106
SEISMIC BASE ISOLATION BY MEANS OF NONLINEAR MODE LOCALIZATION BY YUMEI WANG B.ENGR., Tongji University, 1991 M.ENGR., Tongji University, 1999 THESIS Submitted in partial fulfillment of the requirements for the degree of Master of Science in Civil and Environmental Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2003 Urbana, Illinois

SEISMIC BASE ISOLATION BY MEANS OF NONLINEAR MODE

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

SEISMIC BASE ISOLATION BY MEANS OF NONLINEAR MODE LOCALIZATION

BY

YUMEI WANG

B.ENGR., Tongji University, 1991 M.ENGR., Tongji University, 1999

THESIS

Submitted in partial fulfillment of the requirements for the degree of Master of Science in Civil and Environmental Engineering

in the Graduate College of the University of Illinois at Urbana-Champaign, 2003

Urbana, Illinois

Abstract

This study assesses the performance of a nonlinear base isolation system composed of a

nonlinear foundation tuned in 1:1 internal resonance to a flexible mode of the main structure that

is to be isolated. Smooth stiffness nonlinearity of the third degree is initially considered. Under

this condition it is shown that nonlinear mode localization occurs, whereby a localized nonlinear

normal mode (NNM) is induced in the system, that confines energy to the foundation and away

from the structure to be protected. The application of nonlinear localization to seismic isolation

distinguishes this study from other base-isolation studies in the literature. After reviewing the

literature in the field of seismic base isolation and NNMs, a numerical study of the NNMs of the

main structure and nonlinear foundation under consideration is carried out, and the existence of

the localized NNM that is the basis of the proposed design is confirmed. A numerical study with a

Matlab – based model is then performed, considering ground motions representing near-field

seismic effects. The responses of the system with the proposed nonlinear foundation are compared

to linear designs, and the improved seismic isolation performance of the proposed system is

established. Additional simulations are performed by replacing the third-order smooth stiffness

nonlinearity with a clearance. It is found that the introduction of the simple clearance nonlinearity

leads to significant reduction of seismic energy transmitted to the main structure, resulting in

significant improvement of the seismic isolation effectiveness of the foundation.

iii

Acknowledgements

I would like to express my gratitude to all those who made this thesis possible.

In the first instance, I want to thank the Department of Civil and Enviromental Engineering,

for recruiting me and for providing me with advanced education. I am also obliged to the

Department of Aerospace Engineering for the working environment they provided and everything

they did to facilitate this work.

I am deeply indebted to my advisor, Professor Lawrence A. Bergman. His continuous

support, guidance, encouragement, and tolerance helped me in all the time of this research work.

His broad knowledge and deep insight always make him give stimulating suggestions at the first

moment. He is a respectable mentor in both research and life, and a sage director in difficult

times.

I also want to thank my other advisor, Professor Alexander F. Vakakis, who stimulated my

interest in nonlinear dynamics and opened me a window to a field beyond traditional Civil

Engineering. I appreciated his help as well as his working style: promoting self-confidence and

autonomy. His enthusiasm in research work and exploration for truth set an example to me.

My deepest appreciation would forward to Dr. D. Michael McFarland, who closely read my

numerous revisions, corrected my grammatical errors, offered valuable information for

improvement, supplied lots of tips and materials, and helped make sense of confusions. I sincerely

thank him for his consistent kindness, patience, and help in striving everything for excellence.

This research project was supported by the National Science Foundation Grant CMS-00060.

This sponsorship is greatly acknowledged. The study, with all the efforts of this Thesis, would

have been impossible without this support.

Finally, thanks to my family and my friends who endured this long process with me, always

offering support and love. I couldn't have made it without you. All other people who contributed

to this Thesis in many ways but I did not mention; thank you from the bottom of my heart!

iv

Table of Contents

Chapter 1: Introduction to Linear and Nonlinear Passive Vibration Control…………1

Chapter 2: Dynamic Response of Tuned Linear Systems……………………………….6 2.1 Introduction…………………………………………………………………………………...6 2.2 Analysis of a Classical 2DOF Tuned System……………………………………..………….7 2.2.1 Free Vibration………………………………………………………………….………..8 2.2.2 Forced Vibration……………………………………………………………………….10

2.3 Analysis of Our Proposed, Weakly Coupled MDOF System……………………………….11 2.3.1 Tuning Algorithm………………………………………………..…………………….12 2.3.2 Vibration Analysis……………………………………………..………………………13

Chapter 3: Estimation of the Undamped Nonlinear Normal Modes …………………20 3.1 Description of the System…………………………………………………………………...20 3.2 Undamped NNMs Using the Complexification-Averaging Method ………………………..22 3.3 Numerical Results for Undamped NNMs…………………………………..………….…....25 3.3.1 Physical Parameters for Estimation…………………………………………….……....25 3.3.2 Solutions of the Algebraic Equations…….…………………………………………….26 3.3.3 Figures and Discussion….……...……………….…………………………...…………27

Chapter 4: Model Building for System Analysis in Simulink…………...……………..31 4.1 Introduction………………………………………………………………………………….31 4.2 Elements of a Model…………..…………………………………………………………….31 4.3 Scalar System: the Nonlinear Isolation System……………………………………………..32 4.4 Vector (State-Space) Linear System: the Superstructure…………………………………....34

Chapter 5: Ground Motions Considered in This Study………….……………………...37 5.1 Introduction………………………………………………………………………………….37 5.2 Makris and Chang’s Model………………………………………………..………………...38 5.3 He’s Model…………………………………………………………………………………..39 5.4 Ground Motions Employed in This Thesis……………………………..…………………...41 5.4.1 Full Sine Wave Ground Velocity………………………………………………………41 5.4.2 Historic Earthquake Record……………………………………………………………42 5.4.3 Approximation by He’s Method………………………………………………………..43

Chapter 6: Numerical Simulation of the Undamped System…………...………………45 6.1 Approximation of the Cubic Nonlinearity by a Clearance Nonlinearity.…………………....45 6.2 Response to Full Sine-Wave Ground Velocity Pulses……………………………………....46 6.2.1 Shock Spectra……………………………………………..……………………………47 6.2.2 Response in the Time Domain……………………………………………..……..……48 6.2.3 Response in the Frequency Domain…………………………………………………....50

6.3 Response to Scaled Histories of Earthquake Records………….………………….…...……51 6.3.1 Response in the Time Domain…………………………………………………………51 6.3.2 Response in the Frequency Domain……………………………………………………53

6.4 Response to the Analytical (He) Model of Ground Velocity…………………….………….54 6.4.1 Response in the Time Domain………………………………………………………....55 6.4.2 Response in the Frequency Domain…………………………………………………...55

v

Chapter 7: Numerical Simulation of the Damped System………………………….…..57 7.1 Construction of the Viscous Damping Matrix……………………………….……………....57 7.1.1 Classical Damping Matrix for the Superstructure……………………………………...57 7.1.2 Damping in the Base and Subfoundation…………...………………………………….58

7.2 Response to the Analytical (He) Model of Ground Velocity………………………………..60 7.2.1 Damping in the Superstructure Only ( %2=ς s )……………………………………….60 7.2.2 Damping in the Subfoundation Only ( %2=ςc )……………………………………....62

7.3 Parametric Analyses………………………………………………………………………....63 7.3.1 Clearance Size and Clearance Spring Stiffness………………………………………...64 7.3.2 Damping at the Base Only……………………………………………………………...68

Chapter 8: Comparison of the Nonlinear Isolation System with Conventional Base Isolation……………..…….………………………………………………………..73 8.1 Conventional Base Isolation…………………………………………………………….…...73 8.2 Response to the Scaled Velocity History……….…………………………………………...74 8.2.1 Velocity History Scaled to the First Deformational Mode ( 1.0=gT s)…………...…...74 8.2.2 Velocity History Scaled to the Fundamental Mode ( 45.0=gT s)…….………….…….76

8.3 Some Final Thoughts………………………………………………………………………...78

Chapter 9: Conclusions and Recommendations for Further Study.………...…………80

List of References……………………………..…….………………………………………....81

Appendix A Simulink Model: Nonlinear Base Isolation System………..…..……..…86

Appendix B Matlab Code: Main Program….……………………………...……………..89

Appendix C Matlab Code: Full-sine Ground Velocity Pulse…………..……...…....…96

Appendix D Matlab Code: Envelope of the Control Force………….……...…….…...97

Appendix E Simulink Model: SDOF Oscillator.……….…………………..……………98

Appendix F Matlab Code: Response Spectra.……………………………...…………….99

Appendix G Simulink Model: Conventional Isolation System………..……….……101

vi

Chapter 1: Introduction to Linear and Nonlinear Passive Vibration

Control

The protection of structures from transient excitations has been a very important issue in

vibration engineering for many years. It represents an area with many and broad applications,

encompassing both industrial and infrastructural problems.

Methods for the design of shock protective systems can be found in standard vibration

handbooks and texts (Crede, 1951, Thomson, 1972). In most applications, system designs based

on passive control are the approach of choice. Usual passive control devices include vibration

isolators, vibration dampers, and vibration absorbers.

Vibration isolators depend on decoupling the structure from ground motions. This

decoupling is achieved by increasing the flexibility of the system, together with providing

appropriate damping (Skinner et al., 1993). As a result, the fundamental period is lengthened;

input energy content at higher structural frequencies that produce deformation in the structure

cannot be transmitted into the structure due to mode orthogonality. In many, but not all,

applications, the isolation system is mounted beneath the structure and is referred to as “base

isolation”. The concept of base isolation has now become a practical reality and has been applied

to many structures and types of equipment (Naeim and Kelly, 1999, Nagarajaiah et al., 2000).

Vibration dampers are devices that are capable of producing reactive forces proportional to

powers of input displacement and/or velocity. Common examples are viscoelastic, piezoelectric,

and fluid dampers. Such a damping device can also be a mass-spring system (Reed, 1961, Jones et

al., 1967) or a centrifugal pendulum (Shaw and Alsuwaiyan, 2000). Another important type is the

impact damper, which might consist of loosely-supported (Liu and Marr, 1987) or freely-moving

masses or particles. These devices dissipate energy and, by so doing, reduce system response.

Grubin (1956), Masri (1973), and Yokomichi et al. (1996) investigated the response of systems

equipped with acceleration dampers which consist of particles constrained to move in a container

with a certain clearance. The dissipation mechanism for these dampers is the momentum transfer

by collision and the conversion of mechanical energy into heat; the damping effect varies with the

acceleration level the particles achieve between collisions. Masri (1973) presented an exact

solution for two-symmetric-impact motion. Yokomichi et al. (1996) noticed that accelerations of

the particles are proportional to the modal amplitude of the container’s location. He thus proposed

a mass ratio that is modified by the modal vector for the system analysis.

Another type of passive vibration controller is the vibration absorber. Unlike vibration

dampers that dissipate energy, the vibration absorber, damped or undamped, functions as a

discrete, tuned, resonant energy transfer device (Nashif et al., 1985). This also differs from

isolation technology in that the isolation system does not absorb vibrational energy but, rather,

deflects it through the dynamics of the system (Nashif et al., 1985). A linear vibration absorber is

capable of protecting the main structure away from a selected range of excitation frequencies. The

effective bandwidth is dictated by the damping in the absorber, and a trade-off exists between

attenuation efficiency and bandwidth. Since its invention in 1909 by Frahm, the concept has

attracted the special and continued attention of many researchers. Den Hartog (1956) lucidly

described the working principle of the device in his monograph. Warburton and Ayorinde (1980)

extended the solution repertoire by examining several classes of excitation and responses to be

controlled. Recently, Sadek et al. (1996) extended the work of Villaverde (1985) to find optimum

parameters of a tuned mass damper by making the modal damping ratios of the first two modes of

vibration equal for the reduction of seismic responses of structures. The use and limitation of

these formulas with multi-degree-of-freedom systems was also discussed, and a study of the

possibility of controlling multiple structural modes with the multi-tuned mass dampers through

tuning each damper to the corresponding mode was carried out (Li, 2002).

Each strategy mentioned above has its limitations and reasonable range of application. In a

base-isolated structure, damping is introduced in the isolator to limit excessive displacements as

well as suppress possible resonances (Kelly, 1997). Fail-safe systems have been designed such

that they can be activated when ground shock levels at a site are exceeded. In some, the building

impacts against a stop when the design displacement is exceeded (Tsai, 1997, Malhotra, 1997). In

another approach, a sliding surface is provided with a small clearance; beyond this displacement,

the surface is contacted and vertical load transferred from the bearings to the sliders thereby

reducing the potential for collapse and increasing damping through friction (Kelly et al., 1980).

However, high damping at the base and fail-safe impact amplify high-frequency responses and

accelerations, which adversely affects performance.

The major limitation of linear vibration absorbers is that they are effective only in the

neighborhood of a single frequency. This narrow-band effectiveness poses problems when the

excitation is not fixed, and the resulting response at resonant frequencies can adversely affect

overall absorber performance. The problem is exacerbated by the fact that the excitation is seldom

a pure sinusoid at a single frequency but often contains secondary harmonics. Damping can be of

modest help in the magnification region but is a hindrance in the isolation region.

The fact that linear vibration absorbers can be vibration amplifiers provided the motivation

to investigate the performance of nonlinear absorbers. In his paper, Roberson (1952) gave a

2

parametric analysis showing that, for some optimal parameters, a cubicly nonlinear absorber

offers a wider frequency “suppression band” than the corresponding linear absorber. Pipes (1953)

showed that a vibration absorber with a nonlinear spring modeled as a hyperbolic sine function

can prevent sharp resonant peaks and introduce odd harmonic components of relatively small

amplitude. Arnold (1955) studied the effect of hardening and softening nonlinearities in a

vibration absorber. A jump phenomenon about the crossing frequency in this system indicated the

possibility of a mode of vibration for which the main mass is substantially at rest. Some recent

localization work also showed the effectiveness of nonlinear absorbers in reducing the resonant

response in some cases. Specifically, Leo and Inman (1999) compared the performance of a

passive isolation system and an active-passive vibration isolation system using a quadratic

optimization method. Cuvalci et al. (2002) experimentally studied a passive, tuned vibration

absorber under sinusoidal and random excitations and determined the parameters most effective

for absorption.

However, these studies do not provide insight into the nonlinear dynamics and so cannot

provide a general methodology for design. The search for such a paradigm led researchers to the

concept of motion confinement. The theory behind the phenomenon of passive motion

confinement is mode localization. Studies of mode localization initiated from linear systems with

weak structural irregularity. It was known that random imperfections in substructures may lead to

passive confinement of vibration in weakly coupled linear periodic systems (Anderson 1978,

Hodges, 1982, Pierre et al., 1987, Pierre, 1988). Later, it was found that mode localization could

also occur in perfectly symmetric nonlinear systems, with the only prerequisite being weak

coupling between subsystems (Vakakis et al, 1996). For example, Luongo (1991) showed that

geometric nonlinearity in systems with high modal density has the same effect on localization that

structural imperfections have in linear theory. In fact, the occurrence of nonlinear mode

localization is not an artifact of the symmetry of the system. It is caused by the amplitude-

dependent frequencies in the nonlinear systems. “Mistuning” of the nonlinear system results in a

mere perturbation of existing modes (Vakakis, 1993).

Nonlinear mode localization can be studied in the framework of nonlinear normal modes

(NNMs) and gives rise to a variety of nonlinear dynamic phenomena that can be used to develop

robust shock and vibration isolation designs for certain engineering systems. The concept of

NNMs was first introduced by Rosenberg (1960, 1963, 1964) and was later applied to the study of

nonlinear oscillations by other researchers (Vakakis, 1990, Vakakis and Rand, 1992a, b; Shaw

and Pierre, 1991, 1993). Rosenberg (1963) defined NNMs as “vibration in unison”; i.e.,

synchronous periodic motions during which all coordinates of the system vibrate equiperiodically.

Vakakis (1996) defined NNMs as free oscillations of discrete or continuous undamped nonlinear

3

systems where all coordinates vibrate in unison reaching their maximum and minimum values

simultaneously. From these definitions, it can be concluded that NNMs are extensions of the

classical normal modes of linear vibration theory. However, in contrast to the linear case, NNMs

can exceed in number the degrees of freedom of the nonlinear system due to NNM bifurcations

(Caughey et al, 1990, Rand et al, 1992). This was shown in Anand’s (1972) investigation on the

natural modes of a nonlinear system with 2 DOF. The system was shown to possess three modes

of vibration for both hardening and softening nonlinearities. The stability analysis indicated seven

different modal stability patterns, depending on the values of the parameters of nonlinearity.

Apparently these modal patterns do not have counterparts in the linear theory. NNMs also differ

from linear normal modes in that a general transient nonlinear response cannot be expressed as a

linear superposition of NNMs. A localized NNM is a subclass that leads to passive vibration

confinement. The existence of such localized normal modes was rigorously proved by MacKay

and Aubry (1994) in weakly coupled nonlinear chains.

The use of NNMs and nonlinear localization phenomena for vibration isolation has been the

subject of copious research in recent years. Most of the research on nonlinear modal interactions

focuses on internal or auto-parametric resonance, because internal resonance may provide a

coupling in a system, and under certain conditions energy transfer can occur between modes at

different frequencies (Nayfeh et al., 1994). Nayfeh et al. (1997) used this principle for the

optimization of a vibration isolation system. Nonlinear localization used for isolating structures

from earthquake-induced motions was also studied. Vakakis et al. (1999) designed a nonlinear

device, a spring-mass subfoundation, as a nonlinear absorber for a continuous beam with a base

mass. The parameters were selected such that a 1:1 internal resonance is induced between the

device and the primary structure. As a result, nonlinear mode localization occurred in the system;

seismic excitation was confined to the substructure, and the primary structure was protected.

These studies, and many others, revealed a variety of exclusively nonlinear phenomena that

cannot be modeled by linear or even linearized methodologies. They also proved that suitable

placement of nonlinear elements in a linear system can alter its modal properties, introducing new

stable modes. Nonlinear attachments, if properly designed, can act as passive vibration absorbers

to prevent unwanted disturbance. Due to the properties of NNMs, it is possible to design a base

isolation system using the localization property.

The system that will be examined in this thesis is a base isolation system, with a tuned

subfoundation as the nonlinear vibration absorber. Our study has two main objectives: First, to

validate the concept of nonlinear localization in the context of shock isolation of flexible

structures subjected to transient base inputs; and second, to apply these results to design and

4

performance assessment. The goal of this work is to compare the performance of the nonlinear

system with that of the corresponding linear system and to assess the potential of the former.

In our design, we synthesize each passive element in the above-mentioned three passive

vibration controllers: the structure to be isolated is weakly coupled with the subfoundation mass

and, thus, realizes a lengthened fundamental period of vibration, as in common base-isolated

structures; the vibro-impact element (the impact damper or the impact fail-safe mechanism in

some base isolation systems) is used to induce nonlinearity in the system; internal resonance is

realized by tuning the subfoundation to the primary structure based on linear dynamics, so in the

absence of the nonlinear element the subfoundation is a linear vibration absorber. However, the

theory behind our design relies on nonlinear normal modes and mode localization, so these

elements have different functions in our system. The weak coupling generates energy

redistribution (internal resonance) necessary for mode localization to occur, and the clearance

impact is to provide nonlinearity and “mistuning”, or perturbation, to trigger the NNMs.

The non-smooth nonlinearity we are using is provided by two springs, one having a gap.

This non-smooth nonlinearity, namely a clearance nonlinearity, is an approximation of a high-

order smooth nonlinearity. It is seen that there are only moderate quantitative differences

separating them. Its effectiveness has been shown in previous studies (Vakakis et al, 1999,

McFarland et al., 2001a,b, Wang et al., 2002, Jiang, 2002). Apart from the fact that such non-

smooth elements induce strong nonlinearity in the system, they are rather easy to implement in

practical settings since they are realized by means of linear stiffnesses. This makes the design

easy to validate experimentally with realistic forcing conditions.

This thesis contains nine chapters. First, we review general internal resonance and the

beating phenomenon in a linear system. Then we deduce analytically the undamped NNMs of the

system with a third-order nonlinearity using the complexification-averaging method. In the two

chapters that follow, we introduce the simulation model of the building and ground motions

considered in this study. Chapters 6 and 7 present numerical simulation results for undamped and

damped systems as well as a sensitivity analysis. Then we compare the performance between this

system with the conventional isolation system. Chapter 9 summarizes conclusions. We will see

that the simple introduction of the clearance nonlinearity leads to significant reduction of energy

transmitted to the primary structure, and localization of this energy to the secondary structure.

5

Chapter 2: Dynamic Response of Tuned Linear Systems

2.1 Introduction

Predicting the response of a primary system (building, nuclear plant, etc.) with one or more

appended secondary systems has received considerable attention over the years. Special emphasis

has been given to secondary systems that are tuned to a natural frequency of the primary system,

sometimes referred to as vibration absorbers or tuned mass dampers. Tuning of secondary systems

is of great concern to the designer since the configuration, if misdesigned, can result in extremely

large response of the entire system. Thus, exact tuning of the primary and secondary systems is

generally assumed, fully recognizing that frequencies apparently identical are actually in close

proximity.

Since the secondary system is generally light in comparison with the main structure, one way

to design the secondary system is to assume that the secondary system does not perturb the

motion of the primary system. The equations of motion then decouple into two smaller sets of

equations that can be solved in succession (Nakhta, 1973, Khachian et al., 1998, Der Kiureghian

et al., 1999).

A common approach for the design of linear systems subjected to transient inputs employs

the notion of a shock (or response) spectrum, which represents a plot of the absolute value of the

maximum response of a damped, linear, single-degree-of-freedom (SDOF) oscillator versus the

natural frequency (or period) of transient input of a specified shape. When a multiple-degree-of-

freedom system is tuned by a SDOF subsystem as described above, however, the validity of the

shock spectrum is questionable since modal interactions can be significant. The presence of a

tuned subsystem can also make the use of experimental modal analysis somewhat problematic, as

such systems can have two eigenfrequencies very close to the frequency of tuning. While these

modes often contribute significantly to the response, the close spacing of the frequencies makes it

difficult to compute the modal data or infer the joint response of the modes (Ruzicka, 1981).

Caughey (1965) extensively studied this problem in the context of vibration absorbers.

A number of analytical methods have been devised to account for the dynamic interaction

between components (Penzien and Chopra, 1965, Thomson, 1972, Ruzicka, 1981). It is known

that beating occurs when two natural frequencies of the system are close. This condition usually

occurs when the coupling between two subsystems is very weak. Examples abound in the

literature of vibration absorbers, coupled pendulum systems, electrical systems with parallel

capacitances, and fluid couplings with two cylinders. Transfer of energy takes place in the

6

coupled system which can induce vibrations in the primary structure instead of suppressing them.

The beat phenomenon has been discussed in many classical vibration texts and papers (see, e.g.,

Den Hartog, 1956, Yalla et al., 2000). This chapter focuses on understanding the phenomenon

from a mathematical point of view. Analysis is presented to elucidate the beat phenomenon in a

simple two degree-of-freedom (DOF) system, one being the vibration absorber. The presence of

beating in our proposed 4DOF tuned linear system subjected to a harmonic ground excitation is

also validated analytically.

2.2 Analysis of a Classical 2DOF Tuned System

Ruzicka (1981) studied the tuned 2DOF system shown in Figure 2.1. He examined the

response of the secondary system ma and gave analytical as well as approximate solutions. In this

section, we will focus on the response of the entire structure.

ya(t)+ Xg(t) yp(t)+ Xg(t) Xg(t)

kskp

mamp

Figure 2.1: Two-degree-of-freedom system

The equation of motion of this system in matrix form is

M (2-1) )()()( tXtt &&&& M1KYY −=+

where

M , , , (2-2)

=

a

p

mm0

0

−+=

aa

aap

kkkkk

K

=)()(

)(tyty

ta

pY

=11

1

and subscripts p and a represent primary and attached subsystems, respectively. It is assumed that

the attached mass and spring are tuned to the primary system, and that the ratio of the attached

mass to the primary mass is very small, giving

7

2ω==a

a

p

p

mk

mk

, 1<<==εp

a

p

a

kk

mm (2-3)

Substituting (2-3) into (2-1) and premultiplying both sides by yields 1−M

(2-4) )(111

001 2 tX

yy

yy

a

p

a

p &&&&

&&

−=

εε−ε−ε+

ω+

ε

2.2.1 Free Vibration

Let the right hand side of eq. (2-4) be null. The free vibration solution of (2-3) can then be

written as

(2-5) ∑=

+=2

1)Ωsin()(

iiiii θtAt φY

where is the ith natural frequency of the system, φ is the ith mode shape, θ are

phase angles, and are constants determined by initial conditions. The natural frequencies and

mode shapes are determined from the homogeneous algebraic equation

=ai

pii φ

φi

iA

(2-6) 0)(

)1(det 222

222

=

Ω−ωεεω−εω−Ω−ε+ω

where, for each nontrivial solution , ω

(2-7)

=

Ω−ωεεω−εω−Ω−ε+ω

00

)()1(

222

222

a

p

φφ

For ε <<1, the natural frequencies and mode shapes can be approximated by

ω∆−ω=ε−ω≈Ω )1(1 , ω∆+ω=ε+ω≈ )1(2Ω (2-8a)

+≈

ε1211

1φ ,

−≈

ε1211

2φ (2-8b)

8

Substituting, the response in free vibration is given by

])sin[(1211

])sin[(1211

)()(

2211 θ+ω∆+ω

ε−+θ+ω∆−ω

ε+=

tAtA

tyty

a

p

(2-9)

Notice that the responses of both the primary and secondary systems are the sum of two

sinusoidal oscillations which are close in frequency. This gives rise to a classical beat

phenomenon, in which the amplitude of the combined oscillation rises and falls as the component

oscillations drift in an out of phase. To see this, set AAA == 21 , 021 =θ=θ . This gives

tAtAtyty

a

p )sin(1211

)sin(1211

)()(

ω∆+ω

ε−+ω∆−ω

ε+=

ωω∆ε

ωω∆−≈

ωω∆ε

−ωω∆

ωω∆−=

ttA

ttA

ttAtt

ttA

cos][sin2

cos][cos2

cossin2sincos

coscos2 (2-10)

0.5 1 1.5 2 2.5t

-1

-0.5

0.5

1disp . Response of the Primary Mass

0.5 1 1.5 2 2.5 t

-40

-20

20

40disp . Response of the Secondary Mass

Figure 2.2: Beat phenomenon in free vibration

The motion described by equation (2-10) is the solid curve in Figure 2.2. The bracketed

terms , and their negatives are, respectively, the envelopes in the top and bottom

plots, represented by the dashed curves, where we chose

ω∆cos ω∆sin

2−=A , 50=ω , ε , and 01.0= 5=ω∆

for illustration. The rate of oscillation of the envelopes (the “beating”) is governed

by ω<<εω=ω∆ . ω∆ is referred to as the beat frequency, and the corresponding period

ω∆π= 2T is the beat period.

9

2.2.2 Forced Vibration

The steady-state solution of the forced vibration problem (2-4) can be obtained using

Duhamel’s integral

∑ ∫=

ττ−ΩτΩ

−=

=2

1 0

)(sin)()()(

)(i

t

ii

ii

a

p dtXP

tyty

t &&φY (2-11)

where i

Ti

Ti

iPMφφM1φ

= are the modal participation factors, φ is the modal matrix, φ is the ith

mode, the coefficient

i

i

iPΩ

is the magnification factor of the ith mode, and the integral is the ith

modal response. Substituting (2-8b) into Pi, it follows that

2

11

ε+≈P ,

21

2ε−

≈P (2-12)

Finally, substituting (2-12) and (2-8a) into (2-11) and integrating, responses of the primary and

the attached mass can be written in the form

)sin(),,(cos)(sin)()( 21 ppppp ttEttyttyty θ+ωω∆ω=ω+ω= (2-13a)

)sin(),,(cos)(sin)()( 21 aaaaa ttEttyttyty θ+ωω∆ω=ω+ω= (2-13b)

where and are slow-frequency envelopes of the fast-frequency

responses that are given by and , respectively, where

),,( tE p ω∆ω ),,( tEa ω∆ω

)(ty pi )(tyai

∫∫ ττ−ω∆ωττω

ε+ττ−ω∆ωττ

ω−≈

tt

p dtXdtXty00

1 )(sin]sin)([2)(cos]cos)([1)( &&&& (2-14a)

∫∫ ττ−ω∆ωττω

ε+ττ−ω∆ωττ

ω≈

tt

p dtXdtXty00

2 )(sin]cos)([2)(cos]sin)([1)( &&&& (2-14b)

∫∫ ττ−ω∆ωττω

−ττ−ω∆ωττεω

≈tt

a dtXdtXty00

1 )(cos]cos)([25)(sin]sin)([1)( &&&& (2-14c)

10

∫∫ ττ−ω∆ωττω

+ττ−ω∆ωττεω

≈tt

a dtXdtXty00

2 )(cos]sin)([25)(sin]cos)([1)( &&&& (2-14d)

It can be seen that , , , and all represent oscillations of the system

with slow frequency

)(1 ty p )(2 ty p )(1 tya )(2 tya

ω∆ , with the bracketed terms as external forces. For example, the first term

of (2-14a) represents the slow-frequency “velocity” response subjected to acceleration

, while the second term represents the slow-frequency “displacement” response

subjected to support acceleration . As a result,

ttX ωcos)(&&

tωsintX )(&& ),,( tE p ω∆ω and are both

harmonic functions. Thus, the forced vibration of the tuned system is also characterized by the

presence of beating. What makes a difference in the forced responses is that the beating frequency

is not , but a frequency determined by the forcing frequency and

),,( tEa ω∆ω

ω∆ ω∆ .

When , the forced response of the system can be readily calculated numerically

with Mathematica. For , ,

ttX ω= sin)(&&

A 2−= 50=ω 01.0=ε , and 5=ω∆ , the response is given by Figure

2.3.

0.5 1 1.5 2 2.5t

-0.002

-0.001

0.001

0.002disp . Response of the Primary Mass

0.5 1 1.5 2 2.5t

-0.04

-0.02

0.02

0.04disp . Response of the Secondary Mass

Figure 2.3: Beat phenomenon in forced vibration

2.3 Analysis of Our Proposed, Weakly Coupled MDOF System

To analytically solve for the eigenvalues and eigenvectors of a multiple degree-of-freedom

system is, in most cases, intractable for n , as the order of the polynomial characteristic

frequency equation is twice the number of DOF n. Design of the secondary system involves

choosing its parameters such that tuning to a resonant frequency of the existing primary system is

achieved. This further increases the complexity of the problem.

3>

11

The system we chose to examine has a total (primary plus secondary) of four degrees of

freedom. The primary system is connected to the secondary system through a very weak spring kb

(Figure 2.4). Due to this weak connection, the two subsystems can be regarded as decoupled when

selecting parameters for tuning. This makes it possible to readily find a partial analytical solution

for this system.

2.3.1 Tuning Algorithm

Consider the undamped linear system (spring kn does not exist) in Figure 2.4. Without loss of

generality, assume the masses of the superstructure to be m and the stiffness coefficients to be k.

The masses of the base and the subfoundation are 1.5m. The coupling stiffness coefficient is

, where ε . kkb ε= 1<<

)(g tx&&

m2

mb

GROUND

k1

mc

x1

m1

k2

kc kb

x2

Primary Structure

ee kn

kn

Figure 2.4: Our proposed linear system

Stiffness kc is to be tuned to one mode of the primary structure, which consists of the base

and superstructure. Accounting for the weak stiffness of kb, the lowest system natural frequency

will be nearly zero, corresponding to the almost rigid-body mode of the primary structure. The

second and third natural frequencies will be non-zero, and one of them will be used to tune the

subfoundation. In this study, the first deformational frequency of the primary structure, 2pω , is

chosen to be tuned with the subfoundation. On physical grounds, the resonant frequency is chosen

to be the frequency of the first mode in which the base-structure combination would experience

significant inter-story displacements.

12

Considering the linear system consisting of the base and superstructure, we can readily

compute its natural frequencies from the stiffness and mass matrices of (2-15):

, (2-15)

=

mm

m

p

5.1000000

M

ε+−−−

−=

kkkkk

kk

p

)1(02

0K

where subscript p represents the primary system. Frequencies pω are found by solving

. The determinant is reduced to the polynomial equation (2-16),

where λ .

0|||| 1 =λ−=λ− − IKMID ppppp

2pp ω=

02)()68()()211(3 223 =ε+λε++λε+−λ−mk

mk

ppp (2-16)

This equation can be solved if ε is given a value. Let 033.0=ε , i.e. 301 ; then

mk

p 0093.01 =λ , mk

p 8301.02 =λ , mk

p 8483.23 =λ (2-17)

This ε is small enough for the beating phenomenon to occur. It will be used through this Thesis.

The tuning spring coefficient is determined from 2pc ω=ω , where the resonant frequency of

the isolator degree of freedom is c

bcc m

kk +=ω . Thus,

k kkm pcc 2122.12 =ε−λ= (2-18)

and then tuning process is complete.

2.3.2 Vibration Analysis

Once kc is selected, the natural frequencies and mode shapes of the entire structure can be

determined. The mass and stiffness matrices of the system are

13

M , K (2-19)

=

mm

mm

5.100005.100000000

+εε−ε−ε+−

−−−

=

kkkkk

kkkkk

)2122.1(00)1(0

0200

The dynamical matrix D of the full structure is, for 033.0=ε ,

mk

−−−

−−−

== −

8301.0022.000022.06887.06667.0001210011

1KMD (2-20)

Setting | = 0 gives the eigenvalues |ID λ−

mk00906.01 =λ ,

mk81488.02 =λ ,

mk84547.03 =λ ,

mk84982.24 =λ (2-21)

The circular frequencies, λ=Ω , are

mk0952.01 =Ω ,

mk9027.02 =Ω ,

mk9195.03 =Ω ,

mk6880.14 =Ω (2-22)

and the natural frequencies in Herze. are πΩ 2 , or

mkf 0151.01 = ,

mkf 1437.02 = ,

mkf 1463.03 = ,

mkf 2687.04 = (2-23)

Note that and ( and ) are very close, which will lead the system to beat. Frequencies

and Ω can be rewritten as

3

3Ω 2f 3f

2202 ω∆−ω=Ω , 2203 ω∆−ω=Ω (2-24)

14

where mk9111.020 =ω ,

mk0084.02 =ω∆ . Note that 20ω is nearly equal to the tuning

frequency mk

mk

pp 9111.08301.022 ≈=λ=ω .

Solving for the modal vectors φ from [ ] 0φID =λ− gives

φ , , , (2-25)

=

2607.09729.09909.01

1

−−

=

1311.17806.0

18511.01

−=

1733.18216.0

1545.01

−=

0062.05707.08493.11

whereφ is the modal matrix and φ is the ith modal vector. Figure 2.5 graphically depicts the

mode shapes, which have all been normalized to unity at story 2 for uniformity.

i

Figure 2.5: Mode shapes of the linear 4 DOF system

Again, free vibration is the linear superposition of the four modes. Around the tuned

frequencies 2202 ω∆−ω=Ω , 2203 ω∆−ω=Ω ,

])sin[(])sin[()( 222032122021 θ+ω∆+ω+θ+ω∆−ω= tAtAt φφY (2-26)

Assume , and 121 == AA 021 =θ=θ

15

ttttt 2022320232 cos]sin)[(sin]cos)[()( ωω∆−+ωω∆+= φφφφY

≈ (2-27)

ωω∆−ωω∆+

tttt

20223

20232

cos]sin)[(sin]cos)[(

φφφφ

Equation (2-27) indicates that free vibration of the tu

( ) vibration modulated by slow-frequency (20ω 2ω∆ ) v

The forced response, on the other hand, is the su

factors and modal responses of each mode as

∑ ∫=

τΩ

−=4

1 0

sin)()(i

t

i

ii X

Pt &&φY

where y represents displacement with respect to the gro

, )()()( txtytx g+=

=

(((

)( 1

2

yyyy

t

c

b

Y

and i

Ti

Ti

iPMφφM1φ

= are the modal participation fa

, , and 0255.1 2 =P1 =P 4350.0− 4101.03 =P 4 −=P

factors are km10.7721 ,

km0752.2 ,

km446.0 0.446

From the above analysis, we can see that if the s

frequency content close to the tuned frequency 2pω , it

to the resonance effect, while its peak responses fall o

periodic vibration. In this condition, the first and the

combined effect is small. This can be

that are shown in Figure∫ ττ−Ωτ=t

ii dtXtA0

)(sin)()( &&

primary structure

subfoundation

ned 4DOF system is also a fast-frequency

ibration.

mmation of the products of magnification

ττ−Ω )(i dt (2-28)

und; i. e.,

(2-29)

))))(

tttt

ctors. In this example with 033.0=ε ,

0005.0 , the corresponding magnification

0, and km0003.0 , respectively.

ystem is subjected to ground motions with

will vibrate at the fast tuned frequency due

n the envelope of the slow-frequency ( ω∆ )

fourth modes can be ignored, because their

verified from the modal responses

2.6.

16

1 2 3 4 5 6t

-0.0075-0.005-0.0025

0.00250.0050.0075

A1 1st Modal Amplitude

1 2 3 4 5 6 t

-0.75-0.5-0.25

0.250.50.75

A2 2nd Modal Amplitude

1 2 3 4 5 6t

-0.75-0.5-0.25

0.250.50.75

A3 3rd Modal Amplitude

1 2 3 4 5 6t

-0.01

-0.005

0.005

0.01A4 4th Modal Amplitude

Figure 2.6: Resonant modal responses to harmonic ground excitation

It can be seen that, although the magnification factor of the 1st mode is five times that of the

2nd mode and 24 times that of the 3rd mode, the modal amplitudes of the 2nd and the 3rd modes are

100 times that of the 1st mode. So the response is mainly determined by the 2nd and the 3rd modes,

such that

∫∫ ττ−ΩτΩ

−ττ−ΩτΩ

−=t

g

t

g dtxP

dtxPt0

33

33

02

2

22 )(sin)()(sin)()( &&&& φφY

∫∫ ττ−ω∆−ωτΩ

−ττ−ω∆+ωτΩ

−t

g

t

g dtxP

dtxP

0220

3

33

0220

2

22 )])(sin[()()])(sin[()( &&&& φφ=

ttqPPtq

PP2023

3

32

2

213

3

32

2

2 sin)]()()()[ ωΩ

−Ω

−Ω

− φφφφ=

ttqPPtq

PP2043

3

32

2

233

3

32

2

2 cos)]()()()[( ωΩ

−+Ω

φφφφ+

(2-30)

where

17

q , q (2-31a) ∫ ττ−ω∆τωτ=t

g dtxt0

2201 )(coscos)()( && ∫ ττ−ω∆τωτ=t

g dtxt0

2202 )(sinsin)()( &&

q , q (2-31b) ∫ ττ−ω∆τωτ=t

g dtxt0

2203 ])(cossin)()( && ∫ ττ−ω∆τωτ=t

g dtxt0

2204 )(sincos)()( &&

Again, the motion Y(t) can be rewritten in the form

]sin[)()( 20 θ−ω= ttt EY (2-32)

where E(t) are slow-frequency envelopes limiting the peaks of fast-frequency (ω) oscillation

, and θ are the phase angles of the oscillation E(t) at each moment. ]sin[ 20 θ−ω t

Let , k =3500, and m =0.3. The solutions above can be readily calculated

numerically in Mathematica and are shown in Figure 2.7.

ttX ω= sin)(&&

1 2 3 4 5 6t

-0.006-0.004

-0.002

0.0020.0040.006

disp . Response of the 2nd story

1 2 3 4 5 6 t

-0.001

-0.0005

0.0005

0.001disp . Response of the 1st story

1 2 3 4 5 6 t

-0.004

-0.002

0.002

0.004

disp . Response of the Base

1 2 3 4 5 6t

-0.003-0.002-0.001

0.0010.0020.003

disp . Response of the Subfoundation

Figure 2.7: Resonant responses of the linear system subjected to harmonic ground excitation

18

It is apparent from these plots that there is energy transfer between the subfoundation and the

primary system.

For more complicated cases, where resonance does not occur, i. e., , the 1pg ω≠ω st mode

and the 4th mode cannot be ignored. Analysis for these conditions carried out numerically in

Matlab and Simulink. Model building in Simulink will be discussed in Chapter 4, and numerical

results will be given in Chapter 6.

19

Chapter 3: Estimation of the Undamped Nonlinear Normal Modes

3.1 Description of the System

The model of the nonlinear base-isolated structure is depicted in Figure 3.1. It is essentially

identical to the four-degree-of-freedom linear model discussed in the previous chapter, except that

the linear spring k with clearance has been added between ground and the subfoundation m .

The primary structure to be isolated consists of a 2-DOF superstructure plus a base mass. The

base is weakly and linearly coupled to an intermediate or secondary system which itself is

connected to ground through a bilinear hardening spring, which is simply the parallel combination

of two linear springs, one having a clearance nonlinearity. The intermediate system consists of the

subfoundation m

n e c

c and linear spring kc, which acts in conjunction with the clearance spring to

localize the ground input energy away from the primary structure. This is achieved by tuning the

absorber to a mode of the superstructure to be isolated so that internal resonance occurs. The

clearance spring kn acts to excite the appropriate localized nonlinear normal mode, while the

clearance e is a design parameter which determines the onset of nonlinear behavior.

)(g tx&&

m2

mb

GROUND

k1

mc

x1

m1

k2

kc kb

x2

Primary Structure

ee kn

kn

Figure 3.1: Schematic of the proposed system

The desired response of the complete structure to ground motion would resemble a nonlinear

normal mode (NNM) in which motion is almost entirely confined to the intermediate

subsystem—a localized mode in which the base and superstructure participate very little. We will

demonstrate that this nonlinear design is capable of localizing energy away from the primary

structure by examining the modal structure in the region of the internal resonance. To facilitate

20

the analysis to follow, a smooth stiffness to ground, composed of the sum of linear and cubic

terms, is presumed, as analysis employing non-smooth nonlinearity is beyond the scope of this

thesis.

Let x represent displacement with respect to a fixed reference frame and y displacement

with respect to ground, and let subscripts g and c represent ground and subfoundation,

respectively. Then, the equations of motion of the system are

m

(3-1)

−−=−+−+−+=−+−+−+−+=−+−+−+−+

=−+−+

)()()()(0)()()()(0)()()()(

0)()(

1111

212112121111

12212222

gcbcbbcbgcccc

bcbbbcbbbb

bb

xxFxxkxxcxxcxmxxkxxkxxcxxcx

xxkxxkxxcxxcxmxxkxxcxm

&&&&&&

&&&&&&

&&&&&&

&&&&

For free vibration, equations (3-1) can be rewritten as

m

(3-2)

−=−+−+−+=−+−+−+−+=−+−+−+−+

=−+−+

)()()()(0)()()()(0)()()()(

0)()(

1111

212112121111

12212222

cbcbbcbgcccc

bcbbbcbbbb

bb

yFyykyycxycymyykyykyycyycy

yykyykyycyycymyykyycym

&&&&&&

&&&&&&

&&&&&&

&&&&

where the restoring force provided by the cubic nonlinearity is given by

(3-3) 3)( ukukuF nc +=

while the restoring force provided by the clearance nonlinearity is given by

− (3-4)

++

−+=

)(

)()(

eukukuk

eukukuF

nc

c

nc

eueue

eu

−<≤≤

>

We replace this by the smooth nonlinearity in (3-3) for the purpose of this analysis.

The linear spring coefficient kc is used to tune the subfoundation or intermediate subsystem

to one of the modes of the primary structure; kc will be determined through the tuning process, as

has been discussed in Chapter 2. To ensure that the design embodies the requirements of weak

linear coupling between the base and the intermediate subsystem, we set

21

cb k1

=k , cn kk 2α= (3-5)

where α1 and α2 are values greater than one. Through this tuning procedure, all parameters

needed for the design of the intermediate system can be specified.

3.2 Undamped NNMs Using the Complexification-Averaging Method

The objective of this chapter is to find analytical forms of the NNMs of the undamped

system, as their structure will tell us much about the response of the damped, forced system.

Here, we find the steady-state NNMs for the smooth cubic nonlinearity.

We note that both linear and nonlinear systems admit single-frequency exponential solutions.

So, in this context, NNMs can be regarded as solutions that possess exponential temporal

dependence; i.e., solutions that consist of a “fast” oscillation with frequency ω modulated by a

“slow” envelope. The existence of the strongly nonlinear term in the fourth equation of (3-2)

prohibits the use of standard perturbation techniques for analyzing the dynamics. However, as

shown in Manevitch (2001), Vakakis (2001), and Jiang (2002), the dynamics of (3-2) can be

studied analytically by complexification and averaging. This simplifies the associated

computations and leads to results that are physically meaningful. Averaging relies on the

transformation of the equations of motion to an averaged set, which is more easily examined.

To this end we introduce the complex quantities

U , V)()()( tyityt bb ω+= & )()()( 11 tyityt ω+= & ,W )()()( 22 tyityt ω+= & , )()()( tyitytS cc ω+= &

(3-6)

In terms of these variables, the accelerations, velocities, and displacements can be expressed as

ω+

=+

=+

ω−=

ω−

=+

=+

ω−=

ω−

=+

=+

ω−=

ω−

=+

=+

ω−=

itStStytStStytStSitSty

itWtWtytWtWtytWtWitWty

itVtVtytVtVtytVtVitVty

itUtUtytUtUtytUtUitUty

ccc

bb

2)()()(,

2)()()(,

2)()()()(

2)()()(,

2)()()(,

2)()()()(

2)()()(,

2)()()(,

2)()()()(

2)()()(,

2)()()(,

2)()()()(

***

*

2

*

2

*

2

*

1

*

1

*

2

***

2

&&&&

&&&&

&&&&

&&&&

(3-7)

22

where i is the imaginary unit and * denotes the complex conjugate. Substituting (3-7) into (3-2),

the equations of motion are complexified and assume the form

−−+

+−

++

ω−

−−+

++

ω−

−+ω

=+

+−

−+ω

−−

ω−

+

+−

++

ω−

−−++

ω−

=+

+−−

−−

ω−

+

+−

++

ω−

−−++

ω−

=+

+−−

−−

ω−

++ω

0]2

)([)22

(

]2

)(2

)([22

)()](2

[

0)22

(]2

)(2

)([

)22

(]2

)(2

)([)](2

[

0)22

(]2

)(2

)([

)22

(]2

)(2

)([)](2

[

0)22

(]2

)(2

)([)](2

[

3***

*****

****

**

1

**

1*

**

1

**

1

**

2

**

2*

1

**

2

**

2*

2

SSikUUSSc

UUiSSikSScSSikSSiSm

UUSScUUiSSik

VVUUcVViUUikUUiUm

VVUUcVViUUik

WWVVcWWiVVikVViVm

WWVVcWWiVVikWWiWm

nb

bccc

bb

b

&

&

&

&

(3-8)

Since we seek steady-state vibrations of approximate frequency ω, we express the complex

variables in polar form

(3-9)

ωµ+µ=⇒µ=⇒ϕ=ωλ+λ=⇒λ=⇒ϕ=

ωσ+σ=⇒σ=⇒ϕ=ωϕ+ϕ=⇒ϕ=⇒ϕ=

ωωω−ω

ωωω−ω

ωωω−ω

ωωω−ω

titititi

titititi

titititi

titititi

etiettSettSettSetiettWettWettWetiettVettVettVetiettUettUettU

)()()()()()()()()()()()()()()()()()()()()()()()()()()()(

**

**

**

**

&&

&&

&&

&&

where , , and are complex amplitudes representing slowly varying complex

envelopes of the steady state motion. Substituting (3-9) into (3-8) and retaining only terms

containing the fast frequency ω, we obtain the complex modulation equations that govern the

slow dynamics (i.e., the envelope) of the response.

)(tϕ )(tσ )(tλ )(tµ

µµ−

ϕ−

µ+

ωϕ

+ωµ

−+µ

+ωµ

−µω

−+ωϕ

−ωµ

−+ωσ

+ωϕ

−+ϕω

−σ

+ωσ

−ωϕ

−+ωλ

+ωσ

−+σω

−λ

+ωλ

−ωσ

+λω

08

3)

22()

22(

22)

2(

0)22

()22

()22

()22

)2

(

0)22

()22

)22

()22

()2

(

0)22

()22

()2

(

3

*2

11

11221

222

nbbccc

bbb

ikciikcikim

ciikciikim

ciikciikim

ciikim

&

&

&

&

(3-10)

23

In deriving the above equations we have omitted higher harmonics in the fast oscillation.

There can exist subharmoic or superharmonic steady state responses, but the error resulting from

this approximation is not expected to be large for the fundamental frequency-amplitude plots that

will be discussed here.

For steady-state motions, the amplitudes of the above complex variables must vary much

more slowly than those with fast frequencies. Thus, we set their derivatives equal to zero in (3-

10); i.e., 0)( =tϕ& , 0)( =tσ& , , and 0)( =tλ& 0)( =tµ& , and we obtain the set of equations that

governs the amplitudes and phases of the envelope modulations,

µµ−ϕ−

ω+µ++

ω−

ω−ω

=µ−ω

+σ−ω

+ϕ++ω

−ω

−ω

=ϕ−ω

+λ−ω

+σ++ω

−ω

−ω

=σ−ω

+λ+ω

−ω

04

3)()(

0)()()(

0)()()(

0)()(

3

*2

1111

112212211

22222

nbbbcbcc

bbbbb

ikcikccikikmi

cikcikccikikmi

cikcikccikikmi

cikcikmi

(3-11)

To analyze (3-11) we express the complex variables in terms of their real and imaginary

components,

, σ , )()()( 21 tixtxt +=ϕ )()()( 21 tiytyt += )()()( 21 tiztzt +=λ , )()()( 21 titt ν+ν=µ (3-12)

Substituting (3-12) into (3-11), and separately setting equal to zero the real and imaginary parts,

we obtain eight real algebraic equations governing the eight new coordinates. These eight

equations are solved numerically in Mathematica.

From (3-5), (3-8) and (3-12), we recover the relationships between the complex amplitudes

and original coordinates; e. g.,

)sin))(cos()(()()()()( 1122 tittzitzettxitxtW ti ωω+ωω+=λ=ω+= ω&

Thus,

)sin(||1)(2 λθ+ωλω

= ttx , )cos(||)(2 λθ+ωλ= ttx& (3-13a)

Similarly,

24

)sin(||1)(1 σθ+ωσω

= ttx , )cos(||)(1 σθ+ωσ= ttx& (3-13b)

)sin(||1)( ϕθ+ωϕω

= ttxb )cos(||)( ϕθ+ωϕ= ttxb& (3-13c)

)sin(||1)( µθ+ωµω

= ttxc , )cos(||)( µθ+ωµ= ttxc& (3-13d)

where amplitudes are given by 22

21| zz +=λ| , and phase angles by

1

21tanzz−

λ =θ , etc.

This completes the theoretical analysis.

3.3 Numerical Results for Undamped NNMs

3.3.1 Physical Parameters for Estimation

We need to choose parameters of the system for numerical solution of equations (3-11). The

values selected are given in Table 3.1. These are consistent with the tuning criteria for the

isolation system outlined above. Damping has been neglected at this stage.

Table 3.1: Physical parameters of the system

Superstructure Base Intermediate System

m1, m2 0.2 kg mb 0.3 kg mc 0.3 kg

k1, k2 3500 N/m kb 116 N/m kc 4242 N/m

kn 2×106kc N/m3

To replace the cubic nonlinearity with a bilinear clearance element, we chose the stiffness

coefficient of the clearance spring k cn k10= and the clearance 0009.0=e m. The resulting

smooth nonlinear restoring force, F , and its bilinear counterpart, are shown in

Figure 3.2.

3ukn)( uku c +=

25

-0.002 -0.001 0.001 0.002u

-75

-50

-25

25

50

75FHuL

Figure 3.2: Nonlinear restoring force

3.3.2 Solutions of the Algebraic Equations

Of the eight algebraic equations, six are linear and two are cubic. Using Guassian

elimination, the eight equations are reduced to two nonlinear equations

0)])((1[ 22

211 =+ω+ zzfz

(3-14) 0)])((1[ 22

212 =+ω+ zzfz

where is a function of . )(ωf ω

It can be seen from (3-14) that, for each ω ,

)(

122

21 ω

−=+f

zz (3-15)

where 22

21 zz + is the slow oscillation amplitude of the 2nd story. Other complex variables and,

hence, amplitudes in (3-12) can all be deduced linearly from z and . Thus, this is the only

amplitude solution for this problem. Equation (3-15) indicates that z and can be purely

imaginary for some values of ω . This is acceptable, as inspection reveals that, at these

frequencies, purely imaginary solutions are equivalent to purely real solutions, and vice versa.

1 2z

1 2z

From (3-15) we see that, though 22

21 zz + is a constant for each ω , and can be

arbitrary. For this problem, it is sufficient, then, to consider only the case z , since the

remaining solutions can be found by rotating this one by

1z

1

2z

0=

)|)(|

1arctan(ω

±f

in the plane. )2z,( 1z

26

3.3.3 Figures and Discussion

Figure 3.3 depicts the amplitude versus frequency plots for the system with the cubic

nonlinearity. Figure 3.4 presents the enlargement of the plots in Figure 3.3 near the internal

resonance point. The straight vertical lines represent the linear bending modes ( 0=nk ), at

frequencies rad/s, rad/s, and 415.119=ω 636.121=ω 296.223=ω rad/s. The dotted line is the

1st bending mode of the primary structure (the mode to which the subfoundation is tuned), where

. This coincides with the frequency of the tuning algorithm discussed in Chapter 2.

The mode shapes of the linear system (without the stiffness nonlinearity) are independent of

frequency, and can be found only to within a multiplicative constant at their respective natural

frequencies.

523.120=ω

The curves of Figures 3.3 and 3.4 represent the nonlinear normal modes of the undamped

system with the local nonlinearity, i.e., the free, synchronous periodic solutions of the unforced

system (Vakakis et al., 1996). It can be seen that NNM shapes (i.e., the relative displacements of

the masses of the system) are functions of frequency. In the region near a tuning frequency, one

notes that both stories of the main structure, the base and the subfoundation, undergo resonance-

type behavior (they possess sharp local maxima versus frequency, though, since no external

harmonic excitation exists, one cannot technically characterize this local behavior as ‘resonance’),

with the two stories and the base vibrating at relatively large amplitudes compared to the

subfoundation. It is concluded, therefore, that near the tuning region ‘inverse localization’ of the

free vibration to the main structure and away from the foundation occurs. In fact, the point where

the amplitude of story 1 tends to zero can be used to distinguish between the different branches of

NNMs of the system, which in the plots are labeled from 1 to 4. In this system due to the localized

nature of the nonlinearity there exist four NNMs (equal in number to the modes of the

corresponding linear structure), and they can be considered as analytic continuations for the

nonlinear case of the linear modes of the system with no stiffness nonlinearity; in other words, no

NNM bifurcations were detected in the system examined.

Outside of the tuning region, however, NNM 4 becomes localized at the subfoundation for

increasing frequency. This conclusion is drawn by the increase in the amplitude of the

subfoundation with simultaneous decrease of the amplitudes of the two stories. It is this high-

frequency nonlinear localization of NNM 4 to the subfoundation that gives rise to the motion

confinement phenomena discussed in the next sections.

27

-10

-8

-6

-4

-2

0

2

4

0 1000 2000 3000 4000 5000

ω

Loga

rithm

ic a

mpl

itude

s of

slo

wly

va

ryin

g en

velo

pes

s

3 NNM 4

-5

-4

-3

-2

-1

0

1

60

Loga

rithm

ic a

mpl

itude

s of

slo

wly

var

ying

en

velo

pes

-5

-4

-3

-2

-1

0

1

116

Loga

rithm

ic a

mpl

itude

s of

slo

wly

var

ying

enve

lope

s

NNMs 1,2,

110 160 210 260 310 360 410

NNM 2

NNM 4

s NNM 1

F

NNM 3

ω

121 126 131 136

ω

s

NNM 4

NNM 1

S

igure 3

NNM 2

tory 2 Story 1 Ba

.3: Amplitudes of slow

NNM 3

se Subfoundation

-frequency oscillation

rad/

rad/

rad/

28

-3

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

116 118 120 122 124 126

log(

| λ|)

-4

-3.5

-3

-2.5

-2

-1.5

-1

-0.5

0116 118 120 122 124 126

log(

| σ|)

s

ω ωNNM 1 NNM 2 NNM 3 NNM 1 NNM 2 NNM

(a) Amplitude of the 2nd story (b) Amplitude of the 1st story

-3

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

116 118 120 122 124 126

ω

log(

| ϕ|)

-3

-2.5

-2

-1.5

-1

-0.5

0116 118 120 122 124

ω

log(

| µ|)

s

NNMNNM 2 NNM 1NNM 3NNM 2 NNM 1

(c) Amplitude of the base (d) Amplitude of the subfoundation

Figure 3.4: Amplitudes of slow frequency oscillation near the internal resonan

A last remark is made regarding the limit of the frequency-amplitude plots as the

the nonlinearity tends to zero. In that limit the system becomes linear and all curves de

four vertical lines emanating from the three linearized natural frequencies; these repres

(nonlocalized) modes of the linear structure. Hence, as the strength of the nonlinear

zero, one can imagine the various curves in Figure 3.3 and 3.4 ‘dipping’ to zero amp

the resonance-type peaks degenerating to vertical lines (in the plots only three

linearized natural frequency peaks are depicted, since the lowest linearized natura

occurs at below the range of the plots).

In Section 2.3.2, we noted that the maximum amplitudes of the various masses d

free vibration are determined by components of the mode shapes. In the vicinity of the

spaced natural frequencies, the ratios of the maximum amplitudes of the three sto

primary structure to that of the subfoundation are constants, since these two modes d

rad/s

rad/

3

126

s

c

e

i

o

l

u

t

o

rad/

rad/

3

e

strength of

generate to

nt the four

ty tends to

litudes and

f the four

frequency

ring linear

wo closely

ries of the

minate the

29

free response and the linear mode shapes are invariant. Thus, the amplitude ratios are straight

lines in configuration space. In Figure 3.5, we plot both linear and nonlinear modal amplitudes of

the three stories of the primary structure versus that of the subfoundation for comparison

purposes.

0

0.003

0.006

0.009

0.012

0.015

0 0.004 0.008 0.012 0.016 0.02 0.024 0.028 0.032

| µ |

| σ |

0

0.01

0.02

0.03

0.04

0.05

0 0.004 0.008 0.012 0.016 0.02 0.024 0.028 0.032

| µ || λ

|

(a) 1st story vs. the subfoundation (b) 2nd story vs. the subfoundation

0

0.01

0.02

0.03

0.04

0.05

0 0.004 0.008 0.012 0.016 0.02 0.024 0.028 0.032

| µ |

| φ |

(c) The base vs. the subfoundation

Figure 3.5: Modal amplitudes of the primary structure vs. the subfoundation

The straight lines in Figure 3.5 represent the linear normal modes. Curves above the straight

lines are nonlinear normal modes near the resonant frequency, while curves below the straight

lines are outside of the resonant frequency range. Figure 3.5 shows that the nonlinear normal

modes in this system are curves instead of straight lines. In the terminology of Rosenberg (1960)

these are nonsimilar NNMs.

30

Chapter 4: Model Building for System Analysis in Simulink

4.1 Introduction

From the analysis in Chapter 2, we see that, even for an idealized linear 2DOF system

(Figure 2.1) subjected to simple harmonic excitation, the analytical solution can be somewhat

complicated. In view of the fact that we wish to explore relatively large-scale MDOF nonlinear

structures, finding solutions analytically is not practical. Thus, we resort to available analysis

tools, such as Matlab and Simulink, to perform numerical calculations. In this chapter we will

discuss how we built simulation models corresponding to the system of Figure 3.1 in Simulink.

Simulink is a sophisticated graphical user interface (GUI) that complements Matlab for with

modeling, simulating, and analyzing dynamical systems. It supports linear and nonlinear systems,

modeling in continuous or discrete time or a hybrid of the two. Its modeling environment uses

familiar block diagrams, so systems illustrated in texts can be easily implemented in Simulink

(Dabney and Harman, 1998).

4.2 Elements of a Model

Like a usual system analysis flow chart, a Simulink model consists of three types of

elements: sources, the system being modeled, and sinks. Figure 4.1 illustrates the relationship

among the three elements. The central element, the system, is the Simulink representation of a

block diagram of the dynamical system being modeled. The sources are the inputs to the

dynamical system. Sources include constants, function generators such as sine waves and step

functions, and signals created in Matlab. The output of the system is received by sinks. Examples

of sinks are graphs, oscilloscopes, and output files. The simulation results can be put into the

Matlab workspace for post-processing and visualization.

System Block

Diagram Sinks Sources

Figure 4.1: Elements of a Simulink model

31

Simulink includes a comprehensive library of sinks, sources, linear and nonlinear

components, and connectors. One can also customize and create blocks. After a model is defined,

one can simulate it, using a choice of solvers, and then execute it either from the Simulink menus

or by entering commands in the Matlab command window. In our Simulink code, the isolation

system is represented by a scalar model while the two-story building is described by a vector

(state-space) model. These will be discussed in the next section.

4.3 Scalar System: the Nonlinear Isolation System

Consider the nonlinear oscillator of the proposed isolation system, depicted in Figure 4.2.

mb ekn

yb(t)+xg(t) yc(t)+xg(t) xg(t)

kbkc

mc

Figure 4.2: The scalar nonlinear system

Let y represent displacement with respect to the ground and z, displacement with respect to the

base mass mb. Then, with subscripts g and c representing ground and subfoundation, respectively,

gcg xyzxyx ++=+= (4-1)

Referring to equations (3-1), the equations of motion of the subfoundation mc and base mb may be

written as

gc

cg

c

ccbcbbcbcccc

cc x

mk

xmc

yFxxkxxcxkxcm

x ++−−−−−−−= &&&&&& )]()()([1 (4-2)

0)]()()()([1111 =−+−+−+−−= cbcbbbcbb

bb xxkxxkxxcxxc

mx &&&&&& (4-3)

The restoring force in (4-2) provided by the nonlinear spring-gap element is given by

32

− (4-4)

++

−+=

)(

)()(

eukukuk

eukukuF

nc

c

nc

eueue

eu

−<≤≤

>

The equations have been rewritten so that the system inputs are ground velocity and displacement

instead of commonly used ground acceleration, as will be explained in detail in Chapter 5. Figure

4.3 shows the Simulink model of the subfoundation.

System Block Diagram Sinks Sources

Figure 4.3: The Simulink model of the subfoundation

Oval blocks represent inputs and outputs, rectangular blocks (with + and - signs),

summation, and triangular blocks, gains. Scalar input signals to the subfoundation mc include

ground velocity and displacement, linear spring restoring forces and damping forces produced by

their relative motion to base and to ground, and the nonlinear restoring force. The outputs of the

system are scalar signals of accelerations, velocities and displacements and are displayed by

oscilloscopes. The central diagram is where the dynamic differential equation is solved. The

velocity and the displacement are obtained by numerically integrating the acceleration. We set the

initial conditions to zero, because the system is at rest before the ground moves. The nonlinear

restoring force is modeled by a dead-zone block and a function block.

The Simulink model of the base can be built in the same way as that of the subfoundation,

except that it is linear and its input forces result from its motion relative to the subfoundation and

the first story mass.

33

4.4 Vector (State-Space) Linear System: the Superstructure

For linear MDOF systems, it is often convenient to use vector signals, as they provide a more

compact and easier-to-understand model. The state-space approach is particularly useful for

modeling linear systems, because we can take advantage of matrix notation to describe very

complex systems in a compact form. Additionally, we can compute the system response using

matrix arithmetic. Before discussing model building using vectors, let’s discuss the concept of

state variables and describe the state-space block in Simulink first.

The general form of the state-space model of a dynamical system is

x ),,( tuxf=& (4-5)

where is the state vector, u the input vector, and t time. Equation (4-5) is called the system

state equation. We also define the system output equation to be

x

y ),,( tuxg= (4-6)

In linear spring-mass-damper systems, the system state equation and the output equation can

be written as linear combinations of the system states and inputs as

x BuAx +=& (4-7) y DuCx += (4-8) where matrix A is called the system matrix, B the input matrix, C the output matrix, and D the

direct transmittance matrix. Equations (4-7) and (4-8) comprise the state-space block in Simulink.

To use state-space block to model the linear superstructure, we first rewrite the equations of

motion of the two-story superstructure with the base velocity and the base displacement as input

0)()( 12212222 =−+−+ xxkxxcx &&&&m (4-9)

m bb xcxkxxkxkxxcxcx &&&&&& 11212112121111 )()( +=−++−++ (4-10)

The matrix equation representing the dynamics of the two-story superstructure is

m )()()()( tttt uDkxxcx 1−=++ &&& (4-11)

34

where

m , c , , (4-12)

=

2

1

00

mm

−+=

22

221

ccccc

−+=

22

221

kkkkk

k

x , , (4-13)

=)()(

)(2

1

txtx

t

=

0011

1

ckD

=)()(

)(txtx

tb

b

&u

The state-space equation of motion for the superstructure of this system is

(4-14) )()()(

)()( 1 t

tt

tt

u0

Dxx

m00k

xx

0mmc

=

+

&&&

&

Note that the boldface zeros in (4-14) represent 22× null matrices.

Define the state vector ξ , and so ξ . It follows that the system state

equation is

=)()(

)(tt

txx&

=)()(

)(tt

txx&&

&&

ξ (4-15) BuAξ +=&

where

A , B (4-16) 211AA −= 1

11BA −=

A , A , (4-17)

=

0mmc

1

=m00k

2

=0

DB 1

1

If the output variables are the system state variables, then ξy = , and the output equation is

ξ DuCξ += (4-18)

where

C , D (4-19) I=

=

00000000

This is shown in Figure 4.4, block “StateSpace1”.

35

If the velocity and the acceleration are output variables, then ξ becomes the

output vector, such that y . Referring to equation (4-15) where ξ , we see that for

this case C and . This is shown in Figure 4.4, block “StateSpace2”.

=)()(

)(tt

txx&&

&&

BuAξ +=ξ&=

B=

&

A= D

System BlockDiagram

Sources Sinks

Figure 4.4: The Simulink model of the linear superstructure

The two input scalar signals (base displacement and base velocity) are combined to form a vector

signal by using the multiplexer (Mux) block, while the demultiplexer blocks, Demux1 and

Demux2, split vector signals into sets of scalar signals.

The full nonlinear system is then the combination of the isolation system and the

superstructure. Refer to Appendix A to see details of the programming.

36

Chapter 5: Ground Motions Considered In This Study

5.1 Introduction

The dynamic response of a structure depends on its mechanical characteristics and the nature

of the excitation. A protective system that efficiently controls the response of the structure when

subjected to certain classes of inputs may not be efficient when subjected to other classes. For

example, long-period pulses contained in near field earthquake records decrease the performance

of passive viscous dampers (Malhotra, 1999). Hence, the usual method of using randomly chosen

ground motions or assuming simple filtered white noise excitation models may not lead to an

effective protective system.

Previous studies have demonstrated that near field earthquakes contain coherent long-period

pulses with some overriding high-frequency fluctuations (Anderson et al., 1986, Iwan and Chen,

1994). These long-period pulses are sometimes distinguishable even in the acceleration time

history. Peak ground acceleration (PGA) is the most commonly used measure of earthquake

potential. However, studies indicated that what makes near field ground motions particularly

destructive to flexible structures is not their PGA but their long-duration pulse, which represents

the incremental velocity that the above-ground mass has to reach (Anderson and Bertero, 1986,

Hall et al., 1995, Iwan, 1997). The peak ground displacement (PGD) is not a reliable measure

either, since the displacement history is the second integral of the acceleration history, and the

long-period components may be filtered out (He, 2003). The peak ground velocity (PGV) seems

to be a better representative measure of earthquake destructiveness as it represents the cumulative

effect of the seismic energy radiating from the fault for near field ground motions (Somervill and

Graves, 1993).

These long-period pulses present a challenge to the concept of base isolation for earthquake

protection. Although recent design codes have incorporated near field effects in the design

spectrum, the design procedure is the same as that for ordinary ground motions. Furthermore, in

addition to earthquakes, structures will experience a number of transient excitations throughout

their lives, some of which may also exhibit these characteristics. Hence, it is imperative to study

the kinematic characteristics of near field effects from available records for shock-protective

systems.

Makris (1997) and Makris and Chang (1998) proposed closed-form velocity-pulse models

for near field ground motions. Based on these studies, He (2003) proposed an analytical model for

velocity pulses and systematically studied the properties of the proposed pulse model. It was

37

verified from response spectra and curve-fitting analyses that this model could capture the major

characteristics of near field earthquakes. Thus, in this study we use velocity pulses and velocity

time histories other than the usual accelerograms as the inputs to account for near-field effects.

We will introduce these models briefly, and then explain which of them will be employed in this

study.

5.2 Makris and Chang’s Model

Makris and Chang (1998) grouped velocity pulses present in near field ground motions into

three types of pulses: the type A pulse, the type B pulse, and type Cn pulses, represented by a

unique set of tri-geometric functions. Type A is a forward-only pulse; type B is a full sine wave

pulse that approximates a forward-and-back motion; and type Cn pulses approximate an n-cycle

ground displacement. The formulations of these velocity pulse types are given by

Type A: )cos1(2

)( tv

tx pp

p ω−=& , pTt ≤≤0 (5-1)

Type B: tvtx ppp ω= sin)(& , pTt ≤≤0 (5-2)

Type Cn: ϕ−ϕ+ω= sin)sin()( pppp vtvtx& , pTnt )21(

πϕ

−+≤≤0 (5-3)

where subscript p denotes “pulse”.

There is a permanent ground displacement at the end of the Type A pulse, while the ground

displacement of type B fully recovers to its initial value. In order to have a zero ground

displacement at the end of a type Cn pulse, the condition

(5-4) 0)(0

=∫ dttxT

g&

must be satisfied. Here, T is the duration of a type Cn pulse

ppp TnTnT )21(2)21( πϕ−+=ωϕ−+= (5-5)

with pulse frequency pp Tπ=ω 2 . Substitution of (5-3) and evaluation of the integral (5-4) gives

38

cos[( 0cossin])12[(])12 =ϕ−ϕϕ−π++ϕ−π+ nn (5-6)

Solution of the transcendental equation (5-6) gives the value of the required phase angle ϕ .

For example, for a type C1 pulse, π=ϕ 0697.0 ; whereas, for a type C2 pulse, π= 0410.0ϕ . Figure

5.1 shows the velocity shapes of type A, B, C1 and C2 pulses, where the values 0.1=pT and

were used. 0.1=pv

Figure 5.1: Ground velocity pulse types proposed by Makris and Chang (1998)

To model a particular ground motion, only one function can be used. However, it has been

observed that actual ground motion pulses are combinations of A, B and/or Cn pulses. As a result,

there are significant disagreements between recorded and modeled ground motion pulses. These

simplified approximations involve only parameters of the pulse period and peak amplitude.

However, ground motion energy is a building up and decaying process. The model proposed by

He and Agrawal accounts for these characteristics of the ground motions.

5.3 He’s Model

He’s model is expressed as

; bteCttx atnp sin)( −=& ppa ως−= ; 21 ppb ς−ω= (5-7)

39

Equation (5-7) presumes that the velocity pulse is an enveloped sinusoidal function. The

parameters a and n determine the shape of the pulse envelope, pς is the decay factor, pω is the

frequency, and C is the amplitude constant of the pulse. Figure 5.2 shows the versatility of the

proposed ground velocity pulses generated by different values of parameters n, a and b. It is seen

that the model is capable of generating various types of pulses reminiscent of those observed in

near field earthquake ground motions.

Figure 5.2: Ground velocity pulse models proposed by He (2003)

The ground acceleration is obtained by differentiating equation (5-7) and is given by

]cossin)[()( btbbtatneCttx atn

p +−= −&& (5-8)

While the ground displacement can be obtained by integrating equation (5-7), its explicit form is

quite complicated despite n being an integer.

For the special case of n , the velocity pulse (5-7) is a decaying sine-wave, and the

corresponding ground displacement is

0=

22

)cossin()(ba

bbtbbtaeCtxat

p +++−

=−

(5-9)

40

The velocity pulse in equation (5-7) can be defined in terms of the vector of parameters

as ],,,[ banC=Θ

),()( Θtxtx pp && = (5-10)

To obtain parameters for a particular ground motion, the start time of the pulse, t0, should

also be included in the vector of parameters, i.e., ],,,,[ 0tbanC=Θ . In his thesis, He (2003) also

gave a method to find the parameter vector Θ that best fits a recorded ground velocity . This

is achieved by minimizing the sum of squares of the difference between x and at

discrete time instants t

)(txg&

,( ip tx&)( ig t& )Θ

i,

S , i = 1, 2, 3,…, N (5-11) ∑=

−=N

iipig txtx

1

2)],()([)( ΘΘ &&

where the ti are the discrete times at which the time history is recorded.. )( ig tx&

The minimization of S was done by Newton’s method in this analysis. Using the

minimization procedure, He (2003) obtained optimal parameter vectors for 36 recorded ground

motions from around the world.

)(Θ

5.4 Ground Motions Employed in This Thesis

5.4.1 Full Sine Wave Ground Velocity

The first input we will examine is a ground velocity shock pulse, which imparts a large

amount of energy to the structure in a short time. From Makris and Chang, we adopt the type B

model, i.e., the forward-and-back sine-wave, because it is simple, and the final displacement will

be zero at the end of the ground motion. This type of velocity pulse, and its corresponding

acceleration and displacement, can be expressed as

tT

vtxp

ppπ

=2sin)(& , pTt ≤≤0 (5-12)

tT

vT

txp

pp

pππ

=2cos2)(&& , pTt ≤≤0 (5-13)

41

)2cos1(2

)( tT

vTtx

p

ppp

π−

π= , pTt ≤≤0 (5-14)

Figure 5.3 shows the acceleration, velocity and displacement time histories of the pulse. The

Matlab code for this velocity pulse is given in Appendix C.

Figure 5.3: Acceleration, velocity and displacement time histories

of a Type B pulse with 0.5=pT , 0.1=pv

5.4.2 Historic Earthquake Record

In order to substantiate the results of performance of our proposed nonlinear protective

system, a dynamic analysis of a system subjected to a historic earthquake records was performed.

The record used is the NS component of the Erzincan, Turkey earthquake, which occurred

03/22/92. It was selected because its velocity and displacement time histories both have long-

period pulses. The long-period pulse is also distinguishable in the acceleration time history. It is

an example of a typical near field earthquake and was recommended to the writer by Professor

Wilfred Iwan of the California Institute of Technology (Iwan, 2002). The record employed was

adjusted for base-line position, and the velocity was assumed to have a linear variation between

two consecutive time points. Zero ground velocities were added at the end of the record to

account for free vibration response in the maximum response calculation. The duration of this

zero amplitude ground velocity was arbitrarily chosen.

42

5.4.3 Approximation by He’s Method

We chose to use a function to simulate ground motions containing near field effects in our

analyses. As He (2003) has already provided optimal model parameters for 36 historic

earthquakes around the world, we employed his results for the Erzincan earthquake record in (5-7)

C cm/s, , 57.70−= 00.5=n 80.1=a , 75.2=b , 6.00 =t s (5-15)

Figure 5.4: Erzincan earthquake time histories and their approximations

The recorded and approximate earthquake ground motions for the first 15 seconds are plotted

in Figure 5.4, where it is apparent that the simulated closed-form velocity time history agrees well

with that of the actual earthquake record. While the displacement time histories are also

comparable, the corresponding acceleration time histories differ, verifying the conjecture that the

approximate velocity can only represent the dominant low frequency component, resulting in

underestimation of peak ground acceleration (He and Agrawal, 2002).

To examine how well the approximated pulse model captures the characteristics of the actual

earthquake, numerical simulations were conducted to compare the dynamic responses of a SDOF

oscillator subjected to the recorded ground motions and those of the He pulse model. Consider a

SDOF oscillator subjected to a ground acceleration . Its equation of motion is )(txg&&

(5-16) )()()(2)( 2 txtytyty g&&&&& =ω+ςω+

43

where is viscous damping ratio, ς ω is the natural frequency of the oscillator, and y is the

relative displacement. The validity of the analytical model is verified by comparing the response

spectra of the two ground motions. The response spectrum is a plot of a maximum response of a

SDOF oscillator versus its natural period or frequency. The acceleration, velocity and

displacement response spectra for both the earthquake and its analytical model, for 0.5% and 2%

damping, are plotted in Figure 5.5. Matlab code and Simulink models for calculating response

spectra are given in Appendice E and F, respectively.

)(t

Figure 5.5: Erzincan earthquake response spectra

We see that the spectra of the pulse model match well with those of the recorded ground

motion except for the acceleration spectra at short periods. This is because the analytical model

excludes the high frequency (low period) components of the recorded ground motion. It is known

that the response of long-period structures is more sensitive to the low-frequency components in

the ground motions, while the high-frequency components affect the acceleration response of

short-period structures. Thus, the proposed pulse model may be a good approximation with which

to investigate the response of flexible structures subjected to ground motions with near field

effects.

44

Chapter 6: Numerical Simulation of the Undamped System

In Chapter 3, the nonlinear restoring force-displacement relation employed was given by

, where kN/m, and kN/m3)( ukukuF nc += 4228=ck cn kk 6102×= 3. Sometimes, a higher odd-

ordered nonlinearity can be needed to achieve the desired performance. However, increasing the

order of the stiffness nonlinearity raises implementation issues. Rather, we replaced the third

order smooth nonlinearity by a non-smooth clearance nonlinearity and then examined numerically

the dynamics of the system subjected to three different ground motions: the full sine wave ground

velocity pulse, the ground velocity modeled by He’s formulation, and a scaled near-field historic

earthquake record. Some typical results from the numerical simulations are presented. The

effectiveness of the nonlinear isolation system is demonstrated through shock spectra, modal

amplitudes, and system responses in both the time and frequency domains. Responses of the

linear and nonlinear systems are plotted together for comparison. We used Matlab and Simulink

to perform the simulations. The solver employed was ode4 (4th order Runge-Kutta), and the

integration time step was 0.00025 seconds. Appendix B gives the Matlab code of the Simulation.

6.1 Approximation of the Cubic Nonlinearity by a Clearance Nonlinearity

Figure 6.1: Restoring force-displacement relationships

45

To replace the cubic nonlinearity with a bilinear clearance element, we chose the stiffness

coefficient of the clearance spring k cn k10= and the clearance 0009.0=e m. For relative

displacement of the subfoundation with respect to ground less than or equal to 0.002 m, the force-

displacement relationships for the linear, cubic nonlinear, and bilinear clearance springs are

plotted in Figure 6.1. It can be seen that the bilinear clearance nonlinearity and the cubic

nonlinearity are very close for the selected parameters. For the same parameters used in Chapter

3 and the newly specified kn, the natural frequencies and periods of the system and its component

parts are shown in Table 6.1.

Table 6.1: Natural frequencies and periods of the system

Superstructure Superstructure +Base

Linear system (kn = 0)

Linear system (e = 0)

2.259 2.203 2.229 18.9 19.21 13.01 19.21 19.42 35.54

f ( Hz )

34.07 35.54 35.54 62.77 0.4427 0.4539 0.4489

0.0527 0.0521 0.0769 0.0521 0.0515 0.0304

T ( s )

0.0294 0.0281 0.0281 0.0159

1188..99999 1199..22111199..4422 3355..5544

00..00552277 00..0055221100..00551155 00..00330044

Here, we see that the addition of the base and weak spring results in a near rigid-body mode

at 2.259 Hz in the conventionally isolated structure. This becomes 2.203 Hz in the combined

structure incorporating the subfoundation. The first superstructure natural frequency of 19.21 Hz

in the conventionally isolated structure is significantly higher than the lowest fixed-base natural

frequency of 13.01 Hz, demonstrating the effect of a standard isolation system on the

deformational natural frequencies of the structure. The introduction of the subfoundation and

tuning results in two very close natural frequencies at 18.99 Hz and 19.42 Hz, which provide the

internal resonance in the linear system required for localization.

6.2 Response to Full Sine-Wave Ground Velocity Pulses

As described in Chapter 5, we first use the Makris and Chang Type B pulse model (i.e., a full

sine-wave pulse) to examine the dynamics of the proposed system.

tvtx ppp ω= sin)(& , pTt ≤≤0 (5-2)

46

The amplitude of the velocity shock pulse used is pg gv ω= 1 ; thus the corresponding amplitude

of acceleration is 1g.

6.2.1 Shock Spectra

In order to compare the performance of a specified system under the action of different

transient inputs, we plotted the maximum value of a various response quantities versus the

transient pulse duration T . This plot is known as the shock spectrum in the general vibration

literature and the response spectrum in the earthquake engineering literature. The shock spectra

for the nonlinearly isolated building and for the same system with k (i.e., a fully linear

isolation system incorporating the subfoundation) are shown in Figure 6.2. The clearance e used

here is always one-fourth of the peak value of from the linear case.

g

0=n

)(tyc

Figure 6.2: Shock spectra of various responses to a full sine ground velocity pulse

It can be seen that the motion of the subfoundation is limited because of the clearance

nonlinearity and that the nonlinear isolation system suppresses the response in the range of the

two closely coupled linear normal modes (when the pulse width is small) and subharmonic

resonances (when the pulse width is large). In the following, we employ a high frequency (small

period) input pulse to investigate the difference in performance between the linear and nonlinear

47

isolation systems. Hence, T is chosen to be 0.052 s, about the second natural period of the

primary structure.

g

6.2.2 Response in the Time Domain

To study the underlying nonlinear behavior, it is necessary to solve for the modal amplitudes

of the primary structure. The modal amplitudes in terms of the displacements of the building and

the base with respect to the subfoundation are expressed

A (6-1) ∑=

=3

1)()(

i

Ti tt zϕ

Here, are modal amplitudes of the three masses in the primary structure at moment t,

, the transpose of the ith normalized mode of the primary structure with , and z ,

the displacement vector at moment t. is obtained from Matlab simulation results. The

clearance we used here is 0.027 m, which is one-tenth of the peak value of from the linear

case.

)(tATiϕ 0)( =tyc

)(tyc

)(t

)(tz

Figure 6.3:(a) Beat phenomenon in the linear system;(b)mode localization in the nonlinear system

48

Note that, in the linear case, the system was designed such that the natural frequency of the

uncoupled subfoundation is equal to one of the natural frequencies of the primary structure. As a

result, when weak coupling between the two subsystems is introduced, we expect a beat

phenomenon to occur, whereby energy from one subsystem transfers to the other and back again.

This is shown in Figure 6.3(a), in which the second and third modal coordinates of the primary

structure are plotted together.

Now, consider the corresponding nonlinear system under the action of the same ground

velocity pulse. Previous studies tell us that the nonlinear system with internal resonance and weak

coupling between the subsystems has a localized NNM. As a result, energy is mainly confined to

the secondary subsystem, and a relatively small portion of this energy ‘leaks’ to the superstructure

(Vakakis et al, 1999). This is presented in Figure 6.3(b). Note the localization effect, the very

small 2nd modal amplitudes that occur in the motion of the primary structure. A2 has been reduced

by 2 orders of magnitude in the nonlinear case. Where did the energy go? It will be shown later

that energy mainly goes to the higher nonlinear modes of the structure. Modal amplitudes of

higher modes are usually very small, as A3 in Figure 6.3.

Figure 6.4: Time history responses to the full sine-wave velocity pulse

Linear Nonlinear

We can examine the dynamics in the displacement and acceleration response time histories.

These are shown in Figure 6.4. Here, the peak responses of the primary structure and linear

subfoundation are out of phase, indicating energy transfer between the two subsystems.

Examination of the nonlinear responses shows that both the inter-story drifts (left column) and the

49

story accelerations (right column) of the building and the base are drastically reduced, while the

motion of the subfoundation is confined in the dead zone, and the subfoundation mass vibrates

strongly at a higher frequency than in the linear case. This reflects the localization of motion to

the subfoundation, as expected.

6.2.3 Response in the Frequency Domain

It is instructive to examine the responses of both the linear and nonlinear systems in order to

determine how energy is redistributed in the latter. For this, we study the power spectra of the

responses as a function of the frequency. The power spectrum of an ergodic time series is the

square of the modulus of the formal Fourier transform of its covariance function.

The differences in the responses noted above are similarly distinctive in the frequency domain.

The PSDs of the time histories of Figure 6.4 are shown in Figure 6.5. The most prominent feature

of these PSDs is the frequency shifting from the tuned mode, i.e., the second mode, to a higher,

nonlinear mode.

Figure 6.5: Power spectra of the system subjected to the full sine-wave velocity pulse

Referring to the nonlinear normal mode (NNM) results of Chapter 3, the high-frequency localized

peak of the subfoundation response is due to excitation by the localized NNM 4, which as we

showed is partially localized in the subfoundation. No such localization can exist in the linear

50

system since for that system all linear normal modes are spatially extended. Hence, we are able to

prove that the enhanced isolation properties of the nonlinear design can be solely attributed to the

excitation by the external shock of the high-frequency localized NNM of the combined system.

6.3 Response to Scaled Histories of Earthquake Records

The simulation above demonstrates the principle of using nonlinear localization for isolation

of a system subjected to a full sine velocity pulse input. However, the single pulse does not

accurately model a general near-field earthquake. In order to further substantiate this concept, we

extend the simulation to a historic near field earthquake input. The following are results of

simulations employing the previously discussed Erzincan records. To examine the isolation

efficiency of our proposed system, which has a second natural period of 0.052 s, we scaled the

ground velocity history using

acturalgVscaledg xSFx && = (6-2)

t acturaltscaled tSF= (6-3)

with and , resulting in a reduction in record length to 2 seconds from the

original 40 seconds. SF and are scale factors for amplitude and time step, respectively. The

factor is chosen based upon the dominant period of the earthquake velocity record, which is

1.91 seconds, approximately 37 times the tuning period of 0.052 seconds. SF scaling

corresponds to 0.096-second period earthquake, approximately twice the tuning frequency, which

is rather typical in full scale. Simulation results are presented in the following sections.

0.1=vSF

tSF

05.0=tSF

v tSF

05.0=t

6.3.1 Response in the Time Domain

Examination of the modal responses shown in Figure 6.6 reveals that the tuned mode nearly

disappears in the nonlinear system (the modal amplitude decreased by 25 times), while the third

mode experiences a minor increase. Meanwhile, the energy exchange phenomenon is again seen

in the linear system responses: beating envelopes of the primary structure and subfoundation are

complementary to each other. The time histories (Figure 6.7) further illustrate the outstanding

performance of the nonlinear system: the building barely moves compared with the response of

51

the linear system; and the accelerations of the building and base decrease significantly, while the

acceleration of the subfoundation increases accordingly.

Figure 6.6:(a) Beat phenomenon in the linear system;(b)mode localization in the nonlinear system

Figure 6.7: Time history responses to the scaled earthquake velocity

Linear Nonlinear

52

6.3.2 Response in the Frequency Domain

Figure 6.8: Power spectra of the system subjected to the scaled Erzincan earthquake

The time domain results are confirmed in the frequency domain. From Figure 6.8, we see

that the second mode is completely suppressed, with energy moved to a higher mode at nearly 50

Hz.

6.4 Response to the Analytical (He) Model of Ground Velocity

A systematic investigation of seismic isolation systems applied to reasonably complex

structures subjected to ground motions cannot be accomplished with a limited number of historic

records. It is, therefore, necessary to develop analytical representations of ground motion for use

in parameter studies or Monte Carlo simulation. Here we examine the performance of our

proposed nonlinear isolation system when the structure is subjected to an analytical representation

of the ground velocity determined from the Erzincan earthquake, as described earlier. It was

shown that the analytical formulation (He, 2003) can capture the major characteristics of the

earthquake record, except for high frequency components, which results in underestimation of

acceleration response at high frequencies. As shown previously, the Erzincan earthquake velocity

record is optimally represented by

53

(cm/s) (6-4) tettx tp 75.2sin57.70)( 8.15 −−=&

Because the earthquake is scaled to accommodate the tuned frequency of the structure, the

time series obtained from equation (6-4) is similarly scaled.

6.4.1 Response in the Time Domain

Examination of the modal amplitudes of the linear system in Figure 6.9 reveals the beat phenomenon seen earlier employing the full sine velocity pulse and the historic earthquake. The second modal amplitude of the nonlinear system is now one-seventh that of the linear system. Localization in the nonlinear system is apparent from the time histories of the displacements and accelerations (Figure 6.10). The second mode is nearly completely suppressed by the nonlinear subsystem, while the subsystem itself responds strongly in higher modes, absorbing more energy than in the linear case.

Figure 6.9:(a) Beat phenomenon in the linear system;(b)mode localization in the nonlinear system

54

Figure 6.10: Time history responses to the analytical model of earthquake velocity

Linear Nonlinear

6.4.2 Response in the Frequency Domain

Results in the frequency domain, shown in Figure 6.11, are similar to those obtained

previously for the Erzincan earthquake.

Figure 6.11: Power spectral densities of the system subjected to the full sine-wave velocity pulse

55

Interestingly, the isolation system appears to perform better when subjected to the historic

earthquake record than for its decaying harmonic representation. This is likely attributable to the

fact that the actual earthquake, with its greater energy content, more readily drives the system into

the NNM and resulting localized response. However, the analytical model proves to be useful for

certain kinds of analyses.

56

Chapter 7: Numerical Simulation for the Damped System

7.1 Construction of the Viscous Damping Matrix

7.1.1 Classical Damping Matrix for the Superstructure

Damping in a structure is generally a function of its configuration, materials, construction,

working stresses, environment, boundary conditions, and so forth. The pragmatic analyst usually

presumes that the structure is classically damped, thus avoiding many complicating issues and

facilitating analysis by linear methods. The viscous damping matrix that results is diagonalizable

by the same transformation that uncouples the undamped linear equations of motion.

The norm in the analysis of civil engineering structures, particularly tall buildings, is to

specify the modal damping ratios for a number of modes. A procedure to determine the viscous

damping matrix from modal damping ratios can readily be derived from

ϕ (7-1) Cc =ϕT

where C is the diagonal damping matrix with the nth diagonal element equal to the generalized

modal damping

)2( nnnn M ως=C (7-2)

Here, ς is the nth modal damping ratio, M is the nth modal mass, and ω is the nth circular

natural frequency. Equation (7-1) can also be written as

n n n

c (7-3) 11)( −−= ϕϕ CT

If the nth modal damping ratio ς is not specified, the resulting c assumes a zero damping ratio in

the nth mode.

n

There is a problem associated with the use of the damping matrix given by equation (7-3).

We need story damping for the construction of the damping matrix of the combined (primary plus

isolation) system, as well as for the convenience of numerical simulation. In equations of motion

such as (3-1), the damping matrix is banded, reflecting the close-coupling of the structure; that is,

57

individual dampers only couple neighboring degrees of freedom. However, c given by equation

(7-3) can be a fully populated matrix with elements in the band not corresponding to story

damping. Thus, in what follows, we will employ stiffness-proportional damping of the form

C kβ= (7-4)

where k is stiffness matrix of the superstructure (see equation (4-12)).

Consider the modal matrix normalized such that . Then, ϕ Im =ϕϕT

ϕ (7-5) )diag()diag(2)diag( 2i

Tiii

T ωββωςc ==== ϕϕϕ kc

It follows that

2

ii

ωβ=ς (7-6)

Physically, this can be interpreted as inserting, in parallel with each spring, a dashpot with

damping proportional through β to the spring stiffness. Generally, the damping ratio is specified

for the fundamental mode defining β , and the remaining damping ratios are given by (7-6).

Clearly, this is overly conservative, as the modal damping ratios increase linearly with frequency,

which is not borne out by experimental observations. Still, it serves our purpose here given the

small number of modes involved in the computation. Story damping coefficients can then be

obtained from the matrix given by (7-4). Further damping assumptions and calculations are

discussed in the next section.

7.1.2 Damping in the Base and Subfoundation

The assumption of classical damping may not be appropriate for a base isolated system, even

if the structure itself is classically damped (Chopra, 1995). The nonclassical damping matrix

associated with the entire system is constructed by first evaluating the classical damping matrix

for the structure alone from modal damping ratios. The damping contributions of the isolation

system are then included to obtain the damping matrix of the complete system.

Refer to Figure 3.1. Suppose that linear viscous damping c connects the base mass and the

subfoundation with the lateral stiffness k . Two parameters, T and , are introduced to

characterize the isolation system

b

b b bς

58

b

b ωπ2

=T (7-7)

bbbb mmmc ςω++= )(2 21 (7-8)

where b

bb mmm

k++

=21

ω .

We can interpret T as the natural vibration period and ς as the damping ratio of the base-

structure with the assembly assumed to be a rigid body. With the physical parameters of the

system given in Table 3.1, it follows that

b b

4443.0=b

.01

T s. Compared with the period computed in

Table 7.1, where the nearly rigid-body period is 4427=nT s, we see that the fundamental period

of the base-structure is changed only slightly by the flexibility of the structure.

Damping for the subfoundation can also be specified alone. Assuming its damping ratio is

, the damping coefficient is computed by cς

cccc mc ςω= 2 , where c

bcc m

kkω

+= (7-9)

Table 7.1: Damping coefficients and modal damping ratios

%2=ςb , %1=ςc %5=ςb , ς %1=c

Damping Superstructure Superstructure + Base

Linear system where kn = 0

Superstructure + Base

Linear system where kn = 0

1.86 4.65 1.96 1.51 4.88 2.05 2 3.17 2.21 3.57 3.90

ς (%)

5.24 5.48 5.49 5.53 5.53 c1 1.1723 c2 1.1723 cb 0.3960 0.9899

c (Ns/m)

cc 0.724 0.724

Assuming damping of the fundamental mode of the superstructure to be 2% of critical, the

remaining damping coefficients are calculated and shown in Table 7.1. If damping ratios of the

base and foundation are, respectively, %5=bς and %1=cς , it follows that c Ns/m

and c Ns/m. After all of the damping coefficients are determined, modal damping ratios

9899.0=b

724.0=c

59

for the assembled structure and its component parts can be estimated. These are also presented in

Table 7.1.

Observe that the damping in the first mode of the superstructure-base assembly is, 1.96%

and 4.88%, corresponding to the isolation-rigid body system damping of 2% and 5%,

respectively. Damping in the structure has little influence on the base modal damping, because the

structure behaves almost as a rigid body in this mode. By contrast, high damping of the isolation

system has increased the damping in the first mode of the superstructure from 2% to 3.17% and

3.57%, respectively, for the two values of base isolation system damping. This is a consequence

of nonclassical damping in the assembled system. The coupling terms are nonzero, and so the

modal equations are coupled. As we neglect the coupling terms when calculating the modal

damping ratios, modal analysis provides approximate results.

We see that the modal damping ratios for the fixed-base superstructure increase linearly with

natural frequency because the damping is stiffness proportional. However, as we are concerned

primarily with the tuned modes in the 4DOF isolated system, where the 1st and 2nd modal

damping ratios are approximately 2%, the values chosen are appropriate.

7.2 Response to the Analytical (He) Model of Ground Velocity

The actual earthquake record contains not only low-frequency pulses, but also higher-

frequency components. As shown in Chapter 5, the analytical model by He (2003) captures the

major characteristics of ground motions with near-field effects. Therefore, this model will be

employed to investigate the dynamics of the damped system. The effects of damping in both the

superstructure and subfoundation on the system performance will be investigated. Damping at the

base will also be included in the parametric analysis.

7.2.1 Damping in the Superstructure Only ( 2%ς s = )

Comparing the damped modal responses of Figure 7.1 with the undamped responses of

Figure 6.9, we observe that a small level of damping in the superstructure decreases the amplitude

of the linear 2nd mode response by 2.5 times while the envelopes of the building and

subfoundation remain complementary to each other; in the nonlinear system, the degree of

localization remains nearly the same as in the undamped system. This is expected, as the

necessary conditions for nonlinear localization to occur are internal resonance and suitable

nonlinearity in the isolation system, which are unaffected by damping in the superstructure. This

60

is confirmed by the time history responses in Figure 7.2, where damping in the superstructure

changes the linear superstructure response, while nonlinear responses and responses of the base

and the subfoundation are almost identical to those of the undamped case, with only minor

diminished amplitudes because of energy dissipation.

Figure 7.1: Modal responses of the linear and the nonlinear system (ς ) %2=s

Figure 7.2: Damped time history responses ( %2=sς )

Linear Nonlinear

61

Figure 7.3 Power spectra ( %2=sς )

7.2.2 Damping in the Subfoundation Only ( 2%ς c = )

Figure 7.4 Modal responses of the linear and the nonlinear system ( ) %2ς =c

62

Figure 7.5 Damped time history responses ( ς %2=c )

Linear Nonlinear

Interestingly, 2% damping parallel to the tuning spring is allocated equally to the tuned

modes: the 2nd and 3rd modal damping ratios are both 1%, while the 1st and 4th are nearly zero.

Notice that the ground motion is also scaled to the tuning period, so it is expected that this

damping helps dissipate ground energy and structural energy in the tuned modes. As a result, the

response of the tuned modes decreases and that of the 1st mode increases. The performance of the

nonlinear system compared with that of the linear system is not as impressive, because nonlinear

localization only contributes partly to suppression of the tuned modes, the remainder occurring

through dissipation in common with the linear system.

From Figure 7.4, we see that the maximum amplitude of the 2nd nonlinear modal response

(the tuned mode) is comparable to that of the undamped case, while the maximum amplitude of

the 2nd linear modal response is approximately 40% of that of the undamped case. From the time

history responses (Figure 7.5), it is apparent that both linear and nonlinear systems perform

adequately.

7.3 Parametric Analyses

The clearance e and the clearance spring stiffness kn are design parameters that determine the

strength of the nonlinearity. The smaller gap e and larger kn, the stronger the nonlinearity. In this

63

section, we will discuss how sensitive the performance of the system is to these parameters. It is

desirable to find optimal values of the full parameter set to ensure the best possible design.

However, many factors must be included in the analysis for the nonlinear system to make the

study tractable. Thus, only some qualitative results with examples will be used to illustrate the

necessary steps in this analysis. In additional to the two parameters mentioned previously,

damping at the base level is also included to show the sensitivity of the system to it.

7.3.1 Clearance Size and Clearance Spring Stiffness

In this system, strong nonlinearity helps to drive the input energy into high frequency modes

of the structure, while weak nonlinearity results in a smaller frequency shift. We desire strong

mode localization; thus it seems desirable to shift as much input energy to high frequencies as

possible. In other words, stronger nonlinearity seems desirable. However, this is not the case. By

contrast, excessively large kn decreases the effect of localization and amplifies accelerations; and

making e too small actually results in an untuned linear system which is very sensitive to high-

frequency ground motions. These effects are illustrated in Figures 7.6 through 7.11.

Figure 7.6: Modal responses with a smaller clearance

Figures 6.9 to 6.11 in Chapter 6 showed the undamped responses to the modeled ground

velocity with 008.010/|)_max(| == linyce m. To examine the system sensitivity to e, we give

the modal responses and response power spectra for identical systems incorporating, first, a

64

smaller value of clearance 004.020/|)_max(| == linyc

021.04/|)_max(|

e m, shown in Figures 7.6 and 7.7, and a

larger value of clearance, == linyce m, in Figures 7.8 and 7.9.

Figure 7.7: Power spectra for system with smaller clearance

Comparing Figures 7.6 and 7.7 with Figures 6.9 and 6.11, we see that a smaller clearance

results in better localization; it eliminates the beating in the 3rd mode as well, and shifts the energy

from 35 Hz to 42 Hz.

On the other hand, a large clearance decreases the localization effect for this specific kn.

Though the amplitude of the 2nd mode decreases, the beat phenomenon is still apparent, and the

amplitude of the 3rd mode is large. In Figure 7.8 and 7.9, we see that energy shifts to a frequency

only modestly higher than the resonant frequency. In fact, for cn kk 10= , the best clearance is zero

--- an untuned linear system. It is apparent that the good performance of such a system is not a

result of the tuning-localization algorithm; it just happens that the system is not sensitive to this

particular ground motion.

65

Figure 7.8: Modal responses with a larger clearance

Figure 7.9: Power spectra for a system with larger clearance

The effects of kn are given in Figures 7.10 and 7.11 for cases when k and cn k15= cn kk 2= ,

respectively.

66

Figure 7.10: Modal responses with a larger kn

Figure 7.11: Modal responses with a smaller kn

Here, it can be seen that neither smaller nor larger kn results in improved performance. We

conclude, then that the system performance is not sensitive to kn in this range. However, if

and approaching zero, the system will become the linear tuned system, and the nonlinear

localization effect disappears. For the specified clearance

cn kk 2<

008.010/|)_max(| == linyce m, the

range is reasonably good for nonlinear localization, likely because the resulting

impact forces don’t undergo drastic change.

cn kk 10<<ck2

67

7.3.2 Damping at the Base Only

Figure 7.12: Modal responses of the linear and the nonlinear system ( ) %2ς =b

Figure 7.13: Damped time history responses ( %2ς =b )

Linear Nonlinear

Figure 7.12 shows the linear and nonlinear modal responses with ς . It is apparent

here that damping at the base interrupts the energy interchange between the two subsystems,

%2=b

68

leading to an irregular pattern of beating in the linear case. However, the modal amplitudes do not

experience much of a decrease. In the nonlinear system, localization occurs only in the early

phase when ground motion is large. Time history responses (Figure 7.13) also exhibit this abrupt

change in behavior. In the first phase, the beating phenomenon can be observed in the linear

system, and displacements and accelerations both decrease significantly in the nonlinear system.

In the second phase, the beating is absent in the linear case, and in the nonlinear system

displacement and acceleration responses grow larger, with less energy transfer to the

subfoundation.

It may seem counterintuitive that responses can grow with time in a damped system, while

the external excitation decays to nearly zero. We will examine where the energy causing this

comes from by analyzing the “control” force in the isolation system. By control force, we mean

the impact force, or force generated by impacts in the clearance spring kn. The force in the

clearance spring is expressed by

(7-10)

+

−=

)(0

)()(

euk

eukuF

n

n

c

eueue

eu

−<≤≤−

>

For the undamped response to the modeled ground velocity mentioned in Chapter 6, the control

force and the absolute value of its envelope are shown in Figure 7.14. Matlab code for the force

envelope is given in Appendix D.

Figure 7.14: Control force and force envelope in the undamped system

69

From Figure 7.14, it can be seen that energy is regularly transmitted into the clearance spring

through impact for the duration of the input in the undamped system. It is this sequence of

impacts that changes the physical property of the system; through them, the ground energy

transfers through the clearance spring to the subfoundation, where it is confined and prohibited

from “leaking” into the superstructure. In the meantime, the system displays an entirely different

behavior than the linear system. It vibrates in a stable, localized mode that has no counterpart in

the linear system. This behavior of this nonlinear system sets it apart from the more usual

“impact damping” configurations that rely on collisions between particles and/or components to

dissipate energy.

For 0.5% damping at the base level (i.e., %5.0=ςb ), the force and force envelope vary with

time as shown in Figure 7.15. Interestingly, the control force diminishes with time, until there is

only one-sided impact. When the impact force reaches zero, the nonlinear mechanism no longer

occurs, there is no interaction between the linear structure and the nonlinear isolation system, and

the system responds linearly, without localization.

Summarizing, with the proposed nonlinear isolation system, energy is not dissipated but,

rather, is transferred into the clearance spring through repetitive impacts, and the system is driven

into a localized nonlinear normal mode. In the system with damping at the base, energy is

dissipated, weakening the effect of the clearance nonlinearity with time. Eventually, the control

force diminishes, and the structure vibrates in a non-localized linear mode.

Figure 7.15: Control force and force envelop with %5.0=ςb

70

This effect of damping at the base is further examined in Figure 7.16, where force envelopes

for different damping levels are plotted.

Figure 7.16: Control force envelop for varying damping levels

Here, it is clear that even a small change in damping at the base can greatly decrease the

control force. However, if the clearance e is decreased, thus maintaining the control force, the

performance is likely to improve. To examine this hypothesis, we halved the clearance used in the

example of Figures 12 and 13, for 004.020/|)_max(| == linyce m.

Figure 7.17: Control force envelop for %2=ςb and a smaller clearance

71

In Figure 7.17, we compare the control forces corresponding to base damping of 2% for the

two clearance levels. Clearly, the impacts persist for the smaller clearance, thus prolonging the

nonlinear behavior and resulting localization.

Finally, examination of the response of the system in Figure 7-18 with 2% base damping and

smaller clearance reveals that the second mode response has been further suppressed by a modest

amount, at the expense of the third mode response, which increases slightly. Note the growth in

second mode response once the control force, shown in the previous plot, goes to zero.

Figure 7.18: Modal responses for %2=ςb and a smaller clearance

72

Chapter 8: Comparison of the Nonlinear Isolation System with

Conventional Base Isolation

In the last two chapters, we examined the responses of the nonlinear system subjected to the

scaled ground velocity history that models the Erzincan earthquake. In this chapter, we will

examine the performance differences between this nonlinear isolation system and the

conventional system insofar as their ability to suppress undesired higher modes of response.

Furthermore, we will assess the performance they achieve when subjected to ground motions with

different frequency content. To facilitate this study, we will still use the analytical model of He

(2003) for simulation. The effect of damping on the performance of conventional base isolation

systems has been well documented (see, for example, Chopra, 1995), and its effect upon the

performance of our nonlinear system was discussed in Chapter 7. Thus, in the brief study to

follow, we will consider only the undamped case in order to assess the relative merits of each.

8.1 Conventional Base Isolation

As is well known (Skinner et al., 1993), conventional base isolation relies on linear mode

orthogonality to suppress higher modes of response, so that the isolated structure moves primarily

in its fundamental (soft first story) mode. The conventionally isolated system corresponding to

our novel nonlinear system is shown in Figure 8.1.

)(g tx&&

GROUND

m2

mb

k1

x1

m1

k2

x2

kb=ε k1

xb

Figure 8.1: Conventionally isolated structure

The system consists of a two-story building (the superstructure) plus a base mass (the foundation).

The base mass is connected directly to the ground through a compliant spring , here assumed to bk

73

be linear. As stipulated in Chapters 6 and 7, 301

≈ε . The structure then has a fundamental period

of 0.443 s, and a second natural period of 0.052 s.

8.2 Response to the Scaled Velocity History

From Equations (5-7) and (5-15) in Chapter 5, we see that the period used to capture the

main pulses of Erzincan earthquake is

912.1222

=+

π=

bagT s (8-1)

This period is inconsistent with our simulation models. In previous chapters, we employed a

scaled velocity history to examine the dynamics around the tuned periods of the structure.

Scaling was accomplished by reducing the time step to 0.00025 s, 1/20 of the original value of

0.005 s. This results in a scaled period of the ground motion of approximately 0.1 s, near the 2nd

peak of the linear shock spectrum (Figure 6.2).

We also examine the performance of both isolation systems when subjected to excitations

with frequency content at the (near) rigid-body mode of the system. To scale the earthquake to the

rigid-body mode, the scale factor must be approximately 1/5; i.e., the time step becomes 0.001 s.

In Subsections 8.2.1 and 8.2.2, a comparison will be made of the performance of the two isolation

systems subjected to both scaled earthquake records.

8.2.1 Velocity History Scaled to the First Deformational Mode (Tg = 0.1 s )

Figure 8.1 shows the modal responses for the first three modes of the two isolated structures.

Conventional isolation is better than the nonlinear system at suppressing the second and third

modes. Neither system is able to substantially affect the rigid-body mode. This is further

illustrated in the time history responses (Figure 8.2), where the response curves for conventional

isolation are smooth, and those for nonlinear isolation have high-frequency components

overriding the low-frequency responses. The ability of the nonlinear system to reduce the effect of

the second mode can be improved by decreasing the value of k . Theoretically, the nonlinear

system could eliminate the tuned modes entirely, provided that the coupling spring stiffness is

sufficiently small, not entirely practical for real structures.

b

74

Figure 8.1: Modal responses of the linear and the nonlinear system

Figure 8.2 Time history responses

The power spectra of the relative displacement response are consistent with prior

observations. The undamped linear isolation system suppresses the high-frequency responses by

mode orthogonality, while the nonlinear system suppresses the tuned frequency and shifted

energy to higher non-localized modes. Here, the spectra are calculated from response time

histories. The spectra for the nonlinear system are of displacements with respect to the

subfoundation, while the spectra for the conventional isolation system are of displacements with

75

respect to ground. Note that the response at the frequency corresponding to the “rigid body” mode

is higher for the conventional system, as some of this energy is shifted to higher frequencies in the

nonlinearly isolated structure.

Figure 8.3: Power spectra

8.2.2 Velocity History Scaled to the Fundamental Mode (Tg = 0.45 s)

Neither isolation system is designed to affect the near rigid-body response of the structure.

The reader should be cautioned that the use of undamped responses for this evaluation may prove

somewhat deceiving, in that responses due to the higher (deformational) modes are four to five

orders of magnitude smaller than the near rigid-body response. Damping in the isolator would

provide some control on the rigid-body behavior and likely provide a better benchmark. This,

however, will be left to verify in continuing work.

76

Figure 8.4: Modal responses of the linear and the nonlinear systems

Figure 8.5: Undamped time history responses

Comparing the modal responses of Figure 8.4, we observe that both systems have nearly

equal first mode response (10,000 times greater than the second mode response). This is a direct

result of tuning the excitation to the near rigid-body mode. This is also perfectly clear from the

time history response of Figure 8.5. Finally, examination of the power spectra of Figure 8.6 for

the two systems, as well as for the linear system resulting from setting the spring constant of the

77

clearance spring to zero, reveals that which was obvious from the beginning, that long-period

excitation does not drive the system to its localized mode. Thus, no advantage is gained from the

nonlinear isolation system when the energy content of the input is concentrated at frequencies

well below the deformational modes of the structure.

Figure 8.6: Power spectra

8.3 Some Final Thoughts

While comparisons of the undamped responses are not compelling, it is possible to note

some likely trends. For example, the nonlinear isolation system that we propose herein will be

efficacious when the frequency content of the earthquake is concentrated near the deformational

modes of the structure. This will most likely be the case in high-rise buildings, for which the

conventional base isolation is not suitable. Furthermore, the nonlinear system utilizes the large

initial energy of the near field earthquake to quickly drive the system to its localized response,

where the subfoundation is the primary vibrating element while the remainder of the structure

moves nearly as a rigid body. This, of course, minimizes interstory displacements and resulting

stresses in the building structure.

The localization phenomenon can be graphically observed in the PSD plots of absolute

acceleration response, shown in Figure 8.7. Here, we focus on the first case, where the energy of

the input is concentrated near the first deformational mode of the structure.

78

Figure 8.7: Power spectra of absolute acceleration

Linear Nonlinear Conventional

Examination of the relative magnitudes of these spectra at the 37 Hz peak reveals the localization,

with the response of the subfoundation being orders of magnitude greater than that of the

remaining degrees of freedom. This is sharp contrast to the peak near 20 Hz, which represents a

perturbation to the linear mode and is localized to the structure and not the subfoundation.

The fundamental issue, then, is, can sufficient input energy be shifted from lower modes,

especially the near rigid-body mode, to the higher mode, localized to the subfoundation, to

significantly affect the overall response of the primary structure to the point where the new

isolation system outperforms the conventional system. The results of this brief study are

inconclusive in this regard.

79

Chapter 9: Conclusions and Recommendations for Further Study

In this thesis, a novel seismic base isolation concept has been introduced which makes use of

internal resonance and nonlinear localization to confine seismic energy to a nonstructural element,

the subfoundation, which is a noncritical part of the overall system. We have shown that, with

proper tuning, any deformational mode of the primary structure can be suppressed, with resulting

reductions in response. When compared with a nearly identical system without the additional

clearance spring (thus making the system linear), the nonlinear system performs well. The

response of the first deformational mode of the structure is almost fully suppressed, and system

response to the full-sine velocity pulse, the historic Erzincan earthquake, and the He model of the

Erzincan earthquake showed peak reductions in the interstory displacements of up to 61% and in

the absolute accelerations of up to 65%.

The potential advantages of such a system are many: First, the additional stage required, the

subfoundation, serves no purpose other than as a tuning mass. Thus, it can be fabricated

inexpensively and is not weight critical. Second, a clearance spring not only is easy to implement

but also can be specified during the design and construction stage. Thus, this simple additional

step can provide the needed strong nonlinearity. Third, this system can suppress the response of

any individual deformational mode and provide relatively broad-frequency protection. It is

similarly efficient for acceleration and displacement inputs, as well. Fourth, as stated in this

thesis, the working principle for this nonlinear element is well understood and, thus, its

performance can be predicted and assured. The degree of nonlinearity, damping location and

damping ratio can be specified according to the designer’s preference to achieve the desired level

of vibration reduction.

However, this manifestation of the system is not ideal in that, like a traditional base isolation

system, it cannot suppress the rigid-body mode, which overshadows performance when the

ground motion contains long-period pulses. It is also deficient in that it possesses only one

localized NNM. By adding additional nonlinear elements, one is able to induce more localized

NNMs in the system, thus enhancing its capacity to confine seismic energy in the subfoundation

and provide better isolation to the main structure. In addition, the design might be improved if the

subfoundation were designed with additional tunable degrees of freedom, if the nonlinear

elements were rearranged, if additional nonlinear damping elements were introduced, or if

damping and/or nonlinearities in the primary structure were considered.

These are likely candidates for future investigations.

80

List of References

Anand, G. V., 1972, “Normal Modes of a Coupled Nonlinear System”, International Journal of Nonlinear Mechanics, 7(1), 81-91. Anderson, P. W., 1978, “Localized Moments and Localized States”, Reviews of Modern Physics, 50(2), 191-201. Anderson, J. G., and Bertero, V., 1986, “Uncertainties in Establishing Design Earthquake”, Journal of Structural Engineering, ASCE, 113, 1709-1724. Anderson, J. G., Bodin, P., Brune, J. N., Prince, J. and Singh, S. K., 1986, “Strong Ground Motion from the Michoagan, Mexico Earthquake”, Science, 233, 1043-1049. Arnold, F. R., 1955, “Steady-State Behavior of Systems Provided With Nonlinear Dynamic Vibration Absorbers”, Journal of Applied Mechanics, 22, 487-492. Caughey, T. K. and O’Kelly, M. E. J., 1965, “Classical Normal Modes in Damped Linear Dynamic Systems”, Journal of Applied Mechanics, 32(3), pp583-588. Caughey, T. K., Vakakis, A. F., and Sivo, J., 1990, “Analytic Study of Similar Normal Modes and Their Bifurcations in a Class of Stronghly Nonlinear Dynamic Systems”, International Journal of Nonlinear Mechanics, 25(5), pp521-533. Chopra, A. K., 1995, Dynamics of Structures, Theory and Applications to Earthquake Engineering, Prentice-Hall, Inc. NJ. Crede, C. E., 1951, Vibration and Shock Isolation, Wiley and Sons, New York. Cuvalci, O., Ertas A. Ekwaro-Osive, S., and Cicek, I., 2002, “ Nonlinear Vibration Absorber for a System Under Sinusoidal and Random Excitation: Experiments”, Journal of Sound and Vibration, 249(4), 707-718. Dabney, J. B., and Harman, T. L., 1998, The Student Edition of SIMULINK: Dynamic System Simulation for MATLAB, Prentice Hall, New Jersey. Den Hartog, J. P, 1956, Mechanical Vibrations (4th edition), McGraw-Hill, New York. Der Kiureghian, A., Sackman, J. L. and Hong, K-J., 1999, “Interaction in Interconnected Electrical Substation Equipment Subjected to Earthquake Ground Motions”, Report PEER 1999/01, Pacific Earthquake Engineering Research Center, University of California, Berkeley, CA, February. Grubin, C., Monica S., 1956, “On the Theory of the Acceleration Damper”, ASME Journal of Applied Mechanics, 23, 373-378. Hall, J. F., Heaton, T. H., Halling, M. W., and Wald, D. J., 1995, “Near-Source Ground Motions and Effects on Flexible Buildings”, Earthquake Spectra, 11, 569-605. He, W., 2003, Smart Energy Dissipation Systems for Protection of Civil Infrastructures from Near-Field Earthquakes, PhD Thesis, The City University of New York, 2003.

81

He, W., and Agrawal, A. K., 2002, “A Closed-Form Approximation of Near-Fault Ground Motion Pulses for Flexible Structures”, to be submitted. Hodges, C, 1982, “Confinement of Vibration by Structural Irregularity”, Journal of Sound and Vibration, 82(3), 441-424. Iwan, W. D., and Chen, X. D., 1994, “Important Near-Field Ground Motion Data from the Landers Earthquake”, Proceedings of 10th European Conference of Earthquake Engineering, Vienna. Iwan, W. D., 1997, “Drift Spectrum: Measure of Demand for Earthquake Ground Motions”, Journal of Structural Engineering, 123(4), 397-404. Iwan, W. D., 2002, Personal communication with the author on International Conference on Advances and New Challenges in Earthquake Engineering Research, August 15-17, 2002, Harbin, China. Jiang, X., 2002, “Theoretical and Experimental Studies of Steady-State Nonlinear Localization and Energy Pumping in Systems of Coupled Oscillators”, Ph.D. Thesis, University of Illinois at Urbana-Champaign. Jones, D. I. G., Nashif A. D., and Adkins R. L., 1967, “Effect of Tuned Dampers on Vibrations of Simple Structures,” AIAA Journal, 5(2). Khachian, E., Der Kiureghian, A., and Tonoyan, A., 1998, “Dynamic Properties Of Primary-Secondary Systems”, Journal of Structural Engineering Association, Yerevan, Armenia, 7(24), 1-4 (in Armenian). Kelly, J. M., 1997, Earthquake-Resistant Design With Rubber, 2nd ed., Berlin and New York: Springer-Verlag. Kelly, J. M., Beucke, K. E., and Skinner, M. S., 1980, “Experimental Testing of a Friction Damped Aseismic Base-Isolation System with Fail-Safe Characteristics”, Report No. UCB/EERC-80/18, Earthquake Engineering Research Center, University of California, Berkley, CA, 1980. Leo, D. J., and Inman, D. J., 1999, “A Quadrative Programming Approach to the Design of Active-Passive Vibration Isolation Systems”, Journal of Sound and Vibration, 220(5), 807-825. Li, C, 2002, “Optimum Multiple Tuned Mass Dampers for Structures Under the Ground Acceleration Based on DDMF and ADMF”, Earthquake Engineering and Structural Dynamics, 31: 897-919. Liu, T. H., Marr, A., 1987, Seismic Engineering: Recent Advances in Design, Analysis, Testing and Qualification Methods, presented at the 1987 Pressure Vessels and Piping Conference, San Diego, California, June 28-July 2, 1987, PVP Series, 127, ASME, USA. Luongo, A. 1991, “On the Amplitude Modulation and Localization Phenomena in Interactive Buckling Problems”, International Journal of Solids and Structures, 27(15), 1943-1954.

82

MacKay, R. S., and Aubry, S., 1994, Proof of Existence of Breathers for Time-Reversible or Hamiltonian Networks of Weakly Coupled Oscillators, Nonlinear Systems Laboratory Preprint 20/1994, Mathematics Institute, University of Warwick, Coventry, UK. Makris, N., 1997, “Rigidity-Plasticity-Viscosity: Can Electrorheological Dampers Protect Base-Isolated Structures from Near-Source Ground Motions?” Earthquake Engineering and Structural Dynamics, 26, 571-591. Makris, N., and Chang, Shih-Po., 1998, “Effect of Damping Mechanisms on the Response of Seismically Isolated Structures”, PEER Report 1998/06, Pacific Earthquake Engineering Research Center, University of California, Berkeley, November 1998. Malhotra, P. K., 1997, “Dynamics of Seismic Impacts in Base-Isolated Buildings”, Earthquake Engineering and Structural Dynamics, 26, 797-813. Malhotra, P. K., 1999, “Response of Building to Near-Field Pulse-Like Ground Motions”, Earthquake Engineering and Structural Dynamics, 28, 1309-1326. Masri, S. F., 1973, “Steady-State Response of a Multidegree System With an Impact Damper”, ASME Journal of Applied Mechanics, 40, 127-132. McFarland, D. M., Wang, Y., Vakakis A. F., and Bergman L. A., 2002a, “A Parametric Analysis Of A Novel Shock Isolation System”, Proceedings of the 15th ASCE Engineering Mechanics Conference EM 2002, 2-5 June, 2002, Columbia University, New York, NY. CD-ROM, Paper no. 337, 8 pages. McFarland, D. M., Wang, Y., Vakakis, A. F., and Bergman, L. A., 2002b, “Model Testing of a Nonlinear Base Isolation Concept”, Proceedings of the International Conference on Advances in Building Technology, Hong Kong, 4-6 December 2002, 513-520. Meirovitch, L. I., 2001, “Description of Localized Normal Modes in the Chain of Nonlinear Coupled Oscillators Using Complex Variables”, Nonlinear Dynamics 25, 95-109. Naeim, F., and Kelly, J. M., 1999, Design of seismic isolated structures: From theory to practice, Wiley, Chichester, England. Nagarajaiah, S., and Xiaohong S., 2000, “Response of Base-Isolated USC Hospital Building in Northridge Earthquake”, Journal of Structural Engineering, 126(10), 1177-1186. Nakhta, T., 1973, Approximate Dynamic Response of Light Secondary Systems, Ph.D. Thesis, University of Illinois at Urbana-Champaign. Nashif, A. D., Jones, D. I. G., Henderson, J. P., 1985, Vibration Damping, Wiley and Sons, New York. Nayfeh, S. A., and Nayfeh, A. H., 1994, “Energy Transfer from High- to Low-Frequency Modes in a Flexible Structure via Modulation”, Journal of Vibration and Acoustics, 116, 203-207. Nayfeh, T. A., Emaci, E., and Vakakis A. F., 1997, “Application of Nonlinear Localization oo ohe Optimization of a Vibration Isolation System”, AIAA Journal, 35(8), 1378-1386.

83

Penzien, J. and Chopra, A., K., 1965, “Earthquake Responses of Appendage on a Muti-Story Building”, Proceedings of the Third World Conference on Earthquake Engineering, New Zealand. Pierre, C. and Dowell, E. H., 1987, “Localization of vibrations by Structural irregularity”, Journal of Sound and Vibration, 114(3), 549-564. Pierre, C. 1988, “Mode localization and Eigenvalue Loci Veering Phenomena in Disordered Structures”, Journal of Sound and Vibration, 126(3), 485-502. Pipes, L. A., 1953, “Analysis of a Nonlinear Dynamic Vibration Absorber”, Journal of Applied Mechanics, 20, 515-518. Rand, R. H., Pak, C. H., and Vakakis, A., F., 1992, “Bifurcation of NNMs in a Class of Two DOF Systems,” Acta Mechanics, (Suppl), 3, 129-145. Reed, F. E., 1961, “Dynamic Vibration Absorbers and Auxiliary Mass Dampers,” Shock and Vibration Handbook (eds. Harris C. M. and Crede C. E.), McGraw-Hill, New York. Roberson, R. E., 1952, “Synthesis of a Nonlinear Dynamic Vibration Absorber”, Journal of Applied Mechanics, 27, 263-268. Rosenberg, R. M., 1960, “Normal Modes of Nonlinear Dual-Mode Systems”, Journal of Applied Mechanics, 27, 263-368. Rosenberg, R. M., 1963, “The Ateb(h)-Functions and Their Properties”, Quart. Applied Mathematics, 21(1), 37-47. Rosenberg, R. M., and Kuo, J., 1964, “Nonsimiliar Normal Mode Vibrations in a Nonlinear System Having Two Degrees of Freedom”, Journal of Applied Mechanics, 9, 155-242. Ruzicka, G. C., 1981, Dynamic response of tuned secondary systems, Ph.D. Thesis, University of Illinois at Urbana-Champaign. Sadek, F., Mohraz, B, Taylor, A. W., and Chung, R. M., 1997, “A Method of Estimating the Parameters of Tuned Mass Dampers for Seismic Applications”, Earthquake Engineering and Structural Dynamics, 26: 617-635. Shaw, S., and Alsuwaiyan, A., 2000, “Torsional Vibration Reduction Using Passive Nonlinear Absorbers”, Smart Structures and Materials 2000, Damping and Isolation, Proceedings of SPIE, 388-399. Shaw, S., and Pierre, C., 1991, “Nonlinear Normal Modes and Invariant Manifolds”, Journal of Sound and Vibration, 150(1), 170-173. Shaw, S., and Pierre, C., 1993, “Normal Modes for Nonlinear Vibration Systems”, Journal of Sound and Vibration, 164(1), 85-124. Skinner, R. I., Robinson, W. H., McVerry G. H., 1993, An Introduction to Seismic Isolation, Willey, Chichester, England.

84

Somerville, P. G. and Graves, R. W., 1993, “Conditions That Give Rise to Unusually Large Long Period Ground Motions”, The Structural Design of Tall Buildings, 2, 211-232. Thomson, W. T., 1972, Theory of Vibration with Applications, Prentice-Hall, Inc., Englewood Cliffs, N. J., 1972. Tsai, H., 1997, “Dynamic Analysis of Base-Isolated Shear Beams Bumping Against Stops”, Earthquake Engineering and Structural Dynamics, 26, 515-528. Vakakis, A. F., 1990, “Analysis and Identification of Linear and Nonlinear Normal Modes in Vibrating Systems”, PhD Thesis, California Institute of Technology, Pasadena, CA. Vakakis, A. F., 1993, “ Analytic Study of Linear and Nonlinear Mode Localization”, Journal of Applied Mechanics, 60(2), 555-557. Vakakis, A. F., 2001, “Inducing Passive Nonlinear Energy Sinks in Linear Vibrating Systems”, Journal of Sound and Vibration, 123(3), 324-332. Vakakis, A. F., Manevitch, L. I., Mikhlin, Y. V., Pilipchuk, V. N., and Zevin, A. A. (1996). Normal Modes and Localization in Nonlinear Systems, John Wiley & Sons, New York. Vakakis, A. F., Kounadis, A. N. and Raftoyiannis, I. G., 1999, “Use Of Nonlinear Localization for Isolating Structures from Earthquake-Induced Motions”, Earthquake Engineering and Structural Dynamics, 28, 21-36. Villaverde, R., 1985, “Reduction in Seismic Response With Heavily-Damped Vibration Absorber”, Earthquake Engineering and Structural Dynamics, 13: 33-42 Wang, Y., McFarland, D. M., Vakakis, A. F., and Bergman, L. A., 2002, “Efficacy of a nonlinear base isolation system subjected to near-field earthquake motions”, Proceedings of International Conference on Advances and New Challenges in Earthquake Engineering Research, August 15-17, 2002, Harbin, China. Warburton, E. O., and Ayorinde, G. B., 1980, “Minimizing Structural Vibration with Absorbers”, Earthquake Engineering and Structural Dynamics, 8: 219-236. Yalla, S. K. and Kareem, A., 2000, “On the Beat Phenomenon in Coupled Systems”, 8th ASCE Conference on Probablistic Mechanics and Structural Reliability, PMC2000-238. Yokomichi, I., Araki, Y., Jinnouchi, Y., Inoue, J., 1996, “Impact Damper With Granular Materials for Multibody System”, Journal of Pressure Vessel Technology, 118, 95-103.

85

Appendix A Simulink Models: Nonlinear Base Isolation System

Simulink model of the 4DOF nonlinear base isolation system

86

Subsystem model of the superstructure

87

Subsystem model of the isolation system

88

Appendix B Matlab Code: Main Program VERBOSE = 0; path(path, 'F:/earthquakes/erzican') load Erz_ns.mat; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% MDOF system (superstructure) %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% zeta_ss = 0.00; zeta_b = 0.00; zeta_c = 0.00; %% Undamped %% Frequency of first mode of superstructure (building) with base fixed f_ss_1 = 13.012; w_ss_1 = 2.0 * pi * f_ss_1; T_ss_1 = 1.0 / f_ss_1; %% First story m1 = 0.2; k1 = 0.5 * (3.0 + sqrt(5.0)) * m1 * w_ss_1^2; %% Second story m2 = m1; k2 = k1; %% Superstructure matrices m_ss = diag([m1, m2]); k_ss = [[k1 + k2, -k2], [-k2, k2]]; wn_ss = sqrt(eig(k_ss, m_ss)); fn_ss = wn_ss ./ (2.0 * pi); Tn_ss = 1.0 ./ fn_ss; [mode_u, lamda_u] = eig(k_ss, m_ss); M_U = mode_u' * m_ss * mode_u; K_U = mode_u * k_ss * mode_u'; beta = 2* zeta_ss/wn_ss(1); c_ss = beta* k_ss; c2 = c_ss(2,2); c1 = c_ss(1,1) - c_ss(2,2); if VERBOSE fprintf('\nSuperstructure physical parameters:\n') m1 m2 c1 c2 k1 k2 fprintf('\nNatural frequencies of 2-DOF superstructure with base fixed:\n') wn_ss fn_ss Tn_ss End %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Ground shock %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

89

global vpulse Vg Tg vpulse = 'typeB'; % The shock pulse is a complete wave Tg = 0.052 Vg = 9.81 / (2* pi/Tg); if VERBOSE if isglobal(pulseshape) fprintf('\nThe variables\n') pulseshape Ag Tg fprintf('have been declared global.\n') end vpulse Vg Tg end t0 = 0.0; tf = 3; timestep = 0.00025; simoptions = simset('FixedStep', timestep); simoptions = simset(simoptions, 'Solver', 'ode4'); %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Unisolated system %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% n = 2; d_ss = [[k1, c1], [0, 0]]; A1_ss = [[c2, m2], [m2, 0]]; A2_ss = [[k2, 0], [0, -m2]]; B1_ss(1:2,1:2) = d_ss; A_ss = -inv(A1_ss) * A2_ss; B_ss = inv(A1_ss) * B1_ss; C_ss = eye(2*n-2); D_ss = zeros(2*n-2, 2); mb = m1; cb = c1; kb = k1; m1_disp_index = 1; m1_vel_index = n; sim('conventional_vp', [], simoptions); yb_u = yb_ss; ys_u = zs_ss + yb_ss; xs_u_tt = xs_ss_tt; xb_u_tt = xb_ss_tt; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Traditional Isolation System %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% n = 3; mb = 0.3; kb = k1 / 30.0; m_ssb = diag([mb, m1, m2]);

90

k_ssb = [[k1 + kb, -k1, 0.0], [-k1, k1 + k2, -k2], [0.0, -k2, k2]]; wn_ssb = sqrt(eig(k_ssb, m_ssb)); fn_ssb = wn_ssb./ (2.0 * pi); Tn_ssb = 1.0 ./ fn_ssb; [mode_b, lamda_b] = eig(k_ssb, m_ssb); M_C = mode_b' * m_ssb * mode_b; K_C = mode_b' * k_ssb * mode_b; wb = sqrt(kb /(m1+ m2+ mb)); cb = 2* zeta_b* wb* (m1+ m2+ mb); c_b(2:3,2:3) = c_ss; c_b(1,2)= -c1; c_b(2,1)= -c1; c_b(1,1)= c1 + cb; c = mode_b'* c_b* mode_b; for i = 1:3 zeta_ssb(i) = c(i,i)/ (2* M_C(i,i) * wn_ssb(i)); end A1_ss = [[c_ss, m_ss], [m_ss, zeros(size(m_ss))]]; A2_ss = [[k_ss, zeros(size(m_ss))], [zeros(size(m_ss)), -m_ss]]; B1_ss(1:2,1:2) = d_ss; B1_ss(3:4,1:2) = zeros(2,2); A_ss = -inv(A1_ss) * A2_ss; B_ss = inv(A1_ss) * B1_ss; C_ss = eye(2*n-2); D_ss = zeros(2*n-2, 2); m1_disp_index = 1; m1_vel_index = n; if VERBOSE mb kb cb wn_ssb fn_ssb Tn_ssb end sim('conventional_vp', [], simoptions); yb_c = yb_ss; ys_c(:,1) = zs_ss(:,1) + yb_ss; ys_c(:,2) = zs_ss(:,2) + yb_ss; xs_c_tt = xs_ss_tt; xb_c_tt = xb_ss_tt; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% MDOF superstructure plus base %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Matrices of superstructure plus base (with rigid-body mode) mb = 0.3; kb = k1 / 25.0;

91

k_ssc = [[k1 + kb, -k1, 0.0], [-k1, k1 + k2, -k2], [0.0, -k2, k2]]; wn_ssc = sqrt(eig(k_ssc, m_ssb)); fn_ssc = wn_ssc ./ (2.0 * pi); Tn_ssc = 1.0 ./ fn_ssc; [mode_c, lamda_c] = eig(k_ssc,m_ssb); M_C = mode_c' * m_ssb * mode_c; K_C = mode_c' * k_ssc * mode_c; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Isolation system %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Can mc = 0.3; kc = mc * wn_ssc(2)^2 - kb; %% Nonlinear spring e = 0.1e-3; kn = 10.0 * kc; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Superstructure plus isolation system %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% k_ssiso = [[kc + kb, -kb, 0.0, 0.0], [-kb, kb + k1, -k1, 0.0], [0.0, -k1, k1 + k2, -k2], [0.0, 0.0, -k2, k2]]; m_ssiso = diag([mc, mb, m1, m2]); wn_ssiso(:,1) = sqrt(eig(k_ssiso, m_ssiso)); fn_ssiso(:,1) = wn_ssiso(:,1) ./ (2.0 * pi); Tn_ssiso(:,1) = 1.0 ./ fn_ssiso(:,1); [mode_l, lamda_l] = eig(k_ssiso,m_ssiso); M_L = mode_l' * m_ssiso * mode_l; K_L = mode_l' * k_ssiso * mode_l; k_ssiso = [[kc + kn + kb, -kb, 0.0, 0.0], [-kb, kb + k1, -k1, 0.0], [0.0, -k1, k1 + k2, -k2], [0.0, 0.0, -k2, k2]]; m_ssiso = diag([mc, mb, m1, m2]); wn_ssiso(:,2) = sqrt(eig(k_ssiso, m_ssiso)); fn_ssiso(:,2) = wn_ssiso(:,2)./ (2.0 * pi); Tn_ssiso(:,2) = 1.0 ./ fn_ssiso(:,2); if VERBOSE fprintf('\nPhysical parameters of isolator:\n') mc mb cc cb kc kb kn

92

e wn_ssc fn_ssc Tn_ssc fprintf('\nModal parameters of 2-DOF superstructure plus isolation:\n') wn_ssiso fn_ssiso Tn_ssiso end wc = sqrt((kb+ kc)/mc); cc = 2* zeta_c* wc* mc; c_l(2:4,2:4) = c_b; c_l(1,2)= -cb; c_l(2,1)= -cb; c_l(1,1)= cc + cb; c = mode_l'* c_l* mode_l; for i = 1:4 zeta_ssiso(i) = c(i,i)/ (2* M_L(i,i) * wn_ssiso(i)); end %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Simulation %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% kn = 0.0; sim('eqvp', [], simoptions); zs_lin = zs; yb_lin = yb; yc_lin = yc; ys_lin(:,1) = zs(:,1)+ yb; ys_lin(:,2) = zs(:,2)+ yb; xs_lin = xs; xb_lin = xb; xc_lin = xc; zbc_lin = yb-yc; zsb_lin = zs; zsc_lin(:,1) = zs(:,1)+ yb -yc; zsc_lin(:,2) = zs(:,2)+ yb -yc; zbc_lin = yb -yc; xs_lin_t = xs_t; xb_lin_t = xb_t; xc_lin_t = xc_t; xs_lin_tt = xs_tt; xb_lin_tt = xb_tt; xc_lin_tt = xc_tt; U_lin = U; kn = 10.0 * kc; e = max(abs(yc_lin)/10) sim('eqvp', [], simoptions); zs_n = zs; yb_n = yb; yc_n = yc; ys_n(:,1) = zs(:,1)+ yb; ys_n(:,2) = zs(:,2)+ yb; xs_n = xs;

93

xb_n = xb; xc_n = xc; zsc_n(:,1) = zs(:,1)+ yb -yc; zsc_n(:,2) = zs(:,2)+ yb -yc; zbc_n = yb -yc; yc_n = yc; xs_n_t = xs_t; xb_n_t = xb_t; xc_n_t = xc_t; xs_n_tt = xs_tt; xb_n_tt = xb_tt; xc_n_tt = xc_tt; %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Modal Amplitudes %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% for i =1:length(tout) A_c(i,:) = ((mode_b' * m_ssb) *[yb_c(i),ys_c(i,1:2)]')'; end [mode, Value] = eig(k_ssb,m_ssb); IM = mode' * m_ssb; for i =1:length(tout) A_lin(i,:) = (IM*[zbc_lin(i),zsc_lin(i,1),zsc_lin(i,2)]')'; A_n(i,:) = (IM*[zbc_n(i),zsc_n(i,1),zsc_n(i,2)]')'; end %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Power Spectral Densities %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% num = tf/timestep; [Pys_c(:,1),ff] = periodogram(ys_c(1:num,1),[],num,2000); [Pys_c(:,2),ff] = periodogram(ys_c(1:num,2),[],num,2000); [Pyb_c,ff] = periodogram(yb_c(1:num),[],num,2000); [Pzsc_lin(:,1),ff] = periodogram(zsc_lin(1:num,1),[],num,2000); [Pzsc_lin(:,2),ff] = periodogram(zsc_lin(1:num,2),[],num,2000); [Pzbc_lin,ff] = periodogram(zbc_lin(1:num),[],num,2000); [Pyc_lin,ff] = periodogram(yc_lin(1:num),[],num,2000); [Pzsc_n(:,1),ff] = periodogram(zsc_n(1:num,1),[],num,2000); [Pzsc_n(:,2),ff] = periodogram(zsc_n(1:num,2),[],num,2000); [Pzbc_n,ff] = periodogram(zbc_n(1:num),[],num,2000); [Pyc_n,ff] = periodogram(yc_n(1:num),[],num,2000); %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Envelope of Absolute Control Force %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% [FF,Ti] = Force_Env(U,kn,tout); set(0,'DefaultAxesColorOrder',[0 0 1;1 0 0;0 1 0]) %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% figure(1) subplot(2,1,1) plot(tout, A_lin(:, 2),tout, A_lin(:, 3)) legend('A2','A3'); ylabel('Modal Amplitudes') title('Linear system'); subplot(2,1,2) plot(tout, A_n(:, 2),tout,A_n(:, 3)) legend('A2','A3'); ylabel('Modal Amplitudes')

94

title('Nonlinear system'); figure(3) subplot(4, 2, 1) plot(tout,zs_lin(:,2)-zs_lin(:,1),tout,zs_n(:,2)-zs_n(:,1)) ylabel('x2-x1 (m)') title('Displacements') subplot(4, 2, 3) plot(tout,zs_lin(:,1),tout,zs_n(:,1)) ylabel('x1-xb (m)') subplot(4, 2, 5) plot(tout,yb_lin,tout,yb_n) ylabel('xb-xg (m)') subplot(4, 2, 7) plot(tout,yc_lin,tout,yc_n) ylabel('xc-xg (m)') xlabel('t (s)') subplot(4, 2, 2) plot(tout,xs_lin_tt(:,2),tout,xs_n_tt(:,2)) ylabel('x2-tt (m/s/s)') %legend('linear','nonlinear'); title('Accelerations') subplot(4, 2, 4) plot(tout,xs_lin_tt(:,1),tout,xs_n_tt(:,1)) ylabel('x1-tt (m/s/s)') subplot(4, 2, 6) plot(tout,xb_lin_tt,tout,xb_n_tt) ylabel('xb-tt (m/s/s)') subplot(4, 2, 8) plot(tout,xc_lin_tt,tout,xc_n_tt) ylabel('xc-tt (m/s/s)') xlabel('t (s)') figure(4) subplot(2, 2, 1) semilogy(ff(1:240),[Pzsc_lin(1:240,2),Pzsc_n(1:240,2)]) ylabel('PSD of zc2 (m2/Hz)') subplot(2, 2, 2) semilogy(ff(1:240),[Pzsc_lin(1:240,1),Pzsc_n(1:240,1)]) legend('linear','nonlinear'); ylabel('PSD of zc1 (m2/Hz)') subplot(2, 2, 3) semilogy(ff(1:240),[Pzbc_lin(1:240),Pzbc_n(1:240)]) ylabel('PSD of zbc (m2/Hz)') subplot(2, 2, 4) semilogy(ff(1:240),[Pyc_lin(1:240),Pyc_n(1:240)]) ylabel('PSD of yc (m2/Hz)') xlabel('frequency (Hz)') figure(5) plot(tout,Fn) ylabel('F(u) (N)') xlabel('t (s)') hold on; plot(Ti,FF,'r') legend('Control Force','Force Envelope'); hold off;

95

Appendix C Matlab Code: Full-sine Ground Velocity Pulse function p = groundvel(t) global vpulse Vg Tg zeta_g switch vpulse case 'typeA' if 0.0 < t & t < Tg p = Vg * sin(pi * t / Tg); else p = 0.0; end case 'typeB' if 0.0 < t & t < Tg p = Vg * sin(2* pi * t / Tg);%+ Vg * sin(120.7095/2*t)/ 4; else p = 0.0; end case 'typeD' wd_g = 2*pi/Tg * sqrt(1 - zeta_g^2); if t >0 &t < Tg p = Vg * exp(-zeta_g *2*pi/Tg * t) * sin(wd_g * (t - 0)); else p = 0.0; end end

96

Appendix D Matlab Code: Envelope of the Control Force

function [FF,Ti] = Force_Env(U,kn,tout) count1 = 0; count2 = 0; for i = 1:length(U)-1 if U(i)==0 & U(i+1)~=0 count1 = count1 + 1; arohead(count1) = i+1; end if U(i)~= 0 & U(i+1)==0 count2 = count2 + 1; aroend(count2) = i; end end if U(length(U))~=0 count2 = count2 + 1; aroend(count2) = length(U); end for j = 1:length(arohead) for i = 1:length(U) if i>=arohead(j) & i<=aroend(j) [F,I] = max(abs(U(arohead(j):aroend(j)))); FF(j) = F * kn; Ti(j) = tout(I + arohead(j)-1); end end end

97

Appendix E Simulink Model: SDOF Oscillator

Simulink Model of SDOF oscillator used for calculating response spectra

98

Appendix F Matlab Code: Response Spectra

m = 1; zeta1 = 0.005; zeta2 = 0.02; tf = 44.905; timestep = 0.005; simoptions = simset('FixedStep', timestep); simoptions = simset(simoptions, 'Solver', 'ode4'); Tmax = 8.0; dT = 0.005; Tvec = 0.005 : dT : Tmax; for i = 3:length(Tvec) T = Tvec(i) k = (2 * pi)^2 * m / T^2; c = 2 * sqrt(k * m) * zeta1; sim('oscillator_a', [], simoptions); sd_05(i) = max(abs(disp)); sv_05(i) = max(abs(vel)); sa_05(i) = max(abs(acc)); c = 2 * sqrt(k * m) * zeta2; sim('oscillator_a', [], simoptions); sd_2(i) = max(abs(disp)); sv_2(i) = max(abs(vel)); sa_2(i) = max(abs(acc)); % c = 2 * sqrt(k * m) * zeta1; % sim('oscillator_v', [], simoptions); % sd_05(i) = max(abs(disp)); % c = 2 * sqrt(k * m) * zeta2; % sim('oscillator_v', [], simoptions); % sd_2(i) = max(abs(disp)); end figure(2) plot(Tvec,sv_2) ylabel('velocity,cm/s') xlabel('period,sec') title('Response Spectra') figure(3) plot(Tvec,sd_2) ylabel('displacement,cm') xlabel('period,sec') title('Response Spectra') load Erz_ns.mat figure(1) subplot(3,2,1) plot(1./ERZ_NS_005(11:112,2),ERZ_NS_005(11:112,6)*9.81) legend('Recorded','Approximated') ylabel('Acceleration') Title('Damping 0.5%') subplot(3,2,3) plot(11./ERZ_NS_005(11:112,2),ERZ_NS_005(11:112,4)/100) ylabel('Velocity')

99

subplot(3,2,5) plot(1./ERZ_NS_005(11:112,2),ERZ_NS_005(11:112,3)/100) ylabel('Displacement') xlabel('T (s)') subplot(3,2,2) plot(1./ERZ_NS_020(11:112,2),ERZ_NS_020(11:112,6)*9.81) legend('Recorded','Approximated') ylabel('Acceleration') Title('Damping 2%') subplot(3,2,4) plot(1./ERZ_NS_020(11:112,2),ERZ_NS_020(11:112,4)/100) ylabel('Velocity') subplot(3,2,6) plot(1./ERZ_NS_020(11:112,2),ERZ_NS_020(11:112,3)/100) ylabel('Displacement') xlabel('T (s)')

100

Appendix G Simulink Model: Conventional Isolation System

Simulink model of the 3DOF conventional base isolation system

101