Interpolation Methods for the Model Reduction of Bilinear ...

188
Interpolation Methods for the Model Reduction of Bilinear Systems Garret M. Flagg Dissertation submitted to the Faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mathematics Serkan Gugercin, Chair Joseph A. Ball Christopher A. Beattie Jeffrey T. Borggaard April 30, 2012 Blacksburg, Virginia Keywords: Nonlinear systems, Model reduction, Interpolation theory, Rational Krylov subspace methods Copyright 2012, Garret M. Flagg

Transcript of Interpolation Methods for the Model Reduction of Bilinear ...

Interpolation Methods for the Model Reduction of Bilinear Systems

Garret M. Flagg

Dissertation submitted to the Faculty of the

Virginia Polytechnic Institute and State University

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in

Mathematics

Serkan Gugercin, Chair

Joseph A. Ball

Christopher A. Beattie

Jeffrey T. Borggaard

April 30, 2012

Blacksburg, Virginia

Keywords: Nonlinear systems, Model reduction, Interpolation theory, Rational Krylov

subspace methods

Copyright 2012, Garret M. Flagg

Interpolation Methods for the Model Reduction of Bilinear Systems

Garret M. Flagg

(ABSTRACT) Bilinear systems are a class of nonlinear dynamical systems that arise in

a variety of applications. In order to obtain a sufficiently accurate representation of the

underlying physical phenomenon, these models frequently have state-spaces of very large

dimension, resulting in the need for model reduction. In this work, we introduce two new

methods for the model reduction of bilinear systems in an interpolation framework. Our first

approach is to construct reduced models that satisfy multipoint interpolation constraints

defined on the Volterra kernels of the full model. We show that this approach can be

used to develop an asymptotically optimal solution to the H2 model reduction problem for

bilinear systems. In our second approach, we construct a solution to a bilinear system

realization problem posed in terms of constructing a bilinear realization whose kth-order

transfer functions satisfy interpolation conditions in Ck. The solution to this realization

problem can be used to construct a bilinear system realization directly from sampling data on

the kth-order transfer functions, without requiring the formation of the realization matrices

for the full bilinear system.

Dedication

For Sheena.

“Her children arise up, and call her blessed; her husband also, and he praiseth her. Many

daughters have done virtuously, but thou excellest them all. Favour is deceitful, and beauty

is vain: but a woman that feareth the LORD, she shall be praised. Give her of the fruit of

her hands; and let her own works praise her in the gates.” Proverbs 31: 28-31

iii

Acknowledgments

I am very grateful for all of the support and encouragement I have received from many people

in the course of completing this work. I would first like to acknowledge the formative role

that Serkan Gugercin has had in my mathematical training. He has been an ideal advisor,

steadfastly working with and advocating for me. It been a great privilege and pleasure to

be one of his first Ph.D students, and all his future students have much to look forward

to. Many thanks to Christopher Beattie for all of our good conversations over the past few

years, and for introducing me to several lovely areas of mathematics. I would also like to

thank Joe Ball for teaching me complex analysis and always willingly offering me his insight

into many difficult problems. Thanks also to Jeff Borggaard for serving on my committee.

I am grateful to Kapil Ahuja, Sara Wyatt, Hans-Werner van Wyk, Idir Mechai and Caleb

Magruder for their comradery. This work is as much my wife Sheena’s as it is mine. She has

graciously laboured alongside me in all humility and wisdom, providing encouragement and

inspiration, and spurring me to run the race set before us both in faith. My children James,

Marigold, and Rosemary are my joy and delight, and they have made all the sacrifices of

the last few years mere trifles in comparison to the wonderful gift of getting to know each

them. I also want to thank my siblings Heather, Jeannine, Melanie and Ian for their love and

friendship all these years. Finally I want to offer heartfelt thanks to my parents, Michael and

Brenda. It was their hard work, done in faith and love, that set me on the firm foundation

iv

which is Jesus Christ, and taught me to love him foremost. Unto the Lord Jesus Christ be

all glory, honor, and praise.

v

Contents

1 Introduction 1

2 Bilinear Systems 6

2.1 Volterra series representation of the input-output operator. . . . . . . . . . . . 7

2.2 Bilinear system stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.3 System grammians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.4 Bilinear system norms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.5 Approximation of nonlinear systems . . . . . . . . . . . . . . . . . . . . . . . . . 33

3 Model Reduction and Interpolation 44

3.1 The Petrov-Galerkin model reduction framework . . . . . . . . . . . . . . . . . . 45

3.2 Interpolation-based model reduction . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.3 Subsystem Interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.4 Volterra Series Interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4 H2 Optimal Model Reduction 61

vi

4.1 Alternatives to H2 Optimal Bilinear Model Reduction . . . . . . . . . . . . . . 99

5 Solving the Bilinear Sylvester Equations 107

5.1 Direct Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

5.2 Iterative Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

5.3 Krylov projection-based approximation of ordinary Sylvester equations . . . . 113

5.4 Krylov projection-based methods for the approximation of the bilinear Lya-

punov equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6 Data-Driven Model Reduction of SISO Bilinear Systems 127

6.1 Classical Bilinear Realization Theory . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.2 The structure of the interpolation data . . . . . . . . . . . . . . . . . . . . . . . . 136

6.3 Construction of the Bilinear Realization . . . . . . . . . . . . . . . . . . . . . . . 139

6.4 Volterra kernel sampling methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

7 Conclusions 163

7.1 A summary of contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

7.2 Directions for future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

Bibliography 166

vii

List of Figures

2.1 Comparison of the steady-state behavior for the linear, quadratic, and fourth

order polynomial heat-transfer systems . . . . . . . . . . . . . . . . . . . . . . . . 42

4.1 Comparison of the relative H2 error for B-IRKA and TB-IRKA approxima-

tions to the Fokker-Planck system . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

4.2 Comparison of average time per iteration using B-IRKA and TB-IRKA[13

terms] for the Fokker-Planck system . . . . . . . . . . . . . . . . . . . . . . . . . 90

4.3 Comparison of the relative H2 error for B-IRKA and TB-IRKA approxima-

tions to the nonlinear heat-transfer system . . . . . . . . . . . . . . . . . . . . . 91

4.4 Steady state response of nonlinear heat-transfer system and unscaled bilinear

B-IRKA and TB-IRKA approximations of order 12 . . . . . . . . . . . . . . . . 92

4.5 Comparison of TB-IRKA and B-IRKA approximations of nonlinear heat trans-

fer system scaled with α = 5 × 104 . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

4.6 Steady state response of nonlinear heat-transfer system and scaled bilinear

B-IRKA and TB-IRKA approximations of order 12 . . . . . . . . . . . . . . . . 94

4.7 Comparison of TB-IRKA and B-IRKA approximations of Burgers’ equation

control system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

viii

4.8 Comparison of average time per iteration for TB-IRKA[2,4] and B-IRKA ap-

plied to Burgers’ equation control system . . . . . . . . . . . . . . . . . . . . . . 96

4.9 Comparison of average time per iteration in TB-IRKA and B-IRKA for several

orders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

4.10 Comparison of TB-IRKA and B-IRKA approximations of heat transfer control

system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

4.11 Convection-diffusion problem: Comparison of the relative H2 error in the

B-IRKA and TB-IRKA[2, 3 and 6 terms] approximations taking p0 = 1 and

varying over the parameter range for p1 and p2 . . . . . . . . . . . . . . . . . . . 99

4.12 Convection-diffusion problem: Comparison of the relative H2 error in the

B-IRKA and TB-IRKA[2 terms] approximations taking p0 = 0.5 and varying

over the parameter range for p1 and p2 . . . . . . . . . . . . . . . . . . . . . . . . 100

4.13 Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem

interpolation response to the true response for the input u(t) = e−t . . . . . . . 103

4.14 Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem

interpolation error for the input u(t) = e−t . . . . . . . . . . . . . . . . . . . . . 104

4.15 Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem

interpolation response to the true response for the input u(t) = (cos(πt/10)+1)/2105

4.16 Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem

interpolation error for the input u(t) = (cos(πt/10) + 1)/2 . . . . . . . . . . . . 105

4.17 Burgers’ Equation: A comparison of the TB-IRKA and scaled B-IRKA

error for the input u(t) = e−t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

ix

4.18 Burgers’ Equation: A comparison of the TB-IRKA and scaled B-IRKA

error for the input u(t) = sin(20t) . . . . . . . . . . . . . . . . . . . . . . . . . . 106

5.1 Relative error in the 2-norm as r varies for the EADY Model . . . . . . . . . . 120

5.2 Relative error in the 2-norm as r varies for the Rail Model . . . . . . . . . . . . 121

5.3 Comparison of the relative error in the L-norm for pseudo-H2 projection sub-

space and SVD approximations of the heat transfer model . . . . . . . . . . . . 126

5.4 Comparison of the relative error in the Frobenius norm for pseudo-H2 projec-

tion subspace and SVD approximations of the heat transfer model . . . . . . . 126

x

List of Symbols

Rm×n set of real matrices of size m by n

Cm×n set of complex matrices of size m by n

s a complex number

∣s∣ modulus of s

AT transpose of A

A∗ complex conjugate transpose of A

diag(a1, . . . , ak) diagonal matrix with diagonal elements a1, . . . , ak

∥A∥p induced p-norm of a matrix

∥A∥F Frobenius norm of a matrix

I the identity matrix of appropriate size

ı√−1

λi(A) the ith eigenvalue of A

A⊗B the Kronecker product of A and B.

Σ a linear time-invariant dynamical system

ξ a generic nonlinear dynamical system

ζ a bilinear time-invariant dynamical system

ζ a reduced-dimension bilinear system

ζN a polynomial system of degree N

ζN a reduced-dimension polynomial system of degree N

hk(t1, . . . , tk) the kth order Volterra kernel of ζ in the time domain

Hk(s1, . . . , sk) the transfer function of the kth order Volterra kernel of ζ

Hk(s1, . . . , sk) the transfer function of the kth order Volterra kernel of ζ

∥ζ∥H2 H2 norm of a bilinear system

xi

Chapter 1

Introduction

High fidelity modeling of complex physical phenomena frequently results in dynamical sys-

tems with very large complexity. These cumbersome models often outstrip the computational

resources available for using the models in applications like system control, simulations and

data assimilation. A wide variety of model reduction techniques have been developed to

ameliorate this problem for linear time invariant (LTI) dynamical systems. See [2] and the

references therein for further information on model reduction of LTI systems. The options

are fewer for nonlinear dynamical systems, and they naturally depend heavily on the partic-

ular class of nonlinear systems under consideration. For highly nonlinear phenomenon where

little is known analytically about the system dynamics, principle orthogonal decomposition

(POD) and its variants, such as the Discrete Empirical Interpolation Method (DEIM) are

the main approach to model reduction [31]. More can be said for systems whose nonlinear

dynamics are analytic functions of the state and input. Under small perturbations, such as

inputs with a small magnitude, the input-output map for systems of this kind can be accu-

rately represented as a Volterra series [81, 27]. Nonlinear systems that admit a Volterra series

representation are frequently referred to as weakly nonlinear systems. Bilinear systems are an

1

Garret M. Flagg Chapter 1. Introduction 2

important class of weakly nonlinear systems that are well-suited to accurately representing

nonlinear phenomenon resulting from inputs of small magnitude, and their simple algebraic

structure makes it possible to obtain a deeper insight into their properties. A bilinear system

with m inputs and p outputs is characterized by the following set of equations

ζ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) =Ax(t) +m

∑k=1Nkx(t)uk(t) +Bu(t)

y(t) =Cx(t),

(1.1)

where A,Nk ∈ Rn×n for k = 1, . . .m, B ∈ Rn×m and C ∈ Rp×n.

For fixed inputs ζ is linear in the state, and for a fixed state it is linear in the input, hence

the name bilinear systems. The nonlinear properties of the system are due to multiplicative

coupling of the state and the input through the terms Nk. In the not so distant 1970’s,

bilinear systems received a flurry of attention due in large part to their many applications–

described in [73, 74, 81, 72, 28]– together with the momentum gained by the complete

algebraic characterization of linear dynamical systems in the work of Kalman [62, 63, 64]

that resulted in many system-theoretic results related to classical realization theory being

completely generalized to bilinear systems in work done by d’Alessandro, Isidori, Brockett,

Frazho, Fliess and Sontag [36, 26, 27, 46, 47, 48, 49, 88]. More recently, the subject of

bilinear realization theory has been revisited in the work of Petreczky [77] for switched

bilinear systems.

Bilinear systems arise as natural models for physical systems ranging from nuclear fission to

DC brush motors. They can also be used to appromixate generic weakly nonlinear dynamical

systems of the form

ξ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) = f(x(t), t) + g(x(t), t)u(t),

y(t) =Cx(t)(1.2)

Garret M. Flagg Chapter 1. Introduction 3

where f and g are analytic functions of the state, and continuous in t. In this context,

an approximation technique called the Carleman linearization can be used to construct a

bilinear system approximation to ξ that matches N terms in the Taylor series expansion of

f and g around some equilibrium state. Bilinear systems have been applied in this context

to the modeling of nonlinear RC circuits, and microelectromechanical systems (MEMS) such

as parallel-plate electrostatic actuators [6].

Certain types of linear stochastic differential equations also have the form of (1.3). For ex-

ample, a bilinear system results from the spatial discretization of the Fokker-Planck equation

in [29]. Bilinear systems coming from the Carlemann linearization or stochastic differential

equations frequently have very large order. If the nonlinear system (1.2) has k states, and the

Carleman linearization matches N terms in the Taylor series, the resulting bilinear system

approximation is order n = k + k2 + ⋯ + kN . Hence, there is a real need for a theory and

techniques of model reduction of bilinear systems. Given a bilinear system ζ of order n, the

goal of the model reduction is to construct a bilinear system

ζ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

˙x(t) = Ax(t) +m

∑k=1Nkxuk(t) + Bu(t)

y(t) = Cx(t)

(1.3)

such that A, Nk ∈ Rr×r and B ∈ Rr×p, C ∈ Rp×r, for some r ≪ n. Throughout the remainder

of this work, all reduced-order quantities will always be denoted with tildes, unless otherwise

specified.

There are SVD-based approaches to bilinear model reduction that suitably generalize bal-

anced truncation for linear systems [16], [29]. These approaches have proven to be very accu-

rate, but as in the linear case they fall prey to computational challenges involved in the solu-

tion of large scale generalized Sylvester equations. We will therefore focus on interpolation-

based approaches to bilinear model reduction which can produce accurate models at a much

Garret M. Flagg Chapter 1. Introduction 4

lower cost. Petrov-Galerkin projection and its connection with rational interpolation the-

ory provides a powerful theoretical framework for the model reduction of linear dynamical

systems. Interpolation-based Petrov-Galerkin techniques for bilinear model reduction were

developed in [5], [6],[24], [78],[34]. Recently, necessary conditions for optimality in the H2

norm for bilinear systems were given in [20]. In this work, a new interpolation framework for

bilinear systems is introduced that places these necessary conditions firmly within the inter-

polation framework by explicitly reformulating them as multipoint interpolation conditions

on the Volterra kernels in the frequency domain. Expressions for the H2 norm are derived

which generalize the familiar linear expressions, illuminating the connections between bilin-

ear H2 optimality and the poles and residues of the bilinear subsystem transfer functions.

In chapter 5 we will consider a generalization of the classical bilinear realization theory that

makes it possible to construct bilinear realizations directly from data on the kernels of the

Volterra series representation of the bilinear system sampled anywhere in their domain of

definition. The construction we develop also generalizes the results on univariate rational

interpolation to rational functions in k variables having a very special (and simple) type of

polar set.

A few words on notational conventions and some frequently used concepts from linear algebra

are now in order. We will frequently make use of the vec operator and Kronecker product

–both important tools in linear algebra. The Kronecker product of two matrices A∈ Cm×n

and B ∈ Cu×v is denoted A⊗B and is defined as

A⊗B =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

a1,1B a1,2B . . . a1,nB

a2,1B a2,2B . . . a2,nB

⋮ ⋮ ⋮ ⋮

am,1B am,2B . . . am,nB

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

∈ Cum×nv

Garret M. Flagg Chapter 1. Introduction 5

The binary operator ⊗ will be used to denote the Kronecker product unless explicitly stated

otherwise. The vec operator is a group isomorphism from Rm×n to Rnm defined by simply

stacking the columns of a matrix M ∈ Rm×n into one long vector. One of the more useful

algebraic properties of the vec operator that we will use frequently is

vec(MPT ) = (T T ⊗M)vec(P ).

All matrix-valued quantities will be indicated with bold-face type and denoted by captial

letters, and all vector-valued quantities will be denoted with lower-case letters and bold-face

type. All scalar quantities will be represented in ordinary type-face. ζ will denote a bilinear

system and Σ will denote an LTI system.

Chapter 2

Bilinear Systems

In this chapter we will examine some of the basic system-theoretic properties of bilinear

time-invariant dynamical systems. In §2.1 we first consider the external representation of a

bilinear dynamical system as a nonlinear operator B ∶ U → Y mapping admissible inputs u ∈ U

to outputs y ∈ Y, and derive the Volterra series representation of B. Section 2.2 deals with the

internal representation of ζ in terms of its realization parameters (A,N1, . . . ,Nm,C,B) and

consider properties such as system controllability and observability formulated in terms of

the realization of ζ. In §2.4 we introduce bilinear system norms and derive a new expression

for the H2 norm of a bilinear system in terms of the transfer function representation of

ζ. Finally in §2.5 we will consider the use of bilinear systems in approximating nonlinear

dynamical systems more broadly and introduce the Carleman linearization technique.

6

Garret M. Flagg Chapter 2. Bilinear Systems 7

2.1 Volterra series representation of the input-output

operator.

The output y(t) of the bilinear system

ζ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) =Ax(t) +m

∑k=1Nkx(t)uk(t) +Bu(t)

y(t) =Cx(t),

(2.1)

can be constructed as a Volterra series with Volterra kernels defined explicitly by the coef-

ficient matrices A, Nk for k = 1 . . . ,m, B, C. For inputs uk(t) that are bounded on a time

interval [0, T ] this equation is Lipschitz continuous in the state x and continuous in t, so

the Picard-Lindelof theorem guarantees that (1.3) has a solution on any finite time interval

[0, T ] [99]. Let N = [N1, . . . ,Nm] and define the change of variable z(t) = e−Atx(t). Apply

this change of variable to (2.1) to obtain the equivalent system

z(t) = N(t)(Im ⊗ z(t))u(t) + B(t)u(t)

y(t) = Cz(t),z(0) = 0,(2.2)

where N(t) = e−AtN(Im ⊗ eAt), B = e−AtB, and C = CeAt. The solution for z(t) is now

constructed by applying the Picard iteration [99]. First, write

z(t) = ∫t

0N(σ1)(Im ⊗ z(σ1))u(σ1)dσ1 + ∫

t

0B(σ1)u(σ1)dσ1 (2.3)

Next, write

z(σ1) = ∫

σ1

0N(σ2)(Im ⊗ z(σ2))u(σ2)dσ2 + ∫

σ1

0B(σ2)u(σ2)dσ2 (2.4)

Garret M. Flagg Chapter 2. Bilinear Systems 8

and substitute (2.4) into (2.3) to get

z(t) =∫t

0∫

σ1

0N(σ1)(Im ⊗ [N(σ2)(Im ⊗ z(σ2))u(σ2)]u(σ1))dσ2 dσ1

+ ∫

t

0∫

σ1

0N(σ1)[Im ⊗ B(σ2)u(σ2)]u(σ1)dσ2 dσ1+ (2.5)

t

0B(σ1)u(σ1)dσ1 (2.6)

= ∫

t

0∫

σ1

0N(σ1)[(Im ⊗ N(σ2))(Im ⊗ (u(σ2)⊗ z(σ2))]u(σ1)dσ2 dσ1 (2.7)

+ ∫

t

0∫

σ1

0N(σ1)[u(σ1)⊗ B(σ2)u(σ2)]dσ2 dσ1+ (2.8)

t

0B(σ1)u(σ1)dσ1 (2.9)

(2.10)

= ∫

t

0∫

σ1

0N(σ1)[(Im ⊗ N(σ2))(Im ⊗ Im ⊗ z(σ2))]u(σ1)⊗u(σ2)dσ2 dσ1 (2.11)

+ ∫

t

0∫

σ1

0N(σ1)[Im ⊗ B(σ2)]u(σ1)⊗u(σ2)dσ2 dσ1+ (2.12)

t

0B(σ1)u(σ1)dσ1 (2.13)

(2.14)

Garret M. Flagg Chapter 2. Bilinear Systems 9

Continuing this process, after N steps gives

z(t) =∫t

0∫

σ1

0⋯∫

σN−1

0N(σ1)(Im ⊗ N(σ2))⋯

(Im ⊗ Im⋯⊗´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶N−1 times

N(σN))(Im ⊗ Im⋯Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

N times

⊗z(σN))u(σ1)⊗u(σ2)⊗⋯⊗u(σN)dσn⋯dσ1+

(2.15)

N

∑k=1∫

t

0∫

σ1

0⋯∫

σk−1

0N(σ1)(Im ⊗ N(σ2))⋯(Im ⊗⋯⊗ Im

´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶k−1 times

N(σk−1)) (2.16)

⋅ (Im ⊗ Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k times

⊗B)(σk))u(σ1)⊗u(σ2)⊗⋯⊗u(σk)dσk⋯dσ1 (2.17)

By assumption, N(t), z(t), and uk(t) are bounded on [0, T ], so there exists some K > 0 s.t.

K > max sup0<t<T

{∥N(t)∥, ∥u∥, ∥z(t)∥} (2.18)

Therefore

∣∫

t

0∫

σ1

0⋯∫

σN−1

0N(σ1)(Im ⊗ N(σ2))⋯(Im ⊗ Im ⊗⋯⊗

´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶N−1 times

N(σN))

⋅(Im ⊗ Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

N times

⊗z(σN))u(σ1)⊗u(σ2)⊗⋯⊗u(σN)dσN⋯dσ1∣ <K2N+1tN

N !.

Thus, letting N →∞ and changing back to the original variables yields a uniformly conver-

gent Volterra series representation for the solution of y(t) as

Garret M. Flagg Chapter 2. Bilinear Systems 10

y(t) =∞

∑k=1∫

t

0∫

σ1

0⋯∫

σk−1

0CeA(t−σ1)N(Im ⊗ eAσ1−σ2N)⋯(Im ⊗⋯⊗ Im

´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶k−2 times

⊗eAσk−1−σkN)

⋅ (Im ⊗ Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗B)u(σ1)⊗u(σ2)⊗⋯⊗u(σk)dσk⋯dσ1 (2.19)

The form of the Volterra kernels is simpler and possibly more enlightening in the case of

single-input-single-output (SISO) bilinear systems. In the SISO case, a realization of the

bilinear system (2.1) is given by (A, N , b, c) where now N ∈ Rn×n is a single matrix and

b,cT ∈ Rn are vectors. In the SISO case, the Volterra series in (2.19) reduces to

y(t) =∞

∑k=1∫

t

0∫

σ1

0⋯∫

σk−1

0ceA(t−σ1)NeA(σ1−σ2)N

⋯NeA(σk−1−σk)bu(σk)u(σk−1)⋯u(σ1)dσk⋯dσ1

This representation is given in terms of the so-called triangular Volterra kernels:

hk(σ1, σ2, . . . , σn) = ceA(t−σ1)NeA(σ1−σ2)N⋯NeA(σn−1−σn)b.

Our interest in the Volterra kernels will lie predominantly in their frequency domain repre-

sentation. To analyze them in that setting we introduce the multivariate Laplace transform.

Definition 2.1. Given a function hk(t1, . . . , tk) defined on Rk+, define its Laplace transform

Hk(s1, . . . , sk) by

Hk(s1, . . . , sk) =

∫0

∫0

h(t1, . . . , tn)e

k

∑j=1

tjsjdt1⋯dtk (2.20)

Garret M. Flagg Chapter 2. Bilinear Systems 11

The multivariate Laplace transform of the triangular kernels yields expressions that are

difficult to analyze . To gain some clarity, it is useful to make the change of variable t = σ0,

tn−i = σi − σi+1. The Volterra kernels can then be written in the so-called regular form as

hk(t1, t2, . . . , tk) = ceAtnNeAtk−1N⋯NeAt1b.

Provided the matrix A of the bilinear system is Hurwitz, the k-variate Laplace transform of

hk is

Hk(s1, s2, . . . , sk) = c(skI −A)−1N(sn−1I −A)−1N⋯N(s1I −A)−1b (2.21)

Returning to the MIMO case, similar expressions can be derived for the input-output rela-

tionship in terms of the regular kernels of the Volterra series (2.19) as

y(t) =∞

∑i=1∫

t

0∫

t1

0⋯∫

ti−1

0h(t1, t2, . . . , tk)(u(t −

i

∑k=1

tk)⊗⋯⊗u(t − ti))dtk⋯dt1. (2.22)

and the regular Volterra kernels are given as

h(t1, t2, . . . , tk) =CeAtkN(Im ⊗ eAtk−1)(Im ⊗ N)⋯

(Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗eAt2)(Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗N) (2.23)

⋅ (Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗eAt1)(Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗B).

The multivariable Laplace transform Hk(s1, . . . , sk) of the degree k regular kernel (2.23) of

Garret M. Flagg Chapter 2. Bilinear Systems 12

ζ is given by

Hk(s1, . . . , sk) =C(skI −A)−1N [Im ⊗ (sk−1I −A)−1](Im ⊗ N)⋯

⋅ [Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗(s2I −A)−1](Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗N) (2.24)

⋅ [Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗(s1I −A)−1](Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗B).

Let Ik =k

⨉j=1

{1,2, . . . ,m} be the k-fold Cartesian product of the indices j = 1, . . . ,m, so

that each element i ∈ Ik corresponds to some possible k-tuple combination of the indices

j = 1, . . . ,m. Upon inspection of the definition of Hk(s1, . . . , sk) in equation (2.24), it is

clear that it may be decomposed into Mk = mk−1 matrix-valued rational functions in Cp×m

as

Hk(s1, . . . , sk) = [C(skI −A)−1Ni1(k−1)⋯Ni1(1)(s1I −A)−1B,

C(skI −A)−1Ni2(k−1)⋯Ni2(1)(s1I −A)−1B, . . . ,

C(skI −A)−1NiMk(k−1)⋯NiMk(1)(s1I −A)−1B]

(2.25)

where each ij ∈ Ik is distinct for j = 1, . . . ,Mk.

2.2 Bilinear system stability

The standard formulation of stability for a linear system on [0,∞) is the following.

Definition 2.2. The linear system Σ is bounded-input-bounded-output (BIBO) stable if

for any bounded input, the output is bounded on [0,∞).

For a linear system to be BIBO stable it is sufficient for A to be Hurwitz, that is, for

Garret M. Flagg Chapter 2. Bilinear Systems 13

maxi(Re(λi(A))) < 0. Due to the action of N on trajectory, this is no longer the case for

bilinear systems. For example, let ζ be a SISO system with a constant input u(t) ≡ α ∈ R.

For this input, ζ can be written as the linear system

Σ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) = (A + αN)x(t) + bα

y(t) = cx(t)(2.26)

and therefore its stability depends on the eigenvalues of A + αN . For any nontrivial N , α

can be chosen sufficiently large so that maxi(Re(λi(A+αN))) > 0, and hence for this input

ζ will have an unbounded output.

It follows that in general, this definition is too strong a formulation of BIBO stability for

bilinear systems, since it is only satisfiable for linear systems. The next theorem, due to Siu,

and Schetzen in [87] provides sufficient conditions to guarantee that all sufficiently bounded

inputs yield bounded outputs.

Theorem 2.1. Suppose there exists an M > 0 so that the input ∥u(t)∥ =

√m

∑k=1

∣uk(t)∣2

satisfies ∥u(t)∥ ≤M for all t > 0. Let Γ =m

∑k=1

∥Nk∥. Then the output of ζ given by (2.1) with

inputs uk(t) is bounded on [0,∞) if there exists scalars β > 0 and 0 < α ≤ −maxi(Re(λi(A))),

such that ∥eAt∥ ≤ βe−αt, t ≥ 0 and Γ < α/Mβ.

In light of these considerations, BIBO stability of a bilinear system makes sense inside of the

ball ∥u∥∞ < α/(Γβ). Outside of this ball, no guarantee can be made on the boundedness of

the system’s outputs.

Garret M. Flagg Chapter 2. Bilinear Systems 14

2.3 System grammians

The concepts of controllability and observability for bilinear systems were first considered in

[36] and [26] and generalize these notions for linear systems in a straightforward way.

Definition 2.3. A state x of system (2.1) is reachable from the origin if there exists an

input function u(t) ∈ L2(Rm)[0, T ] that maps the origin of the state space into the state x

in time t ≤ T .

Due to the nonlinearity of ζ, the set of reachable states does not generally form a subspace

of Rn. As a result, reachability is formulated as a somewhat weaker condition on the span

of reachable states.

Definition 2.4. [81] The bilinear system (2.1) is called span reachable if the set of reachable

states spans Rn.

The reachability of an LTI system can be completely characterized by the Krylov subspace

Kn = span[B,AB, . . . ,An−1B]. If the rank of Kn is equal to n, the system is completely

reachable, and the subspace of reachable states is the image of Kn. In a similar manner,

define P1 = B, and Pi = [APi−1, N(Im ⊗Pi−1)] for i = 1, . . . , n. The span reachability of

ζ is determined by the span of Range(Pn). In particular, ζ is span reachable if and only if

rank(Pn) = n. See [81] for further details.

Unobservable states are also defined in the usual way.

Definition 2.5. The state x0 ≠ 0 is unobservable if the response y(t) from x(0) = x0 is

equal to the response from x(0) = 0 for all inputs u(t) ∈ L2[0, T ]

Definition 2.6. The bilinear system (2.1) is observable provided it has no unobservable

states.

Garret M. Flagg Chapter 2. Bilinear Systems 15

Unlike the set of reachable states, the set of observable states is a linear subspace of Rn.

Define QT1 = CT , N⊕T = [NT

1 , . . . ,NTm] and QT

i = [ATQTi−1 N

⊕T (Im ⊗QTi−1)]. Then the

subspace of unobservable states is equal to N (Qn) [81], where N (⋅) denotes the nullspace of

an operator.

Alternative characterization of span reachability and observability can be given in terms of

the controllability and observability grammians of bilinear systems. Following D’Alessandro,

Isidori, and Ruberti [36], first define

p1(t1) = eAt1B (2.27)

plk−1,...,l1

(t1, . . . , tk) = eAtkNlk−1e

Atk−1Nlk−2⋯Nl1eAt1B. (2.28)

The reachability grammian is then defined as

P =∞

∑k=1

∫0

∫0

m

∑lk−1=1

⋯m

∑l1=1

plk−1,...,l1

pTlk−1,...,l1

dt1⋯dti (2.29)

Similarly, define

q1(t1) = ceAt1 (2.30)

qlk−1,...,l1

(t1, . . . , tk) = ceAtkNlk−1e

Atk−1Nlk−2⋯Nl1eAt1 . (2.31)

Then the observability grammian is defined as

Q =∞

∑k=1

∫0

∫0

m

∑lk−1=1

⋯m

∑l1=1

qTlk−1,...,l1

qlk−1,...,l1

dt1⋯dti (2.32)

Theorem 2.2. [1] Provided they exist, P and Q solve the following generalized Lyapunov

equations.

Garret M. Flagg Chapter 2. Bilinear Systems 16

AP +PAT +n

∑k=1

NkPNTk +BBT = 0 (2.33)

ATQ +QA +n

∑k=1

NTk QNk +C

TC = 0 (2.34)

For the sake of completeness, we sketch the proof here.

Proof. It suffices to show the result for (2.33). The result follows similarly for (2.34).

P1 =

∫0

p1pT1 dt1, (2.35)

where p1 is defined in equation (2.27) solves

AP1 +P1AT +BBT = 0,

Continuing, define

P2 =

∫0

∫0

m

∑l1=1

eAt2Nl1eAt1BBT eA

T t1NTl1eA

T t2dt1dt2

Then P2 solves

AP2 +P2AT +

m

∑j=1

NjP1NTj = 0

Garret M. Flagg Chapter 2. Bilinear Systems 17

and for k > 2

Pk =

∫0

∫0

m

∑lk−1=1

⋯m

∑l1=1

plk−1,...,l1

pTlk−1,...,l1

dt1⋯dti (2.36)

=

∫0

eAtk(m

∑lk−1=1

Nlk−1Pk−1NTlk−1)e

AT tkdtk (2.37)

(2.38)

and hence Pk solves

APk +PkAT +

m

∑j=1

NjPk−1NTj = 0.

Summing these equations for k = 1, . . . ,N gives

A(N

∑k=1

Pk) + (N

∑k=1

Pk)AT +

m

∑j=1

Nj(N−1

∑k=1

Pk−1)NTj +BBT = 0.

Letting N →∞ yields the desired result.

As Zhang and Lam have pointed out in [105], solutions P , Q of the the generalized Lyapunov

equations (2.33), (2.34) may exist even though the integrals defining P and Q diverge. The

next theorem clarifies the conditions under which the solutions of the generalized Lyapunov

equations (2.33), (2.34) are equal to P and Q, respectively.

Theorem 2.3. [36],[105] Suppose that A is Hurwitz and that P and Q uniquely solve (2.33),

(2.34). Then

1.) P = P ≻ 0 iff ζ is span reachable.

2.) Q = Q ≻ 0 iff ζ is observable.

Garret M. Flagg Chapter 2. Bilinear Systems 18

Let x(t,x0, u) denote the solution of 2.1 at time t with input u(t) and x(0,x0, u) = x0. For

a given x0 ∈ Rn, and some bound α > 0 on the L2 norm of the inputs, define the input and

output energy functionals

Ec(x0) = infu∈L2(−∞,0]x(−∞,x0,u)=0

0

∫−∞

∣u(t)∣2dt, (2.39)

Eα0 (x0) = max

u∈L2[0,∞)

∥u∥L2<α

∫0

∣y(⋅,x0,0)∣2dt (2.40)

If ζ is linear (N = 0), then by the stability arguments given above, we may drop the

dependency on α, and the grammians provide information on the minimum energy required

to drive a system from a state x0 to 0, and the maximum possible energy for an output

observed from initial state x0. Let P♯ be the Moore-Penrose inverse of P . Then Ec(x) is

related to P by

Ec(x) = xTP ♯x (2.41)

and Eo(x0) is given by the quadratic form

Eo(x0) = xT0Qx0 (2.42)

A concept closely connected to the energy functionals Ec and Eo is the balanced realization

of a system.

Definition 2.7. The realization (A, N1, . . . ,Nm, B, C) of the bilinear system (2.1) is said

Garret M. Flagg Chapter 2. Bilinear Systems 19

to be balanced if P =Q = Σ solves

AΣ +ΣAT +n

∑k=1

NkΣNTk +BBT = 0

ATΣ +ΣA +n

∑k=1

NTk ΣNk +C

TC = 0,

where Σ > 0 is a diagonal matrix with diagonal entries σ1 > σ2 > ⋅ ⋅ ⋅ > σn > 0. The quantities

σi for i = 1, . . . , n are the singular values of the bilinear system.

Remark 2.1. In general, the singular values of the bilinear system (2.1) are given as

σi =√λi(PQ) =

√λi(QP) for i = 1, . . . , n

Remark 2.2. Given P ,Q > 0 then a balancing transformation T is given in terms of

P = LLT and LTQL = UΣ2UT as T = LUΣ−1/2.

When the system Σ is linear (Nk = 0) and balanced, equations (2.41), (2.42) indicate that

the states which require the smallest amount of energy to control also correspond to the

initial states that yield the largest output energy, a situation that only occurs when the

system is balanced. A linear system with a balanced realization yields information about

which states are most important for capturing the dominant system dynamics. The situation

is more complicated for bilinear systems, but for sufficiently bounded inputs, the grammians

do provide estimates on the controllability and observability energies for a given state close

to the origin.

Theorem 2.4. [16] Given a bilinear system ζ assume that

P = Q = diag(σ1, . . . , σn), σi > 0. Then there exists an ε > 0, so that for all canonical unit

vectors ei, the inequalities

Ec(εei) > ε2eTi P−1ei = ε

2/σi, (2.43)

Garret M. Flagg Chapter 2. Bilinear Systems 20

and

Eo(εei) < ε2eTi Qei = ε2σi (2.44)

The controllability and observability grammians as we have defined them here provide a

natural generalization to their counterparts in linear systems theory. We mention here that

balanced truncation for more general classes of nonlinear systems has also been studied by

Scherpen, Fujimoto and collaborators. See [50] and the references therein for further details.

As we have seen, under suitable hypotheses the grammians can be construed similarly as

providing information on the dominant local dynamics of ζ. There are, however, alternative

formulations of the system grammians that are worth considering briefly here, given their

general algebraic resemblance to relations that we will derive later. Recently Couchman et.

al. observed that the Q, P defined in equations (2.33) and (2.34) are not invariant under

varying time-scales [35]. For the system ζ given by (2.1), make the time transformation

τ = 1/αt for some α > 0. This results in the bilinear system

ζ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(τ) = αAx(t) + αm

∑k=1Nkx(τ)uk(τ) + αBu(τ)

y(τ) =Cx(τ),

(2.45)

And therefore after the time transformation, assuming that ζ is span reachable, P will be

given as the solution of the equation

αAP + αPA + α2m

∑k=1

NTk PNk = −BB

T (2.46)

As a result, the states which are most reachable/observable in the standard formulation

depend on the time scale involved, which is an undesirable result for dynamical systems

Garret M. Flagg Chapter 2. Bilinear Systems 21

where the dominant dynamics occur at different time-scales. In [35] Couchman et. al.

propose an alternative formulation of controllability and observability for bilinear systems

of a slightly different form

ζ ∶

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) =Ax(t) +m

∑k=1Nkx(t)uk(t) +Bw(t)

y(t) =Cx(t),

(2.47)

The only difference between (2.47) and (2.1) is that the forcing term w(t) ∈ Rm in equation

(2.47) may differ from the inputs uk(t) that are coupled with the state x(t). Consider the

set of inputs U = {u ∶ [0,∞) → R∣ supt

∣u(t)∣ ≤ 1}. One possible way to make the grammians

invariant under varying time-scales is to consider matricesPD,QE which satisfy the following

theorem.

Theorem 2.5. [35] Given a bilinear system ζ with realization (A, N1, . . . ,Nm, B, C) and

matrices PD,QE ≻ 0 then

(A +m

∑k=1

uk(t)Nk)PD +PD(A +m

∑k=1

uk(t)Nk)T +BBT ≺ 0, (2.48)

(A +m

∑k=1

uk(t)Nk)TQE +QE(A +

m

∑k=1

uk(t)Nk) +CTC ≺ 0 (2.49)

holds for all t ∈ [0,∞) and uk ∈ U for k = 1, . . . ,m if and only if

(A +m

∑k=1

Nk)PD +PD(A +m

∑k=1

Nk)T +BBT ≺ 0 (2.50)

(A +m

∑k=1

Nk)TQE +QE(A +

m

∑k=1

Nk) +CTC ≺ 0 (2.51)

Moreover, PD, QE satisfy the following energy inequalities:

1. If w = 0, then the energy in the output y for initial condition x0 is bounded from above

Garret M. Flagg Chapter 2. Bilinear Systems 22

by

maxu∈B(0,α)

∥y∥2L2

< xT0QEx0 (2.52)

2. The minimum energy of the disturbance w for all input sequences u ∈ B(0, α) required

to drive the system from x(−∞) = 0 to x(0) = x0 is bounded from below according to

∀u ∈ B(0, α),∀x0 ∈ Rn ∶ minw∈L2[−∞,0)

∥w∥2L2

> xT0P−1D x0 (2.53)

The matrices PD and QE are called D-grammians and E-grammians respectively. They

resolve the problem of time-scale dependence, since any time-scale transformation τ = 1/αt

corresponds to scaling the matricesPD andQE by 1/α, and therefore the dominant dynamics

of the system interpreted in terms of these grammians does not change with the time scale.

The D and E grammians are clearly nonunique, but given their interpretation in terms of

bounds on the input/output energy of the system, they can be used to develop a balanced

truncation approach in the usual manner. Moreover, balanced truncation applied to these

grammians yields the following error bounds.

Theorem 2.6. [35] Assume that PD = QE = diag(σ1, . . . , σn), σi > σi+1 for i = 1, n − 1,

satisfying the conditions of Theorem 2.5 are balanced, and that ζr is the reduced order model

computed by truncating after the σrth singular value. Then

maxu∈B(0,α)

maxw∈L2[0,∞)

(∥y − yr∥L2[0,∞)

∥w∥L2[0,∞))

2

≤ 2n

∑j=r+1

σj (2.54)

In order to minimize this bound as much as possible, an LMI based approach can be used

to minimize the functional

f(PDQE) = trace(PDQE) (2.55)

as done in [35]. The cost involved in the ensuing minimization program will make this

Garret M. Flagg Chapter 2. Bilinear Systems 23

approach intractable for very large order bilinear models, but it constitutes a unique and

theoretically powerful alternative to the current formulation of bilinear system grammians.

2.4 Bilinear system norms

We turn now to the definition of system norms for the bilinear system ζ given in (2.1).

The Lp norms, with p ≥ 1, generalize to bilinear systems in a natural way. Recall that the

Lp[0,∞) norm of a linear system is defined on the impulse response h(t) as

∥Σ∥Lp = (

∫0

∥h(t)∥pp)1/p

. (2.56)

The Lp norm of a bilinear system is similarly defined on the kth order Volterra kernels.

Definition 2.8. For p ≥ 1, the Lp norm of ζ is

∥ζ∥Lp = (∞

∑i=1

∫0

∫0

∥hi(t1, . . . , ti)∥ppdt1⋯dti)

1/p

(2.57)

With this definition, the L2 norm of a bilinear system can be expressed in terms of the

bilinear system grammians.

Proposition 2.1. [105] Let ζ=(A, N1, . . . ,Nm, B, C) have a finite L2 norm. Then

∥ζ∥2L2

= trace(CPCT ) = trace(BTQB)

Proof. Let hk1,...,ki

(t1, . . . , ti) =CeAtiNki⋯Nk2eAt1bk1 . From equation (2.23),

hi(t1, . . . , ti)hi(t1, . . . , ti)T =

m

∑k1=1

⋯m

∑ki=1

hk1,...,ki

(t1, . . . , ti)hk1,...,ki(t1, . . . , ti)T ,

Garret M. Flagg Chapter 2. Bilinear Systems 24

and

∫0

∫0

m

∑k1=1

⋯m

∑ki=1

hk1,...,ki

(t1, . . . , ti)hk1,...,ki(t1, . . . , ti)T

=

∫0

∫0

m

∑k1=1

⋯m

∑ki=1

CeAtiNki⋯Nk2eAt1bk1b

Tk1eA

T t1NTk2⋯NT

kieA

T tiCT

=CP iCT

Summing over the P i and taking the trace gives the desired equality. The proof involving

Q is done analogously, using the fact that trace(hTi hi) = trace(hihTi ).

The Hp Hardy spaces can also be generalized to bilinear systems. Let s = x + ıy ∈ C with

x, y ∈ R. In the linear case, the Hp norm of Σ is defined on its transfer function H(s) as

∥Σ∥Hp = supx>0

(

∫−∞

∥H(x + ıy)∥ppdy)1/p

, (2.58)

where ∥H(s)∥p = (m

∑i=1σi(H(s))p)

(1/p)

is the Schatten p-norm of H(s).

Definition 2.9. For p ≥ 1, the Hp norm of ζ in (2.1) is

∥ζ∥Hp = (∞

∑i=1

supx1>0,...,xi>0

∫−∞

∫−∞

∥Hi(x1 + ıy1, . . . , xi + ıy1)∥ppdy1⋯dyi)

1/p

, (2.59)

where ∥Hi(s1, . . . , si)∥p is the Schatten p-norm of Hi(s1, . . . , si).

If H1(s1) is analytic in C+, then trace(H1(s1)TH1(s1)) =

m

∑i,j=1

∣H i,j1 (s1)∣

2 is subharmonic

and satisfies zero-order growth asymptotics on C+, since H i,j1 (s1) is a proper rational matrix

function with poles in the left half-plane. This means that for any ε > 0, there exists Aε > 0

so that trace(H1(s1)TH1(s1)) ≤ Aεeε∣s∣, s ∈ C+. For the special cases p = ∞ and p = 2 we

Garret M. Flagg Chapter 2. Bilinear Systems 25

may therefore apply the Phragmen-Lindelof Principle to H(s) on the domain D = C+, which

says that the maximum occurs on the boundary of D, see [7]. Thus, the H2 and H∞ norms

reduce to

∥H1(s1)∥∞ = supω∈R

maxi=1,...m

σi(H1(ıω)),

and

∥H1(s1)∥H2= (

∫−∞

trace(HT1 (−ıω)H1(ıω))dω)

1/2

For transfer functions satisfying such conditions, the H2-norm and the frequency-domain L2

norm are equivalent. To obtain a similar result for the bilinear H2 norm requires a slightly

deeper analysis of the kth-order transfer functions. Recall that the transfer function of the

kth-order homogenous subsystem of a SISO system is given as

Hk(s1, s2, . . . , sk) = c(skI −A)−1N(sk−1I −A)−1N⋯N(s1I −A)−1b

Writing (siI −A)−1 as the classical adjoint over the determinant, it is readily seen that

Hk(s1, s2, . . . , sk) =P (s1, s2, . . . , sk)

Q(s1)Q(s2)⋯Q(sk)(2.60)

Where Q(sj) = det(sjI −A) for j = 1, . . . , k, and P (s1, s2, . . . , sk) is a k-variate polynomial

with maximum total degree kn. Thus Hk(s1, s2, . . . , sk) is a proper k-variate rational function

with singularities of a very simple analytic variety. This allows the for the extension of the

equivalence result.

Theorem 2.7. Assume that A is Hurwitz. Then

∥ζ∥L2[0,∞) = ∥ζ∥L2(ıR) = ∥ζ∥H2 (2.61)

Garret M. Flagg Chapter 2. Bilinear Systems 26

Proof. The first equality

∥ζ∥L2[0,∞) = ∥ζ∥L2(ıR)

directly follows from the application of the Plancherel’s theorem in n-variables [22]. For a

fixed variable si, the second equality follows by applying the Phragmen-Lindelof Principle

to each variable separately in the expression for the H2 norm.

The Hardy space H2 norm of a linear system ζ ∶=(A, 0, b, c) can be written in terms of the

poles and residues of the system’s transfer function. The following theorem describes this

result.

Theorem 2.8. [53] Let H(s) =m

∑k=1

φks−λk

be an asymptotically stable linear system. Then

∥H∥H2 = (m

∑k=1

φkH(−λk))1/2

(2.62)

The fact that the polar sets (the analytic varieties of the singularities) of Hk(s1, . . . , sk) are

separable into k − 1 dimensional hyperplanes makes it possible to give a partial fraction

expansion of H(s1, . . . , sk) that avoids all the intricacies of k-variate residue theory. To this

end, define the following quantities.

Definition 2.10. For a kth-order homogenous subsystem H(s1, . . . , sk), let

φl1,...,lk

= limsk→λlk

(sk − λlk) limsk−1→λlk−1

(sk−1 − λlk−1)⋯ lims1→λl1

(s1 − λl1)H(s1, . . . , sk) (2.63)

We now prove our first result, a pole-residue decomposition for the kth-order transfer function

of a bilinear system.

Theorem 2.9 (Pole-Residue Formula for H(s1, . . . , sk)).

Let H(s1, . . . , sk) =P (s1, . . . , sk)

Q(s1)Q(s2)⋯Q(sk)where P (s1, . . . , sk) is a polynomial in k variables of

Garret M. Flagg Chapter 2. Bilinear Systems 27

total degree k(n − 1) and Q(si) is a polynomial of degree n in the variable si with simple

zeros at the points λ1, . . . , λn ∈ C. Then

H(s1, . . . , sk) =n

∑l1=1

⋯n

∑lk=1

φl1,...,lk

k

∏i=1

(si − λli)

(2.64)

Proof. Since H(s1, . . . , sk) =P (s1, s2, . . . , sk)

Q(s1)Q(s2)⋯Q(sk), the function

F (s1, . . . , sk) =H(s1, . . . , sk)Q(s2)⋯Q(sk) is holomorphic on Ck ∖ ∪ni=1{λi} ×Ck−1. The sets

Ai = {λi} ×Ck−1 are analytic varieties given by the functions f(s1, . . . , sk) = (s1 − λi) respec-

tively. Note that by Hartog’s Extension Theorem, (see [65] for details) (s1 −λi)F (s1, . . . , sk)

extends to a holomorphic function on Ai, so that in a neighborhood of any point p =

(λi, p2, . . . , pk) ∈ Ai,

(s1 − λi)F (s1, . . . , sk) =∞

∑∣j∣=0

αj(s1 − λi)j(1)(s2 − p2)

j(2)⋯(sk − pk)j(k)

where the sum is taken over all k-tuples j ∈ Zk+ and ∣j∣ =k

∑i=1j(i). This implies that

F (s1, . . . , sk) =∞

∑∣j0∣=0

αj0(s2 − p2)j0(2)⋯(sk − pk)j0(k)

s1 − λi+G(s1, . . . , sk), (2.65)

where G is holomorphic on Ai and the indices j0 satisfy j0(1) = 0. Let Q−1(λi) =n

∏j=1j≠i

(λi −λj).

Then from the definition of F (s1, . . . , sk) in (2.65), it follows that

lims1→λi

(s1 − λi)F (s1, . . . , sk) =P (λi, s2, . . . , sk)

Q−1(λi)=

∑∣j0∣=0

αj0(s2 − p2)j0(2)⋯(sk − pk)

j0(k) (2.66)

Garret M. Flagg Chapter 2. Bilinear Systems 28

Thus, on Ai,

F (s1, . . . , sk) =P (λ1, s2, . . . , sk)

Q−1(λi)(s1 − λ)+G(s1, . . . , sk) (2.67)

Subtracting each of the “principal parts”P (λi, s2, . . . , sk)

Q−1(λi)(s1 − λi)from F and combining terms gives

U(s1, . . . , sk) = F (s1, . . . , sk) − (n

∑i=1

Li(s1)P (λi, s2, . . . , sk))/Q(s1) (2.68)

where Li(s1) is the Lagrange polynomial determined by the points λj, j ≠ i for j = 1, . . . , n.

U is entire on Ck, so we now show that U ≡ 0. Now note that by assumption, the maximum

degree of s1 is n − 1, so

P (s1, . . . , sk) =n−1

∑j=0

sj1αj(s2, . . . , sk) (2.69)

where the coefficients αj(s2, . . . , sk) are polynomials. For any values of the coefficients αj,

the polynomial in s1 is uniquely determined by the points λ1, . . . , λn; thus

P (s1, . . . , sk) =n

∑i=1

Li(s1)P (λi, s2, . . . , sk)

and therefore, by the definition of F , U ≡ 0. So we now have that

P (s1, . . . , sk)/Q(s1) ≡n

∑i=1

Li(s1)P (λi, s2, . . . , sk)/(Q(s1)

on Ck.

Thus

P (s1, . . . , sk)/Q(s1)Q(s2) =n

∑i=1

P (λi, s2, . . . , sk)

Q−1(λi)(s1 − λi)Q(s2)(2.70)

Garret M. Flagg Chapter 2. Bilinear Systems 29

Now repeatedly applying the same argument as above to the functions

P (λl1 , λl2 , . . . , λli−1 , si, . . . , sk)/Q(si) for l1, . . . , li−1 = 1, . . . , n

gives the desired result:

H(s1, . . . , sk) =n

∑l1=1

⋯n

∑lk=1

φl1,...,lk

k

∏i=1

(si − λli)

(2.71)

Remark 2.3. The partial fraction decomposition can be derived directly from the state-space

representation of H(s1, . . . , sk) = c(skI −A)−1N⋯N(s1I −A)−1b. H(s1, . . . , sk) is invariant

under state-space representations, so we may assume the A = diag(λ1, . . . , λm). Expanding

the state-space representation directly gives the decomposition desired. Here, the first few

Garret M. Flagg Chapter 2. Bilinear Systems 30

steps in the expansion are shown:

H(s1, s2, . . . , sk) =

c(skI −A)−1N⋯

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

n1,1 n1,2 . . . n1,m

n2,1 n2,2 . . . n2,m

⋮ ⋱ ⋮

nm,1 . . . nm,m

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

b1/(s1 − λ1)

b2/(s1 − λ2)

bm/(s1 − λm)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

= c(skI −A)−1N⋯

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

n1,1 n1,2 . . . n1,m

n2,1 n2,2 . . . n2,m

⋮ ⋱ ⋮

nm,1 . . . nm,m

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

m

∑l1=1

n1,l1bl1

(s2−λ1)(s1−λl1)

m

∑l1=1

n2,l1bl1

(s2−λ2)(s1−λl1)

m

∑l1=1

nm,l1bl1(s2−λm)(s1−λl1)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⋮ after k − 1 steps

= [c1/(sk − λ1) c2/(sk − λ2) . . . cr/(sk − λm)]

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

m

∑lk−1=1

⋯m

∑l1=1

n1,lk−1nlk−1,lk−2⋯nl3,l2nl2,l1bl1k−1∏i=1

(si−λli)

m

∑lk−1=1

⋯m

∑l1=1

n2,lk−1nlk−1,lk−2⋯nl3,l2nl2,l1bl1k−1∏i=1

(si−λli)

m

∑lk−1=1

⋯m

∑l1=1

nm,lk−1nlk−1,lk−2⋯nl3,l2nl2,l1bl1k−1∏i=1

(si−λli)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(2.72)

=n

∑l1=1

⋯n

∑lk=1

φl1,...,lk

k

∏i=1

(si − λli)

The pole-residue decomposition of the transfer functions can be used to derive an expression

Garret M. Flagg Chapter 2. Bilinear Systems 31

for the H2 bilinear system norm. This expression was also given independently by Breiten

and Benner in [14], though our derivation of it here is new.

Theorem 2.10 (H2 norm expression). Let ζ be a SISO bilinear system with a finite H2

norm. Then ∥ζ∥2H2

=∞

∑k=1

n

∑l1=1

n

∑l2=1

⋯n

∑lk=1

φl1,...,lk

H(−λl1 , . . . ,−λlk)

Proof. From Theorem 2.9

H(s1, . . . , sk) =n

∑l1=1

⋯n

∑lk=1

φl1,...,lkk

∏i=1

(si − λli)

, (2.73)

and from Theorem 2.7

∥ζ∥2H2

= ∥ζ∥2L2(ıR)

=∞

∑k=1

1

(2π)k

∫−∞

∫−∞

Hk(−ıω1, . . . ,−ıωk)Hk(ıω1, . . . , ıωk)dωkdωk−1⋯dω1 (2.74)

Substituting (2.73) for Hk(ıω1, . . . , ıωk) at the kth term in the series (2.74) and considering

this term alone gives

=1

(2π)k

∫−∞

ı∞

∫−∞

n

∑l1=1

⋯n

∑lk=1

φl1,...,lkHk(−ıω1, . . . ,−ıωk)k

∏i=1

(ıωi − λli)

dωkdωk−1⋯dω1

=n

∑l1=1

⋯n

∑lk=1

1

(2π)k

∫−∞

∫−∞

φl1,...,lkHk(−ıω1, . . . ,−ıωk)k

∏i=1

(ıωi − λli)

dωkdωk−1⋯dω1

=n

∑l1=1

⋯n

∑lk=1

φl1,...,lkHk(−λl1 , . . . ,−λlk) (2.75)

The expression in (2.75) is an application of Cauchy’s formula in k-variables, in the following

way. Consider the contours γRj = [−ıRj, ıRj] ∪ {z = Rjeıθ for π/2 ≤ θ ≤ 3π2 } for j = 1, . . . , k

in the complex plane, and let Γ =k

⨉j=1γRj be the distinguished boundary of the polycylinder

Garret M. Flagg Chapter 2. Bilinear Systems 32

given by the set of points DR1,...,Rj = {(s1, . . . , sk)∣sj ∈ intγRj for j = 1, . . . k}, where “int”

denotes the interior of the contour. For all sufficiently large Rj, j = 1, . . . , k all the points

(λl1 , . . . , λlk) ∈ DR1,...,Rk for l1, . . . , lk = 1, . . . , n. But the functions Hk(−s1, . . . ,−sk) are holo-

morphic on DR, and so by Cauchy’s formula (see [80] for details on extending Cauchy’s

formula to polycylinders)

Hk(−λl1 , . . . ,−λlk) =1

(2πı)k ∫γR1

⋯∫γRk

Hk(−s1, . . . ,−sk)k

∏i=1

(si − λli)

dskdsk−1⋯ds1

=1

(2πı)k ∫γR1

⋯∫γR2

(

3π/2

π/2

−ıHk(−s1, . . . ,−Rke−ıθk)Rkıeıθk

k−1

∏i=1

(si − λli)(Rkeıθk − λlk)

dθk

+

Rj

∫−Rj

ıH(−s1, . . . ,−ıωk)k−1

∏i=1

(si − λli)(ıωk − λlk)

dωk)dsk−1⋯ds1

Letting Rj →∞, the term

3π/2

π/2

−ıHk(−s1, . . . ,−Rke−ıθk)Rkıeıθk

k−1

∏i=1

(si − λli)(Rkeıθk − λlk)

dθk∣→ 0,

since Hk(−s1, . . . , sk) is a proper rational function in the variable sk. Thus,

Hk(−λl1 , . . . ,−λlk) =1

(2πı)k ∫γR1

⋯∫γRk

Hk(−s1, . . . ,−sk)k

∏i=1

(si − λli)

dskdsk−1⋯ds1

=1

(2πı)k ∫γR1

⋯∫γR2

∫−∞

ıH(−s1, . . . ,−ıωk)k−1

∏i=1

(si − λli)(ıωk − λlk)

dωk)dsk−1⋯ds1

Garret M. Flagg Chapter 2. Bilinear Systems 33

Repeating this argument k − 1 times yields the desired result that

Hk(−λl1 , . . . ,−λlk) =1

(2π)k

∫−∞

∫−∞

Hk(−ıω1, . . . ,−ıωk)k

∏i=1

(ıωi − λli)

dωkdωk−1⋯dω1 (2.76)

Since this holds for every k, returning to our original goal we now have that

∑k=1

1

(2π)k

∫−∞

∫−∞

Hk(−ıω1, . . . ,−ıωk)Hk(ıω1, . . . , ıωk)dωkdωk−1⋯dω1

=∞

∑k=1

n

∑l1=1

⋯n

∑lk=1

φl1,...,lkHk(−λl1 , . . . ,−λlk)

2.5 Approximation of nonlinear systems

We conclude the background discussion for bilinear systems by considering to what extent

finite dimensional bilinear systems may be used to approximate nonlinear systems generally,

and how such approximations can be constructed. The latter topic consists of a presentation

of the standard Carleman linearization technique. The approximation capabilities of bilinear

systems were considered independently by Sussman [92] and Fliess [47]. Let F (u) ∶ U ⊂

Rm → R be any functional that maps m inputs u in the set of admissible inputs U to the real

numbers. Let B(u) denote specifically any such mapping determined by a bilinear system.

The results of Sussmann and Fliess are summarized as follows.

Theorem 2.11. Suppose that F is causal, and that all admissible inputs are bounded on some

finite time interval [0, T ]. Moreover, assume that F is continuous in the weak∗ topology on

the input semigroup S(U) defined by the semigroup operation of concatenation. For every

Garret M. Flagg Chapter 2. Bilinear Systems 34

ε > 0, there exists a bilinear system Bε such that

sup0≤t≤T

∣F (u)(t) −Bε(u)(t)∣ < ε

for all inputs u ∈ U .

So the output behavior of any weakly continuous, causal, input-output map may be approx-

imated arbitrarily close by a bilinear system. Frequently such an input-output map can be

characterized by a set of first-order nonlinear differential equations in the form

x(t) = f(x(t), t) +m

∑k=1g[k](x(t), t)uk(t), x(0) = x0

y(t) = c(x(t), t)

(2.77)

Assume that the vector-valued functions functions f ,g[k] for k = 1, . . . ,m, c are analytic

in x and continuous in t. Systems of this kind are called linear-analytic, because they are

linear in the input and analytic in the state. Linear-analytic systems are considered weakly

nonlinear, and can be described by a Volterra series for inputs of small magnitude, a result

which is summarized in the following theorem.

Theorem 2.12. [81, 27] Suppose a solution to the unforced linear-analytic system exists for

t ∈ [0, T ]. Then there exists an ε > 0 such that for all inputs satisfying ∣∣u(t)∣∣ < ε, there is a

Volterra system representation of the input-output mapping that converges on [0, T ].

The Carleman linearization applies to the class of linear-analytic systems. The exposition

of the Carleman linearization technique presented here closely follows the development by

Rugh in [81]. To simplify matters we make the slightly stronger assumption that y(t) is a

continuously differentiable function of t. We can then write

Garret M. Flagg Chapter 2. Bilinear Systems 35

y(t) = ∇xc(x, t)T x(t) +

∂tc(x, t) (2.78)

with y(0) = c(x0,0). Since y(t) is a linear-analytic state-equation, y(t) can be appended to

the state-vector x(t) to form a new n + 1 dimensional state vector x(t), and yielding a new

linear-analytic system

˙x(t) = f(x(t), t) +m

∑k=1g[k](x(t), t)uk(t), x(0) = x0

y(t) = cx(t)

(2.79)

where now c = [0 ⋯ 0 1], and the n + 1th entry of f is

fn+1(x(t), t) = ∇xc(x, t)T x(t) +

∂tc(x, t)

and the n + 1th entry of the vector-valued function g[k] is

g[k]n+1(x(t), t) = ∇xc(x, t)

Tg[k](x(t), t) (2.80)

Thus, we can always rewrite (2.77) so that y(t) is a linear function of the state. Moreover,

(2.77) can always be simplified further so that x0 = 0 and that f(0, t) = 0. If x0 ≠ 0, then

let x(t) = x(t) − z(t), where z(t) is the solution to (2.79) with zero forcing term and initial

condition x0. Then

Garret M. Flagg Chapter 2. Bilinear Systems 36

˙x(t) = ˙x(t) − z(t)

= f(x, t) − f(z, t) +m

∑k=1

g[k](x(t), t)uk(t)

= f(x + z, t) − f(z, t) +m

∑k=1

g[k](x + z, t)uk(t)

= f(x, t) +m

∑k=1

g[k](x(t), t)u(t) (2.81)

y(t) = cx(t) + cx0, x(0) = 0

So it is sufficiently general to consider all linear analytic systems of the form

x(t) = f(x(t), t) +m

∑k=1g[k](x(t), t)uk(t), x(0) = 0

y(t) = cx(t) + y0(t)

(2.82)

To proceed with the Carleman linearization, we will need to keep track of all the terms

in an n-variate Taylor series expansion. To simplify this bookkeeping task, let x(i) =

x⊗x⊗⋯⊗x´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

i−1 times

∈ Rni . Using this notation, one can write the Taylor series expansion of

an analytic function about the point x = 0 as

f(x) = f(0) +F1x(1) +F2x

(2) + . . . +Fix(i) + . . . (2.83)

Applying this expansion to the linear-analytic state equations in (2.84) and truncating after

N terms in each series yields

Garret M. Flagg Chapter 2. Bilinear Systems 37

x(t) =N

∑i=1Fi(t)x(i)(t) +

m

∑k=1

N−1

∑i=0G

[k]i (t)x(i)(t)u(t), x(0) = 0

y(t) = cx(t) + y0(t)

(2.84)

The crucial step in the linearization is developing a differential equation for each of the terms

x(j). Consider x(2) first.

d

dtx(2)(t) = x⊗x +x⊗ x

= (N

∑i=1

Fi(t)x(i)(t) +

m

∑k=1

N−1

∑i=0

G[k]i (t)x(i)(t)uk(t))⊗x

+x⊗ (N

∑i=1

Fi(t)x(i)(t) +

m

∑k=1

N−1

∑i=0

G[k]i (t)x(i)(t)uk(t))

= (N

∑i=1

Fi(t)⊗ In + In ⊗Fi(t) +m

∑k=1

N−1

∑i=0

G[k]i (t)⊗ In + In ⊗G

[k]i (t))x(i+1)

To continue with the approximation procedure, all terms x(i) with i > N are dropped from

the sum in Fi and all terms with i > N − 1 in the sum with the Gi. This leaves

d

dtx(2)(t) = (

N−1

∑i=1

Fi(t)⊗ In + In ⊗Fi(t) +m

∑k=1

N−2

∑i=0

G[k]i (t)⊗ In + In ⊗G

[k]i (t))x(i+1) (2.85)

Proceeding in the same manner for N ≥ i > 2 yields the differential equations

d

dtx(i)(t) =

N−i+1

∑j=1

Fi,j(t)xi+j−1(t) +

m

∑k=1

N−i

∑j=0

G[k]i,j (t)x

(i+j−1)uk(t) (2.86)

Where F1,j(t) = Fj(t), and for i > 1

Fi,j(t) = Fj(t)⊗ In ⊗ In⋯⊗ In + In ⊗Fj(t)⊗ In ⊗⋯⊗ In +⋯ + In ⊗⋯⊗ In ⊗Fj(t) (2.87)

Garret M. Flagg Chapter 2. Bilinear Systems 38

so that there are i − 1 total Kronecker products in each term, and i terms in the total sum.

Similarly when i=1, define G1,j(t) =Gj(t) for j = 0, . . . ,N − 1 and for i > 1

Gi,j(t) =Gj(t)⊗ In ⊗ In⋯⊗ In + In ⊗Gj(t)⊗ In ⊗⋯⊗ In +⋯+ In ⊗⋯⊗ In ⊗Gj(t) (2.88)

Stacking the x(i) in a vector x = [x(1) x(2) x(3) . . . x(N)]T

∈ R∑Nj=1 nj , yields the following

bilinear system approximation to the linear-analytic system in (2.84)

x(t) =A(t)x(t) +m

∑k=1Nk(t)x(t)uk(t) +B(t)u(t)

y(t) = cx(t) + y0(t),

(2.89)

where

A(t) =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

F1,1 F1,2 ⋯ F1,N

0 F2,1 ⋯ F2,N−1

0 0 ⋯ F3,N−2

⋮ ⋮ ⋮ ⋮

0 0 ⋯ FN,1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, Nk(t) =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

G[k]1,1 G

[k]1,2 ⋯ G

[k]1,N−1 0

G[k]2,0 G

[k]2,1 ⋯ G

[k]2,N−2 0

0 G[k]3,0 ⋯ G

[k]3,N−3 0

⋮ ⋮ ⋮ ⋮

0 0 ⋯ G[k]N,0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

for k = 1, . . . ,m,

B(t) =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

G[1]1,0 . . . G

[m]

1,0

0 . . . 0

0 . . . 0

⋮ ⋮

0 . . . 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, c = [c 0 ⋯ 0]

(2.90)

Garret M. Flagg Chapter 2. Bilinear Systems 39

A bilinear model of the Fitzhugh-Nagumo equations

The FitzHugh-Nagumo equations are a simplified version of the Hodgkin-Huxley model for

the activation and deactivation dynamics of a spiking neuron.

They are given as

v(t) = v(v − κ)(v − 1) −w + i(t)

τw(t) = σv − γw

y(t) = v(t)

(2.91)

where σ, γ are positive constants, v is the membrane potential and w the density of a chemical

substance, and 0 < κ < 1. i(t) is the excitation current input to the system. Taking v = x1

and w = x2 and x = [x1, x2]T , this system can be written as

x(t) = f(x(t)) + e1i(t)

y(t) = eT1 x,(2.92)

where f has the obvious definition. Then f(0, t) = 0, so the system is linear analytic, and

in the form specified by (2.84). As a simple, low dimensional illustration of the Carleman

linearization, let us rewrite system (2.92) as a bilinear system. Since the nonlinearities in

(2.92) are quadratic, the bilinear system representation will match the input-out map of the

system (2.92) exactly for initial conditions at zero. Expanding f about zero gives

F1 =

⎡⎢⎢⎢⎢⎢⎢⎣

−κ −1

σ −γ

⎤⎥⎥⎥⎥⎥⎥⎦

, F2 =

⎡⎢⎢⎢⎢⎢⎢⎣

−2(1 + κ) 0 0 0

0 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎦

, F3 =

⎡⎢⎢⎢⎢⎢⎢⎣

6 0 . . . 0

0 0 . . . 0

⎤⎥⎥⎥⎥⎥⎥⎦

, (2.93)

and G1,0 = e1. For j > 0, Gj = 0.

The other terms in the bilinearization are generated from Fj andGj, j > 0 as in the equations

Garret M. Flagg Chapter 2. Bilinear Systems 40

(2.87) and (2.88), yielding a bilinear system of dimension 2 + 22 + 23 = 14 that matches the

exact dynamics of the original system.

Nonlinear Heat Transfer Model

In this next example a novel bilinear model for a nonlinear heat transfer problem first

introduced by Yousefi et. al. in [102] is contstructed. The physical system to be modeled is

the heat transfer along a 1D beam with length L, cross sectional area A, and nonlinear heat

conductivity represented by a polynomial in temperature T (x, t) of arbitrary degree N

κ(T ) = a0 + a1T + ⋅ ⋅ ⋅ + anTN (2.94)

The right end of the beam (at x = L) is fixed at ambient temperature. The model has

two inputs: a time-dependent uniform heat flux u1(t) at the left end (at x = 0) and a

time-dependent heat source u2(t) distributed along the beam. Including the nonlinear heat

conductivity in the differential form of the heat transfer equation gives

−∇ ⋅ (κ(T )∇T ) + ρcpT = u2(t). (2.95)

Where ρ is the material density, and cp is the heat capacity. Applying the definition of κ(T )

to this equation yields the heat transfer system governed by the equations

−N

∑i=0

ai∇ ⋅ (T i∇T ) + ρcpT = u2(t) (2.96)

By applying the Ritz-Galerkin orthogonality requirements to (2.96) in the weak formulation

on a test-space of linear 1D finite elements leads to the following finite-element discretization

Garret M. Flagg Chapter 2. Bilinear Systems 41

of (2.96):

KT + ρcpMT =Bu + k(T ) (2.97)

Where T ∈ Rn is the spatially discretized temperature, K,M ∈ Rn×n, B ∈ Rn×2 and k(T ) ∶

T → Rn collects together all the nonlinear terms. The matrices M and K are invertible,

tridiagonal, linear mass and stiffness matrices defined as

M = A`

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1/3 1/6

1/6 2/3 1/6

⋱ ⋱ ⋱

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

K = a0A/`

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 −1

−1 2 −1

⋱ ⋱ ⋱

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

where ` = L/n, the length of a single element on the beam. B and k are defined as

B =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

A A`/2

0 A`

0 Al

⋮ ⋮

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

k(T ) = A/l

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

N

∑i=i

ai(Ti+11 −T i+12 )

i+1

N

∑i=1

−ai(Ti+11 −2T i+12 +T i+13 )

i+1

N

∑i=1

−ai(Ti+1k−1−2T i+1k +T i+1k+1)

i+1

N

∑i=1

−ai(Ti+1n−1−2T i+1n )

i+1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(2.98)

The function g(T ) is analytic in T , and its Taylor series terminates after N +1 terms. In this

application the heat conductivity is a 4th order polynomial, so the polynomials in g(T ) are all

5th order. Strictly speaking, a bilinear system could be constructed that exactly matches the

nonlinear equations starting from zero initial conditions for all inputs. If the finite element

discretization has n0 elements, this would result in a bilinear realization of order n =5

∑j=1nj0,

Garret M. Flagg Chapter 2. Bilinear Systems 42

which grows large much too fast to be of any practical use. However, the coefficients of the

polynomials in the heat conductivity decay very rapidly. For example a reasonable choice of

polynomial coefficients sets a0 = 144.495, a1 = −0.5434, a2 = 9.27496×10−4, a3 = −8.28691×10−7

and a4 = 3.18727 × 10−10. Thus, it is reasonable to anticipate that the terms up to degree

2 in g dominate the system dynamics, and truncate the Taylor series expansion after the

second term. This expectation is confirmed for typical inputs of interest into the system.

When the system order is n = 400 Figure 2.1 shows the response of the nonlinear system

(2.97), the quadratic approximation (setting the coefficients a2 = a3 = a4 = 0), and the linear

approximation to a constant heat flux input of u1(t) = 5 × 104W /m2, while fixing u2(t) ≡ 0.

The response is measured at the tenth node on the beam for each model. As the figure

illustrates, the quadratic approximation closely matches the steady-state behavior of the

original nonlinear system, whereas the linear system only provides a crude approximation to

the true response.

0 50 100 150 200 250 300−5

0

5

10

15

20

25

30

35

Time [s]

Tem

pera

ture−3

00 [K

]

Original ModelQuadratic Approx.Linear Approx.

Figure 2.1: Comparison of the steady-state behavior for the linear, quadratic, and fourthorder polynomial heat-transfer systems

Thus, we will derive the bilinear realization for a quadratic approximation to (2.97), taking

f(T ) = −KT +k(T ) and g[k](T , t) =Bk as in the form of the linear analytic system (2.84).

Garret M. Flagg Chapter 2. Bilinear Systems 43

Note that since the lowest degree terms in k(T ) are of degree 2, ∇f(T )∣T=0 = −K. Let

a1A2` H = F2 be the matrix that collects together all the second partial terms in the Taylor

series expansion of f as in (2.83). The entries of H are defined as follows:

H1,1 = 2, H1,n+2 = −2

Hn,n2−n−1 = −2, Hn,n2 = 4

Hj,n(j−2)+j−1 = −2, Hj,n(j−1)+j = 4, Hj,nj+j+1 = −2, for j = 2, . . . , n − 1.

Since g[k] for k = 1,2 is constant, G[k]1,0 =Bk are the only nonzero terms in the Taylor series

expansion of g[k].

The bilinear realization of the quadratic system is therefore:

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−K a1A2` H

0 −K ⊗ In + In ⊗ −K

⎤⎥⎥⎥⎥⎥⎥⎦

Nk =

⎡⎢⎢⎢⎢⎢⎢⎣

0 0

Bk ⊗ In + In ⊗Bk 0

⎤⎥⎥⎥⎥⎥⎥⎦

for k = 1,2

B =

⎡⎢⎢⎢⎢⎢⎢⎣

B

0

⎤⎥⎥⎥⎥⎥⎥⎦

C

C is left unspecified because it depends on what will be measured for a given simulation. As

in the example for the constant heat flux input given above, frequently it is the temperature

at a given node, or the average temperature over some collection of the nodes.

Chapter 3

Model Reduction and Interpolation

Petrov-Galerkin projection and its connection with interpolation theory provides a powerful

theoretical framework for the model reduction of linear dynamical systems. Constructing a

low-order interpolant of the full-order transfer function requires solving shifted linear sys-

tems. Typically the realization for the full-order model is sparse, and so solving the shifted

linear systems can be done at a relatively low computational cost. If an optimal, or asymp-

totically optimal, collection of interpolation points can be determined, the cost of comput-

ing an accurate reduced order model can be dramatically decreased when compared with

grammian-based approaches to model reduction, which are highly accurate but require the

solution of the full-order system grammians in full-matrix arithmetic. For the H2 optimal

approximation of LTI systems, locally optimal reduced order models can be constructed via

interpolation using the Iterative Rational Krylov Algorithm (IRKA) of Gugercin, Antoulas

and Beattie [54]. If further information is known about the pole-distribution of the full-

order transfer function, asymptotically optimal interpolation methods have been proposed

by Druskin et. al in [39, 40, 41] . Recently Flagg, Beattie, and Gugercin showed that it is

possible to construct nearly optimal H∞ LTI system approximations starting from an ap-

44

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 45

proximation that is locally H2 optimal [45, 43]. Thus, interpolation-based model reduction

has a demonstrable track-record of producing computationally efficient algorithms that yield

high fidelity reduced models. For bilinear systems, computing the system grammians in or-

der to apply balanced truncation methods is even more costly than for LTI systems, making

it all the more important to develop an interpolation-based alternative. Interpolation-based

Petrov-Galerkin techniques for bilinear model reduction were first developed in [5], [6],[24],

[78],[34]. In this chapter we will first present the Petrov-Galerkin projection framework for

model reduction generally, and demonstrate how interpolation is accomplished in this frame-

work for LTI systems. We then consider two generalizations of interpolation-based model

reduction to bilinear systems.

3.1 The Petrov-Galerkin model reduction framework

Consider r-dimensional subspaces Vr and Wr of the full order state space. We wish to

construct an approximation x(t) ∈ Vr to the true state x(t) so that

˙x(t) −Ax(t) +Nx(t)u(t) − bu(t) ⊥Wr (3.1)

Let V ,W satisfying W T V = Ir be real n × r matrices whose columns form a basis for Vr

and Wr respectively. The Petrov-Galerkin approximation can be constructed by defining

x(t) = V xr(t) for some xr(t) ∈ Rr and enforcing

W T (V xr(t) −AV xr(t) +NV xru(t) − bu(t)) = 0 (3.2)

Enforcing this condition yields the order r bilinear system

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 46

xr(t) = W TAV xr(t) + W TNV xr(t)u(t) − W Tbu(t)

yr(t) = cV xr(t)(3.3)

In this framework, finding an accurate reduced order model is equivalent to finding accurate

projection subspaces Vr and Wr. Both interpolation and balancing methods are subsumed

under the Petrov-Galerkin framework. In the case of balanced truncation, this is readily

seen by considering the role of the balancing transformation T in the model reduction. After

balancing, ζ has the realization (T −1AT ,T −1NT ,T −1b,cTT ). The projection matrices that

yield a balanced truncation approximation are then given as W = (T Tr )−1, V = Tr, where

Tr is the first r columns of T and (T Tr )−1 is the first r columns of (T T )−1. Interpolation-

based model reduction methods explicitly define the subspaces Vr and Wr based on some

underlying function subspace, such as polynomials, as in the case of Krylov subspaces, or

rational functions as in the case of rational Krylov subspaces. In the next section we will

begin by briefly reviewing interpolation-based model reduction of linear systems, and then

present two natural, but distinct generalizations to bilinear systems.

3.2 Interpolation-based model reduction

LetH(s) ∈ Cp×m be the transfer function of an order n linear system. Given r right-tangential

point-direction pairs (σ1,r1), . . . , (σr,rr) where σj ∈ C and rj ∈ Rm and r left-tangential

point-direction pairs (µ1, `1), . . . , (µr, `r) with µj ∈ C and `j ∈ Rp, the rational tangential

interpolation problem is to construct Hr(s) of order at most r so that

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 47

`TjH(µj) = `TjHr(µj), for j = 1, . . . , r and, (3.4)

H(σj)rj =Hr(σj)rj, for j = 1, . . . , r. (3.5)

Interpolation via projection was first proposed by Skelton et al. [96, 103, 104]. Later,

Grimme [52] showed how to construct a reduced-order interpolant, using a method of Ruhe

[82]. Rational tangential interpolation for MIMO linear dynamical systems as it is presented

here was more recently developed by Gallivan et al. [51]

Theorem 3.1. Let Σ be a linear system with the realization (A,B,C). Given r right-

tangential point-direction pairs (σ1,r1), . . . , (σr,rr) and r left-tangential point-direction pairs

(µ1, `1), . . . , (µr, `r), construct V ,W ∈ Rn×r

so that

Range(V ) = span([(σ1I −A)−1Br1, . . . , (σrI −A)−1Brr]) (3.6)

Range(W ) = span([µ1I −AT )−1CT`1, . . . , (µ1I −A

T )−1C`r]) (3.7)

and W T V = Ir.

Define the reduced-order linear system ζr with the realization

A = W TAV Br = WTB Cr =CV (3.8)

Then ζr satisfies

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 48

`TjH(µj) = `TjHr(µj), for j = 1, . . . , r and, (3.9)

H(σj)rj =Hr(σj)rj, for j = 1, . . . , r. (3.10)

Moreover, if σj = µj for i = 1, . . . r, then

`TjH′(σj)rj = `

TjH

′r(σj)rj, for j = 1, . . . , r. (3.11)

Proof. [51]

Hr(σj)rj =CV (σjIr − A)−1Brrj

=CV (σjIr − A)−1W T (σjIn −A)(σjIn −A)−1Brj

=CV (σjIr − A)−1W T (σjIn −A)V uj, for some uj ∈ Rr

=CV (σjIr − A)−1(σjIr − A)uj

=CV uj

=C(σjIn −A)−1Brj

=H(σj)rj

The statement for the left-tangential vectors can be proved similarly. The statement con-

cerning derivatives is proved as follows

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 49

`TjH′r(σj)rj = `

Tj CV (σjIr − A)−2Brrj

= `Tj CV (σjIr − A)−1W T V (σjIr − A)−1Brrj

= qTj WT V uj, for some uj,qj ∈ Rr

= `Tj C(σjIn −A)−1(σjIn −A)−1Brj

= `TjH′(σj)rj

The construction of the interpolant in Theorem 3.1 underscores the central component of

all projection-based interpolation methods. The basic idea is to project the state onto the

subspace spanned by the system transfer function evaluated at some collection of frequencies,

whenever this makes sense. As we shall see, there are two different ways to push this strategy

forward in the case of bilinear systems. The first approach, which we will call subsystem

interpolation, is to place the interpolation information for a finite number of subsystems in

the span of the projection basis. Recall that for a bilinear system ζ the transfer function

for the kth order homogeneous subsystem in the frequency domain is a k-variate complex-

valued rational function. The total response can be computed by summing over all the

kth-order homogeneous subsystems. Assuming that the contributions to the response from

higher order subsystems is negligible, a good system approximation may be achieved by

matching the low-order subsystems, together with some specified number of their derivatives,

at several different points. The alternative is to enforce matching conditions on the whole

Volterra series. This latter approach is a kind of multipoint interpolation scheme that carries

information from point evaluations at every term in the Volterra series. This approach will be

considered in greater detail after subsystem interpolation, but we note here that the Volterra

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 50

series interpolation approach allows for a construal of the solution of the H2 optimal model

reduction problem for bilinear systems as an interpolation problem, and as such provides

insight into developing other interpolation strategies.

3.3 Subsystem Interpolation

The subsystem interpolation problem is posed in terms of the multimoments of the transfer

functions Hk(s1, . . . , sk).

Definition 3.1 (Multimoments). Let ζ be a bilinear system with realization (A, N1, . . . ,Nm,

B, C). Then for some point (σ1, . . . , σk) ∈ Ck, together with nonnegative integers m1, . . . ,mk,

a multimoment H(m1,...,mk)k (s1, . . . , sk) of the kth order transfer function Hk(s1, . . . , sk) is

defined as

H(m1,...,mk)k (s1, . . . , sk) =C(skI −A)−mkN [Im ⊗ (sk−1I −A)−mk−1](Im ⊗ N)⋯

⋅ [Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗(s2I −A)−m2](Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−2 times

⊗N) (3.12)

⋅ [Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗(s1I −A)−m1](Im ⊗⋯⊗ Im´¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¸¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¹¶

k−1 times

⊗B).

The subsystem interpolation problem was originally introduced by Phillips in [78] for SISO

bilinear systems as a method for matching moments around infinity of the transfer functions

for the kth order homogeneous subsystems. It was later extended to the problem of matching

moments around zero and infinity by Bai and Skoogh [6] , and finally extended to all of Cn by

Breiten and Damm [24]. The subsystem interpolation problem was then further generalized

to the case of multi-input-multi-output systems by Lin, Bao, and Wei [69] for multimoments

around the origin, and requires the construction of block Krylov subspaces. Here we will

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 51

present the most general case of tangential interpolation on the first through the kth order

homogenous subsystems of a bilinear system ζ defined by (2.1).

Theorem 3.2 (Subsystem Interpolation). Let ζ:=(A, N1, . . . ,Nm, B, C) be given, to-

gether with the sequences {σj}kj=1,{γj}kj=1 ⊂ C and vectors cT ∈ Cp and b ∈ Cm Define

bj = oj ⊗ b, and N⊕T = [NT1 , . . . ,N

Tm], where oj is a column of mj−1 ones. Let Kq(M ,x) =

span{x,Mx, . . . ,M q−1x} the standard polynomial Krylov subspace. To construct a reduced

order system ζ that matches all the multimoments H(l1,...,lj)j (σ1, . . . , σj)bj) and

cH(l1,...,lj)

k (γj, . . . , γ1) for j = 1, . . . , k and l1 . . . lj = 1, . . . , q construct the matrices V and W

as follows:

span{V (1)} =Kq((σ1I −A)−1, (σ1I −A)−1Bb) (3.13)

span{W (1)} =Kq((γ1I −A)−∗, (γ1I −A)−∗C∗c∗) (3.14)

span{V (j)} =Kq((σjI −A)−1, (σjI −A)−1N(Im ⊗V (j−1))) for j = 2, . . . , k (3.15)

span{W (j)} =Kq((γjI −A)−T , (γjI −A)−TN⊕T (Im ⊗W (j−1))) for j = 2, . . . , k (3.16)

spanV =span{k

⋃j=1

span{V (j)}} (3.17)

spanW =span{k

⋃j=1

span{W (j)}}. (3.18)

Provided W T = (W TV )−1W T is defined, the system Σr:=( A=W TAV , N=W TNV ,

C=CV , B=W TB) satisfies

H(l1,...,lj)

k (σ1, . . . , σj)bj = H(l1,...,lj)j (σ1, . . . , σj)bj

and

cH(l1,...,lj)j (γj, . . . , γ1) = cH

(l1,...,lj)j (γj, . . . , γ1),

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 52

for j = 1, . . . ,K and l1 . . . lk = 1, . . . , q.

The proof given here reduces to the proof given by Breiten and Damm in [24] for the SISO

result, so I will only include the novel portion.

Proof. Let b=Bb, and c = cC. Fix j ∈ {1, . . . , k}. Define the indexing set Ij =j

⨉`=1

{1,2, . . . , j},

so that each i ∈ Ij is a j-tuple with some combination of the integers 1, . . . , j. Finally, let

Mj =mj−1. In light of equation (2.25),

H(l1,...,lj)(σ1, . . . , σj)bj = ∑i∈Ij

C(σjI −A)−ljNi(j−1)⋯Ni(1)(σ1I −A)−l1 b (3.19)

cH(l1,...,lj)(γ1, . . . , γj) = [c(sjI −A)−1Ni1(j−1)⋯Ni1(1)(s1I −A)−1B, (3.20)

c(γjI −A)−1Ni2(j−1)⋯Ni2(1)(γ1I −A)−1B, . . . , (3.21)

c(γjI −A)−1NiMj (j−1)⋯NiMj (1)(γ1I −A)−1B] (3.22)

Therefore it is sufficient to show that for any i ∈ Ij,

(σjI −A)−ljNi(j−1) . . .Ni(1)(σ1I −A)−l1 b

= V (σjIr − A)−lkNi(j−1) . . . Ni(1)(σ1Ir − A)−l1W T b, and (3.23)

c(γ1I −A)−lkNi(k−1)⋯Ni(1)(γkI −A)−l1

= cV (γ1Ir − A)−ljNi(j−1)⋯Ni(1)(γjIr − A)−l1(W TV )−1W T (3.24)

The vectors

(σjI −A)−ljNi(j−1) . . .Ni(1)(σ1I −A)−l1 b and (c(γ1I −A)−ljNi(j−1)⋯Ni(1)(γjI −A)−l1)T

lie in the range of V andW by construction, so the the proof of equalities 3.23, 3.24 proceeds

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 53

precisely as the proof given by Breiten [24],[25] for the SISO case.

Assuming the subspaces are linearly independent and constructed from a total of ν sequences,

denoted σβ,1, . . . , σβ,N and µβ,1, . . . µβ,ν for β = 1, . . . , ν, the interpolating system ζr will have

dimension r = ν(p1 + 1) + 2ν(p1 + 1)(p2 + 1) + . . . + NνN

∏j=1

(pj + 1), which can clearly grow

large rather quickly. The main advantage of subsystem interpolation is that it does not

depend on convergence of the Volterra series, which in general may be very difficult to

establish. As it stands however, subsystem interpolation is unable to satisfy any known

optimality conditions. Nevertheless, it still makes sense to use this approach if the Volterra

series can be well approximated by truncating the series, as several examples we provide

later demonstrate. This approach also clearly introduces a large number of parameters

that must be determined in order to construct an approximation, and in the absence of a

theory of optimal subsystem interpolation, how to best choose the parameters is anyone’s

guess. In practice, most applications simply assume that interpolation of the first and second

subsystems is sufficient to provide a good characterization of the system, and this mostly

because it is simple.

3.4 Volterra Series Interpolation

Consider the response of a bilinear system ζ to the Heaviside function H(t). The Heaviside

function is defined as the step function

H(t) =

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

0 if t < 0

1 if t ≥ 0(3.25)

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 54

For the inputs uk(t) = κH(t) for k = 1, . . . ,m and t > 0 the bilinear system ζ can be written

as

ζ =

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

x(t) =Ax(t) + κm

∑k=1Nkx(t) +Bκom

y =Cx(t),

(3.26)

where om is a column of m ones, and therefore the output of ζ is equivalent to a that of the

linear system Σκ with realization (A+κm

∑k=1Nk,B, C). Hence, an accurate approximation of

the system for these inputs corresponds to accurately approximating its linear counterpart.

Let Hκ(s) =C(sI − (A+ κm

∑k=1Nk))

−1B be the transfer function of the system Σκ. Suppose

Σr is approximation to Σκ satisfying some rational tangential interpolation conditions at

points σj and along right and left directions rj and `j, respectively. Does this system satisfy

any interpolation properties interpreted in the context of the underlying bilinear system?

The interpolation conditions for the linear system are given as H(σj)rj =Hr(σj)rj, where

H(σj)rj =C(σjI −A − κm

∑k=1

Nk)−1Brj

=C(I + κ(σjI −A)−1m

∑k=1

Nk)−1(σjI −A)−1Brj. (3.27)

Assuming that ∥(κ(σjI −A)−1m

∑k=1Nk)∥2 < 1, (3.27) can be rewritten as the Neumann series

H(σj)rj =C∞

∑k=0

(κ(σjI −A)−1m

∑k=1

Nk)k(σjI −A)−1Brj (3.28)

=∞

∑k=0

κkHk(σj, . . . , σj)rj, (3.29)

where rj = om ⊗ om ⊗ rj.

Hence, interpolation of the linear system for the Heaviside function corresponds to matching

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 55

the frequency domain Volterra series weighted by κ along the sequences {σj, σj, σj, . . .}.

Moreover, considering the interpretation of the H2 norm as a weighted sum over all possible

combinations of point evaluations in the mirror image of the spectrum of A summed over all

homogeneous subsystems suggests that we may pose a more general multipoint interpolation

problem for which the reduced and full-order systems match on weighted sums of the Volterra

series kernels.

So consider the following interpolation problem. Given two sets of points σ1, σ2, . . . , σr to-

gether with a matrix U ∈ Rr×r and µ1, . . . , µr ∈ C together with a matrix S ∈ Rr×r, fix some

j ∈ {1,2, . . . , r} and define the weighted series

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

H(σl1 , σl2 , . . . , σj) <∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

H(µj, µl2 , . . . , µlk) <∞

where l1, l2, . . . , lk = 1, . . . , r. The weights ηl1,l2,...,lk−1,j

are given in terms of the entries of

U = {ui,j} as

ηl1,l2,...,lk−1,j

= uj,lk−1ulk−1,lk−2⋯ul2,l1 for k ≥ 2 and ηl1 = 1 for l1 = 1, . . . , r. (3.30)

For example, η1,2,3 = u3,2u2,1. Thus, the weights ηl1,l2,...,lk−1,j

are generated by multiplying

sequences of the entries of U together in the combinations determined by the index li.

The weights ηl1,l2,...,lk−1,j

are defined in the same way in terms of the entries of S. Note

that for the interpolation conditions in σj, sk = σj, whereas s1, . . . , sk−1 may take any other

value σ ∈ {σ1, . . . , σr} for all the transfer function evaluations in the series. Analogously,

for the interpolation conditions in µj, s1 = µj, whereas s2 . . . , sk−1 may take any other value

µ ∈ {µ1, . . . , µr} for all the transfer function evaluations in the series. Given this data, the

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 56

goal is to construct a reduced order system ζr ∶=(A, N , b, c) of order r so that for each

j = 1, . . . r

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(Hk(σl1 , σl2 , . . . , σj) − Hk(σl1 , σl2 , . . . , σj)) = 0 (3.31)

and

∑k=1

r

∑l2

r

∑l3

⋯r

∑lk

ηl1,l2,...,lk−1,j

(Hk(σj, σl2 , . . . , σlk) − Hk(σj, σl2 , . . . , σlk)) = 0 (3.32)

The solution to this problem is given as follows.

Theorem 3.3 (Volterra Series Interpolation). Let ζ ∶=(A, N , b, c) be a bilinear system of

order n. Suppose that for some r < n, points σ1, . . . , σr ∈ C and µ1, . . . , µr ∈ C, together with

U ,S ∈ Rr×r are given so that the series

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(σjI −A)−1N(σlk−1I −A)−1N⋯N(σl1I −A)−1b (3.33)

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(µjI −AT )−1NT (µlk−1I −A

T )−1NT⋯NT (µl1I −AT )−1cT

converge for each σj, and µj. Let Λ = diag(σ1, . . . , σr), and M = diag(µ1, . . . , µr) and let V ,

W ∈ Rn×r solve the generalized Sylvester equations

V Λ −AV −NV UT = beT (3.34)

WM −ATW −NTWST = cTeT .

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 57

If W T V ∈ Rr×r is invertible, then the reduced order model ζr of order r defined by

A = (W T V )−1W TAV , N = (W T V )−1W TNV ,

b = (W T V )−1W Tb, c = cV (3.35)

satisfies (3.31) for each σj, µj, j = 1, . . . , r.

Proof. We first show that the jth column of V is equivalent to (3.33). Let V (1) ∈ Rn×r solve

V (1)Λ −AV (1) = beT (3.36)

and for k ≥ 2, let V (k) ∈ Rn×r be the solution to

V (k)Λ −AV (k) =NV (k−1)UT (3.37)

Then V =∞

∑k=0V (k). Let vk,j denote the jth column of V (k). Then v1,j = (σjI −A)−1b, and

in general for k ≥ 2

vk,j = (σjI −A)−1fk−1,j (3.38)

where fk−1,j is the jth column of NV (k−1)U . We show by induction on k that

fk−1,j =r

∑lk−1

r

∑lk−2=1

⋯r

∑l1

ηl1,l2,...,lk−1,j

N(σlk−1I −A)1N(σlk−2I −A)−1N⋯N(σl1I −A)−1b (3.39)

So let k = 2. Then

f1,j =r

∑l1=1

uj,l1Nv

1,l1=

r

∑l1=1

ηl1,jN(σl1I −A)−1b. (3.40)

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 58

Now suppose the statement holds for k > 2. Then

vk,j = (σjI −A)−1fk−1 (3.41)

=r

∑lk−1

r

∑lk−2=1

⋯r

∑l1=1

ηl1,l2,...,lk−1,j

(σjI −A)−1N(σlk−1I −A)−1N⋯(σl1I −A)−1b (3.42)

and therefore

fk,j =r

∑lk

uj,lkNvk,lk (3.43)

=r

∑lk

r

∑lk−1

⋯r

∑l1

uj,lkηl1,l2,...,lk−1,lkN(σlkI −A)−1N⋯N(σl2I −A)−1b (3.44)

=r

∑lk

r

∑lk−1

⋯r

∑l1

ηl1,l2,...,lk−1,lk,jN(σlkI −A)−1N⋯N(σl2I −A)−1b (3.45)

Now define the skew projector P = V (W T V )−1W T . Then

P (V Λ −AV −NV U − beT ) =V (Λ − A − NU − beT ) = 0 (3.46)

Since V is full rank, it follows that Γ = Ir solves the projected Sylvester equation

ΓΛ − AΓ − NΓUT = beT .

By the same construction as above, the jth column of Γ can be represented as

γj =∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(σjIr − A)−1N(σlk−1Ir − A)−1N⋯N(σl1Ir − A)−1b

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 59

And therefore

V γj = vj =∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

V (σjIr − A)−1N(σlk−1Ir − A)−1N⋯N(σl1Ir − A)−1b

(3.47)

=∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(σjI −A)−1N(σlk−1I −A)−1N⋯N(σl1I −A)−1b (3.48)

Multiplying equation (3.47) on the left by c gives the desired result in terms of the inter-

polation conditions on σj. For the interpolation conditions in the points µj, observe that

precisely the same construction of the columns of W follows from the proof given above

applied to the equation

WM −ATW −NTWS = cTo

Now P T = W (V TW )−1V T is a skew projection onto the range of W , and

P T (WM −ATW −NTWS − cToT ) (3.49)

= W (V TW )−1((V TW )M − AT (V TW ) − NT (V TW )S − cToT ) (3.50)

= 0 (3.51)

Since W (V TW )−1 is full rank, this implies that Ξ = V TW ∈ Rr×r solves

ΞM − ATΞ − NTΞST − cToT = 0

Again, by the construction given above, the columns ξj ∈ Rr of Ξ for j = 1, . . . , r can be

represented as

Garret M. Flagg Chapter 3. Model Reduction and Interpolation 60

ξj =∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(µjIr − AT )−1NT (µlk−1Ir − A

T )−1NT⋯NT (µl1Ir − AT )−1cT

And therefore

W (V TW )−1ξj = wj (3.52)

=∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

W (V TW )−1(µjIr − AT )−1NT (µlk−1Ir − A

T )−1NT⋯ (3.53)

NT (µl1Ir − AT )−1cT (3.54)

=∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

(µjI −AT )−1NT (µlk−1I −A

T )−1NT⋯NT (µl1I −AT )−1cT

(3.55)

for j = 1, . . . , r. Taking the transpose of these equations and multiplying on the right by b

yields the desired result for the interpolation points in µj and weights in S.

Chapter 4

H2 Optimal Model Reduction

In this chapter two different approaches to solving the H2 optimal model reduction problem

are presented. The first approach aims to generalize the structured-orthogonality conditions

first introduced by Gugercin, Beattie, and Antoulas in [54] to the case of bilinear systems.

For LTI systems they showed that certain Hilbert-space orthogonality conditions provided

a unifying framework for all previously derived first-order necessary conditions for the H2

optimal model-reduction problem. The second approach is to directly derive an expression

for the H2 error that can be differentiated with respect to the reduced-order model param-

eters. This approach is carried out in [105] and [20]. For LTI systems both approaches

are equivalent, but in the bilinear case our results show that they are not. Interestingly

enough, the fact that these approaches are not equivalent is grounded in the fact that satis-

fying structured orthogonality conditions would require a subsystem interpolation approach,

whereas the Breiten-Benner [20] and Zhang-Lam [105] conditions require satisfying Volterra

series interpolation conditions. First, consider the structured-orthogonality conditions for

LTI systems summarized in the following theorem.

61

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 62

Theorem 4.1 (Structured Orthogonality conditions). [54] If Hr has simple poles and sat-

isfies ∥H −Hr∥H2 ≤ ∥H − H(ε)r ∥H2 for any order r system H

(ε)r satisfying ∥Hr − H

(ε)r ∥ ≤ Cε,

then

⟨H −Hr,Hr ⋅H1 +H2⟩H2 = 0 (4.1)

for all real dynamical systems H1, H2 having simple poles at the same locations as Hr.

This inner-product formulation of the optimality conditions can be used to derive the Meier-

Luenberger conditions [71], as well as the Wilson conditions for H2 optimality [101]. In this

sense, they constitute a unifying framework for the H2 model reduction problem in the linear

case, which is the motivation for extending these results to the bilinear case. The Meier-

Luenberger conditions stated here provide a foreshadowing of the type of interpolation-based

necessary conditions we aim to derive for the bilinear case. Here we present them in their

generalized form for MIMO LTI-systems, although they were originally derived for SISO

LTI-systems.

Theorem 4.2 (Meier-Luenberger conditions). [54] Let Σ be an LTI system identified with

its transfer function H(s). Suppose that H(s) =r

∑j=1

1s−λj

cj bTj with bj ∈ Rm, cj ∈ Rp is a

locally H2 optimal approximation to H(s). Then

1.) cTjH(−λj) = cjH(−λj)

2.) H(−λj)bj = H(−λj)bj

3.) cTjH′(−λj)bj = cTj H

′r(−λj)bj, for j = 1, . . . , r.

Recently Gugercin, Beattie, and Antoulas introduced the Iterative Rational Krylov Algo-

rithm (IRKA) as a interpolation-based method for constructing reduced-order models that

are H2 optimal [54]. The algorithm is outlined below.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 63

Algorithm 4.1 (Iterative Rational Krylov Algorithm). [54]

1. Make an initial r-fold shift selection: {σ1, . . . , σr} that is closed under conjugation (i.e.

{σ1, . . . , σr} = {σ1, . . . , σr}) and initial tangent directions b1, . . . , br and c1, . . . , cr, also

closed under conjugation.

2. V = [(σ1I −A)−1Bb1 . . . (σrI −A)−1Bbr]

W = [(σ1I −AT )−1CT c1 . . . (σrI −AT )−1CT cr]

3. While (not converged)

a. A = W TAV , E = W T V , B = W TB, C =CV .

b. Compute Y∗AX = diag(λi) and Y

∗EX = Ir, where Y

∗and X are the left and right

eigenvectors of λE − A.

c. σj ←Ð −λj(A, E) for j = 1, . . . , r, b∗j ←Ð ejY∗B and cj ←Ð CXej.

d. V = [(σ1I −A)−1Bb1 . . . (σrI −A)−1Bbr]

W = [(σ1I −AT )−1CT c1 . . . (σrI −AT )−1CT cr]

4. A = (W T V )−1W TAV , B = (W T V )−1W TB, C =CV

The Wilson conditions forH2 optimality are given in terms of conditions on the controllability

and observability grammians of the error system denoted Pe, Qe respectively. Here we

present them in the general form derived by Zhang and Lam [105]. The Wilson conditions

for LTI systems are recovered by simply taking N = 0 in the equations below.

Theorem 4.3 (Generalized Wilson Conditions [105]). Suppose ζr is a locally optimal ap-

proximation to ζ in the H2 norm. Let Pe, Qe be the grammians of ζ − ζr, and partition

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 64

Pe,Qe conformably with the dimensions of A, and A as:

Pe =

⎡⎢⎢⎢⎢⎢⎢⎣

P P0

PT0 Pr

⎤⎥⎥⎥⎥⎥⎥⎦

Qe =

⎡⎢⎢⎢⎢⎢⎢⎣

Q Q0

QT0 Qr

⎤⎥⎥⎥⎥⎥⎥⎦

Then

QT0P0 +QrPr = 0, QrNPr +QT

0NP0 = 0

QT0B +QrB = 0, CPr −CP0 = 0

In order to carry the structured-orthogonality approach forward to the bilinear case, let us

first consider how the analysis will carry forward for SISO bilinear systems. First, define the

Hilbert space

F =∞

⊕k=1

L2((ıR)k).

The inner product on F is defined as the sum of the inner products on each of the com-

ponents, and the elements of F consist of all sequences of functions for which the bilin-

ear H2 norm is finite. In particular, this contains all H2 bilinear systems. Let ⊕H =

(H1(s1),H2(s1, s2),H3(s1, s2, s3), . . . ) be any sequence of proper rational functions of the

form Hk(s1, s2, . . . , sk) =Pn(s1,s2,...,sk)

Qn(s1)Qn(s2)⋯Qn(sk), and assume that ⊕H ∈ F . Let λ1, . . . , λn be

the zeros of Qn, and fix a k ∈ N. Now let the set Jk index the elements of the Cartesian

product Lk =k

⨉i=1

{λ1, . . . , λn}. Lk contains all possible k-tuple combinations of the zeros

of Qn. With each tuple λj ∈ Lk, associate the quantity φk,j as in Definition 2.10. Now

let H{σ1,...,σr} be the subspace of F for which each element ⊕H = (H1(s1), H2(s1, s2), . . . )

satisfies Hk(s1, . . . , sk) = Pr(s1, s2, . . . , sk)/Qr(s1)Qr(s2)⋯Qr(sk), is a strictly proper ra-

tional function and Qr(s`) is a polynomial of degree r with simple zeros at the points

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 65

σ1, . . . , σr ∈ C. Let Irk index the elements σi of the Cartesian product Sk =k

⨉i=1

{σ1, . . . , σr}.

For any ⊕Hr ∈ H{σ1,...,σr} let φk,i’s be defined for Hk(s1, . . . , sk) also as in Definition 2.10.

Theorem 4.4. If ⊕H⋆ is the optimal approximation to ⊕H out of H{σ1,...,σr}, then for all

k ∈ N and each σi ∈ Sk, ⊕H⋆ satisfies

Hk(−σi(1),−σi(2), . . . ,−σi(k)) = H⋆k (−σi(1),−σi(2), . . . ,−σi(k)) (4.2)

Proof. H{σ1,...,σr} is a closed subspace of F , so by the standard Hilbert projection theorem

⊕H − ⊕H⋆ ⊥ H{σ1,...,σr}. So for any ⊕G ∈ H{σ1,...,σr}

⟨⊕H − ⊕H⋆,⊕G⟩F =

∑k=1

⟨Hk(s1, s2, . . . , sk) − H⋆k (s1, s2, . . . , sk),Gk(s1, s2, . . . , s∞)⟩L2(Rn)

=∞

∑k=1

⎡⎢⎢⎢⎢⎣

∑j∈J n

k

βk,j(H(−σj(1),−σj(2), . . . ,−σj(k)) − H(−σj(1),−σj(2), . . . ,−σj(k)))

⎤⎥⎥⎥⎥⎦

(4.3)

= 0 (4.4)

For each k, there are rk residues βk,j determined by Gk. So expression 4.3 follows from the

pole-residue expansion of Gk, together with an application of the residue calculus in each

variable, as in the proof of Theorem 2.10. Provided the residues βk,j of Gk are chosen small

enough to guarantee convergence, each residue can be chosen to guarantee that the each

term in the sum (4.3) is greater than zero. So to force the sum to 0 over all choices of the

residues of the functions Gk, we must have that

Hk(−σi(1),−σi(2), . . . ,−σi(k)) = H⋆k (−σi(1),−σi(2), . . . ,−σi(k)) (4.5)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 66

For a given choice of reduced-order poles, it is possible to construct a sequence of ratio-

nal functions which satisfy the necessary interpolation conditions given by (4.5). However,

in general it is not be possible to find a finite dimensional bilinear realization for this se-

quence, yielding an infinite dimensional solution to the model reduction problem! This is

because there is not a one-to-one correspondence between a minimal bilinear realization and

a sequence of recognizable rational functions. Indeed, it is not possible to carry forward

the structured orthogonality conditions in general, because one can readily show that the

set of r-dimensional bilinear realizations having poles at σ1, . . . , σr is not a closed subspace

of F , although it is certainly a subset of H{σ1,...,σr}. It is worth noting, however, that for

any truncated Volterra series, it is possible to construct a finite dimensional realization of

H⋆(s1, s2, . . . , s∞) which satisfies all of the interpolation conditions.

Now that we have seen that it is not possible to generalize the structured-orthogonality

conditions to the bilinear model reduction problem we return to the more standard approach,

which is to write out the H2 norm of the error system in terms of the realization parameters

A, Nk, B, C, and A, B, Nk C and then differentiate the resulting expression with respect

to the reduced-order model parameters. The next theorem, due to Breiten and Benner [20],

shows how to write the H2 error so that computable necessary conditions can be derived.

Theorem 4.5. [20] Let ζ ∶= (A,N1, . . . ,Nm,B,C) of order n be approximated by a system

ζr ∶= (A, N1, . . . , Nm, B, C) of order r < n. Then

∥ζ − ζr∥2H2

= vec(Ip)T ([C −C]⊗ [C −C])× (4.6)

( −

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−m

∑k=1

⎡⎢⎢⎢⎢⎢⎢⎣

Nk 0

0 NTk

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Nk 0

0 Nk

⎤⎥⎥⎥⎥⎥⎥⎦

)−1

×

(

⎡⎢⎢⎢⎢⎢⎢⎣

B

BT

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

B

B

⎤⎥⎥⎥⎥⎥⎥⎦

)vec(Im) (4.7)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 67

where RΛR−1 is the spectral decomposition of A, and B = BTR−T , C = cR, Nk =

RTNkR−T .

Differentiating this expression with respect to the parameters Λ, B, C, and Nk yields the

following necessary conditions for H2 optimality.

Theorem 4.6 (Necessary conditions for H2 optimality). [20] Suppose ζr is a locally H2

optimal approximation of order r to the full order system ζ. Let RΛR−1 be the spectral

decomposition of A, and let B = BTR−T , C = CR, Nk = RT (N)TR−T for k = 1, . . . ,m.

Then ζr satisfies the following conditions:

vec(Ip)T (eie

Tj ⊗C)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(BT ⊗B)vec(Im)

= vec(Ip)T (eie

Tj ⊗ C)( − Λ⊗ Ir − Ir ⊗ A −

m

∑k=1

(Nk)T ⊗ Nk)

−1

(BT ⊗ B)vec(Im), (4.8)

vec(Ip)T (C ⊗C)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(ejeTi ⊗B)vec(Im)

= vec(Ip)T (C ⊗ C)( − Λ⊗ Ir − Ir ⊗ A −

m

∑k=1

(Nk)T ⊗ Nk)

−1

(ejeTi ⊗ B)vec(Im), (4.9)

vec(Ip)T (C ⊗C)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(eieTi ⊗ In)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(BT ⊗B)vec(Im)

=vec(Ip)T (C ⊗ C)( − Λ⊗ Ir − Ir ⊗ A −

m

∑k=1

(Nk)T ⊗ Nk)

−1

(eieTi ⊗ In)( − Λ⊗ Ir − Ir ⊗ A −

m

∑k=1

(Nk)T ⊗ Nk)

−1

(BT ⊗ B)vec(Im), and (4.10)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 68

vec(Ip)T (C ⊗C)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(ejeTi ⊗ N)( − Λ⊗ In − Ir ⊗A −

m

∑k=1

NTk ⊗Nk)

−1

(BT ⊗B)vec(Im)

=vec(Ip)T (C ⊗ C)( − Λ⊗ Ir − Ir ⊗ A −

m

∑k=1

(Nk)T ⊗ Nk)

−1

(ejeTi ⊗

¯N)( − Λ⊗ Ir − Ir ⊗ A −m

∑k=1

(Nk)T ⊗ Nk)

−1

(BT ⊗ B)vec(Im) (4.11)

The necessary conditions given in Theorem 4.6 can be used to generalize the Iterative Ra-

tional Krylov Algorithm (IRKA) to the bilinear case. The algorithm below, developed by

Breiten and Benner describes the Bilinear Iterative Rational Krylov Algorithm (B-IRKA).

Algorithm 4.2 (B-IRKA). [20]

Input: A, Nk for k = 1 . . .m, B, C, A, Nk for k = 1, . . . ,m, B, C

Output: Aopt, N optk for k = 1, . . . ,m, Bopt, Copt

1. While: Change in Λ > ε do:

2. RΛR−1 = A, B = BTR−T , C =CR, Nk =RTNTk R

−T for k = 1, . . . ,m.

3. Solve

V (−Λ) −AV −m

∑k=1

NkV NTk =BBT

and

W (−Λ) −AT V −m

∑k=1

NTk W NT

k =CT C

4. V = orth(V ), W = orth(W ).

5. A = (W T V )−1W TAV , Nk = (W T V )−1W TNkV for k = 1, . . . ,m,

B = (W T V )−1W TB, C =CV .

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 69

6. end while

7. Aopt = A, N optk = Nk for k = 1, . . . ,m, Bopt = B, Copt = C

By setting Nk = 0 for k = 1, . . . ,m, B-IRKA reduces to the Sylvester equation formulation

of IRKA.

We now present an analysis of the necessary conditions of Theorem 4.6 which connects

them with our new multipoint Volterra series interpolation scheme. Our analysis shows that

the Breiten-Benner necessary conditions construed in terms of multipoint Volterra series

interpolation yields rather satisfying generalizations of the Meier-Luenberger conditions, in

the sense that they describe interpolation conditions that can be characterized completely

in terms of the poles and residues of the reduced-order subsystems.

In order to obtain this result, we first prove the following lemma, which clarifies the relation-

ship between the multi-point Volterra series interpolation conditions and the pole residue

expansion of a bilinear system.

Lemma 4.1. Let the SISO bilinear system ζ have the realization A, N , b, c of order n. Let

ζr be a bilinear system of order r < n with realization A, N , b, c. Let RΛR−1 be the spectral

decomposition of A, and let B = BTR−T , C = CR, N = RT (N)TR−T . Moreover, let the

residues φl1,...,lk for k = 1, . . . ,∞ and lk = 1, . . . , r of the transfer functions Hk(s1, . . . , sk)

corresponding to the kth order homogeneous subsystems of ζr be defined as in Definition

2.10. Let V

V (−Λ) −AV −NV NT = bbT

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 70

Then

c(cV )T =∞

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

(4.12)

Proof. Let U = N , r = b and σj = −λj for j = 1, . . . , r. By applying the construction of the

columns of V given in the proof of Theorem 3.3, we have that

cV (∶, j) =∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

bl1Hk(−λl1 ,−λl2 , . . . ,−λj)

(4.13)

where recall that ηl1,...,lk−1,j = uj,lk−1ulk−1,lk−2⋯ul2,l1 for k ≥ 2 and ηl1 = 1 for l1 = 1, . . . , r. Now

for each j = 1, . . . , r, observe that by the definition of ηl1,...,lk−1,j, for k ≥ 2

ηl1,...,lk−1,j bl1 = N(j, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1 (4.14)

And therefore

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1

ηl1,l2,...,lk−1,j

bl1Hk(−λl1 ,−λl2 , . . . ,−λj)

=∞

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(j, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λj)

(4.15)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 71

for j = 1, . . . , r. Hence,

(cV )T =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(1, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λ1)

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(2, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λ2)

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(r, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λr)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

Now apply the residue derivation given in (2.72) in the obvious way to the product

[c1 c2 . . . , cr]

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(1, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λ1)

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(2, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λ2)

∑k=1

r

∑l1

r

∑l2

⋯r

∑lk−1N(r, lk−1)N(lk−1, lk−2)⋯N(l2, l1)bl1Hk(−λl1 ,−λl2 , . . . ,−λr)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(4.16)

to obtain

=∞

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

Using Lemma 4.12, we now show that the H2 optimal necessary conditions are equivalent to

multipoint Volterra series interpolation conditions with weights given by the reduced order

residues and interpolation points by the reflection of the poles of the reduced order transfer

functions across the imaginary axis.

Theorem 4.7. Let ζ be a SISO system of order n and suppose that ζ ∈ F . Let ζr =

(A, N , b, c) be an H2 optimal approximation of order r. Then ζr satisfies the following

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 72

multipoint Volterra series interpolation conditions.

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

=∞

∑k=1

r

∑l1=1

⋯r

∑lk−1=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk), (4.17)

and

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λk))

=∞

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λlk)) (4.18)

where φl1,...,lk

, and λli are the residues and poles of the transfer functions Hk associated with

ζr.

Proof. Let RΛR−1 be the spectral decomposition of A, and let b = bTR−T , c = cR, N =

RTNTR−T . Moreover, suppose that V and W solve

V (−Λ) −AV −NV NT = bbT (4.19)

W (−Λ) −AT V −NTW NT = cT c (4.20)

By applying the vec operator to equations (4.19) and (4.20), we have that

vec(V ) = ( − Λ⊗ In − Ir ⊗A − NT ⊗N)

−1

(bT ⊗ b).

Thus,

(eTj ⊗ c)vec(V ) = cV (∶, j)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 73

is equivalent to the left-hand side of necessary condition (4.8). By applying Lemma 4.12 to

both sides of (4.8) gives

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

=∞

∑k=1

r

∑l1=1

⋯r

∑lk−1=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

The second equality (4.18) follows from condition (4.10). Simple algebra shows that the right-

hand-side of equality (4.10) is equivalent to the product W (∶, j)T V (∶, j). This is equivalent

to

(∞

∑k=1

r

∑l1=1

⋯r

∑lk−1=1

cl1ηj,lk−1,...,l1c(−λl1In −A)N⋯N(−λlk−1)In −A)−1N(−λjIn −A)−1)⋅

(∞

∑k=1

r

∑r1=1

⋯r

∑rk−1=1

ηr1,...,rk−1,j br1(−λjIn −A)−1N(−λrk−1In −A)−1N⋯N(−λr1In −A)−1b).

(4.21)

Expanding over the first few terms in k is sufficient to establish the general pattern:

W (∶, j)T V (∶, j) = cj bjc(−λjIn −A)−2b +r

∑r1=1

cjηr1,j br1(c(−λjIn −A)−2N(−λr1In −A)−1b

+r

∑l1=1

cl1ηj,l1 bjc(−λl1In −A)−1N(−λjIn −A)−2b)+

+r

∑r1=1

r

∑r2=1

cjηr1,r2,j br1(c(−λjIn −A)−2N(−λr2In −A)−1N(−λr1In −A)−1b

+r

∑l1=1

r

∑l2=1

cl1ηj,l2,l1 bj(c(−λl1In −A)−1N(−λl2In −A)−1N(−λjIn −A)−2b

+r

∑l1=1

r

∑r1=1

cl1ηj,l1ηr1,j br1(c(−λl1In −A)−1N(−λjIn −A)−2N(−λr1In −A)−1b

+ . . . , (4.22)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 74

where the weights ηr1,r2,j

, ηj,l2,l1

etc. are defined in (3.30), and the indices in r and l keep

track of the cases where terms on the right are multiplied by terms on the left and vice versa,

respectively. The expansion of the product for the solution of the reduced order matrices

follows similarly. Thus,r

∑j=1W (∶, j)T V (∶, j) gives the desired expression for the derivatives

as:

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λk))

=∞

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λlk)).

Since all the terms j are equal on both sides of equation (4.10), the second result follows.

Enforcing the multipoint interpolation conditions required by Theorem 4.6 or alternatively

Theorem 4.7 requires solving the bilinear Sylvester equations given in Step 3. of B-IRKA.

The direct solution of these equations has O((nr)2) complexity. This means that as r grows

even moderately large, between say 10 and 30, the computational cost will be large per

iteration of B-IRKA. Theorem 4.7 suggests a new and inexpensive, asymptotically optimal

alternative. Instead of matching the entire series and all its first partials at the mirror image

of the reduced order eigenvalues, it is possible to construct approximants ζr that satisfy

the first-order necessary conditions over the first N terms in the series. Moreover, we will

show that this approximation is optimal in an appropriately defined generalization of the H2

norm for bilinear systems. To see this, let us first consider polynomial systems generated by

truncating the Volterra series of a bilinear system.

Definition 4.1. Given a SISO bilinear system ζ with realization (A, N , b, c), define the

polynomial system ζN to be the operator mapping inputs u(t) to ouputs y(t) determined by

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 75

the following input-output mapping

y(t) =N

∑k=1∫

t

0∫

σ1

0⋯∫

σk−1

0ceA(t−σ1)NeA(σ1−σ2)N

⋯NeA(σk−1−σk)bu(σk)u(σk−1)⋯u(σ1)dσk⋯dσ1

Note that a polynomial system can also be identified with its sequence of transfer functions as

ζN ≡ (H1(s1),H2(s1, s2), . . . ,HN(s1, . . . , sN)) where Hk(s1, . . . , sk) = c(skI−A)−1N⋯N(s1I−

A)−1b, for k = 1, . . .N .

The H2 norm for polynomial systems ζN , denoted HN2 is defined in the obvious way.

Definition 4.2. Let ζN be a polynomial system determined generated by the bilinear system

ζ with realization (A, N , b, c). Then

∥ζN∥HN2=

¿ÁÁÁÀ

N

∑k=1

∫0

∫0

∣hk(t1, . . . , tk)∣2dtk⋯dt1 (4.23)

Enforcing the multi-point interpolation conditions of Theorem 4.7 on the first N terms is in

fact equivalent to constructing an HN2 optimal approximation of the polynomial system ζN .

Let ζNr be a polynomial reduced order model generated by the bilinear system ζr with

realization (A, N , b, c). We will now derive necessary conditions for HN2 optimality in

terms of the realization parameters of ζr.

From Theorem 2.10,

∥ζN − ζNr ∥2H2

=N

∑k=1

n

∑l1=1

⋅ ⋅ ⋅n

∑lk=1

φl1,...,lk

(Hk(−λl1 , . . . ,−λlk) − Hk(−λl1 , . . . ,−λlk))

−n

∑l1=1

⋅ ⋅ ⋅n

∑lk=1

φl1,...,lk

(Hk(−λl1 , . . . ,−λlk) − Hk(−λl1 , . . . ,−λlk)). (4.24)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 76

In terms of the realization parameters of ζ and ζr,

∥ζN − ζNr ∥2 = [c −c](N

∑k=1

Pk)

⎡⎢⎢⎢⎢⎢⎢⎣

c

−c

⎤⎥⎥⎥⎥⎥⎥⎦

, (4.25)

where P0 solves⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

P0 −

⎡⎢⎢⎢⎢⎢⎢⎣

AT 0

0 −ΛT

⎤⎥⎥⎥⎥⎥⎥⎦

P0 =

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

[bT bT ]

and for k > 0, Pk solves

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

Pk −Pk

⎡⎢⎢⎢⎢⎢⎢⎣

AT 0

0 −ΛT

⎤⎥⎥⎥⎥⎥⎥⎦

=

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

Pk−1

⎡⎢⎢⎢⎢⎢⎢⎣

NT 0

0 NT

⎤⎥⎥⎥⎥⎥⎥⎦

This follows as a straightforward application of the construction of the columns of Pk anal-

ogous to the construction given in proof of Theorem 3.3. Applying the vec operator to the

Lyapunov equation for P0 one can write

vec(P0) =⎛

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

−1 ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

(4.26)

and

vec(Pk) =⎛

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

−1 ⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

vec(Pk−1)

(4.27)

Applying the vec operator to the sum (4.25) and successively substituting the expressions

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 77

(4.26) and (4.27) into the sum gives

EN =∥ζN − ζNr ∥2

=([c − c]⊗ [c − c])N

∑k=0

⎡⎢⎢⎢⎢⎣

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(4.28)

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

The differentiation of EN with respect to the optimization parameters is greatly simplified

by using the following lemma, first derived in [20].

Lemma 4.2 ([20]). Let C(x) ∈ Rp×n, A(y),Gk ∈ Rn×n, and K ∈ Rn×m with

L(y) = −A(y)⊗ I − I ⊗A(y) −m

∑k=1

Gk ⊗Gk

and assume that C and A are differentiable with respect to x, and y. Then

∂x[(vec(Ip))

T (C(x)⊗C(x))L(y)−1(K ⊗K)vec(Im)

= 2(vec(Ip))T (

∂xC(x)⊗C(x))L(y)−1(K ⊗K)vec(Im)

and

∂y[(vec(Ip))

T (C(x)⊗C(x))L(y)−1(B ⊗B)vec(Im)]

= 2(vec(Ip))T (C(x)⊗C(x))L(y)−1(

∂yA(y)⊗ I)L−1(y)(K ⊗K)vec(Im).

Another important tool for analyzing the resulting expressions for the derivative of E is the

permutation matrix

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 78

M =

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In

0

⎤⎥⎥⎥⎥⎥⎥⎦

Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

0T

Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎥⎥⎦

(4.29)

first introduced in [20]. Given matrices H ,K ∈ Rr×r and L ∈ Rn×n, the permutation M has

the following property:

MT

⎛⎜⎜⎝

HT ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

L 0

0 K

⎤⎥⎥⎥⎥⎥⎥⎦

⎞⎟⎟⎠

M

= [Ir ⊗ [In0T ] Ir ⊗ [0 Ir]]

⎛⎜⎜⎝

HT ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

L 0

0 K

⎤⎥⎥⎥⎥⎥⎥⎦

⎞⎟⎟⎠

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In

0

⎤⎥⎥⎥⎥⎥⎥⎦

Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

0T

Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎥⎥⎦

= [Ir ⊗ [In0T ] Ir ⊗ [0 Ir]]

⎡⎢⎢⎢⎢⎢⎢⎣

HT ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

L

0

⎤⎥⎥⎥⎥⎥⎥⎦

HT ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

0T

K

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎥⎥⎦

=

⎡⎢⎢⎢⎢⎢⎢⎣

HT ⊗L 0

0 HT ⊗K

⎤⎥⎥⎥⎥⎥⎥⎦

Differentiating EN with respect to the parameters Λ, N , b, c and making use of Lemma 4.2

(taking Gk = 0 for k = 1, . . . ,m and K =N in Lemma 4.2) and the permutation M gives

∂EN∂cj

=2([0,−ej]⊗ [c − c])N

∑k=0

⎡⎢⎢⎢⎢⎣

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(4.30)

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 79

=2([0,−ej]⊗ [c − c])N

∑k=0

⎡⎢⎢⎢⎢⎣

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 −Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

=2( − ej ⊗ [c − c])N

∑k=0

⎡⎢⎢⎢⎢⎣

⎝− Λ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

− Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

N ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k

b⊗

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

=2( − ej ⊗ [c − c])MTN

∑k=0

⎡⎢⎢⎢⎢⎣

⎡⎢⎢⎢⎢⎢⎢⎣

−Λ⊗ In 0

0 −Λ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗A 0

0 Ir ⊗ A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N ⊗N 0

0 N ⊗ N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k

M⎛

⎝b⊗

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

=(−2ej ⊗ c)N

∑k=0

[( − Λ⊗ In − Ir ⊗A)

−1

N ⊗N]

k

(b⊗ b)

+ 2(ej ⊗ c)N

∑k=0

[( − Λ⊗ Ir − Ir ⊗ A)

−1

N ⊗ N]

k

(b⊗ b) (4.31)

∂EN

∂λj=([c − c]⊗ [c − c])

N

∑k=0

k−1

∑j=0

⎡⎢⎢⎢⎢⎣

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

j⎛

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(

⎡⎢⎢⎢⎢⎢⎢⎣

0 0

0 eieTi

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

)⎛

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

−1

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 80

⎡⎢⎢⎢⎢⎣

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k−j−1 ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

=([c − c]⊗ [c − c])N

∑k=0

k−1

∑j=0

⎡⎢⎢⎢⎢⎣

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

j⎛

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(

⎡⎢⎢⎢⎢⎢⎢⎣

0 0

0 eieTi

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

)⎛

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎣

⎝−

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 Λ

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k−j−1 ⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

=( − c⊗ [c − c])N

∑k=0

k−1

∑j=0

⎡⎢⎢⎢⎢⎣

⎝− Λ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

− Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

N ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

j

⋅⎛

⎝− Λ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

− Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(eieTi ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

)

⎝− Λ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

− Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1⎡⎢⎢⎢⎢⎣

⎝− Λ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

In 0

0 Ir

⎤⎥⎥⎥⎥⎥⎥⎦

− Ir ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

A 0

0 A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⋅ N ⊗

⎡⎢⎢⎢⎢⎢⎢⎣

N 0

0 N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k−j−1

(b⊗

⎡⎢⎢⎢⎢⎢⎢⎣

b

b

⎤⎥⎥⎥⎥⎥⎥⎦

)

=( − c⊗ [c − c])MTN

∑k=0

k−1

∑j=0

⎡⎢⎢⎢⎢⎣

⎡⎢⎢⎢⎢⎢⎢⎣

−Λ⊗ In 0

0 −Λ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗A 0

0 Ir ⊗ A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 81

⎡⎢⎢⎢⎢⎢⎢⎣

N ⊗N 0

0 N ⊗ N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

j⎛

⎡⎢⎢⎢⎢⎢⎢⎣

−Λ⊗ In 0

0 −Λ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗A 0

0 Ir ⊗ A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

(

⎡⎢⎢⎢⎢⎢⎢⎣

eieTi ⊗ In 0

0 eieTi ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

)⎛

⎡⎢⎢⎢⎢⎢⎢⎣

−Λ⊗ In 0

0 −Λ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗A 0

0 Ir ⊗ A

⎤⎥⎥⎥⎥⎥⎥⎦

−1

⎡⎢⎢⎢⎢⎣

⎡⎢⎢⎢⎢⎢⎢⎣

−Λ⊗ In 0

0 −Λ⊗ Ir

⎤⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎣

Ir ⊗A 0

0 Ir ⊗ A

⎤⎥⎥⎥⎥⎥⎥⎦

−1 ⎡⎢⎢⎢⎢⎢⎢⎣

N ⊗N 0

0 N ⊗ N

⎤⎥⎥⎥⎥⎥⎥⎦

⎤⎥⎥⎥⎥⎦

k−j−1 ⎡⎢⎢⎢⎢⎢⎢⎣

b⊗ b

b⊗ b

⎤⎥⎥⎥⎥⎥⎥⎦

= − (c⊗ c)N

∑k=1

k−1

∑j=0

[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j

(−Λ⊗ In − Ir ⊗A)−1(eieTi ⊗ In)

(−Λ⊗ In − Ir ⊗A)−1[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j−1

b⊗ b

+ (c⊗ c)N

∑k=1

k−1

∑j=0

[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

j

(−Λ⊗ Ir − Ir ⊗ A)−1(eieTi ⊗ Ir)

(−Λ⊗ In − Ir ⊗ A)−1[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

k−j−1

b⊗ b (4.32)

Simplifying these bloated expressions for the other derivatives requires exactly the same

kinds of steps as in simplifying the derivative of E with respect to the parameters in c and

λi, so we omit the derivations here. The resulting expressions are

∂EN

∂bj=(−2c⊗ c)

N

∑k=0

[( − Λ⊗ In − Ir ⊗A)

−1

N ⊗N]

k

(ej ⊗ b)

+ 2(c⊗ c)N

∑k=0

[( − Λ⊗ Ir − Ir ⊗ A)

−1

N ⊗ N]

k

(ej ⊗ b) (4.33)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 82

∂EN

∂Ni,j

= − 2(c⊗ c)N

∑k=1

k

∑j=1

[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j−1

(−Λ⊗ In − Ir ⊗A)−1(eieTj ⊗ In)

[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j

(−Λ⊗ In − Ir ⊗A)−1b⊗ b

+ 2(c⊗ c)N

∑k=1

k

∑j=1

[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

j−1

(−Λ⊗ Ir − Ir ⊗ A)−1(eieTj ⊗ Ir)

[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

k−j

(−Λ⊗ Ir − Ir ⊗ A)−1b⊗ b (4.34)

Remark 4.1. This derivation of the HN2 optimal necessary conditions for the approximation

of a degree N polynomial system ζN is essentially the partial sums version of the first order

necessary conditions for H2 optimality applied to the bilinear system ζ, where limN→∞

ζN = ζ.

That is, taking N →∞ gives the necessary conditions of Theorem 4.6.

The following theorem retranslates the necessary conditions gained from setting expressions

(4.31)-(4.34) to zero into multipoint interpolation conditions on the truncated Volterra series.

Theorem 4.8. Let ζ = (A,N ,b,c) be an order n bilinear system and ζN be the polynomial

system determined by ζ. Let ζr = (A, N , b, c) be a bilinear system of order r, and define ζNr

as the polynomial system determined by ζr. Suppose that ζNr is an H2 optimal approximation

to ζN . Then ζNr satisfies

N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk)

=N

∑k=1

r

∑l1=1

⋯r

∑lk−1=1

φl1,...,lk

Hk(−λl1 , . . . ,−λk), (4.35)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 83

and

N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λk))

=N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λlk)) (4.36)

where φl1,...,lk

, and λli are the residues and poles of Hk associated with ζNr .

Proof. Setting expression (4.31) to zero requires that

ej ⊗ cN

∑k=0

[( − Λ⊗ In − Ir ⊗A)

−1

N ⊗N]

k

(b⊗ b)

=ej ⊗ cN

∑k=0

[( − Λ⊗ Ir − Ir ⊗ A)

−1

N ⊗ N]

k

(b⊗ b) (4.37)

The left-hand side of (4.37), is equivalent to the product cN

∑k=0Vk(∶, j), where V0 solves

V0(−Λ) −V0A = bbT

and for k > 1, Vk solves

Vk(−Λ) −VkA =NVk−1NT (4.38)

Thus, we may apply Lemma 4.12, noting that by the proof of the lemma, the result also

applies for any of the partial sumsN

∑k=0Vk(∶, j) which correspond to the first N terms in the

solution of the generalized Sylvester equation

V (−Λ) − V A −NV NT = bb.T

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 84

Thus, we conclude that

N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λlk) (4.39)

=N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

Hk(−λl1 , . . . ,−λlk) (4.40)

Now consider (4.32). Setting this expression equal to zero requires that

(c⊗ c)N

∑k=1

k−1

∑j=0

[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j

(−Λ⊗ In − Ir ⊗A)−1(eieTi ⊗ In)

(−Λ⊗ In − Ir ⊗A)−1[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j−1

b⊗ b

=(c⊗ c)N

∑k=1

k−1

∑j=0

[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

j

(−Λ⊗ Ir − Ir ⊗ A)−1(eieTi ⊗ Ir)

(−Λ⊗ In − Ir ⊗ A)−1[(−Λ⊗ Ir − Ir ⊗ A)−1N ⊗ N]

k−j−1

b⊗ b (4.41)

Fixing the summation index k, consider the summation

k−1

∑j=0

(c⊗ c)[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j

(−Λ⊗ In − Ir ⊗A)−1(eieTi ⊗ In)

(−Λ⊗ In − Ir ⊗A)−1[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j−1

b⊗ b (4.42)

The expression

(c⊗ c)[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j

(−Λ⊗ In − Ir ⊗A)−1

corresponds to vec(Wj)T , where for j > 0, Wj solves

Wj(−Λ) −ATWj −NTWj−1N = cT c

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 85

and when j = 0,

Wj(−Λ) −ATWj = cT c

Similarly, the expression

(−Λ⊗ In − Ir ⊗A)−1[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j−1

b⊗ b

corresponds to vec(Vk−j−1), where Vk−j−1 solves the dual Sylvester equations. Therefore the

jth term in the sum (4.42) is equivalent to the product

(Wj(∶, i))TVk−j−1(∶, i)

We can therefore write

(c⊗ c)N

∑k=1

k−1

∑j=0

[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

j

(−Λ⊗ In − Ir ⊗A)−1(eieTi ⊗ In)

(−Λ⊗ In − Ir ⊗A)−1[(−Λ⊗ In − Ir ⊗A)−1N ⊗N]

k−j−1

b⊗ b (4.43)

=N

∑k=1

k−1

∑j=0

(Wj(∶, i))TVk−j(∶, i)) = (

N

∑k=1

(Wj(∶, i))T )(

N

∑k=1

Vk(∶, i)) (4.44)

The equality

N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λk))

=N

∑k=1

r

∑l1=1

⋯r

∑lk=1

φl1,...,lk

(k

∑j=1

∂sjHk(−λl1 , . . . ,−λlk)) (4.45)

now follows by exactly the same argument as in the proof of the derivative result for Theorem

4.17, where the only difference is the that the series terminates after N terms.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 86

The multipoint truncated-Volterra series interpolation conditions derived in Theorem 4.8 can

be enforced via the following algorithm, which we will call truncated B-IRKA or TB-IRKA.

Algorithm 4.3 (TB-IRKA). Input: A, N , b, c, A, N , b, c, N

Output: Aopt, N opt, bopt, copt

1. While: Change in Λ > 0 do:

2. RΛR−1 = A, b = bTR−T , c = cR, N =RTNTR−T

3. Solve

V1(−Λ) −AV1 = bbT

W1(−Λ) −ATW1 = cT cT

4. For j = 2, . . . ,N , solve

Vj(−Λ) −AVj =NVj−1NT

Wj(−Λ) −ATWj =NTWj−1N

T

5. V = (N

∑j=0Vj), W = (

N

∑k=0Wj).

6. A = (W T V )−1W TAV , N = (W T V )−1W TNV , b = (W T V )−1W Tb, c = cV .

7. end while

8. Aopt = A, N opt = N , bopt = b, copt = c

Upon convergence, the reduced order model ζr = (A, N , b, c) satisfies the interpolation

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 87

conditions of Theorem 4.8. This follows from the fact that V and W solve

V (−Λ) −AV −NN

∑j=0

VjNT = bbT (4.46)

W (−Λ) −ATW −NTN

∑j=0

WjN = cT c (4.47)

and as we have seen the reduced matrices A, N , b, cconstructed by projecting onto V and

along W satisfy the multipoint interpolation conditions on the first N terms.

The advantage of TB-IRKA is that it requires only 2Nr sparse linear solves per iteration to

exactly solve the ordinary Sylvester equations given in steps 3. and 4. of TB-IRKA. The

computational complexity of each linear solve is therefore on the order of n2, rather than

(rn)2. Thus as r grows, the total cost due to the reduced dimension r remains negligible,

making the cost of TB-IRKA comparable to IRKA, its linear counterpart . Moreover, we

have observed that in order for bilinear systems to stay in F , the magnitude of the terms in

the Volterra series decays very rapidly, so that comparable results to the algorithm B-IRKA

may be obtained by keeping only the first 3 or 4 terms in the series. The numerical examples

below demonstrate the results of using this approach.

A bilinear model of the Fokker-Planck equations

The following example is an application from stochastic control that was first introduced

by Hartmann et. al in [29] and later used as a test case for B-IRKA in [20]. Consider a

Brownian particle confined by a double-well potential W (x) = (x2−1)2. Assume the particle

is initially in the left well, and is dragged to the right well. The particle’s motion can be

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 88

described by the stochastic differential equation

dXt = −∇V (Xt, t)dt +√

2σdWt,

with σ = 2/3 and V (x,u) = W (x, t) + Φ(x,ut) = W (x) − xu − x. As an alternative to

these equations it is noted in [29] that one can instead determine the underlying probability

distribution function

ρ(x, t)dx = P [Xt ∈ [x,x + dx)]

which is described by the Fokker-Planck equation

∂ρ

∂t= σ∆ρ +∇ ⋅ (ρ∇V ), (x, t) ∈ (a, b) × (0, T ],

0 = σ∇ρ + ρ∇B, (x, t) ∈ {a, b} × [0, T ],

ρ0 = ρ, (x, t) ∈ (a, b) × 0

A finite-difference discretization of the Fokker-Planck equations consisting of 500 nodes in the

interval [−2,2] leads to a SISO bilinear system, where the output matrix c is a discretization

of the (set-theoretic) characteristic function of the interval [0.95,1.05]. Figure 4.1 compares

the relative H2 error in the reduced order models computed from B-IRKA and TB-IRKA

after truncating at the 13th term in the Volterra series. It was necessary to keep this many

subsystems because the Volterra series for this model converged somewhat slowly, and so

H2 error in the approximation decayed slowly as well. As Figure 4.1 demonstrates, there is

relatively little difference between the two approximations for most orders of approximation.

For the orders of approximation r = 2,4, the average time per iteration of TB-IRKA and B-

IRKA was the same, but as the reduced order system grew to between 6 and 24, the average

time per iteration for TB-IRKA was 51% less than for B-IRKA on the average. Figure 4.2

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 89

compares the average time per iteration for several orders of approximation.

2 4 6 8 10 12 14 16 18 20 22 24

10−2

10−1

Rel

ati

ve

H2

Err

or

O rde r of reduced system

TB-IRKA [13 te rms]

B-IRKA

Figure 4.1: Comparison of the relative H2 error for B-IRKA and TB-IRKA approximationsto the Fokker-Planck system

Nonlinear heat transfer model

The model used in this example is the bilinear model for the nonlinear heat transfer system

introduced in §2.2. Recall that the nonlinear part k(t) of the heat transfer system is a

polynomial of degree N +1 in the state variables given by (2.98). For applications of interest

N = 4, but as we have seen the system’s response is well approximated by the polynomial

terms up to degree 2, but poorly approximated by the linearization of the system. In this

example the original nonlinear system is order n = 40, yielding a bilinear system ζ of order

n = 1640 with 2 inputs and 1 output, taken as the second node in the order 40 discretization.

We approximate this bilinear system using both B-IRKA and TB-IRKA. In TB-IRKA we

truncated after only the first order (linear) homogeneous subsystem. Figure 4.3 compares

the relative H2 error for both approximations. As the figure illustrates, both B-IRKA and

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 90

0 5 10 15 20 250

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Order of reduced system

Avera

ge

tim

ep

er

itera

tion

[s]

TB-IRKA [13 te rms]

B-IRKA

Figure 4.2: Comparison of average time per iteration using B-IRKA and TB-IRKA[13 terms]for the Fokker-Planck system

TB-IRKA yield essentially the same accuracy of approximation for all orders, and achieve a

relative error of 9.4 × 10−7 with an order 12 approximation.

However, even with the order 12 approximation, either method applied directly to ζ yields

poor approximations for inputs of interest. This is illustrated in Figure 4.4, which compares

the response for the order 12 B-IRKA and TB-IRKA approximations to the true response for

the inputs u1(t) = 3.5 × 105 and u2(t) = 0. The low accuracy in the response, in spite of the

high accuracy in the reduced order model is due to the fact that that ∥N1∥2 and ∥N2∥ are

both on the order of 10−3, and so for inputs in the unit ball on L∞[0,∞), ζ behaves essentially

like an LTI system. This is reflected in the approximations computed by both B-IRKA and

TB-IRKA, which yield reduced A, B, and c, which are nearly equal to the result of applying

IRKA to the linear part of ζ for the same orders of approximation. In other words, neither

method can “see” the nonlinear terms N1, N2 in ζ. In applications, the magnitude of the

heat-flux inputs is on the order of 105, and so the input-state coupling becomes significant

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 91

2 3 4 5 6 7 8 9 10 11 1210−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Order of reduced system

Rel

ati

ve

H2

Err

or

TB - IRKA

B - IRKA

Figure 4.3: Comparison of the relative H2 error for B-IRKA and TB-IRKA approximationsto the nonlinear heat-transfer system

at this scale. In order to capture the nonlinear portion of the system accurately, we instead

applied B-IRKA and TB-IRKA to the scaled bilinear system ζα given by mappingNk → αNk

for k = 1,2 and B → αB. For this application we chose α = 5 ⋅ 104, which was the lowest

order magnitude for a constant heat-flux input. The truncation in TB-IRKA applied to the

scaled system was done after the second term in the Volterra series. Again, we see from

Figure 4.5 that B-IRKA and TB-IRKA yield essentially equivalent approximations for all

orders of approximation considered. Figure 4.6 shows the response of the true quadratic

nonlinear system compared with the order 12 B-IRKA and TB-IRKA approximations and

for the scaled approximations, the responses match almost perfectly.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 92

0 50 100 150 200 250 3000

50

100

150

200

250

300

Time [s]

Tem

pera

ture−3

00 [K

]

True ResponseB−IRKA [r=12]TB−IRKA [r=12]

Figure 4.4: Steady state response of nonlinear heat-transfer system and unscaled bilinearB-IRKA and TB-IRKA approximations of order 12

Viscous Burgers’ Equation Control System

Another model reduction benchmark is a bilinear control system derived from Burgers’ equa-

tion that was originally introduced in [25]. Consider the viscous Burgers’ equation

∂v

∂t+ v

∂v

∂x= ν

∂2v

∂x2,(x, t) ∈ (0,1) × (0, T )

subject to initial and boundary conditions

v(x,0) = 0, x ∈ [0,1], v(0, t) = u(t), v(1, t) = 0 t ≥ 0

Discretizing Burgers’ equation in the spatial variable using n0 nodes in a standard central

difference finite difference scheme leads to a system of nonlinear ordinary differential equation

where the nonlinearity is quadratic in the state. Measurements of the system are given as the

spatial average of v. The Carleman linearization technique applied to this system yields a

bilinearized system of dimension n = n0 +n20 that exactly matches the input-output behavior

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 93

2 3 4 5 6 7 8 9 10 11 1210−5

10−4

10−3

10−2

10−1

100

Order of reduced system

Rel

ati

ve

H2

Err

or

TB - IRKA

B - IRKA

Figure 4.5: Comparison of TB-IRKA and B-IRKA approximations of nonlinear heat transfersystem scaled with α = 5 × 104

of the original nonlinear system, since the nonlinearity is only quadratic. Here, we take

n0 = 50 and the set the parameter ν = 0.1. Figure 4.7 compares B-IRKA and TB-IRKA

approximations for even orders r = 2, . . . ,20, truncating after the second and fourth term in

TB-IRKA. We denote these truncations TB-IRKA[2], and TB-IRKA[4], respectively. As the

figure shows, there is little difference between TB-IRKA[2] and TB-IRKA[4] for all orders

of approximation. Figure 4.8 shows the average time per iteration for TB-IRKA[2] and TB-

IRKA[4] compared with the average time per iteration in B-IRKA. For this example, there

was a 35% decrease in the average time per iteration in TB-IRKA[2] on the average. However,

as Figure 4.8 shows, TB-IRKA[4] was actually slightly more costly for orders r = 4− 12. For

r = 14 − 20 TB-IRKA[4] had a shorter time per iteration compared with B-IRKA and for

orders r = 14 − 20, the average decrease in time per iteration for TB-IRKA[4] was 30%.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 94

0 50 100 150 200 250 3000

50

100

150

200

250

300

Time [s]

Tem

pera

ture−3

00 [K

]

True ResponseTB−IRKA [r=12]B−IRKA [r=12]

Figure 4.6: Steady state response of nonlinear heat-transfer system and scaled bilinear B-IRKA and TB-IRKA approximations of order 12

Heat transfer model

As a final example we will consider a boundary controlled heat transfer system. This model

has also become a benchmark for testing model reduction methods, and it was first introduced

in [16]. The system dynamics are governed by the heat equation subject to Dirichlet and

Robin boundary conditions.

xt = ∆x in (0,1) × (0,1),

n ⋅ ∇x = 0.8 ⋅ u1,2,3(x − 1) on Γ1,Γ2,Γ3

x = 0.8 ⋅ u4 on Γ4,

where Γ1,Γ2,Γ3 and Γ4 denote the boundaries of the unit square. A carefully constructed

spatial discretization using k2 grid points yields a bilinear system of order n = k2, with two

inputs and one output, chosen to be the average temperature on the grid. Taking k = 100,

we demonstrate TB-IRKA on a bilinear system of order n = 10,000, and compare it with

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 95

1 2 3 4 5 6 7 8 9 1010−4

10−3

10−2

10−1

100

Order of reduced system

Rel

ati

ve

H2

Err

or

B-IRKA

TB-IRKA [2 te rms]

TB-IRKA [4 te rms]

Figure 4.7: Comparison of TB-IRKA and B-IRKA approximations of Burgers’ equationcontrol system

B-IRKA for the same system. Figure 4.10 compares the relative H2 error in TB-IRKA

approximations truncated at two and four terms with the relative error in the B-IRKA

approximation for the same orders. Again, the figure illustrates that their is relatively little

difference between the two approaches for the orders r = 2,16 using even just two terms in

the Volterra series. Using 4 terms in the Volterra series yields TB-IRKA approximations that

are essentially equivalent to the B-IRKA approximations for all orders. Both B-IRKA and

TB-IRKA started from the same initial guess, and we compared average time per iteration

for all orders of approximation in Figure 4.9. For orders r = 2,4, B-IRKA is marginally

faster, but on average when N = 2 there was a 62% decrease in the time per iteration in

TB-IRKA compared to B-IRKA and when N = 4, there was a 30% decrease in the time per

iteration in TB-IRKA compared to B-IRKA.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 96

2 4 6 8 10 12 14 16 18 200

0.5

1

1.5

2

2.5

3

3.5

4

Order of reduced system

Avera

ge

tim

ep

er

itera

tion

[s]

B-IRKA

TB-IRKA [2 te rms]

TB-IRKA [4 te rms]

Figure 4.8: Comparison of average time per iteration for TB-IRKA[2,4] and B-IRKA appliedto Burgers’ equation control system

A parameter-varying convection-diffusion problem

Benner and Breiten showed in [15] that certain classes of parameter-varying linear systems

can be effectively approximated over the desired range of parameters by appropriately re-

formulating the linear system as a bilinear system. Here we carry out this approach for

a parameter-varying convection-diffusion problem first introduced in [11]. The model is

governed by the standard convection-diffusion equations

∂x

∂t(t,ξ) = p0∆x(t,ξ) +

2

∑i=1

pi∇x(t,ξ) + b(ξ)u(t),

ξ ∈ [0,1] × [0,1], t ∈ (0,∞), x(t,ξ) = 0, ξ ∈ ∂([0,1] × [0,1])

and the parameters p0, p1, p2 need to be adjusted to capture the particular physics that is

being modeled.

After a finite-difference discretization in the spatial variable ξ, we obtain the linear parameter-

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 97

0 5 10 15 20 25 300

5

10

15

20

25

30

Order of reduced system

Avera

ge

tim

eper

itera

tion

[s]

TB-IRKA [2 te rms]

TB-IRKA [4 te rms]

B-IRKA

Figure 4.9: Comparison of average time per iteration in TB-IRKA and B-IRKA for severalorders

varying dynamical system

x(t) = p0A0x(t) +2

∑i=1

Aix(t)pi + bu(t) (4.48)

y = cx(t),

where c = on is chosen as the observation matrix. This system can be viewed as a bilinear

system where the parameters p1 and p2 are particular system inputs. We can rewrite system

(4.48) as a bilinear system with three inputs and one output:

x =Ax +3

∑k=1

Nkxuk(t) +Bu(t)

y(t) = cx(t)

with A=p0A0, N1 = A1, N2 = A2, N3 = 0, B = [0,b] ∈ Rn×3, for inputs of interest

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 98

0 5 10 15 20 25 3010−4

10−3

10−2

10−1

100

Order of reduced system

Rel

ati

ve

H2

Err

or

TB-IRKA [2 te rms]

TB-IRKA [4 te rms]

B-IRKA

Figure 4.10: Comparison of TB-IRKA and B-IRKA approximations of heat transfer controlsystem

u(t) = [p1, p2, u(t)]T .

The parameter range of interest is p0 ∈ [0.1,1], p1, p2 ∈ [0,1] [11]. Taking p0 = 1 we computed

TB-IRKA approximations keeping 2,3, and 6 terms in the Volterra series, and compared

them with the B-IRKA approximation to the full bilinear system. Each of these approxima-

tions were of dimension r = 12. To place the reduced bilinear system matrices in the linear

parameter-varying formulation we took A = p0A0 ∈ Rr×r, N1 = A1 ∈ Rr×r and N2 = A2 as the

reduced-dimension matrices that approximate the linear parameter-varying system (4.48).

In order to evaluate the accuracy of the approximations, we varied the parameters p1 and p2

over the whole parameter range of interest, and for each selection of parameters we computed

the relative H2 norm of the error between the full and reduced dimension systems for that

choice of parameters. The surfaces plotted in Figure 4.11 show how the relative H2 error

of the linear systems varies over the parameter values. As Figure 4.11 shows, TB-IRKA[2

terms] actually gives the best approximation error over the parameter space, and the ap-

proximation error increases as the number of terms kept in the Volterra series increases, with

B-IRKA giving, in this case, the largest errors over the parameter space. We believe this is

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 99

due to the fact B-IRKA is actually a better approximation over the whole L2 unit ball of

inputs, and so it gives up some accuracy for these particular inputs. Next we computed a

TB-IRKA[2 terms] approximation and a B-IRKA approximation both of dimension r = 12

to approximate the bilinear system resulting from taking p0 = 0.5. Figure 4.12 shows the

relative H2 error in the linear systems over the parameter range for p0 = 0.5. Again for this

example, TB-IRKA[2 terms] yields a smaller approximation error than B-IRKA, and both

yield nearly uniform error over the range of parameters.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 00.5

18e−5

2.8e−4

4.8e−4

6.84e−48e−4

p 2

p 1

Rel

ativ

e H 2 E

rror

TB−IRKA[6]TB−IRKA[3]TB−iRKA[2]B−IRKA

Figure 4.11: Convection-diffusion problem: Comparison of the relative H2 error in theB-IRKA and TB-IRKA[2, 3 and 6 terms] approximations taking p0 = 1 and varying over theparameter range for p1 and p2

4.1 Alternatives to H2 Optimal Bilinear Model Reduc-

tion

It is not uncommon to end up dealing with a bilinear system which does not have a convergent

H2 norm. For example, the bilinear system approximation ζ, to the nonlinear RC circuit

model first introduced by Skoogh and Bai is a standard benchmark model for testing methods

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 100

00.2

0.40.6

0.81

00.10.20.30.40.50.60.70.80.9110−5

10−4

10−3

10−2

p 1

Relative H2 error when p 0 = 0.5, r = 12

p 2

Rel

ativ

eH

2E

rror TB−IRKA [2 terms]

B−IRKA

Figure 4.12: Convection-diffusion problem: Comparison of the relativeH2 error in the B-IRKA and TB-IRKA[2 terms] approximations taking p0 = 0.5 and varying over the parameterrange for p1 and p2

of model reduction, but ζ /∈ F [6]. Other benchmark models, such as bilinear approximations

to Burgers’ equation are also not H2 for modest Reynolds numbers. In these situations,

there are a few options available. One technique is to scale ζ by the mapping

γ ↦ ζγ ∶= (A, γN , γb, c),

where γ < 1 is chosen sufficiently small so that ∥ζγ∥H2< ∞. H2 optimal model reduction is

carried out on ζγ, and the original input-output map can be recovered by scaling the inputs

u(t) for the original system by u(t)/γ.

Any truncation ζN of the original system has a finite HN2 norm. Computing an HN2 optimal

approximation to ζN using TB-IRKA is therefore another alternative when ζ is not H2.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 101

Frequently an H2 optimal approximation of the first few terms in the Volterra series is

sufficiently accurate to match the output of the ζ.

Yet another approach is to match some combination of subsystem moments in the hopes of

capturing the dominant portion of the Volterra series for the inputs of interest.

Numerical Examples

TB-IRKA and the subsystem interpolation approach is demonstrated on an RC circuit model

of dimension 40,200. Solving the bilinear Sylvester equations required for B-IRKA was

beyond our computational resources for this model. We demonstrate the scaled B-IRKA

and TB-IRKA approaches on a control problem governed by the viscous Burger’s equations

first introduced in §4. The subsystem interpolation approach of Bai and Skoogh [6] that

matches moments about zero for the first and second order transfer functions is used as the

choice of moments in the subsystem interpolation approach.

Nonlinear RC circuit model

A bilinear model for an RC circuit with nonlinear resistors and an independent current

source was originally derived in [6]. This circuit was originally proposed by Chen and White

[32]. Let v1, . . . , vN be the N node voltages in the circuit and u(t) be the input-signal to the

independent current source. This circuit can be modeled by a linear-analytic model of the

form⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

v = f(v(t)) + bu(t)

y(t) = cv(t), (4.49)

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 102

where

f(v) =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−g(v1) − g(v1 − v2)

g(v1 − v2) − g(v2 − v3)

g(vk−1 − vk) − g(vk − vk+1)

g(vN−1 − vN)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

∈ RN b = cT =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1

0

0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

∈ RN (4.50)

and g(v) = e40v(t) + v − 1. Taking a second order approximation of g(v) by matching the 2

terms of its MacLaurin series yields a quadratic nonlinear system. One can then apply the

Carleman linearization to this quadratic nonlinear system to obtain a bilinear model of the

nonlinear circuit. If the original linear-analytic model (4.49) is order N , the bilinear model

approximation is order n = N +N2. For this example, we take N = 200, so n = 40,200. We

compare the TB-IRKA approach with the Bai and Skoogh subsystem interpolation approach

for two standard test inputs: u(t) = e−t and u(t) = (cos(πt/10)+1)/2. Following the example

given in their paper [6] we constructed an order 21 reduced-order model that matched the

following moments about zero:

cA−q1b, for q1 = 1 . . . ,20 (4.51)

cA−1NA−q1b, for q1 = 1, . . . ,20 (4.52)

(4.53)

This approximation is compared to an order 13 approximation constructed from TB-IRKA,

where TB-IRKA was implemented by keeping the first two terms in the Volterra series.

Figures 4.13, 4.15 compares the response of the reduced order models to the true response

for inputs u(t) = e−t and u(t) = (cos(πt/10)+1)/2 respectively. Figures 4.14 and 4.16 compare

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 103

the relative error in the response As the figures illustrate, the true system responses are very

well matched by the TB-IRKA approximation.

0 0.5 1 1.5 2 2.5 3 3.5 40

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0.016

0.018

Time [ s]

y(t

)

True Re sponse

TB - IRKA [2 te rms, r=13]

S ubsy stem Inte rpolat ion [ r=21]

Figure 4.13: Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem inter-polation response to the true response for the input u(t) = e−t

Burgers’ Equation Control System

In this example, we use the bilinear control system derived from Burgers’ equation introduced

§4. Here we take ν = 0.01, corresponding to a Reynolds number of 100 and construct a bilinear

system ζ of order n = 930. ζ is not an H2 system, which can be checked by observing that

the series used to compute its control grammian diverges. For this example we compared

TB-IRKA[2] with the scaled version of B-IRKA. An order r = 9 approximation is used to

compute the response of both methods to the inputs u(t) = e−t and u(t) = sin(20t). The

relative error in the output of using the scaling values γ = 0.4,0.5 for B-IRKA are compared

with the TB-IRKA[2] approximation in Figures 4.17, 4.18. As the figures show, very good

approximation results using B-IRKA can be obtained for the right value of γ, in this case

γ = 0.4 yielded good approximations, but the quality of the approximations is fairly sensitive

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 104

0 0.5 1 1.5 2 2.5 3 3.5 410−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Time [ s]

Rel

ati

ve

Abso

lute

Err

or

Su bsy stem Inte rpolation [ r=21]

TB - IRKA [2 te rm s, r=13]

Figure 4.14: Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem inter-polation error for the input u(t) = e−t

to the choice of γ. For both inputs the TB-IRKA[2] approximation yields a highly accurate

approximation, and indeed is more accurate than B-IRKA on the whole time interval of [0,4]

seconds for the input u(t) = e−t.

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 105

0 1 2 3 4 5 6 7 8 9 100

0.005

0.01

0.015

0.02

0.025

Time [ s]

y(t

)

True Re sponse

TB - IRKA [2 te rm s, r=13]

Su bsy stem Inte rpolation [ r=21]

Figure 4.15: Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem inter-polation response to the true response for the input u(t) = (cos(πt/10) + 1)/2

0 1 2 3 4 5 6 7 8 9 1010−9

10−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

Rel

ati

ve

Abso

lute

Err

or

T ime [ s]

Su bsystem [ r=21]

TB-IRKA [2 te rms, r=13]

Figure 4.16: Nonlinear RC Circuit: A comparison of the TB-IRKA and subsystem inter-polation error for the input u(t) = (cos(πt/10) + 1)/2

Garret M. Flagg Chapter 4. H2 Optimal Model Reduction 106

0 0.5 1 1.5 2 2.5 3 3.5 410−9

10−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

Time [ s]

Rela

tive

Abso

lute

Err

or

TB-IRKA [2 te rms, r = 9]

B-IRKA [γ = .5,r = 9]

B-IRKA [γ = .4, r = 9]

Figure 4.17: Burgers’ Equation: A comparison of the TB-IRKA and scaled B-IRKA errorfor the input u(t) = e−t

0 5 10 15 20 25 30 35 40 45 5010−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

Time [ s]

Rel

ati

ve

Abso

lute

Err

or

TB-IRKA [2 te rms, r = 9]

B-IRKA [γ = .5,r = 9]

B-IRKA [γ = .4, r = 9]

Figure 4.18: Burgers’ Equation: A comparison of the TB-IRKA and scaled B-IRKA errorfor the input u(t) = sin(20t)

Chapter 5

Solving the Bilinear Sylvester

Equations

The solution X of the bilinear Sylvester equation

AX +XB +m

∑k=1

NkXUk +Y = 0 (5.1)

where A, Nk ∈ Rn×n, B,Uk ∈ R`×` and Y ∈ Rn×` has frequently played an important role

in the bilinear model reduction methods discussed so far. Moreover, we have seen how the

bilinear Lyapunov equations

AX +XAT +m

∑k=1

NkXNTk +Y = 0. (5.2)

have important system theoretic interpretations in terms of the bilinear controllability and

observablity grammians given by equations (2.33), and (2.34).

107

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 108

When all Nk or Uk = 0, (5.1) reduces to the standard Sylvester equation

AX +XB +Y = 0. (5.3)

In the next two sections we will discuss both direct and iterative methods for solving large-

scale bilinear Sylvester equations, with special attention to the bilinear Lyapunov equations.

In the third section we will present a new analysis of rational Krylov projection methods

for solving Sylvester equations and their connection to the ADI method. We will then

consider the generalization of these results to bilinear Sylvester equations. These results

provide insight into the advantages and limitations of using Krylov projection methods in

the bilinear case.

5.1 Direct Methods

Perhaps the most brute force approach to solving (5.1) is to transform it into a linear system

by applying the vec operator to it, which yields the nm-dimensional linear system

− (I ⊗A +BT ⊗ I +m

∑k=1

UTk ⊗Nk)vec(X) = vec(Y ). (5.4)

and X can be found by inverting the nm-dimensional linear operator T = −(I⊗A+BT ⊗I +m

∑k=1UTk ⊗Nk). The computational complexity of using this approach directly is O((nm)3),

which makes it untenable for nearly all large-scale applications. This effectively exhausts

all known direct methods for solving the full bilinear Sylvester equations. Unless otherwise

specified, we will therefore consider the bilinear Lyapunov equations exclusively in what

follows.

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 109

A more efficient direct method of solution based on the Sherman-Morrison-Woodbury for-

mula exploits the fact that the Nk are typically low rank. This approach was developed by

E.G. Collins, Jr. et. al. in [33], [79]. The lefthand side of equation (5.2) can be viewed

as the sum of the linear operators LA(X) = AX +XAT and Π(X) =m

∑k=1NkXNT

k . If

rank(Nk) = rk ≪ n, then the rank of Π is bounded above by r0 =m

∑k=1

r2k. Assume that

for each Nk, a rank revealing factorization is available, and that rank(Π) = r ≪ n2. Then

vec(Π) may be factored as P1P2, where P2 ∈ Rr,n2and P1 ∈ Rn2,r. One may then apply the

following lemma, which is a variant of the Sherman-Morrison-Woodbury formula, to obtain

the solution of equation (5.2).

Lemma 5.1. Let L,P ∈ Rn×n, and consider the linear equation

(L +P )x = y. (5.5)

Suppose P = P1P2 where P2 ∈ Rr,n and P1 ∈ Rn,r, and that L is invertible. Then equation

(5.5) is uniquely solvable if and only if I +P2L−1P1 is nonsingular. If w solves

(Ir +P2L−1P1)w = P2L

−1y

then x = L−1(y −P1w) solves equation (5.5).

The direct inversion of LA can be done in O(n3) operations using the Bartels-Stewart al-

gorithm [10]. If a rank revealing factorization of Π is indeed available, then (5.2) may be

solved in max{O(n3r),O(r3)} operations [37]. If in addition the Nk’s, k = 1, . . . ,m, must be

factorized, then (5.2) may be solved in O(mn3+r20n

2+rn3) operations [37]. For an algorithm

that achieves this estimate and details the factorization of the Nk’s and Π, see [37].

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 110

5.2 Iterative Methods

There are several iterative methods for solving the bilinear Lyapunov equations, all of which

are based on the following results on convergent regular splittings due to H. Schneider [84].

The following is a brief summary of the iterative approaches introduced by Damm in [37].

Throughout the discussion, σ(T ) ⊂ C denotes the spectrum of a linear operator T , and

ρ(T ) = max{∣λ∣ ∶ λ ∈ σ(T )} denotes the spectral radius.

Theorem 5.1. Let A ∈ Rn×n and consider linear operators LA,Π ∶ Rn×n → Rn×n, where

LA =AX +XAT , and Π is nonnegative in the sense that Π(X) ≥ 0, whenever X ≥ 0. The

following are equivalent:

a.) ∀Y > 0,∃X > 0 s.t. LA(X) +Π(X) = −Y ,

b.) ∃Y ,X > 0 s.t. LA(X) +Π(X) = −Y ,

c.) ∃Y , with (A,Y ) controllable s.t. ∃X > 0 and LA(X) +Π(X) = −Y ,

d.) σ(LA +Π) ⊂ C−,

e.) σ(LA) ⊂ C− and ρ(L−1A Π) < 1.

If ρ(L−1A Π) < 1, then (LA +Π)−1 is given by the Neumann series

(LA +Π)−1 =∞

∑k=0

(L−1A Π)k(L−1

A ), (5.6)

which leads to the most obvious iterative scheme

Xk+1 = L−1A Π(Xk) +L

−1A (Y ) (5.7)

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 111

that we have already been applying liberally throughout the proofs in the multipoint Volterra

series interpolation discussion. Although this iteration is guaranteed to converge to the

solution, the convergence may be very slow. A more sophisticated iterative method uses

Krylov-subspace methods such as GMRES (generalized minimum residual) to minimize the

residual Rk =Xk+1 −Xk, see [37]. For larger problems, inverting LA may still be too costly.

In this case, one can use the ADI method as a kind of preconditioner. The ADI method

was first introduced by Peaceman and Rachford [75] as a method for solving parabolic and

elliptic PDEs, and was later adapted to solving the Sylvester equation by Wachspress in

[97]. It is a fixed point iteration scheme for approximating the solution X to the ordinary

Sylvester equation

AX +XB +Y = 0. (5.8)

that has been developed extensively, see [76, 18, 17, 68, 55, 90, 89, 57, 83, 97, 93, 98]. Given

two sequences of shifts {α1, α2, . . . , αr, . . .},{β1, β2, . . . , βr, . . .} ⊂ C and an initial guess X0,

the ADI iteration proceeds as follows :

Xi =(A − αiI)(A + βiI)−1Xi−1(B − βiI)(B + αiI)

−1 (5.9)

− (αi + βi)(A + βiI)−1Y (B + αiI)

−1. (5.10)

This iteration can be generalized to the bilinear Lyapunov equation by observing that for a

given shift α ∈ R ∖ {0},

LA(X) =1

2α((A + αI)−1X(A + αI)T − (A − αI)X(A − αI)T ) (5.11)

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 112

Thus, (5.2) can be written in the fixed-point form

X = (A+αI)−1(A−αI)X(A+αI)−T (A−αI)T −2α(A+αI)−1(m

∑j=1

NjXNTj +Y )(A+αI)−T

(5.12)

The ADI iteration can now be interpreted as a preconditioning technique in the following

way. For given ADI-parameters α` < 0 ` = 1, . . . , L and Xk, define the iterations

Xk+ lL= (A + α`I)

−1(A + α`I)Xk+ l−1L(A + α`I)

−T (A − α`I)T

− 2α`(A + α`I)−1(

m

∑j=1

NjXkNTj +Y )(A + α`I)

−T (5.13)

Setting

G0 =L

∏`=1

(A + α`I)−1(A − α`I),

Gp = (L

∏`=p+1

(A + α`I)−1(A − α`I))(A + αpI)

−1, for p = 1, . . . , L

a single iteration on Xk can then be written as

Xk+1 = G0XkGT0 +

L

∑p=1

2αp(m

∑j=1

(GpNj)Xk(GpNj)T +GpY G

Tp ) (5.14)

[37]. The kth step in this iteration is equivalent to applying L steps of the ADI iteration at

the kth term in the series (5.6), and therefore at best it will converge at the same rate as the

iteration in (5.7). The obvious advantage to this approach is that provides an inexpensive

way to approximate L−1A and can therefore be applied as a preconditioning technique in an

iterative method like GMRES [37].

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 113

5.3 Krylov projection-based approximation of ordinary

Sylvester equations

In this section, we will make a brief excursion into the analysis of rational Krylov projection

methods (RKPM) for the ordinary Sylvester equations. In the next section, we will generalize

our results to the bilinear case. Throughout this section Y is the rank one matrix Y =

bc for Sylvester equations and in the context of Lyapunov equations we will take Y =

bb∗. I will present new results that connect the RKPM with the ADI iteration for linear

systems. The RKPM is the presiding alternative to the ADI iteration for solving large-

scale Sylvester equations [58, 61, 38, 60, 85, 91, 42, 9]. In the RKPM, the Sylvester equation

AX+XB+Y = 0 is projected onto the rational Krylov subspaces Kratr (A,b,σ) = span{(σ1I−

A)−1b, . . . , (σrI −A)−1b} and Kratr (B∗,c, µ) where σ= {σ1, . . . σr}, and µ = {µ1, . . . , µr} are

the sets of shifts used to construct the respective rational Krylov spaces and ν denotes the

conjugate of ν. See [12] for further details regarding Kratr (A,b,σ), and constructing an

orthonormal basis via the rational Arnoldi iteration. Let Qr and Ur denote the orthonormal

basis for Kratr (A,b,σ) and Krat

r (B∗,c, µ). Then, the RKPM approximation is constructed

by first solving

Q∗rAQrXr + XrU

∗rB

∗Ur +Q∗rbc

∗Ur = 0 (5.15)

and then approximating X by QrXrU∗r . The solution of the projected Sylvester equation

(5.15) is inexpensive. Like the ADI method, the RKPM method also relies heavily on a

good choice of shifts to produce accurate results. The main result is that the ADI iteration

and rational Krylov projection-based methods are equivalent for the special case of shifts

satisfying condition (1) in Theorem 4.2. I will call these shifts pseudo-H2 optimal shifts.

Recently, Breiten and Benner showed that these shifts are also locally optimal with respect

to a special energy norm related to the Lyapunov equations for Hermitian linear dynamical

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 114

systems [21]. The main equivalence theorem requires the following lemma, which connects

the ADI approximation for the Sylvester equation with rational Krylov subspaces. This

extends an earlier result by Li and White [67] which establishes a similar connection for the

the case of the Lyapunov equation.

Lemma 5.2. Let Q = bc∗, where b ∈ Rn and c ∈ Rm. Let {σ1, . . . , σr} and {µ1, . . . , µr}

be two collections of shifts that satisfy R(µi),R(σi) > 0 for i = 1, . . . , r. Suppose Xr is the

approximate solution to the Sylvester equation (5.3) obtained by applying the pair of shifts

αi = −σi and βi = −µi in the ADI iteration (5.9) for i = 1, . . . , r with X0 = 0. Then there

exist Lr ∈ Cn×r and Rr ∈ Cm×r such that Xr = LrR∗r and colspan(Lr) ⊂ Krat

r (A,b,µ) and

colspan(Rr) ⊂ Kratr (B∗,c, σ)

Proof. The proof is given by induction on i, the iteration step. First note that for i = 1,

X1 = (µ1 + σ1)(A − µ1I)−1bc∗(B − σ1I)−1, so let L1 = [(µ1 + σ1)(A − µ1I)−1b] and R1 =

[(B∗ − σ1I)−1c]. Then L1 and R1 clearly satisfy the hypothesis and X1 = L1R∗1 . Now

suppose that the statement holds for Xi. Then, for j = 1, . . . , i, the jth column of Li is

T(j)i (A)b, where T

(j)i (λ) is a proper rational function that lies in the span of { 1

λ−µ1, . . . , 1

λ−µi}.

Similarly, the jth column ofRi is S(j)i (B∗)c, where S

(j)i (λ) lies in the span of { 1

λ−σ1, . . . , 1

λ−σi}.

Therefore Xi+1 can be written as

Xi+1 =(A + σi+1I)(A − µi+1I)−1LR∗(B + µi+1I)(B − σi+1B)−1

+ (µi+1 + σi+1)(A − µi+1I)−1bc∗(B − σi+1I)

−1

=i

∑j=1

(A + σi+1I)(A − µi+1I)−1T

(j)i (A)bc∗S

(j)i (B)(B + µi+1I)(B − σi+1I)

−1

+ (µi+1 + σi+1)(A − µi+1I)−1bc∗(B − σi+1I)

−1

For j = 1, . . . i, let the jth column of Li+1 be (A + σi+1I)(A − µi+1I)−1T(j)i (A)b and let the

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 115

(i + 1)th column be T(i+1)i+1 (A)b = (µi+1 + σi+1)(A − µi+1I)−1b. Then clearly colspan(Li+1) ⊂

Krati+1(A,b,µ). Similarly, let (B∗ − σi+1I)−1(B∗ + µi+1I)S

(j)i (B∗)c be the jth column of

Ri+1 for j = 1, . . . , i, and S(i+1)i+1 (B∗)c = (B∗ − σi+1I)−1c be the (i + 1)th column. Then

colspan(Ri+1) ⊂ Krati+1(B

∗,c, σ). Finally, we note that by construction, Xi+1 = Li+1R∗i+1.

Next, we present our equivalency result, showing that the approximate solution of the

Sylvester equation (5.3) by ADI and RKPM are indeed equivalent when the shifts are cho-

sen as pseudo-H2 optimal points. This result applied to the special case of the Lyapunov

equations was first presented at the 2010 SIAM Annual Meeting [44] then later published

independently in [38]. Our new result here, on the other hand, is more general than both [44]

and [38] since it tackles the case of Sylvester equation and includes the Lyapunov equation

as a special case. Moreover, while the proof given in [38] for the special case of Lyapunov

equation makes use of a novel connection between the ADI iteration and the so-called Skele-

ton approximation framework first developed in the work of Tyrtyshnikov [94], the proof

we provide here for the more general Sylvester equation case is given directly in terms of

rational Krylov interpolation conditions, and in that sense is simpler.

Theorem 5.2. Given the Sylvester equation (5.3) with Y = bc∗, where b ∈ Rn and c ∈ Rm,

let Qr ∈ Rn×r be an orthonormal basis for the rational Krylov subspace Kratr (A,b,σ) and let

Ur ∈ Rm×r be an orthonormal basis for the rational Krylov subspace Kratr (B∗,c, σ) for a set

of shifts σ = {σ1, . . . , σr} where R(σi) > 0 for i = 1, . . . , r. Let Xr ∈ Rr×r solve the projected

Sylvester equation

Q∗rAQrXr + XrU

∗rBUr +Q

∗rbc

∗Ur = 0, (5.16)

and let Xr ∈ Rn×m be computed by applying the shifts αi = −σi and βi = −σi to exactly r

steps of the ADI iteration (5.9) for i = 1, . . . , r. Then Xr = QrXrU∗r if and only if either

λ(Q∗rAQr) = −{σ1, . . . , σr} or λ(U∗

rBUr) = −{σ1, . . . , σr}.

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 116

Proof. (⇐) First suppose that λ(Q∗rAQr) = −{σ1, . . . , σr}. The proof remains the same

if we instead suppose that λ(U∗rBUr) = −{σ1, . . . , σr}. Let A = Q∗

rAQr, and b = Q∗rb,

Br = U∗rBUr, and c = U∗

r c. Note that after we apply r steps of the ADI iteration with the

set of shifts αi = βi = −σi to the projected Sylvester equation (5.16), we obtain the exact

solution Xr, since λ(A) = −{σ1, . . . , σr}. By Lemma 5.2, at the rth step of the ADI iteration

Xr = LrR∗r where Lr = [T (1)(A)b, . . . , T (r)(A)b] where T (i)(A)b are rational functions that

lie in Kratr (A, b,σ). Similarly Rr = [S(1)(B∗

r )c∗, . . . , S(r)(B∗

r )c∗] where the S(i)(B∗

r )c are

rational functions that lie in Kratr (B∗

r , c∗, σ). Furthermore, for the same shifts, αi = βi = −σi

for i = 1, . . . , r, applied to r steps of the ADI iteration on the full Sylvester equation (5.3), we

have Xr = LrR∗r and Lr = [T (1)(A)b, . . . , T (r)(A)b] and Rr = [S(1)(B∗)c, . . . , S(r)(B∗)c].

Thus it is sufficient to show that QrLr = Lr and that UrRr =Rr. Without loss of generality

consider just the former equation. This, in turn, amounts to showing that QrT (i)(A)b =

T (i)(A)b. If Ti(A)b are a set of orthogonal rational functions that span Kratr (A,b,σ), then

it is sufficient to show that

QrTi(A)b = Ti(A)b. (5.17)

Equality (5.17) follows readily from the interpolation properties of the Galerkin projection,

which we show below. First, note that due to the interpolation properties of the Galerkin

projection, Qr(σiIr − A)−1b = (σiI −A)−1b. Let V = [(σ1I −A)−1b . . . (σrI −A)−1b]. Then,

for some x ∈ Rr,

V x = Ti(A)b =Qr[(σ1Ir − A)−1b . . . (σrIr − A)−1b]x =QrTi(A)b, (5.18)

which proves (5.17). (⇒) Let Xr be the solution of

Q∗rAQrXr + XrU

∗rBUr +Q

∗rbc

∗Ur = 0 (5.19)

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 117

where Qr is an orthonormal basis for Kratr (A,b,σ) and Ur is an orthonormal basis for

Kratr (B∗,c, σ). Suppose that QrXrU∗

r =Xr. Let Xr be the approximate solution of (5.19)

resulting from applying the shifts αi = βi = −σi for i = 1, . . . , r to exactly r steps of the ADI

iteration (5.9). By the interpolation result given in the proof above, QrXrU∗r = Xr. It

follows from the assumptions that, QrXrU∗r = QrXrU∗

r , so Xr = Xr. But this means that

Xr solves (5.19), and so either λ(Q∗rAQr) = −{σ1, . . . , σr} or λ(U∗

rBUr) = −{σ1, . . . , σr}.

The parameters for which the ADI iteration and the rational Krylov projections coincide

also satisfy orthogonality conditions on the residual for the special case of the Lyapunov

equation

AX +XA∗ + bb∗ = 0 (5.20)

For a given approximation Xr to the solution X, define the residual R as

R =AXr +XrA∗ + bb∗. (5.21)

The following result was first given in [38]. Here we present a new and more concise proof

of the orthogonality result in terms of the special interpolation properties of the pseudo

H2-optimal shifts.

Theorem 5.3. Given AX +XA∗ + bb∗ = 0, let Xr ∈ Rr×r solve the projected Lyapunov

equation

Q∗rAQrXr + XrQ

∗rAQr +Q

∗rbb

∗Qr = 0,

where Qr is an orthonormal basis for the Kratr (A,b,σ) with σ = {σ1, . . . , σr} Let Xr =

QrXrQ∗r .Then Q∗

rR = 0 if and only if λ(Q∗rAQr) = −{σ1, . . . , σr} where R is the residual

defined in (5.21).

Proof. (⇒) Suppose that Q∗rR = 0. Multiplying (5.21) with Q∗

r from the left and then

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 118

transposing the resulting equation leads to

AQrXr +QrXrQ∗rA

∗Qr + bb∗Qr = 0. (5.22)

Let A =Q∗rAQr = TΛT −1 be the eigenvalue decomposition of A where Λ = diag(λ1, . . . , λr).

Plug these expressions into (5.22), and right multiply by T −∗ to obtain

QrXrT−∗Λ∗ +AQrXrT

−∗ + bb∗QrT−∗ = 0 (5.23)

Let ζi be the ith entry of b∗QrT −∗. Then it is straightforward to show that the ith column

of QrXrT −∗ must be (−λiI −A)−1bζi. Thus, it follows that Kratr (A,b,σ) = Krat

r (A,b,−λ),

where λ = {λ1, . . . , λr}. Since both sets σ and λ are closed under conjugation, after an

appropriate reordering, we obtain σi = −λi.

(⇐) Observe that

AXr + XrA∗ +Q∗

rbb∗Qr = 0⇒ (5.24)

AXrT−∗ + XrT

−∗Λ∗ +Q∗rbb

∗QrT−∗ = 0. (5.25)

Thus, the ith column of XrT −∗ is (−λiIr−A)−1Q∗rbζi. But since Qr is an orthonormal basis

for Kratr (A, b,σ), and λi = −σi, this means

Qr(−λiIr − A)−1Q∗rbζi = (−λiI −A)−1bζi = (QrXrT

−∗)ei, (5.26)

where ei is the ith unit vector. Thus,

QrXrT−∗Λ∗ +AQrXrT

−∗ + bb∗QrT−∗ = 0, (5.27)

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 119

which implies

QrXrQ∗rA

∗Qr +AQrXr + bb∗Qr = 0. (5.28)

Transpose this last expression and use the fact that Q∗rQr = Ir to obtain

Q∗rQrXrQ

∗rA

∗ +Q∗rAQrXrQ

∗r +Q

∗rbb

∗ =Q∗rR = 0, (5.29)

which is the desired result.

Numerical examples for LTI systems

Here we present two numerical examples illustrating the accuracy of the RKPM method using

pseudo-H2 optimal shifts applied to LTI systems. Three different approximation methods

are compared for each model.

• Method 1: The RKPM is applied to the a sequence of shifts that alternates between 0

and ∞. The resulting subspace is generally referred to as the extended Krylov subspace.

Its application to RKPM was first introduced by Simoncini in [85].

• Method 2: The RKPM is applied using r pseudo-H2 optimal shifts; or equivalently r

steps the ADI iteration is applied using r pseudo-H2 optimal shifts.

• Method 3: r-steps of the ADI iteration are applied, where the ADI shifts are chosen

via Penzl’s heuristic method [76].

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 120

The Eady model

The first example is an Eady model used for atmospheric storm tracking. The full order

model is order 598. For more information about this model, see [30]. All three methods are

used to approximate the controllability grammian P of the system. The approximations are

compared in both the induced-2 norm, where the SVD the approximation is optimal. It is

clear from Figure 5.1 that the Method 2 outperforms Method 1 and Method 3 for all ranks

of approximation, and that it yields a nearly optimal approximation in the 2-norm.

5 10 15 20 25 30 35 40 45 5010−10

10−8

10−6

10−4

10−2

Rank

||X−

Xk|| 2

/||X

|| 2

πk + 1/π 1

M etho d 1

M e tho d 2

M e tho d 3

Figure 5.1: Relative error in the 2-norm as r varies for the EADY Model

Rail Model

This model arises from a semidiscretized heat transfer problem for the optimal cooling of

steel profiles during a cooling process for a rolling mill. See [19], for further information on

this model. The the order of the full model is n = 1357. In this example, Figure 5.2 shows

that methods 2 and 3 perform comparably, but notably this is in large part due to the fact

that the collection of shifts chosen in Penzl’s heuristic in method 3 distribute themselves

closely to the distribution of pseudo-H2 optimal shifts. Both methods 2 and 3 yield very

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 121

accurate approximations in the matrix 2-norm for each rank.

0 2 4 6 8 10 12 14 16 18 2010−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

101

Rank

||X−

Xk|| 2

/||X

|| 2

π k + 1/π 1

M etho d 1

M etho d 2

M etho d 3

Figure 5.2: Relative error in the 2-norm as r varies for the Rail Model

5.4 Krylov projection-based methods for the approxi-

mation of the bilinear Lyapunov equations

We now generalize the orthogonality results of the previous section to the bilinear Lyapunov

equations. The results are presented for SISO bilinear systems to simplify the presentation,

but can readily be extended to MIMO systems. Let X be any approximation to the solution

X of the bilinear Lyapunov equations

AX +ATX +NXNT + bbT = 0.

Define the residual R as

R =AX + XAT +NXNT +Q = 0.

Theorem 5.4. Consider a SISO bilinear system ζ=(A, N , b, c). Let Qr be an orthonormal

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 122

basis for V which solves the multipoint Volterra series interpolation problem (3.31) for shifts

σi = −λ(QTrAQr) and weight matrix U = T −1QT

rNQrT , where TΛT −1 = A is the eigenvalue

decomposition of A =QTrAQr. Let Xr, solve the projected bilinear Lyapunov equations

AXr + XrA + NXrNT + bbT = 0.

Then the residual Rr of the approximation Xr = QrXrQTr is orthogonal to Qr. Moreover,

if Qr∈ Rn×r is any orthonormal matrix of rank r, define A, N , b, c in the usual way, and let

TΛT −1 = A be the usual eigenvalue decomposition. If Xr = QrXrQTr where Xr solves the

projected bilinear Lyapunov equations and QTrRr = 0, then ζr interpolates ζ in the points

−λj(A) with weights U = T −1QTrNQrT .

Proof. For the first part, we show that the residual is orthogonal to the projection subspace.

By hypothesis,

AXr + XrA + NXrNT + bbT = 0 ⇒

AXrT−T + XrT

−TΛ + NXrT−TUT + bbTT −T = 0

By the construction in the proof of Theorem 3.3, the jth column of XrT −T is given as the

weighted Volterra series (3.33) with the weights and shifts given in the hypotheses, and since

Qr is an orthonormal basis for V , QrXrT −T solves

V (−Λ) −AV −NV UT − bbTT −T = 0

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 123

Therefore

QrXrT−TΛ +AQrXrT

−T +NQrXrT−TUT + bbTT −T = 0 ⇒

QrXrQTrA

TQr +AQrXr +NQrXrQTrN

TQr + bbQTr = 0

Transposing this last equality and using the fact that QTrQr = Ir gives

QTr (AXr +XrA

T +NXrNT + bbT ) = 0

QTr (Rr) = 0

Now suppose that Qr ∈ Rn×r is a rank r orthonormal matrix that satisfies the hypothesis of

the second part. Then

QTr (AXr +XrA

T +NXrNT + bbT ) = 0

AXrQTr + XrQ

TrA

T + NXrQTrN

T + bbT = 0 ⇒

QrXrT−T (−Λ) −AQrXrT

−T −NQrXrT−TUT − bbTr = 0

So again, V =QrXrT −T solves the interpolation problem in the points −λ(A) and weights

U = TNT −1. ζr resulting from the projection onto V and along (V T V )−1V T differs from

the system ζr by the similarity transformation T TX−1r , and therefore ζr interpolates ζ in

the points −λ(A) and weights U .

Note that in the bilinear case the interpolation conditions are necessary and sufficient, but

the projection subspaces are not necessarily determined by the weights in U , and the points

σi. This is due to the fact that there is no underlying function space for the Volterra

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 124

series interpolation problem with functions that are uniquely determined by the choice of

interpolation points and weights, as is the case for rational Krylov subspaces.

These orthogonality conditions can be used to show that the one-sided interpolation condi-

tions are optimal for Hermitian bilinear systems in a special matrix inner product. Suppose

A, N are Hermitian, the pair (A, b) is controllable, and a solution to the bilinear Lyapunov

equation with Y = bbT exists. By Theorem 5.1, σ(LA +Π) ⊂ C−, and since LA +Π is Her-

mitian, T = −(I ⊗A +A⊗ I +N ⊗N) is positive definite. Following Vandereycken in [95],

we can define the inner product

⟨x,y⟩T = yTT x

and define the corresponding norm

∥x∥T =√

⟨x,x⟩T

which we will call the induced energy norm of the bilinear system. The orthogonality con-

ditions given in Theorem 5.4 yield the following optimality result in the the induced energy

norm.

Theorem 5.5. Let ζ have state-space realization (A,N ,b,cT ), and suppose A,N , are Her-

mitian. Let Xr satisfy the hypotheses of Theorem 5.4. Then Xr is optimal over the subspace

Z = {Z ∶ Z =QrZQTr , Z ∈ Rr×r} in the induced energy norm.

Proof. Note that Z is closed and convex. By the standard Hilbert space projection theorem

[66], optimality is equivalent to ⟨vec(X −Xr), vec(Z)⟩T = 0 for any Z ∈ Z. By the definition

of the T inner product

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 125

⟨vec(X −Xr), vec(Z)⟩T = 0

⇐⇒ trace(ZT (LA +Π)(X −Xr)) = 0

⇐⇒ trace(QrZ(QTrRr)) = 0,

and this last equality clearly obtains for the approximation Xr.

Bilinear heat transfer system

The controllability grammian P of the bilinear heat transfer system first introduced in

Chapter 4. of order n = 10,000 is approximated. For this system the operator L is Hermitian

positive definite, so the L-norm of the grammian exists. The pseudo-H2 approximations are

compared with the SVD approximation in both the L-norm and the Frobenius norm in

Figures 5.4 and respectively. As expected from our results in Theorem 5.4, the pseudo-H2

projection subspace outperforms the SVD in the L-norm for each rank of approximation,

but the rank r SVD approximation performs better in the Frobenius norm. In either case,

there is little difference in the quality of the approximations.

Garret M. Flagg Chapter 5. Solving the Bilinear Sylvester Equations 126

0 5 10 15 20 25 30 35 40 45 5010−5

10−4

10−3

10−2

10−1

100

Rela

tive

err

or

inL-

norm

Rank r of Xr

Pseudo-H 2 pro je c t i on

S VD

Figure 5.3: Comparison of the relative error in the L-norm for pseudo-H2 projection subspaceand SVD approximations of the heat transfer model

0 5 10 15 20 25 30 35 40 45 5010−5

10−4

10−3

10−2

10−1

100

Rela

tive

err

or

inFro

beniu

snorm

Rank r of Xr

Pseudo-H 2 pro je c t i on

S VD

Figure 5.4: Comparison of the relative error in the Frobenius norm for pseudo-H2 projectionsubspace and SVD approximations of the heat transfer model

Chapter 6

Data-Driven Model Reduction of

SISO Bilinear Systems

In this section we present a solution to the following bilinear realization problem: Suppose

we have data corresponding to the values of the kth-order transfer functions Hk(s1, . . . , sk)

of a SISO bilinear system evaluated at several values in Ck, for k = 1, . . . n. We want to find

a system ζ with realization (A, N , b, c) such that ζ agrees on all the subsystem data. Note

that in this chapter the realization parameters (A,N ,b,c) are not related to full or reduced

order models, but are intended as the generic notation for the realization parameters. It

turns out that as long as the data has a special structure, it is possible to construct such

a bilinear realization that will satisfy the interpolation data. The structure on the data

essentially corresponds to the data given in the subsystem interpolation problem of Theorem

3.2. Assume for the moment that given the data, their is a bilinear system ζ ∶= (A,N ,b,c)

that satisfies the subsystem matching conditions. Our approach will be to convert the bilinear

interpolation problem into a tangential interpolation problem for a MIMO linear system. The

idea behind this is most easily seen with a simple example. Suppose that a bilinear system

127

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 128

of order 2 has realization A,N ,b,c, and that we have the following subsystem transfer

function data: H2(1,2) = c(2I −A)−1N(I −A)−1b, H2(2,1) = c(I −A)−1N(2I −A)−1b,

H1(1) = c(I −A)−1b, H1(2) = c(2I −A)−1b. We can use this data to construct a MIMO

linear system with realization (A, B, C) by defining

A =A, B = [b N(σ1I −A)−1b] , C = ([cT NT (σ1I −AT )−1cT )]T

.

This MIMO system, which we denote G(s), matches the bilinear subsystem data collection

along correctly chosen bi-tangential directions. For example,

H(1,2) = c(σ2I −A)−1N(σ1I −A)−1b = [1 0]G(2)

⎡⎢⎢⎢⎢⎢⎢⎣

0

1

⎤⎥⎥⎥⎥⎥⎥⎦

As the example suggests, the solution to the linear realization problem will play an important

role in constructing bilinear realizations from the kernel data. Let us therefore first consider

the solution to the realization problem for MIMO LTI systems. In classical linear realization

theory a sequence of Markov moments hk ∈ Rp×m is specified, and the goal is to find a triple

of matrices (A ∈ Rn×n,B ∈ Rn×m,C ∈ Rp×n), called a realization, that satisfies hk =CAk−1B.

The essential object involved in the construction of the realization is the Hankel matrix

H =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

h1 h2 . . . hk hk+1 . . .

h2 h3 . . . hk+1 hk+2 . . .

⋮ ⋮ . ..

⋮ ⋮ . ..

hk hk+1 . . . h2k−1 h2k . . .

hk+1 hk+2 . . . h2k h2k+1 . . .

⋮ ⋮ . ..

⋮ ⋮

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 129

The sequence is realizable if and only if rank(H) = n < ∞ [2]. Moreover, the minimal

dimension of the realization is n. The main tool in the construction of the state-space model

is the column shift operator σ. The action of σ on any column of H is given by shifting right

m columns. Given H of rank n, a realization can be constructed as follows. Let ∆ ∈ Rn×n be

a nonsingular submatrix of H, and let σ∆ be the matrix having the same rows, but columns

resulting from shifting each column of ∆ by m columns. Let Γ ∈ Rn×m have the same rows

as ∆, but the first m columns only, and let Ξ ∈ Rp×n be the submatrix of H composed of the

same columns as ∆, but its first p rows. A realization that matches the Markov parameters

in the sequence hk is then given by

A = ∆−1σ∆, B = ∆−1Γ, C = Ξ

The key object required to generalize the realization problem to tangential interpolation

data associated with a linear dynamical system is the Loewner matrix. The Loewner matrix

was first used to solve rational interpolation problems with unconstrained poles by Belevitch

[13]. It was used to systematically solve interpolation problems in [3], in particular the scalar

rational interpolation problem with unconstrained poles. Later the Loewner matrix approach

to interpolation was extended to the matrix-valued case in [4]. The Loewner matrix encodes

the rational interpolation data in such a way that the interpolant can be characterized by

it. The presentation of the Loewner matrix here is given in terms of tangential interpolation

data on a p ×m rational matrix function. Let H(s) = C(sE −A)−1B, where A,E ∈ Rn×n,

B ∈ Rn×m and C ∈ Rp×n. Tangential interpolation data corresponds to sampling H(s) along

different directions at different points λi ∈ C. The interpolation data can be divided into

right and left tangential interpolation directions.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 130

Following [70], the right interpolation data is presented as follows:

{(λi,ri,wi)∣λi ∈ C, ri ∈ Cm×1,wi ∈ Cp×1, i = 1, . . . , ρ}

Interpreted in terms of H(s), the right interpolation data satisfies

wi =H(λi)ri.

The left interpolation data is presented as

{(µj, `j,vj)∣ µj ∈ C, `j ∈ C1×p,vj ∈ C1×m, j = 1, . . . , ν},

and this data satisfies

`jH(µj) = vj.

Given left and right interpolation data in this form, assume that µi ≠ λj for i, j = 1, . . . , r,

and define the Loewner matrix L ∈ Cν×ρ and shifted Loewner matrix σL ∈ Cν×ρ

Li,j =virj − `iwj

µi − λjσLi,j =

µivirj − λj`iwj

µi − λj

In order to carry out the construction, the following assumption is made on the data:

rank(xL − σL) = rank [L σL] = rank

⎡⎢⎢⎢⎢⎢⎢⎣

L

σL

⎤⎥⎥⎥⎥⎥⎥⎦

= k, x ∈ {λi} ∪ {µj} (6.1)

Theorem 6.1. [70] If assumption (6.1) is satisfied, then for some x ∈ {λi} ∪ {µj}, compute

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 131

the short singular value decomposition

xL − σL = Y ΣX

where rank(xL−σL) = rank(Σ) = size(Σ) = k and Y , ∈ Cν×k, X ∈ Ck×ρ. A minimal realization

(E,A,B,C) of an interpolant is then constructed as follows:

E = −Y ∗LX∗

A = −Y ∗σLX∗

B = Y ∗V

C =WX∗

As a partial solution to the bilinear realization problem, our goal will be to recast the

bilinear interpolation data in the form of a linear tangential interpolation problem, and then

apply Theorem 6.1 to obtain the A-matrix in the bilinear realization. Before we proceed,

let us first consider what approaches in bilinear realization theory have previously been

taken. As we shall see, the classical linear realization theory has been nicely generalized

in a couple of different ways. Issidori directly generalized the Hankel matrix for linear

systems to the bilinear case and showed how to use his generalization to solve the bilinear

realization problem [59]. The approach outlined here was developed by Frazho [49]. Frazho

developed a complete bilinear realization theory in terms of a pair of forward shift operators,

which he applied to discrete-time bilinear systems. His approach was later applied to the

continuous-time case by Rugh [81]. We will present Rugh’s application to continuous-time

bilinear systems. An alternative to Frazho’s approach is presented in the work of Fliess, who

develops a realization theory in terms of power series expansions of noncommuting variables

[46], [48]. A more recent approach to the Fliess realization theory is developed by Ball et.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 132

al. in [8].

6.1 Classical Bilinear Realization Theory

Given a degree k homogenous regular Volterra kernel Hk(s1, . . . , sk) of the continuous-time

bilinear system ζ, expand Hk in a negative power series about the point at infinity in Ck

Hk(s1, . . . , sk) =∞

∑i1=0

⋯∞

∑ik=0

h(i1, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)k

The bilinear realization problem on the sequence of Volterra kernels

(H1(s1),H2(s1, s2), . . . ,Hk(s1, . . . , sk), . . . ) ∈ F is to find a realization ζ ∶= (A,N ,b,c) so

that

cAikNAik−1N⋯NAi1b = h(i1, . . . , ik)

for all k > 0 and all nonnegative integers i1, i2, . . . , ik.

Define the shift operators σk as

σkHk(s1, . . . , sk) =∞

∑i1=0

⋯∞

∑ik=0

h(i1 + 1, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)k

Define the shift operator τk by

τkHk(s1, . . . , sk) =

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

∑i1=0

⋯∞

∑ik−1=0

h(0, i1, . . . , ik−1)s−(i1+1)1 ⋯s

−(ik−1+1)k−1 k > 1

0, k = 1

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 133

The σk are linear operators that may be interpreted as

σkHk(s1, . . . , sk) = s1Hk(s1, . . . , sk) − lims1→∞

s1Hk(s1, . . . , sk),

whenever the limits exist. The τk are linear operators that map a power series in k variables

to a power series in k − 1 variables, and τ kk = 0. It is also straightforward to see that

τkHk(s1, . . . , sk) = lims1→∞

H(s1, . . . , sk)∣ s2=s1s3=s2⋮

sk=sk−1

whenever the limit exists.

Now define the linear operators

σ =∞

⊕k=1

σk (6.2)

τ =∞

⊕k=1

τk. (6.3)

Two more operators are necessary to define an abstract realization of the sequence H. Define

the initialization operators ιk ∶ R→ U by

ιkr =Hk(s1, . . . , sk)r

and the initialization operator on F as ι =∞

⊕k=1ιk. Define the evaluation operators εk ∶ F → R

by

εkHk(s1, . . . , sk) =

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

0, k > 1

h0, k = 1

And the evaluation operator on the sequence as ε =∞

⊕k=1

εk. Next, consider the subspaces

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 134

U1 = span{H, σH, σ2H, . . .}. Let τUi denote the image of Ui under τ , and for i > 1 define

Ui = span{τUi−1, στUi−1, σ2τUi−1, . . .}

Now consider the linear subspace

U = span{U1,U2,U3, . . .}.

Theorem 6.2. [81],[49] (σ, τ, ι, ε) on U is an abstract bilinear realization of the bilinear

system ζ on defined by the sequence of Markov moments h(i1, i2, . . . , ik) for k ≥ 1 and all

i1, . . . , ik > 0.

Proof. For a given ` ≥ 1,

σj1ι = σj1H

σj1H =∞

⊕k=1

(σj1k

∑i1=0

⋯∞

∑ik=0

h(i1, i2, . . . , ik)s−(i1+1)1 s

−(ik+1)k )

=∞

⊕k=1

(∞

∑i1=0

⋯∞

∑ik=0

h(i1 + j1, i2, . . . , ik)s−(i1+1)1 s

−(ik+1)k )

τσj1ι =∞

⊕k=1

(∞

∑i1=0

⋯∞

∑ik=0

h(j1, i1, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)k )

σj2τσj1ι =∞

⊕k=1

(∞

∑i1=0

⋯∞

∑ik=0

h(j1, i1 + j2, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)k )

τσj2τσj1ι =∞

⊕k=1

(∞

∑i1=0

⋯∞

∑ik=0

h(j1, j2, i1, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)k )

σj`τσjk−1τ⋯τσj1ι =∞

⊕k=1

(∞

∑i1=0

⋯∞

∑ik=0

h(j1, j2, j3, . . . , i1 + j`, . . . , ik)s−(i1+1)1 ⋯s

−(ik+1)ik+1 )

εσj`τσjk−1τ⋯τσj1ι = h(j1, j2, . . . , j`)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 135

Constructing a finite dimensional realization with matrices (A, N , b, c) representing the

action of the operators σ, τ, ι, ε depends on the dimension of U . If U is finite dimensional,

then the H has a finite dimensional bilinear realization of the same dimension. Using this

construction, one can also generalize the Hankel matrix for linear systems to a behavior

matrix for bilinear systems. Identify H, σH, τH and so on with their sequences of Markov

moments

H = ((h(0), h(1), h(2), . . .), (h(0,0), h(0,1), h(0,2), . . . , h(1,0), h(1,1), h(1,2), . . .), . . .)

σH = ((h(1), h(2), h(3), . . .), (h(1,0), h(1,1), h(1,2), . . . , h(2,0), h(2,1), h(2,2), . . .), . . .)

τH = ((h(0,0), h(0,1), h(0,2), . . .), (h(0,0,0), h(0,0,1), h(0,0,2), . . . ,

h(0,1,0), h(0,1,1), h(0,1,2), . . .), . . .)

Through this identification, we can then write

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 136

BH =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

H

σH

σ2H

τH

στH

σ2τH

σj`τ⋯τσj1H

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

=

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

h(0) h(1) . . . h(0,0) h(0,1) . . .

h(1) h(2) . . . h(1,0) h(1,1) . . .

⋮ ⋮ ⋮ ⋮ ⋮ ⋮

h(0,0) h(0,1) . . . h(0,0,0) h(0,0,1) . . .

⋮ ⋮ ⋮ ⋮ ⋮ ⋮

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

There are many ways to systematically list the sequences in the Markov moments, and so the

behavior matrix given above is a nonunique description of H. Isidori et. al. [59] constructed

an alternative behavior matrix and used it to construct a bilinear realization directly, without

any discussion of shift operators. The entries of their matrix, properly reordered, correspond

to the BH as it is presented here.

6.2 The structure of the interpolation data

The solution of the interpolation problem posed on the kth order Volterra kernels presented

in Theorem 3.2 provides insight into a natural way to organize the interpolation data. In

effect, all the interpolation data is essentially determined by the highest order subsystem

for which interpolation conditions are stipulated. Assume that the subsystem of order N

is the highest order subsystem for which interpolation constraints are imposed. Similar to

the construction of the Krylov subspaces in Theorem 3.2, we build the notation for the

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 137

collection of interpolation constraints starting from the first order homogeneous subsystem

and working our way up. So first start with the array of interpolation constraints

S1 = {(σ1, φ1), . . . , (σk1 , φk1)∣σi, φi ∈ C for i = 1, . . . , k1}.

Now let σi,j denote the point (σi, σi,j) ∈ C2 and let

S2 ={(σ1,1, φ1,1), (σ1,2, φ1,2), . . . , (σ1,k2 , φ1,k2), (σ2,1, φ2,1), . . . , (σ2,k2 , φ2,k2),

. . . , (σk1,1, φk1,1), . . . , (σk1,k2 , φk1,k2)∣φi,j ∈ C for i = 1, . . . , k1 and j = 1, . . . , k2}

Continuing recursively, let σl1,...,lm = (σl1 , σl1,l2 , . . . , σl1,...,lm) ∈ Cm and define Sm accordingly

as the set of associated interpolation constraints for the collection of points {σl1,...,lm}, lj =

1, . . . , kj for j = 1, . . . ,m ⊂ Cm. Let

S = ∪Nm=1Sm.

The total number of interpolation constraints in S is M = k1 + k1k2 + k1k2k3 + ⋅ ⋅ ⋅ + k1k2⋯kN .

Now let

U1 = {(γ1, µ1), (γ2, µ2), . . . , (γk1 , µk1)}∣γi, µi ∈ C for i = 1, . . . , k1}.

Let γi,j denote the point (γi,j, γi) ∈ C2. Note that in this case the new value γi,j sits in the

first entry of the 2-tuple, rather than in the second as is the case with σi,j. Define

U2 ={(γ1,1, µ1,1), . . . , (γ1,k2 , µ1,k2), (γ2,1, µ2,1), . . . , (γ2,k2 , µ2,k2),

. . . , (γk1,1, µk1,1), . . . , (γk1,k2 , µk1,k2)∣µi,j ∈ C for i = 1, . . . , k1 and j = 1, . . . , k2}

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 138

Define the sets Um for m = 3, . . .N similarly to U2, and let

U = ∪Nm=1Um.

There are also M = k1 + k1k2 + k1k2k3 + ⋅ ⋅ ⋅ + k1k2⋯kN interpolation constraints in U. A last

collection of interpolation constraints is needed to determine the N matrix in the bilinear

realization. Let

σ = [σ1, . . . , σk1 , σ1,1, . . . , σ1,k2 , σ2,1, . . . , σ2,k2 , . . . , σ1,1,...,1, σ1,1,...,2, . . . , σ1,1,...,kN , . . . , σk1,k2,...,kN ]

be a row vector of length M that uniquely lists the whole collection of values enumerated

in the m − tuples defined in Sm m = 1, . . . ,N . Let γ similarly be a row vector of length

M that lists all the values in the m − tuples given by the arrays Um for m = 1, . . . ,N .

Let C = σ ⊗ γ be the tensor product of σ and γ, and with each entry σl1,...,lm ⊗ γr1,...,ru

for m,u = 1, . . . ,N and lm = 1, . . . , km, ru = 1, . . . , ku of the tensor identify the point

(σl1 , σl1,l2 , . . . , σl1,...,lm , γl1,...,lu , γl1,...,lu−1 , . . . , γl1) ∈ Cm+u. Define the last collection of inter-

polation constraints as

T = {(σl1,...,lm ⊗ γr1,...,ru , η(l1,...,lm),(r1,...,ru))∣(σl1,...,lm ⊗ γr1,...,ru) ∈ Cm+u and η(l1,...,lm),(r1,...,ru) ∈ C}

(6.4)

Let

P = S ∪U ∪T. (6.5)

The total number of interpolation constraints stipulated in P is 2M +M2, where again

M = k1 + k1k2 + k1k2k3 + ⋅ ⋅ ⋅ + k1k2⋯kN . We will encode the data corresponding to point con-

catenations in T in a concatenation matrix denoted by τL. First, we place the interpolation

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 139

values η(l1,...,lm),(r1,...,ru) in a long column matrix denoted η in the same order as the entries

of the tensor σ⊗γ. Partition η into M blocks of length M denoted ηi for i = 1, . . . ,M as in

η = [η1 η2 . . . ηM] ∈ C1×M2

(6.6)

Each ηi collects all the interpolation constraints for some point σr1,...,ru concatenated with

each of the points in γ. Now define

τL = [vec(η1) vec(η2) . . . vec(ηM)] ∈ CM×M (6.7)

6.3 Construction of the Bilinear Realization

In order to obtain the bilinear realization, we first use S and U to construct a MIMO linear

realization satisfying tangential interpolation constraints determined by S and U. For the

given data, a MIMO system having q = 1 + k1 + k1k2 + k1k2k3 + ⋅ ⋅ ⋅ + k1k2⋯kN−1 inputs and

outputs will be constructed. Thus, we must first translate the data in S, and U into right

and left tangential interpolation conditions respectively. Let Iq be the identity in Rq×q, and

let the vectors ej for j = 1, . . . , q denote the jth column of the identity. With each pair

(γl1,...,lm , µl1,...,lm) ∈ U associate the tuple (γl1,...,lm ,vMm+Ll1,...,lm, lMm+Ll1,...,lm

= eTJm+Ll1,...,lm−1

),

where

J1 = 0, J2 = 1, and Jm =m

∑j=1

Jlj + k1k2⋯km−2 for m > 2, (6.8)

M1 = 0 and Mm =m

∑j=1

Mj + k1k2⋯km−1, for m > 1 (6.9)

(6.10)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 140

Ll1,...,lm = 1, . . . , k1k2⋯km uniquely enumerates the combinations l1, . . . , lm by first running

through the indices in lm and fixing all the other indices at one, then incrementing the index

in lm−1 by one and lm and keeping l1, . . . , lm−2 fixed and so on. Let

vMm+Ll1,...,lm= [µl1,...,lm , τL(Mm +Ll1,...,lm ,1 ∶MN)] ∈ Cq.

Define the whole collection of left interpolation conditions by

M = diag(γ) ∈ CM×M

L = [lT1 , . . . , lTM]T ∈ CM×q and V =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

v1

vM

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

∈ CM×q.

(6.11)

Next, we construct right tangential interpolation conditions from S. Again for each pair

(σl1,...,lm , φl1,...,lm) associate the tuple (σl1,...,lm ,wMm+Ll1,...,lm,rMm+Ll1,...,lm

= eJm+Ll1,...,lm−1

), where

here

wMm+Ll1,...,lm=

⎡⎢⎢⎢⎢⎢⎢⎣

φl1,...,lm

τL(1 ∶MN ,Mm +Ll1,...,lm)

⎤⎥⎥⎥⎥⎥⎥⎦

(6.12)

Define the full collection of right interpolation conditions by

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 141

Λ = diag(σ) ∈ CM×M

R = [r1, . . . ,rM], and W = [w1, . . . ,wM] ∈ Cq×M .

(6.13)

From this data, the Loewner matrix and shifted Loewner matrices L and σL can be con-

structed as the solutions of the following Lyapunov equations [70]:

LΛ −ML = LW −V R σLΛ −MσL = LWΛ −MVR (6.14)

Assume that for some k ≤M assumption (6.1) is satisfied. Then construct the linear system

Σ=(A ∈ Ck×k, B ∈ Ck×p, C ∈ Cp×k, 0 ∈ Cp×p) according to Theorem 6.1. In order to construct

a bilinear realization, we make the following additional assumptions on the data:

Assumption: Either k =M or k = p and rank(L(1 ∶ k,1 ∶ k)) = k (6.15)

Assuming either case in assumption (6.15) holds, a bilinear system realization (A,N , b, c)

which will satisfy all the interpolation conditions given in P can be constructed directly from

(A, B, C) and the data in τL.

To complete the construction of a bilinear realization, we need to extract some additional in-

formation from the matrix τL to constructN . We break the development of the construction

into the two possible cases k =M or k = p.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 142

The case k =M

When the rank of L is M , A ∈ CM×M , EM×M constructed according to Theorem 6.1 must

also be rank M . Thus E is invertible, and we make the reassignment

A =E−1A

B =E−1B (6.16)

C =C

Define

B =[B(∶,1)oTr ,B(∶,2)oTk2 , . . . ,B(∶,1 + r)oTk2 ,B(∶,2 + r)oTk3 . . .

B(∶,1 + r + rk2)oTk3, . . . ,B(∶ 2 + r + rk2 + ⋅ ⋅ ⋅ + rk2k3⋯km−2)o

Tkm, . . . ,

B(∶, p = 1 + r + rk2 + ⋅ ⋅ ⋅ + rk2k3⋯km−1)oTkm

] ∈ CM×M , (6.17)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 143

C =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

orC(1, ∶)

ok2C(2, ∶)

ok2C(r + 1, ∶)

ok3C(r + 2, ∶)

ok3C(1 + r + k2r, ∶)

okmC(2 + r + rk2 + ⋅ ⋅ ⋅ + rk2k3⋯km−2, ∶)

okmC(∶, p = 1 + r + rk2 + ⋅ ⋅ ⋅ + rk2k3⋯km−1)

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

∈ CM×M (6.18)

Now let U , QT ∈ CM×M be the solutions of

UΛ −AU − B = 0 QTM −AT QT − CT = 0 (6.19)

The matrices U and Q provide a factorization of the Loewner matrix L.

Theorem 6.3. Let L ∈ CM×M be the Loewner matrix formed from the interpolation data Λ,

R,W ,M ,L,V , and assume that rank(L) =M . Then L = −QU .

Remark 6.1. From the assumption that L is full rank, it follows from Theorem 6.3 that Q

and U are full rank and therefore invertible.

Proof. Let {xi,j} =QU and {li,j} = L. Let qi be the ith row of Q and let uj be the jth row of

U for i, j = 1, . . . ,M . By the construction of Q, and U , the ith row of Q corresponds to the

row vector C(i, ∶)(γl1,...,lmI −A)−1, for some γl1,...,lm corresponding to the entry M(i, i) and

by the construction of C, C(∶, i) = C(∶, Jlm−1 + Ll1,...,lm−1). Likewise, uj corresponds to the

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 144

column vector (σr1,...,ruI −A)−1B(∶, j), for some σr1,...,ru corresponding to the entry Λ(j, j)

and by the construction of B, B(∶, j) =B(∶, Jlu−1 +Lr1,...,ru−1). So,

xi,j =C(∶, Jm +Ll1,...,lm−1)(γl1,...,lmI −A)−1(σr1,...,ruI −A)−1B(Ju +Lr1,...,ru−1) (6.20)

= lMm+Ll1,...,lmC(γl1,...,lmI −A)−1(σr1,...,ruI −A)−1BrMu+Lr1,...,ru

(6.21)

=lMm+Ll1,...,lm

(C(γl1,...,lmI −A)−1B −C(σr1,...,ruI −A)−1B)rMu+Lr1,...,ru

σr1,...,ru − γl1,...,lm(6.22)

=vMm+Ll1,...,lm

rMu+Lr1,...,ru

− lMm+Ll1,...,lmwMu+Lr1,...,ru

σr1,...,ru − γl1,...,lm(6.23)

=virj − liwj

σj − γi(6.24)

= −li,j (6.25)

The action of Q on the columns of B corresponds to shifting left along columns of the

concatenation matrix τL, and likewise the action of U on the rows C corresponds to shifting

down along the rows of τL. In this sense, τL functions analogously to the indexing shift

operator τ given in (6.3).

Lemma 6.1. Assume that L is full rank, and let Q, U are given by (6.19), and that A, B,

C, are constructed according to (6.16). Then

1. QB(∶, Ju+1 +Lr1,...,ru) = τL(∶,Mu +Lr1,...,ru), for u = 1, . . . ,N − 1 and M1 = 0, Lr1 = r1,

rm = 1, . . . , km for m = 1, . . . ,N − 1.

2. (C(Jm+1 +Ll1,...,lm)T U = τL(Mm +Ll1,...,lm , ∶), for m = 1, . . . ,N − 1 and M1 = 0, Ll1 = l1,

lm = 1, . . . , km for m = 1, . . . ,N − 1

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 145

Proof. Fix some r1, . . . , ru. For i = 1, . . . ,M

Q(i, ∶)B(∶, Ju+1 +Lr1,...,ru) = C(i, ∶)(γl1,...,lmI −A)BeJu+1+Lr1,...,ru

(6.26)

=C(Jm +Ll1,...,lm−1 , ∶)(γl1,...,lm)I −A)−1BeJu+1+Lr1,...,ru

(6.27)

= lMm+Ll1,...,lmC(γl1,...,lmI −A)−1Be

Ju+1+Lr1,...,ru(6.28)

= vMm+Ll1,...,lm(Ju+1 +Lr1,...,ru) (6.29)

= τL(Mm +Ll1,...,lm ,Mu +Lr1,...,ru) (6.30)

Since each row i =Mm +Ll1,...,lm of Q corresponds to some point γl1,...,lm , the product of the

rows of Q with the column B(∶, Ju+1 +Lr1,...,ru) corresponds to concatenating all the points

in γ with the point σr1,...,ru . By the definition of τL, this is the same thing as the column

τL(∶,Mu +Lr1,...,ru). This proves 1.

For the proof of 2., fix some l1, . . . , lm. Then for j = 1, . . . ,M

C(∶, Jm+1 +Ll1,...,lm)U(∶, j) =C(∶, Jm+1 +Ll1,...,lm)(σr1,...,ruI −A)−1B(∶, j) (6.31)

(from the definition of B).

= eTJm+1+Ll1,...,lmC(σr1,...,ruI −A)−1B(∶, Ju +Lr1,...,ru−1),

= eTJm+1+Ll1,...,lmC(σr1,...,ruI −A)−1BrMu+Lr1,...,ru

(6.32)

=wMu+Lr1,...,ru(Jm+1 +Ll1,...,lm) (6.33)

= τL(Mm +Ll1,...,lm ,Mu +Lr1,...,ru) (6.34)

Since each column j =Mu+Lr1,...,ru of U corresponds to the point σr1,...,ru , the product of the

columns of U with the row C(∶, Jm+1 +Ll1,...,lm) corresponds to concatenating all the points

in σ with the point γl1,...,lm , which is equivalent to (6.34).

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 146

We now have all the components necessary to construct the bilinear realization. Define

b =B(∶,1) (6.35)

c =C(1, ∶) (6.36)

N = Q−1τLU−1. (6.37)

Theorem 6.4. Given the interpolation constraints P defined in (6.5), assume that rank(L) =

M and let A, B, C be defined as in (6.16), U , Q be given by (6.19), and b, c, N be

given by (6.35), (6.36), (6.37) respectively. Then ζ:=(A,N ,b,c) satisfies all the subsystem

interpolation conditions in the array P.

Proof. We first consider the case where the interpolation constraint is in the array S. Suppose

we have the pair (σr1,...,ru , φr1,...,ru) ∈ S. Then

Hu(σr1 , σr1,r2 , . . . , σr1,...,ru)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1N(σr1I −A)−1b

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1Q−1τLU−1U(∶, r1)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1Q−1τLM1+r1

(By Lemma 6.2) (6.38)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1Q−1QB(∶, J2 +Lr1)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1B(∶, J2 +Lr1)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯N(σr1,r2I −A)−1B(∶,Mr2 +Lr2)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯(σr1,r2,...,r3I −A)−1NU(∶,Mr2 +Lr2)

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯(σr1,r2,...,r3I −A)−1Q−1τLU−1U(∶,Mr2 +Lr2)

Then by Lemma 6.2)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 147

=c(σr1,...,ruI −A)−1N(σr1,...,ru−1I −A)−1N⋯(σr1,r2,...,r3I −A)−1Q−1QB(∶, Jr3 +Lr2)

⋮ (Reapplying Lemma 6.2 at each step)

=c(σr1,...,ruI −A)−1NUMu−1+Lr1,...,ru−1

=c(σr1,...,ruI −A)−1Q−1τLU−1UMu−1+Lr1,...,ru−1

=c(σr1,...,ruI −A)−1Q−1QB(∶, Ju +Lr1,...,ru−1)

=c(σr1,...,ruI −A)−1B(∶, Ju +Lr1,...,ru−1)

=eT1C(σr1,...,ruI −A)−1BrMu+Lr1,...,ru

=eT1wMu+Lr1,...,ru

=φr1,...,ru (6.39)

A proof for every interpolation constraint in U follows similarly, also using Lemma 6.2. Recall

that for a pair (σr1,...,ru ⊗ γl1,...,lm , η(r1,...,ru),(l1,...,lm)) ∈ T, σr1,...,ru ⊗ γl1,...,lm corresponds to the

point

(σr1 , σr1,r2 , . . . , σr1,...,ru , γl1,...,lm , γl1,...,lm−1 , . . . , γl1) ∈ Cm+u,

and η(r1,...,ru),(l1,...,lm) is the value of the subsystem Hu+m(s1, . . . , su+m) evaluated at this point.

By the first part of the proof note that

Hu+m(σr1,...,ru ⊗ γl1,...,lm)

=C(Jm +Ll1,...,lm−1 , ∶)(γl1,...,lmI −A)−1N(σr1,...,ruI −A)−1B(Ju +Lr1,...,ru−1)

=Q(Mm +Ll1,...,lm , ∶)Q−1τLU−1U(∶,Mu +Lr1,...,ru)

=eTMm+Ll1,...,lmτLeMu+Ll1,...,lu

=τL(Mm +Ll1,...,lm ,Mu +Lr1,...,ru)

=η(r1,...,ru),(l1,...,lm) (6.40)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 148

The case k = q and rank(L(1 ∶ q,1 ∶ q)) = q

Now we will assume that rank(L) = q <M , and rank(L(1 ∶ q,1 ∶ q)) = q. In this case,A ∈ Cq×q,

E constructed according to Theorem 6.1 must also be rank q. Thus E is invertible, and we

make the reassignment

A =E−1A

B =E−1B (6.41)

C =C

but L ∈ CM×M , so to construct a bilinear realization with conformable dimensions, we will

make use of the information in the submatrix τL(1 ∶ q,1 ∶ q) and the submatrix L(1 ∶ q,1 ∶ q).

First let U , Q∈ Cq×q solve

UΛ(1 ∶ q,1 ∶ q) −AU −B = 0 QTM(1 ∶ q,1 ∶ q) −ATQT −CT = 0 (6.42)

The matrices U , and Q, are a factorization of L(1 ∶ q,1 ∶ q).

Theorem 6.5. Let L ∈ CM×M be the Loewner matrix formed from the interpolation data Λ,

R,W ,M ,L,V , and assume that rank(L) = q <M . Then L(1 ∶ q,1 ∶ q) = −QU .

Remark 6.2. From the assumption that rank(L(1 ∶ q,1 ∶ q) = q, it follows from Theorem

6.3 that Q and U are full rank and therefore invertible.

Proof. Let {xi,j} = QU and {li,j} = L(1 ∶ q,1 ∶ q) for i, j = 1, . . . , q. Let qi be the ith

row of Q and let uj be the jth row of U for i, j = 1, . . . , q. By the construction of Q,

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 149

and U , the ith row of Q corresponds to the row vector C(i, ∶)(γl1,...,lmI −A)−1, for some

γl1,...,lm corresponding to the entry M(i, i). There is a one-to-one correspondence between

the indices i and the indices Jm + Ll1,...,lm for m = 1, . . . ,N − 1 and lm = 1 . . . , km. So we

can write C(∶, i) = C(∶, Jm + Ll1,...,lm−1). Likewise, uj corresponds to the column vector

(σr1,...,ruI−A)−1B(∶, j), for some σr1,...,ru corresponding to the entry Λ(j, j) and using the fact

that there is a one-to-one correspondence between the indices j and the indices Ju+Lr1,...,ru−1

for u = 1, . . . ,N − 1 and ru = 1, . . . ku, B(∶, j) =B(∶, Ju +Lr1,...,ru−1). So,

xi,j =C(∶, Jm +Ll1,...,lm−1)(γl1,...,lmI −A)−1(σr1,...,ruI −A)−1B(Ju +Lr1,...,ru−1) (6.43)

= lMm+Ll1,...,lmC(γl1,...,lmI −A)−1(σr1,...,ruI −A)−1BrMu+Lr1,...,ru

(6.44)

=lMm+Ll1,...,lm

(C(γl1,...,lmI −A)−1B −C(σr1,...,ruI −A)−1B)rMu+Lr1,...,ru

σr1,...,ru − γl1,...,lm(6.45)

=vMm+Ll1,...,lm

rMu+Lr1,...,ru

− lMm+Ll1,...,lmwMu+Lr1,...,ru

σr1,...,ru − γl1,...,lm(6.46)

=virj − liwj

σj − γi(6.47)

= li,j (6.48)

As in Lemma 6.2, the following lemma connects the concatenation matrix τL with the action

of U and Q on B, and C respectively.

Lemma 6.2. Assume that L(1 ∶ q,1 ∶ q) is full rank, and let Q,U be given by (6.42). Let

A, B, Cbe given by (6.41). Then

1. QB(∶, Ju+1 +Lr1,...,ru) = τL(∶,Mu +Lr1,...,ru), for u = 1, . . . ,N − 1 and M1 = 0, Lr1 = r1,

rm = 1, . . . , km for m = 1, . . . ,N − 1.

2. (C(Jm+1 +Ll1,...,lm)TU = τL(Mm +Ll1,...,lm , ∶), for m = 1, . . . ,N − 1 and M1 = 0, Ll1 = l1,

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 150

lm = 1, . . . , km for m = 1, . . . ,N − 1

Proof. The proof is exactly the same as in the proof of Lemma 6.2, except the bars are

removed from all the obvious quantities.

We now have all the components necessary to construct the bilinear realization. Define

b =B(∶,1) (6.49)

c =C(1, ∶) (6.50)

N =Q−1τLU−1. (6.51)

Theorem 6.6. Given the interpolation constraints P defined in (6.5), assume that rank(L) =

q and let A, B, C be defined as in (6.41), U , Q be given by (6.42), and b, c, N be

given by (6.49), (6.50), (6.51) respectively. Then ζ:=(A,N ,b,c) satisfies all the subsystem

interpolation conditions in the array P.

Proof. The proof is the same as in Theorem 6.4, except all the bars are dropped from the

obvious quantities.

The following algorithm summarizes the approach for constructing a bilinear system that

interpolates the data.

Algorithm 6.1 (Bilinear realization from transfer function data).

Input: P = T ∪ S ∪U

Output: A,N ,b,c

1. Construct the matrices σ, γ, and η

τL = [η1, . . . , ηk1 , η1,1, η1,2, . . . , η1,k2 , η2,1, . . . , ηk1,k2 , . . . , η2,1 , . . . , ηk1,...,kN ] from S, U, T.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 151

2. Λ = diag(σ), M = diag(γ), R = [eJm+Ll1,...,lm−1 ], for m = 1, . . . ,N and lm = 1, . . . , km.

L =RT

3. For m = 1, . . . ,N and for lm = 1, . . . , km

V = [V ;µl1,...,lm , τ∗L(Mm +Ll1,...,lm ,1 ∶MN)] (6.52)

W = [W , φl1,...,lm , τL(1 ∶MN ,Mm +Ll1,...,lm)] (6.53)

4. Solve

LΛ −ML = LW −V R σLΛ −MσL = LWΛ −MVR (6.54)

5. Compute

λ1,1L − σL = Y ΣX

where rank(λ1,1L − σL) = rank(Σ) = size(Σ) = k and Y , ∈ Cν×k, X ∈ Ck×ρ.

6. E = −Y ∗LX∗, A =E−1(−Y ∗σLX∗), B =E−1(Y ∗V ), C=WX∗.

7. If (rank(L) =M)

a.) Solve

UΛ −AU = B QTM −AT QT = CT

b.) N = Q−1τLU−1, b =B(∶,1), c =C(1, ∶)

8. Else If (rank(L)=q)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 152

a.) Solve

UΛ(1 ∶ q,1 ∶ q) −AU =B QTM(1 ∶ q,1 ∶ q) −ATQT =CT

b.) N =Q−1τLU−1, b =B(∶,1), c =C(1, ∶)

9. Else

a. Assumption (6.15) is not satisfied, and no bilinear realization can be constructed.

10. Return A, N , b, c

Examples

A few simple examples illustrate the method of Algorithm 6.1.

Example 1

In this example, we reconstruct the bilinear system ζ = (A,N ,b,c) where

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−1 0

0 −2

⎤⎥⎥⎥⎥⎥⎥⎦

N =

⎡⎢⎢⎢⎢⎢⎢⎣

1 1

1 1

⎤⎥⎥⎥⎥⎥⎥⎦

b =

⎡⎢⎢⎢⎢⎢⎢⎣

1

0

⎤⎥⎥⎥⎥⎥⎥⎦

c = [1 2] (6.55)

H1(s1) is sampled at the points H1(0) and H1(1). The second-order transfer function

H2(s1, s2) is sampled at the points σi = (1, i + 1) , and γi = (0, 1i+1) for i = 1, . . . ,4. Thus,

k1 = 1 and k2 = 4. This yields 2M = 2(k1 + k1k2) = 10 left and right interpolation constraints.

The dimension of the input-output space for the MIMO linear system will be q = 1 + k1 = 2.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 153

The interpolation constraints in S and U and T are

S = {(1,1/2), ((1,2),5/12)), ((1,3),13/40), ((1,4),4/15), ((1/5),19/84)},

U = {(0,1), ((0,1/2),4/3), ((0,1/3),3/2), ((0,1/4),8/5), ((0,1/5),5/3)},

and

T ={((1,0),1), ((1,1/2,0),16/15), ((1,1/3,0),33/28), ((1,1/4,0),56/45), ((1,1/5,0),85/66),

((1,2,0),7/12), ((1,2,1/2,0),38/45), ((1,2,1/3,0),11/16), ((1,2,1/4,0),98/135),

((1,2,1/5,0),595/792), ((1,3,0),9/20), ((1,3,1/2,0),12/25), ((1,3,1/3,0),297/560),

((1,3,1/4,0),14/25), ((1,3,1/5,0),51/88), ((1,4,0),11/30), ((1,4,1/2,0),88/225),

((1,4,1/3,0),121/280), ((1,4,1/4,0),308/675), ((1,4,1/5,0),17/36), ((1,5,0),13/42),

((1,5,1/2,0),104/315), ((1,5,1/3,0),143/392), ((1,5,1/4,0),52/135),

((1,5,1/5,0),1105/2772)} (6.56)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 154

The matrices σ,γ, and η, are

σ =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1

2

3

4

5

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

T

γ =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0

1/2

1/3

1/4

1/5

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

T

η =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1

16/15

33/28

56/45

85/66

7/12

28/45

11/16

98/135

595/792

9/20

12/25

297/560

14/25

51/88

11/30

88/225

121/280

308/675

17/36

13/42

104/315

143/392

52/135

407/1021

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

T

. (6.57)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 155

The concatenation matrix τL is

τL =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 7/12 9/20 11/30 13/42

16/15 28/45 12/25 88/225 104/315

33/28 11/16 297/560 121/280 143/392

56/45 98/135 14/25 308/675 52/135

85/66 595/792 51/88 17/36 407/1021

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(6.58)

The left and right tangential interpolation conditions are given below:

Λ = diag([1,2,3,4,5]) R =

⎡⎢⎢⎢⎢⎢⎢⎣

1 0 0 0 0

0 1 1 1 1

⎤⎥⎥⎥⎥⎥⎥⎦

W =

⎡⎢⎢⎢⎢⎢⎢⎣

1/2 5/12 13/40 4/15 19/84

1 35/6 9/20 11/30 13/42

⎤⎥⎥⎥⎥⎥⎥⎦

M = diag([0,1/2,1/3,1/4,1/5]) L =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 0

0 1

0 1

0 1

0 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

V =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 1

4/3 16/15

3/2 33/28

8/5 56/45

5/3 85/66

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

This tangential interpolation data yields

L =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−1/2 −7/24 −9/40 −11/60 −13/84

−2/3 −29/90 −37/150 −1/5 −53/315

−3/4 −5/14 −153/560 −31/140 −73/392

−4/5 17/45 −13/45 −158/675 −62/315

−5/6 −155/396 −79/264 −8/33 −565/2772

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(6.59)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 156

σL =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1/2 5/12 13/40 4/15 19/84

2/3 19/45 49/150 4/15 71/315

3/4 13/28 201/560 41/140 97/392

4/5 22/45 17/45 208/675 82/315

5/6 50/99 103/264 7/22 745/2772

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(6.60)

and rank(L) = rank(σL) = 2. One can easily check that assumption (6.1) is satisfied for

k = 2. Thus, we construct the linear system E ∈ R2×2, A ∈ R2×2, B ∈ R2×2 and C ∈ R2×2

according to Theorem 6.1 that satisfies all the tangential interpolation constraints. E is

invertible, so we give the realization parameters A←E−1A, B ←E−1B, C below:

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−4399/3742 931/4489

506/725 −6827/3742

⎤⎥⎥⎥⎥⎥⎥⎦

B =

⎡⎢⎢⎢⎢⎢⎢⎣

2231/1462 2215/1514

571/442 −2141/1002

⎤⎥⎥⎥⎥⎥⎥⎦

C =

⎡⎢⎢⎢⎢⎢⎢⎣

1825/2307 −229/1428

221/166 −321/13126

⎤⎥⎥⎥⎥⎥⎥⎦

The state-space transformation

T =

⎡⎢⎢⎢⎢⎢⎢⎣

2231/1462 8809/6292

571/442 −2727/490

⎤⎥⎥⎥⎥⎥⎥⎦

applied to this realization gives

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−1 0

0 −2

⎤⎥⎥⎥⎥⎥⎥⎦

B =

⎡⎢⎢⎢⎢⎢⎢⎣

1 1/2

0 1/2

⎤⎥⎥⎥⎥⎥⎥⎦

C =

⎡⎢⎢⎢⎢⎢⎢⎣

1 2

2 2

⎤⎥⎥⎥⎥⎥⎥⎦

The tangential interpolation data falls into the case where the rank(L) = 2 = q, the number

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 157

of inputs, and rank(L(1 ∶ q,1 ∶ q)) = 2. Hence to find N , we solve

UΛ(1 ∶ 2,1 ∶ 2) −AU =B QTM(1 ∶ 2,1 ∶ 2) −ATQT =CT

for U , and Q. This gives

U =

⎡⎢⎢⎢⎢⎢⎢⎣

1/2 1/6

0 1/8

⎤⎥⎥⎥⎥⎥⎥⎦

Q =

⎡⎢⎢⎢⎢⎢⎢⎣

1 1

4/3 4/5

⎤⎥⎥⎥⎥⎥⎥⎦

,

and therefore

N =Q−1τL(1 ∶ 2,1 ∶ 2)U−1 =

⎡⎢⎢⎢⎢⎢⎢⎣

1 1

1 1

⎤⎥⎥⎥⎥⎥⎥⎦

Thus, we assign b =

⎡⎢⎢⎢⎢⎢⎢⎣

1

0

⎤⎥⎥⎥⎥⎥⎥⎦

=B(∶,1) and c = [1 2] =C(1, ∶) and recover the bilinear realization

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−1 0

0 −2

⎤⎥⎥⎥⎥⎥⎥⎦

N =

⎡⎢⎢⎢⎢⎢⎢⎣

1 1

1 1

⎤⎥⎥⎥⎥⎥⎥⎦

b =

⎡⎢⎢⎢⎢⎢⎢⎣

1

0

⎤⎥⎥⎥⎥⎥⎥⎦

c = [1 2] (6.61)

In this example, it was possible to reconstruct the bilinear system ζ exactly from the data.

Our goal was basically to construct the linear MIMO system with realization

A =

⎡⎢⎢⎢⎢⎢⎢⎣

−1 0

0 −2

⎤⎥⎥⎥⎥⎥⎥⎦

B = [b N(I −A)−1b] C =

⎡⎢⎢⎢⎢⎢⎢⎣

c

c(0I −A)−1N

⎤⎥⎥⎥⎥⎥⎥⎦

The system with this realization had exactly 10 free parameters, and we stipulated 10 linearly

independent tangential interpolation conditions, so we were able to construct exactly this

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 158

realization. From this realization, we used the information in τL to reconstruct N .

Example 2

In this example we apply our realization strategy as a model reduction method. The model

ζ to be reduced has the realization

A =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−1 0 1/6 0

0 −2 0 1/4

0 0 −3 0

0 0 0 −4

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

N =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 0 0

2 0 0 0

0 3 0 0

0 0 4 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

b =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1

0

1

0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

c = [1 0 1 0]

The interpolation data was acquired by sampling the subsystem transfer functions

Hk(s1, . . . , sk) of ζ at particular points. The points sampled and the subsystem transfer

function evaluations are

S ={(1,37/48), ((1,1),0), ((1,1/2),0)}

U ={(2,49/40), ((2,2),0), ((4,2),0)}

T ={((1,2),0), ((1,2,2),247/1440), ((1,4,2),382/3373), ((1,1,2),463/2009),

((1,1,2,2),0), ((1,1,4,2),0), ((1,1/2,2),667/2173), ((1,1/2,2,2),0), ((1,1/2,4,2),0)}

For this example, k1 = 1 and k2 = 2, so M = k1 + k1k2 = 3 and there will be q = 1 + k1 = 2

inputs and outputs for the MIMO linear realization. The concatenation matrix is

τL =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 463/2009 667/2173

247/1440 0 0

382/3373 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(6.62)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 159

The left and right interpolation data is

Λ =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 0 0

0 1 0

0 0 1/2

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

R =

⎡⎢⎢⎢⎢⎢⎢⎣

1 0 0

0 1 1

⎤⎥⎥⎥⎥⎥⎥⎦

W =

⎡⎢⎢⎢⎢⎢⎢⎣

37/48 0 0

0 463/2009 917/3299

⎤⎥⎥⎥⎥⎥⎥⎦

M =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

2 0 0

0 2 0

0 0 4

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

L =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 0

0 1

0 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

V =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

49/90 0

0 247/1440

0 382/3373

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

This tangential interpolation data gives

L =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−163/720 0 0

0 −621/10537 −376/5299

0 −79/2022 −405/8606

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, σL =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

229/720 0 0

0 76/675 551/4050

0 356/4799 323/3600

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

The rank of L = 3, so this example falls in the case where rank(L) = M . We construct the

linear system E ∈ R3×3, A ∈ R3×3, B ∈ R3×3 and C ∈ R3×3 according to Theorem 6.1 that

satisfies all the tangential interpolation constraints. E is invertible, so we give the realization

parameters A←E−1A, B ←E−1B, C below:

A =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−229/163 0 0

0 −944/511 83/13260

0 −8399/160 −2122/511

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

, B =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−392/163 0

0 2520/1391

0 8128/217

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

,

C =

⎡⎢⎢⎢⎢⎢⎢⎣

−37/48 0 0

0 295/817 65/436442

⎤⎥⎥⎥⎥⎥⎥⎦

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 160

From the realization data we next compute

U =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−1 0 0

0 659/1033 1745/2266

0 1745/2266 −659/1033

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

Q =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−163/720 0 0

0 775/8402 5/42356

0 272/4447 5/76653

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

Finally we have

N = Q−1τLU−1 =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 −2857/1687 149/1839

−2520/1391 0 0

−8128/217 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

Thus we obtain the reduced order model ζr of order r = 3 with realization A = A, N ,

b = B(∶,1), c = C(1, ∶). Note that one can easily check that the model ζr we constructed

using the realization procedure of Algorithm 6.1 is the same reduced-order model obtained by

constructing a subsystem interpolant at the same collection of interpolation points according

to Theorem 3.2.

6.4 Volterra kernel sampling methods

Let Ψ be a nonlinear dynamical system and assume that it can be well approximated by

the first k terms of a Volterra series of a bilinear system. This situation frequently arises

in the modeling of weakly nonlinear circuits [100], [81], where the dominant characteristics

of the system output are captured by the first few kernels the information in the remaining

kernels is indistinguishable from the measurement error. In this case, multi-tonal inputs can

be used to sample the Volterra kernels and construct a bilinear realization of a model ζ that

approximates Ψ.

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 161

The multi-tonal inputs used are of the form

u(t) =`

∑j=1

αjeıβjt.

The response in the kth regular kernel is given as

yk(t) =`

∑j1

⋯`

∑jk

(αj1⋯α

jk)et(

k

∑i=1

ıβji)Hk(ıβj1 , ıβj1 + ıβj2 , . . . ,k

∑i=1

ıβji) (6.63)

The first step in measuring the Volterra kernels is to separate the response y(t) into its kernel

components. Since Ψ has a Volterra series representation, the mapping y(t) = Ψ(αu(t))

behaves like a low-order polynomial in α. So suppose for some choice of distinct α′is, i =

1, . . . , k we apply the signal αiu(t) to Ψ. Then for the corresponding system responses

ri(t), i = 1, . . . , k we have

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

r1

r2

rk

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

=

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

α1 α21 . . . αk1

α2 α22 . . . αk2

⋮ ⋱ ⋮

αk α2k . . . αkk

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

y1

y2

yk

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

+

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

e1

e2

ek

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

(6.64)

This is nothing but a simple polynomial interpolation problem, and the approach is first

mentioned in [56] and independently in [86]. In general, it is quite accurate for low order

components of the response (the first and possibly the second term) and abysmally inaccurate

for higher-order components. The situation can be improved by using clever frequency

separating techniques [23]. As a very simple example, assume that the input has the form

u(t) = 2R(L

∑`=1

βi exp(ı`t)) (6.65)

Garret M. Flagg Chapter 6. Data-Driven Model Reduction of SISO Bilinear Systems 162

and that it is odd, meaning βi = 0 for i even. Then from equation (6.63), the odd and

even order responses only occur at odd and even order frequencies, respectively. This means

that, for example, to compute the second-order response at an even frequency, there is no

dependence on the odd-order terms, and so the second-order coefficient will dominate and

can be accurately computed. Further details regarding frequency sampling techniques can

be found in [23], and the references therein.

Chapter 7

Conclusions

7.1 A summary of contributions

The two main theoretical contributions of this work are the development of a multi-point

interpolation framework applicable to the entire Volterra series representation of a bilin-

ear system, and the solution of a rational interpolation-based bilinear realization problem.

Practically speaking, our multi-point interpolation method provided the insight necessary

to greatly simplify the construction of H2 reduced order models (in the asymptotic sense)

using the algorithm TB-IRKA. Our analysis shows that TB-IRKA approximations are local

H2 optimal approximations to polynomials systems generated by truncating the Volterra se-

ries representation after N terms. We have shown through several examples that TB-IRKA

yields reduced order models with accuracy comparable to the exact solutions of the H2 op-

timal bilinear model reduction problem computed using the algorithm B-IRKA. Moreover,

TB-IRKA requires the solution of a small number of ordinary Sylvester equations per itera-

tion and is therefore significantly cheaper than B-IRKA as the order r of the reduced order

model increases. In the development of the our new multi-point interpolation framework for

163

Garret M. Flagg Chapter 7. Conclusions 164

bilinear systems, we provided a detailed analysis of the kth order transfer functions of SISO

bilinear systems, deriving their pole-residue decomposition and using this decomposition to

develop a pole-residue expression of the H2-norm that plainly generalizes the expression for

the H2 norm of LTI systems.

The realization problem we posed and solved makes it possible to construct bilinear re-

alizations of weakly nonlinear systems using natural sampling inputs such as multi-tonal

sinusoidal inputs. Since it corresponds to the subsystem interpolation method, it also makes

it possible to construct reduced order model directly from data on the full order system,

without necessarily having to form the projection matrices given in Theorem ??. The

realization results also provide a generalization of the solution to the SISO LTI rational-

interpolation/realization problem given in terms of the Loewner and shifted Loewner matri-

ces to the case of bilinear systems.

We also considered the solution of ordinary and bilinear Sylvester equations. Our analysis

showed that the ADI method for solving ordinary Sylvester equations is in fact exactly

equivalent to the rational Krylov projection method for the specially chosen pseudo-H2

optimal shifts. This analysis let to a a new proof of the result that the Sylvester equation

residual is orthogonal to the projection subspace for the special case of pseudo-H2 optimal

shifts, and that moreover, this choice of shifts yields nearly optimal rank r approximations to

the solution of the Lyapunov equations. We then generalized these orthogonality results to

the bilinear case, showing that the choice of shifts and weights in the multi-point interpolation

problem that correspond to satisfying the first condition of Theorem 4.7 yields a projection

subspace that is orthogonal to the residual in the bilinear Lyapunov equations. Finally, we

derived a new bilinear model of a nonlinear heat transfer problem that we hope will be used

as a test example for further developments in the model reduction of bilinear systems.

Garret M. Flagg Chapter 7. Conclusions 165

7.2 Directions for future work

A major challenge of bilinear model reduction is the ubiquitous requirement of solving the

bilinear Sylvester equations. A further development of rational Krylov projection approach

for solving the bilinear Sylvester equations that more deeply develops optimal choices of

shifts and weights in the multi-point Volterra series problem may help significantly reduce

the cost of solving these equations.

The fuller development of the scope and applicability of bilinear models also needs to be

developed further, now that there are techniques for reducing their dimension. The model

reduction techniques that have been developed for bilinear models so far will hopefully

open up more opportunities to use them appropriately in modeling, and employing them

in parameter dependent partial differential equations as in the Fokker-Planck model is in

interesting route for further investigation.

Finally, a deeper development of the sampling techniques necessary to construct a bilinear

realization, as well as the development of a sampling algorithm that would achieve this goal

on real examples would be an important future work.

Bibliography

[1] S.A. Al-Baiyat and M. Bettayeb. A new model reduction scheme for k-power bilinear

systems. In Decision and Control, 1993., Proceedings of the 32nd IEEE Conference

on, pages 22–27. IEEE, 1993.

[2] A.C. Antoulas. Approximation of Large-Scale Dynamical Systems (Advances in Design

and Control). Society for Industrial and Applied Mathematics, Philadelphia, PA, USA,

2005.

[3] AC Antoulas and B.D.O. Anderson. On the scalar rational interpolation problem.

IMA Journal of Mathematical Control and Information, 3(2-3):61–68, 1986.

[4] AC Antoulas, JA Ball, J. Kang, and JC Willems. On the solution of the minimal

rational interpolation problem. Linear Algebra and its Applications, 137:511–573, 1990.

[5] Z. Bai. Krylov subspace techniques for reduced-order modeling of large-scale dynamical

systems. Applied Numerical Mathematics, 43(1-2):9–44, 2002.

[6] Z. Bai and D. Skoogh. A projection method for model reduction of bilinear dynamical

systems. Linear algebra and its applications, 415(2-3):406–425, 2006.

[7] J. Bak and D.J. Newman. Complex analysis. Springer Verlag, 2010.

166

Garret M. Flagg Bibliography 167

[8] J.A. Ball, G. Groenewald, and T. Malakorn. Structured noncommutative multidimen-

sional linear systems. SIAM Journal on Control and Optimization, 44:1474, 2005.

[9] L. Bao, Y. Lin, and Y. Wei. A new projection method for solving large Sylvester

equations. Applied numerical mathematics, 57(5-7):521–532, 2007.

[10] R.H. Bartels and GW Stewart. Algorithm 432: Solution of the matrix equation AX+

XB= C. Communications of the ACM, 15(9):820–826, 1972.

[11] U. Baur, C. Beattie, P. Benner, and S. Gugercin. Interpolatory projection methods

for parameterized model reduction. SIAM Journal on Scientific Computing, 33:2489,

2011.

[12] B. Beckermann, S. Guttel, and R. Vandebril. On the convergence of rational Ritz

values. SIAM Journal on Matrix Analysis and Applications, 31(4):1740–1774, 2010.

[13] V. Belevitch. Interpolation matrices. Philips Res. Rep, 25:337–369, 1970.

[14] P. Benner and T. Breiten. Krylov-Subspace Based Model Reduction of Nonlinear

Circuit Models Using Bilinear and Quadratic-Linear Approximations. Progress in In-

dustrial Mathematics at ECMI, 2010.

[15] P. Benner and T. Breiten. On h2-model reduction of linear parameter-varying systems.

Proceedings in Applied Mathematics and Mechanics, 11:805–806, 2011.

[16] P. Benner and T. Damm. Lyapunov equations, energy functionals, and model order

reduction of bilinear and stochastic systems. SIAM Journal on Control and Optimiza-

tion, 49:686, 2011.

[17] P. Benner, R.-C. Li., and N Truhar. On the ADI method for Sylvester equations.

Journal of Computational and Applied Mathematics, 233(4):1035 – 1045, 2009.

Garret M. Flagg Bibliography 168

[18] P. Benner, E. S. Quintana-Ortı, and G. Quintana-Ortı. State-Space Truncation Meth-

ods for Parallel Model Reduction of Large-Scale Systems. Parallel Computing, special

issue on “Parallel and Distributed Scientific and Engineering Computing”, 29:1701–

1722, 2003.

[19] P. Benner and J. Saak. Efficient numerical solution of the LQR-problem for the heat

equation. Proc. Appl. Math. Mech, 4(1):648–649, 2004.

[20] Peter Benner and Tobias Breiten. Interpolation-based H2-model reduction of bilinear

control systems. Technical Report MPIMD/11-02, Max Planck Institute Magdeburg

Preprints, June 2011.

[21] Peter Benner and Tobias Breiten. On optimality of interpolation-based low-rank

approximations of large-scale matrix equations. Max Planck Institute Magdeburg

Preprints, December, 2011.

[22] S. Bochner and K. Chandrasekharan. Fourier transforms. Number 19. Princeton

University Press, 1949.

[23] S. Boyd, YS Tang, and L. Chua. Measuring volterra kernels. Circuits and Systems,

IEEE Transactions on, 30(8):571–577, 1983.

[24] T. Breiten and T. Damm. Krylov subspace methods for model order reduction of

bilinear control systems. Systems & Control Letters, 2010.

[25] Tobias Breiten. Krylov Subspace Methods for Model Order Reduction of Bilinear

Control Systems. Master’s thesis, Technical University of Kaiserslautern, Department

of Mathematics, November 2009.

[26] R.W. Brockett. On the algebraic structure of bilinear systems(algebraic lie structure

Garret M. Flagg Bibliography 169

theory of bilinear systems in terms of controllability, observability and equivalent re-

alization). Theory and applications of variable structure systems, (2):153–168, 1972.

[27] R.W. Brockett. Volterra series and geometric control theory. Automatica, 12(2):167–

176, 1976.

[28] C. Bruni, G. Dipillo, and G. Koch. Bilinear systems: An appealing class of nearly

linear systems in theory and applications. Automatic Control, IEEE Transactions on,

19(4):334–348, 1974.

[29] B. Schaeffer-Bung C. Hartmann and A. Zueva. Balanced model reduction of bilinear

systems with applications to positive systems. submitted to SIAM J. Control and

Optimization, 2010.

[30] Y. Chahlaoui and P. Van Dooren. Benchmark examples for model reduction of lin-

ear time-invariant dynamical systems. Dimension Reduction of Large-Scale Systems,

45:381–395, 2005.

[31] S. Chaturantabut and D.C. Sorensen. Nonlinear model reduction via discrete empirical

interpolation. SIAM Journal on Scientific Computing, 32:2737, 2010.

[32] Y. Chen, J. White, et al. A quadratic method for nonlinear model order reduction.

2000.

[33] E.G. Collins Jr and A.S. Hodel. Efficient solution of linearly coupled lyapunov equa-

tions. SIAM Journal on Matrix Analysis and Applications, 18:291, 1997.

[34] M. Condon and R. Ivanov. Krylov subspaces from bilinear representations of nonlinear

systems. COMPEL: The International Journal for Computation and Mathematics in

Electrical and Electronic Engineering, 26(2):399–406, 2007.

Garret M. Flagg Bibliography 170

[35] I.J. Couchman, E.C. Kerrigan, and C. Bohm. Model reduction of homogeneous-in-the-

state bilinear systems with input constraints. Automatica, 2011.

[36] P. D’Alessandro, A. Isidori, and A. Ruberti. Realization and structure theory of bilinear

dynamical systems. SIAM Journal on Control, 12:517, 1974.

[37] T. Damm. Direct methods and adi-preconditioned krylov subspace methods for gener-

alized lyapunov equations. Numerical Linear Algebra with Applications, 15(9):853–871,

2008.

[38] V. Druskin, L. Knizhnerman, and V. Simoncini. Analysis of the Rational Krylov

Subspace and ADI Methods for Solving the Lyapunov Equation. SIAM Journal on

Numerical Analysis, 49(5):1875–1898, 2011.

[39] V. Druskin, L. Knizhnerman, and M. Zaslavsky. Solution of large scale evolutionary

problems using rational Krylov subspaces with optimized shifts. SIAM Journal on

Scientific Computing, 31(5):3760–3780, 2009.

[40] V. Druskin, C. Lieberman, and M. Zaslavsky. On adaptive choice of shifts in ratio-

nal krylov subspace reduction of evolutionary problems. SIAM Journal on Scientific

Computing, 2010.

[41] V. Druskin and V. Simoncini. Adaptive rational krylov subspaces for large-scale dy-

namical systems. Systems & Control Letters, 60:546–560, 2011.

[42] A. El Guennouni, K. Jbilou, and AJ Riquet. Block Krylov subspace methods for

solving large Sylvester equations. Numerical Algorithms, 29(1):75–96, 2002.

[43] G. Flagg, C. Beattie, and S. Gugercin. Interpolatory H∞ model reduction. Arxiv

preprint arXiv:1107.5364, 2011.

Garret M. Flagg Bibliography 171

[44] G. M. Flagg. H2-optimal interpolation: New properties and applications, July, 2010.

Talk given at the 2010 SIAM Annual Meeting, Pittsburgh (PA).

[45] G.M. Flagg, S. Gugercin, and C.A. Beattie. An interpolation-based approach to H∞

model reduction of dynamical systems. In Decision and Control (CDC), 2010 49th

IEEE Conference on, pages 6791–6796. IEEE, 2010.

[46] M. Fliess. Sur la realization des systemes dynamiques bilineaires. CR Acad. Sc. Paris

A, 277:243–247, 1973.

[47] M. Fliess. Series de Volterra et series formelles non commutatives. Comptes Rendus

Acad. Sciences Paris, 280:965–967, 1975.

[48] M. Fliess. A remark on the transfer functions and the realization of homoge-

neous continuous-time nonlinear systems. Automatic Control, IEEE Transactions on,

24(3):507–508, 1979.

[49] A.E. Frazho. A shift operator approach to bilinear system theory. SIAM Journal on

Control and Optimization, 18:640, 1980.

[50] K. Fujimoto and J.M.A. Scherpen. Balanced realization and model order reduction

for nonlinear systems based on singular value analysis. SIAM Journal on Control and

Optimization, 48(7):4591–4623, 2010.

[51] K. Gallivan, A. Vandendorpe, and P.V. Dooren. Model reduction of mimo systems

via tangential interpolation. SIAM Journal on Matrix Analysis and Applications,

26(2):328–349, 2005.

[52] E.J. Grimme. Krylov projection methods for model reduction. PhD thesis, University

of Illinois, 1997.

Garret M. Flagg Bibliography 172

[53] S. Gugercin. Projection methods for model reduction of large-scale dynamical systems.

PhD thesis, Ph. D. Dissertation, ECE Dept., Rice University, 2002.

[54] S. Gugercin, A.C. Antoulas, and C. Beattie. H2 model reduction for large-scale linear

dynamical systems. SIAM Journal on Matrix Analysis and Applications, 30(2):609–

638, 2008.

[55] S. Gugercin, D.C. Sorensen, and A.C. Antoulas. A modified low-rank Smith method

for large-scale Lyapunov equations. Numerical Algorithms, 32(1):27–55, 2003.

[56] A. Halme, J. Orava, and H. Blomberg. Polynomial operators in non-linear systems

theory. International Journal of Systems Science, 2(1):25–47, 1971.

[57] M. Heinkenschloss, D.C. Sorensen, and K. Sun. Balanced Truncation Model Reduction

for a Class of Descriptor Systems with Application to the Oseen Equations. SIAM

Journal on Scientific Computing, 30:1038, 2008.

[58] D.Y. Hu and L. Reichel. Krylov-subspace methods for the Sylvester equation. Linear

Algebra and its Applications, 172:283–313, 1992.

[59] A. Isidori. Direct construction of minimal bilinear realizations from nonlinear input-

output maps. Automatic Control, IEEE Transactions on, 18(6):626–631, 1973.

[60] I.M. Jaimoukha and E.M. Kasenally. Krylov subspace methods for solving large Lya-

punov equations. SIAM Journal on Numerical Analysis, pages 227–251, 1994.

[61] K. Jbilou. Low rank approximate solutions to large Sylvester matrix equations. Applied

mathematics and computation, 177(1):365–376, 2006.

[62] RE Kalman. Mathematical Description of Linear Dynamical Systems. SIAM Journal

on Control, 1:152–192, 1963.

Garret M. Flagg Bibliography 173

[63] RE Kalman. Algebraic structure of linear dynamical systems, I. The module of Σ.

Proceedings of the National Academy of Sciences of the United States of America,

54(6):1503, 1965.

[64] RE Kalman, M. Arbib, and P. Falb. Topics in Mathematical Systems Theory. McGraw-

Hill, New York, 1969.

[65] S.G. Krantz. Function theory of several complex variables. Amer Mathematical Society,

2001.

[66] E. Kreyszig. Introductory functional analysis with applications. Wiley, 1989.

[67] J.R. Li and J. White. Low Rank Solution of Lyapunov Equations. SIAM Journal on

Matrix Analysis and Applications, 24(1):260–280, 2002.

[68] J.R. Li and J. White. Low-rank solution of Lyapunov equations. SIAM review, pages

693–713, 2004.

[69] Y. Lin, L. Bao, and Y. Wei. Order reduction of bilinear mimo dynamical systems using

new block krylov subspaces. Computers & Mathematics with Applications, 58(6):1093–

1102, 2009.

[70] AJ Mayo and AC Antoulas. A framework for the solution of the generalized realization

problem. Linear Algebra and its Applications, 425(2-3):634 – 662, 2007.

[71] L. Meier III and D. Luenberger. Approximation of linear constant systems. Automatic

Control, IEEE Transactions on, 12(5):585–588, 1967.

[72] R.R. Mohler. Natural bilinear control processes. Systems Science and Cybernetics,

IEEE Transactions on, 6(3):192–197, 1970.

[73] R.R. Mohler. Bilinear control processes. Academic Press New York, 1973.

Garret M. Flagg Bibliography 174

[74] R.R. Mohler. Nonlinear systems (vol. 2): applications to bilinear control. Prentice-Hall,

Inc. Upper Saddle River, NJ, USA, 1991.

[75] D.W. Peaceman and HH Rachford. The numerical solution of parabolic and elliptic

differential equations. Journal of the Society for Industrial and Applied Mathematics,

3(1):28–41, 1955.

[76] T. Penzl. A cyclic low rank Smith method for large sparse Lyapunov equations. SIAM

Journal on Scientific Comput, 21(4):1401–1418, 2000.

[77] M. Petreczky. Realization theory for linear and bilinear switched systems: A formal

power series approach. ESAIM: Control, Optimization and Calculus of Variations,

17(02):446–471, 2011.

[78] J.R. Phillips. Projection-based approaches for model reduction of weakly nonlinear,

time-varying systems. Computer-Aided Design of Integrated Circuits and Systems,

IEEE Transactions on, 22(2):171–187, 2003.

[79] S. Richter, L.D. Davis, and E.G. Collins Jr. Efficient computation of the solutions

to modified lyapunov equations. SIAM journal on matrix analysis and applications,

14:420, 1993.

[80] W. Rudin. Function theory in polydiscs. Mathematics Lecture Note Series, 1969.

[81] W.J. Rugh. Nonlinear system theory. Johns Hopkins University Press Baltimore, MD,

1981.

[82] A. Ruhe. Rational krylov algorithms for nonsymmetric eigenvalue problems. ii. matrix

pairs. Linear Algebra and its Applications, 197:283–295, 1994.

[83] J. Sabino. Solution of large-scale Lyapunov equations via the block modified Smith

method. PhD thesis, RICE UNIVERSITY, 2007.

Garret M. Flagg Bibliography 175

[84] H. Schneider. Positive operators and an inertia theorem. Numerische Mathematik,

7(1):11–17, 1965.

[85] V. Simoncini. A new iterative method for solving large-scale Lyapunov matrix equa-

tions. SIAM Journal on Scientific Computing, 29(3):1268–1288, 2008.

[86] RJ Simpson and HM Power. Correlation techniques for the identification of nonlinear

systems. Measurement and Control, 5:316–321, 1972.

[87] T. Siu and M. Schetzen. Convergence of Volterra series representation and BIBO

stability of bilinear systems. International journal of systems science, 22(12):2679–

2684, 1991.

[88] E.D. Sontag. Polynomial Response Maps, volume 13 of Lecture Notes in Control and

Information Sciences. Springer Verlag, 1979.

[89] D.C. Sorensen and A.C. Antoulas. The Sylvester equation and approximate balanced

reduction. Linear algebra and its applications, 351:671–700, 2002.

[90] T. Stykel. Gramian-Based Model Reduction for Descriptor Systems. Mathematics of

Control, Signals, and Systems (MCSS), 16(4):297–319, 2004.

[91] T. Stykel and V. Simoncini. Krylov subspace methods for projected Lyapunov equa-

tions. Applied Numerical Mathematics, 2011.

[92] H.J. Sussmann. Semigroup representations, bilinear approximation of input-output

maps, and generalized inputs. Mathematical systems theory, 131:172–191, 1975.

[93] N. Truhar and R.C. Li. On the ADI Method for Sylvester Equations. Technical

report, Technical Report 2008-02, Department of Mathematics, University of Texas at

Arlington, 2008, available at http://www. uta. edu/math/preprint/rep2008 02. pdf,

2007.

Garret M. Flagg Bibliography 176

[94] E. Tyrtyshnikov. Mosaic-skeleton approximations. Calcolo, 33(1):47–57, 1996.

[95] B. Vandereycken and S. Vandewalle. A riemannian optimization approach for comput-

ing low-rank solutions of lyapunov equations. SIAM Journal on Matrix Analysis and

Applications, 31(5):2553–2579, 2010.

[96] C.D. Villemagne and R.E. Skelton. Model reduction using a projection formulation.

Intl. J. Contr, 46:2141–2169, 1987.

[97] E.L. Wachspress. The ADI minimax problem for complex spectra. Applied Mathematics

Letters, 1(3):311–314, 1988.

[98] E.L. Wachspress. Trail to a Lyapunov equation solver. Computers & Mathematics with

Applications, 55(8):1653–1659, 2008.

[99] Wolfgang Walter. Ordinary Differential Equations. Springer, 1998.

[100] D.D. Weiner and J.F. Spina. Sinusoidal Analysis and Modeling of Weakly Nonlinear

Circuits: With Application to Nonlinear Interference Effects. Van Nostrand Reinhold,

1980.

[101] D.A. Wilson. Optimum solution of model-reduction problem. Proc. IEE, 117(6):1161–

1165, 1970.

[102] A. Yousefi, B. Lohmann, J. Lienemann, and JG Korvink. Nonlinear heat transfer

modelling and reduction. In Proceedings of the 12th IEEE Mediterranean Conference

on Control and Automation, 2004.

[103] A. Yousuff and R.E. Skelton. Covariance Equivalent Realizations with Application to

Model Reduction of Large Scale Systems. Control and Dynamic Systems, 22, 1984.

Garret M. Flagg Bibliography 177

[104] A. Yousuff, D.A. Wagie, and R.E. Skelton. Linear system approximation via covariance

equivalent realizations. Journal of mathematical analysis and applications, 106(1):91–

115, 1985.

[105] L. Zhang and J. Lam. On H2 model reduction of bilinear systems. Automatica,

38(2):205–216, 2002.