Exact model reduction by a slow fast decomposition of nonlinear mechanical systems

Similar documents
1 Lyapunov theory of stability

One Dimensional Dynamical Systems

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

AA242B: MECHANICAL VIBRATIONS

MCE693/793: Analysis and Control of Nonlinear Systems

EN Nonlinear Control and Planning in Robotics Lecture 3: Stability February 4, 2015

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

ENGI 9420 Lecture Notes 4 - Stability Analysis Page Stability Analysis for Non-linear Ordinary Differential Equations

L = 1 2 a(q) q2 V (q).

ME 680- Spring Geometrical Analysis of 1-D Dynamical Systems

Lecture 4. Chapter 4: Lyapunov Stability. Eugenio Schuster. Mechanical Engineering and Mechanics Lehigh University.

7 Pendulum. Part II: More complicated situations

Nonlinear Control Lecture 5: Stability Analysis II

Lyapunov Stability Theory

Chapter III. Stability of Linear Systems

A Model of Evolutionary Dynamics with Quasiperiodic Forcing

Periodic Sinks and Observable Chaos

8.1 Bifurcations of Equilibria

Dynamical Systems & Lyapunov Stability

Control Systems. Internal Stability - LTI systems. L. Lanari

Kinematics. Chapter Multi-Body Systems

7 Planar systems of linear ODE

Appendix A Equations of Motion in the Configuration and State Spaces

Linear Algebra. Min Yan

Nonlinear Autonomous Systems of Differential

Mechanics IV: Oscillations

Dynamics of a mass-spring-pendulum system with vastly different frequencies

Canards at Folded Nodes

Consider a particle in 1D at position x(t), subject to a force F (x), so that mẍ = F (x). Define the kinetic energy to be.

Topic # /31 Feedback Control Systems. Analysis of Nonlinear Systems Lyapunov Stability Analysis

Lecture 9: Eigenvalues and Eigenvectors in Classical Mechanics (See Section 3.12 in Boas)

Stability theory is a fundamental topic in mathematics and engineering, that include every

Suppression of the primary resonance vibrations of a forced nonlinear system using a dynamic vibration absorber

A Study of the Van der Pol Equation

Convergence Rate of Nonlinear Switched Systems

Order Reduction of Parametrically Excited Linear and Nonlinear Structural Systems

Stabilization of a 3D Rigid Pendulum

Structural Dynamics Lecture 4. Outline of Lecture 4. Multi-Degree-of-Freedom Systems. Formulation of Equations of Motions. Undamped Eigenvibrations.

In the presence of viscous damping, a more generalized form of the Lagrange s equation of motion can be written as

A plane autonomous system is a pair of simultaneous first-order differential equations,

Linearization problem. The simplest example

Advanced Vibrations. Elements of Analytical Dynamics. By: H. Ahmadian Lecture One

Multi Degrees of Freedom Systems

Stabilization and Passivity-Based Control

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

LECTURE 8: DYNAMICAL SYSTEMS 7

Robotics. Dynamics. Marc Toussaint U Stuttgart

B5.6 Nonlinear Systems

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

STABILITY. Phase portraits and local stability

Nonlinear dynamics & chaos BECS

Fundamentals of Dynamical Systems / Discrete-Time Models. Dr. Dylan McNamara people.uncw.edu/ mcnamarad

Assignments VIII and IX, PHYS 301 (Classical Mechanics) Spring 2014 Due 3/21/14 at start of class

4. Complex Oscillations

Lagrangian Coherent Structures (LCS)

Hyperbolicity singularities in rarefaction waves

EML5311 Lyapunov Stability & Robust Control Design

Chaotic motion. Phys 420/580 Lecture 10

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Modeling and Analysis of Dynamic Systems

Chaotic motion. Phys 750 Lecture 9

MATH 205C: STATIONARY PHASE LEMMA

CENTER MANIFOLD AND NORMAL FORM THEORIES

Graceful exit from inflation for minimally coupled Bianchi A scalar field models

Linear and Nonlinear Oscillators (Lecture 2)

Chapter 6 Nonlinear Systems and Phenomena. Friday, November 2, 12

Damped harmonic motion

Stabilization of a Specified Equilibrium in the Inverted Equilibrium Manifold of the 3D Pendulum

Singular Perturbation Theory for Homoclinic Orbits in a Class of Near-Integrable Hamiltonian Systems

Seminar 6: COUPLED HARMONIC OSCILLATORS

ROTATIONAL STABILITY OF A CHARGED DIELEC- TRIC RIGID BODY IN A UNIFORM MAGNETIC FIELD

ME8230 Nonlinear Dynamics

3 Mathematical modeling of the torsional dynamics of a drill string

Complex Dynamic Systems: Qualitative vs Quantitative analysis

Global Maxwellians over All Space and Their Relation to Conserved Quantites of Classical Kinetic Equations

Math 1302, Week 8: Oscillations

Available online at ScienceDirect. Procedia IUTAM 19 (2016 ) IUTAM Symposium Analytical Methods in Nonlinear Dynamics

Lecture 4: Numerical solution of ordinary differential equations

The Liapunov Method for Determining Stability (DRAFT)

An homotopy method for exact tracking of nonlinear nonminimum phase systems: the example of the spherical inverted pendulum

Control of Mobile Robots

3. Fundamentals of Lyapunov Theory

Nov : Lecture 22: Differential Operators, Harmonic Oscillators

Autonomous Helicopter Landing A Nonlinear Output Regulation Perspective

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. small angle approximation. Oscillatory solution

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

27. Topological classification of complex linear foliations

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. Oscillatory solution

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

Nonlinear Control Lecture 2:Phase Plane Analysis

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

where x 0 is arbitrary.

Trajectory-tracking control of a planar 3-RRR parallel manipulator

Implicit Functions, Curves and Surfaces

Solutions for B8b (Nonlinear Systems) Fake Past Exam (TT 10)

MCE693/793: Analysis and Control of Nonlinear Systems

The Big Picture. Discuss Examples of unpredictability. Odds, Stanisław Lem, The New Yorker (1974) Chaos, Scientific American (1986)

Modal Decomposition and the Time-Domain Response of Linear Systems 1

Differential Equations and Linear Algebra Exercises. Department of Mathematics, Heriot-Watt University, Edinburgh EH14 4AS

Chapter 14. Oscillations. Oscillations Introductory Terminology Simple Harmonic Motion:

Transcription:

Nonlinear Dyn 07 90:67 647 DOI 0.007/s07-07-3685-9 ORIGINAL PAPER Exact model reduction by a slow fast decomposition of nonlinear mechanical systems George Haller Sten Ponsioen Received: 7 November 06 / Accepted: 8 July 07 / Published online: 8 August 07 Springer Science+Business Media B.V. 07 Abstract We derive conditions under which a general nonlinear mechanical system can be exactly reduced to a lower-dimensional model that involves only the softer degrees of freedom. This slow fast decomposition SFD enslaves exponentially fast the stiffer degrees of freedom to the softer ones as all oscillations converge to the reduced model defined on a slow manifold. We obtain an expression for the domain boundary beyond which the reduced model ceases to be relevant due to a generic loss of stability of the slow manifold. We also find that near equilibria, the SFD gives a mathematical justification for two modal reduction methods used in structural dynamics: static condensation and modal derivatives. These formal reduction procedures, however, are also found to return incorrect results when the SFD conditions do not hold. We illustrate all these results on mechanical examples. Keywords Model reduction Invariant manifolds Slow fast systems Introduction While often hoped otherwise, a typical multi-degreeof-freedom mechanical system cannot necessarily be G. Haller B S. Ponsioen Institute for Mechanical Systems, ETH Zürich, Leonhardstrasse, 809 Zurich, Switzerland e-mail: georgehaller@ethz.ch reduced to a lower-dimensional model. There is often a good reason why the original model involves several degrees of freedom, all of which are essential to reproduce the dynamics at the required level of accuracy. For any multi-degree-of-freedom system, projections to various linear subspaces are, nevertheless, routinely employed for model reduction purposes see Besselink et al. 4 for a review of techniques in structural vibrations, Benner et al. 3 for a more general survey. Most often, however, the accuracy or even the fundamental validity of these procedures is a priori unknown. The main reason is that distinguished subspaces identified from linearization or other considerations are generally not invariant under the nonlinear dynamics. As a consequence, trajectories of the full system do not follow those of a projection-based model, as shown in Fig.. A model reduction principle can be justified in a strict mathematical sense if the reduced model is defined on an invariant set of the full nonlinear system, and hence model trajectories are actual trajectories of the full system. In addition, the invariant set carrying the model dynamics should be robust and attracting for the reduced model to be of relevance for typical trajectories. While numerical or perturbative approximations to such a set will at best be approximately invariant, the attractivity of the actual invariant manifold is expected to keep the impact of non-invariance small, driving trajectories toward the actual invariant set. Motivated by these considerations, we propose here two requirements for mathematically justifiable and

68 G. Haller, S. Ponsioen q n qt q 0 u k E u q q u xt qt Fig. Model reduction by projection of a full trajectory qt onto a k-dimensional subspace E, typically spanned by a few eigenvectors u,...,u k of the linearized system at the origin. The model trajectory xt starting from a point q 0 E is constrained to lie in E, but the full trajectory qt starting from q 0 will generally leave the plane E robust model reduction in a nonlinear, non-autonomous mechanical system: R There exists an attracting and persistent lowerdimensional forward-invariant manifold Mt. Along the manifold Mt, the modeled degrees of freedom with generalized coordinates y and velocities ẏ are smooth functions of the modeling degrees of freedom with generalized coordinates x and velocities ẋ. R General trajectories approaching Mt synchronize with model trajectories at rates that are faster than typical rates within Mt. By the requirement R, the construction of a smooth, lower-dimensional dynamical model should be equivalent to a reduction to a lower-dimensional invariant manifold, as illustrated in Fig. a. The dynamics on this manifold, however, is only relevant for the full system dynamics if nearby motions qt approach model trajectories on Mt, i.e., the manifold has a domain of attraction foliated by stable manifolds of individual model trajectories. In addition, we require Mt to be persistent robust under small perturbations since mechanical models have inherent parameter uncertainties and approximations, and a model reduction should be robust with respect to these. Requirement R ensures that full system trajectories not only approach the set of model trajectories in the phase space, but also synchronize with specific model trajectories. Consider, for example, a linear, twodegree-of-freedom mechanical system with an asymptotically stable fixed point at the origin. The fast stable manifold M of this fixed point cf. Fig. b is invariant, attracting and persistent, even unique see, e.g., Cabre et al. 5. Yet, the dynamics on M fails to act as a faithful reduced-order model for the typical nearequilibrium dynamics. Indeed, the flow on M predicts a fast decay rate that is unobservable along typical trajectories on their way to the fixed point. This is because general trajectories first approach the x, ẋ = 0, 0 subspace, then creep toward the origin more slowly, synchronizing with motions along this subspace, rather than with those in M. In contrast, the invariant manifold M in Fig. b satisfies both R and R and is indeed a good choice for model reduction. This is ensured by a dichotomy of timescales created by the gap in the real part of the spectrum of the eigenvalues of the fixed point. The larger this gap, the more efficient the reduced-order model in predicting typical system behavior. In a nonlinear system, one generally loses the local slow fast dichotomy of timescales that may arise near fixed points, such as the one in Fig. b. A global reduced-order model with the properties R R will, therefore, not exist unless the slow fast timescale difference created locally by the fixed point extends to a larger domain of the phase space. In more mechanical terms, a global model reduction is only feasible when the x variables stay globally stiffer i.e., faster than the y variables. Such a global slow fast partition of coordinates has been assumed in several case studies of mechanical systems, such as an undamped spring coupled to a pendulum Georgiou and Schwartz 9 and its extensions to higher or even infinitely many dimensions Georgiou and Schwartz, Georgiou and Vakakis 0. These studies tacitly assume the existence of a slow manifold without specific consideration to its stability and robustness. Due to a lack of normal hyperbolicity for the limiting slow manifold critical manifold, welldefined invariant slow manifolds do not actually exist in these mechanical models. Recent results guarantee only near-invariant surfaces under certain conditions MacKay, Kristianssen and Wullf 0. These surfaces, however, do not attract trajectories from an open

Exact model reduction by a slow fast decomposition 69 Fig. a Illustration of the geometry of requirements R and R for model reduction in a mechanical system with generalized coordinates q and associated velocities q. The reduced model depends only on a smaller group of degrees of freedom, described by the position vector x and the corresponding velocity vector ẋ. The remaining degrees of freedom are characterized by the positions y and velocities ẏ. b An attracting and persistent invariant manifold M that does not provide a faithful reduced-order model for the full system dynamics Invariant manifold satisfying R and R qt,qt = xt, xt, yt, yt x y, y q t,q t a t x Invariant manifold satisfying R but violating R b y, y x,x neighborhood of the phase space. As a result, their relevance for model reduction is a priori unclear, as they violate the requirement R. As a further case study, a forced and stiff linear oscillator coupled to a softer nonlinear oscillator was considered by Georgiou et al. 8,. As the authors observe, the existence of an attracting, two-dimensional slow manifold in these two studies follows from a globalized version of the center manifold theorem Carr 6 and from the geometric singular perturbation formulation of Fenichel 7, respectively. These approaches are similar in spirit to the work we describe here, but pertain to specific, low-dimensional, soft stiff mechanical models without targeting model reduction issues per se. Related work also includes that of Lubich, who developed a numerical scheme for mechanical systems with very stiff potential forces. In this context, all degrees of freedom are equally fast and hence no oscillatory mode can be enslaved to the rest via model reduction. An exceptional slow manifold which involves coordinates from all degrees of freedom becomes attracting only under the numerical scheme. A numerical procedure is introduced for approximating such slow manifolds in more general but still uniformly stiff mechanical systems by Ariel et al.. In a more mathematical treatment, Stumpp 6 considered general mechanical systems with stiff damping forces and showed the existence of an attracting slow manifold governing the asymptotic behavior of the system. Again, all degrees of freedom are assumed equally stiff, and hence, no modes can be eliminated via model reduction. In contrast to these specific case studies and purely stiff reduction procedures, we consider here general mechanical systems and establish conditions under which stiffer degrees of freedom can be identified and eliminated by reduction to an attracting slow manifold defined over the remaining softer degrees of freedom. We do not assume any specific force or inertia term to be large or small. Rather, we seek the broadest set of conditions under which an exact slow fast decomposition emerges and yields a reduced-order mechanical system. The SFD procedure arising form our analysis satisfies the key requirements R R discussed above. We also establish the maximal domain of SFD and give a specific upper bound on the rate at which general solutions synchronize with those of the reduced-order model. Our includes several classes of mechanical systems and justifies earlier heuristic reduction schemes under certain conditions. In particular, under the SFD conditions, the techniques of static condensation and modal derivatives, respectively, can rigorously be justified as first- and second-order local approximations to a slow manifold near an equilibrium. At the same time, we give examples of these reduction procedures fail when the SFD conditions are not met.

60 G. Haller, S. Ponsioen We illustrate these results on simple mechanical systems, but our formulas are explicit enough to be applied to higher-degree-of-freedom problems. Importantly, determining the eigenvalues and modes shapes is not a prerequisite for the application of SFD. Indeed, the stiffer modes may be fully both linearly and nonlinearly coupled to the rest of the modes. Setup. General form of the mechanical system Consider a n-degree of freedom, non-dimensionalized mechanical system of the form Mq, t q Fq, q, t = 0, where M R n n is a nonsingular mass matrix that may depend on the generalized coordinates q and the time t in a smooth fashion class C r for some r. The internal and external forces acting on the system are contained in the term F R n, which generally depends on q, t and the generalized velocities q R n.. Classic model reduction by projection to a subspace As noted in the Introduction, model reduction for system is generally motivated by an assumed coordinate change q = Ux, withamatrixu R n k and a reduced coordinate vector x R k with k < n see, e.g., Geradin and Rixen 3. Substitution into, followed by a multiplication by U T, then suggests the reduced equations of motion U T MUx, tu ẍ U T FUx, U ẋ, t = 0, 3 the projection of from the full state space R n onto a k-dimensional subspace E, parametrized by the variable x cf. Fig.. The main focus of model reduction studies is then the most expedient choice of the matrix U. It is often forgotten, however, that for Eq. 3 to hold, one must have qt = Uxt for all times, i.e., E must be an invariant plane for. This assumption is practically certain to be violated unless special symmetries are present. Even for unforced and stable structural system i.e., when is autonomous and q = 0 is asymptotically stable, the invariance of modal subspaces is violated when nonlinear terms are present. The mismatch between modal subspaces and nonlinear invariant manifolds emanating from the origin will only be small very close to the origin. In addition, various choices of U may render projected equations that do not capture typical dynamics even close to q = 0 cf. Haller and Ponsioen 5 and Sect. 5 below for examples..3 Slow softer and fast stiffer variables If system is non-autonomous, we assume that its explicit time dependence in M and F is precisely one of the following three types: periodic quasiperiodic with finitely many rationally independent frequencies 3 aperiodic over a finite time interval a, b. Inthe periodic and quasiperiodic cases, we let t T = R, whereas in the aperiodic case, we let t T =a, b. Next, we split the generalized coordinate vector q as x q =, x R s, y R f, s + f = n, y into yet unspecified slow coordinates x and fast coordinates y. This slow fast partition refers to the expected relative speed of variation of the x and y variables. In mechanical terms, we expect x to label relatively softer degrees of freedom as opposed to the relatively stiff degrees of freedoms labeled by the y coordinates. We seek conditions under which a mathematically rigorous model reduction process exists to express yt uniquely as function of xt, at least asymptotically in time, along general trajectories qt of. The x, y partition of q may be suggested by a modal analysis of the linear system or simply by the physics of a mechanical problem. Importantly, q is not assumed to be a set of linear modal coordinates, and hence our procedure does not rely on an a priori identification of a linearized spectrum near an equilibrium point. To allow for a potentially stiff dependence of the system on y, we introduce a small, non-dimensional parameter > 0 and consider M and F as smooth functions of y/ and y for >0. At this point, this represents no loss of generality, given that any smooth function of y and can also be viewed as a smooth function of y/ and for >0because y = y/.

Exact model reduction by a slow fast decomposition 6 Remark The introduction of the small variable through the identity y = y/ above reflects expected stiff behavior with respect to the y = 0 position in the y degrees of freedom. If, instead, one expects stiff behavior small displacements with respect to a more general y = y 0 = 0 position, then one should first let ỹ = y y 0 and then drop the tilde before introducing the dummy variable as above. Using the notation introduced above, we split the mass matrices and forcing terms in by letting M x, y Mq, t =, t; M x, y, t; M x, y, t;, M x, y, t; F x, ẋ, y, ẏ, t; Fq, q, t =, > 0, F x, ẋ, y, ẏ, t; where M R s s, M, M T R s f, M R f f, F R s and F R f. Again, as notated above, this notation is general enough to allow for cases in which M or F depends purely on y, or depends both on y and y/. The corresponding equations of motion are M ẍ + M ÿ F = 0, M ÿ + M ẍ F = 0. 4 Taking appropriate linear combination of these equations, and introducing the matrix M i and forces Q i via M i x, y, t; = M ii M ij M jj M ji, i, j =,, i = j, Q i x, ẋ, y, ẏ, t; = F i M ij M jj F j, i, j =,, i = j, 5 we deduce from 4 the inertially decoupled equations of motion: M x, y, t; ẍ Q x, ẋ, y, ẏ, t; = 0, M x, y, t; ÿ Q x, ẋ, y, ẏ, t; = 0. 6 Importantly, Eq. 6 is fully equivalent to for any choice of the partition q = x, y and for any choice of a scalar parameter > 0. In particular, M x, y, t; R s s and M x, y, t; R f f are nonsingular matrices for all x, y, t and for all >0. As we shall see in later examples, the partition q = x, y will need to be selected in given problems in a way that further assumptions detailed below are satisfied..4 Assumptions of the SFD and illustrating examples We now list assumptions that will be sufficient to guarantee the existence of an exact reduced-order model satisfying the requirements R R. First, using the new variable η = y/, we define the mass-normalized forcing terms P x, ẋ,η,ẏ, t; = M x,η,t; Q x, ẋ,η,ẏ, t;, P x, ẋ,η,ẏ, t; = M x,η,t; Q x, ẋ,η,ẏ, t;, which are, by our assumptions, class C r in their arguments for >0. The following assumptions concern properties of P i in their = 0 limit. A Nonsingular extension to = 0:The functions P and P are at least of class C in their arguments at = 0. In other words, assumption A requires continuous differentiability of the transformed forcing terms P i also in the limit of = 0, when the dummy variable η = y/ is held fixed, independent of. A Existence of a fast zero-acceleration set critical manifold: The algebraic equation Q x, ẋ, η, 0, t; 0 0 can be solved for η on an open, bounded domain D 0 R s R s T. Specifically, there exists a C function G 0 : D 0 R s such that Q x, ẋ, G 0 x, ẋ, t, 0, t; 0 0 7 holds for all x, ẋ, t D 0. We refer to the set M 0 t defined by η = G 0 x, ẋ, t as a critical manifold. Assumption A ensures the existence of a smooth set M 0 t of instantaneous zero-acceleration states critical manifold for the fast coordinates. Here the velocity variable ẋ R s is viewed as arbitrary, and hence unrelated to the actual time derivative of xt along a trajectory qt. As a consequence, these instantaneous zero-acceleration states are not equilibria and do not form an invariant set for system 6. Under assumption A3 below, however, M 0 t will turn out to approximate a slow invariant manifold that carries a reduced-order model satisfying the requirements R R: A3 Formal asymptotic stability of the critical manifold: With the matrices Ax, ẋ, t = ẏ P x, ẋ, G 0 x, ẋ, t, 0, t; 0, Bx, ẋ, t = η P x, ẋ, G 0 x, ẋ, t, 0, t; 0, 8

6 G. Haller, S. Ponsioen the equilibrium solution u 0 R f of the unforced, constant-coefficient linear system u + Ax, ẋ, tu + Bx, ẋ, tu = 0 9 is asymptotically stable for all fixed parameter values x, ẋ, t D 0. Here prime denotes differentiation with respect to an auxiliary time τ that is independent of t. Note that assumption A3 requires the linear unforced oscillatory system 9, posed formally for the dummy fast variable η, to be asymptotically stable. In this context, x, ẋ, t play the role of constant parameters ranging over D 0. Assumption A3 is satisfied, for instance, when A is symmetric, positive semi-definite and B is symmetric, positive definite over D 0. In that case, A represents a damping matrix and B represents a stiffness matrix for all parameter values x, ẋ, t D 0. At this point, A3 is only a formal requirement with no immediately clear mathematical meaning. This is because the critical manifold M 0 t is not invariant under Eq. 6 and hence the arguments of A and B are, in fact, time-varying, and hence do not determine the stability of 9. Example Weakly nonlinear system with external forcing Consider a typical multi-degree-of-freedom mechanical system of the form M ẍ + C ẋ + K x + S x, y = f t, M ÿ + C ẏ + K y + S x, y = f t, 0 with x R s and y R f. Here the M i are symmetric and positive definite constant mass matrices; C i are constant symmetric damping matrices; K i are constant symmetric stiffness matrices; and the functions S i x, y = O x, x y, y model nonlinear coupling terms. By definition 5, for an arbitrary scalar parameter > 0 independent of M i, C i, D i and S i, we specifically have Q x, ẋ, y, ẏ, t; = C ẋ + K x + S x, y Q x, ẋ, y, ẏ, t; = C ẏ + K y f t, + S x, y f t. Therefore, the functions P x, ẋ,η,ẏ, t; = M C ẋ + K x + S x,η f t, P x, ẋ,η,ẏ, t; = M C ẏ + K η + S x,η f t, are differentiable in at the = 0 limit, satisfying assumption A. However, we have P x, ẋ,η,ẏ, t; 0 0. Therefore, while any function G 0 x, ẋ, t satisfies A, both matrices A and B defined in 8 vanish, and hence assumption A3 never holds for system 0.For this assumption to hold, some of the system parameters must be related to the small parameter, as we shall see in the next two examples. Example Partially stiff weakly nonlinear system with very small stiff-inertia and parametric forcing Consider now the slightly modified multi-degree-offreedom mechanical system M ẍ + C ẋ + K x + S x, y = f t, M ÿ + C ẏ + K y + S x, y = f t, with a non-dimensional small parameter. All variables, matrices and functions are the same as in Example,butthey-component of this system generates very small inertial forces and also has large linear stiffness. This time, we have P x, ẋ,η,ẏ, t; = M C ẋ + K x + S x,η f t, P x, ẋ,η,ẏ, t; = M C ẏ + K η + S x,η f t, therefore assumption A is not satisfied, given that P is not differentiable at = 0. Example 3 Paradigm for targeted energy transfer: Weakly nonlinear system with small inertia in its essentially nonlinear component Consider the multidegree-of-freedom mechanical system M ẍ + C ẋ + K x + S x, y = 0, M ÿ + C ẏ + S x, y = 0, 3 with a non-dimensional small parameter. Again, all variables and matrices are the same as in Example

Exact model reduction by a slow fast decomposition 63, but the y-component of 3 generates small inertial forces and no linear stiffness forces. This system is noted as a prototype example of targeted energy transfer cf. Vakakis et al. 7 from the x degrees of freedom to the y degrees of freedom. This energy transfer mechanism suggests the lack of a reduced-order model over the x-degrees of freedom, given that the y-variables display no long-term enslavement to the x-variables. Calculating the quantities in our assumption A, we find P x, ẋ,η,ẏ, t; = M C ẋ +K x +S x,η, P x, ẋ,η,ẏ, t; = M C ẏ + S x,η, are differentiable at = 0, and hence assumption A holds. However, the equation Q x, ẋ,η,0, t; 0 = S x, 0 = 0 cannot be solved for the variable η at any point. As a consequence, even though a set of zero-acceleration states is defined by the equation S x, 0 = 0, this set is not attracting. Indeed, the matrix Bx, v, t defined in assumption A3 vanishes identically and hence the linear system 9 is not asymptotically stable. Example 4 Partially stiff weakly nonlinear system with parametric forcing Consider now the multidegree-of-freedom mechanical system M ẍ + C ẋ + K x + S x, y = f t, M ÿ + C ẏ + K y + S x, y = f t, 4 with the same quantities as in Example. The difference here is that the mass matrix of the y degrees of freedom has small norm for and the stiffness matrix is large in norm in the same equation. In addition, the nonlinear coupling term in the x-equation is assumed to have a stiff dependence on y. In this case, we have P x, ẋ,η,ẏ, t; = M C ẋ + K x + S x,η f t, P x, ẋ,η,ẏ, t; = M C ẏ + K η + S x,η f t, which satisfy assumption A. Solving the equation Q x, ẋ,η,0, t; 0 = 0forη, we find that assumption A is satisfied by the function G 0 x, ẋ, t = K f t S x, 0, x, ẋ, t D 0 = R s R s R, 5 provided that the stiffness matrix K is invertible. In that case, we obtain Ax, ẋ, t = M C, Bx, ẋ, t = M K + y S x,g 0 x, ẋ, t ε=0 = M K, and hence the homogeneous linear oscillatory system in assumption A3 becomes η + M C η + M K η = 0 or, equivalently, M η + C η + K η = 0. 6 The zero equilibrium of this system is asymptotically stable by our assumptions on M, C and K. Therefore, assumption A3 is also satisfied for system 4. 3 Main result: global existence of an exact reduced-order model To state our main result formally, we first define the following functions for all x, ẋ, t D 0 : H 0 x, ẋ, t = x G 0 x, ẋ, tẋ + ẋ G 0 x, ẋ, t P x, ẋ, G 0 x, ẋ, t, 0, t; 0 + t G 0 x, ẋ, t, G x, ẋ, t = D η P x, ẋ, G 0 x, ẋ, t, 0, t; 0 Dẏ P x, ẋ, G 0 x, ẋ, t, 0, t; 0 H 0 x, ẋ, t D η P x, ẋ, G 0 x, ẋ, t, 0, t; 0 D P x, ẋ, G 0 x, ẋ, t, 0, t; 0, H x, ẋ, t = x G x, ẋ, tv + ẋ G x, ẋ, t P x, ẋ, G 0 x, ẋ, t, 0, t; 0 + t G x, ẋ, t. 7 With these quantities, we have the following result: Theorem Under assumptions A A3 and for > 0 small enough: i The mechanical system admits an exact reduced-order model satisfying the requirements R R. ii The reduced-order model is given by ẍ P x, ẋ, G 0 x, ẋ, t, 0, t; 0 = D η P x, ẋ, G 0 x, ẋ, t, 0, t; 0 G x, ẋ, t + Dẏ P x, ẋ, G 0 x, ẋ, t, 0, t; 0 H 0 x, ẋ, t +D P x, ẋ, G 0 x, ẋ, t, 0, t; 0 + O 8

64 G. Haller, S. Ponsioen for all x, ẋ, t D 0. iii If M x,η, is smooth in at = 0, then the multiplication of 8 by M gives a form of the reduced-order model that does not require the inversion of M : M x, G 0 x, ẋ, t, t; 0 ẍ Q x,v,g 0 x, ẋ, t, 0, t; 0 = O. 9 iv Reduced-order models 8 9 describe the reduced flow on a s-dimensional invariant manifold M t along which positions and velocities in the stiff degrees of freedom are enslaved to those in the slow degrees of freedom via y = G 0 x, ẋ, t + G x, ẋ, t + O 3, ẏ = H 0 x, ẋ, t + H x, ẋ, t + O 3. 0 v The xt components of the trajectories of system synchronize with appropriate model trajectories x R t of 8 or 9 at an exponential rate. Specifically, let qt = xt, yt be a full trajectory of system such that at a time t 0,the initial position qt 0 is close enough to the slow manifold carrying the reduced-order model. Then there exists a trajectory x R t of the reduced-order model 8 or 9 such that xt xr t ẋt ẋ R t xt 0 x R t 0 ẋt C 0 ẋ R t 0 yt 0 G 0 x R t 0, ẋ R t 0, t+o e Λ t t0, ẏt 0 H 0 x R t 0, ẋ R t 0, t+o t > t 0, where Λ > 0 can be selected as any constant satisfying max Re λ j x, ẋ, t < Λ <0, j, f,x,ẋ,t D 0 with λ j x, ẋ, t, j =,..., f, denoting the eigenvalues of the associated linear system 75. The constant C > 0 generally depends on the choice of Λ but is independent of the choice of the initial conditions qt 0 and qt 0. Proof See Appendix. In Fig. 3, we illustrate the geometric relation between the reduced model flow on the slow manifold to general trajectories of the full system, as described by Theorem. x qt,qt = xt, yt, xt, yt x R t, x R t y, y Fig. 3 Reduced-order model trajectory x R t, ẋ R t as a projection from the slow manifold M t to the space of the x, ẋ variables. Other nearby trajectories converge to the slow manifold exponentially fast, and hence their projection on the x, ẋ space synchronizes exponentially with trajectories of the reduced-order model Remark If the left-hand side of reduced-order model 8 has structurally stable features cf. Guckenheimer and Holmes 4, then, for >0 small enough, these features persist smoothly under the addition of the O terms of the right-hand side, and hence an explicit computation of these terms is not necessary. For instance, if system 8 has a single attracting fixed point or periodic orbit over the compact domain D 0, then either of these features is robust without the explicit inclusion of the O and higher-order terms on its right-hand side. If, however, system 8 is conservative, then the inclusion of O terms is necessary to obtain a robust, dissipative reduced-order model. If the O terms are also conservative, then explicit evaluation of the O is required following the expansion scheme used in the proof of Theorem. Remark 3 The synchronization expressed by means that both positions and velocities predicted by reduced-order model 8 are relevant for the observed system dynamics as long the time t > t 0 is selected from the domain T. This time-domain is unbounded i.e., T = R for mechanical systems with explicit periodic and quasiperiodic time dependence. For the case of temporally aperiodic time dependence, the times allowed in are restricted to the finite interval T =a, b. O is required following the expansion scheme used in the proof of Theorem. 0 x

Exact model reduction by a slow fast decomposition 65 Example 5 Partially stiff weakly nonlinear system with parametric forcing We recall that the partially stiff system 4 in Example 4 satisfies assumptions A A3 and hence admits an exact, global reducedorder model. The form of the function G 0 from 5 is G 0 x, ẋ, t = K f t S x, 0. The mass matrix M is independent of, and hence equivalent form 9 of the reduced-order model applies and gives M ẍ + C ẋ + K x + S x, K f t S x, 0 = f t + O. The leading-order terms in expressions 0 for the slow manifold are y = G 0 x, ẋ, t + O = K f t S x, 0 + O, ẏ = H 0 x, ẋ, t + O = K f t x S x, 0ẋ + O. If f t is periodic or quasiperiodic in time, then we have the synchronization estimate for all times t > t 0. Specifically, any Λ>0 can be selected such that Λ <0 is a strict upper bound on the real part of the spectrum of the oscillatory system 6. We note that if we had assumed a non-stiff coupling of the form S x, y in Example 4, then assumptions A A4 would still have been satisfied, but the reduced model would simplify to M ẍ + C ẋ + K x + S x, 0 = f t + O, uncoupling completely from the stiff modes at leading order. Convergence estimate would remain valid in this case, too. 4 The boundary of the domain of model reduction In the examples we have discussed so far, the domain D 0 could be selected arbitrarily large. Thus, a reducedorder model exists over arbitrarily large x, ẋ, t values in these problems, as long as is kept small enough. In general, however, D 0 will have a nonempty boundary D 0 over which reduced-order model 8 9 cannot be further extended. Such non-extendibility of the reduced-order model domain arises from a break-down in the solvability of algebraic equation 7 for the critical manifold. By the implicit function theorem, this occurs along points satisfying det η P x, ẋ, G 0 x, ẋ, t, 0, t; 0 = 0, x, ẋ, t D 0. 3 In the generic case, this determinant becomes zero at points where η P has a single zero eigenvalue, i.e., rank η P x, ẋ, G 0 x, ẋ, t, 0, t; 0 = f, x, ẋ, t D 0. 4 Under further nondegeneracy conditions see., e.g., Arnold, a fold develops in the critical manifold along D 0, i.e., M 0 M + 0 ceases to be a locally unique graph over the x, ẋ, t variables. As we pass from M 0 to the newly bifurcating critical manifold branch M 0, the matrices Ax, ẋ, t and Bx, ẋ, t vary smoothly in their arguments, given that one manifold branch is smoothly connected to the other one along a fold. Under nondegeneracy condition 4, precisely one eigenvalue of the matrix Bx, ẋ, t will cross zero in the passage from M + 0 to M 0 along the critical manifold. In this case, the graph segment η = G 0 x, ẋ, t describing the bifurcating branch M 0 9 violates assumption A. As a consequence, the folded slow manifold branch M perturbing from M 0 is unstable and hence irrelevant for reduced-order modeling. In summary, unlike in the setting of the local construction of spectral submanifolds near equilibria cf. Haller and Ponsioen 5, a folding invariant manifold arising in SFD is not a technical limitation to overcome when one is in pursuit of a more global reducedorder model. Rather, a fold in the slow manifold over the plane of slow variables signals precisely the limit beyond which no reduced-order model satisfying R R exists in a given part of the phase space. Example 6 Partially stiff weakly nonlinear system with parametric forcing Consider now the multidegree-of-freedom mechanical system M ẍ + C ẋ + K x + S x, y M ÿ + C ẏ + K y + S x, y = f t, = f t, 5 with the same variables, matrices and functions used in Example 4, except that here the coupling function S also has a stiff dependence on the y variables. We then obtain

66 G. Haller, S. Ponsioen P x, ẋ,η,ẏ, t; = M C ẋ + K x + S x,η f t, P x, ẋ,η,ẏ, t; = M C ẏ + K η + S x,η f t. Thus, assumption A is satisfied again. Condition 3 in this case gives det M K + y S x,η = 0. By the non-singularity of M, this latter condition is equivalent to det K + y S x,η = 0. For instance, when S has only quadratic terms, then this last condition can always be written as det K + Π x + Φη = 0, 6 where Π and Φ are 3-tensors of appropriate dimensions. Suppose now, for simplicity, that Φ 0, the master variable x is a scalar m =, and Π R f f is nonsingular. Requirement 6 then becomes det Π K xi = 0, 7 which implies that x cannot be an eigenvalue of Π K. Consequently, condition 7 fails along the domain boundary { D 0 = x, ẋ, t : j: x = λ j x, ẋ, t : j : } x = λ j Π K, with λ j Π K denoting the jth real eigenvalue of the matrix Π K. To illustrate the geometry of the critical manifold in a simple case, we let s = f = and select the parameters, the coupling and the forcing terms as K = 4, S x,η= x + 4η, f t = sin t, so that the equation P x, ẋ,η,0, t; 0 = 0 takes the form 4η + x + 4η sin t = 0. 8 This equation is solved by η = x = t = 0 and hence the set D 0 is nonempty. The boundary D 0, defined by conditions 3 4, satisfies det η P x, ẋ,η,0, t; 0 = η P x, ẋ,η,0, t; 0 = 4 + 8η = 0 η =, η det η P x, ẋ,, 0, t; 0 = 8 = 0, where the second condition here is the classic nondegeneracy condition for fold bifurcations in the onedimensional case cf. Arnold. Substitution of η = into Eq. 8 gives an explicit definition for D 0 in the x, ẋ, t space as D 0 = { x, ẋ, t : x = + sin t }. 9 A direct solution of Eq. 8 through the quadratic formula confirms that the zero set η = G ± 0 x, ẋ, t = ± x sin t, { } x, ẋ, t D 0 = x, ẋ, t : x < + sin t indeed ceases to be a graph and develops a fold singularity along D 0. The stability of the two branches of G ± 0 x, ẋ, t can be determined by calculating 9 along both branches: A ± x, ẋ, t = M C, B ± x, ẋ, t = M 4 + 8G ± 0 x, ẋ, t = 4 x sin t. Therefore, the critical manifold M + 0 = { x, ẋ, t D 0 : η = G + } 0 x, ẋ, t satisfies assumptions A A3 but develops a fold over D 0 along the boundary curve D 0 defined in 9. The additional branch M 0 = { x, ẋ, t D 0 : η = G } 0 x, ẋ, t emanating from the domain boundary D 0 is unstable, as its associated constant-coefficient linear system cf. assumption A, given by M u + C u 4 x sin tu = 0, is unstable. We show the stable and unstable critical manifolds, as well as the domain boundary D 0,in Fig. 4. 5 Approximate SFD near equilibria: Static condensation and modal derivatives Here we show that at least two formal reduction procedures used in structural dynamics, modal condensation and the method of modal derivatives, can be mathematically justified when the conditions A A3 of the SFD are satisfied. In this case, these two procedures turn out to provide local first- and second-order

Exact model reduction by a slow fast decomposition 67 Fig. 4 The stable critical manifold branch M + 0 and the unstable branch M 0 for nonlinear mechanical system 5 with s = f =. Also shown is the domain boundary D 0 along which the fold in the critical manifold M 0 develops approximations, respectively, to a slow manifold M emanating from an equilibrium point of the unforced mechanical system. To show this, we also assume the following: A4 Independence of critical manifold of the slow velocities: The relation ẋ Q x, ẋ,η,0, t; 0 0, 30 η t holds, i.e, the function P has not explicit dependence on the slow velocities ẋ for = 0 and ẏ = 0. We further assume that the domain D 0, over which the graph η = G 0 x, ẋ, t is defined, contains the line x = 0ofthex, ẋ, t parameter space, i.e, A5 Critical manifold contains an unforced fixed point: We assume {x, ẋ, t : x = 0, ẋ = 0} D 0. 3 This condition is satisfied, for instance, when is a weakly nonlinear system whose unforced part admits a fixed point at q = x, y = 0. The implications of assumptions A4 A5 for the geometry of the critical manifold are illustrated in Fig. 5. Static condensation Geradin and Rixen 3 is a linear reduction procedure applied to a q = x, y partition of the degrees of freedom in system near an equilibrium point. In this reduction method, the inertial terms and velocities are simply ignored in the linearized Fig. 5 The geometry of the critical manifold M 0 t under assumptions A3 A4 at an arbitrary time t equation for the y degrees of freedom. The resulting linear algebraic equation is solved for y, and the result is substituted for y in the x equations, yielding a single second-order differential equation in the x variables. The method of modal derivatives Idelsohn and Cardona 6, Rutzmoser et al. 4, WuandTiso 8 considers a similar q = x, y partition of coordinates near an unforced equilibrium and seeks a quadratic invariant manifold tangent to an eigenspace of the linearized system. The main assumption is that along this quadratic manifold, the y coordinates can be written as purely quadratic functions of the x coordinates, with the coef-

68 G. Haller, S. Ponsioen ficients of these quadratic forms collected in an appropriate modal derivative tensor. The above two reduction methods can be justified in our present setting as follows: Proposition Under assumptions A A5: i The expressions derived for the slow manifold M t in 0 satisfy G 0 x, ẋ, t = Γt + Φtx + Θtx x + O x 3, 3 where the function Γt is the solution of the equation P 0,Γt, 0, t; 0 = 0, and the two-tensor Φt and the three-tensor Θt satisfy Φt = x=0,η=γt,ẏ=0,=0 η P x P, Θt = η P xx P + xη P + ηη P x=0,η=γt,ẏ=0,=0 Φt Φt. 33 ii Assume that P x, ẋ,η,0, t; 0 has no explicit time dependence and x, y are modal coordinates for the linearized system at x, y = 0, i.e., t P x, ẋ,η,0, t; 0 0, x P 0, 0, 0, 0, t; 0 = 0, η P 0, 0, 0, 0, t; 0 = 0. We then obtain Γ = 0, Φ = 0, Θ = η P 0, 0, 0, 0, t; 0 xx P 0, 0, 0, 0, t; 0. 34 iii Under the conditions of statement ii, a linear-inx and zeroth-order-in- approximation to M t yields the modal-condensation-based reduced model ẍ P x, ẋ, 0, 0, t; 0 + O, x 3 = 0 35 for the dynamics on M t iv Under the conditions of statement ii, a quadraticin-x and zeroth-order-in- approximation to M t yields the modal-derivative-based reduced-order model ẍ P x, ẋ, Θtx x, 0, t; 0 + O, x 4 = 0 36 for the dynamics on M t, with Θt generally referred to as the modal derivative tensor. Proof See Appendix. Remark 4 Combining statement iii of Theorem with Proposition gives that if M x,η,is smooth in at = 0, then the static-condensation-based model 35 is equivalent to M x, 0; 0 ẍ Q x, ẋ, 0, 0, t; 0 + O, x 3 = 0, 37 and the modal-derivatives-based reduced model 36is equivalent to M x, Θx x; 0 ẍ Q x, ẋ, Θx x, 0, t; 0 + O, x 4 = 0. 38 Remark 5 The unevaluated higher-order O x 3 and O x 4 terms in Eqs. 37 and 38 generally do not remain uniformly small over the full model reduction domain D 0. Rather, one can only use the leading-order model terms in these equations reliably as long as the slow coordinates are rescaled as x = 3 ξ and x = 4 ξ, respectively. In that case, 37 and 38 can be re-written as M ξ,0; 0 ξ Q ξ, ξ,0, 0, t; 0 + O = 0, 39 M ξ,θξ ξ; 0 ξ Q ξ, ξ,θξ ξ,0, t; 0 + O = 0, 40 respectively. One can then arguably focus on the - independent leading-order terms for > 0 small enough. The static-condensation- and model-derivativebased reductions are, therefore, justified in order O 3 and O 4 neighborhoods of the x = 0 equilibrium, respectively, provided that the assumptions of Proposition are satisfied. Example 7 Localized reduced-order model for a stiff, weakly nonlinear system with parametric forcing We now reconsider the multi-degree-of-freedom mechanical system 4 and assume that f t 0, 4 i.e., that the external forcing on the stiff degrees of freedom vanishes. Using the results from Example 6, we have P x, ẋ,η,ẏ, t; = M C ẋ + K x + S x,η f t, P x, ẋ,η,ẏ; = M C ẏ + K η + S x,η. As already discussed in Example 6, conditions A A3 are satisfied and hence Theorem guarantees a

Exact model reduction by a slow fast decomposition 69 slow manifold and determines its reduced dynamics. Assumption 30 is clearly satisfied, as P does not depend on ẋ. Assumption 3 also holds, as one sees from the expression for G 0 in. Since P has no explicit time dependence, the static condensation and modal derivative formulas in 83 apply and take the specific form Γ 0, Φ 0, Θ = M K M xx S 0, 0 = K xx S 0, 0. 4 Therefore, in a neighborhood of the origin, the reducedorder formulation 39 applies and statement iii of Proposition justifies the static-condensation-based reduced model M ξ + C ξ + K ξ + S x, 0 f t + O = 0 as a leading-order reduced model for the dynamics on M t in an order O 3 neighborhood of the unforced equilibrium x = 0. Similarly, statement iv of Proposition justifies the modal-derivative-based reduced-order model M ξ +C ξ +K ξ +S x, K xx S 0, 0x x f t + O = 0 43 in an order O 4 neighborhood of the unforced equilibrium x = 0. The above example illustrates how Proposition puts static condensation and modal derivatives in a rigorous context under appropriate assumptions. We now also illustrate, however, that these two intuitive reduction methods give incorrect results when the assumptions of Proposition are not satisfied. ẍ + c + μ ac 4c k + k c c + c c 6c c + 4c 3 x ẋ D k ac c 3c c + c + 4k k Example 8 Failure of static modal condensation and model-derivative-based reduction Consider a twodegree-of-freedom nonlinear, coupled oscillator system with amplitude-dependent damping in the first mode, given by the equations ẍ + c +μ x ẋ +k x + axy + bx 3 = 0, x R, ÿ + c ẏ + k y + cx = 0, y R. 44 Note that the linearized system at the x, y = 0, 0 equilibrium is in modal coordinates. For c > c, we obtain slower linear amplitude decay in the twodimensional modal subspace of the x variable than in the modal subspace of the y variable. This suggests a reduction to a model involving only the slower x variables. The argument used in Example, however, shows that 44 violates assumption A3 and hence Proposition does not apply. The static condensation procedure, nevertheless, gives the formal reducedorder model ẍ + c + μ x ẋ + k x + bx 3 = 0, 45 and the method of modal derivates formally gives the formal reduced model ẍ + c + μ x ẋ + k x + b ac x 3 = 0, 46 k modifying 45 at cubic order only. While a global slow manifold is not guaranteed to exist in this example, a unique, two-dimensional analytic invariant manifold tangent to the subspace of the x variables at the origin does exist cf. Haller and Ponsioen 5. This spectral submanifold SSM offers a mathematically rigorous process for model reduction in system 44, providing an exact reduced flow to which 45 and 46 can be compared. As we show in Appendix 3, the reduced model on the slow SSM is of the form + ẋ x D + b ac 4c 4 6c3 c + c c + 5c k c c k + 3k + c k + 8k 6k k + k D + O4 = 0, x 3 47

630 G. Haller, S. Ponsioen where D = c c c + k 4c k 8c c k c c k + 4c k + 6k 8k k + k. 48 A comparison of the exact reduced model 47 with the statically condensed version 45 and with the modalderivatives-based version 46 shows that the latter two heuristic reduction methods miss most terms already in the leading-order cubic nonlinearities. Depending on the specific value of the parameters, the missing terms can significantly impact the nature of the reduced dynamics and hence cannot be omitted. We conclude that neither the Guyan reduction nor the method of modal derivatives gives correct result in this general case. Consider, however, now the parameter range expressed by the scaling c c /, k k /, 49 i.e., the case when the y-degree of freedom is notably stiffer and stronger damped as the x-degree of freedom in system 44. In this parameter domain, system 44 satisfies the assumptions of Proposition and hence the approximation to the slow SSM should coincide with the approximation to the global slow manifold M in this case. Indeed, after the above scaling, formulas 89 for the constants α, β and γ in Appendix 3 simplify to α = c +O 3, β=o 3, γ =O 3, k and hence the exact reduced model 47 simplifies to ẍ + c + μ x ẋ + k x + b ac k / x 3 + O 3 = 0, 50 coinciding with the modal-derivatives-based reducedorder model 46. This agreement, however, only holds in slow fast scaling 49. Even in the conservative limit, when the SSM is replaced by a unique, analytic Lyapunov-subcenter manifold Kelley 9, we obtain a conservative limit of the exact reduced-order model 47 in the form ac ẍ + k k 4k k ẋ x + b ac k k x 3 + Ox 4 = 0, 5 k 4k k filled with nonlinear normal modes periodic orbits. At the same time, the conservative limit of the static condensation procedure gives ẍ + k x + bx 3 = 0, 5 while the modal derivatives-based reduction 46gives ẍ + k x + b ac x 3 + Ox 4 = 0. 53 k Comparing 5 and 53 shows that the method of modal derivatives gives an incorrect reduced-order model up to cubic order, unless we have either a = 0 or c = 0. As shown in Fig. 6, error between the actual actual reduced flow 5 and 53 grows unbounded in the vicinity of the : resonance represented by k = 4k between the two natural frequencies of the undamped limit of system 44. In the limit of an exact : resonance, no invariant manifold tangent to the x-subspace exists, even though the modal derivative approach still suggests the existence a bounded reduced flow on such a manifold. 6 A detailed example: three-degree-of-freedom system with a pendulum damper We consider a the system depicted in Fig. 7, with a mass M hanging on a vertical spring of unstretched length L and linear viscous damping C h. The spring is hardening, with linear stiffness coefficient K h and cubic stiffness coefficient Γ h > 0. The mass is subject to downward external periodic forcing of the form f h t = f h0 sin ω t, as well as to gravity whose constant is g. The downward position of the mass from the unstretched spring position is measured by the coordinate h. The horizontal spring with linear stiffness coefficient K d and natural length D is fixed to the surroundings, thereby introducing geometric nonlinearities. Added in this direction is a viscous damper with damping coefficient C d and an external periodic force f d t = f d0 sin ω t, both acting in the horizontal direction. As indicated in Fig. 7, a pendulum of mass m and length l is attached to the mass M. The angle of the pendulum from the vertical is denoted by γ.the pendulum is also subject to angular viscous damping with coefficient c p, and to an external periodic force f p t = f p0 sin ω t acting on m in a direction normal to the pendulum. The equations of motion for this system are ml γ ml sin γ ḧ + ml cos γ d + c p γ + mgl sin γ = f p tl,

Exact model reduction by a slow fast decomposition 63.5.5 0.5 0.5 ẋ 0 ẋ 0-0.5-0.5 - - -.5 -.5 -.5 - -0.5 0 0.5.5 -.5 - -0.5 0 0.5.5 x x k = 4. k = 4.00 Fig. 6 Trajectories of cubic modal-derivative-based reduction 53 blue and those of exact cubic reduction 5 red to the unique, D analytic invariant manifold over the x, ẋ variables. The remaining parameters are set as k = a = b = c =. Color figure online M + m d + ml cos γ γ ml sin γ γ + C d ḋ + K d D + d Qd, h = f d t + f p t cos γ, K h, Γ h,c h D M K d,c d f h t g c p γ l L f d t d h f p t with the geometric nonlinear term Qd, h D Qd, h =. D + d + h 54 The linearized oscillation frequencies of the uncoupled springs and of the pendulum are Kh ω h = M, ω Kd g d = M, ω p = l, 55 respectively. With the help of these frequencies, we non-dimensionalize the h and d coordinates, the time t, and all system parameters by letting h = h L, d = d D, t = ω pt, m Fig. 7 Three-degree-of-freedom coupled pendulum M + mḧ ml sin γ γ ml cos γ γ + C h ḣ + K h h + K d Qd, hh + Γ h h 3 = M + mg + f h t f p t sin γ, Δ = l L, ρ = D L, β = m M, F ht = f ht Mg, F p t = f pt Mg, π h = F dt = f dt Mg, C h ω p M, π d = C d ω p M, π p = c p ω p ml, q h = ω h ω, q d = ω d p ω, a h = Γ h L p Mω, p G pt = f pt mg,

63 G. Haller, S. Ponsioen which leads to the following definition for the scaled version of Qd, h Q d, h = ρ ρ + d. + h Denoting differentiation with respect to the new time t still by a dot, then dropping all the tildes, we obtain the non-dimensionalized equations of motions Δ γ Δ sin γ ḧ + ρδcos γ d + π p γ + Δ sin γ = Δ G p t, + βḧ βδsin γ γ βδcos γ γ + π h ḣ + q h h + q d hqd, h + a h h 3 = + βδ + F h tδ F p tδ sin γ, + β d + β Δ ρ cos γ γ β Δ ρ sin γ γ + π d ḋ M = M M M M Δ = +β + β cos x γ ρδcos xγ β Δ ρ cos x, γ + β M = M M M M = Q = F M M F = q q + β + β sin, x γ q = π p ẋ γ Δ sin x γ + Δ G p t + Δ sin x γ βδcos xγ ẋγ + β π h ẏ q h y q d y Q x d, y a h 3 y 3 + + βδ + F h tδ F p tδ sin x γ q = β Δ ρ sin x γ ẋ γ π d ẋ d q d + x d Q + F d t Δ ρ + F pt Δ ρ cos x γ x d, y + q d + d Qd, h = F d t Δ ρ + F p t Δ cos γ. 56 ρ 6. Two softer degrees of freedom We are interested in applying the SFD procedure to system 56 to obtain an exact reduced-order model for the dynamics. First, we assume that h is a stiff degree of freedom and γ, d represent the softer degrees of freedom. In that case, using the notation from system, we can write the mass matrix Mq, t; and the forcing term Fq, q, t; as Δ ρδcos x γ Δ sin x γ Mq, t; = β Δ ρ cos x γ + β 0, βδsin x γ 0 + β Q = F M M F =βδcos x γ ẋγ π h ẏ q h y q d y Q x d, y a h 3 y 3 + + βδ + F h tδ F p tδ sin x γ + + β β sin x γ Δ + β sin π p ẋ γ Δ sin x γ x γ + Δ G p t βρ sin x γ cos x γ + β sin β Δ x γ ρ sin x γ ẋγ π d ẋ d q d + x d Q x d, y + F d t Δ ρ + F p t Δ ρ cos x γ. π p ẋ γ Δ sin x γ + Δ G p t β Δ ρ Fq, q, t; = sin x γ ẋγ π d ẋ d q d + x d Qx d, y + F dt Δ ρ + F pt Δ ρ cos x γ βδcos x γ ẋγ π h ẏ h q h y q d y Qx d, y a h 3 y 3. + + βδ + F h tδ F p tδ sin x γ Here we have introduced the coordinates x, y by letting x γ = γ, x d = d, y = h. The modified mass matrices M i and the forcing terms Q i defined in 5 take the specific form These give the following expression for the function P P x,v,η,w,t; = M qs q s, 57

Exact model reduction by a slow fast decomposition 633 q s = π p v γ Δ sin x γ + Δ G p t + Δ sin x γ βδcos xγ vγ + β π hw h q h η q d ηqx d,η a h 3 η 3 + + βδ + F h tδ F p tδ sin x γ, q s = β Δ ρ sin x γ v γ π dv d q d + x d Qx d,η + F d t Δ ρ + F pt Δ ρ cos x γ, with the inverse of M given by M = + β ρδcos x γ Δ β Δ ρ cos x γ + β cos. x γ Δ +β The function P takes the specific form P x,v,η,w,t; + β sin x γ = + β βδcos x γ v γ π hw h q h η q d ηqx d,η a h 3 η 3 + + βδ + F h tδ F p tδ sin x γ + + β β sin x γ Δ + β sin x γ π p v γ Δ sin x γ + Δ G p t βρ sin x γ cos x γ + β sin x γ β Δ ρ sin x γ v γ π dv d q d + x d Qx d,η+ F d t Δ ρ + F p t Δ ρ cos x γ. 58 We observe that lim 0 P x,v,η,w,t; 0, and therefore assumptions A A3 are not satisfied without further assumptions on the parameters that ensure the stiff soft partition of the coordinates. To this end, we express the stiffness of the y degree of freedom by letting Δ = l L = δ, ρ = D L = φ, q d = ω d ω p q h = ω h ω p π h = = Ω h, a h = α h 4, C h ω p M = μ h, π d = C d ω p M = μ d, = Ω d, π p = c p ω p ml = c p ω p m δ l = c pδ ω p m l = μ p. 59 In this parameter range, assumptions A A3 are satisfied, as we show in Appendix 4. The reduced model arising from these calculations is of the form γ = ω p M + m M + m sin c p γ ω γ sin γ + f pt p ml mg ω p M cos γ m M + m sin γ Mω sin γ γ C d p ω p Mlḋ K d Mg d + f dt Mg + f pt Mg cos γ + O, 60 ω d p = DM ml M + m sin γ MDω sin γ γ C d p MDω ḋ P K dl MgD d + f dtl MgD + f ptl MgD cos γ ω p Dm cos γ M + m sin c p γ ω p mdl γ l D sin γ + l f p t + O. 6 D mg We have implemented this model in Mathematica to show how a general trajectory xt of the full system is attracted to reduced model trajectories the slow manifold M. A graphical illustration of this behavior is shown in Fig. 8. For a numerical illustration of the accuracy of the reduced model, we choose the following values for the system parameters: l = D = 6m, L = m, M = m = kg, K h = 600 N/m, Γ h = 0.5 N/m 3, K d = N/m, C d = 0.33 ω p M kg/s, C h = 3 ω p M kg/s, c p = 0.33 ω p m L kg m /s, g = 9.8 m/s, f p t = 0.5 sint N, f h t = f d t = 0.5 sin3t N, = 0 8. We give the full system the initial condition q 0 = γ 0, d 0, γ 0, ḋ 0, h 0, ḣ 0 =.000,.00, 0.000, 0.000, 0.088, 0.00530, which lies off the slow manifold M, then integrate the trajectory starting from this initial condition in forward time. We track the Euclidean distance

634 G. Haller, S. Ponsioen Reduced Dynamics h,h t between the fast variables ht, ḣt and the explicitly computable slow manifold M for the given slow variables γ t, dt, γt, ḋt. When the fast variables are O0 5 close to M after the time value t t = 5.6 s, we take the point x t belonging to the full trajectory and use the slow coordinates γ t, dt, γt, ḋt of this point as an initial position for reduced model 60 and 6. Consecutively, we simulate the reduced model in backward and forward time and compare the results with the results obtained from the full model see Figs. 9, 0,. Fig. 8 Illustration of the attracting slow manifold M for mechanical system 56, graphed over the two slow degrees of freedom x γ and x d and their corresponding velocities. A general trajectory qt is attracted to the slow manifold, synchronizing exponentially fast with a trajectory of reduced dynamics dashed line 6. Two stiffer degrees of freedom We reconsider here the same mechanical system as in Sect. 6., but assume now that both the d and h variables represent stiff degrees of freedom, while γ still describes a softer degree of freedom. In this set- a b Fig. 9 Exponentially fast synchronization of the softer γ, γcoordinates of the full trajectory and of a reduced model trajectory a b Fig. 0 Exponentially fast synchronization of the softer d, ḋ coordinates of the full trajectory and of a reduced model trajectory

Exact model reduction by a slow fast decomposition 635 Fig. Exponentially fast convergence of the fast coordinate h along the full trajectory to the same coordinate along a trajectory of the reduced system ting, the anticipated slow variable x and fast variable y = y d, y h are defined as x γ = γ, y d = d, y h = h, We express the stiffness of the y degree of freedom by letting Δ = l L = δ, q d = ω d ω p π h = π p = = Ω d, q h = ω h ω p a h = α h 4, = Ω h, C h ω p M = μ h, π d = C d ω p M = μ d, c p ω p ml = c p ω p m δ l = c pδ ω p m l = μ p. 6 As we show in Appendix 5, assumptions A A3 are satisfied in the parameter regime represented by the above scaling for 0 <. In the scaled variables, we have δ M = + β. Thus, the mass matrix M associated with the slow degree of freedom is not differentiable at = 0. Therefore, only the more general form 8 of the reduced model is applicable, giving ẍ = P x, ẋ, G 0 x, ẋ, t, 0, t; 0 + O = μ p δ ẋ sin x + G pt + O. Scaling back to the original time, we conclude that at leading order, the exact reduced-order model on the two-dimensional, attracting slow manifold M is given by ẍ + μ p δ ẋ + sin x = G pt + O, or, equivalently, ml ẍ + c p ẋ + mgl sin x = f p tl + O. 63 As for the example treated in Sect. 6., we illustrate numerically that trajectories of the full system synchronize exponentially fast with those of the reduced-order model. For the parameter values l = 6m, L = 3m, M = 0.5 kg, m = 0.5 kg, K h = 000 N/m, Γ h = 0.5 N/m 3, K d = 80 N/m, C d = 3 ω p M kg/s, C h = 3 ω p M kg/s, c p = ω p m L kg m /s, g = 9.8 m/s, f p t = 0.6 sinω p t, f h t = f d t = 0, = 0 8, and the initial condition, x 0 = γ 0, γ 0, h 0, d 0, ḣ 0, ḋ 0 =.000, 0.000, 0.0084, 0.096, 0.000555, 0.00546, we illustrate the convergence of the trajectory to a trajectory of the reduced model on the slow manifold in Fig.. The target model trajectory was identified as earlier in the soft soft stiff version of the same example. 7 Conclusions We have developed a methodology for exact model reduction in multi-degree-of-freedom mechanical systems with softer and stiffer degrees of freedom. This SFD approach allows for a systematic identification of parameter regimes in which an attracting slow manifold exists. On this invariant manifolds, the stiff variables are enslaved to the remaining softer variables. We have derived explicit expressions for the slow manifold and for the first two orders of the reduced flow on this manifold. The latter formulas provide a mathematically exact reduced-order model with which trajectories of the full system synchronize at an a priori predictable exponential rate. We have also identified a domain boundary over which the slow manifold generically loses its stability and hence the dynamics on it no

636 G. Haller, S. Ponsioen Fig. Instantaneous projection of the slow manifold M for the periodically forced, stiff-stiff soft mechanical system at t = 0 s. a A trajectory of the full system is launched at the initial condition x 0 =.000, 0.000, 0.0084, 0.096, 0.000555, 0.00546 andintegratedinforwardtime.displayedinred is the h component of the corresponding full system trajectory that converges to the slow manifold and synchronizes with the reduced-order model dashed, shown up to time t = 5.6 s. b Convergence of the horizontal coordinate d red to the slow manifold, synchronizing with the dynamics of the reduced-order model dashed. Color figure online longer serves as a reduced-order model for the mechanical system. Slow fast reduction has previously been carried out with varying levels of mathematical rigor in several specific mechanical model problems see the Introduction for a review. Our contributions here are: i explicit conditions under which an attracting slow manifold in guaranteed to exist in a general, multi-degreeof-freedom mechanical system; ii readily applicable general formulas for reduced-order models on such manifolds. All these results follow from the application of classic results from geometric singular perturbation theory see, e.g., Fenichel 7, Jones 8. We have found that the SFD conditions yield reduced-order models that satisfy the basic requirements R R we have formulated for a mathematically exact model reduction procedure. As we has shown explicitly in Sect. 5, the formal methods of static condensation and modal derivatives in structural dynamics can only be justified if the conditions of SFD are satisfied. When these conditions do not hold, the reduced-order models produced by these methods are inaccurate or even qualitatively incorrect. Importantly, the SFD approach does not require the explicit identification of eigenfrequencies and normal modes for a linearized system, which is a numerically costly undertaking for high-degree-of-freedom systems. Instead, the SFD can be carried out based on a general identification of stiffer and softer vibratory modes, without an explicit decoupling of these modes. This flexibility for the method enables its application in structural vibrations problems such as those including forced and damped beams cf. Jain et al. 7 for a detailed example involving the von Kármán beam model. An extension of the SFD methodology to stiff soft continuum vibrations described by partial differential equations should also be possible through an appropriate extension of the necessary geometric singular perturbation results to infinite dimensions see, e.g., Menon and Haller 3. Acknowledgements We would like to thank Paolo Tiso, Daniel Rixen and Shobhit Jain for useful conversations and for their insights on the subject of this paper. Appendix : Proof of the main result First-order autonomous form By the nondegeneracy of M, the matrices M and M are necessarily invertible, which enables us to split in the form M M M M M M M M ẍ = F M M F, ÿ = F M M F. The nondegeneracy of M also implies that the two matrices on the left-hand side of this system must be