ + ε) − Gr (σ + ε) = O(ε2 ),
Then Vr and Wr can be chosen to be real matrices and the reduced order system Gr
defined by Ar = (WrT Vr )−1 WrT AVr , br = (WrT Vr )−1 WrT b, cTr = cT Vr is itself real
and matches the first two moments of G(s) at each of the interpolation points σk , i.e.,
G(σk ) = Gr (σk ) and G (σk ) = Gr (σk ) for k = 1, . . . , r.
For Krylov-based model reduction, one chooses interpolation points and then con-
structs Vr and Wr satisfying (2.6) and (2.7), respectively. Note that, in a numerical
implementation, one does not actually compute (σi I − A)−1 , but instead computes
a (potentially sparse) factorization (one for each interpolation point σi ), uses it to
solve a system of equations having b as a right-hand side, and uses its transpose to
solve a system of equations having c as a right-hand side. The interpolation points
are chosen so as to minimize the deviation of Gr from G in a sense that is detailed in
the next section. Unlike Gramian-based model reduction methods such as balanced
truncation (see section 2.2 below and [2]), Krylov-based model reduction requires only
matrix-vector multiplications and some sparse linear solvers, and can be iteratively
implemented; hence it is computationally effective; for details, see also [15, 16].
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
P and Q are called the reachability and observability Gramians, respectively. Under
the assumption that A is stable, both P and Q are positive semidefinite matrices.
Square roots of the eigenvalues of the product PQ are the singular values of the
Hankel operator associated with G(s) and are called the Hankel singular values of
G(s), denoted by ηi (G).
Let P = UUT and Q = LLT . Let UT L = ZSYT be the singular value decompo-
sition with S = diag(η1 , η2 , . . . , ηn ). Let Sr = diag(η1 , η2 , . . . , ηr ), r < n. Construct
where Zr and Yr denote the leading r columns of left singular vectors, Z, and right
singular vectors, Y, respectively. The rth-order reduced order model via balanced
truncation, Gr (s), is obtained by reducing G(s) using Wr and Vr from (2.9).
Another important dynamical systems norm (besides the H2 norm) is the H∞
norm defined as GH∞ := supω∈R |G(ıω)|. The reduced order system Gr (s) obtained
by balanced truncation is asymptotically stable and the H∞ norm of the error system
satisfies G − Gr H∞ ≤ 2(ηr+1 + · · · + ηn ).
The value of having, for reduced order models, guaranteed stability and an explicit
error bound is widely recognized, though it is achieved at potentially considerable
cost. As described above, balanced truncation requires the solution of two Lyapunov
equations of order n, which is a formidable task in large-scale settings. For more
details and background on balanced truncation, see section III.7 of [2].
2.3. The H2 norm. H2 will denote the set of functions, g(z), that are analytic
for z in the open right half plane, Re(z) > 0, and such that for each fixed Re(z) =
x > 0, g(x + ıy) is square integrable as a function of y ∈ (−∞, ∞) in such a way that
∞
sup |g(x + ıy)|2 dy < ∞.
x>0 −∞
H2 is a Hilbert space and holds our interest because transfer functions associated with
stable SISO finite-dimensional dynamical systems are elements of H2 . Indeed, if G(s)
and H(s) are transfer functions associated with real stable SISO dynamical systems,
then the H2 inner product can be defined as
∞ ∞
def 1 1
(2.10)
G, HH2 = G(ıω)H(ıω) dω = G(−ıω)H(ıω) dω,
2π −∞ 2π −∞
with a norm defined as
+∞ 1/2
1
GH2 = |G(ıω)|2 dω
def
(2.11) .
2π −∞
Notice in particular that if G(s) and H(s) represent real dynamical systems, then
G, HH2 =
H, GH2 and
G, HH2 must be real.
There are two alternate characterizations of this inner product that make it far
more computationally accessible.
Lemma 2.3. Suppose A ∈ Rn×n and B ∈ Rm×m are stable and, given b, c ∈ Rn
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
and b, c ∈ Rm , define associated transfer functions,
−1 −1
G(s) = cT (sI − A) b and H(s) =
cT (sI − B) b.
(2.12) If P solves T = 0,
AP + PBT + bb then
G, HH2 = cT P
c.
(2.13) If Q solves QA + BT Q +
ccT = 0, then T Qb.
G, HH2 = b
(2.14) If R solves cT = 0,
AR + RB + b then
G, HH2 = cT Rb.
Gramians play a prominent role in the analysis of linear dynamical systems; refer
to [2] for more information.
Proof. We detail the proof of (2.12); proofs of (2.13) and (2.14) are similar. Since
A and B are stable, the solution, P, to the Sylvester equation of (2.12) exists and is
unique. For any ω ∈ R, rearrange this equation to obtain in sequence
(−ıωI − A) P + P ıωI − BT − bb T = 0,
−1
−1 −1 T
−1
(−ıωI − A) P + P ıωI − BT = (−ıωI − A) bb ıωI − BT ,
−1
−1
cT (−ıωI − A) P c + cT P ıωI − BT
c = G(−ıω)H(ıω),
and finally
L
−1
L
T −1
c T
(−ıωI − A) dω P
c+c P T
ıωI − B dω
c
−L −L
L
= G(−ıω)H(ıω) dω.
−L
Recently, Antoulas [2] obtained a new expression for GH2 based on the poles
and residues of the transfer function G(s) that complements the widely known alter-
native expression (2.15). We provide a compact derivation of this expression and the
associated H2 inner product.
In particular,
• if μk is a simple pole of H(s), then
res[G(−s)H(s), μk ] = G(−μk )res[H(s), μk ];
• if μk is a double pole of H(s), then
res[G(−s)H(s), μk ] = G(−μk ) res[H(s), μk ] − G (−μk ) · h0 (μk ),
where h0 (μk ) = lims→μk (s − μk )2 H(s) .
Proof. Notice that the function G(−s)H(s) has singularities at μ1 , μ2 , . . . , μm
and −λ1 , −λ2 , . . . , −λn . For any R > 0, define the semicircular contour in the left
half plane:
π 3π
ΓR = {z |z = ıω with ω ∈ [−R, R] } ∪ z z = R e with θ ∈
ıθ
, .
2 2
ΓR bounds a region that for sufficiently large R contains all the system poles of H(s)
and so, by the residue theorem,
+∞
1
G, HH2 = G(−ıω)H(ıω) dω
2π −∞
m
1
= lim G(−s)H(s) ds = res[G(−s)H(s), μk ].
R→∞ 2πı Γ
R k=1
Evidently, if μk is a simple pole for H(s), it is also a simple pole for G(−s)H(s) and
res[G(−s)H(s), μk ] = lim (s − μk )G(−s)H(s) = G(−μk ) lim (s − μk )H(s).
s→μk s→μk
If μk is a double pole for H(s), then it is also a double pole for G(−s)H(s) and
d
res[G(−s)H(s), μk ] = lim (s − μk )2 G(−s)H(s)
s→μk ds
d
= lim G(−s) (s − μk )2 H(s) − G (−s)(s − μk )2 H(s)
s→μk ds
d
= G(−μk ) lim (s − μk )2 H(s) − G (−μk ) lim (s − μk )2 H(s).
s→μk ds s→μk
Lemma 2.4 immediately yields the expression for GH2 given by Antoulas [2,
p. 145] based on poles and residues of the transfer function G(s).
Corollary 2.5. If G(s) has simple poles at λ1 , λ2 , . . . , λn , then
n 1/2
GH2 = res[G(s), λk ]G(−λk ) .
k=1
Many researchers have worked on problem (3.1), the optimal H2 model reduction
problem. See [37, 34, 9, 21, 26, 22, 36, 25] and the references therein.
3.1. Structured orthogonality optimality conditions. The set of all stable
rth-order dynamical systems do not constitute a subspace of H2 , so the best rth-
order H2 approximation is not so easy to characterize, the Hilbert space structure
of H2 notwithstanding. This observation does suggest the following narrower though
simpler result.
Theorem 3.1. Let μ1 , μ2 , . . . , μr ⊂ C be distinct points in the open left half
plane and define M(μ) to be the set of all proper rational functions that have simple
poles exactly at μ1 , μ2 , . . . , μr . Then
• H ∈ M(μ) implies that H is the transfer function of a stable dynamical
system with dim(H) = r;
• M(μ) is an (r − 1)-dimensional subspace of H2 ;
• Gr ∈ M(μ) solves
if and only if
(3.3)
G − Gr , HH2 = 0 for all H ∈ M(μ).
multiple minimizers. Indeed there may be “local minimizers” that do not solve (3.1).
A reduced order system, Gr , is a local minimizer for (3.1) if, for all ε > 0 sufficiently
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
small,
(3.4) (ε) H
G − Gr H2 ≤ G − G r 2
with C being a constant that may depend on the particular family G r(ε) considered.
As a practical matter, the global minimizers that solve (3.1) are difficult to obtain
with certainty; current approaches favor seeking reduced order models that satisfy a
local (first-order) necessary condition for optimality. Even though such strategies do
not guarantee global minimizers, they often produce effective reduced order models
nonetheless. In this spirit, we give necessary conditions for optimality for the reduced
order system, Gr , that appear as structured orthogonality conditions similar to (3.3).
Theorem 3.2. If Gr is a local minimizer to G as described in (3.4) and Gr has
simple poles, then
(3.5)
G − Gr , Gr · H1 + H2 H2 = 0
for all real dynamical systems H1 and H2 having the same poles with the same mul-
tiplicities as Gr .
(Gr · H1 here denotes pointwise multiplication of scalar functions.)
Proof. Theorem 3.1 implies (3.5) with H1 = 0, so it suffices to show that the
hypotheses imply that
G − Gr , Gr · HH2 = 0 for all real dynamical systems H
having the same poles with the same multiplicities as Gr .
Suppose that {G r(ε) }ε>0 is a family of real stable dynamical systems with
dim(G r(ε) ) = r and Gr − G r(ε) H < Cε for some constant C > 0. Then for all
2
ε > 0 sufficiently small,
(ε) 2
G − Gr 2H2 ≤ G − G r H2
(ε) )2
≤ (G − Gr ) + (Gr − G r H2
(ε)
≤ G − Gr H2 + 2 G − Gr , Gr − G
2 (ε) )2 .
+ Gr − G
r r H2
H2
with arbitrary real-valued choices for γi , ρi , and τi . Now suppose that μ is a real pole
for Gr and that
Gr (s)
(3.7) G − Gr , = 0.
s − μ H2
pr−1 (s)
Write Gr (s) = (s−μ) qr−1 (s) for real polynomials pr−1 , qr−1 ∈ Pr−1 and define
pr−1 (s)
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
r(ε) (s) =
G ,
[s − μ − (±ε)] qr−1 (s)
where the sign of ±ε is chosen to match that of G − Gr , Gs−μr (s)
H
. Then we have
2
+
G − Gr , H2 (s)H2 = 0.
Theorem 3.2 describes new necessary conditions for the H2 approximation prob-
lem as structured orthogonality conditions. This new formulation amounts to a unify-
ing framework for the optimal H2 problem. Indeed, as we show in sections 3.2 and 3.3,
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
two other known optimality frameworks, namely, interpolatory- [26] and Lyapunov-
based conditions [36, 22], can be directly obtained from our new conditions by using
an appropriate form for the H2 inner product. The interpolatory framework uses the
residue formulation of the H2 inner product as in (2.16); the Lyapunov framework
uses the Sylvester equation formulation of the H2 norm as in (2.12).
3.2. Interpolation-based optimality conditions. Corollary 2.5 immediately
yields an observation regarding the H2 norm of the error system, which serves as a
main motivation for the interpolation framework of the optimal H2 problem.
Proposition 3.3. Given the full-order model G(s) and a reduced order model
Gr (s), let λi and λ i be the poles of G(s) and Gr (s), respectively, and suppose that
the poles of Gr (s) are distinct. Let φi and φ j denote the residues of the transfer
functions G(s) and Gr (s) at their poles λi and λ i , respectively: φi = res[G(s), λi ] for
i = 1, . . . , n and φj = res[Gr (s), λj ] for j = 1, . . . , r. The H2 norm of the error system
is given by
2
n
G − Gr H2 = res[(G(−s) − Gr (−s)) (G(s) − Gr (s)) , λi ]
i=1
r
+ j ]
res[(G(−s) − Gr (−s)) (G(s) − Gr (s)) , λ
j=1
n r
(3.10) = φi G(−λi ) − Gr (−λi ) − j ) − Gr (−λ
φ j G(−λ j ) .
i=1 j=1
The H2 error expression (3.10) is valid for any reduced order model regardless
of the underlying reduction technique and generalizes a result of [20, 18] to the most
general setting.
Proposition 3.3 has the system-theoretic interpretation that the H2 error is due
to mismatch of the transfer functions G(s) and Gr (s) at mirror images of the full-
order poles λi and reduced order poles λ i . This expression reveals that for good H2
performance, Gr (s) should approximate G(s) well at −λi and −λ j . Note that λ i is
not known a priori. Therefore, to minimize the H2 error, Gugercin and Antoulas
[20] proposed choosing σi = −λi (A), where λi (A) are those system poles having big
residuals φi . They have illustrated that this selection of interpolation points works
quite well; see [18, 20]. However, as (3.10) illustrates, there is a second part of the H2
error due to the mismatch at −λ j . Indeed, as we will show below, interpolation at
−λi is more important for model reduction and is a necessary condition for optimal
H2 model reduction; i.e., σi = −λ i is the optimal shift selection.
Theorem 3.4. Given a stable SISO system G(s) = cT (sI − A)−1 b, let Gr (s) =
cr (sI − Ar )−1 br be a local minimizer of dimension r for the optimal H2 model reduc-
T
tion problem (3.1) and suppose that Gr (s) has simple poles at λ i , i = 1, . . . , r. Then
Gr (s) interpolates both G(s) and its first derivative at −λi , i = 1, . . . , r:
(3.11) i ) = G(−λ
Gr (−λ i ) and i ) = G (−λ
Gr (−λ i ) for i = 1, . . . , r.
Proof. From (3.5), consider first the case H1 = 0 and H2 is an arbitrary transfer
i , i = 1, . . . , r. Denote φ i = res[H2 (s), λ
function with simple poles at λ i ]. Then (2.16)
leads to
r
G − Gr , H2 H2 = i ]
res[(G(−s) − Gr (−s)) H2 (s), λ
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
i=1
r
= i ) − Gr (−λ
φ i G(−λ i ) = 0.
i=1
Since this is true for arbitrary choices of φ i , we have G(−λ i ) = Gr (−λ i ). Now
consider the case H2 = 0 and H1 is an arbitrary transfer function with simple poles
i , i = 1, . . . , r. Then Gr (s)H1 (s) has double poles at λ
at λ i , i = 1, . . . , r, and since
G(−λ i ) = Gr (−λ i ) we have
r
G − Gr , Gr · H1 H2 = i ]
res[(G(−s) − Gr (−s)) Gr (s)H1 (s), λ
i=1
r
=− i ] G (−λ
φ i res[Gr , λ i ) − G (−λ
r
i ) = 0,
i=1
where “tr(M)” denotes the trace of the matrix M. The H2 norm is then
+∞ 1/2
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
1
GH2 = G(ıω)2F
def
(3.14) dω ,
2π −∞
def
2 1/2
where FF = ij |Fij | denotes the usual Frobenius matrix norm. As before,
if G(s) and H(s) represent real dynamical systems, then
G, HH2 =
H, GH2 and
G, HH2 is real.
Necessary conditions for H2 optimality built on structured orthogonality parallel-
ing the results of section 3.1 can be derived in this setting as well. In particular, the
residue form for the inner product is a straightforward analogue of Lemma 2.4 and
leads naturally to interpolation conditions. If F(s) is a matrix-valued meromorphic
function with a pole at λ, then F(s) has a Laurent expansion (with matrix coeffi-
cients), and its residue, res[F(s), λ], will be the coefficient matrix associated with the
expansion term (s − λ)−1 . For example, suppose that F(s) has the realization F(s) =
sI − A
C −1 B. If λ is a simple pole of F(s), then we can assume that λ is a simple
eigenvalue of A associated with a rank-1 spectral projector Eλ and then F(s) =
1
s−λ λE +D(s), where D(s) is analytic at s = λ, and res[F(s), λ] = lims→λ (s−λ)F(s) =
λ B.
CE If λ is a double pole, then we can assume that λ is a double eigenvalue of
A associated with a rank-2 spectral projector Eλ and a rank-1 nilpotent matrix Nλ
such that AE λ = λEλ + Nλ . Then F(s) = 1 2 Nλ + 1 Eλ + D(s), where D(s)
(s−λ)
(s−λ)2
is analytic at s = λ, and so res[F(s), λ] = lims→λ ds d
(s − λ) F(s) = CE λ B.
Lemma 3.5. Suppose that G(s) has poles at λ1 , λ2 , . . . , λn and H(s) has poles
at λ 1 , λ
2 , . . . , λ
ñ , with both sets contained in the open left half plane. Then
ñ
(3.15)
G, HH2 = k ] .
tr res[G(−s)HT (s), λ
k=1
A k x
xk = λ k , y k y
=λ
k∗ A k∗ , and k∗ x
y k = 1,
then
k ] =
tr res[G(−s)HT (s), λ k )b
cTk G(−λ k ,
where b T = y and
k∗ B xk .
ck = C
k
• If λk is a double pole of H(s), and λ k is associated with left and right eigen-
k and x
vectors y
k of A, and generalized eigenvectors,
zk and w k , respectively,
A k x
xk =λ k , A w
k = λ k w k , y
k + x =λ
k∗ A k y k∗ ,
z∗k A k
=λ z∗k + y
k∗ ,
and y k∗ x z∗k w
k = 0, k = 0, and z∗k x k∗ w
k = y k = 1,
then
k ] = d
tr res[G(−s)HT (s), λ T G(−λ
k )b
k + k )
cTk G(−λ ek − k )b
cTk G (−λ k ,
k
k and
where b ck are as above and z∗k B
eTk = k = C
and d w k.
(3.16)
cT G(−λ k ) =
cT Gr (−λ k ), and
k k
k )b
cTk G (−λ
k = k )b
cTk Gr (−λ k , for k = 1, . . . , r.
The SISO (m = p = 1) conditions are replaced in the MIMO case by left tangential,
right tangential, as well as bi-tangential interpolation conditions. From the discussion
of section 2.1, if Ran(Vr ) contains (λ k I + A)−1 Bb
k and Ran(Wr ) contains (λ k I +
A)−T CT ck for each k = 1, 2, . . . , r, then the H2 optimality conditions given above
hold. First-order interpolatory MIMO conditions have been obtained recently in other
independent works as well; see [24, 35].
3.2.2. The discrete time case. An nth-order SISO discrete-time dynamical
system is defined by a set of difference equations
x(t + 1) = Ax(t) + b u(t)
(3.17) G: or G(z) = cT (zI − A)−1 b,
y(t) = cT x(t)
where P11 , Q11 ∈ Rn×n and P22 , Q22 ∈ Rr×r . Wilson [36] showed that the reduced
order model Gr (s) = cTr (sI−Ar )−1 br can be defined in terms of a Galerkin framework
as well by taking
and the resulting reduced order model satisfies the first-order conditions of the optimal
H2 problem. It was also shown in [36] that WrT Vr = I. The next result states the
Lyapunov-based Wilson conditions for H2 optimality and shows their equivalence to
our structured orthogonality framework.
Theorem 3.6. The Wilson conditions for H2 optimality,
A 0 1
P P 1 T b T
(3.27) 2 + P 2 Ar + br b = 0,
0 Ar P
1 and P
Notice that P 2 are independent of we must have
c, so for each choice of b
1 − cT P
cT P
r 2 = 0.
then we have for every b
T [Q
1, Q
2] b
G − Gr , H2 H2 =b = 0.
br
1, Q
Similarly to the first case, [Q so for each choice of
2 ] is independent of b, c we
must have
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
1b + Q
Q 2 br = 0,
1 and P
where the use of P 2 is intended to indicate that they solve (3.27) as well.
Then
!1
W 1
P cr
G − Gr , Gr · H2 H2 = [c , −cr ]
T T
= 0.
!2
W 2
P 0
1 and Q
(Q 2 here also solve (3.28)) and
1
Q 2
Q b
G − Gr , Gr · H2 H2 T ]
= [0, b = 0.
1
Y 2
Y br
and since Y
Since this last equality is true for all b, 1 and Y
2 are independent of b,
we see that Y1 b + Y2 br = 0. We know already that Q1 b + Q2 br = 0, so
Q 1 Q 2 b 0
1 Y 2 = .
Y br 0
1 Q
2
P
1
1
Define Q = R 1 = 0. Premultiply (3.27) by
Q
1 Y
2 2 2 . We
P
will show that R
1
1 Q
2 Y P R
, postmultiply (3.29) by , and subtract the resulting equations to get
Y1 Y2 P2
1 AT − AT R
R 1 = 0 and 2 AT − AT R
R 2 = 1.
cbTr R
r r r r
The first equation asserts that R 1 commutes with AT , and since AT has distinct
r r
eigenvalues, R 1 must have the same eigenvectors as AT . Let y i , x
i be left and right
r
i (respectively, right and left eigenvectors of AT ):
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
x 2y
Ti R i λ i x
i − λ 2y
Ti R Ti
i = x 1y
cbTr R i ,
T
T
0= x i c br y i di .
Either di = 0 or one of x Ti i must vanish, which would then imply that
c and bTr y
1 = 0, which
either dim H < r or dim Gr < r. Thus di = 0 for all i = 1, . . . , r and R
proves the final Wilson condition (3.24).
The converse is omitted here since it follows in a straightforward way by reversing
the preceding arguments.
Hyland and Bernstein [22] offered conditions that are equivalent to the Wilson
conditions. Suppose Gr (s) defined by Ar , br , and cTr solves the optimal H2 problem.
Then there exist positive nonnegative matrices P, Q ∈ Rn×n and two n × r matrices
Fr and Yr such that
(3.30) PQ = Fr MYrT , YrT Fr = Ir ,
where M is similar to a positive definite matrix. Then Gr (s) is given by Ar , br ,
and cTr with Ar = YrT AFr , br = YrT b, and cTr = cT Yr such that, with the skew
projection Π = Fr YrT , the following conditions are satisfied:
(3.31) rank(P) = rank(Q) = rank(PQ),
(3.32) Π AP + PAT + bbT = 0,
T
(3.33) A Q + QA + ccT Π = 0.
Note that in both [36] and [22], the first-order necessary conditions are given
in terms of (coupled) Lyapunov equations. Both [36] and [22] proposed iterative
algorithms to obtain a reduced order model satisfying these Lyapunov-based first-
order conditions. However, the main drawback in each case is that both approaches
require solving two large-scale Lyapunov equations at each step of the algorithm. [40]
discusses computational issues related to solving associated linearized problems within
each step.
Theorems 3.4 and 3.6 show the equivalence between the structured orthogonal-
ity conditions and Lyapunov- and interpolation-based conditions for H2 optimality,
respectively. To complete the discussion, we formally state the equivalence between
the Lyapunov and interpolation frameworks.
Lemma 3.7 (equivalence of Lyapunov and interpolation frameworks). The first-
order necessary conditions of both [22] as given in (3.31)–(3.33) and [36] as given
in (3.23) are equivalent to those of [26] as given in (3.11). That is, the Lyapunov-
based first-order conditions [36, 22] for the optimal H2 problem are equivalent to the
interpolation-based Meier–Luenberger conditions.
We note that the connection between the Lyapunov and interpolation frameworks
has not been observed in the literature before. This result shows that solving the op-
timal H2 problem in the Krylov framework is equivalent to solving it in the Lyapunov
framework. This leads to the Krylov-based method proposed in the next section.
necessary conditions (3.11). Effectiveness of the proposed algorithm results from the
fact that we use rational Krylov steps to construct a Gr (s) that meets the first-order
conditions (3.11). No Lyapunov solvers or dense matrix decompositions are needed.
Therefore, the method is suited for large-scale systems where n 1000.
Several approaches have been proposed in the literature to compute reduced order
models that satisfy some form of first-order necessary conditions; see [37, 34, 9, 21,
26, 22, 36, 25]. However, these approaches do not seem to be suitable for large-scale
problems. The ones based on Lyapunov-based conditions, e.g., [36, 22, 34, 37], require
solving a couple of Lyapunov equations at each step of the iteration. To our knowledge,
the only methods that depend on interpolation-based necessary conditions have been
proposed in [25] and [26]. The authors work directly with the transfer functions of
G(s) and Gr (s); make an iteration on the denominator [25] or poles and residues
[26] of Gr (s); and explicitly compute G(s), Gr (s), and their derivatives at certain
points in the complex plane. However, working with the transfer function, its values,
and its derivative values explicitly is not desirable in large-scale settings. Indeed, one
will most likely be given a state space representation of G(s) rather than the transfer
function. And trying to compute the coefficients of the transfer function can be highly
ill-conditioned. These approaches are similar to [30, 31], where interpolation is done
by explicit usage of transfer functions. On the other hand, our approach, which is
detailed below, is based on the connection between interpolation and effective rational
Krylov iteration, and is therefore numerically effective and stable.
Let σ denote the set of interpolation points {σ1 , . . . , σr }; use these interpolation
points to construct a reduced order model, Gr (s), that interpolates both G(s) and
1 , . . . , λ
G (s) at {σ1 , . . . , σr }; let λ(σ) = {λ r } denote the resulting reduced order
poles of Gr (s); hence λ(σ) is a function from Cr → Cr . Define the function g(σ) =
λ(σ) + σ. Note that g(σ) : Cr → Cr . Aside from issues related to the ordering of
the reduced order poles, g(σ) = 0 yields λ(σ) = −σ; i.e., the reduced order poles
λ(σ) are mirror images of the interpolation points σ. Hence, g(σ) = 0 is equivalent
to (3.11) and is a necessary condition for H2 optimality of the reduced order model,
Gr (s). Thus one can formulate a search for optimal H2 reduced order systems by
considering the root-finding problem g(σ) = 0. Many plausible approaches to this
problem originate with Newton’s method, which appears as
(4.1) σ (k+1) = σ (k) − (I + J)−1 σ (k) + λ σ (k) .
The starting point for another initialization strategy is the H2 expression pre-
sented in Proposition 3.3. Based on this expression, it is appropriate to initiate the
proposed algorithm with σi = −λi (A), where λi (A) are the poles with big residues,
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
where Λ = diag(λ).
4.3. A Newton framework for IRKA. As discussed above, Algorithm 4.1
uses the successive substitution framework by simply setting J = 0 in the Newton
step (4.1). The Newton framework for IRKA can be easily obtained by replacing step
3(b) of Algorithm 4.1 with the Newton step (4.1). The only point to clarify for the
Newton framework is the computation of the Jacobian, which measures the sensitivity
of the reduced system poles with respect to shifts.
Given A ∈ Rn×n and b, c ∈ Rn , suppose that σi , i = 1, . . . , r, are r distinct
points in C, none of which are eigenvalues of A, and define the complex r-tuple
σ = [σ1 , σ2 , . . . , σr ]T ∈ Cr together with related matrices:
(4.2) Vr (σ) = (σ1 I − A)−1 b (σ2 I − A)−1 b . . . (σr I − A)−1 b ∈ Cn×r
and
⎡ ⎤
cT (σ1 I − A)−1
⎢ cT (σ2 I − A)−1 ⎥
⎢ ⎥
(4.3) WrT (σ) = ⎢ .. ⎥ ∈ Cr×n .
⎣ . ⎦
cT (σr I − A)−1
∂λ i
(4.4) =x Ti ∂j WrT AVr x i − λ i Vr x
i + x i x
Ti WrT A − λ Ti WrT ∂j Vr x i ,
∂σj
∂
where ∂j WrT = ∂σ j
WrT = −ej c(σj I−A)−2 and ∂j Vr = ∂σ∂j Vr = −(σj I−A)−2 b eTj .
Proof. With Vr (σ) = Vr and Wr (σ) = Wr defined as in (4.2) and (4.3), both
WrT AVr and WrT Vr are complex symmetric matrices. Write λ for λ
i and x
for x
i ,
so
(4.5) (a) WT AVr x T Vr x
= λW and (b) x T WT AVr = λ x T W T Vr .
r r r r
We reduce the order to r = 6, 5, 4, 3 using IRKA and compare our results with
GFM, OPM, BTM, and the method proposed in [25], denoted by LMPV.
• FOM-3: Example 1 in [34]. Transfer function of FOM-3 is given by
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
s2 + 15s + 50
G(s) = .
s4 + 5s3 + 33s2 + 79s + 50
We reduce the order to r = 3, 2, 1 using IRKA and compare our results with
GFM, OPM, BTM, and the method proposed in [34], denoted by SMM.
• FOM-4: Example 2 in [34]. Transfer function of FOM-4 is given by
10000s + 5000
G(s) = .
s2 + 5000s + 25
We reduce the order to r = 1 IRKA and compare our results with GFM,
OPM, BTM, and SMM.
G(s)−Gr (s) H2
For all these cases, the resulting relative H2 errors G(s) H2 are tabulated in
Table 5.1 below, which clearly illustrates that the proposed method is the only one
that attains the minimum in each case. More importantly, the proposed method
achieves this value in a numerically efficient way staying in the Krylov projection
framework. No Lyapunov solvers or dense matrix decompositions are needed. The
Table 5.1
Comparison.
only arithmetic operations involved are LU decompositions and some linear solvers.
Moreover, our method does not require starting from an initial balanced realization,
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
as suggested in [37] and [22]. In all these simulations, we have chosen a random initial
shift selection, and the algorithm converged in a small number of steps.
To illustrate the evolution of the H2 error throughout the iteration, consider the
model FOM-2 with r = 3. The proposed method yields the following third-order
optimal reduced model:
2.155s2 + 3.343s + 33.8
G3 (s) = .
s3 + 7.457s2 + 10.51s + 17.57
1 = −6.2217 and λ
Poles of G3 (s) are λ 2,3 = −6.1774 × 10−1 ± ı1.5628, and it can be
shown that G3 (s) interpolates the first two moments of G(s) at −λ i for i = 1, 2, 3.
Hence, the first-order interpolation conditions are satisfied. This also means that if
we start Algorithm 4.1 with the mirror images of these Ritz values, the algorithm
converges at the first step. However, we will try four random, but bad, initial selec-
tions. In other words, we start away from the optimal solution. We test the following
four selections: S1 = {−1.01, − 2.01, − 30000}, S2 = {0, 10, 3}, S3 = {1, 10, 3},
and S4 = {0.01, 20, 10000}. With selection S1 , we have initiated the algorithm with
some negative shifts close to system poles, and consequently with a relative H2 error
bigger than 1. However, in all four cases including S1 , the algorithm converged in 5
steps to the same reduced model. The results are depicted in Figure 5.1.
0.8
0.6
0.4
0.2
0
1 2 3 4 5 6 7
k: Number of iterations
Fig. 5.1. H2 norm of the error system vs. the number of iterations.
global minimum when it was started with an initial pole smaller than −0.47. We have
observed exactly the same situation in our simulations. When we start from an initial
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
shift selection smaller than 0.48, IRKA converged to the local minimum. However,
when we start with any initial shift bigger than 0.48, the algorithm converged to the
global minimum in at most 3 steps. Therefore, for this example we were not able the
avoid the local minimum if we started from a bad shift. These observations perfectly
agree with the discussion of section 4.2. Note that the transfer function of FOM-4
can be written as
10000s + 5000 0.99 9999
G(s) = 2 = + .
s + 5000s + 25 s + 0.0050 s + 5000
The pole at −5000 is the one corresponding to the large residue of 9999. Therefore, a
good initial shift is 5000. And if we start the proposed algorithm with an initial shift
at 5000, or close, the algorithm converges to the global minimum.
5.2. CD player example. The original model describes the dynamics between
a lens actuator and the radial arm position in a portable CD player. The model has
120 states, i.e., n = 120, with a single input and a single output. As illustrated in [4],
the Hankel singular values of this model do not decay rapidly and hence the model is
relatively hard to reduce. Moreover, even though the Krylov-based methods resulted
in good local behavior, they are observed to yield large H∞ and H2 error compared
to balanced truncation.
We compare the performance of the proposed method, Algorithm 4.1, with that of
balanced truncation. Balanced truncation is well known to lead to small H∞ and H2
error norms; see [4, 19]. This is due mainly to global information available through
the two system Gramians, the reachability and observability Gramians, which are
each solutions of a different Lyapunov equation. We reduce the order to r, with r
varying from 2 to 40; and for each r value, we compare the H2 error norms due
to balanced truncation and due to Algorithm 4.1. For the proposed algorithm, two
different selections have been tried for the initial shifts. (1) Mirror images of the
eigenvalues corresponding to large residuals, and (2) a random selection with real
parts in the interval [10−1 , 103 ] and the imaginary parts in the interval [1, 105 ]. To
make this selection, we looked at the poles of G(s) having the maximum/minimum
real and imaginary parts. The results showing the relative H2 error for each r are
depicted in Figure 5.2. The figure reveals that both selection strategies work quite
well. Indeed, the random initial selection behaves better than the residual-based
selection and outperforms balanced truncation for almost all the r values except r =
2, 24, 36. However, even for these r values, the resulting H2 error is not far away from
the one due to balanced truncation. For the range r = [12, 22], the random selection
clearly outperforms the balanced truncation. We would like to emphasize that these
results were obtained by a random shift selection and staying in the numerically
effective Krylov projection framework without requiring any solutions to large-scale
Lyapunov equations. This is the main difference our proposed algorithm has with
existing methods and what makes it numerically effective in large-scale settings.
To examine convergence behavior, we reduce the order to r = 8 and r = 10 using
Algorithm 4.1. At each step of the iteration, we compute the H2 error due to the
current estimate and plot this error versus the iteration index. The results are shown
in Figure 5.3. The figure illustrates two important properties for both cases r = 8
and r = 10: (1) At each step of the iteration, the H2 norm of the error is reduced.
(2) The algorithm converges after 3 steps. The resulting reduced models are stable
for both cases.
−1
BT
Large φi
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
10
Random
−2
Relative H2 Error
10
−3
10
−4
10
−5
10
0 4 8 12 16 20 24 28 32 36 40
r: Order of the reduced system
3
r = 10
2.9 r=8
2.8
2.7
H2 error
2.6
2.5
2.4
2.3
2.2
2.1
1 2 3 4 5 6 7 8 9 10
k: Number of iterations
Fig. 5.3. H2 norm of the error system vs. the number of iterations.
SISO system that associates the sixth input of this system with the second output.
We apply our algorithm and reduce the order to r = 6. Amplitude Bode plots of
G(s) and Gr (s) are shown in Figure 5.4. The output response of Gr (s) is virtually
indistinguishable from G(s) in the frequency range considered. IRKA converged in
7 iteration steps in this case, although some interpolation points converged in the
first 2–3 steps. The relative H∞ error obtained with our sixth order system was
7.85 × 10−3 . Note that in order to apply balanced truncation in this example, one
would need to solve two generalized Lyapunov equations (since E = I) of order 20209.
This presents a severe computational challenge, though there have been interesting
approaches to addressing it (e.g., [5]).
−1
10
H(s)
Hr(s)
−2
10
| H(jω)|
−3
10
−4
10
−5
10 −8 −6 −4 −2 0 2
10 10 10 10 10 10
freq ω (rad/sec)
x 10
4 Shift Evolution− Successive Substutition Framework
1
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
0.5
Shift
−0.5
−1
0 5 10 15 20 25 30 35 40 45 50
k: Number of iterations
0
10
1 2 3 4 5 6 7 8 9 10
k: Number of Iterations
0
Shift
−5
−10
−15
10 20 30 40 50 60 70 80 90 100
k: Number of Iterations
Shift Evolution: Newton Framework with σ0 = 2000
2000
Shift
1 2 3 4 5 6 7 8 9 10 11
k: Number of Iterations
and easily show that this reduced model interpolates G(s) and its derivative at σ =
0.2727272. We initiate Algorithm 4.1 with σ0 = 0.27, very close to the optimal shift.
We initiate the Newton framework at σ0 = 2000, far away from the optimal solution.
Convergence behavior of both models is depicted in Figure 5.6. The figure shows
that for this example, the successive substitution framework is divergent and indeed
∂λ
∂σ ≈ 1.3728. On the other hand, the Newton framework is able to converge to the
optimal solution in a small number of steps.
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php
Fix a contour Γ contained in the open left half plane so that the interior of Γ
contains all eigenvalues of M. Then
1 −z
−M(ω 2 I + M2 )−1 = (zI − M)−1 dz.
2πı Γ ω + z 2
2
Thus,
L
1 L
−z
lim (−M)(ω 2 I + M2 )−1 dω = lim dω (zI − M)−1 dz
L→∞ −L 2πı Γ L→∞ −L ω2 + z2
1
= π (zI − M)−1 dz = πI.
2πı Γ
REFERENCES
[1] A.C. Antoulas, Recursive modeling of discrete-time time series, in Linear Algebra for Control
Theory, P. Van Dooren and B. W. Wyman, eds., IMA Vol. Math. Appl. 62, Springer-Verlag,
New York, 1993, pp. 1–20.
[2] A.C. Antoulas, Approximation of Large-Scale Dynamical Systems, Adv. Des. Control 6,
SIAM, Philadelphia, 2005.
[3] A.C. Antoulas and J.C. Willems, A behavioral approach to linear exact modeling, IEEE
Trans. Automat. Control, 38 (1993), pp. 1776–1802.
[4] A.C. Antoulas, D.C. Sorensen, and S. Gugercin, A survey of model reduction methods
for large scale systems, in Structured Matrices in Mathematics, Computer Science, and
Engineering, I (Boulder, CO, 1999), Contemp. Math. 280, AMS, Providence, RI, 2001,
pp. 193–219.
[5] J. Badia, P. Benner, R. Mayo, E.S. Quintana-Ortı́, G. Quintana-Ortı́, and J. Saak,
Parallel order reduction via balanced truncation for optimal cooling of steel profiles, in
Proceedings of the 11th International European Conference on Parallel Processing, EuroPar
2005, Lisbon, J. C. Cunha and P. D. Medeiros, eds., Lecture Notes in Comput. Sci. 3648,
Springer-Verlag, Berlin, 2005, pp. 857–866.
pp. 44–59.
[8] P. Benner and J. Saak, Efficient numerical solution of the LQR-problem for the heat equation,
Proc. Appl. Math. Mech., 4 (2004), pp. 648–649.
[9] A.E. Bryson and A. Carrier, Second-order algorithm for optimal model order reduction, J.
Guidance Control Dynam., 13 (1990), pp. 887–892.
[10] A. Bunse-Gerstner, D. Kubalinska, G. Vossen, and D. Wilczek, H2 -optimal Model Re-
duction for Large Scale Discrete Dynamical MIMO Systems, ZeTeM Technical Report 07-
04, University of Bremen, 2007; available online from http://www.math.uni-bremen.de/
zetem/reports/reports-liste.html#reports2007.
[11] C. De Villemagne and R. Skelton, Model reduction using a projection formulation, Internat.
J. Control, 40 (1987), pp. 2141–2169.
[12] P. Feldman and R.W. Freund, Efficient linear circuit analysis by Padé approximation via a
Lanczos method, IEEE Trans. Computer-Aided Design, 14 (1995), pp. 639–649.
[13] P. Fulcheri and M. Olivi, Matrix rational H2 approximation: A gradient algorithm based
on Schur analysis, SIAM J. Control Optim., 36 (1998), pp. 2103–2127.
[14] D. Gaier, Lectures on Complex Approximation, Birkhäuser, Boston, 1987.
[15] K. Gallivan, E. Grimme, and P. Van Dooren, A rational Lanczos algorithm for model
reduction, Numer. Algorithms, 2 (1996), pp. 33–63.
[16] K. Gallivan, P. Van Dooren, and E. Grimme, On some recent developments in projection-
based model reduction, in ENUMATH 97 (Heidelberg), World Scientific, River Edge, NJ,
1998, pp. 98–113.
[17] E.J. Grimme, Krylov Projection Methods for Model Reduction, Ph.D. thesis, University of
Illinois, Urbana-Champaign, Urbana, IL, 1997.
[18] S. Gugercin, Projection Methods for Model Reduction of Large-Scale Dynamical Systems,
Ph.D. thesis, Rice University, Houston, TX, 2002.
[19] S. Gugercin and A.C. Antoulas, A comparative study of 7 model reduction algorithms, in
Proceedings of the 39th IEEE Conference on Decision and Control, Sydney, 2000, pp. 2367–
2372.
[20] S. Gugercin and A.C. Antoulas, An H2 error expression for the Lanczos procedure, in
Proceedings of the 42nd IEEE Conference on Decision and Control, 2003, pp. 1869–1872.
[21] Y. Halevi, Frequency weighted model reduction via optimal projection, in Proceedings of the
29th IEEE Conference on Decision and Control, 1990, pp. 2906–2911.
[22] D.C. Hyland and D.S. Bernstein, The optimal projection equations for model reduction
and the relationships among the methods of Wilson, Skelton, and Moore, IEEE Trans.
Automat. Control, 30 (1985), pp. 1201–1211.
[23] J.G. Korvink and E.B. Rudyni, Oberwolfach benchmark collection, in Dimension Reduction of
Large-Scale Systems, P. Benner, G. Golub, V. Mehrmann, and D. Sorensen, eds., Lecture
Notes in Comput. Sci. Engrg. 45, Springer-Verlag, New York, 2005.
[24] D. Kubalinska, A. Bunse-Gerstner, G. Vossen, and D. Wilczek, H2 -optimal interpolation
based model reduction for large-scale systems, in Proceedings of the 16th International
Conference on System Science, Wroclaw, Poland, 2007.
[25] A. Lepschy, G.A. Mian, G. Pinato, and U. Viaro, Rational L2 approximation: A non-
gradient algorithm, in Proceedings of the 30th IEEE Conference on Decision and Control,
1991, pp. 2321–2323.
[26] L. Meier and D.G. Luenberger, Approximation of linear constant systems, IEEE Trans.
Automat. Control, 12 (1967), pp. 585–588.
[27] B.C. Moore, Principal component analysis in linear system: Controllability, observability and
model reduction, IEEE Trans. Automat. Control, 26 (1981), pp. 17–32.
[28] C.T. Mullis and R.A. Roberts, Synthesis of minimum roundoff noise fixed point digital
filters, IEEE Trans. Circuits Systems, CAS-23 (1976), pp. 551–562.
[29] T. Penzl, Algorithms for model reduction of large dynamical systems, Linear Algebra Appl.,
415 (2006), pp. 322–343.
[30] L.T. Pillage and R.A. Rohrer, Asymptotic waveform evaluation for timing analysis, IEEE
Trans. Computer-Aided Design, 9 (1990), pp. 352–366.
[31] V. Raghavan, R.A. Rohrer, L.T. Pillage, J.Y. Lee, J.E. Bracken, and M.M. Alaybeyi,
AWE inspired, in Proceedings of the IEEE Custom Integrated Circuits Conference, 1993,
pp. 18.1.1–18.1.8.
[32] F. Riesz and B. Sz.-Nagy, Functional Analysis, Ungar, New York, 1955.
[33] A. Ruhe, Rational Krylov algorithms for nonsymmetric eigenvalue problems II: Matrix pairs,
Linear Algebra Appl., 197 (1994), pp. 283–295.
[34] J.T. Spanos, M.H. Milman, and D.L. Mingori, A new algorithm for L2 optimal model re-
duction, Automat., 28 (1992), pp. 897–909.
[35] P. Van Dooren, K.A. Gallivan, and P.-A. Absil, H2 optimal model reduction of MIMO
Downloaded 07/01/12 to 198.82.177.50. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php