Balanced realization and model order reduction for nonlinear systems based on singular value analysis

Similar documents
Weighted balanced realization and model reduction for nonlinear systems

Model reduction for linear systems by balancing

The Important State Coordinates of a Nonlinear System

Balanced Truncation 1

arxiv: v1 [math.oc] 29 Apr 2017

1 Relative degree and local normal forms

Stabilization and Passivity-Based Control

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

EE363 homework 7 solutions

Chap. 3. Controlled Systems, Controllability

Balancing of Lossless and Passive Systems

Gramians based model reduction for hybrid switched systems

3 Gramians and Balanced Realizations

Observability. Dynamic Systems. Lecture 2 Observability. Observability, continuous time: Observability, discrete time: = h (2) (x, u, u)

FEL3210 Multivariable Feedback Control

Operator based robust right coprime factorization and control of nonlinear systems

Order Reduction of a Distributed Parameter PEM Fuel Cell Model

ME 234, Lyapunov and Riccati Problems. 1. This problem is to recall some facts and formulae you already know. e Aτ BB e A τ dτ

Input to state Stability

Modern Optimal Control

Nonlinear Control Systems

Applied Math Qualifying Exam 11 October Instructions: Work 2 out of 3 problems in each of the 3 parts for a total of 6 problems.

Solution of Linear State-space Systems

1 The Observability Canonical Form

Linearization problem. The simplest example

Output Feedback and State Feedback. EL2620 Nonlinear Control. Nonlinear Observers. Nonlinear Controllers. ẋ = f(x,u), y = h(x)

Equivalence of dynamical systems by bisimulation

Zeros and zero dynamics

Nonlinear Control Systems

16.31 Fall 2005 Lecture Presentation Mon 31-Oct-05 ver 1.1

Discrete and continuous dynamic systems

CDS Solutions to the Midterm Exam

DESIGN OF OBSERVERS FOR SYSTEMS WITH SLOW AND FAST MODES

only nite eigenvalues. This is an extension of earlier results from [2]. Then we concentrate on the Riccati equation appearing in H 2 and linear quadr

QUALIFYING EXAMINATION Harvard University Department of Mathematics Tuesday September 21, 2004 (Day 1)

LMI Methods in Optimal and Robust Control

ACM/CMS 107 Linear Analysis & Applications Fall 2016 Assignment 4: Linear ODEs and Control Theory Due: 5th December 2016

CDS Solutions to Final Exam

Modeling and Analysis of Dynamic Systems

An introduction to Birkhoff normal form

ME Fall 2001, Fall 2002, Spring I/O Stability. Preliminaries: Vector and function norms

10 Transfer Matrix Models

The first order quasi-linear PDEs

Prof. Krstic Nonlinear Systems MAE281A Homework set 1 Linearization & phase portrait

Converse Lyapunov theorem and Input-to-State Stability

Linear System Theory

Module 07 Controllability and Controller Design of Dynamical LTI Systems

Module 03 Linear Systems Theory: Necessary Background

Reflected Brownian Motion

APPPHYS217 Tuesday 25 May 2010

ẋ n = f n (x 1,...,x n,u 1,...,u m ) (5) y 1 = g 1 (x 1,...,x n,u 1,...,u m ) (6) y p = g p (x 1,...,x n,u 1,...,u m ) (7)

ECEN 605 LINEAR SYSTEMS. Lecture 7 Solution of State Equations 1/77

Grammians. Matthew M. Peet. Lecture 20: Grammians. Illinois Institute of Technology

1 Controllability and Observability

Linear dynamical systems with inputs & outputs

Model Reduction for Unstable Systems

Introduction to Geometric Control

Control Systems Design

6.241 Dynamic Systems and Control

Robust Control 2 Controllability, Observability & Transfer Functions

Trajectory Tracking Control of Bimodal Piecewise Affine Systems

Chapter III. Stability of Linear Systems

Nonlinear Observers. Jaime A. Moreno. Eléctrica y Computación Instituto de Ingeniería Universidad Nacional Autónoma de México

Math Ordinary Differential Equations

1. Find the solution of the following uncontrolled linear system. 2 α 1 1

Published in: Proceedings of the 2nd IFAC Workshop on Lagrangian and Hamiltonian Methods for Nonlinear Control

Chapter 6 Balanced Realization 6. Introduction One popular approach for obtaining a minimal realization is known as Balanced Realization. In this appr

6 Linear Equation. 6.1 Equation with constant coefficients

3118 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 57, NO. 12, DECEMBER 2012

Nonlinear Control Lecture # 14 Input-Output Stability. Nonlinear Control

Theorem 1. ẋ = Ax is globally exponentially stable (GES) iff A is Hurwitz (i.e., max(re(σ(a))) < 0).

2. Dual space is essential for the concept of gradient which, in turn, leads to the variational analysis of Lagrange multipliers.

NOTES ON LINEAR ODES

ECE504: Lecture 8. D. Richard Brown III. Worcester Polytechnic Institute. 28-Oct-2008

Identification Methods for Structural Systems

Deterministic Dynamic Programming

2. Review of Linear Algebra

Pseudospectra and Nonnormal Dynamical Systems

Continuous Dynamics Solving LTI state-space equations גרא וייס המחלקה למדעי המחשב אוניברסיטת בן-גוריון

ECEEN 5448 Fall 2011 Homework #5 Solutions

Numerical Algorithms as Dynamical Systems

Empirical Gramians and Balanced Truncation for Model Reduction of Nonlinear Systems

Decay rates for partially dissipative hyperbolic systems

Department of Mathematics, University of California, Berkeley. GRADUATE PRELIMINARY EXAMINATION, Part A Fall Semester 2016

10. Smooth Varieties. 82 Andreas Gathmann

16. Local theory of regular singular points and applications

Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling

Nonlinear Control Systems

Review: control, feedback, etc. Today s topic: state-space models of systems; linearization

Optimal Control. Lecture 18. Hamilton-Jacobi-Bellman Equation, Cont. John T. Wen. March 29, Ref: Bryson & Ho Chapter 4.

DYNAMICAL SYSTEMS

Kalman Decomposition B 2. z = T 1 x, where C = ( C. z + u (7) T 1, and. where B = T, and

Model Reduction via Projection onto Nonlinear Manifolds, with Applications to Analog Circuits and Biochemical Systems

5.2.2 Planar Andronov-Hopf bifurcation

We are IntechOpen, the world s leading publisher of Open Access books Built by scientists, for scientists. International authors and editors

Chap 3. Linear Algebra

Bisimilar Finite Abstractions of Interconnected Systems

Lecture 9 Nonlinear Control Design

Disturbance Decoupling Problem

MA5510 Ordinary Differential Equation, Fall, 2014

Transcription:

Balanced realization and model order reduction for nonlinear systems based on singular value analysis Kenji Fujimoto a, and Jacquelien M. A. Scherpen b a Department of Mechanical Science and Engineering Graduate School of Engineering Nagoya University Furo-cho, Chikusa-ku, Nagoya, 464-8603, Japan fujimoto@nagoya-u.jp b Faculty of Mathematics and Natural Sciences University of Groningen ITM, Nijenborgh 4, 9747 AG Groningen, The Netherlands J.M.A.Scherpen@rug.nl Keywords : balanced realization, model reduction, singular value analysis, Hankel operators, nonlinear control systems. Abstract This paper discusses balanced realization and model order reduction for both continuoustime and discrete-time general nonlinear systems based on singular value analysis of the corresponding Hankel operators. Singular value analysis clarifies the gain structure of a given nonlinear operator. Here it is proved that singular value analysis of smooth Hankel operators defined on Hilbert spaces can be characterized by simple equations in terms of their states. A balanced realization and model order reduction procedure is derived based on it, and several important properties such as stability, balanced form, Hankel norm, controllability and observability of the original system are preserved. The work improves the earlier results of [6] and then continues with new balancing and model reduction results. 1 Introduction In the theory of stable linear systems, the system Hankel operator plays an important role in a number of problems. The relation with the state-space concept of balanced realizations, where the Hankel singular values are important is well understood nowadays, [3], and is providing us a powerful tool for model reduction of linear control systems. In this paper we propose a framework for a general class of stable nonlinear systems where balanced realizations are directly related to the Hankel operator of the nonlinear system. This on its turn provides a tool for model reduction of the nonlinear system where properties as stability, the Hankel norm and the balanced form of the system are preserved. Our approach builds further upon the earlier developments in [6]. A first nonlinear extension of the linear state-space concept of balanced realizations has been introduced in [4], mainly based on studying the past input energy and the future output energy. Since then, many results on nonlinear state-space balancing, related minimality considerations, balancing near invariant manifolds, computational issues for model reduction, flow balancing, trajectory piecewise linear balancing and empirical balancing for nonlinear systems have appeared in the literature, e.g. [10, 11, 7, 8, 1, 14, 16, 1, 3, 5, 8, 30, 31, 7].

In our earlier work, the relation of the state-space notion of balancing for finite dimensional, continuous time, input affine nonlinear systems with the nonlinear Hankel operator has been considered, see e.g. [10, 6, 5]. In particular, the singular value functions of [4], which can be viewed as a nonlinear state-space extension of the Hankel singular values in the linear case, can be related to nonlinear Hankel theory, [4, 6]. However, for obtaining the latter relation, a new characterization of (Hankel singular value functions for nonlinear systems was proposed in [6], resulting in the definition of the so-called axis singular value functions. These functions have a close relationship to the gain structure of the Hankel operator, and are characterized by singular value analysis, [3], of the Hankel operator. Although the axis singular value functions are defined without a direct relationship with the state-space notion of singular value functions of [4], it was shown that they coincide at the coordinate axes when the system has a special statespace realization, hence the name axis singular value functions. In [5, 6], this special state-space realization was adopted and characterizes a nonlinear input-normal/output-diagonal realization. However, the latter realization only has a balance between the coordinate axes of the state-space, whereas the balanced realization of a linear system also balances the relationship between the input-to-state behavior and the state-to-output behavior. From a realization and numerical point of view related to singular value decomposition, the latter property for linear systems is quite important, [1]. A first objective of this paper is to generalize the main results of [6] to a larger class of nonlinear systems, as well as to provide a shorter and more elegant proof. With this, we establish a nonlinear singular value analysis directly related to the Hankel operator of finite and infinite dimensional, continuous and discrete time nonlinear systems. The proof is based upon nonlinear operators for this large class of nonlinear systems. The corresponding result of [6] is only valid for finite dimensional, continuous time systems, and the proof is, though constructive, very long. The use of general nonlinear operators offers the possibility to shorten the proof and make it more insightful. A second objective of this paper is to provide a truly balanced realization for nonlinear finite dimensional continuous and discrete time systems, that offers a tool for model order reduction along the lines of the methods for linear systems. The starting point is the input-normal/output diagonal realization that can be obtained almost immediately from the extended result mentioned above, but now restricted to finite dimensional nonlinear systems. From there, nonlinear balanced realizations are proposed. They provide a balance of the complete part of the state space that we consider, as opposed to only providing a balance among the coordinate axes as in [6]. The balancing method is applicable to discrete time systems as well. Furthermore, the method offers a tool for model reduction proposed in this paper, along similar lines as linear balanced model order reduction methods. It is shown that properties such as balanced form, axis singular value functions, stability, and the Hankel norm are preserved for the reduced order model obtained via the proposed model reduction procedures. As mentioned above, we also propose balancing and order reduction methods for discrete time nonlinear systems in this paper. So far, within our nonlinear balancing framework started in [4] for continuous time input affine systems, only characterizations and computations of the controllability and observability functions of discrete-time nonlinear systems have been reported in [17]. Typical nonlinear discrete-time systems are not input-affine and the earlier results of [6] for continuous-time input-affine systems are not directly applicable to discrete-time systems. Since the proposed approach of our current paper builds on nonlinear operators in stead of state space realizations, our new approach is also valid for discrete time systems. Our results basically built further on our earlier work in [4], and [6]. This means that our approach is valid in a neighborhood of an equilibrium point, and depending on the system, the neighborhood can be large or small. Other methods, such as originally presented in [8], and further developed in [30, 31, 7] for continuous time systems, and in [9] for discrete time systems, are based on the flows of a system. These methods consider linearization around trajectories, and use sliding time windows for the calculation of the reachability and observability Gramians. Then a return to the original nonlinear system is only possible for a limited class of systems. However, more generally, reachability and controllability Gramians in the sliding time window setting can be calculated approximately for the whole state space, thus yielding the basis for a

balancing procedure of a large part of the state space. Nevertheless, relations with minimality and the Hankel operator are less clear than in our approach, among others due to the approximation step in the flow balancing procedures. See [8, 30, 31, 7] for the details. Furthermore, in [14] an approach based on the balancing method of [4] with polynomial approximations is treated by applying a balancing procedure to the different degrees of the polynomials separately. Also here relations with minimality and the Hankel operator are less clear, with one of the reasons the approximation step in the procedure. Model order reduction based on balancing is a method based on singular value decompositions. However, there is also quite a bit of research effort in model order reduction methods based on Krylov methods and moment matching because of their computational advantages. See [1] for an overview for linear systems. Recently, a first extension of moment matching to the nonlinear case was obtained by []. Combinations with and relations to balancing are not yet developed. The outline of this paper is as follows. Section treats preliminaries and the problem setting. The linear systems case and the results of [6] are reviewed, and a nonlinear operator setting for Hankel analysis is introduced. Section 3 provides a singular value analysis of the Hankel operator, including an extension of a main result from [6] to general, finite/infinite dimensional, continuous/discrete-time systems. It also provides a new elegant proof for the finite dimensional continuous time result of [6]. This result is then used to determine balanced realizations for finite dimensional continuous and discrete time nonlinear systems in Section 4. The developments of Section 4 are then used in Section 5 as a tool for model reduction based on the balanced realizations for both continuous and discrete time systems. It is shown that for continuous-time systems balanced truncation is a suitable method for preserving certain balanced realization properties. However, for discrete time systems, balanced truncation does not preserve these balancing properties, hence other order reduction strategies are considered. It is shown that order reduction based on singular perturbations analysis of the balanced realization does preserve the desired balanced realization properties. Finally, in Section 6 we end with some conclusions. Notation: The mathematical notation used throughout is fairly standard. If x R n, the norm is given as x = (x T x 1/. If x L [a, b] the norm is given as x = ( b a x(t dt 1/. Similarly, for discrete time signals if x l [a, b] then the norm is given as x = (Σ b k=a x(k 1/. Note that the type of norm is given by the space of the signal. The symbols R and Z denote the set of real numbers and the set of integers, respectively. Further, the half subsets of them are defined by R + := [0, R, R := (, 0] R, Z + := {0, 1,,...} Z and Z := {0, 1,,...} Z, respectively. A condition about 0 means that this conditions holds for a neighborhood of 0. Finally, x(± is an abbreviation for lim t ± x(t. Throughout this paper by smooth we generally mean C, unless stated otherwise. Preliminaries and problem setting This section refers to the preliminary results on balanced realization for both linear and nonlinear systems, and explains the problem setting for singular value analysis of nonlinear Hankel operators, which is the basic framework for balancing and model reduction for nonlinear control systems..1 Linear systems as a paradigm Here, we briefly review linear balancing theory, see e.g. [3]. The presentation is such that the line of thinking in the nonlinear case is clarified. Consider a causal linear input-output system Σ : L m (R + L r (R + with a state-space realization u y = Σ(u : { ẋ = Ax + Bu y = Cx (1 where x(0 = 0. The corresponding Hankel operator is given by the composition of the observability and controllability operators H = O C, where the observability and controllability operators, 3

O : R n L r (R + and C : L m (R + R n, respectively, are given by x 0 y = O(x 0 := Ce At x 0 u x 0 = C(u := 0 e Aτ Bu(τ dτ. The Hankel, controllability and observability operators are closely related to the observability and controllability Gramians, i.e., Q = O O and P = C C. Furthermore, from e.g. Theorem 8.1 in [3], we have the following relation. Theorem 1 [3] The operator H H and the matrix QP have the same nonzero eigenvalues. The square roots of the eigenvalues of QP are called the Hankel singular values of the system (1 and are denoted by σ i s where σ 1 σ... σ n. They are the singular values of the Hankel operator and the largest singular value equals the Hankel norm Σ H of the system Σ Σ H := H(u L sup = σ 1. ( u L (R + u L u 0 Further, using a similarity transformation (linear coordinate transformation, we can diagonalize both P and Q so that P = Q = diag(σ 1, σ,..., σ n. (3 This state-space realization is called a balanced realization. The system is balanced in two senses: (i P and Q are in a diagonal form, and (ii P = Q which means that the relationship between the input-to-state behavior and the stateto-output behavior is balanced. The property (i plays an central role in model reduction, and (ii is important, since it corresponds to a singular value decomposition that has certain numerical properties, and since it is important for realization algorithms, e.g., Chapter 4, and Chapter 7, Section 7.3 and 7.4 of [1].. Hankel operators for nonlinear systems In this paper, we consider a Hankel operator H : U Y for a nonlinear system defined on Hilbert spaces U and Y. Here, as in the linear case, we suppose that H can be decomposed as H = O C (4 with the controllability operator C : U X and the observability operator O : X Y where C is surjective and X is also a Hilbert space. In the next examples we study H for particular dynamical systems. See [6] for the details. Example 1 Consider an L -stable finite dimensional continuous-time nonlinear system { ẋ = f(x, u, t. (5 y = h(x, u, t The corresponding controllability operator C : L m (R + R n and observability operator O : R n L p (R + are defined by x 0 = C(u : y = O(x 0 : { ẋ = f(x, u, t x( = 0 x 0 = x(0 { ẋ(t = f(x, 0, t x(0 = x 0 y = h(x, 0, t (6. (7 The Hankel operator is given by the composition (4 with U = L m (R +, X = R n and Y = L p (R +. 4

Example Consider an l -stable finite dimensional discrete-time nonlinear system { x(t + 1 = f(x(t, u(t, t. y(t = h(x(t, u(t, t Here we suppose that the x(t + 1 = f(x(t, u(t is invertible with respect to x(t. The corresponding controllability operator C : l m (Z + R n and observability operator O : R n l p (Z + are defined by x 0 = C(u : y = O(x 0 : { x(t 1 = f(x(t, u(t, t x( = 0 x 0 = x(0 { x(t + 1 = f(x(t, 0, t x(0 = x 0. y(t = h(x(t, 0, t The Hankel operator is given by the composition (4 with U = L m (Z +, X = R n and Y = L p (Z +. At this moment, we do not restrict ourselves to one of the above classes of systems, even though the results in [4, 6] are limited to finite dimensional continuous-time time-invariant systems. Here, we study a much wider class of nonlinear systems including time-varying systems, input-non-affine systems, and discrete-time systems. The controllability and observability functions L c : X R + and L o : X R + with respect to the Hankel operator H given in Equation (4 are defined by L c (x 0 := inf C(u=x 0 u U If the pseudo-inverse C : X U of C : U X defined by 1 u (8 L o (x 0 := 1 O(x0. (9 C (x 0 := arg inf u (10 C(u=x 0 u U exists, then L c can be written as L c (x 0 = 1 C (x 0. (11 In the linear case, we have the following relationship with the controllability and observability Gramians P and Q For continuous-time input-affine nonlinear system of the form L c (x 0 = 1 x0t P 1 x 0 (1 L o (x 0 = 1 x0t Q x 0. (13 { ẋ = f(x + g(xu Σ : y = h(x, it is proved that the controllability and observability functions L c (x and L o (x are characterized by the solutions of a Hamilton-Jacobi equation and a Lyapunov equation [4]. L c x f(x + 1 L c T x g(xg(xt c = 0 x (14 L o x f(x + 1 h(xt h(x = 0 (15 5

where 0 is an asymptotically stable equilibrium of ẋ = f gg T ( L c (x/ x T in a neighborhood of the origin. Furthermore, if the system is linear as in Equation (1, and strict positivity of the solutions is assumed, then these partial differential equations reduce to the Lyapunov equations AP + P A T + B B T = 0 QA + A T Q + C T C = 0 for the Gramians P and Q as given in Equations (1 and (13..3 Singular value analysis of Hankel operators For deriving a balanced realization of a given nonlinear system we study the gain structure of the related Hankel operator H, i.e., we examine H(u σ max (c := sup u =c u (16 H(u v max (c := arg sup u =c u. (17 Here we assume the existence of v max U. We add the constraint u = c > 0 because we are interested in the maximizing input for each input magnitude c. Here we suppose that the Hankel operator H is (Fréchet differentiable 1. Since v max defined in Equation (17 is a critical point of ( H(u / u, u = v max needs to satisfy ( H(u d (du = 0 subj. to u = c. (18 u Doing the derivation of the first equation in Equation (18, we obtain ( H(u 0 = d (du u u d( H(u (du H(u d( u (du = u = ( u / H(u H(u, dh(u(du ( H(u / u u, du u = ( u / H(u (dh(u H(u, du ( H(u / u u, du u = (dh(u H(u ( H(u / u u, du. (19 u H(u On the other hand, differentiating the constraint in Equation (18 reduces to Hence we can rewrite the problem (18 into u, du = 0. (dh(u H(u ( H(u / u u, du = 0, du s.t. u, du = 0. Finally, we obtain an alternative formulation of (18 as follows. (dh(u H(u = λ u, λ R. (0 1 Here the operator d( denotes the Fréchet derivative. Fréchet derivative df of a given function f : X Y with Banach spaces X and Y satisfies and df(x(ξ is linear in ξ. f(x + ξ f(x = df(x(ξ + o( ξ X 6

Equation (0 characterizes all critical inputs u as well as the maximizing input v max. Note that this equation does not contain the parameter c anymore. This means that the solutions to this equation will be parameterized by the parameter c implicitly. Essentially, this fact implies that the solution set are curves in the input signal space which characterize the coordinate axes of the balanced coordinates. Then, consequently, we can obtain the nonlinear balanced realization. In order to characterize the balanced realization, we are interested in characterizing the states (at t = 0 achieving the critical points of H(u / u. Hence it is natural to restrict our problem (0 to a subset ImC of the input signal space U since its elements have one-to-one correspondence to those of the state space X. Therefore we will solve Equation (0 with u Im C (1 in what follows. Let us define the solutions for u in the above equations (0 and (1 by v. We call investigation of the solution v and λ for the above equations singular value analysis of H. Singular value analysis proposed here was called differential eigenstructure of Hankel operators in the authors former paper [6]. It should be noted that the singular vector v is an eigenvector of the operator (dh(u H(u. The vector v is a singular vector and the corresponding scalar σ defined by σ(v := H(v ( v is called a singular value of H. See [3] for the details of singular value analysis of nonlinear operators. Furthermore, it also follows from Equation (19 that the critical points of H(u / u without the constraint u = c ( H(u d (du = 0 (3 u are characterized by (dh(u H(u = σ u (4 with the singular value σ as defined in Equation (. Namely, λ in Equation (0 coincides with the square of the singular value σ at the critical points of H(u / u without constraint..4 Nonlinear input-normal realization This section briefly reviews the authors preliminary results on input-normal/output-diagonal realizations for time-invariant input-affine nonlinear systems reported in [6]. Consider a smooth input-affine nonlinear system Σ { ẋ = f(x + g(xu u y = Σ(u : y = h(x (5 with x(t R n, u(t R m and y(t R r, with x = 0 and equilibrium point for u = 0. We introduce the following technical assumptions in order to state the main results of [6]. Assumption B1 Suppose that the system Σ in Equation (5 is asymptotically stable about the origin, that there exist a neighborhood of the origin where the operators O, C and C exist and are smooth. Assumption B Suppose that the Hankel singular values of the Jacobian linearization of the system Σ around x = 0 are nonzero and distinct. The nonlinear state-space developments of [4] give an input-normal/output-diagonal form of system (5 as follows. Theorem [4] Consider the operator Σ with the asymptotically stable state-space realization (5. Suppose that Assumptions B1 and B hold. Then there exists a neighborhood W of the 7

origin and a smooth coordinate transformation x = Φ(z on W converting Σ into an inputnormal/output-diagonal form, where L c (Φ(z = 1 zt z (6 L o (Φ(z = 1 n zi τ i (z (7 with τ 1 (z... τ n (z being the so called smooth singular value functions on W. The above input-normal/output-diagonal realization is important for what follows. However, this result is incomplete in the sense that the properties (i and (ii explained below the equation (3 are not exactly fulfilled as explained in Section 4. Indeed this realization and the functions τ i s are not unique [10] and, consequently, the corresponding model reduction procedure gives different reduced models according to the choices of different sets of singular value functions. In [6] this issue is tackled by considering the relation with the Hankel operator, and a more precise input-output characterization for the input-normal/output-diagonal realization is given. The Hankel operator for the system (5 is given as in Example 1 where we restrict the plant system to be input-affine. Instead of considering the eigenstructure of H H as in the linear case given in Theorem 1, the solution pair λ R and v L (R + of the singular value analysis of H characterized by Equations (0 and (1 is considered. Using the latter singular value analysis we have the following result. Theorem 3 [6] Consider the Hankel operator H in Equation (4. Suppose that Assumptions B1 and B hold. Then there exists a neighborhood S 0 R of 0, n smooth functions σ i : S 0 R + s, i {1,,..., n} such that i=1 min{σ i (s, σ i ( s} max{σ i+1 (s, σ i+1 ( s} holds for all s S 0 and all i {1,,..., n 1} and such that there exist n distinct smooth curves ξ i : S 0 X satisfying ξ i (0 = 0 and with In particular, if S 0 = R, then L c (ξ i (s = s, L o(ξ i (s = σ i (s s L o x (ξ i(s = λ i (s L c x (ξ i(s λ i (s := σ i (s + s sup u U u 0 H(u u dσ i (s. ds = sup s S 0 σ 1 (s. Here the parameter c in Equation (18 is given by c = s with scalar parameter s parameterizing the solutions in the theorem. Furthermore, based on the above theorem, a more precise version of the input-normal/output-diagonal realization was derived. Theorem 4 [6] (Theorem 8 Consider the operator Σ with the state-space realization (5. Suppose that Assumptions B1 and B hold. Then there exist a neighborhood W of 0 and a coordinate transformation x = Φ(z on W converting the system an input-normal form (6 and (7 satisfying the following properties. z i = 0 L c(φ(z z i = 0 L o(φ(z z i = 0 (8 8

holds for all i {1,,..., n} on W. Furthermore τ i (0,..., 0, z i }{{} i-th τ i z (0,..., 0, z i }{{} i-th, 0,..., 0 = σ i (z i, 0,..., 0 = (0,..., 0, dσ i(z i, 0,..., 0 dz }{{ i } i-th holds for all i {1,,..., n}. In particular, if W = R n, then Σ H = sup τ 1 (z 1, 0,..., 0. z 1 R By this theorem, we can obtain an input-normal/output-diagonal realization which has a close relationship to the Hankel operator. In fact, this theorem gives the nonlinear version of the property (i explained below Equation (3. However, the nonlinear version of characterization (ii was not obtained so far. This is one of the problems considered in the remainder of the present paper. It is also noted that Theorems 3 and 4 are only for continuous-time input-affine nonlinear systems and their proofs given in [6] are quite long. In the following section, we derive the result for a wider class of (finite dimensional nonlinear systems, with a much simpler proof based on the analysis of Hankel operator H. 3 Singular value analysis and observability and controllability functions The objective of this section is to relate the singular value analysis of the Hankel operator H of Equations (0 and (1 for a general class of nonlinear systems to the controllability and observability operator independent of the state space representation. In this section we extend the result of Lemma 4 of [6] to the general class of finite or infinite, continuous or discrete time nonlinear systems. In the same time, the extension provides a new, briefer, and more elegant proof than in [6] for the existing result for finite dimensional, continuous time, input affine systems given by Theorem 4 in this paper. Let us consider a system with an input signal space U, an output signal space Y and a statespace X with the controllability and observability operators C : U X and O : X Y. The Hankel operator H is defined by Equation (4. Assumption A1 the operators C : U X, O : X Y and C : X U exist and are differentiable. Under this assumption, we can obtain an alternative characterization of singular value analysis of the Hankel operator on the signal space X, i.e., different from Equation (0. Theorem 5 Suppose that Assumption A1 holds. Assume moreover that there exist λ R and ξ X satisfying dl o (ξ = λ dl c (ξ. (9 Then u = v U defined by v := C (ξ (30 satisfies Equations (0 and (1, i.e., the equations for singular value analysis of H Proof. As preparation for the proof of the theorem, we need to clarify some properties of the signal space ImC given in Equation (1. By Assumption A1, both C and C exist and are differentiable. Hence the constraint (1 can be characterized by singular value analysis of C C since C C is a projector onto ImC. That is, any element in ImC is characterized by C C(u arg sup u U u (31 9

with the maximum singular value 1, since C C(u u C C(u u = 1 u Im C < 1 otherwise hold for the definition of C in Equation (10. Therefore, any elements v ImC satisfies the critical points condition without constraint as in Equation (4 with the (maximum singular value σ = C C(v / v = 1, i.e., which reduces to (d(c C(v (C C(v = 1 v, (dc(v (dc (C(v C C(v = v, (3 where v is a singular vector. Now we can prove the theorem using Equation (3. Suppose that there exist λ R and ξ X satisfying Equation (9 and define the corresponding input v U by Equation (30. Then v is an element of ImC by its definition so Equation (3 holds with the signal v thus defined. Substituting L o and L c in Equations (9 and (11 for (9 yields since Due to the definition of v, (do(ξ (O(ξ = λ (dc (ξ (C (ξ, (33 dl c (x(dx = C (x, dc (x(dx = (dc (x (C (x, dx, (34 dl o (x(dx = O(x, do(x(dx = (do(x (O(x, dx. (35 holds. Substitute this for Equation (33 then we obtain ξ = C(v (36 (dc (C(v C C(v = 1 λ (do(c(v O C(v (37 Then, by further substituting this equation for Equation (3, we have ( 1 (dc(v λ (do(c(v O C(v = v. Since (dc(v is a linear operator, this reduces to (dc(v (do(c(v O C(v = λ v. (38 On the other hand, substituting H in Equation (4 for Equation (0 yields (dc(v (do(c(v O C(v = λ v (39 which coincides with Equation (38. Hence the input v defined by Equation (30 satisfies the equation for singular value analysis (0. Also Equation (1 trivially follows from the definition of v in Equation (30. This proves the theorem. This theorem gives us a sufficient condition for singular value analysis of H characterized in Equations (0 and (1 for a nonlinear system of which the state-space is not specified yet. Condition (9 is easier to check than the Equations (0 and (1, if the dimension of the intermediate state-space X is smaller than that of the input signal space U, e.g., U = L and X = R n. 10

Note that the corresponding singular value σ defined in Equation ( is given by σ(c (ξ = H(C (ξ C = O C C (ξ (ξ C (ξ (1/ O(ξ = (1/ C (ξ = L o (ξ L c (ξ. = O(ξ C (ξ In particular, if we can characterize all ξ i s of Equation (9 and let σ i s denote the corresponding singular values, then clearly we can obtain the Hankel norm, which is the gain of the Hankel operator, in the following way. sup u U u 0 H(u u = max i sup σ i (C (ξ i. Example 3 Suppose that our plant system is the linear dynamical system given in Section.1. Then Equation (9 yields ξ T Q = λ ξ T P 1 with the controllability and observability Gramians P and Q, which is equivalent to ξ i 0 P Q ξ = λ ξ. That is, ξ is the eigenvector of P Q and all eigenvectors of P Q form the basis for the balanced realization, i.e., after the balancing transformation the eigenvectors are transformed into the new coordinate axes of the system.furthermore, λ = σ, where the σ s are the Hankel singular values. Example 4 Suppose that our plant system is the dynamical system given in Example 1 or. Then the solution of singular value analysis of the corresponding Hankel operator can be characterized by an algebraic equation L o x (ξ = λ L c (ξ. (40 x In comparison to the linear case mentioned above in Example 3, the set of ξ s plays the role of the eigenvectors, and thus they can be viewed as the axes of the balanced coordinates. Note that we do not require any state-space realization of the operators here. Hence, Theorem 5 is applicable to very general nonlinear systems including both continuous and discrete time, finite and infinite dimensional, and input-affine and input-non-affine dynamical systems. 4 Balanced realization We now study balanced realizations based on the solutions of Equation (9, that is, balanced realizations whose coordinate axes coincide with the ξ s are investigated. The result on the singular value analysis of Hankel operators given in Theorem 5 holds with any nonlinear system such as continuous-time and discrete-time systems. This allows one to obtain the balanced realization of both continuous-time and discrete-time input-non-affine nonlinear systems. Here, we first extend the input-normal/output-diagonal balancing procedure given in Section to more general systems, and then we study balanced realizations based on the latter extended procedure. We now do restrict ourselves to systems with a finite dimensional state-space. 4.1 Input-normal/output-diagonal balancing In order to generalize Theorem 3 and 4 to general systems with finite dimensional state-space, we need to employ the following assumption. 11

Assumption A Suppose that X R n, that ( L c (x/ x (0 and ( L o (x/ x (0 are positive definite and that the eigenvalues of (( L c / x (0 1 (( L o / x (0 are distinct. Under Assumption A, we can prove the existence of n independent solutions ξ i s of Equation (9 (or (40. Theorem 6 Consider a nonlinear system with a Hankel operator H in Equation (4. Suppose that Assumptions A1 and A hold. Then the same statements as in Theorem 3 hold. Proof. The proof of Theorem 3 does not use a specific state-space realization of the plant and it only depends on Assumption B, which can trivially be replaced by Assumption A. This proves the theorem. This result can be used for input-normal balanced realization for both continuous-time and discrete-time nonlinear systems as follows. Theorem 7 Consider a nonlinear system with a Hankel operator H in Equation (4. Suppose that Assumptions A1 and A hold. Then the same statements as in Theorem 4 hold. Proof. The proof follows the same line as the proof of Theorem 5. Hence, if we assume that the system has a finite-dimensional state-space, then the inputnormal/output-diagonalization procedure given in Theorems 6 and 7 is applicable to the finitedimensional continuous and discrete time state-space systems as given in Examples 1 and. Note that, in contrast to the long, and less elegant proofs of the continuous time result in [6], now we do not require explicit descriptions of the dynamics of the system. 4. Balanced realization So far, we have focused on extending results from [6] to a more general class of systems. The obtained input-normal/output-diagonal representation fulfills item (i given below Equation (3. However, item (ii given below Equation (3 is not fulfilled. For a truly balanced realization, item (ii should also be fulfilled. In this section we propose a new truly balanced realization in the sense that item (ii is also fulfilled. The proposed realization clarifies the input-to-state and state-to-output behavior of nonlinear dynamical systems, i.e., the relationship given in the item (ii below Equation (3. We first obtain a new input-normal/output-diagonal form for a -dimensional system. The latter form plays a key role for obtaining a balanced form for a finite dimensional systems of order n. Lemma 1 Consider a nonlinear system with a Hankel operator H in Equation (4, fulfilling Assumptions A1 and A. Furthermore, assume that X R. Then there exist a neighborhood W X of the origin and a coordinate transformation x = Φ(z on W converting the system into the following form L c (Φ(z = 1 ( z 1 + z L o (Φ(z = 1 ( (z1 σ 1 (z 1 + (z σ (z. Proof. Since the proof is long and technical, we refer to Appendix. Using this lemma recursively and repeatedly along similar lines as the proof of Theorem 8 in [6], a new input-normal/output-diagonal realization can be obtained. In the new realization, all the coordinate axes of the state-space appear decoupled in the observability and controllability functions. 1

Theorem 8 Consider a nonlinear operator with a Hankel operator H in Equation (4. Suppose that Assumptions A1 and A hold. Then there exist a neighborhood W X of the origin and a coordinate transformation x = Φ(z on W converting the system into the following form In particular, if W = X, then Proof. See Appendix. L c (Φ(z = 1 zt z (41 L o (Φ(z = 1 Σ H = n (z i σ i (z i. (4 i=1 sup σ 1 (z 1. z 1 Φ(z 1,0,...,0 X Example 5 Let us take an example from [6]. Consider a nonlinear system { ẋ = f(x + g(xu y = h(x (43 with f(x = g(x = h(x = ( 9x1 x 5 1 x 3 1x x 1 x 4 9x x 4 1x x 1x 3 x 5 ( 18 + x 4 1 + 4x 1 x + x4 0 0 18 + x 4 1 + 4x 1 x + x4 (6x 1 x 5 1 4x3 1 x x1x4 18+x 4 1 +4x 1 x +x4 1+x 4 1 +x 1 x +x4 (3x x 4 1 x x 1 x3 x5 18+x 4 1 +4x 1 x +x4. 1+x 4 1 +x 1 x +x4 It is shown in [6] that this system is balanced in the sense of Theorems 4 and 7, that is, its controllability and observability functions L c (x and L o (x are L c (x = 1 xt x (44 L o (x = 1 ( x 1 τ 1 (x + x τ (x (45 with τ 1 (x = 4(9 + x4 1 + x 1x + x 4 1 + x 4 1 + x 1 x + x4 = 4(9 8x 4 1 16x 1x 8x 4 + o( x 7 (46 τ (x = 9 + x4 1 + x 1x + x 4 1 + x 4 1 + x 1 x + = 9 8x 4 x4 1 16x 1x 8x 4 + o( x 7. (47 They satisfy the relationship (8. Since the dimension of the system is, the coordinate transformation obtaining the input-normal form characterized in Theorem 8 is given by x = Θ 1 Ψ Θ(z in the proof of Lemma 1. In order to compute it explicitly, we employ the Taylor series approximation up to the 5-th order in a similar way to [8]. First of all, it follows from Theorems 4 and 7 that the singular value functions σ i ( s are given by σ 1 (x 1 = 9 + x 4 1 τ 1 (x 1, 0 = 1 + x 4 = 6 8 1 3 x4 1 + o(x 7 1 (48 σ (x = 9 + x 4 τ (0, x = 1 + x 4 = 3 4 3 x4 + o(x 7 (49 13

Therefore, the balanced realization with the state z in Theorem 8 should have the following controllability and observability functions L c (Φ(z = 1 zt z + o( z 6 (50 L o (Φ(z = 1 ( (z1 σ 1 (z 1 + (z σ (z + o( z 6. (51 Substituting Equations (44 (49 for Equations (50 and (51, we obtain a pair of equations for the coordinate function Φ(z = (ϕ 1 (z, ϕ (z as follows. ϕ 1 (z + ϕ (z = z 1 + z + o( z 6 ( 9 8ϕ1 (z 4 16ϕ 1 (z ϕ (z 8ϕ (z 4 (ϕ 1 (z + 4ϕ (z = (6z 1 8 3 z3 1 + (3z 4 3 z3 + o( z 6 Solve the above equations for ϕ 1 (z and ϕ (z with the Taylor series approximation up to the order 5, we obtain the following solution ( ( ϕ1 (z z1 + 4 x = Φ(z = = 3 z3 1z + 8 9 z 1z 4 ϕ (z z 4 3 z4 1z 8 9 z 1z 3 + o( z 5. The transformed system is described in the coordinate z by { ż = ˆf(z + ĝ(zu y = ĥ(z (5 with ˆf(z = ĝ(z = ĥ(z = ( 9z1 z1 5 + 46z1z 3 + 31z 1 z 4 9x 49x 4 1x 34z1z 3 z 5 + o( z 5 ( 3 + 6 z4 1 35 3 z1z 5 z4 8 z1z 3 3 3 z 1 z 3 8 z1z 3 3 3 z 1 z 3 3 + 5 6 z1 4 + 5 3 z1z + ( 18 z 1 3 z1 5 z1z 3 7 z 1 z 4 9 z 47 z1z 4 31 z1z 3 3 z 5 + o( z 5. 6 z4 + o( z 4 It is easy to verify that this system has the controllability and observability functions (50 and (51, that is, it is balanced in the sense of Theorem 8. Once we obtain the observability and controllability functions which are decoupled on the coordinate axes, it is easy to obtain the balanced realization, i.e., a realization with a balance between the input-to-state behavior and the state-to-output behavior. Theorem 9 Consider a nonlinear system with a Hankel operator H in (4. Suppose that Assumptions A1 and A hold. Then there exist a neighborhood W of the origin and a coordinate transformation x = Φ( z on W converting the system into the following form In particular, if W = X, then L c ( Φ( z = 1 L o ( Φ( z = 1 Σ H = n z i (53 σ i ( z i n z i σ i ( z i. (54 i=1 i=1 sup σ 1 ( z 1. z 1 Φ( z 1,0,...,0 X 14

Proof. First of all, let us apply the coordinate transformation of Theorem 8 to obtain the z coordinate. Next we apply another coordinate transformation z = Φ 1 Φ( z = ( ϕ 1 ( z 1, ϕ ( z,..., ϕ n ( z n 1 with z i = ϕ i (z i := z i σi (z i to this system. Then we obtain a state space realization with the controllability and observability functions as in Equations (53 and (54 with which proves the theorem. σ i ( z i := σ i ( ϕ i ( z i We call the state-space realization described in the new coordinates z in Theorem 9 a balanced realization of the given nonlinear system. In fact, the controllability and observability functions can be rewritten as L c ( Φ( z = 1 zt P ( z 1 z, L o ( Φ( z = 1 zt Q( z z P ( z = Q( z = diag( σ 1 ( z 1, σ ( z,..., σ n ( z n which are very similar to those of the linear balanced realization (3. The coordinates could now be called uncorrelated, a terminology that was previously used in e.g., [4, 8]. The functions σ i s and σ i s have the same value, and are the singular values of the Hankel operator H. Furthermore, it is easily seen that for all realizations given in Corollary 7 and Theorems 8 and 9, the singular value functions are uniquely determined, even though the coordinates themselves are not necessarily uniquely obtained. Example 6 Consider the system (5 in Example 5. Let us compute the coordinate transformation z = Φ 1 Φ( z and the singular value functions σ i ( z i s according to the proof of Theorem 9. The solutions are obtained as follows. z = Φ 1 Φ( z = ( 6 6 z 1 + 6 97 z5 1 3 3 z + 3 43 z5 σ 1 ( z 1 = 6 7 z4 1 + o( z 4 σ ( z = 3 4 7 z4 + o( z 4 + o( z 5 This transformation converts the system (5 into the following form. { z = f( z + ḡ( zu y = h( z with f( z = ḡ( z = h( z = ( 9 z1 + 7 36 z5 1 + 3 9 z3 1 z + 31 9 z 1 z 4 9 z 49 36 z4 1 z 17 9 z 1 z 3 + 7 + o( z 5 9 z5 ( 6 3 19 3 108 z4 1 35 3 7 z 1 z 5 3 9 z4 4 6 9 z3 1 z 3 6 7 z 1 z 3 8 3 9 z3 1 z + 16 3 7 z 1 z 3 3 6 + 5 6 16 z4 1 + 5 6 54 z 1 z 19 6 ( 6 3 z 1 19 3 108 z5 1 11 3 7 z3 1 z 7 3 7 z 1 z 4 3 6 z 47 6 16 z4 1 z 31 6 54 z 1 z 3 19 + o( z 5. 6 54 z5 This system has the controllability and observability functions L c ( Φ( z = 1 ( z 1 σ 1 ( z 1 + z + o( z 6 σ ( z 54 z4 (55 + o( z 4 L o ( Φ( z = 1 ( z 1 σ 1 ( z 1 + z σ ( z + o( z 6 of the form (53 and (54, that is, it is balanced in the sense of Theorem 9. 15

5 Model order reduction In this section we propose balanced truncation and singular perturbation model order reduction procedures based on the balanced realizations given in the previous section, which is applicable to continuous-time and discrete-time nonlinear systems, respectively. It is shown that the proposed procedures result in reduced order models that preserve the balanced form and stability We consider the plant systems as given in Examples 1 and. 5.1 Model order reduction for continuous-time systems Consider the smooth time-invariant version of the continuous-time nonlinear system of Example 1 Σ : { ẋ = f(x, u y = h(x, u (56 with asymptotically stable equilibrium point x = 0 for u = 0, and with the Hankel operator H as defined in Example 1. Suppose that Assumptions A1 and A hold and that we already have the coordinate transformation z = Φ(x for one of the realizations obtained in either Theorem 7, Theorem 8 or Theorem 9. Note that all of those realizations are obtained under Assumptions A1 and A. In the new coordinates z, the system can be described as follows. { Σ z ż = f : z (z, u y = h z (z, u Here the system functions f z and h z, and the controllability and observability functions L z c and L z o in the coordinate z are described by f z (z, u := Φ 1 (x x f(φ(z, u (57 x=φ(z The singular value functions σ i (z i s are in order h z (z, u := h(φ(z, u (58 L z c(z := L c (Φ(z (59 L z o(z := L o (Φ(z. (60 max ±c σ i(s > max ±c σ i+1(s (61 in a neighborhood of the origin. Now let us consider the case where max ±c σ k(s max ±c σ k+1(s (6 holds for a certain k {1,..., n}. Then the state components z 1,..., z k are more important in terms of the Hankel operator than z k+1,..., z n due to the ordering of the singular value functions σ i s, i.e., z 1,..., z k cost less control energy to be reached asymptotically, and generate more output energy than z k+1,..., z n. Divide the coordinates into two parts corresponding to the division (6 as z = (z a, z b R n (63 z a := (z 1,..., z k R k (64 z b := (z k+1,..., z n R n k (65 ( f f z (z, u = a (z, u f b. (66 (z, u 16

Next, divide the system Σ into two subsystems by balanced truncation (i.e., by setting either parts of the coordinates equal to zero accordingly as follows Σ a : Σ b : { ża = f a ( (z a, 0, u y = h z ( (z a, 0, u { żb = f b ( (0, z b, u y = h z ( (0, z b, u (67. (68 Let H a and H b denote the Hankel operators related to the divided state-space systems (67 and (68, let L z c, L z o, L a c, L a o, L b c and L b o denote the controllability and observability functions of Σ in the coordinate z, and of Σ a and Σ b, respectively. Let σ i, i = 1,..., n, σ a i, i = 1,..., k and σb i, i = 1,..., k n, denote the singular values of the original system Σ, the reduced order systems Σ a and Σ b, respectively. Then we obtain the following properties, similar to the balanced truncation results for linear systems, [, 9, 3]. Theorem 10 Consider a continuous-time nonlinear system Σ in Equation (56 with a Hankel operator H. Suppose that Assumptions A1 and A hold and obtain a balanced realization on the neighborhood W as in Theorem 7 (or Theorem 8 or Theorem 9. Then the controllability and observability functions for the reduced order systems satisfy and the singular value functions of the reduced systems satisfy L a c (z a = L z c(z a, 0 (69 L a o(z a = L z o(z a, 0 (70 L b c(z b = L z c(0, z b (71 L b o(z b = L z o(0, z b, (7 σ a i (z a i = σ i (z a i i {1,,..., k} (73 σ b i (z b i = σ i+k (z b i i {1,,..., n k}. (74 The state-space systems Σ a and Σ b are also balanced in the sense of Theorem 7 (Theorem 8 or Theorem 9, respectively. Furthermore, there exists a neighborhood of zero U W U for which the Hankel norm is preserved as sup u U W u 0 H a (u u = sup u U W u 0 H(u u. (75 Proof. First of all, for all realizations in Theorem 7, (Theorem 8 or Theorem 9, we have the following property. z i = 0 L c(φ(z z i = 0 L o(φ(z z i = 0. (76 As in Equation (15 (see [4] for the detail, the observability function L z o of the system Σ in the coordinate z is given by a solution of a Lyapunov equation L z o(z f z (z, 0 + 1 z hz (z, 0 T h z (z, 0 = 0. Substituting z = (z a, 0 for this equation, we obtain 0 = ( L z o (z a, z b z a, L z o(z a, z b ( f a ((z a, 0, u z b f b ((z a, 0, u z b =0 = Lz o(z a, 0 z a f a ((z a, 0, u + 1 hz ((z a, 0, 0 T h z ((z a, 0, 0 17 + 1 hz ((z a, 0, 0 T h z ((z a, 0, 0

because of Equation (76. Clearly, this equation coincides with the Lyapunov equation for the observability function L a o of Σ a. That is, we have proved the relation (70. The relation (7 can be obtained in the same way. Next we consider the controllability function L z c. By the definition of C and L c in Equations (6 and (8, it can be observed that L z c(z can be obtained by solving a Hamilton-Jacobi equation L z c(z f z (z, u (z + 1 z u (z T u (z = 0 (77 which is related to the optimal control problem in Equation (8 (see Equation (14 in the inputaffine case. Here u (z is the solution of u = f z (z, u u T L z c(z z T. (78 The existence and smoothness of C in Assumption A1 implies the existence of the solution u = u (z here. Substituting z = (z a, 0 for Equation (78 yields u = f a ((z a, 0, u u T L z c(z a, 0 which is equivalent to the constraint equation for the controllability function L a c. Obviously, u = u (z a, 0 is also the solution of this equation. Further, substituting z = (z a, 0 for the equation (77, we obtain ( L z 0 = c (z a, z b L z z a, c(z a, z b ( f a ((z a, 0, u (z a, 0 z b f b ((z a, 0, u z b =0 (z a + u, 0 (z a, 0 T u (z a, 0 = Lz c(z a, 0 z a f z ((z a, 0, u (z a, 0 + u (z a, 0 T u (z a, 0 which coincides with the Hamilton-Jacobi equation for the controllability function L a c (z a for Σ a. That is, we have the relation (69. The relation (71 can be obtained in the same manner. Since the realization given in Theorem 7 (Theorem 8 or Theorem 9 is characterized only by the controllability and observability functions, the systems Σ a and Σ b are also balanced. Then Equations (73-(75 follow immediately from Theorem 7 (Theorem 8 or Theorem 9. This completes the proof. Theorem 10 reveals several properties of the proposed model reduction method: This model reduction procedure derives balanced reduced order models. Singular value functions are preserved and, in particular, the gain of the related Hankel operator (which is called Hankel norm is preserved. Since the controllability and observability functions are preserved, properties related to these functions, such as stability, etc., [4, 5], of the original system are preserved. These properties are natural nonlinear generalization of the linear case result [0], relating the Hankel theory to state-space balanced realizations, and truncation. Thus, these results are far more general than the state-space balanced realization and truncation presented in [4]. Example 7 The three systems (43, (5 and (55 are all balanced in the sense of Theorems 4 and 7. Therefore, we can apply the balanced truncation procedure stated in Theorem 10 to them. As an example, it is applied to the first one (43. Note that it also works for the other systems (5 and (55 in the same way. Since the dimension of the original system (43 is, the dimension of the reduced order model should be k = 1. Then we obtain the following 1 dimensional system { ẋa = f a (x a + g a (x a u y = h a (x a z a T 18

with f a (x a = 9x a (x a 5 g a (x a = ( 18 + (x a 4 0 ( (6x a (x a 5 18+(x a 4 h a (x a = 1+(x a 4 0. The controllability, observability and the Hankel singular value functions for this system can be computed as follows. L a c (x a = 1 (xa which confirms the outcome of Theorem 10. L a o(x a = 1 (xa τ1 a (x a = 1 (xa σ1(x a a σ1(x a a = σ 1 (x a 9 + (x = a 4 1 + (x a 4 5. Model order reduction for discrete-time systems This section proposes model order reduction based in the balanced representation of Section 4 for discrete-time nonlinear systems. Consider the time-invariant version of the discrete-time nonlinear system of Example { x(t + 1 = f(x(t, u(t Σ : (79 y(t = h(x(t, u(t with asymptotically stable equilibrium point x = 0 for u = 0, and with the Hankel operator H as defined in Example. As in the example, we suppose that the x(t + 1 = f(x(t, u(t is invertible with respect to x(t, that is, there exists a function f 1 satisfying f(f 1 (x, u, u = x, x R n, u R m. As a preparation for the model order reduction for discrete-time systems, we need to characterize the observability and controllability functions L o (x and L c (x by algebraic equations which are similar to the Hamilton-Jacobi equations in the continuous-time case. We introduce a modified version of Lemma 4.1 and Theorem 4.4 in [17]. Lemma Suppose that x = 0 of the system x(t + 1 = f(x(t, 0 is asymptotically stable. Then the observability function L o (x in Equation (9 exists if and only if Ľ o (f(x, 0 Ľo(x + 1 h(x, 0T h(x, 0 = 0, Ľ o (0 = 0 (80 has a solution Ľo(x. If they exist, then L o (x = Ľo(x holds. Proof. Necessity is proved first. Suppose that the observability function L o (x exists. Then the 19

definition of the observability function (9 implies that L o (x(0 = 1 = 1 h(x(t, 0 T h(x(t, 0 t=0 h(x(t, 0 T h(x(t, 0 + 1 h(x(0, 0T h(x(0, 0 t=1 = L o (x(1 + 1 h(x(0, 0T h(x(0, 0 = L o (f(x(0, 0 + 1 h(x(0, 0T h(x(0, 0. This equation has to hold for an arbitrary initial state x(0, that is, it satisfies Equation (80 with Ľ o (x = L o (x since L o (0 = 0. This proves necessity. Next, sufficiency is proved. Suppose that Equation (80 has a smooth solution Ľo(x. Using a notation F (x := f(x, 0, Equation (80 implies that Ľ o (x = Ľo(F (x + 1 h(x, 0T h(x, 0 = Ľo(F (F (x + 1 h(x, 0T h(x, 0 + 1 h(f (x, 0T h(f (x, 0 ( = lim Ľ o (F k (x + 1 k h(f i (x, 0 T h(f i (x, 0 k i=0 = lim Ľ o (F k (x + L o (x k = L o (x. The last equation holds because the system x(t+1 = F (x(t is asymptotically stable and because Ľ o (0 = 0. This completes the proof. This result is a natural nonlinear generalization of the linear case result. In the linear case, the dynamics (5 is given by Σ : { x(t + 1 = Ax(t + Bu(t y(t = Cx(t + Du(t (81 with A, B, C and D of appropriate size. Then the observability function is quadratic, i.e., L o (x = 1 xt Q d x. Equation (80 reduces down to the following Lyapunov equation A T Q d A Q d + C T C = 0 where Q d is the observability Gramian of the linear discrete time system. A similar result for the controllability function is obtained as follows. optimal control problem minimizing a cost function Let us consider an L c (x 0 = 1 min v l (Z x( =0, x(0=x 0 v(t (8 t=0 for the dynamics of C x(t + 1 = f(x(t, u(t, u(t = v( t 1, t = 1,,. 0

Let us denote the input u achieving the minimization in Equation (8 by u(t = u (x(t + 1 which depends on x(t + 1 since this is an optimal control problem with respect to the reverse time. Then the dynamics of C : x 0 v becomes { C x(t + 1 = f(x(t, u : (x(t + 1, x(0 = x 0, t = 1,, v( t 1 = u (x(t + 1 which reduces to { C x( t + 1 = f : ( x( t, u ( x( t, v( t = u ( x( t x(0 = x 0, t = 0, 1,, with t = t and x( t = x(t. Now a modified version of Lemma 4. and Theorem 4.5 in [17] is given as follows. Lemma 3 Suppose that x = 0 of the feedback system x(t + 1 = f 1 (x(t, u (x(t is asymptotically stable. Then a controllability function L c (x in Equation (8 exists if and only if Ľ c (f 1 (x, u (x Ľc(x + 1 u (x T u (x = 0, Ľ c (0 = 0 (83 has a solution Ľc(x. If they exist then L c (x = Ľc(x holds. Proof. This lemma can be proved as a corollary of Lemma by substituting C for O. These results are also natural generalization of the continuous-time Lyapunov/Hamilton-Jacobi equations (80 and (83 that characterize the observability and controllability functions. The characterization of the discrete-time observability and controllability function in the above two lemmas are useful for model order reduction. However, performing balanced truncation as in the continuous time case will not result in reduced order systems that preserve the balanced realization properties. A way to circumvent this, is by considering singular perturbation model reduction based on the balanced representation, similar to the linear case, e.g., [13, 15]. As in the previous section, let us now suppose that Assumptions A1 and A hold and that we already have the coordinate transformation z = Φ(x on the neighborhood W for one of the realizations obtained in Theorems 8 and 9. Consider the system Σ in Equation (5 and suppose that the system is balanced in the sense of Theorems 8 or 9. The original dynamics can be described in the z coordinates as follows. Σ : { z(t + 1 = f z (z(t, u(t y(t = h z (z(t, u(t (84 Here the system functions f z and h z, and the controllability and observability functions L z c and L z o in the coordinate z are described by Equations (58 (60 and f z (z, u := Φ 1 f(φ(z, u. As in the continuous-time case, suppose that the singular value functions are in order as in Equations (61 and (6 and divide the state-space as in Equations (63 (66. Then, accordingly, we obtain two reduced order systems by a singular perturbation method Σ a : Σ b : z a (t + 1 = f a (z a (t, z b (t, u(t z b (t = f b (z a (t, z b (t, u(t y(t = h(z a (t, z b (t, u(t z a (t = f a (z a (t, z b (t, u(t z b (t + 1 = f b (z a (t, z b (t, u(t y b (t = h(z a (t, z b (t, u(t. 1