Eerc 05 01

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 244

REPORT NO.

EERC 2005-01 EARTHQUAKE ENGINEERING RESEARCH CENTER

ESTIMATING THE SEISMIC RESPONSE


OF BASE-ISOLATED BUILDINGS
INCLUDING TORSION, ROCKING
AND AXIAL-LOAD EFFECTS

By

Keri L. Ryan
Anil K. Chopra

A Report on Research Partially Conducted Under


Grants No. CMS-9812414 and 9812531 from the
National Science Foundation

COLLEGE OF ENGINEERING

UNIVERSITY OF CALIFORNIA, BERKELEY


Estimating the Seismic Response of Base-Isolated
Buildings Including Torsion, Rocking, and Axial-Load
Effects

By

Keri L. Ryan
Department of Civil and Environmental Engineering
Utah State University
Logan, UT

and

Anil K. Chopra
Earthquake Engineering Research Center
University of California, Berkeley

A Report on Research Conducted under National Science Foundation Grants


CMS-9812414 and CMS-9812531, NSF Graduate Student Fellowship, and a
California State Legislative Grant

EERC Report 2005/01


Earthquake Engineering Research Center
University of California, Berkeley
June 2005
ABSTRACT

Although base-isolated buildings are designed using dynamic analysis, current code provi-

sions require that design values of the earthquake-induced lateral deformations and forces do not

fall below minimum values determined by a simpler procedure. Isolation systems are strongly

nonlinear, but this procedure estimates the isolator deformation by approximate methods based

on equivalent-linear systems.

Therefore, a rational procedure to estimate the peak response of a lead-rubber bearing

isolation system, based on rigorous nonlinear analysis, is developed in Chapter 3. The procedure

offers an alternative to the iterative equivalent-linear methods used by current U.S. building

codes. The governing equation is reduced to a form such that the median normalized deforma-

tion of the system due to an ensemble of ground motions with given corner period Td is found

to depend on only two parameters: the natural isolation period, defined from the post-yield

stiffness, and the normalized strength, or strength normalized by peak ground velocity. The

dispersion of normalized deformation for an ensemble of ground motions is shown to be small,

implying that the median normalized deformation is a meaningful estimate of response. The

simple trends shown by the median normalized deformation led to the development of suitable

design equations for isolator deformation. These design equations reflect a 13% increase when

the excitation is biaxial (two lateral components of ground motion) compared to single com-

ponent excitation. For comparison, deformations estimated by the equivalent-linear method

are unconservative by up to 50% compared to those found from the more accurate nonlinear

spectrum, and building codes include at most a 4.4% increase for a second component.

Subsequent chapters extend the procedure to more complex models of the isolated build-

ing. Chapter 4 extends the procedure to asymmetric-plan systems, deriving equations to es-

timate the peak deformation among all isolators in an asymmetric building. A variety of

asymmetric-plan systems are represented by an idealized, rectangular system, where eccentric-

ity is introduced by varying the stiffnesses and strengths of individual isolators. The idealized

system is shown to approximate the peak deformation in asymmetric-plan systems with less

than 1% error.

i
In addition to the isolation period and normalized strength, the median normalized de-

formation of asymmetric-plan systems depends on the torsional-to-lateral frequency ratio and

the normalized stiffness eccentricity. However, the influence of each of these parameters, ex-

cept for eccentricity, on the deformation ratio – the ratio of peak deformations in asymmetric

and corresponding symmetric systems – is shown to be negligible. Therefore, the equation

developed to estimate the largest deformation ratio among all isolators depends only on the

stiffness eccentricity and the distance from the center of mass to the outlying isolator. This

equation, multiplied by the design equation for symmetric systems in Chapter 3, gives the peak

deformation of asymmetric systems. This design equation conservatively estimates the peak

deformation among all isolators, but is generally within 10% of the ‘exact’ value.

In Chapter 5, the procedure is applied to buildings isolated with the friction pendulum

system. Friction pendulum (FP) isolators can also be modeled by a nonlinear force-displacement

relation, but with yield displacement on the order of 0.05 cm compared to 1 cm for rubber bear-

ings. Design equations for the peak slip, or displacement, in FP isolators reflect the significant

– 20 to 38% – increase when the excitation is biaxial. Equivalent-linear methods are shown to

underestimate by up to 30% the exact median displacement determined by nonlinear response

history analysis for one component of ground motion, and building codes include at most a

4.4% increase for a second component.

Discussed in Chapter 6, existing models for isolation bearings neglect certain aspects

of their response behavior. For instance, rubber bearings have been observed to decrease in

stiffness with increasing axial load, as well as soften in the vertical direction at large lateral

deformations. The yield strength of lead-rubber bearings has also been observed to vary with

axial load, such that a lightly loaded bearing may not achieve its theoretical strength.

A series of bearing models are developed to include these observed behaviors, referred to

as “axial-load effects”. The models are considered to be most accurate for lead-rubber bear-

ings. Extending an existing two-spring model (shear spring plus rotational springs) developed

from linear stability theory of multi-layer bearings, the constant-strength model is achieved

by incorporating a nonlinear constitutive model for the shear spring. Numerically, this model

ii
is implemented by solving its equilibrium and kinematic equations (a system of five nonlinear

equations) by Newton’s method for the bearing forces, and taking differentials of these equations

to derive the instantaneous bearing stiffness matrix. An empirical equation is developed that

can be calibrated to match the experimentally observed varying yield strength in lead-rubber

bearings; this effect is included in the variable-strength model. The response behavior of these

new models is confirmed by comparison with unpublished experimental data.

Not considered in previous chapters, the peak axial forces in individual isolators need

also be estimated for design due to code testing requirements. For this purpose, the isolated

structure model is modified in Chapter 7 to include rocking about one axis and incorporate the

improved bearing models of Chapter 6. However, rocking of the structure and bearing axial-

load effects are found to have little influence on the peak lateral bearing deformation, whereas

even if rocking is neglected entirely, median response spectra are within, perhaps, 10% of those

when rocking and axial-load effects are included. Furthermore, bearing axial-load effects can

usually be neglected in determining the maximum and minimum bearing axial forces. Based

on analyses that indicate more than 10% error in neglecting such effects, the variable strength

model is recommended only when the normalized strength exceeds 0.75 and the rocking-to-

vertical frequency ratio is less than 0.75.

The design equation for peak lateral deformation is updated to include the slight influence

of rocking, and a design equation for peak axial force is developed for the first time. It is not

surprising that the peak lateral deformation follows trends from earlier chapters and does not

depend on any additional parameters. The design equation for axial force depends on the

isolation period, normalized strength, and rocking-to-vertical frequency ratio. It is shown how

to use this design equation to predict and subsequently eliminate bearing tension early in

the design process. Decreasing the normalized strength of the system is the simplest way to

eliminate tension, but at the expense of larger lateral deformation in the bearings. Since the

peak axial forces do not vary with the vertical-to-lateral frequency ratio, the use of high shape

factor bearings does not provide the expected benefit of avoiding instability.

Chapter 8 integrates all previous modeling concepts in a final three-dimensional (3D)

iii
analysis of the response of the isolated block. The concept of accidental torsion due to axial-

load effects is introduced, where variation of the axial forces and hence stiffnesses and strengths

of individual bearings can induce a time-varying eccentricity. However, accidental torsion in

the isolation system from this source is found to be insignificant. Design equations are updated

to estimate the peak bearing deformation and axial forces in both nominally symmetric and

asymmetric-plan isolation systems due to biaxial excitation. Determined from these design

equations, the peak deformation in a symmetric system is again found to increase by 13%

when the excitation is biaxial, and the further increase in deformation for asymmetric-plan

systems again depends on the eccentricity and the distance from the center of mass to an

outlying corner. Additional parameters that relate to the distance from the center of mass to

the outermost edge in each rocking direction are applied to the axial force design equation for

symmetric-plan systems. These parameters provide only marginal improvement to the axial

force estimates in symmetric systems, but significant improvement to the force estimates in

asymmetric-plan systems, which otherwise increase compared to symmetric systems by a factor

that depends only on eccentricity.

Although not obviously applicable when rocking of the system is included, the earlier

approach of representing general asymmetric-plan systems by an idealized, rectangular plan

is justified since the design equations are as accurate for general asymmetric-plan systems as

for the idealized systems for which they were developed. However, these design equations are

overall less effective than in previous chapters, perhaps indicating the limit of application of the

simplified procedure. While equations to estimate deformations are simple and accurate, those

estimating axial forces are complicated and can err on the order of -25 to 25% for asymmetric-

plan systems.

iv
ACKNOWLEDGEMENTS

This research is funded by National Science Foundation Grants CMS-9812414 and CMS-

9812531 and a Graduate Student Fellowship to the first author, as well as a California state

legislative grant. We gratefully acknowledge this support. Except for editorial changes, this

report is the same as the doctoral dissertation of Keri Lynn Ryan at the University of California,

Berkeley. We especially thank Professor Jim Kelly for his guidance throughout the project. We

also wish to thank Professor Nicos Makris for his guidance in defining the scope of the project

and Professor Douglas Dreger for reviewing the manuscript.

v
TABLE OF CONTENTS

1 INTRODUCTION 1

2 GROUND MOTIONS AND STATISTICAL ANALYSIS 5

2.1 Ground Motions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.2 Statistical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 ESTIMATING ISOLATOR DEMANDS BASED ON NONLINEAR ANALY-

SIS 8

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3.2 System and Governing Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3.2.1 System Considered . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3.2.2 Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.3 System Parameters and Normalization . . . . . . . . . . . . . . . . . . . . . . . . 11

3.3.1 Initial Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.3.2 Normalized Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3.3.3 Isolator Deformation and Force . . . . . . . . . . . . . . . . . . . . . . . . 17

3.4 Comparison to Nonlinear Structural Analysis . . . . . . . . . . . . . . . . . . . . 18

3.5 Earthquake Response Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.5.1 Normalized Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.5.2 Design Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.6 Governing Equations for Bidirectional Excitation . . . . . . . . . . . . . . . . . . 23

3.6.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3.6.2 Normalized Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3.7 Design Equations for Bidirectional Excitation . . . . . . . . . . . . . . . . . . . 25

3.8 Application to Analysis and Design . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.8.1 Estimating Isolator Deformation . . . . . . . . . . . . . . . . . . . . . . . 28

3.8.2 Selecting Isolator Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

vi
3.9 Comparison to Code Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.9.1 Estimating Isolator Deformation . . . . . . . . . . . . . . . . . . . . . . . 31

3.9.2 Evaluation of Equivalent Linear Procedure . . . . . . . . . . . . . . . . . 32

3.9.3 Extension to Bidirectional Excitation . . . . . . . . . . . . . . . . . . . . 33

3.10 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

Appendix 3A: Regression Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3A.1 R2 statistic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3A.2 Error Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3A.3 Evaluation of β Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . 39

Appendix 3B: Peak Force due to Bidirectional Excitation . . . . . . . . . . . . . . . . 42

Appendix 3C: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4 ESTIMATING THE PEAK DEFORMATION IN AN ASYMMETRIC-PLAN

SYSTEM 46

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4.2 Asymmetric Base-Isolated System . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2.1 System Considered . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2.2 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.2.3 Characteristics of Response Histories of Asymmetric System . . . . . . . 51

4.2.4 Peak Response to Bidirectional Excitation . . . . . . . . . . . . . . . . . . 53

4.2.5 Normalized Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . 54

4.2.6 Computation of Median Response . . . . . . . . . . . . . . . . . . . . . . 55

4.3 Modeling Asymmetric-Plan Systems . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.3.1 Parameter Ranges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.3.2 Geometry and Planwise Layout of Isolators . . . . . . . . . . . . . . . . . 58

4.3.3 Two-way Asymmetric Systems . . . . . . . . . . . . . . . . . . . . . . . . 60

4.4 Median Response of Asymmetric Systems . . . . . . . . . . . . . . . . . . . . . . 62

4.5 Estimation of Peak Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.5.1 Equations to Estimate uro . . . . . . . . . . . . . . . . . . . . . . . . . . 65

vii
4.5.2 Equations to Estimate Corner Deformation . . . . . . . . . . . . . . . . . 67

4.5.3 Application to Various Plans . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.5.4 Comparison to IBC Equation . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

Appendix 4A: Maximization of ur (t) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

Appendix 4B: Regression Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4B.1 Considerations for Choosing Model . . . . . . . . . . . . . . . . . . . . . . 77

4B.2 Selecting a Model for ûro . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4B.3 Selecting a Model for ûco . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Appendix 4C: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

5 ESTIMATING THE PEAK DISPLACEMENT OF A FRICTION PENDU-

LUM ISOLATOR 86

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5.2 System, Governing Equations and Normalization . . . . . . . . . . . . . . . . . . 86

5.2.1 System Considered . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5.2.2 Normalized Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . 88

5.2.3 Dispersion of Normalized Displacement . . . . . . . . . . . . . . . . . . . 89

5.3 Peak Earthquake Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.3.1 Median Response Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.3.2 Design Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5.4 Evaluation of Equivalent Linear Procedure . . . . . . . . . . . . . . . . . . . . . . 93

5.4.1 Extension to Bidirectional Excitation . . . . . . . . . . . . . . . . . . . . 96

5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Appendix 5A: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

6 MODEL FOR AXIAL-LOAD EFFECTS IN LEAD-RUBBER BEARINGS 100

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

6.2 Stability Analysis of Multi-Layer Bearings . . . . . . . . . . . . . . . . . . . . . . 101

viii
6.2.1 Approximate Force-Deformation Relation Based on Linear Two-Spring

Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

6.2.2 Effect of Bearing Shape Factors and Compressibility . . . . . . . . . . . . 107

6.3 Variation of Yield Strength with Axial Load . . . . . . . . . . . . . . . . . . . . 108

6.4 Verification of Theory Based on Experimental Data . . . . . . . . . . . . . . . . 109

6.4.1 Axial-Load Varied Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6.4.2 Vertical Characteristic Tests and Offset Tests . . . . . . . . . . . . . . . 112

6.5 Nonlinear Extension of Two-Spring Model . . . . . . . . . . . . . . . . . . . . . 114

6.5.1 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

6.5.2 Numerical Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.6 Observed Response of New Bearing Models . . . . . . . . . . . . . . . . . . . . . 119

6.6.1 Lateral and Vertical Force-Deformation Trends . . . . . . . . . . . . . . . 119

6.6.2 Response to a Seismic Pulse . . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

Appendix 6A: Interpretation of Experimental Data . . . . . . . . . . . . . . . . . . . 127

Appendix 6B: Numerical Implementation of Nonlinear Bearing Model . . . . . . . . . 134

6B.1 Unidirectional Rate-Independent Plasticity . . . . . . . . . . . . . . . . . 134

6B.2 Return Mapping Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . 134

6B.3 Return Mapping for Variable-Strength Model . . . . . . . . . . . . . . . . 135

6B.4 Complete Bearing Routine . . . . . . . . . . . . . . . . . . . . . . . . . . 136

Appendix 6C: Energy Conservation in Coupled Linear Model . . . . . . . . . . . . . . 137

Appendix 6D: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7 ESTIMATING ISOLATOR DEFORMATIONS AND FORCES CONSIDER-

ING LATERAL-ROCKING RESPONSE 143

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

7.2 Governing Equations and Parameters . . . . . . . . . . . . . . . . . . . . . . . . . 144

7.2.1 System Considered . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

7.2.2 Bearing Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

ix
7.2.3 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

7.2.4 System Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

7.2.5 Rigid Structure Approximation . . . . . . . . . . . . . . . . . . . . . . . 151

7.3 Response Histories for Different Models . . . . . . . . . . . . . . . . . . . . . . . 152

7.4 Normalization and Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

7.4.1 Normalized Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

7.4.2 Computation of Median and Dispersion . . . . . . . . . . . . . . . . . . . 165

7.4.3 Dispersion of Various Responses . . . . . . . . . . . . . . . . . . . . . . . 165

7.5 Median Response Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

7.5.1 Median Response Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

7.5.2 Influence of Frequency Ratios . . . . . . . . . . . . . . . . . . . . . . . . 169

7.5.3 Significance of Bearing Axial-Load Effects . . . . . . . . . . . . . . . . . 172

7.6 Design Equations to Estimate Peak Response . . . . . . . . . . . . . . . . . . . . 175

7.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

Appendix 7A: Parameter Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

Appendix 7B: Rigid Structure Approximation . . . . . . . . . . . . . . . . . . . . . . 182

Appendix 7C: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

8 ESTIMATING BEARING DEFORMATIONS AND FORCES IN SYMMETRIC-

AND ASYMMETRIC-PLAN SYSTEMS INCLUDING TORSION AND ROCK-

ING 187

8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

8.2 System Model and Governing Equations . . . . . . . . . . . . . . . . . . . . . . 188

8.2.1 System Considered . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

8.2.2 Bearing Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

8.2.3 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

8.2.4 Normalized Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

8.3 Representative System and Parameter Selection . . . . . . . . . . . . . . . . . . . 194

8.3.1 Plan Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

x
8.3.2 Uncoupled Nonlinear Bearing Model . . . . . . . . . . . . . . . . . . . . . 196

8.3.3 System Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198

8.4 Symmetric Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

8.4.1 Median Response Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

8.4.2 Accidental Torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

8.4.3 Influence of Rocking-to-Vertical Frequency Ratios . . . . . . . . . . . . . 203

8.4.4 Increase in Response due to Bidirectional Excitation . . . . . . . . . . . . 204

8.4.5 Design Equations to Estimate Peak Response . . . . . . . . . . . . . . . 205


   
dyz
8.4.6 Interpretation of Regression Parameters ymax
rx rx and xmax dxz
ry ry . . . 208

8.5 Asymmetric-Plan Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211

8.5.1 Trends for Deformation and Force-Increment Ratios . . . . . . . . . . . . 211

8.5.2 Design Equations to Estimate Peak Response . . . . . . . . . . . . . . . . 212

8.5.3 Application to Three Asymmetric-Plan Systems . . . . . . . . . . . . . . 215

8.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

Appendix 8A: Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222

References 225

xi
1: INTRODUCTION

Although base-isolated buildings are designed using dynamic analysis, the current Inter-

national Building Code (IBC) provisions require that design values of the earthquake-induced

lateral deformations and forces do not fall below minimum values determined by a simpler

procedure [1, 2]. This procedure assumes that the response is dominated by the first mode,

and estimates the isolator deformation from a design spectrum by approximate methods based

on equivalent-linear systems. Isolation systems are strongly nonlinear, however, due to added

energy dissipation devices, such as high-damping fillers, lead cores, and friction devices [2, 3].

Therefore, use of the simple procedure requires defining an equivalent-linear system with period

and damping ratio based on the nonlinear force-deformation relation of the isolator. Because

the period and damping ratio of the equivalent-linear system depend on the absolute maximum,

or peak, deformation, this procedure is iterative.

The two horizontal components of ground motion acting simultaneously should be con-

sidered to determine the peak deformation of the isolation system that can occur in any lateral

direction. The IBC is ambiguous, making no mention of bidirectional excitation in its static

lateral response procedure. The usual 100 + 30 rule – the response to 100% of the ground

motion applied in the most critical direction and 30% applied in the transverse direction –

is used for response spectrum analysis, resulting in at most a 4.4% deformation increase for

bidirectional excitation.

Furthermore, lateral deformations are larger in the corner isolators of asymmetric-plan

buildings due to torsion. To account for this additional deformation, the IBC increases the

design displacement by a factor that is derived from static analysis of an equivalent-linear

system. Again, the justification for this increase is questionable since isolation systems are

strongly nonlinear.

Clearly, preliminary design procedures for base-isolated buildings, as presently applied in

the IBC, leave room for improvement. A rational approach to estimate isolator deformations

should be developed that accounts for the nonlinear force-deformation relation of the isolators

and bidirectional excitation.

1
Initially, rocking or overturning of the structure was not a concern for designing isolation

systems, which were traditionally implemented in short, squat structures. However, recently

isolation has been extended to taller buildings, such as the 32-story LA City Hall, the 18-story

Oakland City Hall, and numerous projects in Japan [4, 5].

A number of approaches have been taken to accommodate the time-varying axial forces

in an isolation system. Due to uncertainty whether isolation bearings and sliding isolation

systems could sustain tension, unrestrained uplift of the structure was allowed in early designs.

Concern about this uplift led researchers to experiment with various restraint systems, which

both limited uplift of the structure and provided tensile resistance through an alternate load

path in the isolation system [6, 7, 8]. It was also proposed that uplift in sliding isolation systems

be accommodated at the foundation level rather than the isolation level [9].

Even with eventual evidence that elastomeric bearings can sustain moderate tensile forces

[10, 11], designers still wish to minimize or avoid tension altogether. Sometimes extreme mea-

sures are taken to meet this goal, such as the “loose-bolt” connection used in the designs of LA

City Hall (retrofit) [12], San Bernadino Medical Center and LAC/USC Medical Center. This

questionable connection detail delays engagement of the bearing in tension by allowing a small

amount of unrestrained uplift, but forces the bolt to transfer shear while acting as a cantilever

element [13].

Underlying the desire to eliminate tension in isolation bearings is the following: First,

the tensile force that can be withstood by elastomeric bearings without excessive damage is

unknown. Second, for lead-rubber (LRB) bearings, the absence of confining pressure when the

bearing is in tension appears to compromise the lateral resistance and diminish the energy dis-

sipation that is relied on to control lateral deformation. Third, satisfying the IBC requirement

that the bearings be tested laterally under the maximum and minimum axial loads predicted

by dynamic analysis is difficult since most testing machines cannot apply simultaneous tension

and shear [2]. Thus, accurate determination of overturning forces on the isolators is essential

to the design process.

To further complicate this matter, the behavior of isolation bearings appears to change

2
considerably when the variation of axial forces on the bearings is taken into account [13, 14,

15, 16, 17, 18]. Existing nonlinear bearing models do not account for the influence of axial

forces on the response of the bearing, nor allow for its consideration in dynamic analysis of the

system. If the behavior of the isolation bearings depends on axial force, among other things,

the time-variation of the strengths and stiffnesses of individual bearings can induce torsion in

a symmetric-plan system, known as accidental torsion.

The overall objective of this dissertation is to develop a procedure suitable for estimating

the peak lateral deformation and axial forces – needed for preliminary design of the isolation

system – for a given design spectrum. Although the structure is treated as rigid, an advanced

model of the isolation system is used to achieve accuracy. This includes characterization of

each individual isolator by a new nonlinear model that incorporates the effects of axial force

on the bearing response. Design equations to estimate the responses of interest are developed

by regression analysis of the peak response computed by nonlinear analysis to an ensemble of

motions representative of the spectrum. The equations of motion are normalized such that the

deformation response of the system is insensitive to ground motion intensity, thereby minimizing

its statistical variation over the ground motion ensemble. This makes the design equations more

generally applicable over a wide range of ground motion intensity.

Chapter 2 presents the ground motion ensemble and statistical analysis equations that will

be used throughout the dissertation. Using a lateral force-deformation relation representative of

LRB bearings, the normalization procedure is first developed in Chapter 3 for an isolated block

subjected to a single component of excitation, and then extended to bidirectional excitation.

The equivalent-linear procedure used by the IBC is shown, for a given design spectrum, to be

generally unconservative in estimating the peak lateral deformation. Chapter 4 extends the

normalization procedure to asymmetric-plan buildings, and demonstrates that the increase in

lateral deformation due to plan-asymmetry can be accounted for by a simple factor. This factor

is applicable to a wide range of systems, due to special properties of base-isolated buildings

that minimize the dependence of the peak response on the plan layout and individual bearing

properties. In Chapter 5, the procedure is specialized for friction-pendulum (FP) isolation

3
systems by changing the yield deformation in the force-deformation relationship.

A suitable model to account for the interaction between lateral response and axial force

in LRB bearings is developed in Chapter 6, as well as a numerical implementation of this model

for dynamic analysis. Using the new bearing model, the isolated structure is modified to include

rocking about one axis in Chapter 7. Design equations for the lateral deformation are updated

for this planar rocking analysis, and design equations to estimate the maximum and minimum

axial forces are developed for the first time. In Chapter 8, the lateral-rocking analysis of

Chapter 7 and lateral-torsional analysis of Chapter 4 are integrated in a final three-dimensional

(3D) analysis of the response of the isolated block. Design equations are extended to estimate

the peak bearing deformation and axial forces in both nominally symmetric – leading to the

concept of accidental torsion due to axial-load effects – and asymmetric-plan isolation systems.

Estimating the axial forces in asymmetric-plan buildings proves to be difficult, perhaps signaling

the limit of application of this simplified procedure.

4
2: GROUND MOTIONS AND STATISTICAL ANALYSIS

2.1 Ground Motions

An ensemble of 20 ground motions has been selected for this study, recorded from four California

earthquakes: (1) 1971 San Fernando (Mw = 6.6), (2) 1987 Superstition Hills (Mw = 6.7),

(3) 1989 Loma Prieta (Mw = 6.9), and (4) 1994 Northridge (Mw = 6.7). These motions,

obtained from PEER Strong Motion Database [19], are listed in Table 2.1. The data includes

the recording site, the earthquake, closest distance of the site to fault rupture, and peak ground

acceleration (ügo ), velocity (u̇go ), and displacement (ugo ) for the two horizontal components

of motion. For each motion, the component having the larger peak ground velocity (PGV) is

defined as the “strong” component, and the other as the “weak” component. Designated the

Large Magnitude Small Distance (LMSR) ensemble, these motions are representative of ground

shaking from a large magnitude earthquake at a site near the fault rupture. Recorded on firm

soil (USGS site class C) in the far-field of these earthquakes, they are broad frequency band

excitations. This LMSR ensemble is adapted from an ensemble of 20 single-component motions

first used by H. Krawinkler [20].

Figure 2.1 shows the median linear response spectra for both the strong and weak-

component ensembles as four-way log plots. In constructing these spectra, each stronger com-

ponent of motion was scaled to a PGV of 35 cm/s, and the same scale factor was applied to the

weaker component; which amplifies the median velocities of 29.4 and 20.9 cm/s before scaling to

35 and 24.9 cm/s for the strong and weak-component ensembles, respectively. The spectra have

been subdivided into spectral regions: acceleration-sensitive region Tn < Tc , velocity-sensitive

region Tc < Tn < Td , and displacement-sensitive region Tn > Td [21, Section 6.8]; Fig. 2.1 also

shows the idealized spectra upon which these spectral regions are based.

2.2 Statistical Analysis

By the procedures of the following section, the dynamic response of an isolation system to

each motion in the ensemble was determined and from this the median value x̂ and dispersion

measure δ were computed. For n observed values xi , the median, defined as the geometric

5
Table 2.1: Characteristics of excitations in the LMSR ground motion ensemble.

Strong Component Weak Component


No. Site Eq R ügo u̇go ugo ügo u̇go ugo
(km) (g) (cm/s) (cm) (g) (cm/s) (cm)
1 LA - Hollywood Stor Lot 1 21.2 0.210 18.9 12.4 0.174 14.9 6.3
2 Brawley 2 18.2 0.116 17.2 8.6 0.156 13.9 5.3
3 El Centro - ICC 2 13.9 0.357 46.4 17.6 0.258 40.9 20.2
4 Plaster City 2 21.0 0.186 20.6 5.4 0.121 9.5 1.9
5 Westmorland Fire Sta 2 13.3 0.211 31.0 20.3 0.172 23.5 13.1
6 Agnews State Hospital 3 28.2 0.172 26.0 12.6 0.159 17.6 9.8
7 Capitola 3 14.5 0.529 35.0 9.1 0.443 29.3 5.5
8 Gilroy Array #3 3 14.4 0.367 44.7 19.3 0.555 35.7 8.2
9 Gilroy Array #4 3 16.1 0.417 38.8 7.1 0.212 37.9 10.1
10 Gilroy Array #7 3 24.2 0.323 16.6 3.3 0.225 16.4 2.5
11 Hollister City Hall 3 28.2 0.215 45.0 26.1 0.246 38.5 17.7
12 Hollister Diff. Array 3 25.8 0.269 43.9 18.5 0.278 35.6 13.0
13 Sunnyvale - Colton 3 28.8 0.207 37.3 19.1 0.209 36.0 16.9
14 Canoga Park - Topanga 4 15.8 0.420 60.8 20.2 0.356 32.1 9.1
15 Faring Rd 4 23.9 0.242 29.8 4.7 0.273 15.8 3.3
16 LA - Fletcher 4 29.5 0.240 26.2 3.6 0.162 10.7 2.9
17 Glendale - Las Palmas 4 25.4 0.357 12.3 1.9 0.206 7.4 1.8
18 LA - Hollywood Storage 4 25.5 0.358 27.5 3.0 0.231 18.3 4.8
19 La Crescenta - New York 4 22.3 0.178 12.5 1.1 0.159 11.3 3.0
20 Northridge - Saticoy 4 13.3 0.477 61.5 22.1 0.368 28.9 8.4

mean, is
 n 
i=1 ln xi
x̂ = exp (2.1)
n
and the dispersion measure is
 n 1/2
− ln x̂)2
i=1 (ln xi
δ= (2.2)
n−1
For small values, δ is close to the coefficient of variation of x. The median plus one standard

deviation (median+σ), which corresponds to a non-exceedance probability of 84.1%, is given

by

x84% = x̂ exp δ (2.3)

These definitions assume the data is sampled from a lognormal distribution, which has been

found to be realistic of earthquake response of systems.

6
(a) Strong-Component Ensemble (b) Weak-Component Ensemble
Acc Sensitive Vel Sens Disp Sensitive Acc Sensitive Vel Sens Disp Sensitive
2 2
)

)
10 10
m

m
10

10
(c

(c
A

A
D

D
(g

(g
10

10
)

)
0

0
1

1
V (cm/s)

V (cm/s)

1 1
10

10
10 10
1

1
0.

0.
01

01
1

1
0.

0.
0 0
0.

0.

10 10
Td =2.06 s

Td =2.35 s
1

1
Tc =.45 s

Tc =.39 s

−1 −1
10 10
−2 −1 0 1 2 −2 −1 0 1 2
10 10 10 10 10 10 10 10 10 10
Tn (sec) Tn (sec)

Figure 2.1: Median linear response spectra for 5% damping (a) strong-component ensemble,
median ügo = 0.324 g, u̇go = 35 cm/s and ugo = 10.2 cm, and (b) weak-component ensemble,
median ügo = 0.274 g, u̇go = 24.9 cm/s and ugo = 7.6 cm. Also shown are idealized spectra
with spectral regions.

7
3: ESTIMATING ISOLATOR DEMANDS BASED ON NONLINEAR
ANALYSIS

3.1 Introduction

Preliminary analysis is an important step in the design of a base-isolated building. How, then,

given the nonlinear behavior typical of most isolation systems, to estimate the peak lateral

deformation of the isolators based on a specific design ground motion or design spectrum? This

deformation estimate could be based on nonlinear analysis due to a single ground motion but

would be sensitive to the particular motion selected, unlike deformation determined from a

design spectrum. A better estimate could be obtained from nonlinear analyses to an ensemble

of ground motions that represent the design spectrum, but a preliminary design procedure that

requires multiple nonlinear analyses is impractical. An additional difficulty is that the response

of a nonlinear isolation system is much more sensitive to ground motion intensity than a linear

system; thus the ground motions in an ensemble must be of similar intensity for the analyses

to be meaningful.

Unlike building frames that resist lateral forces primarily in one direction, the peak de-

formation in any direction – due to simultaneous application of two lateral components of

excitation – controls the design of the isolation system. Furthermore, isolator nonlinearity

causes bidirectional interaction consistent with the assumption of a circular interaction surface,

as shown by extensive testing of both high damping and lead-rubber bridge bearings [22]. An-

alytical models that included this interaction usually estimated the measured response of an

isolated block shaken by various ground motions more accurately than those that neglected it

[22]. Bidirectional excitation and bidirectional yield-interaction should be considered to achieve

a realistic estimate of deformations and forces for design.

The objective of this chapter is to develop a procedure suitable for estimating isolator

deformations for a given design spectrum. The procedure developed is based on rigorous non-

linear analysis to an ensemble of motions representative of the spectrum. The procedure is

effective because the governing equation for the system is rewritten such that its normalized

8
fb
(a) (b)
Q kb
m ub (t) ki
ub
kb , Q

üg (t)

Figure 3.1: (a) Single isolator supporting a rigid mass, and (b) force-deformation relation for
a bilinear isolator.

deformation is insensitive to ground motion intensity, and statistical variation of the normalized

deformation to an ensemble of ground motions is minimized. The procedure, initially devel-

oped for a mass-isolator system subjected to one component of excitation, is then extended for

bidirectional excitation. The equivalent-linear procedure for estimating isolator deformations

and forces is evaluated against the nonlinear procedure developed in this chapter.

3.2 System and Governing Equation

3.2.1 System Considered

The system analyzed is idealized as a rigid mass mounted on a single isolator (Fig. 3.1). The

mass m represents the total mass above the isolator, which includes the structure mass and the

base mass. The isolator has bilinear force-deformation, parameterized by the postyield stiffness

kb , the yield strength Q, and initial stiffness kI or yield deformation uy , where uy is

Q
uy = (3.1)
kI − kb

In a lead-rubber bearing, which will be the focus here, kb is the stiffness of rubber and Q is the

strength of the lead core. The force in the isolator is determined from the force-deformation

relation shown in Fig. 3.1b and represented by the following equation:

fb = kb ub + Qz(kI , ub , u̇b ) (3.2)

where ub is the isolator deformation and z represents the fraction of the yield strength applied.

This function z, which depends on the initial stiffness, deformation, and velocity; equals ±1 on

9
300 20
(a) (b)
150 10
üg (cm/s2 )

ub (cm)
0 0

−150 −10

−300 −20
0 10 20 30 40 0 10 20 30 40
t (sec) t (sec)
2
(c) 0.2 (d)
1
0.1

fb /w
0 0
z

−1 −0.1
−0.2
−2
0 10 20 30 40 −20 −10 0 10 20
t (sec) ub (cm)

Figure 3.2: (a) Strong component of ground motion for LMSR Record No. 11, (b) isola-
tor deformation, (c) yield function z(t), and (d) isolator force-deformation relation. System
parameters are Tb = 2.5 seconds and µ = 0.07.

the upper and lower bounding surfaces – represented by dashed lines in Fig. 3.1b – and varies

linearly between these bounding surfaces.

3.2.2 Equation of Motion

The deformation of the isolator that supports mass m (or weight w), subjected to ground

acceleration üg (t), is governed by

üb (t) + ωb2 ub (t) + µgz(t, kI , ub , u̇b ) = −üg (t) (3.3)


where ωb = kb /m and µ = Q/w. Equation 3.3 is solved for a single ground acceleration (see

Fig. 3.2a) to obtain the deformation history ub (t) (Fig. 3.2b). Also shown is the yielding history

through z(t) (Fig. 3.2c) and the variation of the isolator force coefficient fb /w with deformation

(Fig. 3.2d).

The parameters ωb and µ in Eq. 3.3 characterize the nonlinear system. Because the

system vibrates mostly at its postyield stiffness (see Fig. 3.2d and the time durations when

z = ±1 in Fig. 3.2c), ωb is an appropriate frequency and Tb = 2π/ωb is an appropriate period

to characterize the isolation system; ωb and Tb are generally known as the isolation frequency

10
and period, respectively. Known as the characteristic strength ratio, µ quantifies the strength

of the system relative to the structure weight w, and target ranges of µ provide a basis for

designing the yield strength of the isolation system. Equation 3.3 indicates that besides the

isolation period and the strength ratio, the deformation depends on the initial stiffness kI and

is sensitive to ground motion intensity. We will show how to reduce the number of parameters

that markedly influence the response from 4 to 2.

3.3 System Parameters and Normalization

3.3.1 Initial Stiffness

Unlike the postyield stiffness, which is associated with a physical property of the isolator, the

initial stiffness is open to interpretation, because the transition from initial to yielded state

in actual bearing tests is gradual. Naeim and Kelly [2, pg106] suggest that initial stiffness be

modeled as a fixed multiple of the postyield stiffness. To demonstrate the effect of fixing kI /kb

to a constant, peak deformation is plotted against the isolation period Tb for various strength

coefficients µ in Fig. 3.3a. This figure shows that an increase in yield strength leads to an

increase in deformation at long periods, which is contrary to the expectation that the lead core

should limit deformation in proportion to its size.

Using a different model for initial stiffness, Makris and Chang [23] concluded that the

response is insensitive to initial stiffness. With this model, the yield deformation – rather than

the ratio of the initial to postyield stiffness – is fixed; and the initial stiffness depends on the

yield strength instead of the postyield stiffness. For yield deformation fixed at 1 cm with all

other parameters remaining the same, the peak deformation tends to decrease as yield strength

increases (Fig. 3.3b), consistent with expectations. Although the response may not be sensitive

to small changes in initial stiffness, Fig. 3.3 shows that it is sensitive to the large discrepancies

in initial stiffness resulting from these two disparate models. Thus, selecting a model that best

represents the behavior of the system is important.

If the lead core is press fit into a hole that is slightly too small, causing it to extrude into

the rubber and lock with the steel plates, it deforms primarily in shear throughout its volume

11
30
(a) (b)

25

20
ubo (cm)

15

10 µ=.04
.06
.08
5 .10
.12
0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.3: Peak deformation ubo vs Tb for different values of µ for (a) constant kI /kb = 10
and (b) constant uy = 1 cm; excitation is the strong component of LMSR Record No. 8.

when subjected to load [3]. Consider the following model for this behavior: The initial stiffness

is kI = kl + kb , where kl is the stiffness of lead (prior to yielding) and kb is the stiffness of

rubber. These stiffnesses depend on their respective shear moduli Gl and G, areas Al and A,

and total thickness tr of rubber in the bearing, according to the equations kl = Gl Al /tr and

kb = GA/tr . The yield strength is Q = Al σyl , where σyl is the yield stress of the lead core.

These definitions lead to the following expressions:

kI Gl Al
=1+ (3.4)
kb GA

and from Eq. 3.1:


Q σyl tr
uy = = (3.5)
kl Gl
The stiffness kb and, hence, isolation period Tb may be altered by modifying G or A,

which leave the yield deformation unchanged (Eq. 3.5), or by modifying tr , which leaves the

ratio kI /kb unchanged (Eq. 3.4). On the other hand, the yield strength Q at a fixed period

Tb may be altered only by modifying Al , leaving the yield deformation unchanged. Thus,

although the yield deformation may vary with period, such variations should be independent

of yield strength; therefore, fixing kI /kb to a constant is inadequate because it causes the yield

deformation to vary with yield strength.

12
Because practical concerns dictate that adjustment to the isolation period Tb over a wide

range requires modifications to not only A and G, but also tr , Eq. 3.5 suggests that a model

where yield deformation is dependent to a certain degree on the isolation period is appropriate.

Unfortunately, because these properties do not vary with period in a predictable way, devising

such a model is difficult. Given this obstacle, fixing the yield deformation to 1 cm seems

reasonable and shall be adopted throughout the study, as consideration of multiple values of

deformation does not seem to provide meaningful information.

3.3.2 Normalized Strength

To obtain meaningful statistical analysis of responses to an ensemble of ground motions, the

strength of the system should be defined relative to the intensity of individual ground motions.

Peak Ground Velocity Normalization

Consider a∗y = Q/m = µg, equal to the acceleration at yield of a rigid system with strength Q,

and

u∗y = Q/kb = a∗y /ωb2 (3.6)

which is related to the true yield deformation uy of the system by u∗y = (kI /kb − 1)uy . Dividing

the equation of motion (Eq. 3.3) by u∗y leads to the following equation:

üg
¨b + ωb2 ūb + ωb2 z(ub , u̇b ) = −ωb2
ū (3.7)
a∗y

where ūb = ub /u∗y .

The normalized strength, which characterizes the system strength relative to the PGV

u̇go , is defined as
a∗y µg
η= = (3.8)
ωd u̇go ωd u̇go
The frequency ωd , included to make η a unitless parameter, corresponds to the period Td mark-

ing the transition from the velocity-sensitive to the displacement-sensitive region of the median

spectrum (Fig. 2.1). Peak ground velocity is the preferred measure of ground motion intensity

for base-isolated structures with isolation period Tb typically in the velocity or displacement-

sensitive spectral regions.

13
10 2
(a) (b)
5 1
¨g (1/s)

ūb
0 0

−5 −1

−10 −2
0 10 20 30 40 0 10 20 30 40
t (sec) t (sec)

Figure 3.4: (a) Strong component of normalized ground acceleration for LMSR Record No. 11
and (b) normalized isolator deformation ūb (t) for a system with Tb = 2.5 seconds and η = 0.5.

Incorporating η into the equation of motion results in

ωb2
¨b + ωb2 ūb + ωb2 z(ub , u̇b ) = −
ū ¨g
ū (3.9)
ηωd

¨g = üg /u̇go has been normalized such that its corresponding velocity
where the acceleration ū
¨g (t)
varies from -1 to 1. Figure 3.4 demonstrates the normalized ground acceleration history ū

of the strong component of Record No. 11 and the resulting normalized deformation ūb (t), in

contrast to Fig. 3.2 that presented their unnormalized counterparts.

Equation 3.9 implies that if Tb and η are fixed, the intensity of the ground motion has no

effect on the peak normalized deformation ūbo . This important property permits meaningful

statistical analysis of the responses to an ensemble of motions with common frequency content

but variable intensity. Dispersion in ūbo is expected to be small, allowing the peak response to

be estimated with higher confidence than if the governing equation had not been normalized.

While not dependent on intensity, ūbo appears to depend on the system parameters Tb and η,
¨g (t).
the ground motion ensemble parameter ωd , and the frequency content of the excitation ū
¨g (t) have been normalized to a common intensity, their
However, because the excitations ū

variability is limited to that inherent in a random process. Thus, for a given ensemble, the

median of the normalized deformation ūbo of a system depends on only two parameters: the

isolation period Tb and the normalized strength η, as was promised earlier.

Alternative Normalizations

To demonstrate that normalizing the strength by the PGV effectively reduces the dispersion in

the response, two alternatives were considered:

14
1. No normalization: the system is characterized by its isolation period Tb and strength

coefficient µ.

2. Normalization by peak ground acceleration (PGA): the system is characterized by its

isolation period Tb and normalized strength

µg
η = (3.10)
ügo

The dispersion (Eq. 2.2) of peak normalized deformation ūbo (where applicable) and peak

deformation ubo due to the strong-component ground motion ensemble is presented in Fig. 3.5

for each alternative: no normalization (Fig. 3.5a), PGA normalization (Figs. 3.5b and c), and

PGV normalization (Figs. 3.5d and e). For consistency, µ and η  are the medians of their

corresponding values computed to match η for each ground motion; µ was determined from

Eq. 3.8 and η  from Eq. 3.10. Thus, for five selected values of η: {0.25, 0.5, 0.75, 1.0, 1.5}, the

matching values for µ and η  that have been rounded off are: µ = {.023, .045, .068, .09, .137}

and η  = {.085, .17, .25, .34, .52}.

The dispersion of normalized deformation ūbo is usually smallest using the PGV normal-

ization, demonstrating its superiority over the alternatives. Compare the dispersion of ūbo for

the PGV normalization, ranging from 0.3 to 0.6 (Fig. 3.5d), to that of the PGA normalization,

ranging from 0.4 to 1.0 (Fig. 3.5b), and finally to the dispersion of ubo for no normalization

(since ūbo is undefined), ranging from 0.5 to 1.0 (Fig. 3.5a). If the normalization is effective,

the dispersion of normalized deformation ūbo , which is independent of ground motion intensity,

should be smaller than the dispersion of deformation ubo , which is influenced by individual

ground motion intensity. Note that the PGA normalization provides little benefit because the

dispersion of normalized deformation (Fig. 3.5b) is as large as the dispersion of deformation

(Fig. 3.5c). For the PGV normalization, however, the dispersion of normalized deformation

(Fig. 3.5d) is lower than the dispersion of deformation (Fig. 3.5e), demonstrating the effective-

ness of this approach.

15
1.2
(a)
1

Dispersion of ubo
0.8

0.6

0.4 µ=.023
.045
0.2 .068
.090
.137
0
0 1 2 3 4 5
Tb (sec)

1.2 1.2
(b) (c)
1 1
Dispersion of ūbo

Dispersion of ubo

0.8 0.8

0.6 0.6

0.4 η =.085 0.4


0.17
0.2 0.25 0.2
0.34
0.50
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

1.2 1.2
(d) η=.25 (e)
1 0.5 1
Dispersion of ūbo

Dispersion of ubo

0.75
1.0
0.8 1.5 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.5: Dispersion δ of various responses: (a) ubo for no normalization, (b) ūbo and (c)
ubo for PGA normalization, (d) ūbo and (e) ubo for PGV normalization.

16
3.3.3 Isolator Deformation and Force

The peak isolator deformation ubo and isolator force coefficient fbo /w can be readily obtained

from normalized deformation ūbo . The isolator deformation is

ubo = ūbo u∗y (3.11)

where u∗y (Eq. 3.6) is related to the normalized strength η by Eq. 3.8, indicating that the

deformation demand depends, as it should, on PGV in addition to the isolation period and the

normalized strength.

Assuming the system yields, the isolator force is fbo = Q + kb ubo , leading to a correspon-

dence between force and normalized deformation:

fbo
= 1 + ūbo (3.12)
Q

from which the force coefficient fbo /w can be expressed as

fbo
= µ(1 + ūbo ) (3.13)
w

where µ is related to η by Eq. 3.8, indicating that the isolator force coefficient also depends on

PGV, isolation period and normalized strength.

In Fig. 3.6, the deformation ubo and force coefficient fbo /w due to the strong component

of Record No. 8 are plotted as functions of the isolation period for several values of normalized

strength. Figure 3.6a presents data similar to Fig. 3.3b, but for specified values of η = {0.25,

0.5, 0.75, 1.0, 1.5} instead of µ. Equation 3.8 gives the corresponding values of µ equal to

{.035, .069, .104, .139, .208}, using u̇go = 44.7 cm/s for this excitation and ωd = 3.05 rad/s

(corresponds to Td = 2.06 sec) for the median spectrum of the strong-component ensemble.

While increasing the isolator yield strength reduces the deformation demand over a wide range

of isolation periods (Fig. 3.6a), it may increase the isolator force at longer periods (Fig. 3.6b).

This can be understood by recognizing that at long isolation periods Tb the system approaches

an elastic-perfectly plastic system, in which case the deformation has no limiting value, and the

force coefficient equals the strength coefficient µ (Fig. 3.6b).

17
30 0.5
η=.25
0.5
25 0.75
0.4
1.0
1.5
20
0.3
ubo (cm)

fbo /w
15
0.2
10

0.1
5

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.6: (a) Isolator deformation ubo and (b) force coefficient fbo /w vs Tb for several values
of η due to strong component of LMSR Record No. 8.

3.4 Comparison to Nonlinear Structural Analysis

The preceding formulation of the earthquake response of a nonlinear base-isolation system is

reminiscent of the standard approach for analyzing nonlinear structural response. While similar

in some aspects, the two differ significantly, as described next.

The structural problem is usually formulated as follows: consider a structure with an

elasto-plastic force-deformation relation, initial stiffness kn or initial period Tn , damping ratio

ζ, yield strength fy , and yield deformation uy . The yield strength fy of the structure corresponds

to the yield deformation uy whereas the yield strength Q of the isolation system is the y-intercept

of the fb -ub relation (Fig. 3.1b). If fo is the minimum strength required for the structure to

remain elastic, uo = fo /k is the corresponding deformation, and f¯y = fy /fo is the normalized

strength of the system. In contrast to the isolator strength that was normalized relative to

the intensity of the ground motion (Eq. 3.9), the structural strength is normalized relative to

the elastic demand on the structure. Although the isolator strength η (Eq. 3.8) has no upper

bound, the structural strength f¯y is bounded between 0 and 1.

Table 3.1 compares the governing equations for the structural system and the isolation

system. In step 1, each equation is written in terms of a relevant system frequency: the

initial, small-oscillation frequency ωn for the structure, and the isolation frequency ωb for the

18
Table 3.1: Governing equations for nonlinear structural systems vs. isolation systems.

Structural systems Isolation systems

(1) Equation of motion: (1) Equation of motion:

fy Q
ü + 2ζωn u̇ + z(u, u̇) = −üg (t) üb + ωb2 ub + z(ub , u̇b ) = −üg (t)
m m

(2) Divide by uy : (2) Divide by u∗y :

üg (t) üg (t)


µ̈s + 2ζωn µ̇s + ωn2 z(µs , µ̇s ) = −ωn2 ¨b + ωb2 ūb + ωb2 z(ub , u̇b ) = −ωb2

ay a∗y

(3) Note that ay = f¯y An , giving: (3) Define η so that a∗y = ηωd u̇go , giving:

ω 2 üg (t) ωb2 üg (t)


µ̈s + 2ζωn µ̇s + ωn2 z(µs , µ̇s ) = − ¯n ¨b + ωb2 ūb + ωb2 z(ub , u̇b ) = −

fy An ηωd u̇go

isolation system. In step 2, the structural equation, which is normalized by uy , becomes a

differential equation for the ductility factor µs = u/uy , where ay = fy /m = ωn2 uy . In contrast,

the isolation equation is normalized by u∗y , which is more appropriate for the isolation system

because it depends on the isolation frequency rather than the initial frequency. Aside from the

normalizing parameter, the approaches for the two systems are parallel.

Thereafter, however, the approaches for the two systems diverge to reflect differing defi-

nitions of normalized strength. In step 3, the structural equation of step 2 has been rewritten

in a form similar to the isolation equation by relating ay to f¯y and the pseudo-spectral acceler-

ation An of the reference linear system: ay = ωn2 uy = ωn2 f¯y uo = f¯y A; thus the ground motion

in the structural equation has been normalized by spectral acceleration An . Unable to find a

reference linear system for the isolation system that consistently reflected an appropriate level

of damping, we instead normalized its strength by a ground motion intensity parameter. While

such an approach has also been suggested for structures [24, 25, 26], it has not been widely

adopted. An elastic-plastic structure worked best in the preceding comparison; however, the

ideas are valid for bilinear and other models.

19
3.5 Earthquake Response Spectra

As demonstrated earlier, the PGV normalization reduces dispersion of the normalized deforma-

tion ūbo , allowing the response to an ensemble of ground motions to be estimated with a high

degree of confidence. For the normalization to be useful, nonlinear spectra based on statistical

analysis of response data are needed. Such spectra are developed next for the strong-component

ensemble.

3.5.1 Normalized Deformation

The median response spectrum for normalized deformation ūbo is constructed by performing

the following sequence of steps:

1. Select ranges for the isolation period and normalized strength. We chose Tb from 1 to

5 seconds and five values of η: {0.25, 0.5, 0.75, 1.0, 1.5}. Using the median properties

of the strong-component ensemble – ωd = 3.05 rad/s and median u̇go = 29.4 cm/s for

the unscaled motions – and a generous range for µ from 0.04 to 0.12, Eq. 3.8 led to an

estimated range of η from about 0.44 to 1.31. The five values of η chosen cover a slightly

larger range to account for some variation in intensity among ground motions.

2. Determine the normalized deformation ūbo by nonlinear response history analysis (RHA)

of Eq. 3.9, repeating over the desired range of Tb and η and for each ground motion of

the ensemble. Figure 3.4 presented such results for one time history.

3. Compute the median normalized deformation (Eq. 2.1) over all ground motions for each

value of Tb and η.

Implementing steps 2 and 3 for the strong-component ensemble led to the median normalized

deformation spectrum shown as solid lines in Fig. 3.7.

3.5.2 Design Equations

Figure 3.7b shows that the median normalized deformation ūbo varies linearly with Tb on a log

scale plot. This suggested fitting a regression equation to ln (ūbo ) that is linear in the parameters

ln (Tb ) and ln (η). Regression analysis on a data set more comprehensive than the one shown

20
2
20 10
(a) exact median (b)
design spectrum
1
15 10

η=0.25
0
ūbo

ūbo
10 10
η=0.25
0.5
−1 0.75
5 0.5 10 1.0
0.75 1.5
1.0
1.5 −2
0 10
0 1 2 3 4 5 0 1
10 10
Tb (sec) Tb (sec)

Figure 3.7: Median normalized deformation ūbo for strong-component ensemble along with
the design spectrum (Eq. 3.14b), plotted in two formats: (a) linear scales and (b) log scales.

in Fig. 3.7 resulted in the equation

ln ūbo = 0.65 − 1.81 ln Tb − 1.55 ln η − 0.08(ln η)2 (3.14a)

or

ūbo = 1.91 Tb−1.81 η (−1.55−0.08 ln η) (3.14b)

with coefficients estimated by the method of least squares. The design spectrum given by this

equation – shown as dashed lines in Fig. 3.7 – is very close to the exact (median) spectrum and,

hence, is suitable for design and analysis of isolation systems.

Because the dispersion δ of ūbo is essentially independent of Tb and η (except for small

values of η, Fig. 3.5d), regression analysis can be used to estimate a single value for δ applicable

to all Tb and η. This method of computing dispersion, alternative to applying Eq. 2.2 for each

value of Tb and η, allows estimation of the median+σ, or 84th percentile response, without

further regression analysis. By such methods, the dispersion of ūbo was estimated to be 0.38;

further details are provided in Appendix 3A. Thus, the 84th percentile (Eq. 2.3) of ūbo is

approximated by

ūbo,84% = 1.46 ūbo (3.15)

This equation could be used in place of Eq. 3.14b whenever a more conservative estimate

corresponding to this lower exceedance probability is desired.

21
(a) (b)
15

10 η=0.25 η=0.25
ubo (cm)

0.5 0.5
0.75 0.75
5 1.0 1.0
1.5 1.5
exact median exact median
design equation simplified equation
0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

(c) η=0.25 (d)


20 0.5
0.75
1.0
1.5
10
% Discrepancy

−10

−20
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.8: Comparison of design equations for ubo (a) Eq. 3.16, (b) Eq. 3.18 (simplified equa-
tion) to the exact median by nonlinear RHA with design PGV of 35 cm/s; percent discrepancy
in (c) Eq. 3.16 and (d) Eq. 3.18 relative to the exact median.

A design equation for the median deformation is obtained by substituting Eq. 3.14b for

ūbo and Eqs. 3.6 and 3.8 (with ωd = 3.05) for u∗y into Eq. 3.11:

5.83 0.19 (−0.55−0.08 ln η)


ubo = T η u̇go (3.16)
4π 2 b

This equation provides the deformation for a given median PGV u̇go that reflects the intensity

of the design ground motions. The deformation given by Eq. 3.16 is shown to be close to its

exact median determined by nonlinear RHA of the system for the strong-component ensemble

(Fig. 3.8a), and the percent discrepancy between this design equation and the exact median is

small (Fig. 3.8c).

Although Eq. 3.14b is a better fit to the normalized deformation, a simplified equation

22
that neglects quadratic variation with η, and hence may be desirable for code applications is

ūbo = 1.89 Tb−1.81 η −1.46 (3.17)

This equation for normalized deformation leads to a simplified equation (compared to Eq. 3.16)

for the isolator deformation ubo :

5.76 0.19 −0.46


ubo = T η u̇go (3.18)
4π 2 b

Plotted against the exact median in Fig. 3.8b, the deformation estimated by Eq. 3.18 results in

only a slight loss of accuracy compared to Eq. 3.16 (Fig. 3.8a). Although the simplified equation

(Eq. 3.18) leads to a slightly larger discrepancy relative to the exact median on average, the

error bounds are about the same as for Eq. 3.16 (Fig. 3.8d vs. 3.8c).

The force coefficient could be determined by substituting the normalized deformation

(Eq. 3.14b or Eq. 3.17) into Eq. 3.13, or directly from the deformation (Eq. 3.16 or Eq. 3.18)

by Eq. 3.2 with z = 1.

3.6 Governing Equations for Bidirectional Excitation

3.6.1 Equations of Motion

Consider the system of Fig. 3.1a with the isolator properties – postyield stiffness kb , yield

strength Q, and yield deformation uy – identical in both the x and y-directions, subjected to

bidirectional excitation. The x and y-components of bearing force are given by a generalization

of Eq. 3.2: ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎨fbx ⎪
⎪ ⎬ ⎪
⎨ubx ⎪
⎬ ⎨zx (kI , ub, u̇b)⎪
⎪ ⎬
= kb +Q (3.19)
⎩fby ⎪
⎪ ⎭ ⎪
⎩uby ⎪
⎭ ⎩zy (kI , ub , u̇b)⎪
⎪ ⎭

where ub = ubx , uby T , u̇b = u̇bx , u̇by T and z = zx , zy T are the isolator deformation, velocity

and yield function in vector form.

Generalizing Eq. 3.3 gives the deformation of the isolator in the x and y-directions due

to ground acceleration components ügx and ügy :


⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎨übx ⎪
⎪ ⎬ ⎪
⎨ubx ⎪
⎬ ⎪
⎨zx ⎪
⎬ ⎪
⎨ügx ⎪

2
+ ωb + µg =− (3.20)
⎩üby ⎪
⎪ ⎭ ⎪
⎩uby ⎪
⎭ ⎪
⎩zy ⎪
⎭ ⎪
⎩ügy ⎪

23
with ωb and µ, the isolation frequency and characteristic strength ratio, as defined before.

Equation 3.20 is solved for a system with Tb = 2.5 seconds and µ = 0.07, with the weak and

strong components of Record No. 11 applied in the x and y-directions, respectively, resulting

in the evolution of deformation traced in Fig. 3.9a. Identified in this figure are ubyo , the peak

deformation in the y-direction, and the peak deformation in any direction, subsequently referred

to as peak deformation:

ubo = max (ubx (t)2 + uby (t)2 ) (3.21)
t

Implied by Eq. 3.19, the components of the yield function interact, governed by the con-

straint that |z| ≤ 1, which results in a circular yield surface. Yielding of the system (and

vibration at the postyield stiffness kb ) is defined by |z| = 1; the system is elastic otherwise.

This interaction is also evident in the force versus deformation plots for the x and y-directions

(Fig. 3.9c and d), which appear erratic compared to comparable plots for unidirectional excita-

tion (Fig. 3.2d). Examples comparing a system with no yield-function interaction (rectangular

yield surface) to a system with yield-function interaction as assumed here are available in Fenves

et al [27] and Huang [22].

3.6.2 Normalized Equations

To extend the normalization procedure of Sec. 3.3.2 to systems excited by two components of

ground motion, the equations of motion (Eq. 3.20) are rewritten as


⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫

⎨ū ⎪
⎬ ⎪
⎨ ⎪
⎬ ⎪
⎨ ⎪
⎬ ⎪¨ ⎪
¨bx ūbx zx ωb2 ⎨ū gx

+ ωb2 2
+ ωb =− (3.22)
⎪ ¨by ⎪
⎩ū ⎭ ⎪
⎩ūby ⎪⎭ ⎪
⎩zy ⎪⎭ ηωd ⎪ ¨gy ⎪
⎩ū ⎭

This has been achieved by first dividing Eq. 3.20 by u∗y , such that ūbx = ubx /u∗y and ūby = uby /u∗y

are normalized deformations in the x and y-directions, and then substituting the normalized

strength into the right-hand side. The definition of normalized strength in Eq. 3.8 is still valid

where ωd (Fig. 2.1) and u̇go refer to the stronger component of ground motion, which is applied

in the y-direction; thus u̇go is replaced by u̇gyo and Eq. 3.8 becomes

a∗y µg
η= = (3.23)
ωd u̇gyo ωd u̇gyo

24
20 2
(a) (b)
ubo =18.7 cm
uby (cm) 10 1

zy
0 ubyo = 0
15.7 cm

−10 −1

−20 −2
−20 −10 0 10 20 −2 −1 0 1 2
ubx (cm) zx

0.2 0.2
(c) (d)

0.1 0.1
fbx /w (g)

fby /w (g)

0 0

−0.1 −0.1

−0.2 −0.2
−20 −10 0 10 20 −20 −10 0 10 20
ubx (cm) uby (cm)

Figure 3.9: Bidirectional excitation of a system with Tb = 2.5 sec and µ = 0.07 to LMSR
Record No. 11 (strong component applied in y-direction) results in: (a) deformations ubx vs
uby , (b) yield functions zx vs zy , (c) force-deformation in x and (d) y-directions

Therefore, the components of ground acceleration in Eq. 3.22 are normalized by the PGV in
¨gx = ügx /u̇gyo and ū
the y-direction, i.e., ū ¨gy = ügy /u̇gyo , such that the corresponding velocity

in the y-direction varies from -1 to +1, while the magnitude of velocity in the x-direction is

strictly less than 1.

3.7 Design Equations for Bidirectional Excitation

Similar to the procedure outlined in Sec. 3.5.1, a median spectrum for the deformation ubo

(or other deformation of interest) of a system due to bidirectional excitation is obtained by

repeating the following steps over a suitable selection of Tb and η: (1) determine the peak

normalized deformation ūbo by nonlinear RHA of Eq. 3.22 for every motion in the ensemble, (2)

compute the median of ūbo by Eq. 2.1, and (3) convert ūbo to the deformation ubo by multiplying

by u∗y (Eq. 3.6), a function of η, PGV and ωd for the strong-component ensemble.

25
15 15
(a) (b)

10 10
ubyo (cm)

ubo (cm)
η=0.25
η=0.25
0.5 0.5
0.75 0.75
5 1.0 5 1.0
1.5 1.5
unidirectional excitation unidirectional excitation
bidirectional excitation bidirectional excitation
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.10: Median deformation ubo for unidirectional excitation compared to the median
deformations (a) ubyo in y-direction and (b) ubo – peak in any direction – for bidirectional
excitation, both due to the LMSR ensemble with median PGV of 35 cm/s.

Figure 3.10 compares the median deformation spectra for bidirectional and unidirectional

excitations; the latter case was already shown in Fig. 3.8. The similarity between the comparable

deformations ubyo due to bidirectional excitation and ubo due to unidirectional excitation implies

that adding a transverse component of excitation has little influence on the median deformation

in the direction considered (Fig. 3.10a). However, the peak deformation ubo of the isolator

(Eq. 3.21) is significantly larger when the excitation is bidirectional (Fig. 3.10b).

Because the y-direction deformation due to bidirectional excitation is very similar to the

deformation for unidirectional excitation, both sets of design equations (Eqs. 3.14b and 3.16,

or Eqs. 3.17 and 3.18) are expected to be valid even for bidirectional excitation. The first set

is shown to be a good fit in Fig. 3.11a, where Eqs. 3.14b and 3.16 (shown as dashed lines) are

compared against the exact median values of ūbyo and ubyo due to bidirectional excitation.

Regression analysis of the response data for bidirectional excitation assuming an equation

of the form of Eq. 3.14b or of the simplified Eq. 3.17 led to coefficients for Tb and η that are

suitably close to those for unidirectional excitation. Therefore, they have been constrained to

be identical, and ūbo is estimated to be a constant 13% greater for bidirectional excitation than

for unidirectional excitation:

ūbo = 1.13 ūbo(unidirectional) (3.24)

26
(a)
2
10 15
exact median
1 design equation
10
10

ubyo (cm)
η=0.25
ūbyo

0
10 η=0.25 0.5
0.5 0.75
−1 0.75 5 1.0
10 1.0 1.5
1.5
−2
10 0
0 1 0 1 2 3 4 5
10 10
Tb sec Tb sec
2
(b)
10 15
exact median
1 design equation
10
10 η=0.25
ubo (cm)

0 0.5
ūbo

10 η=0.25 0.75
0.5 1.0
0.75 5
−1
10 1.5
1.0
1.5
−2
10 0
0 1 0 1 2 3 4 5
10 10
Tb sec Tb sec

Figure 3.11: Design equations for (a) y-direction deformations ūbyo (Eq. 3.14b) and ubyo
(Eq. 3.16) and (b) peak deformations ūbo (Eq. 3.24) and ubo (Eq. 3.25) compared to their exact
medians determined by nonlinear RHA due to the LMSR ensemble with PGV of 35 cm/s.

which applies for both versions of ūbo(unidirectional) (Eq. 3.14b and Eq. 3.17). (Because of this

constraint, the estimated dispersion δ is the same as for unidirectional excitation, and the same

factor of 1.46 [see Eq. 3.15] can be applied to Eq. 3.24 to achieve an 84th percentile estimate

of response. Further details of the regression methods used and statistical significance of these

equations are given in Appendix 3A.)

Similarly, design equations for the peak isolator deformation reflect a 13% increase over

those for unidirectional excitation, and are written out here in their entirety for future reference:

6.59 0.19 (−0.55−0.08 ln η)


ubo = T η u̇gyo (3.25)
4π 2 b

and the simplified equation:


6.51 0.19 −0.46
ubo = T η u̇gyo (3.26)
4π 2 b

27
Design Eqs. 3.24 (with Eq. 3.16 for ūbo(unidirectional) ) and 3.25 are shown to be excellent fits for

the exact median data in Fig. 3.11b.

Because of the interaction of the yield functions zx and zy in the x and y-directions, the

peak force in any direction cannot be directly related to the peak deformation for bidirectional

excitation, and Eq. 3.13 is not strictly valid. However, it is shown that Eq. 3.13 is an upper

bound for the actual peak force, and is exact if the yield function z is in the same direction as

the deformation ub when its peak occurs (Appendix 3B). Because z and ub are likely to be

close to the same direction, Eq. 3.13 is a good upper bound; even when the directions differ by

as much as 60 degrees, Eq. 3.13 is at most 15% conservative.

3.8 Application to Analysis and Design

3.8.1 Estimating Isolator Deformation

If the isolation period Tb and strength coefficient µ of a base-isolation system are known, estimat-

ing the isolator deformation from a design equation, such as Eq. 3.25 or 3.26, is straightforward.

First, however, the normalized strength η, which depends on design PGV, must be determined.

For Eq. 3.25, the calculations are demonstrated below in Example 1.

Example 1

Determine the seismic deformation of an isolated building with Tb = 2.5 seconds and µ = 0.06

due to the LMSR ground motion ensemble and a design PGV u̇gyo = 35 cm/s in the direction

of strong excitation.

1. From Eq. 3.23, η = 0.06 ∗ g/(3.05 ∗ 35) ≈ 0.55

2. From Eq. 3.25, ubo = (6.59/4π 2 ) ∗ (2.5)0.19 ∗ (0.55)(−0.55−0.08 ln 0.55) ∗ 35 = 9.39 cm

For comparison, the deformation computed from the median of ūbo that was obtained by non-

linear RHA is ubo = 9.56 cm.

3.8.2 Selecting Isolator Strength

The design equation for deformation could also be used to find suitable choices for the isolation

period and strength coefficient, such that the deformation and force demands do not exceed

28
allowable values. For example, suppose the design problem is posed as follows: Given an

isolation period Tb and a design PGV u̇gyo , select the strength coefficient µ of the system to

meet the following constraints:

ubo ≤ uallowable (3.27)


 
fbo f
≤ (3.28)
w w allowable
For the first constraint, deformation controls the height of the bearing needed to limit

the shear strain, while the height controls the diameter of the bearing needed for stability. By

limiting the deformation, the overall size of the bearing can be reduced, ultimately making

it easier to achieve the desired flexibility. The second constraint may be imposed to limit the

forces transferred to the structure; for instance, it may be necessary to keep the structure elastic

when designing an isolation system for a building retrofit.

A procedure to find strength coefficients such that the constraints are satisfied is sum-

marized.

1. Find a lower bound ηL such that the deformation demand does not exceed the allowable

deformation. For the LMSR ensemble, substitute Eq. 3.25 into Eq. 3.27 to give the

constraint:
6.59 0.19 (−0.55−0.08 ln η)
T η u̇gyo ≤ uallowable (3.29)
4π 2 b
Solve this nonlinear equation numerically or by trial and error to obtain the lower bound.

2. Find an upper bound ηU such that the isolator force demand does not exceed the allowable

force coefficient. With Eq. 3.14b for ūbo and Eq. 3.23 for µ, Eq. 3.13 is substituted into

Eq. 3.28 as an estimate for the force coefficient, giving

ηωd u̇gyo   f 
−1.81 (−1.55−0.08 ln η)
1 + 1.91 Tb η ≤ (3.30)
g w allowable

Solving this nonlinear equation numerically or by trial and error gives a conservative

estimate of ηU , because the left side is an upper bound to the true force.

3. If ηL < ηU , then both constraints (Eqs. 3.27 and 3.28) can be met by any η within the

bounds. Convert the bounds for η to bounds for µ using Eq. 3.23, giving µL < µ < µU .

29
Select a strength coefficient µ, and compute the deformation demand as demonstrated in

Example 1 and the approximate force demand from Eq. 3.13.

4. If ηL > ηU , then allowable deformation and force cannot both be satisfied, thus requiring

modification of the isolation period Tb or the design criteria.

Example 2

For an isolation system with Tb = 2.5 seconds, select the strength coefficient to sustain bidirec-

tional shaking represented by the LMSR ensemble and a design ground velocity of 35 cm/s in

the direction of strong excitation, so that the deformation does not exceed 10 cm and the force

coefficient does not exceed 0.14.

1. Equation 3.29 is solved by trial and error to obtain ηL = 0.48.

2. Equation 3.30 is solved by trial and error to obtain ηU = 0.83.

3. An acceptable range for η is 0.48 to 0.83, equivalent to µ from 0.052 to 0.09 (Eq. 3.23).

Selecting µ = 0.07, or η = 0.64, leads to deformation and (approximate) force demands,

calculated from Eq. 3.25 and Eq. 3.13, of ubo = 8.72 cm and fbo /w = 0.126. These

demands are below the allowable limits and the design is acceptable.

3.9 Comparison to Code Procedure

The International Building Code 2000 [1] and previous editions of the Uniform Building Code

estimate the design displacement of an isolation system based on an equivalent-linear system,

whose properties are generally determined by iteration. This section first evaluates the accuracy

of this method to estimate the isolator deformation for unidirectional excitation, defined by

the strong-component spectrum in Fig. 2.1a, and then evaluates a code rule to increase the

deformation for bidirectional excitation. Recall that the this spectrum was developed by scaling

the strong component of each motion to a PGV of 35 cm/s (Sec. 2.1), representing the same

shaking as for the preceding design examples.

30
3.9.1 Estimating Isolator Deformation

Suppose the deformation of a system with known isolation period Tb and strength coefficient µ

due to unidirectional excitation is to be estimated from a linear response (or design) spectrum.

An iterative procedure, which determines the equivalent-linear system for a nonlinear isolation

system, is summarized as follows:

1. Using initial guesses for the effective period Tef f and effective damping ζef f of the

equivalent-linear system, estimate deformation demand ubo from the spectrum.

2. Evaluate the isolator force from Eq. 3.2, taking z to be 1: fbo = Q + kb ubo .


3. Update estimates of effective period: Tef f = 2π w/(kef f g), and effective damping:

ζef f = (2/π)Q(ubo − uy )/(kef f u2bo ), where kef f = fbo /ubo .

4. Repeat steps 1-3 with updated values of Tef f and ζef f , until successive estimates of ubo

converge.

Example 3

Estimate the deformation for an isolated building with Tb = 2.5 seconds and µ = 0.06 due to

unidirectional excitation defined by the median spectrum of Fig. 2.1a for 5% damping. Spectral

displacements for other damping ratios, needed in the iterative procedure, are determined by

additional response history analyses. Select the weight of the structure as w = 1600 kN, giving

Q = 0.06 ∗ 1600 = 96 kN, and kb = (2π/2.5)2 ∗ (1600/g) = 10.29 kN/cm. Assume uy = 1 cm to

evaluate effective damping.

1. For initial guesses of Tef f = 2 seconds and ζef f = 0.15, the median ubo = 9.27 cm.

2. The isolator force, computed from Eq. 3.2, is fbo = 96 + 10.29 ∗ 9.27 = 191.39 kN.

3. The new effective stiffness is kef f = 191.39/9.27 = 20.65 kN/cm, leading to updated

effective period Tef f = 2π 1600/(20.65 ∗ g) = 1.765 sec, and damping ζef f = (2/π) ∗ 96 ∗

(9.27 − 1)/(20.65 ∗ 9.272 ) = 0.285.

31
4. The deformation demand for the updated equivalent-linear system is ubo = 6.75 cm.

Steps 1-3 are repeated until convergence is attained, with the iterations summarized in

Table 3.2.

This equivalent-linear procedure converges after six iterations, giving deformation demand

ubo = 5.88 cm. This deformation is unconservative compared to the exact median deformation

determined by nonlinear RHA of Eq. 3.9 for all 20 motions: ubo = 8.38 cm.

3.9.2 Evaluation of Equivalent Linear Procedure

The accuracy of using a linear spectrum to estimate the deformation of a system subjected to

unidirectional excitation is evaluated over a broader range of isolation period Tb and normalized

strength η. Figure 3.12a plots the median deformation for a system with η = 0.5 due to the

strong-component ensemble, as computed by both nonlinear and equivalent-linear procedures.

The nonlinear deformation is based on the exact median of ūbo determined by nonlinear RHA

instead of the design equation (Eq. 3.16). These plots indicate that the equivalent-linear pro-

cedure fails to recognize the trend of increasing deformation with period, thereby increasingly

underestimating the deformation at longer periods.

Alternatively, percent discrepancy in the deformation estimated by the equivalent-linear

procedure is demonstrated in Fig. 3.12b for several values of η; by convention, discrepancy is

positive when the equivalent-linear estimate exceeds the nonlinear deformation. These plots

confirm that this estimate is unconservative for a wide range of isolation period and normalized

strength. The isolator deformation is underestimated by 20 to 50% for most practical systems,

Table 3.2: Computation of iterations for the equivalent-linear analysis procedure of Example 3.

Teff (s) ζeff ubo (cm) fbo (kN) fbo/w keff (kN/cm)
2.000 0.150 9.27 191.39 0.120 20.65
1.765 0.285 6.75 165.47 0.103 24.51
1.620 0.315 6.12 158.99 0.099 25.98
1.574 0.322 5.94 157.13 0.098 26.45
1.559 0.323 5.89 156.62 0.098 26.59
1.555 0.324 5.88 156.52 0.098 26.62
1.555 0.324 5.88

32
10 50
(a) (b) η=0.25
0.5
8 0.75

% Discrepancy in ubo
25 1.0
1.5
6
ubo (cm)

0
4

Nonlinear −25
2 Equiv-Linear

0 −50
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 3.12: (a) Comparison of deformation determined by nonlinear and equivalent-linear


procedures for a system with η = 0.5, and (b) discrepancy in deformation estimated by
equivalent-linear procedure relative to nonlinear procedure, for several values of η.

becoming increasingly unconservative for longer isolation periods and higher strengths.

The inability of equivalent-linear methods to determine accurate demands for nonlinear

systems has been documented by several researchers [28, 29, 30]. For example, the ATC-40

procedure, an equivalent-linear procedure to estimate the demands of a inelastic structure, has

been shown to be unconservative compared to deformations determined from nonlinear RHA or

inelastic design spectra [28]. Thus, it is not surprising that the isolator deformations determined

by the equivalent-linear procedure are so unconservative.

3.9.3 Extension to Bidirectional Excitation

Although the equivalent-linear procedure of the IBC does not address the issue of bidirectional

excitation[1], the 100% + 30% rule is recommended for response spectrum analysis. Specifi-

cally, the maximum deformation of the system is calculated as the vector sum of orthogonal

deformations resulting from 100% of the ground motion applied in the most critical direction

and 30% applied orthogonally. Adopting this rule for the IBC equivalent-linear procedure, the

ratio of peak deformations due to bidirectional and unidirectional excitations is given by:

ubo 
= 12 + 0.32 = 1.044 (3.31)
ubo(unidirectional)

33
This estimate is based on the conservative assumption that both components of ground motion

have the same intensity. Compare this to a 13% increase in the median deformation due to

bidirectional excitation given in Eq. 3.24. The estimated increase in isolator deformation by the

100% + 30% rule is only about one-third of the exact increase determined by nonlinear RHA

to bidirectional excitation; hence, another possibility for improvement to the code procedure.

3.10 Conclusions

This investigation to develop a procedure to estimate the isolator deformation and force of a

base-isolated building, based on nonlinear response history analysis, has led to the following

conclusions:

1. For a bilinear isolation system, the common modeling strategy of fixing the ratio of the

initial to the postyield stiffness is inadequate when the response is to be compared over a

wide range of parameters of the system. This strategy, which causes the yield deformation

to grow in proportion to the yield strength, does not correctly represent the behavior of

such systems. An improved strategy is to fix the yield deformation.

2. The median – over an ensemble of ground motions – of the normalized deformation ūb

depends only on the isolation period Tb and normalized strength η. This was achieved by

defining the normalized strength as the system strength ÷ ωd u̇go , where u̇go is the peak

ground velocity and ωd is the corner frequency separating the velocity and displacement-

sensitive regions of their median spectrum.

3. The deformation was normalized to be independent of ground motion intensity in order to

minimize the dispersion of the normalized deformation to an ensemble of ground motions.

This implies that the peak response can be estimated with a high degree of confidence

by the median response of the system to the ground motion ensemble. For the strong-

component ensemble, the dispersion of normalized deformation was shown to be small,

especially when compared to alternative normalization procedures.

34
4. The approach for isolation systems presented here differs from the preferred approach for

structural systems as follows: the controlling frequency, which determines the normalized

deformation, is the isolation frequency rather than the initial frequency, because it has

proved unfeasible to define a linear system that accurately reflects the energy dissipation

of the isolators; the strength is normalized by a ground motion intensity parameter rather

than the response of a corresponding linear system.

5. The median normalized deformation of the isolation system showed simple trends across

isolation period Tb and normalized strength η, and was fit to an equation by regres-

sion analysis, which led to a design equation for the isolator deformation. Comparable

equations for the peak deformation due to unidirectional excitation and the peak defor-

mation in any direction due to bidirectional excitation differed by a constant, reflecting

a 13% increase due to the second component of ground motion. Simplified versions of

these equations that may be preferable for code applications were also included. Exam-

ples demonstrated the use of these equations to determine the deformation of a system

with given parameters, or to select the yield strength of a system such that the isolator

deformation and force are within allowable limits.

6. For a given spectrum, equivalent-linear procedures in building codes were shown to under-

estimate the isolator deformation by 20 to 50% compared to the exact median deformation

determined by nonlinear response history analysis. Furthermore, the common 100% +

30% rule allows for at most a 4.4% increase in deformation due to bidirectional excitation,

compared to the exact increase of 13% determined by nonlinear response history analysis.

This investigation has provided a framework for an alternative procedure that could be

integrated into the base-isolation code to replace the current equivalent-linear procedure.

35
Appendix 3A: Regression Statistics

The R Project, an open-source environment for statistical computing available through GNU,

was used for regression analysis that led to the design equations. Excerpts of the R Project

output for Eq. 3.14b (Table 3A.1a) and Eq. 3.17 (Table 3A.1b) are listed, which also apply

to Eq. 3.24. Before this output is discussed, a terminology for linear least-squares regression

is developed. Any textbook on statistical methods [31, 32, 33] could serve as a reference for

the following concepts. For the general problem, the observed data Y is fit to an equation

containing regressors X1 , X2 , etc. The vector Y contains individual observations yi , and the

matrix X contain the values xi1 , xi2 , etc., corresponding to each observation yi . With linear

regression, the model or predictor for Y , Y̌ , is simply a linear function of the regressors, i.e.,

Y̌ = Xβ. Each observation yi satisfies:


yi = y̌i + ei = xik βk + ei (3A.1)
k

where ei is the residual error between the observed value yi and the predicted value y̌i . By the

least-squares method, the coefficients β are selected to minimize the norm of the residual error,

such that the solution can be shown to satisfy

X  (Y − Xβ) = 0 (3A.2)

Applying the above notation to the model of Eq. 3.14a, which is linear by logarithmic

transformation of the variables, the observed data is Y = ln (ūbo ), and the regressors are X0 = 1

(a vector of ones), X1 = ln (Tb ), X2 = ln (η) and X3 = (ln η)2 . Table 3A.1 shows an additional

regressor Ind , an indicator variable that distinguishes observations of ūbo due to bidirectional

excitation (Ind = 1) from the single-direction responses ūbo(unidirectional) and ūbyo (Ind = 0).

This indicator allows the three different responses to be fit to the same model, whereby ubo

due to bidirectional excitation is permitted to increase by a constant relative to the other two.

Compared to the model for the design equation (Eq. 3.14a, Table 3A.1a), the model for the

simplified equation (Table 3A.1b) omits the regressor (ln η)2 .

36
Table 3A.1: Excerpts of R Project output for analysis of response to both unidirectional and
bidirectional excitation

(a) Regression output for design equation


ln ūbo = β0 + β1 ln Tb + β2 ln η + β3 (ln η)2 + β4 · Ind
where Ind (indicator) =1 for 2-direction response, =0 otherwise
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.645836 0.007287 88.63 <2e-16 ***
lnTb -1.806950 0.006137 -294.46 <2e-16 ***
lneta -1.550846 0.010057 -154.21 <2e-16 ***
lneta2 -0.083941 0.008295 -10.12 <2e-16 ***
ind 0.124334 0.006050 20.55 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3788 on 17635 degrees of freedom


Multiple R-Squared: 0.91, Adjusted R-squared: 0.91
F-statistic: 4.457e+004 on 4 and 17635 DF, p-value: < 2.2e-16
--------------------------------------------------------------------

(b) Regression output for simplified equation – neglects quadratic term


(β3 )
ln ūbo = β0 + β1 ln Tb + β2 ln η + β4 · Ind
where Ind (indicator) =1 for 2-direction response, =0 otherwise
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.636306 0.007246 87.81 <2e-16 ***
lnTb -1.806950 0.006154 -293.61 <2e-16 ***
lneta -1.461657 0.004858 -300.90 <2e-16 ***
ind 0.124334 0.006068 20.49 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3799 on 17636 degrees of freedom


Multiple R-Squared: 0.9095, Adjusted R-squared: 0.9095
F-statistic: 5.906e+004 on 3 and 17636 DF, p-value: < 2.2e-16

3A.1 R2 statistic

Listed in Table 3A.1, the ‘Multiple R-Squared’ statistic is a comprehensive measure of the

goodness of fit. R2 is computed as



(y̌i − Ȳ )2 β  X  Y + n(Ȳ )2
2

R = i = (3A.3)
i (yi − Ȳ )
2 Y  Y + n(Ȳ )2

where Ȳ is the average of Y over the number of observations, n. Interpreting this equation, R2

is the fraction – between 0 and 1 – of total variability in Y accounted for when Y̌ is used to

37
predict Y . Exactly equal to 1 when all data points lie on the regression line, R2 should ideally

be close to 1; however, there is no line between acceptable and unacceptable values [31].

With resulting values of R2 equal to 0.91 and 0.9095 (Table 3A.1), both models (for

Eqs. 3.14b and 3.17) led to good fits of our data. This is especially encouraging since three

different responses (ūbo(unidirectional) , ūbyo and ūbo ) were merged into a common model. Fur-

thermore, the negligible difference in R2 for the models leading to the design equation and the

simplified equation suggests that the additional regressor (ln η)2 accomplishes little. However,

the decision to keep this regressor is based on a different statistic.

3A.2 Error Distribution

The residual errors are assumed to be independent and normally distributed with mean 0 and

constant variance σ 2 (independent of the regressor values), which is often indicated:

ei ∼ N (0, σ 2 ) (3A.4)

When these assumptions are satisfied, the user can go beyond simple data fitting to calculating

distribution parameters, constructing confidence intervals, etc. A zero mean for the residual

errors is an intrinsic property of linear least-squares regression. For the data considered here,

typical normality tests [31, 32, 33] verified that the observed data ūbo is lognormally distributed,

or Y = ln ūbo is normally distributed, and hence the residual errors are also normally distributed.

Furthermore, the residual errors were nearly constant when plotted against individual regressors

(not shown here).

When this assumption (Eq. 3A.4) is satisfied, the variance σ 2 is estimated from the

regression analysis as follows:


 
2 − ē)2
i (ei i (ei )
2
e e (Y − Xβ) (Y − Xβ) Y  (Y − Xβ)
σ = = = = = (3A.5)
d−1 d−1 d−1 d−1 d−1

where d degrees of freedom is equal to n minus the number of regressors. In this derivation, the

residual error mean ē = 0, and X  (Y − Xβ) = 0 (Eq. 3A.2). Listed as the ‘Residual standard

error’ (Table 3A.1), the square root of the variance or σ was estimated to be about 0.38 for

both regression models. In Sec. 3.5.2, this value was referred to as the dispersion of ūbo , which

38
is the standard deviation of Y (= ln ūbo ) (Eq. 2.2). The dispersion of ūbo equals σ because the

predictor Y̌ has zero variance (Y̌ = Xβ), which means that the variance of Y is identical to

the variance of the residual error e (Eq. 3A.1).

3A.3 Evaluation of β Coefficients

The significance of an individual regressor is determined by comparing its estimated coefficient

βk and corresponding standard error, estimated as


σβk = σ ckk (3A.6)

where ckk is the kth diagonal element of (X  X)−1 . The output of Table 3A.1 lists four statistics

for each regressor: the ‘estimate’ of βk , its ‘std. error’ (σβk ), the ‘t-value’, equal to βk ÷σβk , and

a probability ‘Pr(>|t|)’. The probability tests the hypothesis that βk is zero. Specifically, the

t-value results from substituting zero with mean βk and standard deviation σβk into a student’s

t-distribution, which reduces to a standard normal distribution when many degrees-of-freedom

are used. Pr(>|t|) lists the probability of the current t-value occurring if βk = 0 (or βk is

within an interval that contains zero). Thus, when the t-value for βk is sufficiently large or its

corresponding probability sufficiently small, the hypothesis that βk is zero is rejected, meaning

that the kth regressor belongs in the equation.

Applying these concepts to our output, all regressors for the model of Table 3A.1a have

large t-values and small probabilities of their coefficients being zero. In this sense, each coef-

ficient is determined to be in the range of maximum significance ‘∗∗∗’ (highest probability of

being nonzero). These high t-values justify keeping each regressor in the equation, including

the questionable regressor (ln η)2 . However, because the simplest reasonable model may be

preferred for practical application, and the R2 statistic (Eq. 3A.3) indicated little improvement

when (ln η)2 was included (Table 3A.1), Eq. 3.17 was offered as an alternative.

The merging of responses ūbo(unidirectional) , ūbyo and ūbo into one regression equation was

justified by analysis of their regression coefficients determined independently. Table 3A.2 shows

that the corresponding coefficients for the three responses are close in value. To improve on

this, a 95% confidence interval for each coefficient was computed from its estimated standard

39
Table 3A.2: R Project output for independent analysis of various responses

(a) Regression output for response due to unidirectional excitation


ln ūbo = β0 + β1 ln Tb + β2 ln η + β3 (ln η)2
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.66259 0.01243 53.305 < 2e-16 ***
lnTb -1.81138 0.01089 -166.278 < 2e-16 ***
lneta -1.51842 0.01785 -85.052 < 2e-16 ***
lneta2 -0.07261 0.01472 -4.931 8.42e-07 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3882 on 5876 degrees of freedom


Multiple R-Squared: 0.9047, Adjusted R-squared: 0.9046
F-statistic: 1.859e+004 on 3 and 5876 DF, p-value: < 2.2e-16
------------------------------------------------------------------

(b) Regression output for y-direction response due to bidirectional exci-


tation
ln ūbyo = β0 + β1 ln Tb + β2 ln η + β3 (ln η)2
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.63924 0.01238 51.648 < 2e-16 ***
lnTb -1.80681 0.01085 -166.571 < 2e-16 ***
lneta -1.57386 0.01778 -88.536 < 2e-16 ***
lneta2 -0.10195 0.01466 -6.953 3.96e-12 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3866 on 5876 degrees of freedom


Multiple R-Squared: 0.9067, Adjusted R-squared: 0.9066
F-statistic: 1.903e+004 on 3 and 5876 DF, p-value: < 2.2e-16
------------------------------------------------------------------

(c) Regression output for peak response due to bidirectional excitation


ln ūbo = β0 + β1 ln Tb + β2 ln η + β3 (ln η)2
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.76001 0.01155 65.825 < 2e-16 ***
lnTb -1.80266 0.01012 -178.149 < 2e-16 ***
lneta -1.56026 0.01658 -94.088 < 2e-16 ***
lneta2 -0.07727 0.01368 -5.649 1.68e-08 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3606 on 5876 degrees of freedom


Multiple R-Squared: 0.9183, Adjusted R-squared: 0.9182
F-statistic: 2.201e+004 on 3 and 5876 DF, p-value: < 2.2e-16

40
Table 3A.3: 95% confidence intervals for coefficients

Regressor 95% Confidence Interval Estimate

ūbo : 1-direction ūbyo : 2-directions ūbo : 2-directions

Intercept: 1D (0.6383...0.6868) (0.6150...0.6635) 0.6458

Intercept: 2D (0.7374...0.7826) 0.7701

ln Tb (-1.8327...-1.7900) (-1.8281...-1.7855) (-1.8225...-1.7828) -1.8070

ln η (-1.5534...-1.4834) (-1.6087...-1.5390) (-1.5928...-1.5278) -1.5508

(ln η)2 (-0.1015...-0.0438) (-0.1307...-0.0732) (-0.1041...-0.0505) -0.0839

error; these intervals are listed in Table 3A.3 and compared to the coefficients from analysis

of the combined responses (Table 3A.1a). The combined model should not be used if it gives

coefficients outside the confidence intervals for any of the individual models. However, the

estimated coefficients from Table 3A.1a are within all confidence intervals, justifying acceptance

of the combined model.

41
Appendix 3B: Peak Force due to Bidirectional Excitation

The force in the system (Eq. 3.19) can be written as

fb = Qz + kb ub (3B.1)

where fb = fbx , fby T , ub = ubx , uby T , and z = zx , zy T are vectors containing the x and

y-components. The magnitude of the vector fb is computed by the following:

|fb|2 = fb · fb

= (Qz + kb ub) · (Qz + kb ub)

= Q2 z · z + 2Qkb z · ub + kb2 ub · ub

= Q2 |z|2 + 2Qkb |z||ub| cos θ + kb2 |ub|2 (3B.2)

Presumably, the peak force occurs when the system is fully yielded (|z| = 1) and the deformation

is near its peak value (|ub| ≈ ubo ). Thus, with cos θ taking on any value from -1 to +1, the

limiting values of the force are estimated by

(Q − kb ubo ) ≤ fbo ≤ (Q + kb ubo ) (3B.3)

These upper and lower bounds correspond to the cases when the vectors z and ub are collinear

and point in the opposite directions, respectively. In general, the hysteretic contribution to the

force (Qz) is in the direction of instantaneous velocity [34], which is unlikely to be collinear

with the deformation. However, as the isolator deformation increases toward a maximum, the

yield surface translates outward such that the direction of instantaneous velocity is close to the

direction of deformation. The angle between the vectors is certainly less than 90 degrees, which

represents a translation along the yield surface that results in no deformation increase. If the

directions of z and ub vary by as much as 60 degrees, the upper bound estimate of fbo is about

15% conservative when Q and kb ubo are equal, and less conservative when they are unequal.

This indicates that the upper bound equation

fbo ≈ Q + kb ubo

is likely to be a good estimate of the peak force in the isolator.

42
Appendix 3C: Notation

An pseudo-spectral acceleration of a linear structure

Al , A cross-section areas of lead core and rubber, respectively, in a lead-

rubber bearing or isolation system

ay yield acceleration of fixed-base structure: fy /m

a∗y yield acceleration of isolated block: Q/m

fb (t) lateral force of isolator or isolation system

fbx , fby lateral force in x and y-directions due to bidirectional excitation

fbo peak lateral force in a single direction (unidirectional excitation) or

any direction (bidirectional excitation)

fbo /w isolator force coefficient

fo minimum strength for fixed-base structure to remain elastic

fy yield strength of nonlinear fixed-base structure

f¯y normalized yield strength of fixed-base structure: fy /fo

Gl , G shear moduli of lead core and rubber, respectively, in a lead-rubber

bearing

kn initial stiffness of nonlinear fixed-base structure

kb postyield stiffness of isolator or isolation system; stiffness of rubber

kef f effective stiffness of isolation system for equivalent-linear analysis

kI initial stiffness of isolator

kl stiffness of lead core prior to yielding

m mass of isolated block

Q characteristic strength of isolator or isolation system; yield strength

of lead core

Tb isolation period, corresponds to postyield stiffness

Td corner period separating velocity and displacement-sensitive regions

of linear spectrum

Tef f effective period of isolated block for equivalent-linear analysis

43
Tn initial period of nonlinear fixed-base structure

tr total thickness of rubber in a lead-rubber bearing

ub (t) or ub lateral deformation history of isolator or isolation system

ubx , uby lateral deformation in x and y-directions due to bidirectional excita-

tion

ubo peak lateral deformation in a single direction (unidirectional excita-

tion) or in any direction (bidirectional excitation)

ubo,84% median plus one standard deviation estimate of ubo

ubyo peak lateral deformation in the y-direction

ūb (t) or ūb normalized deformation of isolator: ub /u∗y

ūbo peak normalized deformation in a single direction (unidirectional ex-

citation) or in any direction (bidirectional excitation)

üg (t) or üg ground acceleration history

ügx , ügy components of ground acceleration for bidirectional excitation


¨g (t)
ū ground acceleration normalized by peak ground velocity

u̇go peak ground velocity for unidirectional excitation

u̇gyo peak ground velocity in y-direction for bidirectional excitation

ügo peak ground acceleration

uo peak deformation if structure remains elastic: fo /kn

uy yield deformation of isolator or structure

u∗y Q/kb

w weight of isolated block

z(t) or z yield function; dimensionless plastic variable governing yielding of

the isolator or isolation system

zx , zy x and y-components of yield function due to bidirectional excitation

δ dispersion of peak response over ensemble of ground motions

ζ damping ratio of fixed-base structure

ζef f effective damping ratio of isolated block for equivalent-linear analysis

44
η strength of isolator or isolation system normalized by peak ground

velocity: a∗y /(ωd u̇go )

η alternative normalized strength: a∗y /ügo

µ characteristic strength ratio of isolated block: Q/w

µs ductility of nonlinear fixed-base structure, or deformation relative to

yield deformation

ωb isolation frequency, corresponds to postyield stiffness of isolator

ωd corner frequency separating velocity and displacement-sensitive re-

gions of linear spectrum

ωn initial frequency of nonlinear structure

45
4: ESTIMATING THE PEAK DEFORMATION IN AN
ASYMMETRIC-PLAN SYSTEM

4.1 Introduction

The earthquake response of asymmetric, base-isolated structures has been studied using analyt-

ical, numerical and experimental techniques. Analysts working with a simplified model of the

system – a rigid structure on linear isolators – derived closed-form equations for the absolute

maximum or peak response to an acceleration impulse [35, 36]. When the system’s torsional

and lateral frequencies were equal, the corner deformation was predicted to be uninfluenced

by torsion because the peak rotation, although large, occurred with a significant time lag after

the peak lateral deformation [35, 36]. On the other hand, when the torsional frequency of the

system was much greater than its lateral frequency, the corner deformation was estimated by

summing the peak lateral and (now smaller) rotational deformations, which were likely to occur

near the same time [36]. Compared to this estimate, a code factor to account for the increase

in deformation in asymmetric-plan systems was unconservative for systems with torsional-to-

lateral frequency ratios less than 1.485, which, in our view, includes all isolated systems. A later

investigation of the same simplified system [37], which compared results from response history

analysis and response spectrum analysis to this static code factor, concluded that when the

torsional frequency is greater than the lateral frequency, response spectrum analysis is accurate

and the code factor is very conservative.

Researchers employing numerical techniques [38, 39, 40, 41] incorporated bilinear force-

deformation of the isolators and superstructure models of varying detail. Some included yield-

surface interaction of the isolators and varied system parameters over wide ranges [40, 41].

Despite the sophistication, earlier studies [38, 39] sought only to show that the superstructure

forces and deformations are reduced compared to a fixed-base structure, as had already been

accepted for symmetric systems. Clearly, torsion leads to significantly increased isolator defor-

mations only if large rotations and large lateral deformations at the center of mass (CM) occur

at or very near the same time during the earthquake, but some of these studies considered the

46
peak rotations and lateral deformations independently [38, 41]. One investigation [40] related

the increase in deformation at the corner relative to the CM to system parameters: superstruc-

ture or isolation eccentricity, superstructure period, and isolation or structure frequency ratios

[40], but the conclusions were based on only two ground motions.

Shake-table testing of an asymmetric base-isolated structure to triaxial base-excitation

[18] resulted in large rotations of the structure, which contributed significantly to the corner

deformation. However, eccentricities in the system, imposed by an asymmetric arrangement

of a small number of lead-rubber and natural-rubber bearings, may have been unrealistically

large: ebx /r = 0.20 and 0.39.

Although the analytical studies provide insight into the behavior of the system, they are

limited by oversimplification, and previous numerical studies lacked sufficient data to determine

reliably the influence of system parameters on corner deformation. Armed with faster computers

and a growing database of recorded ground motions, response trends can be determined based

on statistical analysis of response due to many ground motions, paving the way for better

preliminary estimates of response.

The objective of this chapter is to extend the techniques of Chapter 3 to asymmetric-

plan systems, thereby developing equations to estimate the peak lateral deformation among all

isolators. These equations are based on global properties of the system, which have been limited

to ranges most applicable to base-isolated buildings. Properties unique to isolation systems

allow one equation, determined from analysis of an idealized system subjected to bidirectional

excitation, to be used to estimate the deformation of many systems with varying planwise layout

of isolators, even though these systems are highly nonlinear. The results demonstrate that the

deformation of a corner isolator in an asymmetric system increases, relative to the deformation

of a symmetric system, by a factor that depends only on eccentricity between the CM and

center of rigidity (CR) and distance from the CM to the outermost corner. The code static

procedure for determining the design displacement is evaluated in light of the results presented.

47
CM axis

Rigid slab w/ mass m


and radius of gyration r y
CR (ebx , eby )
CM x

Isolator i
Position (xi , yi )
Properties kbi , Qi
Deformations ubxi , ubyi

Figure 4.1: A rigid slab supported on a system of isolators

4.2 Asymmetric Base-Isolated System

4.2.1 System Considered

Consider a rigid slab with mass m uniformly distributed in plan, and radius of gyration r,

supported on a system of isolators (Figure 4.1). Each isolator is distinguished by its position

(xi , yi ) from the CM, postyield stiffness kbi and yield strength Qi . The properties kbi and Qi

are independent of the direction of loading. The x and y-components of force in the ith isolator,

extended from Eq. 3.19, are


⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎨fbxi ⎪
⎪ ⎬ ⎪
⎨ubxi ⎪
⎬ ⎪
⎨zxi ⎪

= kbi + Qi (4.1)
⎩fbyi ⎪
⎪ ⎭ ⎪
⎩ubyi ⎪
⎭ ⎪
⎩zyi ⎪

where ubxi and ubyi are the x and y-components of isolator deformation, and zxi and zyi are the x

and y-components of the yield function. These components interact when loaded bidirectionally

(Sec. 3.6.1), such that the magnitude of the yield function equals 1 when the isolator is yielding

and is less than 1 otherwise. Yielding depends on the initial stiffness, determined for every

isolator from an assumed yield deformation of 1 cm (Sec. 3.3.1).

48
4.2.2 Equations of Motion

Equations governing ubx and uby , the x and y-components of deformation at the CM, respec-

tively, and rotation θbz about a vertical axis are



mübx + fbxi = −mügx (4.2a)
i

mr 2 θ̈bz + (−fbxi yi + fbyi xi ) = 0 (4.2b)
i

müby + fbyi = −mügy (4.2c)
i

Equation 4.1 is substituted for fbxi and fbyi , with isolator deformations expressed in terms of

CM deformations: ubxi = ubx − θbz yi and ubyi = uby + θbz xi , resulting in


⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪ ⎪

⎪ ü ⎪
bx ⎪ k −k
eby
0 ⎪
⎪ u ⎪
bx ⎪ ⎪
⎪ Q z ⎪
⎪ ⎪
⎪mü gx ⎪


⎨ ⎪ ⎢
⎬ ⎢
b b r ⎥ ⎪
⎨ ⎪
⎬ ⎨ ⎪ i i xi ⎪
⎬ ⎪
⎨ ⎪

⎥  

m r θ̈bz + ⎢−kb eby 1 e bx ⎥ + − y i x i = −
⎪ ⎪ kbθz r 2 kb r ⎥ ⎪ bz ⎪ ⎪
rθ Q
i i r xiz + Q z
i i r yi ⎪ ⎪ 0 ⎪

⎪ ⎪
⎪ ⎣ r
⎦⎪⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪


⎩ ü ⎭ ⎪ ebx ⎪
⎩ ⎪
⎭ ⎪
⎩  ⎪
⎭ ⎪
⎩ ⎪

by 0 kb r kb uby i Q i z yi mügy
(4.3)

where kb is the lateral postyield stiffness of the isolation system, kbθz is the torsional postyield

stiffness about the CM, and ebx and eby are the x and y-components of the stiffness eccentricity,

all defined as follows:


   
kb = kbi kbθz = kbi x2i + yi2 (4.4a)
i i
1  1 
ebx = kbi xi eby = kbi yi (4.4b)
kb kb
i i
Identical to the postyield stiffness for the single isolator system (Sec. 3.2.1), kb applies in any

lateral direction. The stiffness eccentricity defines the location of the CR relative to the CM,

where the CR is the point about which the first moment of isolator stiffnesses is zero, and a

static horizontal force applied through this point causes no rotation.

Strength properties analogous to the stiffness properties of Eq. 4.4 include the lateral

strength Q (defined identically in Sec. 3.2.1), the torsional strength about the CM, Qθz , and

the strength eccentricity, with components epx and epy in the x and y-directions, all defined

below:
 
Q= Qi Qθz = Qi (x2i + yi2 ) (4.5a)
i i

49
1  1 
epx = Qi xi epy = Qi y i (4.5b)
Q Q
i i
Together, these parameters describe the distribution of strength in the system, but unlike the

stiffness parameters, cannot be separately identified in the equations of motion (Eq. 4.3) due

to the system nonlinearity. Previous studies apparently ignored strength eccentricity [38, 41],

considered strength eccentricity instead of stiffness eccentricity [40], or considered both [39].

Later assumptions will clarify how the strength parameters should be defined.

Dividing Eq. 4.3 through by m gives


⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪
⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ügx ⎪

eby

⎪ übx ⎪⎪ ⎢ 1 − r 0 ⎥⎪ ⎪ ubx ⎪
⎪ ⎪
⎪ i γi zxi ⎪
⎪ ⎪
⎪ ⎪
⎨ ⎬ ⎢ ⎥ ⎨ ⎬ ⎨ ⎬ ⎨ ⎪ ⎬
  
+ ωb2 ⎢ eby
⎢− r Ω2θz erbx ⎥ yi
⎥ ⎪rθbz ⎪ + µg ⎪ i γi − r zxi +
xi = − (4.6)
⎪ r θ̈b ⎪
⎪ ⎪ ⎣ ⎦ ⎪ ⎪ ⎪
z
r yi ⎪
⎪ ⎪

0 ⎪


⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪ ⎪
⎩ü ⎪ ⎭ 0 ebx
1

⎩u ⎪ ⎭ ⎪
⎩ ⎪
⎭ ⎪
⎩ü ⎪ ⎭
by r by i γi zyi gy

where
ωbθz
Ωθz = (4.7)
ωb
and  
kb kbθz
ωb = ωbθz = (4.8)
m mr 2
are the lateral and torsional isolation frequencies. As Ωθz , the ratio of the torsional to the lateral

frequency, increases, the system becomes stiffer in torsion relative to its lateral stiffness. To

obtain Eq. 4.6, the nonlinear terms in Eq. 4.3 were rewritten by introducing γi = Qi /Q, allowing

Q to be factored out; µg was substituted for Q/m. The isolation frequency ωb and strength µ

are identical to corresponding properties that govern the response of the single isolator system

(Sec. 3.2.2); for a given asymmetric system the single isolator system with the same values of

ωb and η shall be known as the corresponding symmetric system.

Equation 4.6 indicates that the deformations ubx , rθbz and uby of the asymmetric system

depend on ωb , µ, Ωθz , ebx /r and eby /r, which shall be referred to as the global properties of

the system, and on the planwise layout of isolators. However, the response of a linear system

(µ = 0) depends only on the global properties and is independent of the isolator layout.

50
4.2.3 Characteristics of Response Histories of Asymmetric System

Response histories are presented for the above defined inelastic, asymmetric system for the

special case of unidirectional (y-direction) excitation ügy (t) and one-way asymmetry (eby =0),

resulting in ubx (t) = 0. The deformations at locations x = ± r from the CM:

u±ry (t) = uby (t) ± rθbz (t) (4.9)

depend only on the system’s global properties and planwise layout of isolators, and thus not

explicitly on the distance r (Eq. 4.6). When eccentricity ebx /r is positive, u−ry refers to the

‘flexible’ side, which tends to deform more than the ‘stiff’ side.

Consider a rectangular-plan system with a length (x-dimension) to width (y-dimension)

ratio of 2 that is asymmetric due to an imbalance of isolator stiffnesses; this idealized system

is developed in detail in Sec. 4.3. Figure 4.2 presents the response histories for a linear system

(µ = 0) with Tb = 2.5 seconds and ebx /r = 0.1 for two different frequency ratios, Ωθz = 1 and

1.25, subjected to a selected ground motion. Stiffness proportional damping that would reduce

to ζ = 0.05 if the system were symmetric is included. The deformation ub of the corresponding

symmetric system is shown dotted for reference. The response of this single isolator system

was studied in detail in Chapter 3. Also shown are the lateral deformation uby at the CM, the

rotational deformation rθbz , and resulting deformations u+ry and u−ry .

The response histories of this linear system are influenced significantly by the frequency

ratio. When the torsional and lateral frequencies are equal (Ωθz = 1, Fig. 4.2a), the peak

rotational deformation rθbz is greater than 50% of the peak lateral deformation ubyo . Concurrent

to a slow build-up of the rotational component, the lateral component dies out, and thus differs

significantly from the deformation ub of the symmetric system. However, when the torsional

frequency exceeds the lateral frequency (Ωθz = 1.25, Fig. 4.2b), the peak rotational deformation

rθbzo is only 20% of the peak lateral deformation ubyo , and absent a rotation build-up, the lateral

component appears to be indistinguishable from the deformation ub of the symmetric system.

Analytical studies of response to a ground acceleration impulse [36] demonstrated that

for systems with equal lateral and torsional frequencies (Ωθz = 1), the time lag of the peak

51
(a) (b)
25
−25
ub (t) ub (t)
0
25

ubo =26.37 cm −25 ubo =26.37 cm


0
25
uby (t) uby (t)
−25
0

25
Deformation (cm)

−25 ubyo =26.68 cm


ubyo =25.36 cm 25
0 rθbzo =13.87 cm
rθbz (t) rθbzo =5.37 cm rθbz (t)
0
−25
−25
25 u+ryo =26.71 cm 25
u+ry (t) u+ry (t)
0 0

−25 −25 u+ryo =24.06 cm


u−ryo =27.89 cm 25 u−ryo =29.53 cm
u−ry (t) u−ry (t)
25
0
0
−25
−25
0 10 20 30 40 0 10 20 30 40
Time (s) Time (s)

Figure 4.2: Deformation ub of a linear symmetric system and deformations uby , rθbz , u+ry
and u−ry of a linear, asymmetric system with Tb = 2.5 sec, ebx /r = 0.1, and (a) Ωθz = 1 and
(b) Ωθz = 1.25, subjected to the strong component of LMSR Record No. 12

rotational deformation, observed in Fig. 4.2a, prevented it from contributing to the corner de-

formation. For systems with torsional frequency much greater than lateral frequency (Ωθz  1),

although the rotational deformation was much smaller, the peak lateral and rotational deforma-

tions were assumed to occur at the same time to allow an estimate of the corner deformation.

These features are discernible in Fig. 4.2, although the conclusions of [36], based on impulse

excitation, apply only partially for realistic ground motions, which is not surprising.

An important effect of nonlinearity is to reduce the influence of frequency ratio, as il-

lustrated in Fig. 4.3 for a system with µ = 0.06. The rotational deformation does not build

up over time for Ωθz = 1, and though its maximum is still larger for Ωθz = 1 than 1.25, the

difference is less significant. In contrast to linear systems, both the lateral (uby ) and rotational

52
(a) (b)
25 25
ubo =15.45 cm ubo =15.45 cm
ub (t) ub (t)
0 0

−25 −25
25 25
ubyo =15.52 cm ubyo =15.11 cm
uby (t) uby (t)
0 0
Deformation (cm)

−25 −25
25 25
rθbz (t) rθbzo =1.20 cm rθbz (t)
0 0
rθbzo =2.02 cm
−25 −25
25 25
u+ryo =13.94 cm u+ryo =14.49 cm
u+ry (t) u+ry (t)
0 0

−25 −25
25 25
u−ryo =17.14 cm u−ryo =15.80 cm
u−ry (t) u−ry (t)
0 0

−25 −25

0 10 20 30 40 0 10 20 30 40
Time (s) Time (s)

Figure 4.3: Deformation ub of a nonlinear symmetric system and deformations uby , rθbz , u+ry
and u−ry of a nonlinear asymmetric system with Tb = 2.5 sec, µ = 0.06, ebx /r = 0.1, and (a)
Ωθz = 1 and (b) Ωθz = 1.25, subjected to the strong component of LMSR Record No. 12

(rθbz ) deformations die out quickly for the nonlinear system, and the ratio of peak rotational to

lateral deformations is smaller; compare rθbzo /ubyo = 0.13 for Ωθz = 1 and 0.08 for Ωθz = 1.25

(Fig. 4.3) to 0.55 and 0.20 for the linear system (Fig. 4.2). Thus, the lateral deformation history

uby of the asymmetric system closely resembles ub of the corresponding symmetric system.

4.2.4 Peak Response to Bidirectional Excitation

The deformations u±ry , applicable for one component of excitation, can be generalized to a

more meaningful measure of deformation for bidirectional (two-component) ground excitations.

Let uro be the peak deformation over all rx , ry  located at distance r from the CM, thereby

satisfying:

rx2 + ry2 = r 2 (4.10)

53
The x and y-components of deformation at location rx , ry  are, respectively:

urx (t) = ubx (t) − ry θbz (t) ury (t) = uby (t) + rx θbz (t) (4.11)

which is an extension of Eq. 4.9, and the magnitude of deformation is



ur (t) = u2rx (t) + u2ry (t) (4.12)

By the method of Lagrange multipliers, ur (t) is maximized subject to the constraint of Eq. 4.10,

the details of which are provided in Appendix 4A. The simple result is

uro = max ( ub (t) + r|θbz (t)| ) (4.13)


t

where ub (t) = [u2bx (t) + u2by (t)]1/2 is the magnitude of the deformation at the CM. Implicit in

Eq. 4.13, the deformation is first maximized over the circle of radius r at each time instant and

then maximized over time. Like u±ry , uro depends on the global properties and planwise layout

of isolators; however, dependence on the latter is shown to be negligible in Sec. 4.3.

4.2.5 Normalized Equations of Motion

Next, the equations of motion for the asymmetric system are normalized comparable to the

equation for the single isolator system (Sec. 3.3.2). Let a∗y and u∗y for an asymmetric system be

defined as for the single isolator system: a∗y = µg and

Q a∗y
u∗y = = 2 (4.14)
kb ωb

The equations of motion (Eq. 4.6) are divided by u∗y , resulting in


⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪
⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ügx ⎪

eby


¨
ūbx ⎪ ⎪ ⎢ 1 − r 0 ⎥⎪ ⎪ ūbx ⎪
⎪ ⎪
⎪ i γi zxi ⎪
⎪ ⎪
⎪ ⎪

⎨ ⎬ ⎢ ⎥ ⎨ ⎬ ⎨ ⎬ ⎨ ⎬
   ω 2
r θ̄¨bz ⎪ + ωb ⎢ ⎥ −
2 eby
⎢− r Ω2θz ebx + ωb2 yi
i γi − r zxi +
xi = b
⎪ r ⎥ ⎪r θ̄bz ⎪ ⎪ z
r yi ⎪ a∗y ⎪
0 ⎪

⎪ ⎪
⎪ ⎣ ⎦⎪⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪


⎩ ū ⎪ ⎪
⎩ ū ⎪ ⎪  ⎪ ⎪
⎩ü ⎪
¨by ⎭ 0 ebx
r 1 by
⎭ ⎩
i γi zyi

gy

(4.15)

where x̄ notation denotes response x normalized by u∗y . Identical to the definition for bidirec-

tional analysis of a single isolator system (Eq. 3.23), η is the strength of the isolation system

normalized relative to ground motion intensity:


a∗y
η= (4.16)
ωd u̇gyo

54
with u̇gyo and ωd referring to the stronger component of ground motion, always applied in the

y-direction. Defining a∗y in terms of η, Eq. 4.15 becomes


⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪

⎪ ¨
ūbx ⎪ ⎪ e
− rby 0 ⎥⎪ ⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ¨gx ⎪⎪


⎨ ⎪
⎬ ⎢ 1 ⎪ ūbx ⎪ ⎪ i γi zxi ⎪ ⎪ ū ⎪

2 ⎢ eby
⎥⎨ ⎬ ⎨
  
⎬ ω 2 ⎨ ⎬
¨ ebx ⎥ + ωb2 yi − b
⎢− r i γi − r zxi +
xi
⎪ r θ̄ bz ⎪ + ω Ω2θz r ⎥ ⎪r θ̄bz ⎪ ⎪ z
r yi ⎪
= 0
⎪ ⎪
b
⎣ ⎦⎪ ⎪ ⎪ ⎪ ηωd ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪

⎩ ū
¨by ⎭ 0 ebx
1 ⎩ ūby ⎭ ⎩ ⎭ ⎩ū
¨gy ⎭
r i γi zyi
(4.17)
¨gx = ügx /u̇gyo and ū
with ū ¨gy = ügy /u̇gyo ; the preceding normalization implies that larger of the

two velocity components, ū˙ gy (t), varies between -1 and +1.

Similar to the single isolator system (see Secs. 3.3.2 and 3.6.2), Eq. 4.17 indicates that any

median normalized deformation of interest (such as ūbx , r θ̄bz and ūby ) over an ensemble of ground

motions depends primarily on the system’s global properties that were defined previously, with

strength now characterized by η instead of µ. Furthermore, the median normalized deformation

is weakly affected by ground motion intensity, and the dispersion of normalized deformation over

the ensemble is expected to be small, as proven for the single isolator system in Sec. 3.3.2. These

properties permit meaningful statistical analysis of the relevant deformations of the asymmetric

system.

4.2.6 Computation of Median Response

For given global properties, the median deformation over an ensemble of ground motions, with

weak and strong-component ensembles applied in the x and y-directions, respectively, is deter-

mined by the following steps: (1) Compute the peak normalized deformation by nonlinear RHA

of Eq. 4.17 to every motion in the ensemble, (2) compute the median value (over the ensemble)

of the normalized deformation by Eq. 2.1, and (3) convert the normalized deformation to actual

deformation by multiplying by u∗y (Eq. 4.14), expressed as a function of η (Eq. 4.16), design

PGV u̇gyo and the corner frequency ωd for the strong-component ensemble.

55
4.3 Modeling Asymmetric-Plan Systems

4.3.1 Parameter Ranges

To focus this study, the ranges of global properties most applicable to base-isolated buildings are

determined first. Based on Chapter 3, the isolation period Tb is varied from 1 to 4 seconds and

the normalized strength η from 0.25 to 1.5. A linear system (η = 0) with stiffness proportional

damping that reduces to 5% in a symmetric system, is occasionally considered for comparison.

Previous studies considered the frequency ratio Ωθz varying from 0.8 to 1.7 [40], and normalized

eccentricities ebx /r and eby /r upwards of 0.2 [18, 36, 40, 41]; however, these are unrealistically

wide ranges of parameters, as will be shown presently.

Consider a system with uniformly distributed mass and identical isolators located regu-

larly over the building plan. Since isolators are mass-produced by manufacturers, limiting to

one prototype is economical, justifying the assumption of identical isolators. Exceptions occur,

for example, when the isolators in the interior of the plan are larger and stiffer than those on

the exterior to support larger tributary mass, or elastomeric bearings are used in the interior

and lead-rubber bearings on the exterior to increase the torsional resistance. With identical

isolators, the planwise distributions of strength and stiffness are identical, implying that the

ratio of torsional to lateral strengths Qθz /Q (Eq. 4.5a) equals the ratio of the torsional to lat-

eral stiffness kbθz /kb (Eq. 4.4a), and the strength eccentricities epx and epy (Eq. 4.5b) equal the

stiffness eccentricities ebx and eby (Eq. 4.4b).

Given these assumptions, as the number of isolators approaches infinity, Eq. 4.4 becomes

 
kbθz ⇒ k̄b x2 + y 2 dA (4.18a)
A
 
1 1
ebx ⇒ k̄b xi dA eby ⇒ k̄b yi dA (4.18b)
kb A kb A

where k̄b is the stiffness per unit area. When the stiffness and mass are both uniformly

distributed, Eq. 4.18a results in kbθz = kb r 2 with r equal to the mass radius of gyration, Ωθz

(Eq. 4.7) reduces to 1, and the eccentricities (Eq. 4.18b) become ebx = eby = 0. Thus, for a

system containing many isolators, the frequency ratio is close to 1 and the eccentricity is close

to 0; the former conclusion was also reached by Lee [38].

56
(a) Rectangular Plan (b) L-plan
1.5
b/d=1 b/t=2
1.4 2 3
3 5
1.3 5 10
10
Ωθz

1.2

1.1

1
0 50 100 150 200 0 50 100 150 200
Number of isolators Number of isolators
(c) L-plan
0.06
b/t=2
0.05 3
5
ebx /r = eby /r

0.04 10
0.03

0.02

0.01

0
0 50 100 150 200
Number of isolators

Figure 4.4: Torsional to lateral frequency ratio Ωθz for (a) rectangular-plan system with
aspect ratios b/d = {1, 2, 3, 5, 10} and (b) asymmetric L-plan system with length-to-thickness
ratios b/t = {2, 3, 5, 10}; and (c) eccentricity for L-plan system, all presented as a function of
plan-dimension ratio and number of isolators

These conclusions are confirmed by numerical results for a symmetric rectangular-plan

system and an asymmetric L-plan system that satisfy the above assumptions. Figure 4.4a and

b show the frequency ratios Ωθz for a rectangular plan with various aspect (length-to-width)

ratios b/d, and an L-plan (Plan A1 in Fig.4.5a) with various length-to-thickness ratios b/t. This

figure confirms a lower bound frequency ratio of 1, and suggests an upper bound of 1.3 for any

system with more than 10 isolators, thus Ωθz is selected as {1 - 1.3}. Figure 4.4c shows that the

maximum normalized eccentricity in either direction (ebx /r, eby /r) for the L-plan is less than

0.06. The frequency ratios for two other examples, Plans A2 and A3 in Fig. 4.5a, fall within

the range suggested above. The normalized eccentricity is largest (ebx /r = 0.108) for Plan

A2, which has the smallest number of isolators and irregular spacing. Because it was hard to

57
(a)
Plan A1: Plan A2: Plan A3:
ebx /r = eby /r = 0.049 ebx /r = 0.108, eby /r = 0 ebx /r = 0, eby /r = 0.057
Ωθz = 1.180 Ωθz = 1.280 Ωθz = 1.172
t 3
3

3 4
d 4
b d
2 1
2
t
1
2 1
b
d (=b) b

(b) Idealized System:


Vary isolator stiffnesses to Isolators
achieve desired properties
CM
CR

Figure 4.5: (a) Configuration of 3 asymmetric-plan systems, and (b) configuration of idealized
system

contrive even these examples, normalized eccentricity will typically be smaller than 0.1. These

bounds derived for frequency ratio and eccentricity are not intended to be firm, but simply

suggest meaningful ranges upon which to focus analysis and conclusions.

4.3.2 Geometry and Planwise Layout of Isolators

As discussed earlier, for a linear isolation system, uro is independent of the planwise layout of

isolators. Although individual elements typically factor into the response of a nonlinear system,

the planwise layout of isolators has little influence on the deformations of a nonlinear isolation

system. To demonstrate, consider an idealized, rectangular-plan system (Fig. 4.5b) with an

aspect ratio b/d = 2, where the stiffnesses of individual isolators are varied to match the global

properties Tb , η, Ωθz , ebx /r and eby /r of Plans A1-A3, and the strength of each isolator is varied

in the same way as its stiffness.

58
(a) Plan A1 (b) Plan A2 (c) Plan A3
30
Plan A1 Plan A2 Plan A3
Idealized Idealized Idealized
25

20 η=0 η=0 η=0


uro (cm)

0.25 0.25 0.25


15
0.5 0.5 0.5
10
1.0 1.0 1.0
5 1.5 1.5 1.5

0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Tb (sec) Tb (sec) Tb (sec)

Figure 4.6: Comparison of median deformations uro of systems with Plans A1-A3 and the
idealized, rectangular-plan (b/d = 2) system with identical global properties

Computed for the LMSR ground motion ensemble with median PGV of 35 cm/s (Sec. 4.2.6),

Fig. 4.6 compares median uro for each of Plans A1-A3 and their corresponding idealized sys-

tems. The deformations of a linear system with η=0, known to be identical for the actual

and idealized systems, are shown for reference. The deformations of the two systems appear

identical, with discrepancies less than 1.5% over all values of Tb and η for all three plans.

These examples suggest that the idealized system is representative of various geometries with

different isolator layouts. The discrepancies for different plans are minimized because isolators

have identical resisting properties in any direction and tend to be arranged at regular intervals

regardless of the plan.

To verify that the idealized, rectangular system with b/d = 2 can represent a sufficiently

wide range of asymmetric plans, this comparative analysis is extended to the extremes, a square

plan with b/d = 1 and a slender plan with b/d = ∞. Figure 4.7 compares median uro , computed

as above, for these extreme plans and the idealized system with identical global properties. The

deformations of a linear system, identical for the actual and idealized systems, are shown again

for reference. Even though these geometries result in greater differences in isolator layout than

before, the deformations compared in Fig. 4.7 still appear identical, with discrepancies less than

2% over the set of global properties considered.

59
30
(a) b/d=1 (b) b/d=∞
Idealized Idealized
25

20 η=0 η=0
uro (cm)

0.25 0.25
15
0.5 0.5
10
1.0 1.0

5 1.5 1.5

0
0 1 2 3 4 0 1 2 3 4
Tb (sec) Tb (sec)

Figure 4.7: Comparison of median deformations uro of systems with extreme aspect ratios [(a)
b/d = 1 and (b) b/d = ∞] and the idealized, rectangular-plan system (b/d = 2) with identical
global properties: ebx /r = 0.1 and Ωθz = 1.25

(a) (b) y
y

x
CR ügx
x
ügx CM CM CR

ügy ügy

Figure 4.8: (a) Two-way asymmetric system with ground motions applied in the x and y  -
directions and (b) idealized system with same ground motions applied in the x and y-directions;
both with identical global properties Tb , η, Ωθz , and magnitude of eccentricity eb /r (not to scale)

Given the outcome of these tests, we conclude that the median deformation, such as uro ,

based only on the global properties of the system is sufficiently accurate, differences in geometry

and isolator layout are unimportant, and the idealized, rectangular-plan system with b/d = 2

(Fig. 4.5b) can represent a wide range of asymmetric plans.

4.3.3 Two-way Asymmetric Systems

Consider a two-way asymmetric system with respect to a rotated x -y  coordinate system,

where the x -axis is directed toward the CR. The eccentricities in the x and y  -directions are

ebx = e2bx + e2by and eby = 0. Thus, a two-way asymmetric system can be viewed after axis

60
(a) Ωθz = 1 (b) Ωθz = 1.25
30
2-way asymmetric
Idealized
25

20 η=0 η=0
uro (cm)

15 0.25 0.25

0.5 0.5
10
1.0 1.0

5 1.5 1.5

0
0 1 2 3 4 0 1 2 3 4
Tb (sec) Tb (sec)

Figure 4.9: Comparison of median deformations uro of a 2-way asymmetric system with
rotated axes and the idealized, rectangular-plan system with identical global properties; ebx /r =
0.1 and (a) Ωθz = 1 and (b) Ωθz = 1.25

rotation as a one-way asymmetric system. The other global properties of the system Tb , η, and

Ωθz are unaffected by axis rotation. Because isolators have identical resisting properties in any

direction, Tb and η remain the same; Ωθz (Eq. 4.7) is unchanged because kbθz (Eq. 4.4) and r

are computed about the vertical CM axis.

Consider the idealized system of Fig. 4.5b with global properties identical to those of the

two-way asymmetric system with rotated axes; the eccentricity ebx /r of the idealized system

equals the eccentricity ebx /r of the two-way asymmetric system. If the system is linear (η = 0),

the response of this idealized system to ground motions applied in the x and y-directions is

identical to the response of the two-way asymmetric system to the same motions now applied in

the x and y  -directions (Fig. 4.8). For nonlinear isolation systems, Sec. 4.3.2 suggests that this

one-way asymmetric idealized system should give a highly accurate estimate to the response.

Figure 4.9 compares median uro , computed as before, for a two-way asymmetric system (rectan-

gular plan as in Fig. 4.8a) with eccentricities ebx /r = 0.086 and eby /r = 0.05 and the idealized

system (Fig. 4.8b) with identical global properties and ebx /r = (0.086)2 + (0.05)2 = 0.1. The

deformations of the two systems appear to be identical, with discrepancies limited to 1.1% over

the set of global properties considered.

61
These results demonstrate that the response of a two-way asymmetric system is predicted

accurately by a one-way asymmetric model with the same magnitude of eccentricity. Therefore,

the results and conclusions presented in subsequent sections, while based on the idealized, one-

way asymmetric system, are applicable to any two-way asymmetric system with eccentricity

equal to the distance between its CM and CR.

4.4 Median Response of Asymmetric Systems

The median deformations uro (Eq. 4.13) of the asymmetric system and ubo of its corresponding

symmetric system – both systems subjected to the same bidirectional ground excitation – are

compared, as well as the deformation ratio:

uro
ûro = (4.19)
ubo

a measure of the increase of deformation due to torsion. Results are presented over the ranges

of global properties (Tb , η, Ωθz and eb /r) defined in Sec. 4.3.1, for the idealized, rectangular-

plan system of Fig. 4.5b with a one-way, x-direction eccentricity eb /r. The subscript x in ebx /r

is dropped because, as demonstrated in the previous section, the direction of eccentricity is

unimportant. The median deformation is computed for the LMSR ensemble with a median

PGV of 35 cm/s for the strong component of excitation, by the procedure of Sec. 4.2.6.

Figure 4.10 presents the median deformations uro and ubo and median deformation ratio

as a function of Tb for several different normalized strengths η and fixed Ωθz = 1.25 and

eb /r = 0.1. Included are the deformations of a linear system with η = 0. The deformation uro

in Fig. 4.10a of nonlinear systems displays smoother variation with Tb compared to the linear

system (η = 0). In Fig. 4.10b, the mostly unsystematic variations of ûro with Tb and η fall

between 1.05 and 1.15, implying that for an eccentricity eb /r of 0.1, plan asymmetry leads to

a 5 to 15% increase in the deformation uro at a distance r from the CM, over the deformation

of the corresponding symmetric system.

When the results for uro , ubo and ûro are presented as a function of normalized strength

η for different values of Tb , the deformations of the asymmetric and corresponding symmetric

system are seen to be similar, decreasing monotonically with increase in strength (Fig. 4.11a),

62
30 1.5
(a) Symmetric (b) η=0
Asymmetric 0.25
25 0.5
1.4
1.0
20 η=0 1.5
ubo , uro (cm)

1.3
0.25

ûro
15
0.5 1.2
10
1.0
1.1
5 1.5

0 1
0 1 2 3 4 0 1 2 3 4
Tb (sec) Tb (sec)

Figure 4.10: (a) Median asymmetric system deformations uro and deformations ubo of a
corresponding symmetric system, and (b) median deformation ratio ûro , both plotted against
Tb ; Ωθz = 1.25 and eb /r = 0.1

20 1.5
(a) Symmetric (b) Tb =1.5 s
Asymmetric 2.5 s
1.4 3.5 s
15
ubo , uro (cm)

1.3
ûro

10
1.2

5
1.1

0 1
0 0.5 1 1.5 0 0.5 1 1.5
η η

Figure 4.11: (a) Median asymmetric system deformations uro and deformations ubo of corre-
sponding symmetric system for Tb = 2.5 sec, and (b) median deformation ratio ûro for varying
Tb , both plotted against strength η; Ωθz = 1.25 and eb /r = 0.1

and the deformation ratio ûro varies unsystematically, with weak dependence on strength

(Fig. 4.11b).

Figure 4.12 shows the median uro against frequency ratio for several values of η, eb /r =

0.1, and Tb = 2 and 4 and seconds; wherein the most relevant range of 1.0-1.3 for Ωθz is indicated.

The deformation ubo of the corresponding symmetric system, unaffected by frequency ratio, is

included for reference. The median deformations of the linear asymmetric system, included

63
(a) Tb = 2.0 sec (b) Tb = 4.0 sec
30
Symmetric
Asymmetric η=0
25

20
ubo , uro (cm)

η=0
15 0.25
0.25 0.5
10
0.5 1.0

5 1.0
1.5
1.5

0
0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
Ωθz Ωθz

Figure 4.12: Median deformations uro and ubo of the asymmetric and corresponding symmetric
system, respectively, plotted against frequency ratio Ωθz ; eb /r = 0.1, (a) Tb = 2 sec, and (b)
Tb = 4 sec

for comparison, display a slight dip in the neighborhood of Ωθz = 1. However, in a nonlinear

system these deformations are essentially independent of the frequency ratio, even outside the

range most relevant to base-isolated buildings (Fig. 4.12), verifying our observations of the

response histories (Sec. 4.2.3). This suggests that intentionally increasing the relative torsional

stiffness of the system, as is sometimes done in design, is unlikely to have much effect on the

response. This contradicts the conclusion of [40] that a decrease in frequency ratio (with all

other properties fixed) leads to increased corner deformation. Figures 4.10-4.12 together leave

the impression that the increase in deformation in an asymmetric system, remarkably, depends

only weakly on each of Tb , η, and Ωθz . Developing a design factor to estimate the deformation

ratio ûro should be straightforward.

Since all results so far were for fixed eccentricity, Fig. 4.13 presents the variation of median

uro and ubo , and ûro against normalized eccentricity eb /r for a system with Tb = 2.5 seconds,

Ωθz = 1.25, and several values of η. Eccentricities eb /r up to 0.2, beyond the range identified

earlier, are considered, to admit special design situations that we may have missed. Figure 4.13a

shows that uro increases with increasing eccentricity, and suggests that the increase is roughly

linear, which Fig. 4.13b confirms, except perhaps for large values of η. This figure also shows

that the increase of ûro with eccentricity varies only slightly with normalized strength.

64
20 1.5
(a) Symmetric (b) η=0.25
Asymmetric 0.5
1.4 0.75
15 1.0
η=0.25
1.5
ubo , uro (cm)

1.3
0.5

ûro
10
0.75
1.0 1.2

5 1.5
1.1

0 1
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Eccentricity ebx /r Eccentricity ebx /r

Figure 4.13: (a) Median asymmetric system deformations uro and deformations ubo of corre-
sponding symmetric system and (b) median deformation ratio ûro , both plotted against eb /r;
Tb = 2.5 sec and Ωθz = 1.25

4.5 Estimation of Peak Deformation

4.5.1 Equations to Estimate uro

An equation for the deformation ratio ûro (Eq. 4.19) is now developed by regression analysis

of the data presented in Sec. 4.4; this equation multiplied by the deformation ubo (Sec. 3.7) of

a corresponding symmetric system will provide an estimate of the resultant deformation uro of

the asymmetric system.

The regression equation for ûro is required to be linear upon logarithmic transformation,

consistent with Eq. 3.14a for ubo (Appendix 4B). The chosen regression equation should also

reduce to ûro = 1 for symmetric systems (eb /r = 0) and approach this limit continuously. To

satisfy this theoretical requirement, the equation for ln (ûro ) cannot vary linearly with ln (Tb ),

ln (η) or ln (Ωθz ), which is permissible since the median of ûro depends only weakly on each

of these properties (Sec. 4.4), and supplementary regression analyses showed the coefficients

associated with these terms to be insignificant (Appendix 4B).

The variation of ûro with normalized eccentricity eb /r (Fig. 4.13) suggests ln (1 + eb /r)

be included in the regression equation; a factor of 1 is added so that this term reduces to zero

when the eccentricity is zero. Because the other global properties are not included as variables

65
(a) Tb = 2.0 sec (b) Tb = 4.0 sec
20
exact median
design equation

15 η=0.25
η=0.25
0.5
uro (cm)

10 0.5 0.75
0.75 1.0
1.0
1.5
1.5
5

0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Eccentricity eb /r Eccentricity eb /r

Figure 4.14: Design equation for uro (Eq. 4.22) compared to exact median computed by
nonlinear RHA to LMSR ground motion ensemble with median PGV = 35 cm/s, for systems
with Ωθz = 1.15; (a) Tb = 2 sec, (b) Tb = 4 sec

in the regression equation, its form is simple:


 eb   eb β1
ln (ûro ) = β1 ln 1 + ⇐⇒ ûro = 1 + (4.20)
r r

and, by omitting the constant coefficient, the theoretical requirement that ûro = 1 when eb /r =

0 is satisfied. This model is shown to be suitable in Appendix 4B, and the value of β1 resulting

from regression analysis on ûro is 0.756.

A more general model to estimate corner deformation, which includes the effect of distance

from the CM, will be introduced in the next section. For consistency with this general model,

β1 is rounded to 0.88, which for small eccentricities (eb /r < 0.1) increases ûro by less than 1.2

percent. This leads to the following design equations for ûro :


 eb 0.88
ûro = 1 + (4.21)
r

and the deformation uro :


 eb 0.88
uro = ubo 1 + (4.22)
r
where the deformation ubo of a corresponding symmetric system is given by Eq. 3.25, or, its

simpler version, Eq. 3.26.

66
20
exact median
design spectrum

15

uro (cm)
η=0.25
10
0.5
1.0
5 1.5

0
0 1 2 3 4
Tb (sec)

Figure 4.15: Design equation for uro (Eq. 4.22) compared to exact median computed by
nonlinear RHA to LMSR ground motion ensemble with median PGV = 35 cm/s, for systems
with Ωθz = 1.25 and eb /r = 0.1

A comparison of Eq. 4.22 with the exact median – over the LMSR ensemble – demon-

strates that the design equation generally follows the increase in deformation with increasing

eccentricity (Fig. 4.14). In this and later figures, the median deformation, computed by the

procedure of Sec. 4.2.6, assumes u̇go of 35 cm/s, as does the deformation given by the compa-

rable design equation. While overall an excellent fit, some discrepancy between Eq. 4.22 and

the median occurs because the design equation for ubo does not match its median exactly. This

can be seen by comparing the fit of uro for fixed eccentricity, presented in Fig. 4.15, to the

comparable fit for ubo in Fig. 3.11b. This comparison suggests that the design equation for uro

(Eq. 4.22) fits the median of uro as well as the design equation for ubo (Eq. 3.25) fits its median.

Regression models that were rejected even though they provided better fits to the median of

uro are discussed in Appendix 4B.

4.5.2 Equations to Estimate Corner Deformation

All that remains is to find a simple way to apply the preceding concepts to estimate the largest

deformation among all isolators, assumed to occur at the outermost corner of the building.

Unlike uro , which depends little on plan geometry, the deformation of the outermost corner

depends on its location, specified as a distance c and direction from the CM. However, a practical

method that is independent of plan geometry should be feasible if a reasonable estimate for the

67
3

b/d=1
1
2
3
y/r

0 CM ∞

−1

−2

−3
−3 −2 −1 0 1 2 3
x/r

Figure 4.16: Rectangular plans of various aspect ratios with corners on a circle of radius c/r

deformation of the outermost corner can be attained from the deformation maximized over a

circle of radius c, as given by the following:

c
uco = max ( ub (t) + r|θbz (t)| ) (4.23)
t r

which has been adapted from Eq. 4.13 for uro .

This equation will give essentially the same deformation for all rectangular plans with

uniformly distributed mass, for two reasons. First, c/r = 3 regardless of plan aspect ratio
 
b/d (r = (b2 + d2 )/12 and c = (b2 + d2 )/2). Second, b/d has little influence on median

uro (Sec. 4.3.2), implying little influence on the CM deformations ub and rθbz . Equation 4.23

determines the largest deformation over a circle of radius c instead of the largest deformation

at any of the four corners (Fig. 4.16), and is clearly conservative.

The accuracy of Eq. 4.23 is evaluated for rectangular plans with different aspect ratios in

Fig. 4.17, which compares the median deformation computed by Eq. 4.23 to the exact median

of corner deformation, largest among the four corners. Because aspect ratio has little influence

on the CM deformations, the approximate (Eq. 4.23) and exact values of corner deformation

are both computed from the CM deformation ub and rotation rθbz of the idealized system with

68
20 11.5
(a) uro (b)
uco
b/d=1
15 η=0.25 b/d=2
11 b/d=3
b/d=∞
uco (cm)

uco (cm)
0.5
10 0.75
1.0
1.5 10.5
5

0 10
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15
Eccentricity eb /r Eccentricity eb /r

Figure 4.17: Median corner deformations uco for rectangular plan with various aspect ratios
compared to conservative estimate given by Eq. 4.23; uro included for reference, Tb = 2.5 sec
and Ωθz = 1.15: (a) several values of η, (b) η = 0.5

b/d = 2 (Fig. 4.5b) rather than a different model for each aspect ratio. The locations where the

exact corner deformations are sampled are meaningless for the idealized system but appropriate

for the systems they represent. A cluster of curves for each of several values of η appears in

Fig. 4.17a; in each cluster the exact corner deformations for different aspect ratios lie between uco

determined from Eq. 4.23 and uro . A single value of strength, η = 0.5, is the focus of Fig. 4.17b,

where median corner deformation is seen to be largest for plans with b/d = 2 (though nearly

equal for b/d = 1, 2 and 3) and smallest for plans with b/d = ∞. Equation 4.23 overestimates

the deformation by at most 2% when b/d = 2 and 5% when b/d = ∞. Therefore, deformation

maximized over the circle of radius c (Eq. 4.23) should estimate the corner deformation with

sufficient accuracy.

In an asymmetric plan, the relative distance c/r from the CM to the outermost corner

varies according to its geometry; the smallest possible c/r is 2 for a circular plan, and no

theoretical maximum exists. However, c/r generally falls within a narrow range; for example,

for Plans A1, A2, and A3 (Fig. 4.5a), c/r = 1.81, 2.05, and 1.83, respectively. To examine how

deformation varies with relative distance from the CM, uco was computed for different values of

c/r by Eq. 4.23 from ub (t) and rθbz (t) of the idealized system. The median – over all excitations

in the ensemble – of uco is plotted as a function of c/r (Fig. 4.18) for a system with Ωθz = 1.15,

69
(a) Tb = 2.0 sec (b) Tb = 4.0 sec
20
exact median
design equation
η=0.25
η=0.25
15
0.5
uco (cm)

0.5
0.75
10 0.75 1.0
1.0 1.5
1.5
5

0
1 1.5 2 2.5 1 1.5 2 2.5
Distance to Corner c/r Distance to Corner c/r

Figure 4.18: Design equation for uco (Eq. 4.25) compared to exact median computed by
nonlinear RHA to LMSR ground motion ensemble with median PGV = 35 cm/s, for systems
with Ωθz = 1.15 and eb /r = 0.15; (a) Tb = 2 sec, (b) Tb = 4 sec

eb /r = 0.15, and varying Tb and η; dot markers indicate where c/r was sampled.

With these results, the pieces are in place to develop an equation to estimate deformation

at the outermost corner. To summarize, the equation will be based on regression of uco for dif-

ferent c/r, which is the deformation maximized over a circle of radius c, shown to conservatively

estimate the deformation at the outermost corner as a function of its distance c from the CM.

To extend the present regression model (Eq. 4.21) to include corner distance, we hypothesized
c eb
that ln (ûco ) varies linearly with ln (1 + r r ). This particular model gave nearly the best fit

among regression models considered (see Appendix 4B for details). Linear regression on re-

sponse data for ln (ûco ), shown by dot markers in Fig. 4.18, resulted in a regression coefficient

of 0.880, giving the following design equation for ûco :


 c eb 0.88
ûco = 1 + (4.24)
r r

Multiplying this by the deformation ubo of a corresponding symmetric system (Eq. 3.25) gives

the design equation for corner deformation:


 c eb 0.88
uco = ubo 1 + (4.25a)
r r
6.59 0.19 (−0.55−0.08 ln η)  c eb 0.88
= T η 1 + u̇go (4.25b)
4π 2 b r r

70
Table 4.1: Corner deformations for asymmetric Plans A1-A3 computed by nonlinear RHA
and estimated by design equation (Eq. 4.25)

Plan Tb (s) η eb/r c/r (Crn) uco (Crn) Eq. 4.25 % Err
A1 2.0 0.46 0.069 1.81 (1,3) 10.15 (2) 10.80 6.4
A1 4.0 0.46 0.069 1.81 (1,3) 11.53 (2) 12.32 6.9
A1 2.0 0.92 0.069 1.81 (1,3) 7.24 (2) 7.73 6.8
A1 4.0 0.92 0.069 1.81 (1,3) 7.38 (3) 8.82 19.5
A2 2.0 0.46 0.108 2.05 (1,4) 10.91 (3) 11.61 6.4
A2 4.0 0.46 0.108 2.05 (1,4) 12.22 (3) 13.24 8.4
A2 2.0 0.92 0.108 2.05 (1,4) 7.95 (3) 8.32 4.7
A2 4.0 0.92 0.108 2.05 (1,4) 8.79 (3) 9.49 8.0
A3 2.0 0.46 0.057 1.83 (1) 10.23 (2) 10.62 3.8
A3 4.0 0.46 0.057 1.83 (1) 11.39 (2) 12.12 6.4
A3 2.0 0.92 0.057 1.83 (1) 7.09 (2) 7.61 7.3
A3 4.0 0.92 0.057 1.83 (1) 7.86 (1) 8.68 10.4

written out here in its entirety (Eq. 4.25b) for completeness. Alternatively, the simplified

equation for ubo (Eq. 3.26) could be used in Eq. 4.25a. Note that (by design) Eq. 4.25a reduces

to the equation for uro (Eq. 4.22) when c/r = 1, and to ubo when eb /r = 0. Figure 4.18

demonstrates that Eq. 4.25 (dashed line) is a good fit to the median of uco ; however, this fit is

not as good as Eq. 4.22 for uro , as discrepancies increase with distance from the CM.

4.5.3 Application to Various Plans

Since the design equation for corner deformation (Eq. 4.25) is developed from the response of

the idealized, rectangular-plan system, the applicability of this equation to various asymmetric-

plan systems is evaluated next. For two different isolation periods (Tb = 2 and 4 seconds) and

strengths (µ = 0.05 and 0.1), Plans A1-A3 (Fig. 4.5a), each with a different eccentricity eb /r

and outermost corner distance c/r, are subjected to the LMSR ground motion ensemble. With

assumed median PGV of 35 cm/s and ωd = 3.05 for the strong-component ensemble, the

normalized strengths (Eq. 4.16) are η = 0.46 and 0.92.

For each plan, the exact deformation at each corner (as numbered in Fig. 4.5a) is the

median over the deformations at that corner, computed from nonlinear RHA for each excitation.

Table 4.1 lists the largest exact corner deformation (among all corners), the value estimated

71
by Eq. 4.25, and the percent discrepancy (% Err) in the latter. A positive discrepancy signifies

that Eq. 4.25 gives a conservative estimate of the largest corner deformation. Also noted are

the outermost corner that enters in the estimate of Eq. 4.25 and the corner that actually sees

the largest deformation; the two are usually different.

Table 4.1 reveals that Eq. 4.25 tends to be conservative, which is not surprising since

the design equation is based on deformation maximized over the circle of radius c while the

exact medians are determined at the corner locations. Another source of conservatism is that

the deformation is not largest at the outermost corner, and the distance in Eq. 4.25 is over-

estimated. However, Eq. 4.25 is typically within 10% of the exact deformation, which is ideal

for an estimate. Successful development of a procedure that generally predicts the median

deformation within 10% was possible specifically because the eccentricities (representative of

isolation systems) and, consequently, the effects of torsion, are small.

4.5.4 Comparison to IBC Equation

An equation in the International Building Code (IBC) [1] increases the design displacement due

to torsion, according to
 
12e
DT D = DD 1+y 2 (4.26)
b + d2
where DT D and DD are the design displacements with and without torsion, respectively, y is

the distance from the CR (not the CM) to the outermost isolator in the direction perpendicular

to the applied load, and b and d are the longest and perpendicular to longest plan dimensions.

This equation increases DD by the rotational deformation resulting from a static torque equal

to DD kef f × eb applied about the CR, where kef f is the effective (equivalent-linear) stiffness at

the peak deformation of the isolation system [2].

Equation 4.26 implies a deformation ratio equal to DT D /DD . Because (b2 + d2 )/12 is

an estimate for r 2 (exact for a rectangular plan), the deformation ratio by this equation is

approximately (1 + yr re ), which is similar to Eq. 4.24. The two equations differ by the coefficient

0.88, and according to their single direction (y and e) or vector magnitude (c and eb ) variables.

Specifically, the code equation gives the deformation ratio resulting from loading in a principal

72
direction, while Eq. 4.24 gives the deformation ratio that results from bidirectional excitation.

However, the two deformation ratios can be compared directly, because an increase in the

deformation (estimated by Eq. 4.24) implies equivalent increases in the component deformations

in each direction (which the code estimates).

The deformation ratios given by the code equation (Eq. 4.26) and Eq. 4.24 are compared

for Plans A1-A3. These data imply a 7.7, 14.9 and 6.1 percent increase in isolator deformation

due to torsion by the code equation, compared to a 10.9, 19.2, and 8.7 percent increase by

Eq. 4.24, for Plans A1-A3 respectively. The torsional increase in deformation by the code

equation is about 30% smaller than that by Eq. 4.24, but this underestimation is less significant

than suggested because the contribution of torsion to deformation is relatively small.

However, the IBC gives an inadequate estimate of DD in a symmetric system (Sec. 3.9).

It was shown that, for unidirectional excitation, equivalent-linear methods used by the IBC un-

derestimate DD by up to 50% compared to nonlinear RHA; and IBC’s at most 4.4% increase in

DD to account for bidirectional excitation is smaller than the 13% increase determined by non-

linear RHA for bidirectional excitation. All factors combined, the IBC grossly underestimates

the total deformation of the asymmetric system for a given design spectrum.

4.6 Conclusions

This investigation to estimate the median value of the peak isolator deformation in an asymmetric-

plan, base-isolated building, subjected to an ensemble of two components of ground motion,

has led to the following conclusions:

1. The median – over the ground motion ensemble – of any normalized deformation of interest

depends primarily on the following global properties of the system, which are attained

by summing over individual isolators: the isolation period Tb , the normalized strength η,

the torsional to lateral frequency ratio Ωθz and the normalized stiffness eccentricity eb /r.

This was achieved by defining the normalized strength as the system strength ÷ ωd u̇gyo ,

where u̇gyo is the peak ground velocity of the strong component of motion and ωd is the

corner frequency separating the velocity and displacement-sensitive regions of the median

73
spectrum for the strong-component ensemble.

2. Because isolators tend to be distributed at regular intervals, and their stiffnesses and

strengths are independent of direction:

• uro (the deformation maximized over a circle of radius r from the CM) is only

weakly dependent on the system geometry and the planwise layout of isolators; thus,

the median of uro of an idealized, rectangular-plan system with global properties

identical to those of the actual asymmetric plan is within 2% of the median of uro

for the actual plan;

• the frequency ratio varies over a narrow range (1-1.3) and the eccentricities tend to

be small (< 0.1), giving focused ranges of these properties for base-isolated systems,

thus permitting sharper conclusions;

• a two-way asymmetric system can be viewed as a one-way asymmetric system by

an axis rotation that does not alter the other global properties, allowing results for

one-way asymmetric systems to be generalized to two-way asymmetric systems.

3. The median ratio ûro (of deformation uro of the asymmetric system to the deformation

ubo of a corresponding symmetric system) depends only weakly on isolation period Tb ,

strength η, and frequency ratio Ωθz , and increases with eccentricity eb /r. Thus, a simple

equation (Eq. 4.21) that depends only on eccentricity was fit to the data for ûro , which

can be combined with an equation to estimate the deformation of the symmetric system

(Eq. 3.25 or Eq. 3.26) to obtain Eq. 4.22 for the deformation uro of the asymmetric system.

4. To extend these concepts to estimate the peak deformation in systems with different plan

geometry, which occurs at one of the corners, it was shown that for rectangular plans, the

deformation maximized over a circle of radius c from the CM provides an excellent, only

slightly conservative estimate of the peak deformation at a corner located at distance c

from the CM. Based on this result, the equation to estimate ûro was generalized, resulting

in Eq. 4.25, which estimates the peak corner deformation as a function of its distance from

74
the CM. For three asymmetric plans, this design equation was always conservative, but

generally within 10% of the exact median value determined by nonlinear response history

analysis for each ground motion in the ensemble.

5. The percent increase in deformation due to torsion determined by the IBC equation is

about 30% smaller than the increase found in this investigation by nonlinear response his-

tory analysis; this underestimation is relatively unimportant given that the contribution

of torsion to deformation is small. Of greater significance, the deformation of a symmetric

system given by the IBC can be unconservative by up to 50% compared to the deforma-

tion determined by nonlinear response history analysis; and this deformation increases

by at most 4.4% to account for bidirectional excitation, smaller than the 13% increase in

deformation determined by nonlinear response history analysis, together implying that,

for a given design spectrum, the IBC grossly underestimates the peak isolator deformation

of an asymmetric system.

75
Appendix 4A: Maximization of ur (t)

From Eq. 4.12, the magnitude of deformation at location rx , ry  is

 1/2
ur (t) = urx (t)2 + ury (t)2
 1/2
= (ubx − ry θbz )2 + (uby + rx θbz )2
 1/2
= u2bx + u2by + 2rx θbz uby − 2ry θbz ubx + r 2 θbz
2
(4A.1)

By the method of Lagrange multipliers, ur is maximized subject to the constraint of Eq. 4.10.

To simplify the math, we maximized the function u2r . In this method, the gradients of both

u2r and r 2 (Eq. 4.10) with respect to coordinates rx and ry are computed and substituted into

∇ur = λ∇r 2 , resulting in two equations:

2
2rx θbz + 2uby θbz = λ2rx
2
2ry θbz − 2ubx θbz = λ2ry

which can be rearranged to obtain rx and ry as functions of λ:

uby θbz ubx θbz


rx = and ry = −
(λ − θbz
2 ) (λ − θbz
2 )

Substituting these equations into the constraint equation (Eq. 4.10) and solving for λ results

in:
θbz 2
2
λ = θbz ± (u + u2by )1/2
r bx
and
uby ubx
rx = ± r and ry = ∓ r (4A.2)
ub ub
where ub , the magnitude of deformation at the CM has been substituted for (u2bx +u2by )1/2 . These

solutions for rx and ry imply that the location of the maximum on the radius r is determined

by the relative values of ubx and uby . One pair of solutions corresponds to a minimum while

the other corresponds to a maximum. Substituting Eq. 4A.2 into Eq. 4A.1 results in:

 
2 1/2
ur = u2b ± 2ub rθbz + r 2 θbz

76
Obviously the maximum results from taking ± to match the sign of θbz , and this equation

simplifies to

ur = ub + r|θbz | (4A.3)

This simple result is a generalization of the response to unidirectional (y-direction) excitation

(Eq. 4.9), for which, recall, deformations were evaluated at x = ±r, locations perpendicular to

the direction of deformation. In the more general case, the x-axis is replaced by a plane through

the CM perpendicular to the direction of instantaneous deformation ub (t), which intersects the

circle of radius r at two points, one being a maximum (ub + r|θbz |) and the other being a

minimum (ub − r|θbz |). Clearly, Eq. 4A.3 reduces to Eq. 4.9 for unidirectional excitation if the

instantaneous deformation is in the y-direction.

Appendix 4B: Regression Analysis

This appendix discusses regression models that were considered in developing design equations

for ûro and ûco , building upon concepts and terminology introduced in Appendix 3A.

4B.1 Considerations for Choosing Model

Because the response data is sampled from a lognormal distribution, a function that is linear

in the logarithmically transformed data – i.e., ln ûro is a linear function of regressors ln Tb , ln η,

etc. – gives a best estimate of the median response (Eq. 2.1). Therefore, regression models for

ûro and ûco were restricted to be of this form, and relaxing the requirement did not lead to any

improvements.

When linearity in the logarithmic-transformed domain is satisfied, the regression equation

for ûro , equal to ūro ÷ūbo , is identical to the ratio of regression equations for ūro and ūbo (assumes

regression models for ūbo , ūro and ûro each contain the same regressors). This suggests that

regression analysis of the deformation ratio ûro , which proves to be simpler and more convenient,

gives identical results to distinct regression analysis of the normalized deformations ūro and ūbo ,

for which dispersion is minimized. The corresponding argument can also be made for the peak

deformation ratio at the outlying corner ûco .

A final consideration, both deformation ratios ûro and ûco should reduce to 1, i.e, the

77
peak deformation reduces to that of a symmetric system as the eccentricity approaches 0.

4B.2 Selecting a Model for ûro

The model for ûro (Eq. 4.20) can be simplified to one that depends only on eccentricity eb /r

by eliminating other regressors. Although the normalized equation of motion (Eq. 3.9) im-

plies that the median value of uro depends on the global parameters Tb , η, Ωθz and eb /r, the

theoretical limit will not be satisfied if the model for ln ûro contains regressors ln Tb , ln η, or

ln Ωθz . Figures 4.10-4.12 demonstrated that the median of ûro is only weakly dependent on

each parameter Tb , η and Ωθz , but it can also be shown conclusively that the corresponding

regressors are statistically insignificant in estimating ûro .

Direct regression analysis of ln ûro implied that ln Tb and ln η were significant. However, to

show that these regressors can be eliminated, we instead compared the coefficients for ln Tb and

ln η determined from separate regression of ln ūro and ln ūbo . This regression analysis utilized a

data set where Tb and η are varied (16 values of Tb by 5 values of η), Ωθz = 1.15 and eb /r = 0.1

(eb /r = 0 for ubo ), with the following model:

ln ūro = β0 + β1 ln Tb + β2 ln η (4B.1)

where results are given in Table 4B.1. Recall that from a statistical perspective, the coefficients

β determined from regression analysis are just estimates, but their true values are likely to fall

within a 95% confidence interval (Appendix 3A). If the coefficients β1 and β2 (corresponding

to ln Tb and ln η) for ln ūro fall within the confidence intervals of the same coefficients for ln ūbo ,

or vice versa, their difference can be considered statistically insignificant. The 95% confidence

interval is about 2 × the “std error” on either side of the estimated coefficient. Since β1 for

ūro (Table 4B.1a) differs from β1 for ūbo (Table 4B.1b) by slightly more than the “std error”,

and β2 are nearly identical for ūro and ūbo , the corresponding coefficients are within confidence

intervals.

Note that the coefficients for ln Tb and ln η in Table 4B.1 differ from those in Appendix 3A

(Table 3A.1) because a smaller data set was used for analysis of asymmetric-plan systems to

limit computation time, and the regressor (ln η) was dropped to make the argument cleaner. The

78
Table 4B.1: Significance of Tb and η (excerpts of R Project output)

(a) Regression output for ūro


ln ūbo = β0 + β1 ln Tb + β2 ln η
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.84421 0.02218 38.06 <2e-16 ***
lnTb -1.82376 0.02286 -79.79 <2e-16 ***
lneta -1.48650 0.01538 -96.66 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.3772 on 1597 degrees of freedom


Multiple R-Squared: 0.9077, Adjusted R-squared: 0.9076
F-statistic: 7855 on 2 and 1597 DF, p-value: < 2.2e-16
--------------------------------------------------------------

(b) Regression output for ūbo


ln ūbo = β0 + β1 ln Tb + β2 ln η
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.74594 0.02176 34.27 <2e-16 ***
lnTb -1.79568 0.02242 -80.08 <2e-16 ***
lneta -1.48965 0.01509 -98.73 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.37 on 1597 degrees of freedom


Multiple R-Squared: 0.9101, Adjusted R-squared: 0.91
F-statistic: 8080 on 2 and 1597 DF, p-value: < 2.2e-16

regression equation for ūbo in Chapter 3 (Eq. 3.14a) is more reliable because it was developed

with a much larger data set; hence the advantage of analyzing ûro and combining with the

accuracy of Eq. 3.14a, rather than directly analyzing ūro .

The regressor ln Ωθz was eliminated by a similar exercise. Since the deformation ubo of the

symmetric system is independent of Ωθz , it is only necessary to test whether the deformation

uro of the asymmetric-plan system depends on Ωθz , via the following model:

ln ūro = β0 + β1 ln Tb + β2 ln η + β3 ln Ωθz (4B.2)

The response data included Tb varied over 3 values, η varied over 4 values, and Ωθz varied over 8

values within its practical range (Fig. 4.12). For this analysis, the t-value for β3 corresponding

to ln Ωθz is small enough (Table 4B.2) that the probability is high that β3 is zero; thus ln Ωθz

79
Table 4B.2: Significance of Ωθz (excerpts of R Project output)

ln ūro = β0 + β1 ln Tb + β2 ln η + β3 ln Ωθz
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Intercept) 0.92507 0.0291407 31.75 <2e-16 ***
lnTb -1.88116 0.0290337 -64.79 <2e-16 ***
lnEta -1.52175 0.0148070 -102.77 <2e-16 ***
lnOmega -0.00084 0.0885179 -0.01 0.992
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.4432 on 1916 degrees of freedom


Multiple R-Squared: 0.8851, Adjusted R-squared: 0.8849
F-statistic: 4920 on 3 and 1916 DF, p-value: < 2.2e-16

should not be included in the regression model.

Once these three regressors had been eliminated, possible regression models for ûro were

limited. The model of Eq. 4.20 included the remaining parameter eb /r and satisfied the lim-

iting behavior (ûro → 1 as eb /r → 0). Regression analysis of this model (Eq. 4.20) with a

comprehensive data set (including variation of Tb over 5 values, η over 5 values, and eb /r over

10 values equally spaced from 0 to 0.15) led to the regression coefficient of 0.756 (Table 4B.3a).

The R2 statistic, which is about 0.6, does not seem high compared to other analyses where it

was above 0.9. However, due to the scatter in ûro as a function of eb /r (Fig. 4.13b), a model

that predicts 60% of this variation must be considered adequate.

To demonstrate that it is usually possible but not necessarily desirable to achieve incre-

mental improvement in the regression equation, one last regression model for ûro is considered.

In this model, Tb and η are included in such a way that the theoretical limits are still satisfied:

ln ûro = (β1 + β2 Tb + β3 η) ln (1 + eb /r) (4B.3)

Resulting analysis with this model (Table 4B.3b) compared to the model of Eq. 4.20 (Ta-

ble 4B.3a) suggests that one of the additional coefficients (β2 ) is significant, and demonstrates

a small improvement in the R2 statistic (0.6072 versus 0.5941). Ultimately, we must make a

judgment call; are the improvements large enough to justify the increased complexity of the

equation? Since the actual coefficients β2 and β3 are small, and it has already been demon-

80
Table 4B.3: Variation of ûro with eccentricity (excerpts of R Project output)

(a) Regression output for ûro


ln ûro = β1 ln (1 + eb /r)
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
ln(1+eb/r) 0.75903 0.00887 85.55 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.05522 on 4999 degrees of freedom


Multiple R-Squared: 0.5942, Adjusted R-squared: 0.5941
F-statistic: 7319 on 1 and 4999 DF, p-value: < 2.2e-16
--------------------------------------------------------------

(b) Regression output for ûro


ln ûro = (β1 + β2 Tb + β3 η) ln (1 + eb /r)
Coefficients:
Estimate Std.Error t-value Pr(>|t|)
ln(1+eb/r) 0.95207 0.02762 34.47 <2e-16 ***
(Tb)ln(1+eb/r) -0.10224 0.00823 -12.43 <2e-16 ***
(Eta)ln(1+eb/r) 0.07818 0.02029 3.85 0.0001 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.05432 on 4997 degrees of freedom


Multiple R-Squared: 0.6074, Adjusted R-squared: 0.6072
F-statistic: 2578 on 3 and 4997 DF, p-value: < 2.2e-16

strated that ûro does not vary systematically with Tb and η (Figs. 4.10b and 4.11b), including

the extra coefficients does not seem worthwhile here.

4B.3 Selecting a Model for ûco

Whereas the additional deformation due to rotation (rθbz ) in uro (Eq. 4.13) was shown to be

accounted for by eb /r, we anticipate that the additional deformation due to rotation (cθbz )

in uco (Eq. 4.23) is accounted for by eb /r multiplied by c/r. In this sense, Eq. 4.24 is the

most logical selection of regression model for ûco . The resulting regression analysis with this

model – labeled Model 1 – (Table 4B.4a), is compared to three other models – Models 2-4

– that were also considered (Table 4B.4b-d). Model 2 (Table 4B.4b), which also contained

only one coefficient and thus arguably no more complex, showed only slight and presumably

inconsequential improvement in its R2 statistic. Analysis suggested that Model 4 (Table 4B.4d),

81
Table 4B.4: Testing models for ûco (excerpts of R Project output)

(a) Model 1: ln ûco = β1 ln (1 + c/r ∗ eb /r)


Coefficients:
Estimate Std.Error t-value Pr(>|t|)
ln(1+c/r*eb/r) 0.88028 0.00230 382.2 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.08164 on 59999 degrees of freedom


Multiple R-Squared: 0.7089, Adjusted R-squared: 0.7089
F-statistic: 1.461e+005 on 1 and 59999 DF, p-value: < 2.2e-16
--------------------------------------------------------------

(b) Model 2: ln ûco = β1 (c/r) ln (1 + eb /r)


Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(c/r)ln(1+eb/r) 0.84608 0.00221 382.5 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.08159 on 59999 degrees of freedom


Multiple R-Squared: 0.7092, Adjusted R-squared: 0.7092
F-statistic: 1.463e+005 on 1 and 59999 DF, p-value: < 2.2e-16
--------------------------------------------------------------

(c) Model 3: ln ûco = β1 ln (1 + eb /r) + β2 ln (c/r)


Coefficients:
Estimate Std.Error t-value Pr(>|t|)
ln(1+eb/r) 1.11735 0.00682 163.9 <2e-16 ***
ln(c/r) 0.05897 0.00111 53.3 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.08386 on 59998 degrees of freedom


Multiple R-Squared: 0.6928, Adjusted R-squared: 0.6928
F-statistic: 6.765e+004 on 2 and 59998 DF, p-value: < 2.2e-16
--------------------------------------------------------------

(d) Model 4: ln ûco = (β1 Tb + β2 η + β3 c/r) ln (1 + eb /r)


Coefficients:
Estimate Std.Error t-value Pr(>|t|)
(Tb)ln(1+eb/r) -0.18063 0.00320 -56.39 <2e-16 ***
(Eta)ln(1+eb/r) 0.25073 0.00815 30.78 <2e-16 ***
(c/r)ln(1+eb/r) 0.99028 0.00595 166.51 <2e-16 ***
---
Signif. codes: 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1

Residual standard error: 0.07916 on 59997 degrees of freedom


Multiple R-Squared: 0.7263, Adjusted R-squared: 0.7262
F-statistic: 5.306e+004 on 3 and 59997 DF, p-value: < 2.2e-16

82
similar to the second model for ûro (Table 4B.3b), might be worth incorporating based on

improvement to its R2 statistic relative to Model 1 and significance of additional regressors

(functions of Tb and η). Once again, we rejected these arguably improved models, preferring

the simplicity of and theoretical justification for Model 1.

Appendix 4C: Notation

a∗y yield acceleration of isolated block: Q/m

b maximum dimension or length of symmetric-plan building

b/d length-to-width ratio, or plan aspect ratio

c distance from CM to outlying corner

d minimum dimension or width of symmetric-plan building

DD code-specified design displacement of isolation system

DT D code-specified design displacement including torsion

ebx , eby stiffness eccentricities in x and y-directions

ebx /r, eby /r dimensionless stiffness eccentricities, normalized by r

eb /r total magnitude of stiffness eccentricity: [(ebx /r)2 + (eby /r 2 )]1/2

epx , epy strength eccentricity in x and y-directions

fbxi , fbyi lateral force of ith isolator in x and y-directions

kbi lateral postyield stiffness of ith isolator

kb lateral postyield stiffness of isolation system (summed over all isola-

tors)

kbθz torsional postyield stiffness of isolation system about CM

kef f effective stiffness of isolation system for equivalent-linear analysis

m mass of isolated block

Qi characteristic strength of ith isolator

83
Q lateral characteristic strength of the isolation system (summed over

all isolators)

Qθz torsional strength of isolation system about CM

r mass radius of gyration about vertical axis

Tb isolation period, corresponds to postyield stiffness of isolation system

ubxi , ubyi lateral deformations of ith isolator in x and y-directions

ubx , uby lateral deformations at CM of asymmetric system in x and y-

directions

ubo peak deformation of corresponding symmetric system

ūbo peak normalized deformation of corresponding symmetric system:

ubo /u∗y

uco peak deformation in any direction maximized at a circle of radius

c from CM; an upper bound to the peak deformation at outlying

corner

ûco peak deformation ratio at circle of radius c: uco /ubo

ügx , ügy x and y-components of ground acceleration

ū ¨gy
¨gx , ū ground acceleration normalized by peak ground velocity

u̇gyo peak ground velocity in y-direction component

u+ry , u−ry peak deformation in the y-direction at ±r from CM, where + is the

stiff side and − is the flexible side for positive eccentricity ebx /r

uro peak deformation in any direction maximized at a circle of radius r

from CM

ûro peak deformation ratio at circle of radius r: uro /ubo

u∗y Q/kb

zxi , zyi dimensionless plastic variables governing yielding of the ith isolator

in the x and y-directions

γi Qi /Q

η strength of isolator or isolation system normalized by peak ground

velocity: a∗y /(ωd u̇go )


84
θbz rotation of asymmetric-plan system about CM

θbzo peak rotation of asymmetric-plan system

µ characteristic strength ratio of isolated block: Q/w



ωb lateral isolation frequency: kb /m

ωbθz torsional isolation frequency: kbθz /mr 2

Ωθz ratio of torsional to lateral isolation frequency: ωbθz /ωb

ωd corner frequency separating velocity and displacement-sensitive re-

gions of linear spectrum

85
5: ESTIMATING THE PEAK DISPLACEMENT OF A FRICTION
PENDULUM ISOLATOR

5.1 Introduction

Recall the procedure developed in Chapter 3, by which isolator deformations due to ground

motions characterized by a design spectrum with its intensity defined by the median peak

ground velocity (PGV) can be estimated. Based on rigorous nonlinear analysis of the isolated

system for an ensemble of ground motions representative of the spectrum, this procedure is

effective because the governing equations for the system were rewritten such that the normalized

deformation was insensitive to ground motion intensity.

Initially developed for a system isolated with lead-rubber bearings, the nonlinear pro-

cedure is general enough to apply to other types of isolation systems. Specifically, friction

pendulum (FP) isolators can also be modeled by a nonlinear force-displacement relation, re-

quiring only a different value for the yield displacement [2]. The yield displacement for FP

isolators is on the order of 0.05 cm [42], compared to 1 cm for lead-rubber bearings (Sec. 3.3.1).

The significant change in the yield displacement suggests that different design equations are

needed for systems with FP isolators.

The objective here is simply to extend earlier techniques to develop design equations to

estimate the peak displacement of a system with FP isolators. Thereafter, an equivalent-linear

procedure to estimate this displacement, upon which the current IBC is based, is evaluated

against the design equations based on rigorous nonlinear analysis.

5.2 System, Governing Equations and Normalization

5.2.1 System Considered

The system analyzed is a rigid mass m supported by a single FP isolator, whereby the weight w

of the system rests on an articulated slider, which slides over a spherical surface in response to

earthquake excitation (Fig. 5.1a). Neglecting the slight rise of the mass on the spherical surface

as done commonly [34], the resisting force in the system is


w
fb = ub + µw sgn(u̇b ) (5.1a)
R

86
fb
(a) (b)
µw w/R
m
ub (t) ub
R, µ

üg (t)

Figure 5.1: (a) Single FP isolator supporting a rigid mass, and (b) rigid plastic force-
displacement of the isolator.

for unidirectional excitation (Fig. 5.1b), and


⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎬ w⎪
⎨fbx ⎪ ⎨ubx ⎪
⎬ ⎪ ⎪
1 ⎨u̇bx ⎬
= + µw (5.1b)
⎭ R⎪
⎩fby ⎪
⎪ ⎩uby ⎪
⎭ u̇b  ⎪
⎩u̇by ⎪

for bidirectional excitation, where ub (ubx , uby ) is the slip in the isolator (x and y-components

of slip for bidirectional excitation), or displacement relative to the ground, and u̇b (u̇bx , u̇by ) the

relative velocity. The resisting force in Eq. 5.1 is composed of a pendular component directed

toward the centered position and a friction component that opposes instantaneous velocity;

thus, the isolator is characterized completely by the radius of curvature R of the spherical

surface and the friction coefficient µ of the sliding interface. The friction coefficient is known to

be sensitive to the sliding velocity, the pressure of the supporting mass, and viscous heating of

the slider over extended excitation [34, 43, 44, 45]; however, these effects are subtle enough that

the time variation of the friction coefficient need not be included in this otherwise simple model,

and varying axial load is irrelevant in a single isolator system. The simple rigid-plastic force-

displacement relation shown in Fig. 5.1b will be modified for the x and y-components when the

excitation is bidirectional, due to interaction of the friction force in the x and y-directions; that

is, the friction force in each direction depends on its velocity component (Eq. 5.1b).

Rigid-plastic force-displacement is difficult to implement in a numerical evaluation of

response. However, the general plasticity model is convenient to implement, and can treat

the small deformation which occurs at the sliding surface prior to slip as a nonzero yield

displacement. This pre-slip deformation is on the order of 0.05 cm [42]. Thus, the force in the

87
FP isolator can be described by Eq. 3.19 with the postyield stiffness kb equal to the pendular

stiffness w/R, the strength Q equal to the friction force µw, and the yield displacement – which

controls the yield functions zx and zy – fixed at 0.05 cm.

With isolator forces represented by the plasticity model, the equations governing the mass

on a single FP isolator (Fig. 5.1a), henceforth referred to as an FP system, subjected to ground

acceleration components ügx and ügy are


⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎨übx ⎪
⎪ ⎬ ⎪
⎨ubx ⎪
⎬ ⎪
⎨zx ⎪
⎬ ⎪
⎨ügx ⎪

+ ωb2
+ µg =− (5.2)
⎩üby ⎪
⎪ ⎭ ⎪
⎩uby ⎪
⎭ ⎪
⎩zy ⎪
⎭ ⎪
⎩ügy ⎪

Unidirectional excitation, which will be discussed throughout alongside bidirectional excitation,

is easily treated by Eq. 5.2 with one of the components identically equal to zero. Equation 5.2

appears identical to the equations for a mass on a single lead-rubber bearing (LRB system)

(Eq. 3.20); the two differ only because they are modeled by different yield displacement or

deformation. The FP system is characterized by its isolation frequency ωb = g/R and friction

coefficient µ (Eq. 5.2). Solving Eq. 5.2 for a given excitation gives the peak displacement – in

any direction:

ubo = max (ubx (t)2 + uby (t)2 ) (5.3)
t

5.2.2 Normalized Equations of Motion

For meaningful statistical analysis of the response to an ensemble of ground motions, the gov-

erning equations are normalized as follows:


⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ¨bx ⎪
⎨ū ⎬ ⎪
⎨ūbx ⎪
⎬ ⎪
⎨zx ⎪
⎬ ⎪ ¨gx ⎪
⎨ū ⎬
ω2
2
+ ωb 2
+ ωb =− b (5.4)
⎪ ¨by ⎪
⎩ū ⎭ ⎪
⎩ūby ⎪
⎭ ⎪
⎩zy ⎪
⎭ ηωd ⎪ ¨gy ⎪
⎩ū ⎭

Identical to Eq. 3.22, this has been formulated by first dividing Eq. 5.2 by u∗y , introducing

ūbx = ubx /u∗y and ūby = uby /u∗y , the normalized displacements in the x and y-directions; and

then substituting the normalized strength into the right-hand side. The following previous

definitions apply: (1) a∗y (equal to the yield acceleration µg), (2) u∗y (Eq. 3.6):

Q a∗y
u∗y = = 2 (5.5)
kb ωb

88
and (3) normalized strength (Eq. 3.23):

a∗y µg
η= = (5.6)
ωd u̇gyo ωd u̇gyo

where ωd corresponds to the transition period Td from the velocity-sensitive to displacement-

sensitive region of the median spectrum (Fig. 2.1) and u̇gyo is the PGV, both referring to

the strong component of motion, which is applied in the y-direction. Hence, the components of

ground acceleration in Eq. 5.4 have been normalized by the PGV u̇gyo , such that the magnitude
¨gx (=ügx /u̇gyo ) is strictly less than 1 and ū
of ū ¨gy (=ügy /u̇gyo ) varies from -1 to +1.

By virtue of Eq. 5.4, the essential benefits of normalization that were found in Chapter 3
¨gx and ū
for LRB systems also hold for FP systems. Specifically, the excitations ū ¨gy have been

normalized to a common intensity, such that their variability is limited to the variability inherent

in different realizations of a random process. Because of this normalization, the median of the

normalized displacement ūbo depends mainly on Tb and η, and its dispersion over many ground

motions is expected to be small.

5.2.3 Dispersion of Normalized Displacement

To verify that the normalization is as effective for FP systems as it was for LRB systems

(Section 3.3.2), the dispersion of ūbo (Eq. 2.2) for FP systems is presented in Fig. 5.2 for

different values of Tb and η. The excitations are limited to unidirectional and are sampled

from the strong-component ensemble, a subset of the LMSR ensemble (Sec. 2.1). While the

dispersion remains small – between 0.4 and 0.6 – for the smaller values of η, which is consistent

with LRB systems (Fig. 3.5d), it escalates rapidly as η increases for reasons identified next.

Because an FP system is initially very stiff, the earthquake response of the system is

controlled by the peak ground acceleration if the friction force is large enough to prevent

slip. Thus, the PGV normalization, shown to be effective for LRB systems (Sec. 3.3.2), is

inappropriate if the friction force is so large that the isolator may or may not be activated by

different ground motions in the ensemble, depending on their peak ground acceleration; this

results in the sharp rise in dispersion at η = 1.5. However, an FP system is ineffective if the

isolator fails to slip, and such large normalized strength is irrelevant to practical application.

89
1.4 η=.25
0.5
1.2 0.75
1.0

Dispersion of ūbo
1 1.5

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5
Tb (sec)

Figure 5.2: Dispersion δ of ūbo to unidirectional excitation by the strong-component ensemble

Thus, in this study, results for FP systems will be restricted to values of η no larger than 1.

5.3 Peak Earthquake Response

5.3.1 Median Response Spectra

When the preceding normalization is used to generate predictions of the median normalized

response and ultimately the median response for a specified intensity of ground motion, its full

benefit is realized. Median response spectra for the normalized displacement ūbo , and the actual

displacement ubo as a function of PGV, are constructed for the FP system as follows:

1. Determine the peak value of the normalized displacement ūbo of the system by nonlinear

response history analysis (RHA) of Eq. 5.4 for a selected value of Tb and η, and a selected

ground motion; repeat for each ground motion in the ensemble.

2. Compute the median normalized displacement (Eq. 2.1) over all ground motions for the

selected value of Tb and η.

3. Convert the normalized displacement to the actual displacement by the following equation:

ubo = ūbo u∗y (5.7)

where u∗y (Eq. 5.5) is a function of η, ωd and PGV (Eq. 5.6); this gives the displacement

for a specified intensity of excitation.

90
(a) Unidirectional Excitation
2
10 15

ubo (cm) (unidirectional)


exact median
ūbo (unidirectional)

1 design equation
10
10
0 η=0.25
10
η=0.25
0.5 0.5
−1
5
10 0.75 0.75
1.0 1.0
−2
10 0
0 1 0 1 2 3 4 5
10 10
Tb (sec) Tb (sec)
(b) Bidirectional Excitation
2
10 15

ubo (cm) (bidirectional)


exact median
design equation
ūbo (bidirectional)

1
10
10 η=0.25
0
10 η=0.25 0.5
0.5 5 0.75
−1
10 0.75 1.0
1.0
−2
10 0
0 1 0 1 2 3 4 5
10 10
Tb (sec) Tb (sec)

Figure 5.3: Design equations for normalized displacement ūbo and displacement ubo for (a)
unidirectional excitation (Eqs. 5.8b and 5.9) and (b) bidirectional excitation (Eqs. 5.10 and
5.11), compared to their exact medians by RHA due to the LMSR ensemble with PGV of 35
cm/s.

4. Repeat steps 1-3 over the desired range of Tb and η.

Determined for the LMSR ensemble (Sec. 2.1), the median normalized displacement ūbo

(step 2 above) and median displacement ubo for a PGV of 35 cm/s (step 3) are plotted (solid

lines) in Fig. 5.3, as a function of Tb and for several values of η. Included are plots for both unidi-
¨gx = 0 in Eq. 5.4, (Fig. 5.3a) and bidirectional excitation (Fig. 5.3b).
rectional excitation, i.e., ū

5.3.2 Design Equations

As already achieved for LRB systems, design equations representative of the median displace-

ments over the range of Tb and η can be determined. Considering unidirectional excitation first,

the response data is presumed to fit the same form of equation used earlier (Eq. 3.14a), which

91
results in the following linear regression equation for the normalized displacement ūbo :

ln ūbo = 0.65 − 1.85 ln Tb − 2.23 ln η − 0.31(ln η)2 (5.8a)

or

ūbo = 1.04 Tb−1.85 η (−2.23−0.31 ln η) (5.8b)

where the coefficients were estimated by the method of least squares. An equation for the

displacement of the isolator is then obtained by substituting Eq. 5.8b into Eq. 5.7 where u∗y

is given by Eqs. 5.5 and 5.6 with ωd = 3.05, which corresponds to Td = 2.06 seconds for the

strong-component ensemble (Fig. 2.1a):

3.17 0.15 (−1.23−0.31 ln η)


ubo = T η u̇gyo (5.9)
4π 2 b

Plotted as dashed lines alongside the exact medians in Fig. 5.3a, these equations (Eqs. 5.8b and

5.9) are shown to be good fits to their corresponding medians, thus suitable for analysis and

design applications. The coefficient -0.31 in the quadratic-ln (η) term is much larger than the

corresponding coefficient -0.08 in Eqs. 3.14b and 3.16 for LRB systems. Therefore, simplification

of Eq. 5.8b by neglecting this term (parallel to Eq. 3.17) is inappropriate.

By similar methods, the following equations were fit to the normalized displacement ūbo

and displacement ubo due to bidirectional excitation:

ūbo = 1.43 Tb−1.86 η (−1.99−0.20 ln η) (5.10)

and
4.36 0.14 (−0.99−0.20 ln η)
ubo = T η u̇gyo (5.11)
4π 2 b
These design equations for bidirectional excitation (Eqs. 5.10 and 5.11) are shown also to be

good fits to their corresponding exact medians (Fig. 5.3b).

Evident from the design equations and as expected, bidirectional excitation causes larger

displacement of the isolator compared to unidirectional excitation. This increase depends signif-

icantly on η, varying from approximately 20% to 38% as η increases from 0.25 to 1.0 (Fig. 5.4).

However, the displacement increase remains essentially independent of Tb , as indicated by the

closeness of the coefficients on Tb .

92
1.5

ubo(bidirectional) /ubo(unidirectional)
1.4

1.3

1.2
η=0.25
1.1 0.5
0.75
1.0
1
0 1 2 3 4 5
Tb (sec)

Figure 5.4: Ratio of design equations for ubo due to bidirectional and unidirectional excitation
(Eq. 5.10 ÷ Eq. 5.8b)

Although the response to bidirectional excitation is of primary interest, developing design

equations for both unidirectional and bidirectional excitation is insightful for comparing FP

systems and LRB systems. The displacement of an FP system is generally much smaller than

the deformation of an LRB system for comparable values of Tb and η (Fig. 5.5). This trend is not

surprising because an FP system, due to its reduced yield displacement, dissipates more energy

than an LRB system at comparable displacements/deformations of the isolator [2, Fig. 5.5], thus

reducing its displacement. The displacements of the FP system shown in Fig. 5.5a range from

50 to 90% as large as the deformations of a comparable LRB system, depending on Tb and η;

clearly the relative difference becomes larger as η increases. However, this discrepancy decreases

for bidirectional excitation (Fig. 5.5b), because the FP-system displacements increase more (22

to 38%) than the LRB-system deformations (constant 13%, see Sec. 3.7) from unidirectional to

bidirectional excitations.

5.4 Evaluation of Equivalent Linear Procedure

The International Building Code [1] and its predecessor, the Uniform Building Code, estimate

the design displacement of an isolation system from a linear design spectrum, with the properties

of the system determined iteratively by equivalent-linear analysis. The accuracy of using a

linear response spectrum to estimate the displacement of an FP system, similar to current code

procedure, is evaluated next. Given a response spectrum for unidirectional excitation, steps to

93
15 15
(a) FP systems (b)
LRB systems
η=0.25
ubo (cm) (unidirectional)

ubo (cm) (bidirectional)


η=0.25
10 10
0.5
0.5
0.75
5 0.75 5 1.0
1.0

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 5.5: Design equations for isolator displacement/deformation compared for FP systems
and LRB systems: (a) unidirectional excitation (Eq. 5.9 vs. Eq. 3.16) and (b) bidirectional
excitation (Eq. 5.11 vs. Eq. 3.25)

iteratively determine the equivalent-linear properties for a given nonlinear isolation system are

as follows:

1. Using initial guesses for the effective period Tef f and effective damping ζef f of the

equivalent-linear system, estimate displacement demand ubo from the spectrum.

2. Evaluate the isolator force as fbo = µw + (w/R)ubo , a specialization of Eq. 5.1 for unidi-

rectional excitation.


3. Update estimates of effective period: Tef f = 2π w/(kef f g), and effective damping:

ζef f = (2/π)µw(ubo − uy )/(kef f u2bo ), where kef f = fbo /ubo .

4. Repeat steps 1-3 with updated values of Tef f and ζef f , until successive estimates of ubo

converge.

The displacement of the FP system is estimated by implementing the above steps for

the median spectrum of the strong-component ensemble (Fig. 2.1a). Spectral displacements for

damping ratios other than 5%, needed in the iterative procedure, are determined by additional

response history analyses. The procedure is repeated for a range of Tb and η, whereby for each

η, the friction force is determined from Eq. 5.6 with the PGV taken to be 35 cm/s. The isolator

94
10 50
(a) Nonlinear (b) η=0.25
Equiv-Linear 0.5
8 0.75
25 1.0

Discrepancy in ubo
6
ubo (cm)

0
4

−25
2

0 −50
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 5.6: (a) Comparison of median displacement determined by nonlinear and equivalent-
linear methods for a system with η = 0.5, and (b) discrepancy in displacement computed by
equivalent-linear method relative to nonlinear method, for several values of η.

displacement estimated by this equivalent-linear procedure is plotted in Fig. 5.6a for a system

with η = 0.5, alongside the exact median displacement determined from nonlinear RHA. In this

case, the equivalent-linear estimate is close to the exact value for short isolation periods, but

the two diverge as the period increases.

To represent a more diverse range of FP systems, the percent discrepancy in the equivalent-

linear estimate relative to the exact value is presented in Fig. 5.6b for several values of η. Based

on this figure, the equivalent-linear procedure overestimates the displacement (positive percent

discrepancy) for very short isolation periods at the low end of the practical range, but under-

estimates it for longer periods representative of practical systems, by up to 30%. While similar

for the three smaller values of normalized strength, the percent discrepancy for η = 1.0 displays

a very different trend, because the FP system tends to be affected by its initial rigidity at

large normalized strengths as discussed earlier. In spite of this, for most isolation periods and

normalized strengths, the displacement of an FP system is significantly underestimated by the

equivalent-linear procedure.

95
5.4.1 Extension to Bidirectional Excitation

Although factors are included to account for bidirectional excitation in the code dynamic pro-

cedures, which include site specific response spectrum analysis and response history analysis,

no change to the displacements specified in the static lateral response procedure is required.

Bidirectional excitation is indirectly accounted for in the larger of the two code specified lev-

els of design. The design displacement, corresponding to the design basis earthquake (DBE),

“act(s) in the direction of each of the main horizontal axes of the structure”, while the maximum

displacement, corresponding to the maximum capable earthquake (MCE), “act(s) in the most

critical direction of response” [1]. Thus, it appears that the maximum displacement reflects a

spectrum increase for bidirectional excitation; however, since the MCE is also a stronger earth-

quake, it becomes difficult to separate the two effects. Nevertheless, the design displacement is

used for most applications, while the applications of the maximum displacement are limited to

stability check and vertical load testing of the isolation system, and allowance for components

that cross the isolation interface.

5.5 Conclusions

A procedure to estimate the displacement of an FP system has been developed based on non-

linear response history analysis, as an extension of earlier procedures for LRB systems, leading

to the following conclusions:

1. The median – over an ensemble of ground motions – of the normalized displacement ūb

depends only on the isolation period Tb and normalized strength η. This was achieved

by defining the normalized strength as the system strength ÷ ωd u̇gyo , where u̇gyo is

the peak ground velocity and ωd is the transition frequency separating the velocity and

displacement-sensitive regions of their median spectrum.

2. The displacement was normalized to be independent of ground motion intensity, with

the intention of minimizing the dispersion of the normalized displacement and ultimately

allowing a higher confidence estimate of the peak displacement. For FP systems, the

dispersion of normalized displacement was shown to be small except for large values of

96
normalized strength, for which the isolator either failed to slip consistently or its movement

was quite limited. However, such large normalized strength is not relevant for practical

application, because the isolation system is ineffective unless the isolator slips.

3. An equation, dependent on isolation period and normalized strength, was fit to the exact

median normalized displacement by regression analysis, ultimately leading to a design

equation for the actual displacement of the FP system. Comparable design equations

for unidirectional and bidirectional excitation predict that the peak displacement in any

direction due to bidirectional excitation is considerably – 22% to 38% – larger, depending

on η, than the peak displacement due to unidirectional excitation. The design equations

also predict that for the same isolation period and normalized strength, the displacement

of an FP system is smaller than the deformation of an LRB system, because more energy

is dissipated.

4. For a given spectrum, equivalent-linear procedures in the International Building Code were

shown to generally underestimate the FP system displacement by up to 30% compared

to the exact median displacement determined by nonlinear response history analysis.

97
Appendix 5A: Notation

a∗y yield acceleration of isolated block: Q/m

fb lateral force of isolator or isolation system

fbx , fby lateral force in x and y-directions due to bidirectional excitation

fbo peak lateral force in a single direction (unidirectional excitation) or

any direction (bidirectional excitation)

kb postyield stiffness of FP isolation system: w/R

kef f effective stiffness of FP system for equivalent-linear analysis

m mass of isolated block

Q characteristic strength of FP system, equal to friction force µw

R radius of curvature of sliding surface in FP isolator

Tb isolation period, corresponds to postyield stiffness kb

Td corner period separating velocity and displacement-sensitive regions

of linear spectrum

Tef f effective period of FP system for equivalent-linear analysis

ub lateral displacement of FP system

ubx , uby lateral displacement in x and y-directions due to bidirectional exci-

tation

ubo peak lateral displacement in a single direction (unidirectional exci-

tation) or in any direction (bidirectional excitation)

ubyo peak lateral displacement in the y-direction

ūb (t) or ūb normalized displacement of FP system: ub /u∗y

ūbo peak normalized displacement in a single direction (unidirectional

excitation) or in any direction (bidirectional excitation)

ügx , ügy components of ground acceleration for bidirectional excitation

ū ¨gy
¨gx , ū components of ground acceleration normalized by peak ground ve-

locity

98
u̇gyo peak ground velocity in y-direction for bidirectional excitation

u∗y Q/kb

w weight of isolated block

zx , zy x and y-components of yield function due to bidirectional excitation

ζef f effective damping ratio of FP system for equivalent-linear analysis

η strength of FP system normalized by peak ground velocity:

a∗y /(ωd u̇gyo )

µ characteristic strength ratio of FP system, friction coefficient of slid-

ing surface

ωb isolation frequency, corresponds to postyield stiffness of FP system

ωd corner frequency separating velocity and displacement-sensitive re-

gions of linear spectrum

99
6: MODEL FOR AXIAL-LOAD EFFECTS IN LEAD-RUBBER
BEARINGS

6.1 Introduction

The method developed in previous chapters to estimate peak isolator deformation, which as-

sumes a shear model of the isolated structure, neglects the overturning associated with lateral

vibration that causes the vertical or axial forces in the isolators to vary with time. Due to

concerns about the stability of rubber isolation bearings under large compressive loads and

their ability to withstand tensile loads, the peak – both maximum and minimum – axial forces

are also important design quantities.

Axial loads have been observed to have a significant effect on the response of isolation

bearings. First, a correlation between lateral stiffness and axial load has been observed in

several types of rubber bearings. In tests of disparate bearings – for example, natural rubber,

high-damping rubber (HDR) and lead-rubber (LRB) bearings – with both dowelled and bolted

connections [14, 15, 16], the secant stiffness decreased with increasing axial load. As an excep-

tion, some HDR bearings showed increased stiffness or hardening at large shear strains that was

magnified by large axial loads [16]. The stiffness reduction appeared to be greatest in bearings

with low shape factors.

Rubber bearings have also been shown to soften in the vertical direction at large lateral

deformations. In pure tension, rubber bearings tend to cavitate, or form small cavities in the

rubber that blow out from negative pressure and link together to form cracks in the rubber

matrix. In one test, this occurred at tensile strains of about .0003, corresponding to tensile stress

of 1.59 MPa (230 psi) [11]. However, in recent projects, bearings under large lateral deformation

were jacked up 12 to 20 mm (0.5 to 0.75 inches) with no evidence of cavitation damage [13].

In other characterization tests, the vertical stiffness of three different HDR bearings decreased

to about 0.7 times the nominal stiffness at lateral deformations near one half their bonded

diameter [46].

Although less documented, the yield strength of an LRB bearing, or strength of the lead

100
core, has been observed to vary with axial load, such that a lightly loaded bearing may not

achieve its theoretical strength. This effect was first noted as an immediate decay in the lateral

force-deformation hysteresis area in response to a sudden drop in axial pressure [17]. While

this undesirable behavior is thought by some to be eliminated with proper confinement of the

lead core, it continues to be observed in current research. In a 3-story structure isolated with

LRB bearings and subjected to triaxial ground excitations, the bearings beneath heavily-loaded

interior columns showed much greater strength and energy dissipation than identical bearings

beneath lightly-loaded exterior columns [18].

In spite of the preceding evidence, existing nonlinear models do not account for the inter-

action between axial loads and lateral/vertical response of the isolators. The objective of this

chapter, therefore, is to develop a model for isolators, specifically LRB bearings, that includes

this interaction, and a numerical implementation of this model for dynamic analysis. The

model is a nonlinear extension of a two-spring model developed from linear stability theory of

multi-layer bearings [36]. Optionally included in the nonlinearity is an empirical representation

of the strength variation of the lead core with axial loads. All behaviors of the model, which

include variation of lateral stiffness and yield strength with axial load, and variation of verti-

cal stiffness with lateral deformation, are confirmed by unpublished experimental data. From

hereafter these behaviors are referred to as axial-load effects. Following development of the

algorithm to implement this bearing model, the response of the new model to a simple seismic

pulse is explored. Unfortunately this model, even without the optional strength variability,

is less suitable for HDR bearings, because it lacks hardening at large shear strains and other

complex behavior typical of such bearings.

6.2 Stability Analysis of Multi-Layer Bearings

In the stability analysis of multi-layer bearings, [36, 47], the bearing is treated as a continuous

composite system in which the steel layers do not deform, allowing prediction of the buck-

ling load and the effective lateral stiffness in the presence of axial load. The stability theory

resembles the linearized theory of an elastic column, but accounts for shear deformation by

considering rotation of the cross-section, which is independent of the lateral deflection [47].

101
δbz
P
P
δbz

Figure 6.1: Response of an axially-loaded multi-layer rubber bearing under shear deformation,
including tilting of the rubber layers, in response to (a) compression loading and (b) tensile
loading. Note that middle reinforcing layers tilt the most.

ub P
fb
PE hb
2 Hinge δbz

θ
)

hb
θ
)

kbo
s
Hinge
PE hb
2

Figure 6.2: Two-spring model of an isolation bearing in the undeformed and deformed con-
figuration.

Also predicted by the stability analysis [47], the multi-layer bearing under simultaneous lateral

and axial loading undergoes an additional vertical displacement beyond that due to material

axial flexibility. Demonstrated visually in Fig. 6.1, the additional displacement δbz , either com-

pressive or tensile depending on the corresponding axial load, is due to tilting of the middle

bearing layers while the bearing is under shear deformation.

102
6.2.1 Approximate Force-Deformation Relation Based on Linear Two-Spring Model

The preceding effects predicted by stability analysis can be represented, approximately, by a

simplified two-spring model of the bearing [36] that results in explicit force-deformation rela-

tions. The two-spring model (Fig. 6.2) is a composition of rigid tees connected by a rotational

spring, subdivided at top and bottom, and a shear spring with frictionless rollers at midheight.

The bottom plate is fixed and the top plate is constrained against rotation. Axial flexibility of

the bearing is included by an additional vertical spring in series (not shown in Fig. 6.2).

Assuming linear material behavior, the nominal shear stiffness of a multi-layer bearing,

also the stiffness of the shear spring in Fig. 6.2, is

GA GAs
kbo = = (6.1)
tr hb

where G is the shear modulus, A the cross-sectional area, and tr the sum thickness of the rubber

layers. In some cases, it is convenient to use modified area As = A(hb /tr ), where hb is the total

height of the bearing, which accounts for the undeforming steel layers. Similarly, the nominal

vertical stiffness of the bearing, i.e., stiffness of the additional vertical spring, is

Ec A Ec As
kbzo = = (6.2)
tr hb

where Ec is the instantaneous compression modulus of the rubber-steel composite bearing. In

this context, the term nominal (denoted by subscript ’o’) means absent axial-load effects.

If the shear stiffness of the two-spring model (Fig. 6.2) were infinite, the rotational stiffness

divided by hb would equal the conventional Euler buckling load PE = (π 2 /h2b )EIs . Here EIs is

the bending stiffness of a multi-layer bearing [36]:

1 hb
EIs = Ec I (6.3)
3 tr

where I is the conventional moment of inertia: πD4 /64 or Arb2 , and rb = D/4 is the bending

radius of gyration in terms of the bearing diameter D. Thus, the rotational stiffness, divided

equally among top and bottom springs (Fig. 6.2), is PE hb .

103
With these simple linear constitutive relations, the equilibrium equations relating the

lateral force fb and axial (compressive) force P to the deformation s across the shear spring

and the rotation θ through the rotational spring (Fig. 6.2) are

fb − kbo s + P θ = 0 (6.4a)

fb hb − PE hb θ + P (s + hb θ) = 0 (6.4b)

which assume small rotation θ. The axial force P and the deformation v across the vertical

spring (not shown in Fig. 6.2) are related by an additional equation:

P − kbzo v = 0 (6.4c)

The kinematic equations relating the total lateral deformation ub and vertical deformation

ubz to the internal deformations s, θ and v, again assuming θ to be small, are:

ub = s + hb θ (6.5a)

ubz = v + δbz
hb 2
= v + sθ + θ (6.5b)
2

In Eq. 6.5b, ubz – positive in compression – is the sum of v, the deformation resulting from

axial flexibility of the bearing, and δbz , the additional vertical displacement that occurs in the

laterally deformed configuration as shown in Figs. 6.1 and 6.2.

Analysis of the linear two-spring model originally presented in [36] is summarized in the

following steps:

1. The system of equations in s and θ (Eqs. 6.4a and b) is solved for the critical buckling

load by setting the matrix determinant to zero:



Pcr ≈ ± PS PE (6.6)

where PS = GAs = kbo hb . Equation 6.6 approximates the buckling load determined from

stability analysis of the multi-layer bearing [36, Eq. 8.12], with a reasonable assumption

that PE  PS .

104
2. The values of the shear deformation s and rotation θ, found by solving the same system

of equations, are substituted into Eq. 6.5a, which, with some approximation, simplifies to

the following lateral force-deformation relation:


  2 
P
fb = kb ub kb = kbo 1 − (6.7)
Pcr

Again, the lateral stiffness kb in Eq. 6.7 for the two-spring model is a good approximation

to the stiffness derived from stability analysis of the multi-layer bearing [36, Fig. 8-4].

3. The values of s and θ, determined as described above, and axial deformation v from

Eq. 6.4c, are substituted into Eq. 6.5b. This leads to the following flexibility equation for

vertical deformation as a function of force:

P (PS + P ) u2b
ubz = + (6.8)
kbzo PE hb

which is inverted to give:


 
PS u2b
P = kbz ubz − (6.9)
PE hb
with incremental, or tangent vertical stiffness kbz given by:
 −1  −1
1 u2b 3u2
kbz = + = kbzo 1 + 2 b2 (6.10)
kbzo PE hb π rb

reduced to its final form by substituting Eq. 6.3 for PE . Thus, the lateral and vertical

force-deformation relations of the bearing are defined by the coupled equations 6.7 and

6.9. If axial-load effects are neglected (take axial load P = 0 in Eq. 6.7 and lateral

deformation ub = 0 in Eq. 6.9), these equations reduce to the familiar uncoupled linear

force-deformation equations with nominal stiffnesses kbo (Eq. 6.1) and kbzo (Eq. 6.2).

Lateral force-deformation curves for different axial loads are shown in Fig. 6.3a, with the

curve corresponding to no axial load (P = 0) representing the nominal stiffness. Each curve

is linear because P is constant. The reduction in lateral stiffness as the axial load increases

can be seen here as well as in Fig. 6.4. The stiffness variation is inconsequential for typical

design values of P/Pcr < 0.2. However, the stiffness rapidly approaches zero as the axial load

approaches the critical load.

105
(a) (b)
0.4 1.2

0.3 P =0
0.2Pcr 1
0.2 0.4Pcr
Lateral force fb /Pst

0.8

Axial force P/Pst


0.6Pcr
0.1 0.8Pcr
0.6
0
0.4
−0.1 ub =0
0.2 5 cm
−0.2 10 cm
15 cm
−0.3 0 20 cm
−0.4 −0.2
−20 −10 0 10 20 −0.2 0 0.2 0.4 0.6 0.8 1 1.2
Lateral deformation ub (cm) Vertical deformation ubz (cm)

Figure 6.3: (a) Lateral force-deformation as a function of P/Pcr and (b) vertical force-
deformation as a function of lateral deformation ub . Lateral and axial forces are normalized by
the static, or gravity load Pst .
Lateral to nominal stiffness ratio: kb /kbo

0.8

0.6

0.4

0.2

0 0.2 0.4 0.6 0.8 1


Axial force P/Pcr

Figure 6.4: Influence of the axial load P on the lateral stiffness kb

Consideration of the negative solution to Pcr (Eq. 6.6) leads to a critical load in tension,

hence the unconventional concept of tension buckling [13]. Because the two solutions to Pcr are

equal and of opposite sign, the bearing buckling behavior in tension mirrors that in compression.

Thus, the stiffness reduction of Eq. 6.7 is the same for both compressive and tensile loads P .

Written in flexibility form (Eq. 6.8), the vertical force-deformation relation shows two

distinct contributions to the vertical deformation, a material effect and a second order geometric

effect. The additional vertical displacement δbz due to geometry of deformation (second term

106
of Eq. 6.8) increases according to the square of the lateral deformation. Note that even with

no axial load on the bearing, the additional displacement δbz is nonzero, albeit small because

PE  PS , as shown in Fig. 6.3b at P = 0 for different lateral deformations ub . The net effect of

this additional displacement is an overall softening of the bearing in the vertical direction, which

depends on the lateral deformation relative to the bending radius of gyration (Eq. 6.10). This

softening is also evident in the vertical force-deformation curves for different lateral deformations

(the curves are linear because lateral deformation is constant) shown in Fig. 6.3b, wherein the

slope of the curve for the laterally undeformed configuration (ub = 0) represents the nominal

vertical stiffness (Eq. 6.2). Physically, the softening occurs as a result of the tilting of bearing

reinforcing layers (Fig. 6.1), meaning the axial loads are resisted in part by shear.

6.2.2 Effect of Bearing Shape Factors and Compressibility

The stability of any bearing is closely related to its geometry, represented by the first and

second shape factors S and S2 . Defined as the ratio of the loaded area to the force free area,

the shape factor S for a circular bearing = D/4t, where t is the thickness of a single rubber

layer. The second shape factor S2 is simply the bearing aspect ratio based only on the rubber

thickness: D/tr for a circular bearing.

The compression modulus Ec , needed in the nominal vertical stiffness (Eq. 6.2), is a

function of the shape factor S [36]:

Ec = 6GS 2 (6.11)

Substituting Eqs. 6.11 and 6.3 into Eq. 6.6 leads to


  
hb 1 2 h
Pcr = GA π (6GS 2 )Arb2
tr 3 tr
πGASS2
= √ (6.12)
2 2

The effect of axial load on the lateral stiffness (Eq. 6.7) is more significant for low shape factor

bearings because Pcr (Eq. 6.12) is smaller, and hence the ratio of design load to Pcr is larger.

The theory and equations presented thus far, assuming rubber to be incompressible,

should be modified for shape factors S > 10 to include the effect of bulk compressibility [36].

107
In terms of the bulk modulus κ, the compression modulus is modified to [48]:
 
1 1 8GS 2

= 1+ (6.13)
Ec Ec κ

and the bending stiffness is modified to [48]:


 
1 1 3GS 2
= 1+ (6.14)
EIs EIs κ

The approximate Eqs. 6.13 and 6.14 are valid for shape factors S ≤ 25, and can be rewritten

as
Ec EIs
Ec = EIs = (6.15)
1 + Kc S 2 1 + Ks S 2
where Kc and Ks depend on the ratio of the shear modulus to bulk modulus G/κ, which ranges

from 1/3000 to 1/2000 for typical values of G and κ. Replacing Ec by Ec and EIs by EIs in

Eqs. 6.12 and 6.10 leads to modified equations for Pcr and kbz :

πGASS2
Pcr =  (6.16)
2 2(1 + Ks S 2 )
 −1
kbz 3u2 (1 + Ks S 2 )
= 1 + 2 b2 (6.17)
kbzo π rb (1 + Kc S 2 )
6.3 Variation of Yield Strength with Axial Load

Experimental evidence [17, 18] of the bearing failing to achieve its full strength when lightly

loaded has been attributed to lack of confinement of the lead plug. In the words of Skinner

[3], “the nominal upper limit of hysteretic force ... should be achieved if there is no vertical

slippage of the plug sides and no horizontal slippage of the plug ends.” Although side slip can

be reduced by decreasing the spacing between steel layers, or using a lead plug slightly larger

than the undeformed cavity to be occupied, the overall slippage unavoidably depends on the

confining pressure and hence the axial load.

The preceding experimental observation has not, to our knowledge, been verified by

mechanical analysis. However, we have developed an empirical equation for the yield strength

as a function of the compressive load P based on experimental data (see Sec. 6.4):
 
Q = Qo 1 − e(−P/Po ) (6.18)

108
1.2

Strength relative to nominal: Q/Qo

P = Po
1

0.8

0.6

0.4

0.2

−0.2
−5 0 5 10
Ratio of P/Po

Figure 6.5: Influence of the axial load P on the yield strength Q

where Qo is the nominal yield strength of the bearing, achievable with an adequate confining

pressure; and Po is the axial load corresponding to about 63% of nominal strength. In Fig. 6.5,

a plot of Eq. 6.18 shows that the strength declines quite rapidly for loads P below Po . The

strength exceeds 95% of its nominal value at loads P ≥ 3Po , therefore we recommend a minimum

of 3Po for the bearing design load Pst , although different scenarios will be explored. When the

bearing is in tension (P < 0), the effective yield strength is taken to be zero because the lead

core tends to rotate instead of shearing.

6.4 Verification of Theory Based on Experimental Data

In conjunction with earthquake simulator tests of a model three-story reinforced concrete build-

ing, prototype HDR bearings (from Bridgestone Corporation and Malaysian Rubber Products

Research Association or MRPRA) and LRB bearings (from Oiles Industory Co.) were tested

to determine their mechanical characteristics. Details of the experimental test program and

some results were published [11, 49]; but much of the test data, especially for MRPRA and

Oiles bearings, was never reported. Although these tests were performed over 10 years ago, the

extensive data is a valuable resource to assess axial-load effects on rubber bearings.

6.4.1 Axial-Load Varied Tests

Sinusoidal displacement-controlled tests of the bearing at a specified shear strain sequence were

repeated at varying axial loads; the postyield stiffness kb and strength Q were measured from

109
the recorded data (Appendix 6A). In reality, the bearing properties vary with shear strain;

thus, the nominal stiffness kbo (at zero axial load) and strength Qo (at the maximum applied

axial load) were observed at each strain level, and the critical load Pcr (Eq. 6.16) – based on

the observed nominal shear modulus Go – was computed at each strain level. Our proposed

equations for postyield stiffness kb (Eq. 6.7) and strength Q (Eq. 6.18) obviously ignore strain

dependence, but so do typical models for LRB bearings because the strain dependence is minor

and properties at large strains control the peak response of the system.

Figure 6.6 compares the normalized postyield stiffness kb /kbo versus P/Pcr observed in

these tests for the three different bearings with Eq. 6.7. For both the Bridgestone (Fig. 6.6a)

and MRPRA bearings (Fig. 6.6b), the theoretical stiffness (Eq. 6.7) agrees well with the exper-

imental data, matching best at larger strains where the data is considered to be most reliable.

Data for the Oiles bearings (Fig. 6.6c) agrees less with Eq. 6.7, partly due to the difficulty of

obtaining a reliable estimate of the nominal stiffness kbo from the test data (Appendix 6A);

however, the slope of the experimental curves appears to match the slope of Eq. 6.7. Con-

sidering the many possibilities for experimental error, the theoretical model and experimental

results seem to be in satisfactory agreement.

On a side note, observe that the range of P/Pcr is different for each bearing, though

they are tested under identical axial loads P . The wide range of shape factors for the bearings

(S between 8.75 and 20 and S2 between 2.86 and 4, listed in Table 6A.1), led to significant

variations in the critical load Pcr (Eq. 6.16). Confirming an earlier assertion, the stiffness

variation due to axial loads is greater in bearings with low shape factors, or thick rubber layers

relative to their size, like MRPRA, than in bearings with high shape factors, like Bridgestone.

The proposed empirical model for strength variation with axial load in LRB bearings

(Eq. 6.18) is based solely on data for the Oiles LRB bearing. To show that this model is

justified by the data, the observed strength ratio Q/Qo as a function of P/Po is compared to

Eq. 6.18 for two different Oiles bearings (Fig. 6.7). The nominal strength Qo at each strain

level is the observed strength of the first bearing at its largest applied load (3Pst = 235 kN),

and the load Po for each bearing was selected visually to best fit the data. Po should be the

110
(a) HDR1 - Bridgestone
Bearing 1 (Pst =78 kN) Bearing 2 (Pst =49 kN)
1.25 1.25
Stiffness relative to nominal: kb /kbo

1 1

0.75 0.75

0.5 Model: 1-(P/Pcr )2 0.5


Obs: strain = 25%
strain = 50%
0.25 strain = 75% 0.25
strain = 100%
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

(b) HDR2 - MRPRA


Bearing 1 (Pst =78 kN) Bearing 2 (Pst =49 kN)
1.25 1.25
Stiffness relative to nominal: kb /kbo

1 1

0.75 0.75

0.5 0.5

0.25 0.25

0 0

0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2

(c) LRB - Oiles


Bearing 1 (Pst =78 kN) Bearing 2 (Pst =49 kN)
1.25 1.25
Stiffness relative to nominal: kb /kbo

1 1

0.75 0.75

0.5 0.5

0.25 0.25

0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Axial load relative to Pcr : |P/Pcr | Axial load relative to Pcr : |P/Pcr |

Figure 6.6: Experimentally observed stiffness ratio kb /kbo vs P/Pcr for (a) Bridgestone HDR,
(b) MRPRA HDR, and (c) Oiles LRB, compared with Eq. 6.7. Data points at P = 0, Pst /2,
Pst , 2Pst , 3Pst , and −Pst /10, where Pst is the design axial load.

111
Bearing 1 (Pst =78 kN) Bearing 2 (Pst =49 kN)
1.2 1.2
Strength relative to nominal: Q/Qo

1 1

0.8 0.8

0.6 0.6
Model 1−exp(−P/Po)
0.4 Obs: strain = 25% 0.4
strain = 50%
strain = 75%
0.2 0.2
strain = 100%

0 0
0 2 4 6 8 0 2 4 6 8
Axial load relative to Po : P/Po Axial load relative to Po : P/Po

Figure 6.7: Strength ratio Q/Qo vs P/Po for Oiles LRB compared with Eq. 6.18. Data points
at P = 0, Pst /2, Pst , 2Pst , and 3Pst , where Pst is the design axial load.

same for both bearings, which are nominally identical; thus the differences in observed values

of 30 and 20 kN for Po may be due to accidental variation and testing under different axial

loads. The associated design loads: Pst = 78 kN and 49 kN, for interior and exterior bearings,

respectively, are just less than the recommended 3Po (Sec. 6.3).

The proposed empirical model (Eq. 6.18), depicted as a solid line, proves to be a reason-

able fit to the bearing data (Fig. 6.7), especially at the larger strains where the data is more

reliable. While the observed strength does not reduce to zero, it does fall exponentially as axial

load is removed. It was documented that at zero axial load the lead core began to extrude out of

both ends of the bearing during testing, seeming to support our claim that the lead core would

be ineffective in tension (Sec. 6.3). However, the Oiles bearings could not be tested in tension

because they were attached by dowelled connections, which may or may not have contributed

to the extrusion of the lead core.

6.4.2 Vertical Characteristic Tests and Offset Tests

In the vertical characteristic tests, a cyclic force-controlled loading in the range Pst ± 0.3Pst

(Pst = 78 or 49 kN) was applied to each bearing, and its vertical deformation was recorded

(Appendix 6A). These tests are used to estimate damping in the vertical direction. The elliptical

appearance of the force-deformation relation (Fig. 6A.4) suggests viscous energy dissipation in

112
vertical motion. A damping coefficient for vertical motion is estimated from the dissipated

energy ED , equal to the area of the force-deformation loop:

1 ED
ζeq = (6.19)
4π 1/2kbz u2bz

where kbz is the secant stiffness and ubz is one-half the deformation, measured as average peak-

to-peak over all cycles (Appendix 6A). From the experimental data, such as Fig. 6A.4, the

average damping coefficient was determined as ζeq = 0.064, 0.093 and 0.079 for the Bridgestone,

MRPRA, and Oiles bearings, respectively. Damping is observed to increase as the bearing shape

factor (Table 6A.1) decreases.

These estimated damping coefficients are not expected to be highly accurate because

techniques for recording vertical deformation were unreliable and the data used to estimate the

energy dissipation was noisy (Appendix 6A). Furthermore, although interpreted as exactly in-

phase, the axial forces and vertical deformations may have been slightly out-of-sync, introducing

a phase lag that falsely increases damping. Although such questions exist, the stated values

provide rough estimates of the damping coefficients, which support the choice of 5% viscous

damping for dynamic analysis.

The offset tests were simply a repeat of the vertical characteristic tests at different im-

posed shear strains, with vertical cycling over the loading ranges Pst ± 0.3Pst or Pst ± Pst .

Taking the nominal stiffness kbzo to be the observed stiffness at zero shear strain, the stiffness

ratio kbz /kbzo as a function of lateral deformation ÷ bending radius (ub /rb ) is compared with

Eq. 6.17 (Fig. 6.8). The theoretical and experimental stiffness ratios are in good agreement for

the Bridgestone bearings (Fig. 6.8a) and the Oiles LRB bearings (Fig. 6.8c). Unfortunately,

the predicted reduction in stiffness does not agree with test data for the MRPRA bearings

(Fig. 6.8b), and currently we have no explanation for this discrepancy. (The difficulties in mea-

suring vertical deformation should not significantly affect the relative stiffness ratio kbz /kbzo ).

However, evaluating the total data set, including both axial-load varied and offset tests, the

experimental data provides a reasonable confirmation of the theory.

113
(a) HDR1 - Bridgestone (b) HDR2 - MRPRA
1.2 1.2
Vertical stiffness ratio: kbz /kbzo

Vertical stiffness ratio: kbz /kbzo


1 1

0.8 0.8

0.6 0.6

0.4 Model: Eq. 6.17 0.4


Br1, Pst ±0.3Pst
0.2
Br1, Pst ±Pst 0.2
Br2, Pst ±0.3Pst
Br2, Pst ±Pst
0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
Deformation/bending radius (ub /rb ) Deformation/bending radius (ub /rb )

(c) LRB - Oiles


1.2
Vertical stiffness ratio: kbz /kbzo

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5
Deformation/bending radius (ub /rb )

Figure 6.8: Vertical stiffness ratio kbz /kbzo vs ub /rb for (a) Bridgestone HDR, (b) MRPRA
HDR, and (c) Oiles LRB. Experimentally observed values for two different bearings and two
different load sequences are compared with Eq. 6.17. Pst =78 kN for Bearing 1 and 49 kN for
Bearing 2.

6.5 Nonlinear Extension of Two-Spring Model

The two-spring model of Sec. 6.2.1 adequately accounts for axial-load effects in a bearing with

linear material behavior. However, most isolation bearings are inherently nonlinear, especially

LRB bearings due to the yielding of the lead core, which will require extension of the two-

spring model to include various constitutive models for the shear spring. Three such models are

included: the coupled linear model uses a linear shear spring, identical to the model of Sec. 6.2.1.

The coupled nonlinear constant-strength model uses a bilinear force-deformation relation for

the shear spring, as does the coupled nonlinear variable-strength model, which additionally

incorporates the strength variation of Eq. 6.18. These latter models are hereafter shortened

114
to constant-strength and variable-strength models. Also defined for reference are comparable

bearing models that neglect axial-load effects, such that the lateral force-deformation relation,

either linear (uncoupled linear) or bilinear (uncoupled nonlinear), is uncoupled from the vertical

force-deformation relation, assumed to be linear. The uncoupled nonlinear model represents

the same lateral force-deformation relation as Eq. 3.2.

6.5.1 Governing Equations

For the coupled linear model, the governing equations (Eqs. 6.4 and 6.5) are unchanged. Note

that we do not use the approximations that led to Eqs. 6.7 and 6.9. To consider a different

constitutive model for the shear spring, Eq. 6.4a becomes:

fb − fs (s) + P θ = 0 (6.20)

where the linear shear spring has been replaced by a general force fs (s).

The force of the bilinear shear spring in the constant-strength model may be deter-

mined numerically by classical rate-independent unidirectional plasticity [50], where the force-

deformation relation is elastic-plastic with kinematic hardening, and thus governed by the

following constitutive law, yield function, flow rule, and hardening law:

fs = kI (s − sp ) (6.21a)

Φ(fs , q) = |fs − q| − fy ≤ 0 (6.21b)

ṡp = γ̇ sgn(fs − q) (6.21c)

q̇ = γ̇ Hsgn(fs − q) (6.21d)

respectively. The constitutive law (Eq. 6.21a) says the spring force equals the initial stiffness

kI times the elastic component of deformation s − sp , sp being the plastic deformation. The

initial stiffness kI = kbo + Qo /sy depends on the nominal stiffness and strength, as well as

the yield deformation sy . The yield function Φ (Eq. 6.21b) determines the set of admissible

forces, where the back force q stores the translation of the yield surface, and the yield force

fy = Qo + kbo sy . The spring response is elastic inside the yield surface (Φ < 0), and plastic

115
flow occurs on the yield surface (Φ = 0), determined by the flow rule (Eq. 6.21c) with constant

slip rate γ̇. Evolution of the back force is governed by the hardening law (Eq. 6.21d), with

hardening stiffness H = kI kbo /(kI − kbo ).

To implement the strength variation due to axial load (variable-strength model), the yield

force fy in Eq. 6.21b is updated consistent with the strength Q (Eq. 6.18) at each time instant.

The initial stiffness kI remains constant, determined by the nominal strength Qo rather than

the current strength Q. For both the constant-strength and variable-strength models, fs (s)

can be computed to satisfy Eq. 6.21 via the return mapping algorithm [50]; details are given in

Appendix 6B.

6.5.2 Numerical Implementation

A numerical routine is presented to implement the bearing model into a typical dynamic analysis

program. Based on the stiffness approach, the program is assumed to compute the deformations

of the system at each time step iteratively. Given the bearing deformations, a local routine for

the bearing computes the bearing forces to satisfy the governing equations (Eqs. 6.20, 6.4a,b

and 6.5) and the bearing stiffness matrix, which are returned for use in the global procedure.

Bearing Forces

Let F = fb , P T represent the vector of independent bearing forces that are to be computed

in the local bearing routine. The resultant moment M1 , found by equilibrium of the bearing in

the deformed configuration, should also be applied at the top of the bearing:

M1 = (fb hb + P ub ) (6.22)

The governing equations (Eqs. 6.20, 6.4b,c, and 6.5a,b) represent a system of five nonlinear

equations in five unknowns: fb , P , s, θ and v. This system of equations is recast in root-finding

116
form: ⎧ ⎫ ⎧ ⎫

⎪ ⎪
g1 ⎪ ⎪
⎪ fb − fs (s) + P θ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ g ⎪
⎪ ⎪
⎪f h − P h θ + P (s + h θ)⎪

⎪ 2
⎨ ⎬ ⎨⎪ ⎪ b b E b b ⎪

g = g3 =
⎪ ⎪ ⎪ ⎪
⎪ P − kbzo v ⎪
(6.23)

⎪ ⎪ ⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ g ⎪
⎪ ⎪
⎪ u − s − h θ ⎪



4

⎪ ⎪

b b ⎪


⎪ ⎪ ⎪ ⎪
⎩g5 ⎭ ⎩ u − sθ − h θ 2 /2 − v ⎪
⎪ ⎪ ⎭
bz b

and solved by Newton’s method, i.e., find x = fb , P, s, θ, vT to satisfy g(x) = 0.

The converged solution x at the previous global iteration serves as an initial guess x(0) .

An improved solution is found by:

x(k) = x(k−1) − [J (x(k−1) )]−1 g(x(k−1) ) (6.24)

for k = 1, 2, . . . , where the Jacobian J(x) (Jij = ∂gi /∂xj ), determined from Eq. 6.23:
⎡ ⎤
∂fs
⎢1 θ − ∂s P 0 ⎥
⎢ ⎥
⎢h (s + h θ) (P − PE )hb 0 ⎥
⎢ b b P ⎥
⎢ ⎥
⎢ ⎥
J (x) = ⎢ 0 1 0 0 −k bzo ⎥ (6.25)
⎢ ⎥
⎢ ⎥
⎢0 −1 −h 0 ⎥
⎢ 0 ⎥
⎣ ⎦
0 0 −θ −(s + hb θ) −1

has a nonzero determinant, and hence is invertible, even when P , s, and θ are zero. Equa-

tion 6.24 is applied repeatedly until the incremental change in the solution is less than a desired

tolerance, that is:  x(k) − x(k−1) < tol. At each iteration the shear spring force fs (s) in

Eq. 6.23 and tangent ∂fs /∂s in Eq. 6.25 are computed by the return mapping algorithm (Ap-

pendix 6B).

Bearing Stiffness Matrix

Next, the bearing stiffness matrix kb, relating a change in the bearing forces dF = dfb , dP T

to the change in deformations dU = dub , dubz T , is derived in three steps:

1. Take differentials of the equilibrium equations (Eq. 6.20, Eq. 6.4b,c), resulting in:

keq dv = T dF (6.26)

117
where dv = ds, dθ, dvT , and the matrices keq and T are given by:
⎡ ⎤ ⎡ ⎤
∂fs
⎢ ∂s −P 0 ⎥ ⎢1 θ ⎥
⎢ ⎥ ⎢ ⎥
keq =⎢
⎢−P (PE − P )hb 0 ⎥
⎥ and T = ⎢hb (s + hb θ)⎥

⎥ (6.27)
⎣ ⎦ ⎣ ⎦
0 0 kbzo 0 1

2. Take differentials of the kinematic equations (Eq. 6.5), resulting in:

dU = T T dv (6.28)

3. Substitute dv from Eq. 6.26 into Eq. 6.28:

 
dU = T T feq T dF (6.29)

where feq = keq −1 . The resultant flexibility matrix fb = T T feq T , which relates the

incremental deformations dU to the incremental forces dF , is given explicitly as:


⎡ ⎤
(PE +P )hb +h2b (∂fs /∂s) PE hb θ+(P +hb (∂fs /∂s))(s+hb θ)
⎢ hb (∂fs /∂s)(PE −P )−P 2 hb (∂fs /∂s)(PE −P )−P 2 ⎥
fb = ⎣ ⎦ (6.30)
PE hb θ+(P +hb ∂fs /∂s)(s+hb θ) (PE +P )hbθ 2 +2P sθ+(∂fs /∂s)(s+hb θ)2 1
hb (∂fs /∂s)(PE −P )−P 2 hb (∂fs /∂s)(PE −P )−P 2
+ kbzo

Given the complexity of Eq. 6.30, the bearing stiffness matrix kb, the inverse of fb, is not

derived explicitly. However, the following observations are relevant:

1. Suppose both the axial load P and deformations s and θ are zero, and the shear spring is

linear (∂fs /∂s = Ps /hb ); then the flexibility matrix (Eq. 6.30) is diagonal with elements:

(PE + PS )hb 1
f11 = and f22 = (6.31)
PE PS kbzo

Typically PE  PS , giving f11 ≈ hb /PS = 1/kbo , such that the lateral and vertical bearing

stiffnesses reduce to their nominal values [k11 ≈ kbo (Eq. 6.1) and k22 = kbzo (Eq. 6.2)].

2. Suppose just the deformations s and θ are zero, and the shear spring is linear; then the

flexibility matrix is diagonal with:

(PE + P + PS )hb 1
f11 = and f22 = (6.32)
PS PE − PS P − P 2 kbzo

118
Inverting f11 and again taking PE  PS gives:
   
(PS PE − P 2 ) P 2
k11 ≈ = kbo 1 − (6.33)
PE hb Pcr

which is identical to the approximate lateral stiffness kb in Eq. 6.7.

3. Suppose hb θ is small relative to the total lateral deformation ub , the shear spring is linear,

and P  Pcr ; then f22 can be approximated as:

1 u2b 1
f22 ≈ + (6.34)
PE hb kbzo

whose inverse is the tangent vertical stiffness kbz derived earlier (Eq. 6.10).

Thus, the exact stiffness matrix kb has been shown with relevant assumptions to lead to the

same approximate force-deformation equations (Eq. 6.7 and 6.9) given earlier, a good check on

its accuracy. An outline of the complete bearing routine, including solving for the force vector

F and stiffness matrix kb, is given in Appendix 6B.

6.6 Observed Response of New Bearing Models

6.6.1 Lateral and Vertical Force-Deformation Trends

Lateral force-deformation behavior for the coupled linear model was already demonstrated in

Fig. 6.3a, where the lateral stiffness was shown to decrease as the applied force P approached

Pcr . The lateral force-deformation relation of the constant-strength model (Fig. 6.9a) is closely

related to the coupled linear model. The postyield stiffness in Fig. 6.9a is essentially identical

to the stiffness of the coupled linear model (Fig. 6.3a), showing the same successive decline as

the axial force is increased toward Pcr . The initial stiffness, and energy dissipated in a single

cycle (loop area), are affected by axial force only negligibly. As a special case, the curve with

P = 0 is equivalent to the uncoupled nonlinear model, which neglects axial-load effects.

The primary effect of including the strength variation of Eq. 6.18 (variable-strength

model) is that the yield strength, and hence energy dissipated in a single cycle, also depends

on the axial force P (Fig. 6.9b). Recall that when P = 0 or small, the strength, and hence

energy dissipated, is also zero or small, as observed in Fig. 6.9b. As axial force becomes large

119
(a) (b)
0.4 0.4
P =0 P =0
0.3 0.2Pcr 0.3 0.05Pcr , 0.38Po
0.4Pcr 0.4Pcr , 1.5Po
0.2 0.6Pcr 0.2 0.6Pcr , 3.75Po
Lateral force fb /Pst

0.8Pcr 0.8Pcr , 6Po


0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

−0.3 −0.3

−0.4 −0.4
−20 −10 0 10 20 −20 −10 0 10 20
Lateral deformation ub (cm) Lateral deformation ub (cm)

Figure 6.9: Lateral force-deformation as a function of P/Pcr for (a) constant-strength model
and (b) variable-strength model.

relative to Po and Pcr , the strength is regained but the postyield stiffness decreases similar

to the constant-strength model (Fig. 6.9a). A well-designed bearing should achieve a balance,

in the sense that P should be large enough to avoid significant strength degradation but well

below Pcr .

For an imposed lateral deformation, the vertical force-deformation relation is independent

of the particular bearing model (linear, constant-strength or variable-strength), and the plots

of Fig. 6.3b for the coupled linear model also represent the constant-strength and variable-

strength models. Clearly, the vertical force-deformation will differ for the different models

when the lateral deformation varies freely during seismic loading.

6.6.2 Response to a Seismic Pulse

For each variation of the bearing model presented in Sec. 6.5, a rigid block supported on two

bearings is subjected to the seismic pulse of Fig. 6.10, and the response of the left exterior

bearing is shown (Figs. 6.11-6.13). These figures demonstrate response histories of the lateral

and vertical deformations and forces, the lateral force-deformation relation, and the vertical

force-deformation relation. For the response histories, the lateral force and the vertical defor-

mation are scaled by factors indicated so that lateral and vertical deformations and lateral and

vertical forces can be viewed at similar amplitudes. These data are shown quantitatively in

120
(a) (b) (c)
150 150 100

100 100
50
50 50
üg (cm/s2 )

u̇g (cm/s)

ug (cm)
0 0 0

−50 −50
−50
−100 −100

−150 −150 −100


0 Tp 0 Tp 0 Tp
Time Time Time

Figure 6.10: (a) Acceleration ügo , (b) velocity u̇go and (c) displacement ugo history of seismic
excitation: approximately 1.5 cycles of a sinusoidal pulse with velocity u̇go = 90 cm/s and
period Tp = 4 seconds. Isolation period Tb for the system excited is 2 seconds.

the force-deformation plots, with forces normalized by the static load Pst . For simplicity the

contribution of bulk compressibility (Eq. 6.15) has been neglected.

The bearing response predicted by the coupled linear model is given (Fig. 6.11), with

the comparable response using the uncoupled linear model shown for reference. The lateral

deformation and axial force of the bearing are nearly in-phase (Fig. 6.11a). As a consequence,

the axial force P increases when the lateral deformation is large and positive, and an associated

drop in stiffness (Eq. 6.7) is observed in the lateral force-deformation relation (Fig. 6.11b).

The lateral stiffness does not appear to change at negative lateral deformations, because the

axial force decreases toward zero, and the lateral stiffness simply defaults to its nominal value

(Fig. 6.4). Including axial-load effects (coupled linear vs. uncoupled linear model) appears to

increase the peak lateral deformation by about 20%.

The vertical deformation is considerably greater when axial-load effects are included (cou-

pled linear model vs. uncoupled linear model, Fig. 6.11a and c). To understand this, recall that

the vertical force-deformation relation depends greatly on the lateral deformation (Eq. 6.9).

Thus, when the axial force is close to its static value, the lateral deformation is close to zero

(recalling the correlation between P and ub observed above) and the vertical force-deformation

121
(a) (b)
0.6
Coupled Linear
Lateral Deformation 0.4 Uncoupled Linear
ub+

Lateral force fb /Pst


0.2
0
0
ub−
−0.2
Vertical Deformation (x15)
ubz
−0.4

0
−60 −40 −20 0 20 40 60
Lateral deformation ub (cm)
Lateral Force (x2)
fb+ (c)
2
0 Axial force P/Pst
1.5
fb−
Axial Force
Pst +P 1

Pst
0.5

Pst -P
0
0 Tp T 0 1 2 3 4 5 6 7
Time (s) Vertical deformation ubz (cm)

Figure 6.11: Response of the left exterior bearing – using the coupled linear model – of a rigid
block subjected to a seismic pulse: (a) lateral and vertical deformation and force histories, (b)
lateral force-deformation, and (c) vertical force-deformation. Also shown is the comparative
response with the uncoupled linear model; Pst /Pcr = 0.4.

is essentially that of the uncoupled linear model. However, as the axial force deviates from

the static force in either direction, the lateral deformation becomes large, resulting in vertical

softening and a considerable increase in vertical deformation. Thus the arc-shaped vertical

force-deformation (Fig. 6.11c), with vertical deformation increasing at axial forces larger or

smaller than Pst . Due to the inherently linear relation between axial force P and deformation

ubz (Eq. 6.9), the total deformation ubz is obviously greater when P > Pst than when P < Pst .

The closed loops exhibited by the above force-deformation relations leave the impression

that energy dissipation occurs (Fig. 6.11b and c), which would violate energy conservation in

an elastic system. However, a simple example in Appendix 6C verifies that energy is conserved.

122
(a) (b)
0.3
Constant-Strength
Uncoupled Nonlinear
Lateral Deformation 0.2
ub+

Lateral force fb /Pst


0.1
0
0
ub−
−0.1
Vertical Deformation (x40)
ubz
−0.2

0
−20 −10 0 10 20
Lateral deformation ub (cm)
Lateral Force (x4)
fb+ (c)
2
0 Axial force P/Pst
1.5
fb−
Axial Force
Pst +P 1

Pst
0.5

Pst -P
0
0 Tp T 0 0.2 0.4 0.6 0.8 1
Time (s) Vertical deformation ubz (cm)

Figure 6.12: Response of the left exterior bearing – using the constant-strength model – of a
rigid block subjected to a seismic pulse: (a) lateral and vertical deformation and force histories,
(b) lateral force-deformation, and (c) vertical force-deformation. Also shown is the comparative
response with the uncoupled nonlinear model; Pst /Pcr = 0.4.

The bearing response predicted by the constant-strength model is plotted in Fig. 6.12

and compared with the response predicted by the uncoupled nonlinear model. Whereas for the

linear models the axial force history and lateral deformation history were in phase (Fig. 6.11a),

for these nonlinear models the lateral force history is more closely in-phase with the lateral de-

formation history (Fig. 6.12a). For the constant-strength model, this causes the lateral stiffness

(initial or postyield) to show the greatest decrease at large positive lateral forces (Fig. 6.12b),

which correspond to the maximum axial forces. The result is a slight increase in peak lateral

deformation compared to the nonlinear model. Because of the changing stiffness, the appar-

ent width of the hysteresis loop changes during the deformation cycle, whereas this width is

123
(a) (b)
0.3
Variable-Strength
Lateral Deformation 0.2 Uncoupled Nonlinear
ub+

Lateral force fb /Pst


0.1
0
0
ub−
−0.1
Vertical Deformation (x40)
ubz
−0.2

0
−20 −10 0 10 20
Lateral deformation ub (cm)
Lateral Force (x4)
fb+ (c)
2
0 Axial force P/Pst
1.5
fb−
Axial Force
Pst +P 1

Pst
0.5

Pst -P
0
0 Tp T 0 0.2 0.4 0.6 0.8 1
Time (s) Vertical deformation ubz (cm)

Figure 6.13: Response of the left exterior bearing – using the variable-strength model – of a
rigid block subjected to a seismic pulse: (a) lateral and vertical deformation and force histories,
(b) lateral force-deformation, and (c) vertical force-deformation. Also shown is the comparative
response with the uncoupled nonlinear model; Pst /Pcr = 0.4 and Pst /Po = 3.

constant for the uncoupled nonlinear model. A high-frequency component is observed in the

axial force cycle (Fig. 6.12a), causing local variations in stiffness and roughness in the lateral

force-deformation relation (Fig. 6.12b).

The vertical force-deformation relation of the constant-strength model is difficult to in-

terpret (Fig. 6.12c). Although quite disordered, it appears hysteretic in nature relative to the

coupled linear model. Perhaps since the axial force variation resembles the lateral force vari-

ation the vertical hysteretic behavior is related to hysteretic behavior in the lateral direction

(Fig. 6.12b). Local variations in the vertical force-deformation relation are related to the high

frequency component of axial force. The peak vertical deformation is significantly less than that

124
of the coupled linear model (Fig. 6.11c), correlating to the reduction of peak lateral deformation.

The bearing response predicted by the variable-strength model, including response his-

tories (Fig. 6.13a) and force-deformation relations (Fig. 6.13b,c) – which are compared with

the response predicted by the uncoupled nonlinear model – is similar to that of the constant-

strength model (Fig. 6.12). The influence of variable strength (Eq. 6.18) is readily apparent

when the axial force is close to zero by comparing Figs. 6.12b and 6.13b. Observe that the

strength decreases and the yield surface contracts at negative lateral forces in the lateral force-

deformation relation (Fig. 6.13b). Compared to the uncoupled nonlinear model, the average

width of the force-deformation loop is smaller, and the associated decrease in energy dissipation

increases the peak lateral deformation, which in turn increases the peak vertical deformation

(Fig. 6.13c). The roughness in the lateral force-deformation increases compared to Fig. 6.12c

as it is affected by local strength variations as well as stiffness variations.

6.7 Conclusions

This investigation of the influence of axial load on the nonlinear response of isolation bearings

has led to the following conclusions:

1. The following axial-load effects have been observed in various tests of both high-damping

rubber and lead-rubber bearings: decreasing lateral stiffness – measured as either the

secant or postyield stiffness – with increasing axial load, decreasing lateral yield strength

with decreasing axial load (lead-rubber bearings only), and decreasing vertical stiffness

with increasing lateral deformation.

2. Based on linear stability theory of an elastic column, a two-spring model of the bearing,

consisting of a shear spring and a rotational spring divided at top and bottom (Fig. 6.2),

accurately accounts for axial-load effects in linear models of bearings. However, bearings

with high-damping fillers or lead cores that provide energy dissipation are nonlinear and

should be modeled as such.

3. Such modeling is achieved by extending the two-spring model to include a nonlinear

constitutive model for the shear spring. Using unidirectional plasticity with kinematic

125
hardening for the shear spring led to the constant-strength model for the bearing. Nu-

merically, this model is implemented by solving a system of five nonlinear equations –

equilibrium (Eqs. 6.20 and 6.4b,c) and kinematic (Eq. 6.5) – by Newton’s method for

the current forces in the bearing, and taking differentials of these equations to derive the

instantaneous stiffness matrix. Although an improvement, the model is an incomplete

representation of high-damping rubber bearings, because it neglects complex effects like

strain hardening.

4. An empirical model (Eq. 6.18) was developed to account for the varying yield strength

in lead-rubber bearings, which can be calibrated to match the experimentally observed

bearing response. This behavior is optionally included by simply updating the yield

force in the plasticity equations to reflect the time-varying yield strength, leading to the

variable-strength model.

5. The behavior of the constant-strength and variable-strength models was demonstrated

in the response of a rigid block supported by two bearings subjected to a seismic pulse.

The effects were most pronounced in the variable-strength model, where the peak lat-

eral deformation increased by about 20% and the peak vertical deformation more than

doubled compared to the uncoupled nonlinear model (without axial-load effects). While

the bearing response predicted by the new models are hereby demonstrated for only one

case, it is desirable to investigate the response for a diverse range of system and bearing

parameters and earthquake excitations, which is the subject of the next chapter.

126
Table 6A.1: Dimensions of Prototype Bearings

Property Bridgestone MRPRA Oiles


Bonded diameter D (mm) 176 140 180
Lead core diameter Dl (mm) - - 25
Rubber thickness per layer t (mm) 2.2 4.0 3.0
Number of rubber layers nr 20 12 21
Total rubber thickness tr (mm) 44 48 63
Thickness per steel shim ts (mm) 1.0 1.6 1.0
Number of steel shims ns 19 11 20
Total bearing height hb (mm) 63 65.6 83
Cross-section area A (mm2 ) 24,388 15,394 25,447
Shape factor S 20 8.75 15
Aspect ratio S2 4 2.92 2.86

Appendix 6A: Interpretation of Experimental Data

This appendix describes how data recorded in the various bearing tests was interpreted. For

reference, the dimensions of the three prototype bearings are listed in Table 6A.1.

In the axial-load varied tests, two bearings from each manufacturer were subjected to the

prescribed lateral deformation, five cycles each at shear strains (lateral deformation relative to

height) of 5, 25, 50, 75 and 100%. The tests were repeated at axial loads of 0, Pst /2, Pst , 2Pst ,

3Pst and -Pst /10, one bearing each for design loads Pst of 78 kN (interior bearing) and 49 kN

(exterior bearing). The tensile tests were omitted for the Oiles LRB bearing, attached by a

dowelled rather than a bolted connection. The lateral force in the bearing in response to the

applied deformation was recorded; this force signal was smoothed using a Gaussian kernel with

full width at half maximum (FWHM) equal to 4*δt. From the deformation and force signals,

the force-deformation relation (bearing hysteresis) for each test was plotted at the various strain

levels. As an example, force-deformation for the Bridgestone bearing at the axial load P = Pst

is shown in Fig. 6A.1.

From the bearing force-deformation, the strength and stiffness at each strain level were

determined independently. Locating the positive and negative y-axis force intercepts in each

cycle and averaging over the cycles, the strength Q was interpreted as half the distance between

the averaged intercepts, which are indicated by dots in Fig. 6A.1. Computed from data points

127
Data File: 911216.07, Bearing: Bridgestone #12, Axial Load Varied: P =78.3 kN
20 20 20
Strain = 5% Strain = 25% Strain = 50%
Bearing force (kN)

10 10 10

0 0 0

−10 −10 −10

−20 −20 −20


−50 0 50 −50 0 50 −50 0 50
Strain = 75% Strain = 100% Complete bearing hysteresis
20 20 20
Bearing force (kN)

10 10 10

0 0 0

−10 −10 −10

−20 −20 −20


−50 0 50 −50 0 50 −50 0 50
Bearing deformation (mm)

Figure 6A.1: Lateral force-deformation at different strain levels and complete bearing hys-
teresis for Bridgestone #12 with applied axial load P = 78 kN. Interpretation of Q and kb given
by dots at the y-intercept and dashed lines.

adjacent to these force intercepts, the postyield stiffness kb was interpreted as the slope at the

force intercepts, averaged over the cycles. For the Bridgestone example, the resultant stiffnesses

are the slopes of the dashed lines drawn through the force intercepts (Fig. 6A.1).

While this procedure worked well for the Bridgestone and MRPRA bearings, some ad-

justments were necessary for the Oiles bearings. A sample force-deformation for Oiles Bearing

1 at P = Pst (Fig. 6A.2) shows a slight pinching of the hysteresis near the origin, i.e., zero shear

strain. In this case, the stiffness was interpreted as the average secant stiffness over a larger

range, from about -63 to +63% of the maximum local strain. Again, the resulting stiffness is

shown as the slope of the dashed lines in Fig. 6A.2.

The Oiles data at zero axial load (P = 0) presented another challenge, as the postyield

stiffness is influenced by the near complete loss of strength (Fig. 6A.3). Although not theoreti-

cally predicted, the interaction between rubber and lead appears to cause the tangent stiffness

128
Data File: 920114.09, Bearing: Oiles #12, Axial Load Varied: P =78.3 kN

20 20 20
Bearing force (kN)

Strain = 5% Strain = 25% Strain = 50%


10 10 10

0 0 0

−10 −10 −10

−20 −20 −20

−50 0 50 −50 0 50 −50 0 50


Strain = 75% Strain = 100% Complete bearing hysteresis

20 20 20
Bearing force (kN)

10 10 10

0 0 0

−10 −10 −10

−20 −20 −20

−50 0 50 −50 0 50 −50 0 50


Bearing deformation (mm)

Figure 6A.2: Lateral force-deformation at different strain levels and complete bearing hys-
teresis for Oiles #12 with applied axial load P = 78 kN. Interpretation of Q and kb given by
dots at the y-intercept and dashed lines.

to decrease at the origin compared to large shear strains. Thus, how should the stiffness,

which will represent the nominal stiffness, be interpreted to give values consistent with those

for nonzero axial loads? Believing the low stiffness near the origin to be an anomaly, we took

the stiffness as the average tangent at about 63% of the maximum local strain, determined

from adjacent data points at that strain. This stiffness, shown drawn through the points about

which it was estimated (Fig. 6A.3), appears indicative of the tangent stiffness away from the

origin. The inconsistency of the methods to determine the stiffness of the Oiles LRB bearing

could explain why it did not match theory as well as the other two (Fig. 6.6).

The vertical characteristic test was applied to 15 each of the Bridgestone and Oiles bear-

ings, and 12 MRPRA bearings, which consisted of a cyclic force-controlled loading over the

range Pst ± 0.3Pst , where Pst was 78 kN or 49 kN. Two bearings of each type were also sub-

jected to the offset tests, a repeat of the vertical characteristic test at imposed offset shear

129
Data File: 920114.07, Bearing: Oiles #12, Axial Load Varied: P =0 kN
20 20 20
Bearing force (kN)

Strain = 5% Strain = 25% Strain = 50%


10 10 10

0 0 0

−10 −10 −10

−20 −20 −20

−50 0 50 −50 0 50 −50 0 50


Strain = 75% Strain = 100% Complete bearing hysteresis

20 20 20
Bearing force (kN)

10 10 10

0 0 0

−10 −10 −10

−20 −20 −20

−50 0 50 −50 0 50 −50 0 50


Bearing deformation (mm)

Figure 6A.3: Lateral force-deformation at different strain levels and complete bearing hys-
teresis for Oiles #12 with applied axial load P = 0 kN. Interpretation of Q and kb given by
dots at the y-intercept and dashed lines.

strains of 0, 50, 100, and 150% and over the ranges Pst ± 0.3Pst and Pst ± Pst .

During the vertical characteristic tests and offset tests, the load signal was applied using

two vertical actuators. The resulting axial force in the bearing was measured by a single load cell

under the bearing, while the vertical deformation was measured by four direct current voltage

transducers (DCDTs) attached at the bearing ‘corners’, and then averaged. The observed

deformations were not very accurate for two reasons: first, the transducers measured the relative

motion between the top and bottom plates of the test machine, which was influenced by plate

bending; second, the vertical bearing deformations were small relative to the resolution of the

transducers. This problem was greatest for the Bridgestone bearings, with peak deformations

on the order of 0.08 mm (.003 in) and a resolution of only 0.025 mm (.001 in). The obvious effect

was incredibly noisy data. A Gaussian kernel was no longer a sufficient filter for smoothing the

data, and better results were obtained by applying a low pass filter that eliminated frequencies

130
Bridgestone Vertical Characteristic Tests: cycles at Pst ± 0.3Pst with Pst=78.3 kN
File: 911120.02, Bearing 1 File: 911204.02, Bearing 2 File: 911205.02, Bearing 3
−40 −40 −40
Vertical force (kN)

−60 −60 −60

−80 −80 −80

−100 −100 −100

−120 −120 −120


−0.05 0 0.05 −0.05 0 0.05 −0.05 0 0.05
File: 911205.05, Bearing 4 File: 911205.08, Bearing 5 File: 911206.02, Bearing 6
−40 −40 −40
Vertical force (kN)

−60 −60 −60

−80 −80 −80

−100 −100 −100

−120 −120 −120


−0.05 0 0.05 −0.05 0 0.05 −0.05 0 0.05
Vertical deformation (mm)

Figure 6A.4: Vertical force-deformation for Bridgestone bearings, as determined from vertical
characteristic tests. Interpretation of vertical stiffness kbz given by dashed lines.

greater than 1 Hz (the test rate was 0.067 Hz).

Samples of the smoothed – though still quite noisy – force-deformation relations from

the vertical characteristic tests of Bridgestone bearings are shown in Fig. 6A.4. The vertical

deformation at the start of the test, which should have indicated the static deformation, was

very inconsistent, and thus the plots were centered at zero deformation. The viscous damping

coefficient, desired from the characteristic tests, was estimated by the following steps: (1) locate

the local extrema – one at each half-cycle – as the maximum of force*deformation (indicated

by dots in Fig. 6A.4), (2) draw a tangent line between each pair of adjacent extreme points, (3)

average the slope of this line over all half cycles to get the vertical stiffness (indicated by dashed

lines in Fig. 6A.4), (4) estimate the energy dissipated in each half cycle by numerically inte-

grating the area between the tangent line and the actual data, (5) average the energy dissipated

over all half cycles and multiply by 2 for energy dissipated per cycle, (6) compute the damping

131
Bridgestone Offset Tests: cycles at Pst ± 0.3Pst (top) or Pst ± Pst (bottom) with Pst=78.3 kN
File: 911218.09, 50% offset File: 911218.12, 100% offset File: 911218.13, 150% offset
−40 −40 −40
Vertical force (kN)

−60 −60 −60

−80 −80 −80

−100 −100 −100

−120 −120 −120


−0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1
File: 911218.10, 50% offset File: 911218.11, 100% offset File: 911218.14, 150% offset
0 0 0
Vertical force (kN)

−50 −50 −50

−100 −100 −100

−150 −150 −150

−0.3 −0.15 0 0.15 0.3 −0.3 −0.15 0 0.15 0.3 −0.3 −0.15 0 0.15 0.3
Vertical deformation (mm)

Figure 6A.5: Vertical force-deformation for a single Bridgestone bearing at offset shear strains
of 50, 100, 150% and load cycles of P ± 0.3P or P ± P . Interpretation of vertical stiffness kbz
given by dashed lines.

coefficient (Eq. 6.19) based on the estimates of vertical stiffness, half peak-to-peak deformation,

and energy dissipated. This damping coefficient was averaged over all characteristic tests for a

particular bearing (Bridgestone, Oiles, MRPRA) to get the values given in Sec. 6.4.2.

The estimates of energy dissipation, in particular, are likely to be influenced by the noisy

data. The numerical integration technique may overestimate the energy dissipation by picking

up noise and/or adding area on both sides of the tangent line as positive energy. For this reason,

the damping coefficients estimated by this technique may be slightly high, but are believed to

give a reasonable impression of vertical damping.

Representative of the offset tests, resulting vertical force-deformation for a Bridgestone

bearing is shown for shear strains of 50, 100 and 150% (Fig. 6A.5). Also shown are the dashed

lines whose slopes represent the vertical stiffness, determined in the same way as for the vertical

characteristic tests. For each load cycle (Pst ± 0.3Pst , Pst ± Pst ), a slight decrease in stiffness

132
Oiles Offset Tests: cycles at Pst ± 0.3Pst (top) or Pst ± Pst (bottom) with Pst=78.3 kN
File: 920115.03, 50% offset File: 920115.05, 100% offset File: 920115.07, 150% offset
−40 −40 −40
Vertical force (kN)

−60 −60 −60

−80 −80 −80

−100 −100 −100

−120 −120 −120


−0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4 −0.4 −0.2 0 0.2 0.4
File: 920115.04, 50% offset File: 920115.06, 100% offset File: 920115.08, 150% offset
0 0 0
Vertical force (kN)

−50 −50 −50

−100 −100 −100

−150 −150 −150

−200 −200 −200


0 5 10 0 5 10 0 5 10
Vertical deformation (mm)

Figure 6A.6: Vertical force-deformation for a single Oiles bearing at offset shear strains of
50, 100, 150% and load cycles of P ± 0.3P and P ± P . Interpretation of vertical stiffness above
and below the design load P given by dashed lines.

with increasing shear strain is observed.

A peculiar effect was observed in the offset tests for the Oiles LRB bearing (Fig. 6A.6).

The response looks normal for load cycles in the range of Pst ± 0.3Pst , but in the larger range of

Pst ± Pst , the bearing softens rapidly as the axial load approaches zero. The effect is magnified

with increase in shear strain. Because the Oiles bearings have dowelled connections, they have a

tendency to rollout at the large shear strains. The axial force couple provides a restoring moment

to resist rollout. Thus, when the axial force is removed, as in these large range offset tests,

significant rollout occurs, and the vertical deformation measurements are affected in unforeseen

ways. The proposed model for the LRB bearing does not really address this behavior, which

is hopefully limited to the increasingly uncommon dowelled connection. We dealt with this

by sampling the tangent vertical stiffness at cycles Pst ± Pst separately above and below Pst

(Fig. 6A.6), and using the stiffness at P > Pst for comparison with the theoretical stiffness

133
(Fig. 6.8).

Appendix 6B: Numerical Implementation of Nonlinear Bearing Model

6B.1 Unidirectional Rate-Independent Plasticity

The basic postulates of unidirectional plasticity were laid out earlier (Eq. 6.21). Recall that

plastic flow (Eq. 6.21c) occurs only when the yield function Φ equals zero (Eq. 6.21b). In

addition, consistency (γ̇ Φ̇ = 0) requires that the force persist on the yield surface for plastic

flow. The slip rate γ̇ can thus be determined by taking the time derivative of the yield function

(Eq. 6.21b), substituting in the constitutive law (Eq. 6.21a), the flow rule (Eq. 6.21c), and the

hardening law (Eq. 6.21d) where appropriate, and rearranging:

sgn(fs − q)kI ṡ
γ̇ = (6B.1)
kI + H

Substituting γ̇ into the flow rule (Eq. 6.21c), the plastic deformation sp is obtained in terms of

the total shear deformation s, which is then introduced into Eq. 6.21a, leading to the deforma-

tion rate of change, or tangent stiffness, during plastic flow:

∂fs kI H
= (6B.2)
∂s kI + H

The tangent stiffness during plastic flow equals the postyield stiffness kbo , thus, from Eq. 6B.2,

the hardening stiffness H is related to kbo : H = kI kbo /(kI − kbo ).

6B.2 Return Mapping Algorithm

Numerical implementation of the nonlinear bearing models necessitates computing the spring

force fs and tangent stiffness ∂fs /∂s for a current estimate of the spring deformation s (Sec. 6.5.2).

This can be accomplished by enforcing the plasticity postulates (Eq. 6.21) using the return map-

ping algorithm [50]. Each of these laws are either discretized or enforced at discrete time steps.

The steps of the algorithm are outlined in pseudocode in Algorithm 1. A trial force

for the spring is computed by taking a trial elastic step (Alg. 1 line 1). If the yield function

(Eq. 6.21b) is satisfied (Φ ≤ 0) then the trial step is confirmed (line 5). If the trial force lies

outside the current yield surface (Φ > 0), the force is returned to the yield surface, through

which the incremental plastic slip ∆γ is defined (line 8). Since plastic flow occurs, the plastic

134
1: fstrial = kI (s − sp ) {Trial elastic step}
2: ξ trial = |fstrial − q|
3: Φtrial = ξ trial − fy {Test yield function for admissibility}
4: if Φtrial ≤ 0 then {Confirm elastic step}
5: (·) = (·)trial
6: ∂fs /∂s = kI
7: else {Plastic step}
8: ∆γ = Φtrial /(kI + H)
9: fs = fstrial − ∆γ kI sgn(ξ trial ) {Return force to the yield surface}
10: sp = sp + ∆γ sgn(ξ trial ) {Update plastic deformation}
11: q = q + ∆γ H sgn(ξ trial ) {Update back force}
12: ∂fs /∂s = kbo
13: end if
14: return: fs , ∂fs /∂s

Algorithm 1: Return Mapping Algorithm

deformation sp and back force q must be updated (Eqs. 6.21c and 6.21d, Alg. 1 lines 10-11).

These variables (q and sp ) are stored separately for each bearing. The tangent stiffness ∂fs /∂s

equals kI during an elastic step (line 6) and kbo during a plastic step (line 12).

6B.3 Return Mapping for Variable-Strength Model

To apply return mapping for the variable-strength model, the yield force fy should be updated

to reflect the variation of strength Q in time: fy = Q + kbo sy . However, if the yield deformation

sy is held constant, the initial stiffness kI varies because fy = kI sy . To avoid complicating the

constitutive law (Eq. 6.21a) with variable initial stiffness, fy is assumed to vary proportional

to Q:
 
fy = fyo 1 − e(−P/Po ) (6B.3)

such that the yield deformation sy also varies in time. Recall that fy = 0 when the bearing is

in tension (P < 0).

Substituting fy (Eq. 6B.3) into Eq. 6.21b, the yield function now reads:
 
Φ(fs , q) = |fs − q| − fyo 1 − e−(P/Po ) (6B.4)

135
1: fstrial = kI (s − sp ) {Trial elastic step}
2: ξ trial = |fstrial − q|
  
3: fy = max 0, fyo 1 − e(−P/Po ) {Update yield force}
4: Φ trial =ξ trial − fy {Test yield function for admissibility}
5: if Φ trial ≤ 0 then {Confirm elastic step}
6: (·) = (·)trial
7: ∂fs /∂s = kI
8: else {Plastic step}
9: ∆γ = Φtrial /(kI + H)
10: fs = fstrial − ∆γ kI sgn(ξ trial ) {Return force to the yield surface}
11: sp = sp + ∆γ sgn(ξ trial ) {Update plastic deformation}
12: q = q + ∆γ H sgn(ξ trial ) {Update back force}
13: ∂fs /∂s = kbo
14: end if
15: return: fs , ∂fs /∂s

Algorithm 2: Return Mapping Algorithm for Variable-Strength Model

Applying the consistency condition, Φ̇ = 0 during plastic flow to this new yield function leads

to:
sgn(fs − q)kI ṡ 1 e−(P/Po ) fyo Ṗ
γ̇ = − (6B.5)
kI + H Po kI + H
The flow rule (Eq. 6.21c) and hardening law (Eq. 6.21d), being functions of the slip rate γ̇,

are indirectly affected by this change in γ̇. But algorithmically, the only change needed is to

update the yield function based on the current axial force P each time the algorithm is called:
 
Φtrial = ξ trial − fyo 1 − e−(P/Po ) (6B.6)

Algorithm 2 lists the complete steps for the variable-strength model.

6B.4 Complete Bearing Routine

Based on Sec. 6.5.2 and above, pseudocode is provided for the complete nonlinear bearing

routine. From a current estimate of the lateral and vertical deformations ub and ubz , the code

returns the force vector F that satisfies equilibrium (Eqs. 6.20 and 6.4b,c) and kinematics

(Eq. 6.5) and the tangent stiffness matrix kb.

136
(a) (b) (c)

(1) (2)
P
ub f3 (3)
Prescribed loads

Vertical Force
Lateral Force
f2 (2)
0
Lateral Deformation

0
Vertical Force 0
(3),v3
0
t0 t1 t2 t3 t4 0 ub 0 v1 v2
Time Lateral Deformation Vertical Deformation

Figure 6C.1: (a) Prescribed load sequence of lateral deformation and axial force, (b) lateral
force-deformation, and (c) vertical force-deformation.

Improved estimates of the bearing forces fb and P , as well as internal deformations s, θ

and v are computed iteratively with Newton’s method (Eq. 6.24) until convergence is attained

(Alg. 3 lines 2-8). In each Newton iteration, the return mapping algorithm (line 3) updates fs

and ∂fs /∂s based on the current value of s. Upon convergence, the matrices feq , T , flexibility

fb and finally stiffness kb, are formed (lines 11-15).

Appendix 6C: Energy Conservation in Coupled Linear Model

To show that energy is conserved in the coupled linear model, consider the following load

sequence: (1) axial load increased linearly from 0 to P , (2) lateral deformation increased linearly

from 0 to ub , (3) axial load decreased linearly from P to 0, (4) lateral deformation decreased

linearly from ub to 0 (Fig. 6C.1a). The resulting lateral force-deformation and vertical force-

deformation are shown in Figs. 6C.1b and c, with each step in the load sequence indicated.

Since the lateral force-deformation does negative work (adds energy to the system) and the

vertical force-deformation does positive work (subtracts energy), energy is conserved if the two

areas are equal.

In terms of the applied load P and deformation ub , the approximate lateral forces are f2 =

PS /hb (1 − P 2 /PS PE )ub and f3 = (PS /hb )ub (Eq. 6.7), and approximate vertical deformations

are v1 = P/kbzo , v2 = P/kbzo + (PS + P )u2b /(PE hb ), and v3 = PS u2b /(PE hb ) (Eq. 6.8). The

137
Require: Previous converged values fb , P , s, θ and v
1: Let x = fb , P, s, θ, vT
2: repeat
3: Solve for⎧fs (x3 ) and ∂fs /∂s with Algorithm ⎫ 1 or 2

⎪ x1 − fs (x3 ) + x2 x4 ⎪


⎪ ⎪


⎪ x h − P h x + x (x + hb 4 ⎪
x )⎪
⎨ 1 b E b 4 2 3 ⎬
4: f (x) = x2 − kbzo x5

⎪ ⎪


⎪ ub − x3 − hb x4 ⎪


⎪ ⎪

⎩ ⎭
ubz − x3 x4 − hb /2x4 − x5
2
⎡ ⎤
1 x4 −∂fs /∂s x2 0
⎢ ⎥
⎢hb (x3 + hb x4 ) x2 (x2 − PE )hb 0 ⎥
⎢ ⎥
5: J(x) = ⎢ ⎢ 0 1 0 0 −kbzo


⎢ ⎥
⎣0 0 −1 −hb 0 ⎦
0 0 −x4 −(x3 + hb x4 ) −1
6: −1
dx = −J (x) · f (x)
7: x = x + dx
8: until norm(dx) < tol
9: Update fb = x1 , P = x2 , s = x3 , θ = x4 , v = x5
10: F = fb , P T
11: det = hb (∂fs /∂s)(PE − P ) − P 2
⎡ (PE −P )hb P

det det 0
⎢ P ∂fs /∂s ⎥
12: feq = ⎣ det det 0 ⎦
1
0 0
⎡ ⎤ kbzo
1 θ
⎢ ⎥
13: T = ⎣hb (s + hb θ)⎦
0 1
14: fb = T T fbT
15: kb = fb−1
16: return: F , kb

Algorithm 3: Local Routine for Constant-Strength or Variable-Strength Model

138
negative work traced by the lateral force-deformation is thus:

1 1 P 2 u2b
W− = ub (f3 − f2 ) = (6C.1)
2 2 PE hb

while the positive work traced by the vertical force-deformation is

1 1 1 P 2 u2b PS P u2b
W+ = P (v2 − v1 ) + P v3 = + (6C.2)
2 2 2 PE hb PE hb

By taking PS  P , which is to the order of accuracy of the approximations, the second term

of Eq. 6C.2 is neglected and the negative and positive work cancel.

139
Appendix 6D: Notation

A cross-sectional area of bearing or isolation system

As cross-sectional area modified for undeforming steel layers: Ahb /tr

D bearing diameter

Ec instantaneous compression modulus of rubber-steel composite bear-

ing

ED energy dissipated in force-deformation loop

EIs effective bending stiffness of a bearing

F force vector in local bearing routine to implement nonlinear models

fb lateral force of bearing or isolation system

fb bearing flexibility matrix used in local routine to implement nonlin-

ear models

fs (s) force of shear spring in two-spring bearing model

fy yield force, for implementation of plasticity

fyo nominal value of yield force

G shear modulus of bearing or isolation system

Go nominal value of shear modulus

H hardening stiffness, for implementation of plasticity

hb total bearing height including steel and rubber layers

I conventional bending moment of inertia

kb instantaneous lateral stiffness of bearing, dependent on axial load

kb bearing stiffness matrix used in local routine to implement nonlinear

models

kbo nominal value of lateral stiffness: GAs /hb

kbz instantaneous vertical stiffness of bearing, dependent on lateral de-

formation

kbzo nominal value of vertical stiffness: Ec As /hb

140
Kc factor by which Ec is multiplied to include bulk compressibility of

rubber

keq stiffness matrix used in local bearing routine, derived from bearing

equilibrium equations

kI initial lateral stiffness of bearing or isolation system

Ks factor by which EIs is multiplied to include bulk compressibility of

rubber

M1 reactionary bearing bending moment

P bearing axial load

Pcr critical load or load that induces bearing buckling

PE Euler buckling load, neglects shear stiffness of the bearing: π 2 EIs /h2b

Po value of strength used to calibrate variable-strength model, typically

about 1/3 of the bearing static load

PS shear stiffness GAs

Pst static load, gravity load, or bearing design axial load

q back force, for implementation of plasticity

Q lateral characteristic strength of bearing or isolation system

Qo nominal value of characteristic strength

rb effective bending radius of bearing

s deformation of shear spring in two-spring bearing model

S bearing shape factor; D/4t for circular bearing

S2 second shape factor or bearing aspect ratio; D/tr for circular bearing

sp plastic shear deformation used in plasticity model

sy yield deformation of shear spring in two-spring bearing model

t thickness of a single bearing layer

Tb isolation period, corresponds to postyield stiffness of isolation system

Tp period of simple seismic pulse

tr total thickness of rubber in a single bearing

ub total lateral deformation of bearing: s + hb θ


141
ubz total vertical deformation of bearing: v + δbz

ug seismic ground displacement, with velocity u̇g and acceleration üg

v vertical deformation due to axial flexibility

γ̇ flow rate, based on plasticity theory

∆γ discrete flow step, for implementation of plasticity

δbz additional vertical deformation due to tilting of bearing layers under

shear deformation

ζeq equivalent viscous damping coefficient

θ rotation of bearing in two-spring bearing model

κ bulk modulus of compressibility

Φ yield function, for implementation of plasticity

142
7: ESTIMATING ISOLATOR DEFORMATIONS AND FORCES
CONSIDERING LATERAL-ROCKING RESPONSE

7.1 Introduction

To estimate the peak axial forces in individual isolators, which are needed for design, it is

necessary to consider rocking or overturning in studying the dynamics of base-isolated struc-

tures. The fundamental coupled lateral-rocking-vertical response of a rigid structure on linear

isolation bearings was explored [51], but the study was limited to vertical ground excitation,

where the vertical degree-of-freedom was coupled to the others through a small eccentricity.

One case study of an isolated building predicted (unrestrained) uplift off several of the bearings

[52], while another determined the uplift of 1994 UBC-designed 6 and 20-story buildings due

to various near-source ground motions [53]. The latter study concluded that resisting uplift

should be a major design consideration; and in a follow-up to that effect, the same 6-story

building was modified to meet the upgraded standards of the 1997 UBC, the bearings were per-

mitted to resist tension, and tensile demands were successfully limited by special measures in

the building frame design [54]. Although making valuable contributions, none of these studies

comprehensively assesses the effects of overturning in base-isolated buildings.

Recall that axial-load effects, implemented in the nonlinear constant-strength and variable-

strength models for lead-rubber bearings, were shown in the previous chapter to influence the

behavior of the isolation bearings. What is the significance of these effects on the overall system

response and pertinent design quantities, such as lateral deformation and axial forces in the

isolators? One previous study incorporated similar axial-load effects into dynamic analysis of

the isolated building [55], showing that neglecting them would increase the error in the lateral

isolator deformation as the factor of safety against buckling was reduced. However, the lateral

behavior was assumed to be linear, and vertical flexibility was neglected such that axial forces

could not be determined accurately, deficiencies that have been remedied by our nonlinear

models.

Given this background information, the objectives of this chapter are (1) to understand

143
d/2 d/2

z, uz
h/2

CM
θx y, uy

h/2
ith bearing
(yi , −h/2)

üg (t)

Figure 7.1: Rigid block rocking on isolation bearings

the response of the isolated building using the improved constant-strength and variable-strength

models for lead-rubber bearings, (2) to estimate the peak lateral deformation and maximum

and minimum axial forces in the isolators for use in preliminary design, via equations based on

response history analysis to an ensemble of ground excitations, and (3) to identify the conditions

when axial-load effects may be neglected in determining the response.

7.2 Governing Equations and Parameters

7.2.1 System Considered

Consider a rigid block with in-plane dimensions d (width) and h (height) supported on isolation

bearings, and subjected to lateral ground excitation üg (t) (Fig. 7.1). Planar motion of the block

is described by the lateral and vertical displacements uy and uz at the CM, and rocking angle

θx about the x-axis. With mass m uniformly distributed over the block, the rocking radius of

gyration is rx = (d2 + h2 )/12. The isolators are assumed to be evenly-spaced across the plan

in both directions. Each isolation bearing is located at (yi , −h/2) from the CM in the y- and

z-directions, respectively.

144
7.2.2 Bearing Models

The bearing models used in this chapter are derived from the equilibrium and kinematic equa-

tions developed previously for the two-spring bearing model – comprised of a shear spring and

a rotational spring (Fig. 6.2). Repeated here for the ith bearing, the equilibrium equations are

fbi − fsi (si ) + Pi θi = 0 (7.1a)

fbi hb − PE hb θi + Pi (si + hb θi ) = 0 (7.1b)

Pi − kbzi vi = 0 (7.1c)

representing the balance of lateral forces, moments, and vertical forces, respectively. The total

lateral and axial forces fbi and Pi are related to (a) the force fsi and deformation si of the shear

spring, and (b) the stiffness PE hb – the Euler buckling load times the bearing height – and

rotation θi through the rotational spring. Not included in the two-spring model (Fig. 6.2), an

additional spring with stiffness kbzi represents the axial flexibility of the bearing that permits

axial deformation vi .

Similarly, the kinematic equations are

ubi = si + hb θi (7.2a)

hb 2
ubzi = vi + si θi + θ (7.2b)
2 i
In the context of the two-spring model (Fig. 6.2), the total lateral deformation ubi is the sum

of the shear deformation si and the rotation θi multiplied by the bearing height (Eq. 7.2a),

while the total vertical deformation ubzi includes the axial deformation vi and an additional

vertical displacement resulting from the laterally deformed configuration (Eq. 7.2b). In contrast

to the global sign convention, the vertical deformation ubzi and force Pi are positive for bearing

compression.

Three models are derived from Eqs. 7.1 and 7.2:

1. coupled linear model: the shear spring is linear (fsi = kbi si in Eq. 7.1a)

145
2. coupled nonlinear constant-strength model: the shear spring follows a bilinear force-

deformation relation with initial stiffness kI , nominal post-yield stiffness kbi , and nominal

yield strength Qi

3. coupled nonlinear variable-strength model: like the constant-strength model, but the yield

strength varies in time according to:


 
Qi (t) = Qi 1 − e(−Pi /Po ) (7.3)

where Po , typically about 1/3 of the bearing static load, can be determined from charac-

teristic tests of prototype bearings

The latter two models will be abbreviated hereafter as the constant-strength and variable-

strength models. Each above model includes various axial-load effects, which cause local cou-

pling of the lateral and axial forces and deformations in individual bearings (Eq. 7.1). These

effects include variation of the total lateral stiffness – whether linear or nonlinear – with ax-

ial load, and variation of the vertical stiffness with lateral deformation. The variable-strength

model includes an additional variation of the yield strength with axial load, which was observed

to be present in lead-rubber bearings. For further details, consult Chapter 6.

When axial-load effects are neglected, Eq. 7.1 simplifies to the following:

fbi = kbi ubi + Qi zi (7.4a)

Pi = kbzi ubzi (7.4b)

where the lateral force-deformation relation (Eq. 7.4a) – defined previously (Eq. 3.2) and now

applied to individual bearings – and the linear vertical force-deformation relation (Eq. 7.4b)

are now uncoupled. Two versions of this uncoupled model:

• uncoupled linear model: the total lateral force-deformation relation is linear with stiffness

kbi , i.e., Qi = 0 in Eq. 7.4a

• uncoupled nonlinear model: the total lateral force-deformation relation is bilinear with

initial stiffness kI , post-yield stiffness kbi , and yield strength Qi , identical to that of the

shear spring in the constant-strength model.

146
are suitable for comparison with the three models defined above. The nominal lateral (kbi ) and

vertical (kbzi ) stiffnesses and strength Qi identified for the various bearing models are presently

assumed to be identical for each bearing in the isolation system.

Viscous damping in the vertical direction is included by an additional axial force com-

ponent for each bearing, fDi = cbzi u̇bzi where cbzi is the damping constant and u̇bzi the axial

bearing velocity. The damping coefficients cbzi are also assumed to be identical for each bearing.

7.2.3 Equations of Motion

When axial-load effects are included, the equations of motion for the system, subjected to

lateral earthquake acceleration üg and gravity loading g are


müy + fbi (ubi , ubzi ) = −müg (7.5a)
i
 h 
mrx2 θ̈x − yi cbzi u̇bzi + fbi (ubi , ubzi ) − yi Pi (ubi , ubzi ) = 0 (7.5b)
2
i i i
 
müz − cbzi u̇bzi − Pi (ubi , ubzi ) = mg (7.5c)
i i

The lateral and axial forces fbi and Pi in the bearing are implicit functions of its deformations

ubi and ubzi (Eqs. 7.1 and 7.2). (The dependence of fbi and Pi on ubi and ubzi is hereafter

implied.) In turn, the bearing deformations are related to the CM displacements (uy , uz and

θx ) by:

h
ubi = uy + θx (7.6a)
2
ubzi = −uz − yi θx (7.6b)

Substituting the time derivative of Eq. 7.6b for u̇bzi into Eq. 7.5, and normalizing Eq. 7.5b

by rx leads to
⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪ ⎪

⎪ y ⎪⎪ ⎪
⎪ · ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪


ü ⎪
⎬ ⎪

0 u̇ y ⎪
⎬ ⎪ ⎨ i fbi ⎪
⎬ ⎪ üg ⎪
⎨ ⎪

 yi2  
i cbzi rx2 · rx θ̇x ⎪
yi = −m 0
2rx fbi −
m rx θ̈x + + h (7.7)
⎪ ⎪ ⎪ ⎪ i i rx Pi ⎪ ⎪ ⎪

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

⎪ ⎪
⎩ ü ⎭ ⎩ ⎪  ⎪
⎭ ⎩ ⎪  ⎪
⎭ ⎩−g⎪
⎪ ⎭
z i cbzi · u̇z − i Pi

147
These equations are coupled in the three CM displacements (uy , θx and uz ), because the force-

deformation equations (Eq. 7.1) for individual bearings are coupled in the lateral and vertical

directions, caused by axial-load effects.

Special Case: Uncoupled Bearing Models

Substituting Eq. 7.4 into Eq. 7.7 gives the equations of motion:
⎧ ⎫ ⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪

⎪ y ⎪
⎪ ü ⎪
⎪ ⎪

⎪ 0 · u̇y ⎪

⎪ ⎢ 1 h
0 ⎥⎪

⎪ u y ⎪⎪
⎪ ⎪

⎪ i i i ⎪
γ z ⎪



⎪ üg ⎪



⎨ ⎬ ⎨ ⎬ ⎢
2rx
⎥ ⎨ ⎬ ⎨ ⎬ ⎨ ⎬
2⎢ h ⎥ 
rx θ̈x ⎪ + 2ω b ζθx Ωθx · rx θ̇x + ω 2
b ⎢ 2rx Ωθx 0 ⎥ ⎪rx θx ⎪ + µg h
i γi zi ⎪
= − 0⎪

⎪ ⎪ ⎪
⎪ ⎪
⎪ ⎣ ⎦⎪ ⎪ ⎪
⎪ 2rx
⎪ ⎪
⎪ ⎪

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎩ üz ⎭ ⎩ ζz Ωz · u̇z ⎭ 0 0 Ω2z ⎩ uz ⎭ ⎩ 0 ⎭ ⎩−g⎪ ⎭
(7.8)

Due to uncoupling at the bearing level, the vertical displacement uz is no longer coupled with

the lateral and rocking displacements uy and θx . The strength coefficient µ:



Q i Qi
µ= = (7.9)
w w

was factored out of the nonlinear terms in Eq. 7.8 by introducing γi = Qi /Q.

Equation 7.8 has been written in terms of several parameters of a linear system (µ = 0):

(1) vertical (to lateral) and rocking (to lateral) frequency ratios:

ωbz ωbθx
Ωz = Ωθx = (7.10)
ωb ωb

where the lateral, vertical and rocking frequencies are


 1/2  1/2    1/2
i kbi i kbzi h2 kbi
i
2
i yi kbzi
ωb = ωbz = ωbθx = + (7.11)
m m 4rx2 m rx2 m

(2) vertical and rocking damping ratios:

cbz cbθx
ζz = ζθx = (7.12)
2mωbz 2mrx2 ωbθx

where the vertical and rocking damping constants are

 
cbz = cbzi cbθx = yi2 cbzi (7.13)
i i

148
7.2.4 System Parameters

The equations of motion (Eq. 7.7) combined with those for the bearing model (Eqs. 7.1 and 7.2)

suggest that the response of the system depends on an extensive list of system and bearing prop-

erties. However, only a short list of parameters, which have been identified by trial and error,

significantly alter the response of the system when varied over their applicable ranges. These

parameters are listed below, and their ranges of variation considered are noted in parentheses:

1. isolation period Tb (from frequency ωb ) (1.5 to 5 seconds)

2. strength coefficient µ (see Sec. 7.4.1)

3. vertical (to lateral) frequency ratio Ωz (20 to 35)

4. rocking (to lateral) frequency ratio Ωθx (5 to 53)

Note that three of these parameters nearly define a linear system (Eq. 7.8 with µ = 0) and the

strength coefficient is relevant for nonlinear systems. As will be shown, this list of parameters

is both relevant and comprehensive when using bearing models that include axial-load effects

(Eq. 7.7).

The isolation period and strength coefficient were defined in Chap. 3. The vertical fre-

quency ratio Ωz (Eq. 7.10) is a function of the lateral and vertical stiffnesses kbi and kbzi

(Eq. 7.11), which are related by the ratio of the compression modulus to shear modulus Ec /G

(Eqs. 6.1 and 6.2). This ratio is a function of the bearing shape factor S (Eqs. 6.11 and 6.15),

where S = D/4t for a bearing with diameter D and thickness t of a single layer. If the shape

factor is identical for all bearings:


 1  1   12
kbzi 2 Ec 2 6S 2
Ωz = i
= = (7.14)
i kbi G 1 + Kc S 2

An applicable range of 10 to 30 for S and Kc = 0.004 (Appendix 7A) led to the range of Ωz

listed above (20 to 35). A typical shape factor of 20 leads to a vertical frequency ratio of 30.4.

Large shape factors (above 20) are commonly used today; such bearings are rather insensitive

to axial loads, but require large diameters to accommodate the design displacement, making

149
it difficult to limit the stiffness for adequate isolation. Bearings with small shape factors such

as 10 (Ωz ≈ 20) may provide some benefit of vertical isolation, but are more sensitive to axial

loads and possibly prone to stability problems.

Substituting ωb and ωbθx (Eq. 7.11) into Eq. 7.10 gives the rocking frequency ratio as
  2  12
1 h2 y kbzi
Ωθx = + 2 i i
(7.15)
4 rx2 rx i kbi

explicitly a function of many factors such as the bearing locations and stiffnesses, building

height, etc. However, it can be simplified when the bearings are identical and spaced evenly on

a grid over the plan, as assumed here:


  12
1 h2 1 d2 (nby + 1) 2
Ωθx = + Ω (7.16)
4 rx2 12 rx2 (nby − 1) z
 2
where i yi has been expressed in terms of nby , the number of bearings spanning the y-direction

(in-plane).

For a fixed total lateral stiffness, the rocking frequency ratio Ωθx decreases as the number

of bearings increases (Fig. 7.2a), thus tending toward a more uniform distribution of stiffness

across the plan and less resistance to rocking. As expected, the rocking frequency ratio changes

most rapidly with number of bearings when this number is small, but seems to reach a plateau

for systems with more than 10 bearings across the plan (Fig. 7.2a). This suggests rocking

analysis of a simple system with only two bearings might not be realistic of more practical

systems with many bearings.

Not surprisingly, the resistance to rocking decreases (Ωθx decreases) for more slender

buildings, as shown in Fig. 7.2a for different building slenderness – or height-to-width – ratios

h/d. The slenderness ratio affects the rocking frequency ratio through the values of h/rx and

d/rx (Eq. 7.16). The applicability of the building model used here to slenderness ratios h/d > 2

may be questionable (see Sec. 7.2.5); however, the application of base-isolation to such buildings

is considered to be rare.

Finally, the rocking frequency ratio Ωθx increases as the vertical frequency ratio Ωz in-

creases (Fig. 7.2b). In the limit as h/d → 0 and nby → ∞, Ωθx → Ωz (Eq. 7.16); thus the

150
(a) (b)
50 40
h/d=0.5
1
Rocking frequency ratio Ωθx

40 2
30 3
4
30
20
20

10
10

0 0
0 1 2 0 10 20 30 40
10 10 10
Number of bearings nby Vertical frequency ratio Ωz

Figure 7.2: Variation of rocking frequency ratio Ωθx with (a) number of bearings across the
plan in-plane (Ωz = 30.4 for S = 20) and (b) vertical frequency ratio Ωz (nby = 10).

rocking frequency ratio is close to the vertical frequency ratio for a squat building with suffi-

ciently large number of bearings, as reflected in Fig. 7.2b for h/d = 0.5. For larger slenderness

ratios, the rocking frequency takes an intermediate value between the lateral frequency and the

vertical frequency.

In summary, over the range of nby (≥ 2), building slenderness ratio h/d (0.5 to 4), and

Ωz (20 to 35), the rocking frequency ratio Ωθx varies from about 5 to 53, where the lower range

reflects slender buildings with many bearings distributed across the plan while the upper range

reflects squat buildings with fewer bearings. For details of how all the parameters in Eq. 7.7

were selected for response history analysis (RHA), consult Appendix 7A.

7.2.5 Rigid Structure Approximation

The validity of the rigid structure approximation comes into particular question for slender,

longer period structures. Assuming the structure to be rigid is accurate only if the fundamental

fixed-base period Ts of the structure is much shorter than the isolation period Tb [21, Chap. 20].

As a guideline, the International Building Code [1] suggests that Tb > 3Ts (required for design

based solely on static analysis), and for this range the rigid structure approximation is certainly

valid. This implies that for the range of 1.5 to 5 seconds considered for Tb , the upper limit for

Ts is 0.5 to 1.66 seconds.

151
An examination of the natural vibration periods of existing buildings measured from their

recorded response during earthquakes (Appendix 7B) indicates that most slender buildings

(h/d > 2) have periods longer than 1.66 seconds. However, these buildings were all fixed-base,

while isolated structures are required by code [1] to be stronger and stiffer. This results in

superstructure periods that are shorter than comparable fixed-base structures, and more likely

to fall within the limit of Ts = 1.66 seconds. The simplified procedure to estimate the bearing

deformation and forces presented later for preliminary design application assumes the structure

to be rigid. However, structural flexibility should be considered in any nonlinear RHA to

evaluate the preliminary design and modify it as necessary.

7.3 Response Histories for Different Models

Response histories are compared for the same base-isolated structure with 5 different bearing

models: uncoupled linear (Fig. 7.3), coupled linear (Fig. 7.4), uncoupled nonlinear (Fig. 7.5),

constant-strength (Fig. 7.6), and variable-strength (Fig. 7.7). The sample structure has slen-

derness h/d = 2, supported by two bearings at the edges. Considered for each bearing model

are (a) harmonic excitation with frequency ω = 4π (chosen to be distinct from the resonant

frequency ωb = π or Tb = 2 sec) and acceleration amplitude ωvo where vo = 40 cm/s; and (b)

the strong component of LMSR Record No. 20, with peak ground velocity u̇go = 61.5 cm/s.

For this demonstration only, viscous damping is omitted from the analyses.

To summarize, the response histories shown are the lateral (uy ), rocking (rx θx ) and

vertical (uz /ust ) displacements of the CM of the block, and the axial deformation (ubz1 /ust )

and force (P1 /Pst ) in the left bearing (representative of a typical exterior bearing), where

all vertical/axial responses have been scaled relative to their static values. Also shown are

power spectra – appropriately scaled complex moduli of the discrete Fourier transforms – of the

response components uy and rx θx , and the estimated variation of the lateral stiffness kb (t)/kb

(Eq. 6.7, summed over both bearings) and vertical stiffness kbz (t)/kbz (Eq. 6.17) relative to

their nominal values.

The most significant outcome of including axial-load effects (coupled linear model, Fig. 7.4)

relative to the uncoupled linear model (Fig. 7.3) is to greatly increase the bearing axial defor-

152
mation ubz1 , especially in compression. For harmonic excitation, this deformation increase

is a result of inciting vertical motion uz at the CM; for the recorded excitation, the rocking

displacement rx θx increases in addition to vertical motion. However, the corresponding left

bearing axial force P1 increases by only a small amount (recorded excitation only), and may

not increase if damping were to be included.

The system with the uncoupled linear bearing model – i.e., linear behavior in both the

lateral and vertical directions with no interaction between them – is well understood [51]. The

frequencies and mode shapes of the lateral and rocking degrees of freedom: ω1 = 3.13, ω2 = 65.0

and φ1 = 0.996, −0.004T , φ2 = 0.004, 0.996T , are nearly uncoupled. With tall spikes near

ω1 and barely visible or nonexistent spikes near ω2 , the power spectra (Fig. 7.3a,b) indicate

that the first mode, which is primarily lateral translation (uy ), dominates. (The second spike

for harmonic excitation reflects the excitation frequency.)

For harmonic excitation, the coupled linear model (Fig. 7.4a) causes the power spectra

and the lateral (uy ) and rocking (rx θx ) displacements to change only negligibly compared to

the uncoupled linear model. Thus, despite the influence of axial load on the lateral stiffness

and the influence of lateral deformation on the vertical stiffness – both small, as reflected in

the plots for kb and kbz – this system vibrates in approximately the same modes as when axial-

load effects are neglected (uncoupled linear model). The variation in vertical stiffness, however,

causes vertical motion of the CM, which contributes to an increase in the left bearing axial

deformation ubz1 but not force P1 .

For the recorded excitation, the power spectra for the coupled linear model (Fig. 7.4b)

reflects a second spike in the rocking displacement at a much lower frequency than the second

mode frequency, an indication that the axial-load effects are significant enough to cause the

“modes” of the system to shift. While the first mode still dominates, it now contributes more

to the rocking component, and combined with participation of the second mode, the total

amplitude of rocking rx θx grows significantly. The modal shift seems to occur because this

strong ground motion simply shakes the system harder, causing large variation in the lateral

stiffness kb due to much larger axial forces, and large variation in the vertical stiffness kbz due

153
(a)
1
5 uy
uy (cm)

0.8 rx θ x
0

Power spectra
−5 0.6
rx θx (cm)

0.025
0.4
0
0.2
−0.025
0
1.5 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1
1

kb (t)/kb
1.3
ubz1 /ust

0.66
1
0.33
0.7
1
kbz (t)/kbz
1.3
P1 /Pst

0.66
1
0.33
0.7
0 5 10 15 20 0 5 10 15 20
Time (sec) (b) Time (sec)
1
100 uy
uy (cm)

0.8 rx θ x
0
Power spectra

−100 0.6
rx θx (cm)

6
0.4
0
0.2
−6

30 0
0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1
1
kb (t)/kb

60
ubz1 /ust

0.66
30
0.33
0
1
kbz (t)/kbz

4
P1 /Pst

0.66
0
0.33
−4
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (sec) Time (sec)

Figure 7.3: Various response histories of the isolated system with rocking, and power spectra
for the lateral and rocking displacements uy and rx θx ; due to lateral excitations: (a) harmonic
with ω/ωb = 4 and (b) LMSR Record No. 20. The system uses the uncoupled linear bearing
model, and has properties Tb = 2 seconds, S = 15, S2 = 3, h/d = 2, nby = 2, and no viscous
damping.

154
(a)
1
5 uy
uy (cm)

0.8 rx θ x
0

Power spectra
−5 0.6
rx θx (cm)

0.025
0.4
0
0.2
−0.025
0
1.5 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1
1

kb (t)/kb
1.3
ubz1 /ust

0.66
1
0.33
0.7
1
kbz (t)/kbz

1.3
P1 /Pst

0.66
1
0.33
0.7
0 5 10 15 20 0 5 10 15 20
Time (sec) Time (sec)
(b)
1
100 uy
uy (cm)

0.8 rx θ x
0
Power spectra

−100 0.6
rx θx (cm)

6
0.4
0
0.2
−6

30 0
0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1
1
kb (t)/kb

60
ubz1 /ust

0.66
30
0.33
0
1
kbz (t)/kbz

4
P1 /Pst

0.66
0
0.33
−4
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (sec) Time (sec)

Figure 7.4: Various response histories of the isolated system with rocking, and power spectra
for the lateral and rocking displacements uy and rx θx ; due to lateral excitations: (a) harmonic
with ω/ωb = 4 and (b) LMSR Record No. 20. The system uses the coupled linear bearing
model, and has properties Tb = 2 seconds, S = 15, S2 = 3, h/d = 2, nby = 2, and no viscous
damping.

155
to a much larger lateral deformation. Also related to the vertical motion at the CM, the left

bearing axial deformation in compression increases to over 60 times its static value, this time

causing an increase in axial force as well.

When comparing the three nonlinear models, the consequences of including axial-load

effects – constant-strength (Fig. 7.6) or variable-strength (Fig. 7.7) models – compared to the

nonlinear uncoupled model (Fig. 7.5) are more subtle than for the linear bearing models. Like

the linear models, an increase in rocking displacement rx θx relative to the lateral displacement

uy is observed, which, when combined with vertical motion at the center of mass, increases

the left bearing axial deformation ubz1 , especially in compression. However, the deformation

increase is much smaller than when the bearings were linear, and the axial forces do not appear

to be affected.

The response of the system with the uncoupled nonlinear bearing model due to harmonic

excitation (Fig. 7.5a) is relatively straightforward, still vibrating in predominantly two modes.

In lateral motion, it reaches a steady state of vibration at the excitation frequency after the

transient excitation associated with the fundamental frequency has damped out. The second

mode influence is stronger here than for the uncoupled linear bearing model, with a higher

frequency reflected in the rocking component rx θx history as well as the power spectra. The

response of the system with the constant-strength model (Fig. 7.6a), is very similar to that with

the uncoupled nonlinear bearing model. Because the lateral deformation is small, the variation

in vertical stiffness kbz is modest, causing only a small increase in the rocking displacement rx θx

and a small vertical motion uz at the CM. Slight increases in the left bearing axial deformation

ubz1 and force P1 appear to be proportional to the increase in rocking displacement, as the

vertical motion uz is comparatively small.

In response to recorded excitation, even the simplest nonlinear model – uncoupled non-

linear model (Fig. 7.5b) – displays much more complicated behavior. The power spectra shows

that instead of two or three frequencies, a range of frequencies are excited, i.e., the response

frequency shifts as the system is excited at different intensities. Consistent with earlier obser-

vation, the rocking component contains higher frequencies reflecting higher mode contribution.

156
The vertical softening or decrease in vertical stiffness kbz with axial-load effects (constant-

strength model, Fig. 7.6b) is more significant here, and similarities with the linear system

response to axial-load effects (Fig. 7.4b) are observed. This includes increased rocking rx θx ,

significant vertical motion uz of the CM, and a marked increase in the peak axial deforma-

tion ubz1 especially on the compression side, with little increase in the corresponding force P1

(Fig. 7.6b), all relative to the uncoupled nonlinear model (Fig. 7.5b).

For the final example – the response of the system using the variable-strength model,

kbz (t) has been replaced by the strength variation Q(t) summed over both bearings and divided

by the nominal strength Q (Fig. 7.7). Since an occurrence of tension that induces a complete

loss of bearing strength is balanced by compression on the other side of the building, Q/Qo

is not likely to fall below 0.5, and indeed does not in this example (Fig. 7.7). Relative to the

constant-strength model (Fig. 7.6), inclusion of the strength variation has a minimal effect on

the response. It causes an increase in the lateral displacement uy due to diminished energy

dissipation, and an increase in the peak axial deformation ubz1 but a decrease in the peak axial

force P1 . These effects are barely visible in Fig. 7.7.

For further understanding, the frequency of the harmonic excitation is varied over a wide

range. The peak lateral uyo and rocking displacement rθxo , and the maximum and minimum

bearing axial deformations ubz± /ust and forces P± /Pst are plotted against the excitation fre-

quency for the two linear (Fig. 7.8) and three nonlinear (Fig. 7.10) bearing models. By extending

the duration of the excitation to two minutes, any resonances that may exist will be apparent.

The duration is irrelevant at non-resonance where the system quickly reaches steady state.

With the uncoupled linear bearing model, the system has a resonance at its first mode

frequency, approximately equal to the isolation frequency ωb (ω/ωb = 1, Fig. 7.8). Including

axial-load effects (coupled linear bearing model) appears to shift the resonance (more dominant

in rocking rx θx ) to a lower frequency (ω/ωb = 0.8), unsurprising since both the lateral and

vertical stiffnesses of the bearings tend to decrease. However, it will be shown that with axial-

load effects, the system does not reach resonance, but instead arrives at a steady state response.

To begin, the displacements uy and rx θx of this system with ω/ωb = 0.8 (Fig. 7.9a and

157
(a)
1
6 uy
uy (cm)

0.8 rx θ x
0

Power spectra
−6 0.6
rx θx (cm)

0.15
0.4
0
−0.15 0.2

0
1.2 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

kb (t)/kb
2
ubz1 /ust

1
0.9
0
1
kbz (t)/kbz

2
P1 /Pst

0.66
1
0.33
0
0 5 10 15 20 0 5 10 15 20
Time (sec) (b) Time (sec)
1
20 uy
uy (cm)

0.8 rx θ x
0
Power spectra

−20 0.6
rx θx (cm)

0.4
0.4
0
0.2
−0.4
0
3 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

1
kb (t)/kb

5
ubz1 /ust

2.5
0.9
0
1
kbz (t)/kbz

2
P1 /Pst

0.66
1
0.33
0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (sec) Time (sec)

Figure 7.5: Various response histories of the isolated system with rocking, and power spectra
for the lateral and rocking displacements uy and rx θx ; due to lateral excitations: (a) harmonic
with ω/ωb = 4 and (b) LMSR Record No. 20. The system uses the uncoupled nonlinear bearing
model, and has properties Tb = 2 seconds, µ = 0.1, S = 15, S2 = 3, h/d = 2, nby = 2, and no
viscous damping.
158
(a)
1
6 uy
uy (cm)

0.8 rx θ x
0

Power spectra
−6 0.6
rx θx (cm)

0.15
0.4
0
−0.15 0.2

0
1.2 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

kb (t)/kb
2
ubz1 /ust

1
0.9
0
1
kbz (t)/kbz

2
P1 /Pst

0.66
1
0.33
0
0 5 10 15 20 0 5 10 15 20
Time (sec) (b) Time (sec)
1
20 uy
uy (cm)

0.8 rx θ x
0
Power spectra

−20 0.6
rx θx (cm)

0.4
0.4
0
0.2
−0.4
0
3 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

1
kb (t)/kb

5
ubz1 /ust

2.5
0.9
0
1
kbz (t)/kbz

2
P1 /Pst

0.66
1
0.33
0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (sec) Time (sec)

Figure 7.6: Various response histories of the isolated system with rocking, and power spectra
for the lateral and rocking displacements uy and rx θx ; due to lateral excitations: (a) harmonic
with ω/ωb = 4 and (b) LMSR Record No. 20. The system uses the nonlinear constant-strength
bearing model, and has properties Tb = 2 seconds, µ = 0.1, S = 15, S2 = 3, h/d = 2, nby = 2,
and no viscous damping.

159
(a)
1
6 uy
uy (cm)

0.8 rx θ x
0

Power spectra
−6 0.6
rx θx (cm)

0.15
0.4
0
−0.15 0.2

0
1.2 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

kb (t)/kb
2
ubz1 /ust

1
0.9
0

2 1
Q(t)/Q
P1 /Pst

1
0.75
0
0 5 10 15 20 0 5 10 15 20
Time (sec) (b) Time (sec)
1
20 uy
uy (cm)

0.8 rx θ x
0
Power spectra

−20 0.6
rx θx (cm)

0.4
0.4
0
0.2
−0.4
0
3 0 20 40 60 80 100
uz /ust

Frequency ω (rad/s)
1

1
kb (t)/kb

5
ubz1 /ust

2.5
0.9
0
2 1
Q(t)/Q
P1 /Pst

1
0.75
0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time (sec) Time (sec)

Figure 7.7: Various response histories of the isolated system with rocking, and power spectra
for the lateral and rocking displacements uy and rx θx ; due to lateral excitations: (a) harmonic
with ω/ωb = 4 and (b) LMSR Record No. 20. The system uses the nonlinear variable-strength
bearing model, and has properties Tb = 2 seconds, µ = 0.1, S = 15, S2 = 3, h/d = 2, nby = 2,
Pst /Po = 3, and no viscous damping.

160
3000 600
∞ Uncoupled linear
2500 Coupled linear
400

ubz+ /ust , ubz− /ust


2000
uyo (cm)

200
1500
0
1000

500 −200

0 −400
−1 0 1 2 −1 0 1 2
10 10 10 10 10 10 10 10

80 60

40
60
P+ /Pst , P− /Pst
20
rθxo (cm)

40 0

−20
20
−40

0 −60
−1 0 1 2 −1 0 1 2
10 10 10 10 10 10 10 10
Excitation to isolation frequency (ω/ωb ) Excitation to isolation frequency (ω/ωb )

Figure 7.8: Various peak responses of system with linear bearing models (uncoupled and
coupled) due to harmonic excitation frequency sweep. System properties are Tb = 2 seconds,
S = 15, S2 = 3, h/d = 2, nby = 2, and no viscous damping.

b) do not grow without bound as would a resonant response, but instead appear to oscillate

between two states. Multiple spikes in the power spectrum of uy (Fig. 7.9c) indicate that

distinct frequencies near the so-called resonance frequency are being excited. To understand

this behavior, the displacement history uy has been split into subintervals at points where its

frequency appears to undergo small changes (Fig. 7.9a). The power spectra of subintervals 1

and 3, and subintervals 2 and 4, respectively, are shown in Fig. 7.9d and e, which reflect distinct

frequencies (albeit only slightly different) for each. Thus, power spectra of the displacement

broken into appropriate subintervals show that a single frequency is excited in each subinterval.

In this sense, the steady state response of this system is defined by the continual transition

between two states defined by the distinct frequencies. Our present understanding is that this

161
(a) (b)
500 100
int 1 int 2

50

rθx (cm)
uy (cm)

0 0

−50
int 3 int 4

−500 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (s) Time (s)

(c) (d) (e)


1 1 1
int 1 int 2
int 3 int 4
0.8 0.8 0.8
ωp = 2.48
Power spectra

ωp = 2.22
0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Frequency ω (rad/s) Frequency ω (rad/s) Frequency ω (rad/s)

Figure 7.9: (a,b) Response histories of lateral displacement uy and rocking rx θx of system with
coupled linear model for isolation bearings to harmonic excitation with frequency ω = 0.8ωb ;
(c) power spectrum of displacement uy , and (d,e) power spectra of displacement uy broken into
subintervals 1 (11.5 - 28.4 sec) and 3 (59 - 76.1 sec), and 2 (31.1 - 56.5 sec) and 4 (78.5 - 103.9
sec), with frequency ωp of peak indicated.

shifting between states is caused by changes in stiffness with response amplitude induced by

axial-load effects. The most important point is that, although composed entirely of linear

materials, this system does not have a resonant frequency.

The peak (over all frequencies) rocking displacement rθxo is two orders of magnitude

smaller for the uncoupled bearing model but the same order of magnitude for the coupled

bearing model as the lateral displacement uyo (Fig. 7.8), indicating that axial-load effects greatly

increase the tendency of the system to rock. While the corresponding peaks of axial deformation

(maximum ubz+ and ubz− ) are much larger for the coupled bearing model, the corresponding

axial forces are larger for the uncoupled bearing model. This implies that the peak forces in

162
100 60
Uncoupled nonlinear
Constant-strength
80 Variable-strength 40

ubz+ /ust , ubz− /ust


uyo (cm)

60 20

40 0

20 −20

0 −40
−1 0 1 2 −1 0 1 2
10 10 10 10 10 10 10 10

6 20

5
10
P+ /Pst , P− /Pst

4
rθxo (cm)

3 0

2
−10
1

0 −20
−1 0 1 2 −1 0 1 2
10 10 10 10 10 10 10 10
Excitation to isolation frequency (ω/ωb ) Excitation to isolation frequency (ω/ωb )

Figure 7.10: Various peak responses of system with nonlinear bearing models (uncoupled
nonlinear, constant-strength and variable-strength) due to harmonic excitation frequency sweep.
System properties are Tb = 2 seconds, µ = 0.1, S = 15, S2 = 3, h/d = 2, nby = 2, Pst /Po = 3,
and no viscous damping.

the bearings may be lower than those predicted by a typical uncoupled linear model, but this

was not true for the earlier example (Figs. 7.3 and 7.4). Notice that for the uncoupled linear

model, the maximum and minimum axial deformations are symmetric about ust (ubz /ust = 1),

likewise with axial forces. But for coupled bearings, vertical displacement of the CM disturbs

the symmetry, and peak deformations and forces in compression tend to be comparatively larger

than those in tension.

The system with the uncoupled nonlinear bearing model again has a resonance very

close to the isolation frequency ωb (Fig. 7.10). This resonance occurs only if the excitation is

sufficiently strong to overcome the hysteretic energy dissipation, which becomes negligible as the

response amplitude grows large. Again, it can be shown that use of either the constant-strength

163
or variable-strength models eliminates this resonance. However, with the uncoupled nonlinear

bearing model, the resonance builds up so slowly that its component responses (even lateral

displacement uyo ) are exceeded by the corresponding non-resonant responses of the system

with either the constant-strength or variable-strength model. Consequently, the possibility

exists that the typical uncoupled nonlinear bearing model could significantly underestimate the

peak response of the system. The various spikes in response that occur at frequencies between

1 and 100*ωb are difficult to explain.

7.4 Normalization and Dispersion

7.4.1 Normalized Equations

The normalization procedure of Chapters 3-5 is extended to the current system, coupled in its

lateral, rocking and vertical DOFs. The normalized equations of motion are:
⎧ ⎫ ⎡ ⎤⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪

⎪ ¨ y ⎪ ⎪ ⎪ ˙ ⎪ ⎪ ⎪ ⎪ ¨g ⎪



ū ⎪
⎬ ⎢
0 0 0 ⎥⎪⎪

ū y ⎪ ⎪




f
i bi ⎪


⎪ ū
⎨ ⎪
⎪ ⎪

⎢  ⎥ ω 2
  ωb2
¨ + ⎢ y
cbzi i
2
⎥ ˙ + b h
− yi = − 0 ⎪ (7.17)
⎪ r θ̄ ⎪ ⎢ 0 0 ⎥ r θ̄ f P
⎦⎪ ⎪ Q ⎪ ⎪ ⎪
x x i m rx 2 x x i 2rx bi i rx i
⎪ ⎪ ⎣ ⎪ ⎪ ⎪ ⎪ ηωd ⎪ ⎪

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪
⎩ ū ¨z ⎭ 0 0 c bz ⎩ ū˙ ⎭ ⎩ − i Pi ⎭ ⎩−ḡ⎪⎭
m z

They were obtained by first dividing Eq. 7.7 by u∗y = Q/kb (Eq. 3.6) such that all degrees of

freedom are normalized – e.g. ūy = uy /u∗y – and then substituting the normalized strength

η = a∗y /ωd u̇go (Eq. 3.8) into the right-hand side. This normalizes the ground acceleration by
¨g = üg /u̇go ), such that the corresponding ground velocity varies between
its peak velocity (ū

-1 and +1. Gravity, applied in the vertical direction, is also normalized (ḡ = g/u̇go ). This

normalization allows the normalized strength η to replace the strength coefficient µ as one of

the key parameters. As in previous chapters, η is varied from 0.25 to 1.5.

The benefit of normalization is not obvious for this system, because in contrast to Chaps. 3

(Eq. 3.9) and 4 (Eq. 4.17), the strength Q does not drop out of the normalized equations

(Eq. 7.17), and ḡ varies from excitation to excitation. However, as will be seen later, the

normalization reduces the dispersion in the response over a ground motion ensemble, and is

therefore effective.

164
7.4.2 Computation of Median and Dispersion

The median and dispersion of a system response r over the ground motion ensemble are deter-

mined by: (1) computing the peak normalized response by nonlinear RHA of Eq. 7.17 to each

ground motion in the ensemble, (2) computing its median (Eq. 2.1) and dispersion (Eq. 2.2)

over the ensemble, and (3) multiplying the median of the normalized response by u∗y to obtain

the median value of the actual response as a function of η and median ground velocity u̇go of

the ensemble.

The responses of interest are the peak bearing deformation ubo , and the maximum and

minimum axial bearing forces relative to the static force (P+ /Pst and P− /Pst ), since actual

forces vary widely depending on the weight of the structure. Because the axial force oscillates

around the static force rather than zero, the positive and negative force increments, defined as

the maximum deviation of force in either direction from the static force:

∆P+ = P+ − Pst (7.18a)

∆P− = Pst − P− (7.18b)

are more meaningful for statistical analysis. Also, using force increments avoids the compli-

cation of a negative value of P− in case of bearing tension, since the equations for median

and dispersion are only meaningful for positive values. The dispersion measure of the force

increments ∆P± represents the variation of the actual forces P± , whose dispersion cannot be

computed, while the median forces are computed from the median force increments (Eq. 7.18).

7.4.3 Dispersion of Various Responses

Compared in Fig. 7.11 are the dispersion measures for actual and normalized response quantities

for a rigid block supported on bearings represented by the variable-strength model. System

parameters Tb and η are varied but the others are held constant, as their variation had little

effect on the dispersion.

Normalization clearly reduces the dispersion of the lateral bearing deformation, which

varies from 0.3 to 0.6 for normalized ūbo and from 0.4 to 1.0 for actual ubo (Fig. 7.11a); these

values are very consistent with the values observed in Chap. 3. The normalized force increments

165
Dispersion of Normalized Responses
1 1 1
(a) (b) η = 0.25 (c)
0.8 0.8 0.5 0.8
0.75
1.0

∆P̄+

∆P̄−
0.6 0.6 0.6
1.5
ūbo

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5

Dispersion of Actual Responses


1 1 1

0.8 0.8 0.8


∆P+

∆P−
0.6 0.6 0.6
ubo

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 7.11: Dispersion over the LMSR ensemble compared for normalized and actual re-
sponses: (a) lateral deformation ubo , (b) positive force increment ∆P+ , (c) negative force in-
crement ∆P− . System parameters are: Ωz = 30.4, Ωθx = 37.2.

∆P̄+ and ∆P̄− (Fig. 7.11b,c) have been reduced to dispersion values varying between 0.1 and

0.5, compared to values varying between 0.4 and 0.8 for the actual force increments. Fortunately,

the added complexity of the system does not nullify the benefit of normalization.

7.5 Median Response Trends

Investigated next is the variation of the peak bearing responses (lateral deformation ubo and

axial forces P± /Pst ) with the parameters identified previously: isolation period Tb , normalized

strength η, and vertical-to-lateral and rocking-to-lateral frequency ratios Ωz and Ωθx . Median

values due to the strong-component ground motion ensemble (Sec. 2.1) with median ground

velocity = 35 cm/s are computed by the method of Sec. 7.4.2. The axial force increment

∆P+ /Pst is maximized over the median values computed for the exterior bearings on each

side of the building; likewise for the minimum force increment ∆P− /Pst . Unless otherwise

indicated, the bearings are represented by the nonlinear variable-strength model, which is the

166
(a) Ωθx = 37.2 (b) Ωθx = 14.6
15 15

η = 0.25 η = 0.25
10 10
ubo (cm)

0.5 0.5
0.75 0.75
1.0 1.0
5 5
1.5 1.5

Rocking
No Rocking
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 7.12: Median response spectra for bearing deformation ubo : (a) Ωθx = 37.2 and (b)
Ωθx = 14.6; Ωz = 30.4.

most rigorous of all models considered.

7.5.1 Median Response Spectra

Rocking of the structure and axial-load effects in bearings have little influence on the lateral

deformation of the bearings, even for taller isolated buildings with significant rocking. Response

spectra of the lateral deformation ubo for a squat building (Ωθx = 37.2, Fig. 7.12a) and a more

slender building (Ωθx = 14.6, Fig. 7.12b) are nearly identical, suggesting that the rocking

properties are irrelevant for determining lateral deformation. Furthermore, these spectra are

very similar to the spectrum determined previously from analysis of the system permitting

lateral motion only, i.e., no rocking (rx θx ) or vertical (uz ) motions (Fig. 3.8); this spectrum is

also shown in Fig. 7.12a and b for comparison.

The deformation increases moderately with an increase in isolation period Tb and de-

creases with an increase in normalized strength η (or energy dissipation), but the maximum

and minimum axial forces (P± /Pst ) have a more complex dependence on Tb and η (Fig. 7.13).

The maximum forces for different normalized strengths are similar when the isolation period Tb

is small (< 2 seconds), but become more disparate as Tb increases, tending to decrease for small

strengths and increase for large strengths. The force spectra also vary a lot for two different

choices of the rocking frequency ratios (Fig. 7.13a vs. b).

167
(a) Ωθx = 37.2 (b) Ωθx = 14.6
2 3
η = 0.25
0.5
1.8 0.75 2.5
1.0
1.6 1.5
P+ /Pst

2
1.4

1.5
1.2

1 1
0 1 2 3 4 5 0 1 2 3 4 5

1 1

0.8
0.5

0.6
P− /Pst

0
0.4

−0.5
0.2

0 −1
0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 7.13: Median response spectra for axial forces P+ /Pst and P− /Pst : (a) Ωθx = 37.2 and
(b) Ωθx = 14.6; Ωz = 30.4.

For given parameters, the maximum and minimum axial force increments ∆P+ /Pst and

∆P− /Pst are nearly identical (but not exactly identical due to vertical motion of the CM and

varying stiffness of the bearings caused by axial-load effects), which causes the minimum axial

forces to appear as a mirror image of the maximum across the line P/Pst = 1 in Fig. 7.13.

Because of their similarity, the maximum and minimum forces may be referred to hereafter

in conjunction as the force extrema. The fact that the force extrema do not always increase

(maximum force increases while minimum force decreases) with increasing isolation period

seems to contradict the expectation that overturning effects are greater in longer period (taller)

buildings. However, because the structure has been modeled as rigid, the building slenderness

is independent of the isolation period, which may be a limitation of this analysis (Sec. 7.2.5).

168
Although increasing the yield strength of the bearings (increasing η) is known to control

lateral bearing deformation (Fig. 7.12), it does so at the expense of increasing the axial force

extrema (Fig. 7.13). For instance, a large normalized strength (η = 1.5) reduces the deformation

ubo by more than half compared to η = 0.25 (Fig. 7.12), but increasing η above 0.8 causes bearing

tension in the more slender building (P− /Pst < 0, Fig. 7.13b). Thus, limiting the normalized

strength η is one means of controlling bearing tension.

7.5.2 Influence of Frequency Ratios

The responses examined are essentially independent of the vertical frequency ratio Ωz , which

depends on the bearing shape factor S and therefore reflects the relative bearing dimensions

(diameter and rubber thickness). This means that the current practice of using high shape

factor bearings to avoid stability problems may be unnecessary, and the benefits of smaller shape

factors – vertical isolation and accommodation of larger deformations – can be realized. Both

lateral bearing deformations and axial forces show almost no variation with vertical frequency

ratio for slenderness ratios within the practical limit of h/d ≤ 2 (Fig. 7.14). Thus, although

some vertical bouncing of the building is predicted by the variable-strength model (Sec. 7.3),

it appears not to influence the axial forces (due to vertical softening, only axial deformations

are affected). For the rare more slender building (h/d = 4, L3 and L6 in Fig. 7.14), the vertical

frequency ratio Ωz begins to influence the bearing axial forces, although with no systematic

trends.

The bearing axial force extrema (P± /Pst ) are influenced by the rocking frequency ratio

Ωθx but not independently by the slenderness ratio h/d or number of bearings across the plan. In

other words, the force extrema are nearly the same in systems with different slenderness ratios

and bearing distributions so long as their rocking frequency ratios are the same. Evidence

is shown by the limited scatter in the forces when plotted against rocking frequency ratio

(Fig. 7.15a), which includes data for many different combinations of slenderness ratio and

number of bearings.

Due to large values and rapid variation of the axial force extrema in the range of low

frequency ratios, this range (say Ωθx < 15) is best avoided. Buildings with lower frequency

169
(a) (b) (c)
15 3.5 1
L1 L4
L2 L5
L3 L6 3 0.5
10
ubo (cm)

2.5 0

P+ /Pst

P− /Pst
2 −0.5
5
1.5 −1

0 1 −1.5
15 20 25 30 35 40 15 20 25 30 35 40 15 20 25 30 35 40
Vertical frequency ratio Ωz Vertical frequency ratio Ωz Vertical frequency ratio Ωz

Figure 7.14: Median values of (a) bearing lateral deformation ubo and (b,c) axial forces P+ /Pst
and P− /Pst plotted against the vertical frequency ratio Ωz . Other parameters are: Tb = 2 sec
and η = 0.5 (L1-L3), or Tb = 4 sec and η = 1.0 (L4-L6); Ωθx , which varies with Ωz , is determined
by h/d = 2 and nby = 2 (L1,L4), h/d = 2 and nby = 15 (L2,L5), or h/d = 4 and nby = 15
(L3,L6).

ratios (flexible in rocking relative to lateral motion) include very slender buildings that are

susceptible to overturning, and/or buildings with many bearings distributed across the plan,

which reduces their resistance to rocking. Rapid variation of the force extrema in this low

frequency range (Fig. 7.15a) makes them difficult to predict, whereas only a small change

in rocking frequency ratio marks the difference between compression and significant tension.

Overall, the force extrema increase approximately quadratically as the rocking frequency ratio

decreases (Fig. 7.15a). When building slenderness h/d and number of bearings nby are varied

independently, the force extrema vary linearly with h/d and increase rapidly with nby when nby

is small and slowly when nby is large (Fig. 7.15b).

A simplified model with only two bearings – at the building edges – should not be used

to estimate the axial forces in edge bearings of realistic systems with many bearings. The axial

force extrema for a two-bearing system are much lower than the others considered (5, 9, and

15, Fig. 7.15b). In fact, the data outliers in Fig. 7.15a are for two bearing systems, and hence

can be disregarded.

As demonstrated earlier (Sec. 7.5.1), the lateral deformation ubo of the bearings is essen-

tially unaffected by rocking properties except in the range of rocking frequency ratios Ωθx < 15

(Fig. 7.15a), or slenderness ratios h/d > 2 (Fig. 7.15b), unimportant since such slender buildings

170
(a) (b)
15 15

10 10
ubo (cm)

5 5

0 0
0 10 20 30 40 50 0 1 2 3 4 5

3.5 3.5
nby =2
3 3 5
9
15
P+ /Pst

2.5 2.5

2 2

1.5 1.5

1 1
0 10 20 30 40 50 0 1 2 3 4 5

1 1

0.5 0.5
P− /Pst

0 0

−0.5 −0.5

−1 −1

−1.5 −1.5
0 10 20 30 40 50 0 1 2 3 4 5
Rocking frequency ratio Ωθx Slenderness ratio h/d

Figure 7.15: Median bearing deformation ubo and axial forces P+ /Pst and P− /Pst plotted
against: (a) rocking frequency ratio Ωθx and (b) slenderness ratio h/d for several values of nby ;
Tb = 2 sec, η = 0.5, and Ωz = 30.4.

are rarely isolated.

We have seen that axial force extrema vary with the rocking frequency ratio Ωθx due

to changes in building slenderness or number of bearings (Fig. 7.15a), but do not vary with

vertical frequency ratio Ωz (Fig. 7.14). Because the rocking frequency is not independent of the

vertical frequency, the force extrema plotted against rocking frequency ratio show a wider band

of scatter when the vertical frequency ratio is varied (Fig. 7.16a), implying that the rocking

frequency ratio alone does not adequately represent the variation of forces. Instead, consider

171
(a) (b)
3 3
Ωz =20.7
26.7
2.5 30.4 2.5
32.7
34.3
P+ /Pst

2 2

1.5 1.5

1 1
0 10 20 30 40 50 60 0 0.5 1 1.5 2

1 1

0.5 0.5
P− /Pst

0 0

−0.5 −0.5

−1 −1
0 10 20 30 40 50 60 0 0.5 1 1.5 2
Rocking frequency ratio Ωθx Rocking-to-vertical frequency ratio Υx

Figure 7.16: For systems with different vertical frequency ratios Ωz , variation of median
axial forces P+ /Pst and P− /Pst with (a) rocking frequency ratio Ωθx and (b) rocking-to-vertical
frequency ratio Υx ; Tb = 2 seconds and η = 0.5.

the force extrema plotted as a function of the rocking-to-vertical frequency ratio Υx :

ωbθx Ωθx
Υx = ≡ (7.19)
ωbz Ωz

(Fig. 7.16b). The new parameter Υx is less sensitive than Ωθx to changes in Ωz , such that the

forces plotted against Υx do not vary with Ωz .

7.5.3 Significance of Bearing Axial-Load Effects

Comparison of the different bearing models demonstrate that the axial-load effects included in

the constant-strength model are unimportant, and only occasional circumstances merit the use

of the most rigorous variable-strength model. Thus, in most cases, the uncoupled nonlinear

bearing model can be used to estimate seismic demands with no loss of accuracy.

172
The peak lateral deformation of the bearing and force extrema are primarily unaffected

by the time variation of lateral and vertical bearing stiffness resulting from axial-load effects.

This claim is shown to be true in Fig. 7.17a and b, which compares the responses given by the

constant-strength and uncoupled nonlinear models for two rocking-to-vertical frequency ratios:

Υx = 1.22 (Fig. 7.17a) and Υx = 0.48 (Fig. 7.17b); and can be extended to all values of Υx

with confidence.

The time variation of bearing strength, which distinguishes the variable-strength model

from the constant-strength model, has only a minor effect on the lateral deformation and

axial force extrema over a wide range of system parameters; however, the variable-strength

model with its additional refinement is recommended for certain combinations of normalized

strength and rocking-to-vertical frequency ratio. Variation in bearing strength causes the lateral

deformation to increase slightly and the force extrema to decrease (Fig. 7.17c,d). The reduction

in the force extrema is greater for large values of the normalized strength η and for smaller

frequency ratios [compare Υx = 0.48 (Fig. 7.17d) to Υx = 1.22 (Fig. 7.17c)]. The increase in

the lateral deformation shows no systematic variation with the parameters.

The axial-load effects in the bearing model increase, thus the uncoupled bearing model be-

comes increasingly inaccurate, as the isolation system gains strength and the building increases

its tendency to rock (η increases and Υx decreases). Figure 7.18 demonstrates that for small

Υx the uncoupled model tends to underestimate the bearing deformation and overestimate the

axial force increments, and hence the force extrema, compared to the variable-strength model

(a positive percent discrepancy indicates the uncoupled model is conservative). The discrep-

ancy in both the lateral deformation and axial force increments is usually less than 10%, which

is not very significant given the other approximations and assumptions: a rigid model of the

superstructure, earthquake excitation in only one direction, etc.

Over the practical range of slenderness ratio, it does not appear necessary to use a bearing

model that includes the axial-load effects. However, time-variation of the strength – included in

the variable-strength model – should be considered if η ≥ 0.75 and Υx ≤ 0.75, which is intended

to include the approximate region where the error in the force increments exceeds 10%. On

173
(a) Υx = 1.22
15 2 1 η=.25
0.5
η=.25 1.8 0.8 0.75
1.0
10 0.5
ubo (cm)

1.5

P+ /Pst

P− /Pst
0.75 1.6 0.6
1.0
1.5 1.4 1.5 0.4
5
1.0
Constant-strength 1.2 0.75 0.2
0.5
Uncoupled η=.25
0 1 0
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
(b) Υx = 0.48
15 3 1

η=.25 1.5 η=.25


2.5 0.5
10 0.5
ubo (cm)

0.5
P+ /Pst

P− /Pst
0.75 1.0
1.0 2 0.75 0 0.75
1.5 1.0
5 0.5
1.5 η=.25 −0.5
Constant-strength 1.5
Uncoupled
0 1 −1
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
(c) Υx = 1.22
15 2 1 η=.25
0.5
η=.25 1.8 0.8 0.75
0.5 1.0
10
ubo (cm)

1.5
P+ /Pst

P− /Pst

0.75 1.6 0.6


1.0
1.5 1.4 1.5 0.4
5 1.0
1.2 0.75 0.2
Variable-strength 0.5
Constant-strength η=.25
0 1 0
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
(d) Υx = 0.48
15 3 1

η=.25 1.5 η=.25


0.5 2.5 0.5
10 0.5
ubo (cm)

P+ /Pst

P− /Pst

0.75 1.0
1.0 2 0 0.75
0.75
1.5 1.0
5 0.5
1.5 −0.5 1.5
Variable-strength η=.25
Constant-strength
0 1 −1
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
Tb (sec) Tb (sec) Tb (sec)

Figure 7.17: Median bearing deformation ubo and axial forces P+ /Pst and P− /Pst for (a,c)
Υx = 1.22 and (b,d) Υx = 0.48; comparison between (a,b) constant-strength model and uncou-
pled nonlinear model (c,d) variable-strength model and constant-strength model.

174
(a) (b) (c)
0 25 25
η=0.25
20 0.5 20

% Discrepancy in ∆P+

% Discrepancy in ∆P−
% Discrepancy in ubo

0.75
15 1.0 15
−5 1.5
10 10
5 5
0 0
−10
−5 −5
−10 −10
−15 −15 −15
0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2
Rocking-to-vertical frequency ratio Υx

Figure 7.18: Percent discrepancy of (a) the median bearing deformation ubo and (b,c) ax-
ial force increments ∆P+ /Pst and ∆P− /Pst with the uncoupled nonlinear model relative to
variable-strength model; Tb = 2 seconds.

the other hand, the variable-strength (or constant-strength) model is needed to accurately

determine the bearing axial deformations (Sec. 7.3), which, usually of lesser concern, have not

been treated in this median response analysis.

7.6 Design Equations to Estimate Peak Response

As in previous chapters, regression equations are fit to the peak deformation ubo and peak axial

forces P± /Pst , which can be used to estimate their values in preliminary design. Regression

analysis utilizes the normalized deformation ūbo and forces P̄± /Pst , which were computed by

RHA of Eq. 7.17 to each ground motion in the LMSR ensemble. The response data varies

over the parameters Tb from 2 to 4.5 seconds, η from 0.25 to 1.5, and Υx from 0.45 to 1.55,

based on Ωθx ranging from 9.5 to 53 and Ωz ranging from 20 to 35. This data set is limited

to the practical range of building slenderness ratios h/d ≤ 2; data for more slender buildings

that display less consistent behavior (Fig. 7.14) are ignored. The regression equations are based

on the responses computed using the variable-strength model, although previous observations

(Fig. 7.17) suggest that similar equations are likely to be produced by data from the constant-

strength or uncoupled nonlinear models.

Extending methods from Chapters 3-5, the regression equation relating the normalized

response to its parameters is assumed to be linear in the logarithmically transformed domain.

175
By testing various possibilities, the best fit to the normalized deformation ūbo was found to be:

ln ūbo = 0.70 − 1.82 ln Tb − 1.43 ln η (7.20a)

ūbo = 2.02Tb−1.82 η −1.43 (7.20b)

which, when multiplied by u∗y (Eq. 3.6) – a function of Tb , η and median (or design) ground

velocity u̇go – results in the following equation for the peak deformation ubo :

6.16 0.18 −0.43


ubo = T η u̇go (7.21)
4π 2 b

Not appearing in this equation, the rocking-to-vertical frequency ratio Υx was found to be

insignificant in regression analysis. This is not surprising since variation in the peak deformation

ubo with either rocking (Ωθx ) or vertical (Ωz ) frequency ratio was limited to instances where the

value of Ωθx was low (Fig. 7.14, Fig. 7.15a), which have been omitted here. For three different

values of Υx , Fig. 7.19 demonstrates that Eq. 7.21 is a good approximation to the exact median

deformation ubo .

The axial force extrema have been shown to depend on isolation period Tb , normalized

strength η, and rocking-to-vertical frequency ratio Υx . As discussed in Sec. 7.4.2, force incre-

ments are preferred for statistical analysis over actual forces, thus, equations are presented for

the force increments ∆P+ /Pst and ∆P− /Pst . Equations were first fit to the normalized force

increment ∆P̄ /Pst , and multiplied by u∗y (Sec. 7.4.1) to obtain the equation for the actual force

increment:
∆P 0.416 (−1.07+0.53 ln Tb ) (0.09+0.18 ln η+0.60 ln Tb ) −2.23
= T η Υx u̇go (7.22)
Pst 4π 2 b
Finally, the axial force extrema P+ /Pst and P− /Pst are computed by:

P+ /Pst = 1 + ∆P/Pst (7.23a)

P− /Pst = 1 − ∆P/Pst (7.23b)

Although the positive and negative force increments can differ due to vertical displacement of

the CM, the two were not statistically distinct when averaged over many ground motions, and

one equation suffices to represent the single force increment ∆P/Pst .

176
(a) Υx = 1.22 (b) Υx = 0.68 (c) Υx = 0.50
15 15 15

η=0.25 η=0.25 η=0.25


10 10 10
ubo (cm)

0.5 0.5 0.5


0.75 0.75 0.75
1.0 1.0 1.0
5 1.5 5 1.5 5 1.5
Exact median
Design equation
0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5

1.5 3 3

1.4
2.5 2.5
1.5
P+ /Pst

1.3
1.5 2 2 1.0
1.2 1.0 1.5 0.75
0.75 1.0 0.5
1.5 0.75 1.5
1.1 0.5 0.5 η=0.25
η=0.25 η=0.25
1 1 1
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5

1 1 1
η=0.25 η=0.25
0.9 0.5 0.5
0.75 0.5 0.5 η=0.25
0.75 0.5
1.0 1.0
P− /Pst

0.8 0.75
1.5 1.5
0 0 1.0
0.7 1.5
−0.5 −0.5
0.6

0.5 −1 −1
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 7.19: Design equation for bearing deformation ubo (Eq. 7.21) and axial forces P+ /Pst
and P− /Pst (Eqs. 7.22 and 7.23), compared to the exact median by RHA to the strong-
component ensemble, for (a) Υx = 1.22 (b) Υx = 0.68 and (c) Υx = 0.50; median/design
ground velocity u̇go =35 cm/s.

More complex than any of the previous equations for lateral deformation, Eq. 7.22 includes

second order coefficients for both Tb and η (i.e., Tbln Tb ) and a coefficient for interdependence of

Tb and η (i.e., η ln Tb ). The variation of the bearing forces with isolation period Tb depends on

the normalized strength η (Fig. 7.13, Fig. 7.19). If Tb and η were independent, the lines for

different values of η would be approximately parallel like they are for the deformation ubo (see

also Fig. 7.12). Interdependence between other parameters – Tb and Υx , η and Υx – was tested

but found to be statistically insignificant.

177
1.5

1.25

Normalized strength η
1

0.75

0.5

0.25
0.2 0.4 0.6 0.8 1
Rocking-to-vertical frequency ratio Υx

Figure 7.20: Conditions that induce bearing tension for a system with isolation period Tb = 3
seconds and PGV u̇go = 35 cm/s.

For three rocking frequency ratios, Eq. 7.23a and b are compared to the exact median

axial force extrema determined by nonlinear RHA in Fig. 7.19 (note different scales in Fig. 7.19a

vs. Fig. 7.19b,c). These design equations are quite accurate for higher frequency ratios, but

less when Υx is as low as 0.5 (Fig. 7.19c). The shape of the spectrum (variation with Tb and η)

appears to depend somewhat on Υx , but we were unable to include this effect without further

complicating the design equation.

The design equations can conveniently predict the occurrence of tension in the bearings.

For example, for design ground velocity u̇go = 35 cm/s and isolation period Tb = 3 seconds,

the shaded portion of Fig. 7.20 indicates the combinations of η and Υx for which Eq. 7.22

predicts at least one exterior bearing to go into tension (P− /Pst < 0 or ∆P/Pst > 1). For

building slenderness ratio h/d ≤ 2, the rocking-to-vertical frequency ratio Υx ≥ 0.45 and

tension is predicted only for normalized strength greater than about 0.75. The particular

combinations that produce tension will vary for other choices of u̇go and Tb . Used in this

manner, Eq. 7.22 provides a valuable tool to identify the potential for bearing tension, which

can then be avoided by modifying the design of the isolation system to ensure that the bearings

remain in compression.

178
7.7 Conclusions

Building codes require that the isolation system be modeled with sufficient detail to assess peak

forces in individual isolator units. Developed in the previous chapter for this purpose, the non-

linear constant-strength and variable-strength models for lead-rubber bearings, which include

axial-load effects, have been implemented in dynamic analysis of the structure. This investi-

gation of the coupled lateral-rocking response of isolated buildings including bearing axial-load

effects, subjected to an ensemble of ground motions, has led to the following conclusions:

1. Rocking of the structure and bearing axial-load effects, i.e., variation of the stiffness and

strength of the bearings with varying axial loads, have little influence on the peak lateral

bearing deformation. The error in neglecting axial-load effects is less than 5% typically

and less than 10% for isolated systems with large strength. Even if rocking is neglected

entirely, response spectra of the lateral deformation are still within 10% of those when

rocking and axial-load effects are included. In these spectra, the deformation increases

slightly with increasing isolation period Tb and decreases with increasing yield strength

of the isolation system (η).

2. Unlike lateral deformation, increasing bearing yield strength η generally causes the axial

force extrema to increase – maximum force increases and minimum force decreases – ex-

cept perhaps for small isolation periods Tb where η has little effect. The force extrema

are essentially independent of the vertical-to-lateral frequency ratio Ωz for a wide range

of systems, implying that use of high shape factor bearings to avoid instability is unnec-

essary and not otherwise beneficial. On the other hand, decreasing the rocking-to-lateral

frequency ratio Ωθx due to changes in slenderness or bearing distribution substantially

increases the force extrema. Because Ωθx and Ωz are interrelated, Ωθx alone cannot ad-

equately predict axial forces when Ωz is varied. The rocking-to-vertical frequency ratio

Υx = Ωθx /Ωz is less sensitive to changes in Ωz , and is thus a better parameter to predict

axial forces for a full range of systems.

179
3. Bearing axial-load effects are usually small enough that they can be neglected in de-

termining the maximum and minimum bearing axial forces. Of these effects, the time-

variation of the lateral and vertical stiffness, as included in the constant-strength model,

cause negligible change to the axial forces. The time-variation of strength included in

the variable-strength model, may, for certain parameter combinations, cause the forces to

vary by more than 10% compared to a typical uncoupled nonlinear model; in this regard

the variable-strength model is recommended when η > 0.75 and Υx < 0.75. Further

investigation is required to determine if axial-load effects are more significant when (1) a

detailed structural model is incorporated, or (2) the excitation is a near-fault motion.

4. By regression analysis of the response computed by nonlinear RHA of the system to the

LMSR ground motion ensemble, equations were developed to estimate the peak lateral

deformation (Eq. 7.21) and axial force extrema (Eqs. 7.22 and 7.23) in the bearings. The

design equation for the force extrema (Eq. 7.22) can predict the occurrence of tension in

the bearings, allowing the design to be modified early to eliminate the possibility of such

tension. The simplest way to reduce the force extrema and eliminate tension is to decrease

the normalized strength of the system, but at the expense of larger lateral deformation

in the bearings.

180
Appendix 7A: Parameter Selection

The following properties are needed to formulate Eq. 7.7:

• System parameters (Eq. 7.7): h/rx , yi /rx for each bearing, and total mass m.

• Bearing properties (Eq. 7.1 and 7.2): initial stiffness kI , post-yield stiffness kbi , strength

Qi , and force Po (variable-strength model only) to determine shear spring force fsi ; Euler

buckling load PE , height hb , vertical stiffness kbzi and damping coefficient cbzi .

We varied the building slenderness ratio h/d from 0.5 to 4, and the number of (in-plane)

bearings nby across the plan from 2 to 15. These two properties uniquely determine h/rx and the

bearing positions yi /rx , by assumption that the mass is uniformly distributed and the bearings

are evenly-spaced.

The linear or bilinear force-deformation relations of the individual bearing shear springs

fsi (si ) (Eq. 7.1a) were determined from the isolation period Tb (kb = ωb2 m with ωb = 2π/Tb ),

characteristic strength µ (Q = µw), and an assumed yield deformation of 1 cm (determines kI ).

The total stiffnesses (kb , kI ) and strength Q were distributed evenly among all bearings. For the

variable-strength model, the additional property Po (Eq. 7.3) was determined from Pst /Po = 3.

These parameters also suffice for the lateral force-deformation of the uncoupled bearing models

(Eq. 7.8).

For the remaining bearing properties, recall that the Euler buckling load PE is related to

the bearing buckling load Pcr by Pcr = PE PS (Eq. 6.6), where PS = GA(tr /hb ). Although the

cross-sectional area A can vary widely, isolation bearings are typically used at design pressures

pst ranging from 5 to 7 MPa (700 to 1000 psi) [36], with shear modulus G between 0.35 and

1.4 MPa (50 and 200 psi) [36]. Based on these typical values, pst /G, equivalent to Pst /GA, was

set to 10. The ratio of bearing height to total rubber thickness, hb /tr , was estimated as 4/3.

Together, these assumptions led to the following for PS and Pcr (Eq. 6.16):

Pst πSS2 Pst


PS = and Pcr =  (7A.1)
7.5 20 2(1 + Ks S 2 )

181
where Pst was the total weight of the rigid block; and finally PE :

2
Pcr 7.5Pst (πSS2 )2
PE = = (7A.2)
PS 800(1 + Ks S 2 )

which was divided by nby to get PE of a single bearing. The bearing height hb was computed:

kbi
hb = (7A.3)
PS

The nominal vertical bearing stiffness kbzi (Eq. 6.2) was determined from the post-yield stiffness

kbi (Eq. 6.1) and the shape factor S (Eqs. 6.11 and 6.15):

6S 2 kbi
kbzi = (7A.4)
(1 + Kc S 2 )

The damping coefficient cbzi was defined from an assumed damping ratio ζz of 5% (Eqs. 7.12

and 7.13). In the above equations, the shape factor S was varied from 10 to 30 (as previously

stated), the second shape factor (bearing aspect ratio) S2 = 3, and Kc = 0.004 and Ks = 0.0015

based on a ratio of shear modulus to bulk modulus G/K = 1/2000.

Appendix 7B: Rigid Structure Approximation

To determine a feasible range of slenderness ratio for which structures can be approximated as

rigid, we investigated the variation of structure period with slenderness ratio. Natural periods

of existing structures vibrating in their linear range have been measured from their vibration

responses recorded during earthquakes, and this data has been assembled for 106 buildings

in California [56]. The observed natural period of these buildings, whose structural systems

include concrete shear wall, R/C and steel moment resisting frames, dual systems, masonry,

and braced frames, are plotted against slenderness ratio in Fig. 7B.1a. The huge scatter in the

data might imply little correlation between slenderness ratio and structure period. However, as

the slenderness ratio increases, it becomes more unlikely that these buildings, if isolated, could

be classified as rigid (Ts ≤ 1.66, Sec. 7.2.5).

The data for the three main structural systems has been separated: concrete shear walls

(Fig. 7B.1b), R/C moment frames (Fig. 7B.1c), and steel moment frames (Fig. 7B.1d). For

each system, the median and 16th percentile (median-σ) curves are also shown, which have

182
(a)
8
concrete SW
R/C MRF
Structure period Ts

6 steel MRF
other

0
0 1 2 3 4 5 6
Slenderness h/d

(b) (c) (d)


6 6 6
transvs dir
5 long. dir 5 5
Structure period Ts

median
4 16th pct 4 4

3 3 3

2 2 2

1 1 1

0 0 0
0 2 4 6 0 2 4 6 0 2 4 6
Slenderness h/d Slenderness h/d Slenderness h/d

Figure 7B.1: Observed fundamental period Ts of existing buildings varying with slenderness
ratio h/d; median and 16th percentile (median-σ) of observed period for (a) all building types,
(b)concrete shear wall, (c) R/C moment resisting frames, and (d) steel moment resisting frame
buildings.

been fit to the associated data by linear regression analysis assuming Ts = α ∗ (h/d)β . The 16th

percentile curve, lower than the median by one standard deviation, may be more representative

of isolated buildings, which are required to be stronger and thus have shorter natural periods Ts

than comparable fixed-base buildings. Allowing the fixed-base structure period to be no greater

than 1.66 seconds to assume rigidity, the maximum slenderness ratios are ∞, 3.41, and 1.64 for

concrete shear walls, R/C moment frames and steel moment frames, respectively, as computed

by the 16th percentile equation. According to the equations, it is most feasible for concrete

shear wall systems to be both slender and sufficiently stiff; however, no data is available for

these buildings with slenderness ratios above 2. Thus, the selection of upper limiting slenderness

183
ratio is a judgment call. For practical application we consider the slenderness ratio for isolated

buildings to be limited to h/d < 2, but for completeness higher values of slenderness ratio are

considered.

Appendix 7C: Notation

As cross-sectional area of bearing or isolation system modified for unde-

forming steel layers: Ahb /tr

cbθx coefficient of damping for rocking about x-axis

cbz coefficient of damping for vertical motion, summed over all bearings

cbzi viscous damping coefficient of single bearing in z-direction

d minimum dimension or width of building

D bearing diameter

Ec instantaneous compression modulus of rubber-steel composite bear-

ing

EIs effective bending stiffness of a bearing

fbi lateral force of ith bearing

fDi vertical damping force of ith bearing

fsi shear spring force (in two-spring model) corresponding to ith bearing

G shear modulus of bearing or isolation system

h building height

hb total bearing height including steel and rubber layers

h/d building slenderness (height-to-width) ratio

kb lateral postyield stiffness of isolation system, summed over all bear-

ings

kbi nominal postyield stiffness of single bearing: GAs /hb

kbz vertical stiffness of isolation system, summed over all bearings

kbzi nominal vertical stiffness of single bearing: Ec As /hb

kI initial lateral stiffness of bearing or isolation system

184
m mass of isolated block

nby number of bearings across the plan in the y-direction

Pcr critical load or load that induces bearing buckling

PE Euler buckling load, neglects shear stiffness of the bearing: π 2 EIs /h2b

Pi axial or vertical force in ith bearing (positive in compression)

P+ maximum axial force observed in any bearing

P− minimum axial force observed in any bearing (negative if tensile)

∆P+ , ∆P− maximum and minimum axial force increments; maximum force de-

viation in either direction from static force

Po value of strength used to calibrate variable-strength model, typically

about 1/3 of the bearing static load

Pst static load, gravity load, or bearing design axial load

Q lateral characteristic strength of isolation system, summed over all

bearings

Qi nominal characteristic strength of a single bearing

rx radius of gyration about x-axis, or axis-of-rocking

si deformation of shear spring (in two-spring model) corresponding to

ith bearing

S bearing shape factor; D/4t for circular bearing

t thickness of a single bearing layer

Tb isolation period, corresponds to nominal postyield stiffness of isola-

tion system

Ts fundamental period of fixed-base structure

ubi lateral deformation of ith bearing

ubo peak lateral deformation

ubzi vertical deformation of ith bearing

ubz+ maximum vertical deformation observed in any bearing

ubz− minimum vertical deformation observed in any bearing

185
∆ubz− , ∆ubz+ maximum deviation of vertical deformation in either direction from

the static value

üg ground acceleration history

u̇go peak ground velocity

ust static vertical deformation due to gravity loading (Pst )

uy lateral deformation of the CM (in the y-direction)

u∗y Q/kb = µg/ωb2

uz vertical deformation of the CM

vi vertical deformation of ith bearing due to axial flexibility

zi yield function, or dimensionless plastic variable governing yielding of

the ith bearing

γi ratio of nominal strength in single bearing to total strength: Qi /Q

ζθx damping ratio for rocking about x-axis

ζz damping ratio for vertical motion

η strength of isolation system normalized by peak ground velocity:

µg/(ωd u̇go )

θx rocking or rotation of block about x-axis

θi rotation of bearing in two-spring model

κ bulk modulus of rubber

µ characteristic strength ratio of isolated block: Q/(mg)

Υx rocking-to-vertical frequency ratio: Ωθx /Ωz

ωb lateral frequency of isolation system, corresponding to postyield stiff-

ness

ωbθx rocking frequency about x-axis

ωbz vertical frequency of isolation system

Ωθx rocking-to-lateral frequency ratio: ωbθx /ωb

Ωz vertical-to-lateral frequency ratio: ωbz /ωb

ωd corner frequency separating velocity and displacement-sensitive spec-

tral regions
186
8: ESTIMATING BEARING DEFORMATIONS AND FORCES IN
SYMMETRIC- AND ASYMMETRIC-PLAN SYSTEMS INCLUDING
TORSION AND ROCKING

8.1 Introduction

Previous chapters have emphasized that reliable estimates of the controlling lateral deformation

and axial forces in the isolation system are key ingredients to an efficient design. For rubber

bearings, correct prediction of the lateral deformation helps the designer balance the tradeoff

between bearing flexibility, stability, and energy dissipation. Axial forces are of particular inter-

est to fulfill testing requirements and achieve an isolation system that remains in compression

under design loads. Thus, initial estimation of pertinent design quantities greatly simplifies the

overall procedure.

For this purpose, Chapter 3 developed a design equation based on nonlinear response

history analysis (RHA) to estimate the peak, or absolute maximum, isolator deformation due

to unidirectional – single component – excitation, which was shown to be much superior to the

equivalent-linear analysis upon which the static lateral response procedure of the International

Building Code (IBC) [1] is based. Subsequently, the design equation was extended to include

the increase in deformation due to bidirectional excitation – two components of ground motion

applied in orthogonal lateral directions. In Chapter 4, a factor that depends on the eccentricity

and the distance from the CM to the outlying isolator was developed to extend the design

equation to asymmetric-plan systems.

The experimentally observed influence of time-variation of axial forces on the bearing

response, known as axial-load effects, motivated the development of improved models for lead-

rubber bearings (Chapter 6). With these models, analysis was extended to lateral-rocking

response due to unidirectional excitation, i.e., unidirectional analysis, and the overall scope

of investigation was expanded to determine bearing axial forces. New design equations for the

peak lateral deformations and maximum and minimum axial forces were based on the improved

models, although the influence of these models was ultimately shown to be small (Chapter 7).

187
Not considered previously, axial-load effects may cause accidental torsion in symmetric-

plan buildings. Time variation of the stiffnesses and strengths of individual bearings, caused by

variation in axial loads, can induce a time-varying eccentricity. Although it can be computed

with suitable analytical models, torsion from this source is accidental because it occurs in

nominally symmetric buildings. Investigation of accidental torsion here is motivated by earlier

work on friction pendulum (FP) isolation systems, where a sophisticated large-deformation

model [57, 58, 59] was used to identify accidental torsion that may occur from varying axial

loads on the isolators [60]. While accidental torsion was found to be insignificant in the isolation

system unless the superstructure, permitted to uplift, reached incipient overturning, it was

noteworthy in the superstructure.

To consider accidental torsion, among other things, the investigations of previous chapters

should be integrated into a comprehensive analysis of the combined lateral-torsional-rocking

response due to bidirectional excitation, applicable to both symmetric and asymmetric-plan

buildings. Using the tools that have been developed, the following questions can be answered

with minimal additional effort: Does rocking influence the factors by which the lateral deforma-

tion increases due to bidirectional excitation and asymmetry of plan? How does rocking about

multiple axes or plan asymmetry influence the maximum and minimum axial forces?

This final chapter simultaneously considers rocking and torsion due to bidirectional exci-

tation in symmetric and asymmetric-plan buildings with the following objectives: (1) to inves-

tigate the possibility of accidental torsion due to axial-load effects and overturning in buildings

isolated with lead-rubber bearings, and (2) to improve previous design equations for estimating

the peak lateral deformation and axial forces in any bearing.

8.2 System Model and Governing Equations

8.2.1 System Considered

Consider a rigid block of arbitrary geometry supported on a two-dimensional grid of isolation

bearings, such that the total system is either nominally symmetric (Fig. 8.1a) or asymmetric

(Fig. 8.1b) in plan. When subjected to orthogonal components of lateral ground excitation

188
z, uz
θz

z, uz

CM x, ux
CM θx x, ux h -y, -uy
h θy
-y, -uy

d ügx Bearings located


b (xi , yi , -h/2) from CM
ügy

Figure 8.1: (a) Symmetric-plan and (b) asymmetric-plan representative systems, for analysis
of lateral-torsional-rocking response due to bidirectional excitation

ügx and ügy , the complete motion of the isolated block is described by six degrees-of-freedom

(DOFs), selected as translations (ux , uy and uz ) and rotations (θx , θy and θz ) about each CM

axis. Specifically, θx and θy refer to rocking about the x and y-axes, while θz refers to twisting

or torsion about the vertical (z) axis. Each isolation bearing is located at xi , yi , −h/2 from

the CM in the x, y and z-directions, respectively.

In this study, symmetric buildings are assumed to be rectangular with plan dimensions

b and d (b/d ≥ 1) and height h, wherein the radii of gyration about their respective axes
  
are easily computed as: rx = (d2 + h2 )/12, ry = (b2 + h2 )/12, and rz = (b2 + d2 )/12.

Computation of these radii for asymmetric-plan buildings is considered case-by-case.

8.2.2 Bearing Models

For this investigation, the variable-strength model of the previous two chapters is extended to

bidirectional motion. To achieve this, the two-spring model (Fig. 6.2), which was the basis for

the variable-strength model, is modified with a shear spring that deforms bidirectionally and a

rotational spring that rotates about any axis in the x − y plane. Equilibrium equations for the

189
ith bearing, representative of the modified spring assemblage are:

fbxi − fsxi (sxi , syi ) + Pi θyi = 0 (8.1a)

fbxi hb − PE hb θyi + Pi (sxi + hb θyi ) = 0 (8.1b)

fbyi − fsyi (sxi , syi ) + Pi θxi = 0 (8.1c)

fbyi hb − PE hb θxi + Pi (syi + hb θxi ) = 0 (8.1d)

Pi − kbzi vi = 0 (8.1e)

Equations 8.1c-e are essentially identical to Eq. 6.4 for unidirectional analysis, and represent

the balance of lateral forces (Eq. 8.1c) and moments (Eq. 8.1d) for motion in the y-direction,

along with the balance of vertical forces (Eq. 8.1e). Analogous equations 8.1a-b are added for

balance of lateral forces and moments in the x-direction. Resultant lateral forces fbxi and fbyi

and axial forces Pi on a bearing are balanced by (a) the force in the nonlinear shear spring

with components fsxi and fsyi acting through deformations sxi and syi , (b) the moment of the

rotational spring with stiffness PE hb acting through rotations θxi and θyi about the respective

axes, and (c) the force of an additional vertical spring with stiffness kbzi acting through axial

deformation vi . With stiffness PE hb – the Euler buckling load times the bearing height – the

rotational spring resists an Euler buckling type of instability; spring rotations θxi and θyi are

positive about the −x and +y-axes, respectively.

The bilinear force-deformation relation of the shear spring is represented by (Eq. 4.1):
⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎨fsxi ⎪
⎪ ⎬ ⎪
⎨sxi ⎪
⎬ ⎪
⎨zxi ⎪⎬
= kbi + Qi (8.2)
⎩fsyi ⎪
⎪ ⎭ ⎪
⎩syi ⎪
⎭ ⎪
⎩zyi ⎪⎭

where kbi and Qi are the nominal postyield stiffness and strength for motion in any lateral

direction. The components zxi and zyi of the yield function interact, with magnitude zi =
2 + z 2 )1/2 ≤ 1, representative of a circular yield surface. The yield deformation of the spring
(zxi yi

is assumed to be 1 cm at the nominal yield strength (Sec. 3.3.1). The actual yield strength

varies in time according to:


 
Qi (t) = Qi 1 − e(−Pi /Po ) (8.3)

190
where Po , typically about 1/3 of the bearing static load, can be determined from characteristic

tests of prototype bearings.

The kinematic equations, relating resultant bearing deformations in the x, y and z-

directions (ubxi , ubyi and ubzi ) to the local shear and axial deformations and rotations are:

ubxi = sxi + hb θyi (8.4a)

ubyi = syi + hb θxi (8.4b)


hb 2 2
ubzi = vi + sxi θyi + syi θxi + (θ + θyi ) (8.4c)
2 xi
Although primarily shear deformation, the resultant lateral deformation (ubxi or ubyi ) in the

respective direction is increased by the bearing rotation acting through its height (Eqs. 8.4a,b).

The total vertical deformation ubzi includes the axial deformation vi and additional vertical dis-

placement incurred in the laterally deformed configuration (Eq. 8.4c). In contrast to the global

sign convention, the vertical deformation ubzi and force Pi are positive for bearing compression.

This bearing model incorporates axial load effects, first identified in Chapter 6, into

the bidirectional response of the bearing. For instance, a variation of the resultant lateral

stiffness with axial load and variation of the resultant vertical stiffness with lateral deformation

(Sec. 6.2.1), are caused by the interaction of shear and rotational springs in the modified spring

assemblage. Variation of the yield strength with axial load (Eq. 8.3) is incorporated directly into

the numerical algorithm for the bilinear shear spring. It is clear from the equilibrium equations

(Eq. 8.1) that the three resultant forces fbxi , fbyi and Pi are coupled, wherein force-deformation

in any one direction depends on the response in the other two.

The equations of motion of the system neglecting bearing axial-load effects will be exam-

ined to identify influential parameters. When axial-load effects are neglected, Eq. 8.1 simplifies

to the following:

fbxi = kbi ubxi + Qi zxi (8.5a)

fbyi = kbi ubyi + Qi zyi (8.5b)

Pi = kbzi ubzi (8.5c)

191
where the lateral force-deformation equations (Eqs. 8.5a and b) and the linear vertical force-

deformation (Eq. 8.5c) are now uncoupled. This lateral force-deformation relation is simply

Eq. 8.2, where the shear deformation now represents the total lateral deformation, and the

bearing strength no longer varies in time.

Viscous damping in the vertical direction is included by an additional axial force com-

ponent for each bearing, fDi = cbzi u̇bzi where cbzi is the damping constant and u̇bzi the axial

bearing velocity.

8.2.3 Equations of Motion

The equations of motion of the system including bearing axial-load effects, subjected to lateral

earthquake accelerations ügx , ügy and gravity loading g are:


müx + fbxi = −mügx (8.6a)
i
 
mrz2 θ̈z − yi fbxi + xi fbyi = 0 (8.6b)
i i

müy + fbyi = −mügy (8.6c)
i
 h 
mrx2 θ̈x − yi cbzi u̇bzi + fbyi − yi Pi = 0 (8.6d)
2
i i i
 h 
2
mry θ̈y + xi cbzi u̇bzi − fbxi + xi Pi = 0 (8.6e)
2
i i i
 
müz − cbzi u̇bzi − Pi = mg (8.6f)
i i

The lateral (fbxi , fbyi ) and axial (Pi ) forces are implicit functions of the bearing deformations

ubxi , ubyi and ubzi (Eqs. 8.1 and 8.4), which can be written in terms of the displacements ux ,

uy , uz , and rotations θx , θy and θz about the CM:

h
ubxi = ux − θy − yi θz (8.7a)
2
h
ubyi = uy + θx + xi θz (8.7b)
2
ubzi = −uz − yi θx + xi θy (8.7c)

Substituting the time derivative of Eq. 8.7c for u̇bzi into Eq. 8.6 and dividing by radii of

192
gyration rx , ry , and rz where appropriate leads to
⎧ ⎫ ⎧ ⎞ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪  ⎪ ⎪

⎪ ü x ⎪ ⎪ ⎪
⎪ u̇ ⎪
⎪ f ⎪
⎪ ⎪
⎪ ügx ⎪



⎪ ⎪
⎪ ⎪

x
⎟ ⎪

i bxi ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎟ ⎪
⎪   ⎪
⎪ ⎪
⎪ ⎪


⎪ r θ̈ ⎪
⎪ ⎪
⎪ r θ̇ ⎟ ⎪
⎪ − yi
f + xi
f ⎪
⎪ ⎪
⎪ 0 ⎪


⎪ z z ⎪
⎪ ⎪
⎪ z z ⎟ ⎪
⎪ i r z
bxi i r z
byi ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎟ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪


⎨ üy ⎬ ⎪ ⎪
⎨ u̇y ⎟ ⎨ ⎟ ⎪ ⎪
⎬ ⎪
⎨ ⎪
fbyi ügy ⎬
m +C ⎟ i
= −m
⎪ ⎪ ⎪ ⎟+⎪ h   yi ⎪ ⎪ ⎪
(8.8)

⎪ r θ̈ ⎪
⎪ ⎪
⎪ r θ̇ ⎟ ⎪
⎪ f − P ⎪
⎪ ⎪
⎪ 0 ⎪⎪


x x⎪
⎪ ⎪

x x⎟ ⎪
⎪ 2rx i byi i rx i ⎪ ⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎟ ⎪
⎪   ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎟ ⎪
⎪ − h xi ⎪
⎪ ⎪
⎪ ⎪


⎪ r y θ̈ y ⎪
⎪ ⎪
⎪ r y θ̇ y ⎟ ⎪
⎪ f bxi + P i ⎪
⎪ ⎪
⎪ 0 ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎠ ⎪

2r y i i ry

⎪ ⎪ ⎪
⎪ ⎪

⎩ ü ⎭ ⎪ ⎪
⎩ u̇ ⎪
⎩  ⎪
⎭ ⎪
⎩ ⎪
z z − i Pi −g ⎭

where
⎡ ⎤
⎡ ⎤  yi2  x i yi  yi
⎢ i rx2 cbzi − i rx ry cbzi i rx cbzi ⎥
⎢0 0 ⎥ ⎢    ⎥
C=⎣ ⎦ and Cz = ⎢ x i yi
⎢− i rx ry cbzi
x2i
i ry2 cbzi − xi ⎥
i ry cbzi ⎥
(8.9)
0 Cz ⎣ ⎦
 yi  
i rx cbzi − xi
i ry cbzi i cbzi

Because of the influence of bearing forces at multiple DOFs in the equations of motion

(Eq. 8.8) and their relation to deformations at the bearing level (Eqs. 8.1 and 8.4), all six global

DOFs of the system are coupled. Thus, a single component of ground excitation can induce

a combined lateral, torsional, rocking, and vertical (bouncing) response. In its most general

form, the damping matrix Cz implies coupling of damping forces in the two rocking and vertical

DOFs. However, if the isolated building is symmetric with stiffness-proportional damping, all

off-diagonal terms of Cz are zero.

8.2.4 Normalized Equations

The normalization procedure of previous chapters is extended to the full three-dimensional

response analysis of the isolated block, leading to the following normalized equations:
⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ ⎪ ⎪ ⎪ ⎪  ⎪ ⎪

⎪ ¨
ū x ⎪ ⎪ ⎪
⎪ ˙
ū x ⎪ ⎪ ⎪
⎪ f ⎪
⎪ ⎪
⎪ ¨gx ⎪
ū ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

i bxi ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪   ⎪
⎪ ⎪
⎪ ⎪


⎪ r ¨
θ̄ ⎪
⎪ ⎪
⎪ r ˙
θ̄ ⎪
⎪ ⎪
⎪− yi
f + xi
f ⎪
⎪ ⎪
⎪ 0 ⎪


⎪ z z ⎪
⎪ ⎪
⎪ z z ⎪
⎪ ⎪
⎪ i rz bxi i rz byi ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪


⎨ ū ⎪ ⎪
¨y ⎬ C ⎨ ū˙ y ⎬ ω 2 ⎨⎪ ⎪ ⎪
⎬ ⎪
⎨ ¨ ⎪

i f byi ω 2 ū gy
+ + b
= − b
(8.10)

⎪ ⎪ m⎪ ⎪ Q ⎪ h   ⎪ ηωd ⎪ ⎪

⎪ rx θ̄¨x ⎪⎪



⎪ rx θ̄˙x ⎪⎪



⎪ fbyi − i ryxi Pi ⎪ ⎪



⎪ 0 ⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

2r x i ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪   ⎪
⎪ ⎪
⎪ ⎪


⎪ r ¨
θ̄ ⎪
⎪ ⎪
⎪ r ˙
θ̄ ⎪
⎪ ⎪
⎪ − h
f + xi
P ⎪
⎪ ⎪
⎪ 0 ⎪



y y ⎪
⎪ ⎪

y y ⎪
⎪ ⎪
⎪ 2r y i bxi i ry i ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪

⎩ ū ¨z ⎭ ⎩ ū˙ z ⎭ ⎩ − i Pi ⎭ ⎩ −ḡ ⎭

193
They were obtained by first dividing Eq. 8.8 by u∗y = Q/kb (Eq. 3.6) such that all DOFs are

normalized – e.g. ūx = ux /u∗y – and then substituting the normalized strength η = a∗y /ωd u̇gyo

(Eq. 3.8) into the right-hand side. This normalizes the ground acceleration by the peak

single-direction velocity, whose corresponding component is always applied in the y-direction


¨gx = ügx /u̇gyo , ū
(ū ¨gy = ügy /u̇gyo ). Thus, the corresponding components of ground velocity vary

between -1 and +1. Gravity, applied in the vertical direction, is also normalized (ḡ = g/u̇gyo ).

Normalization of the equations of motion (Eq. 8.10) parallels that in previous chapters.

Without proof, then, prior assertions regarding the peak normalized responses of the system

are repeated: (1) for a given set of parameters that suitably define the system, the median

normalized response over an ensemble of ground motions is only weakly affected by ground

motion intensity, and (2) the dispersion of the normalized response over the ensemble is small.

Ultimately, this allows meaningful prediction of the peak forces or deformations for varying

intensities of design ground motion.

For given parameters, the median of a response over an ensemble of ground motions,

with weak and strong-component ensembles applied in the x and y-directions, respectively, is

determined by: (1) computing the peak normalized response by nonlinear RHA of Eq. 8.10 to

each ground motion in the ensemble, (2) computing the median (Eq. 2.1) over the ensemble,

and (3) multiplying the normalized median by u∗y to obtain the actual median as a function of

η and design ground velocity u̇gyo .

8.3 Representative System and Parameter Selection

8.3.1 Plan Layout

For a study of asymmetric-plan systems to be useful, the results should be applicable to a

wide variety of systems, and not limited to the plans considered. Formerly, plan asymmetry

was introduced by varying the stiffnesses and strengths of individual bearings in an idealized,

rectangular plan (Sec. 4.3). This approach was justified because the idealized system with key

parameters identical to the actual asymmetric-plan system computed its response with less than

1% error (Sec. 4.3.2). Furthermore, it was shown that a two-way asymmetric system could be

194
Table 8.1: Idealized Plan Layouts for Bidirectional Analysis

Plan No. b/d nbx nby Total nb Ωθz


S1 1 2 2 4 1.73
S2 2 3 2 6 1.48
S3 3 4 2 8 1.34
S4 1 3 3 9 1.41
S5 2 5 3 15 1.27
S6 3 7 3 21 1.18
S7 1 4 4 16 1.29
S8 1 5 5 25 1.23
S9 2 7 4 28 1.18

represented by a one-way asymmetric system, since the two are equivalent under a simple axis

rotation (Sec. 4.3.3).

Several difficulties hinder the use of this approach when rocking of the system is included.

First, because the vertical bearing stiffnesses in the idealized system are varied to introduce

eccentricity, the gravity or static loads Pst on the individual bearings, which are proportional to

these stiffnesses, vary nonuniformly. This particular variation of static loading is unnatural, and

whether it affects the dynamic response factor is unknown. Second, when rocking is included

the properties of an asymmetric-plan system cannot, in general, be replicated in an idealized,

rectangular system, preventing direct comparison of their responses to verify accuracy. Third,

the axis transformation that allowed a two-way asymmetric system to be represented by a

one-way asymmetric system is no longer valid when rocking is present.

Nevertheless, the idealized system approach is pursued because of the tremendous ben-

efit provided if, despite the difficulties mentioned, the results can still be extended to a wide

range of asymmetric-plans with reasonable accuracy. The effectiveness of this approach will

be determined by comparing the final design equations to the response of several systems with

legitimate asymmetry-of-plan.

Nine different rectangular plan layouts have been selected to characterize a wide range

of symmetric and asymmetric buildings. Labeled S1-S9, these 9 plans are defined by their

plan aspect ratios b/d and the number of bearings nbx and nby along the x and y-axes. These

195
properties, as well as the torsional frequency ratio Ωθz , which depends only on the plan layout,

are listed in Table 8.1 for each plan. The rocking frequency ratios Ωθx and Ωθy depend also on

the slenderness – or height-to-width – ratio h/d, which will be varied in analysis. To represent

symmetric buildings, the stiffnesses and strengths of all bearings are modeled as identical,

while they are varied to impose a one-way eccentricity in asymmetric buildings. Plans S1-S3

are limited to symmetric-plan systems because they have unusually large torsional-to-lateral

frequency ratios Ωθz .

8.3.2 Uncoupled Nonlinear Bearing Model

Although the analyses presented here are limited to the variable-strength model, the equations

of motion of the system with the uncoupled nonlinear bearing model help to identify several

parameters that characterize the system even when axial-load effects are present. Substituting

Eq. 8.5 into Eq. 8.10 specializes the normalized equations of motion for the simpler uncoupled

model:
⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧ ⎫
⎪ 

⎪ ūx ⎪
¨ ⎪



⎪ ˙
ūx ⎪ ⎪
⎪ ⎪

⎪ ūx ⎪ ⎪
⎪ ⎪

⎪ i γi zxi





⎪ ¨gx ⎪
ū ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪ ⎪
⎪ ⎪

⎪ r ¨
z z⎪
θ̄ ⎪ ⎪
⎪ r ˙
θ̄ ⎪
⎪ ⎪
⎪ r θ̄ ⎪
⎪ ⎪
⎪ γ (− yi
z + xi
z ⎪
yi ⎪
) ⎪
⎪ 0 ⎪


⎪ ⎪
⎪ ⎪
⎪ z z⎪
⎪ ⎪
⎪ z z⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
i i r z
xi r z

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪  ⎪
⎪ ⎪ ⎪

⎪ ¨y ⎪
⎨ ū ⎬ ⎪
⎨ ˙ ⎪
⎬ ⎪
⎨ ⎪
⎬ ⎪
⎨ ⎪
⎬ ⎪ ¨gy ⎪
⎨ū ⎬
C ūy ūy i γ i zyi ωb2
+ +K  2
+ ωb = − (8.11)

⎪ ⎪ m⎪ ⎪ ⎪ ⎪ ⎪ h  ⎪ ηωd ⎪ ⎪
⎪rx θ̄¨x ⎪
⎪ ⎪



⎪ rx θ̄˙x ⎪




⎪ rx θ̄x ⎪




⎪ i γi zyi





⎪ 0 ⎪



⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

2r x ⎪
⎪ ⎪
⎪ ⎪


⎪ ¨⎪ ⎪ ⎪
⎪ ˙ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ h 

⎪ ⎪
⎪ ⎪


⎪ r θ̄ ⎪
⎪ ⎪
⎪ r θ̄ ⎪
⎪ ⎪
⎪ r θ̄ ⎪
⎪ ⎪
⎪ − γ z ⎪
⎪ ⎪
⎪ 0 ⎪



y y⎪
⎪ ⎪

y y⎪
⎪ ⎪

y y⎪
⎪ ⎪
⎪ 2r y i i xi ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎩ ū¨z ⎭ ⎩ ˙ūz ⎭ ⎩ ūz ⎭ ⎩ 0 ⎭ ⎩ −g ⎪

with stiffness matrix K  :


⎡ ⎤
e
1 − rby 0 0 − 2rhy 0
⎢ z ⎥
⎢ e ⎥
⎢ − by Ω2θz ebx h ebx h eby
0 ⎥
⎢ rz rz 2rx rz 2ry rz ⎥
⎢ ⎥
⎢ ebx ⎥
2⎢
⎢ 0 1 h
0 0 ⎥

K = ωb ⎢
rz 2rx ⎥ (8.12)
⎢ 0 eby ⎥
Ω z rx ⎥
h ebx e
⎢ 2rx rz
h
2rx Ω2θx −Ω2z rxbxy
ry
2

⎢ ⎥
⎢ h eby e ⎥
⎢− 2ry − 2rhy 0 −Ω2z rxbxy Ω2θy −Ω2z erbx ⎥
⎣ rz ry y

eby
0 0 0 Ω2z rx −Ω2z erbx
y
Ω2z

The symbol γi = Qi /Q was introduced, allowing Q to be factored out of the nonlinear terms

and balance the denominator Q in Eq. 8.10.

196
Equation 8.11 has been written in terms of the following parameters of a linear system

(η = 0):

1. vertical (Ωz ), rocking (Ωθx and Ωθy ) and torsional (Ωθz ) – all relative to lateral – frequency

ratios:
ωbz ωbθx ωbθy ωbθz
Ωz = Ωθx = Ωθy = Ωθz = (8.13)
ωb ωb ωb ωb
where the lateral (ωb ), vertical (ωbz ), two rocking (ωbθx and ωbθy ), and torsional (ωbθz )

frequencies are
 1/2  1/2 $  %1/2
h2
i kbi ikbzi 4 kb + i yi2 kbzi
ωb = ωbz = ωbθx =
m m mrx2
$  %1/2  1/2
h2
4 kb + i x2i kbzi i kbi (x2i + yi2 )
ωbθy = ωbθz = (8.14)
mry2 mrz2

2. x and y-direction eccentricities ebx and eby , and a related term ebxy :

1  1  1  1 
ebx = kbi xi ≡ kbzi xi eby = kbi yi ≡ kbzi yi (8.15a)
kb kbz kb kbz
i i i i

1 
ebxy = kbzi xi yi (8.15b)
kbz
i

Most of the above definitions (Eqs. 8.13-8.15) were introduced earlier: Eq. 7.10 (Ωz and Ωθx ),

Eq. 7.11 (ωbz and ωbθx ), and Eq. 4.4b (ebx and eby ).

In the traditional sense, nonzero eccentricities ebx and eby represent the distance between

the CM and center of rigidity, and cause a lateral-torsional coupling of the system in response

to a horizontal force. For this particular system, the same eccentricities are also associated

with coupling of the rocking and vertical DOFs, by assumption that the vertical stiffness of

each bearing is proportional to its lateral stiffness. The eccentricities are zero for a nominally

symmetric system (Sec. 4.2.2), causing many off-diagonal terms to drop out of the stiffness ma-

trix K  (Eq. 8.12), which eliminates the torsional rz θz and vertical uz deformations (Eq. 8.11).

This conclusion does not apply, however, when axial-load effects are included.

197
The term ebxy (Eq. 8.15) refers to a direct coupling between the two rocking deformations

rx θx and ry θy . When ebxy is nonzero, rocking occurs about both x and y-axes even if the

excitation is unidirectional. For every system, however, there exists an orientation of the x and

y-axes for which ebxy = 0. Thus, the value of ebxy does not convey meaningful information

about the system, and is hereafter ignored.

8.3.3 System Parameters

The properties that must be defined to solve the normalized equations of motion (Eq. 8.10)

with Eqs. 8.1 and 8.4 for the bearing model are numerous. Yet the list of parameters that

significantly influence the system response is short. This list combines the parameters identified

previously for unidirectional lateral-rocking analysis (Sec. 7.2.4) and lateral-torsional analysis

(Sec. 4.2.3). However, Ωθx and Ωz have been replaced by the single parameter Υx ; and one

additional parameter – Υy – has been identified.

1. isolation period Tb (from frequency ωb )

2. normalized strength coefficient η

3. rocking-to-vertical frequency ratio about x-axis Υx

4. rocking-to-vertical frequency ratio about y-axis Υy

5. torsional-to-lateral frequency ratio Ωθz

6. normalized eccentricity eb /rz , where eb = (e2bx + e2by )1/2

The isolation period Tb and the normalized strength η are by now very familiar; the remaining

parameters are interpreted and ranges identified as follows.

Rocking-to-Vertical Frequency Ratios Υx and Υy

The rocking-to-vertical frequency ratios Υx and Υy are related to the amount of rocking about

the x and y-axes, respectively. Repeating and extending the definition for unidirectional analysis

(Eq. 7.19):
ωbθx Ωθx ωbθy Ωθy
Υx = ≡ Υy = ≡ (8.16)
ωbz Ωz ωbz Ωz

198
Although Υx and Υy do not appear directly in Eq. 8.10, Sec. 7.5.2 demonstrated that for

undirectional analysis, Υx – the ratio of Ωθx to Ωz – characterized the system response better

than the independent parameters Ωθx and Ωz .

To determine the range of Υx (or Υy ), consider the parameters Ωθx (or Ωθy ) and Ωz that

enter into the calculation (Eq. 8.16). The vertical frequency ratio Ωz has been shown to be a

function of the bearing shape factor S (Eq. 7.14). Since for unidirectional analysis the variation

of Ωz had little influence on the response of the system, Ωz is selected as constant hereafter

(Ωz = 30.4 based on S = 20).

In general, the rocking-to-lateral frequency ratios Ωθx and Ωθy depend on the location and

stiffness of every bearing. However, as was shown in Sec. 7.2.4 (Eq. 7.16), their computation

may be simplified if the bearings are identical and evenly spaced on a grid in plan, and the CM

corresponds to the geometric center of the plan:


 2 1  2 1
1h 1 d2 (nby + 1) 2 2 1h 1 b2 (nbx + 1) 2 2
Ωθx = + Ω Ωθy = + Ω (8.17)
4 rx2 12 rx2 (nby − 1) z 4 ry2 12 ry2 (nbx − 1) z
where nbx and nby are the number of bearings along the x and y-directions.

Using the value of Ωz specified above, along with Eq. 8.17 for Ωθx and Ωθy , Υx and Υy

have been computed as functions of building slenderness ratio h/d for symmetric Plans S1-S9

(Table 8.1). The values of Υx for these plans are identified on a general plot of Υx as a function

of nby for slenderness ratios h/d ranging from 0.5 to 2.0 (Fig. 8.2a), which is an extension

of data that was already shown (Fig. 7.2). While Υx is independent of the plan aspect ratio

b/d, Υy is a function of both b and h (Eq. 8.17), and therefore depends indirectly on both b/d

and the slenderness ratio h/d. Therefore, Υy as a function of nbx is plotted independently for

diferent building aspect ratios b/d=1, 2, or 3 (Figs. 8.2a-c), where values for Plans S1-S9 are

again identified on the appropriate plots.

For square plans with equal number of bearings in each direction (b = d and nbx = nby ),

Υx = Υy , implying that the rocking properties are the same about the x and y-axes. This

applies to Plans S1, S4, S7, and S8 (Fig. 8.2a). For non-square plans (S2, S3, S5, S6, S9,

Fig. 8.2b,c), the rocking frequency ratios differ about the x and y-axes, and typically Υx < Υy .

That is, the building is typically more flexible in rocking about the longer x-axis (Fig. 8.1a).

199
(a) (b) (c)
2 2 2
Plans S1,2,3 Plan S2 Plan S3
Plans S4,5,6 Plan S5 Plan S6
Plans S7,9 Plan S9
1.5 Plan S8 1.5 1.5
Υx (or Υy )

h/d=0.5 h/d=0.5

Υy

Υy
1 h/d=0.5 1 1
1.0 1.51.0
1.5 2.0
1.0
2.0
1.5
0.5 0.5 0.5
2.0

0 0 0
0 1 2 0 1 2 0 1 2
10 10 10 10 10 10 10 10 10
nby (or nbx ) nbx nbx

Figure 8.2: Rocking-to-vertical frequency ratios in rectangular-plan buildings vs. number of


bearings spanning in the direction considered: (a) Υx (and Υy for square plans), (b) Υy for
b/d = 2, and (c) Υy for b/d = 3

Hence, maximum rocking angles are expected to coincide with the direction of application of the

strong component of ground motion. However, Υx may exceed Υy in squat buildings (compare

Υx and Υy for h/d = 0.5, Plans S2, S3, S5, and S6 in Fig. 8.2). In this case, calculations of

Υy relative to Υx are dominated by the increased number of bearings along the x-axis relative

to the y-axis. In general, increasing the number of bearings for given dimensions reduces the

rocking resistance by moving stiffness away from the edge of the plan.

With a maximum of 7 bearings in any direction, Plans S1-S9 do not envelop the complete

ranges of Υx and Υy , which begin to approach lower bounds at say, 20 bearings, but still continue

to decrease as the number of bearings increases up to 100 (Fig. 8.2). The computational expense

prohibits the number of bearings from being further increased for this type of parametric

analysis, but should not be a limitation since variation of the building slenderness ratio allows

Υx and Υy to vary over a wide range. The variation of Υx and Υy in non-rectangular or

asymmetric-plan buildings is similar to that described above, but does not follow any strict

rules.

200
Torsional-to-Lateral Frequency Ratio Ωθz

Only asymmetric-plan systems or symmetric systems with significant accidental torsion are

affected by variation of the torsional frequency ratio Ωθz . For systems with at least 10 bearings

(most systems), Ωθz < 1.3, and Ωθz is bounded below by 1 as the number of bearings increases

(Sec. 4.3.1, Fig. 4.4). The values of Ωθz for Plans S4-S9, used in the analysis of asymmetric-plan

systems, fall between 1.18 and 1.41 (Table 8.1).

Normalized Eccentricity eb /rz

As applied to the idealized systems, the eccentricity is limited to the x-direction, where ebx /rz is

varied from 0 to 0.15 (Sec. 4.3.1). Although not strictly applicable to axial forces, the direction of

eccentricity was shown to be irrelevant for estimating deformations in asymmetric-plan systems

(Sec. 4.3.3). Furthermore, the alignment of eccentricity with the strong-component of ground

motion (applied in the y-direction) is expected to maximize the effect of asymmetry.

8.4 Symmetric Systems

This section presents the combined lateral-torsional-rocking response of the symmetric base-

isolated block. Response quantities presented are: (1) the peak lateral deformation in any

direction measured at the CM (ubo ) (2) the peak lateral deformation in any direction maximized

over the bearings (ucorn ), thus occurring in a corner bearing, and (3) axial forces normalized

by their static forces (maximum P+ /Pst and minimum P− /Pst ), which also occur in corner

bearings. Median values of these response quantities at each corner due to the LMSR ensemble

(median PGV = 35 cm/s) are determined by the method outlined in Sec. 8.2.4, where the

weak component of excitation is applied in the x-direction and the strong component in the

y-direction. Force increments ∆P+ /Pst or ∆P− /Pst (Eq. 7.18) are also computed for statistical

analysis, where median values of maximum and minimum axial forces are derived from the

median force increments. The median deformations and force increments are maximized over

the four corners to obtain absolute peak values. Unless otherwise noted, the variable-strength

model is used for the bearings.

201
8.4.1 Median Response Spectra

Many trends that were observed for unidirectional excitation of the system (Sec. 7.5) extend

to bidirectional excitation. These trends include, but are not limited to the following, stated

without proof: The peak deformation is not affected by rocking-to-vertical frequency ratios Υx

and Υy . Positive and negative axial force increments vary nearly identically. In general, the

force increments increase with increasing normalized strength, but unsystematically at longer

isolation periods. The uncoupled nonlinear bearing model can generally be used for the bearings

in lieu of the variable strength model with little loss of accuracy, where discrepancy between

the two models is usually less than 10%. For the smallest values of η in combination with the

largest values of Υx , the uncoupled model underestimates the peak force increments by more

than 20%, but the force increment magnitude in such instances are small.

8.4.2 Accidental Torsion

Axial forces on the bearings vary in time due to rocking, which, together with axial-load effects

in the variable-strength model, cause the stiffnesses and strengths of individual bearings to

vary in time. This leads to the possibility of torsion in nominally symmetric buildings – usually

known as accidental torsion – when the excitation is bidirectional. The stiffness and strength

variation can induce a dynamically-varying eccentricity in the direction of ground motion, thus

causing a torsional response for excitation in the transverse direction.

However, the resulting accidental torsion turns out to be insignificant, as demonstrated

in Fig. 8.3 where the peak deformations compared at the corner, ucorn , and CM, ubo , are

essentially identical. This conclusion is seen to be true for all combinations of Tb and η, even

though the buildings considered are slender, making them susceptible to overturning and larger

stiffness variations. With Υx and Υy lowest of all systems considered, Plan S8 (Fig. 8.3b) is

most likely to be affected by overturning and hence accidental torsion, but is not. This parallels

the conclusion for FP systems [60] that accidental torsion in the isolation system as a result of

axial load variation is inconsequential.

202
(a) (b)
15 15

η=0.25 η=0.25
Lateral deformation (cm)

10 0.5 10 0.5
0.75 0.75
1.0 1.0

1.5 1.5
5 5

ubo
ucorn
0 0
1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 8.3: Median response spectra for peak corner deformation ucorn compared to peak
deformation at CM ubo for (a) Plan S1 and (b) Plan S8 with slenderness ratio h/d = 2; leads
to Υx = 0.78 and 0.55, Υy = 0.78 and 0.55, respectively.

8.4.3 Influence of Rocking-to-Vertical Frequency Ratios

The axial force extrema (P± /Pst ) are influenced by the rocking-to-vertical frequency ratios

Υx and Υy , but not independently by other rocking-related parameters such as the building

slenderness ratio h/d, the plan aspect ratio b/d, and the number of bearings nb . The single

parameter most closely correlated to the force extrema is Υx , and only minor scatter appears

in the maximum axial force P+ /Pst when plotted against Υx (Fig. 8.4a and c). The scatter

occurs among systems with the same or similar values of Υx , but different h/d, b/d, nb , etc.

Some of this scatter can be accounted for by the additional parameter Υy , which varies

with plan aspect ratio b/d when the remaining properties are held constant. Although the

same axial force data considered above shows considerable more scatter when plotted against

Υy instead of Υx (Fig. 8.4b and d), Υy does provide meaningful information. Consider that for

similar values of Υx , the maximum axial forces tend to be largest for square plans (b/d = 1)

and smallest for plans with b/d = 3 (Fig. 8.4a and c). Since Υy is closely related to plan aspect

ratio, it follows that for Υx held constant the axial forces should have a one-to-one relation to

Υy . Lines of constant Υx in Fig. 8.4b and d show that this is sometimes but not always the

case. As will be seen, Υx and Υy , taken together, provide adequate information to reasonably

203
(a) (b)
2 2
b/d=1 b/d=1
b/d=2 b/d=2
1.8 1.8
b/d=3 b/d=3
Υx const.
1.6 1.6
P+ /Pst

1.4 1.4

1.2 1.2

1 1
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Υx Υy

(c) (d)
2.4 2.4
b/d=1 b/d=1
2.2 b/d=2 2.2 b/d=2
b/d=3 b/d=3
2 2 Υx const.
P+ /Pst

1.8 1.8

1.6 1.6

1.4 1.4

1.2 1.2

1 1
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Υx Υy

Figure 8.4: Maximum axial force P+ /Pst plotted against rocking-to-vertical frequency ratios
(a,c) Υx and (b,d) Υy , where (a,b) Tb = 2 sec, η = 0.5 and (c,d) Tb = 4 sec, η = 1.0. Variations
in Υx and Υy are comprised by data from Plans S1-S9 and varying slenderness ratios.

predict axial force extrema in symmetric systems.

8.4.4 Increase in Response due to Bidirectional Excitation

The complete response of the system subjected to bidirectional excitation, which includes rock-

ing about both lateral axes, is compared to the response when the same system is subjected

to unidirectional excitation. The peak deformation of the bearings in the y-direction, in which

the strong-component ensemble is applied, is essentially unaffected by adding the transverse

component of excitation (Fig. 8.5a), which was true when rocking was neglected (Sec. 3.7). The

peak deformation in any direction, on the other hand, experiences a 10-20% increase when the

excitation is bidirectional (Fig. 8.5b).

204
(a) (b)
15 15

η=0.25 η=0.25
10 10

ucorn (cm)
ubyo (cm)

0.5 0.5
0.75 0.75
1.0 1.0
5 1.5 5 1.5

unidirectional unidirectional
bidirectional bidirectional
0 0
1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec)

Figure 8.5: Median deformations (a) ubyo in the y-direction and (b) ucorn – peak in any direc-
tion – for bidirectional analysis compared to the median deformation ubo from unidirectional
analysis; Υx = Υy = 0.78.

Unlike deformations, increases in bearing axial force extrema due to bidirectional excita-

tion depend on plan layout, and are greatest in square plans (b/d = 1) and minimal in narrow

plans (b/d → ∞). These trends are demonstrated in Fig. 8.6, where the maximum axial force

P+ /Pst is seen to increase – relative to unidirectional excitation – most for Plan S1, which is

square, less for Plan S2 (b/d = 2), and only slightly for Plan S3 (b/d = 3). This reflects the

tendency of a narrow building to rock mostly about its longer axis corresponding to its lower

rocking frequency ratio with minimal influence from the second component of excitation, while

a square plan with equal rocking frequencies about each axis (Υx = Υy ) is influenced by both

components of excitation.

8.4.5 Design Equations to Estimate Peak Response

Comprehensive design equations are developed to estimate the peak responses – deformation

ucorn and maximum and minimum axial forces P± /Pst – in symmetric isolation systems that

account for bidirectional excitation and rocking about both axes. These equations were devel-

oped by regression analysis of the normalized deformations (ūcorn ) and forces (P̄± /Pst ), which

were computed by RHA of Eq. 8.10 to each pair of ground motions in the LMSR ensemble.

Equation 8.10 is solved for building Plans S1-S9 with the variable-strength model for the bear-

205
(a) Plan S1 (b) Plan S2 (c) Plan S3
2 2 2
unidirectional unidirectional unidirectional
bidirectional bidirectional bidirectional
1.8 1.8 1.8

1.6 1.6 1.6


P+ /Pst

1.5 1.5 1.5


1.4 1.0 1.4 1.0 1.4 1.0
0.75 0.75 0.75
0.5 0.5 0.5
1.2 η=0.25 1.2 η=0.25 1.2 η=0.25

1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.6: Maximum axial force P+ /Pst compared for bidirectional and unidirectional analy-
ses. For slenderness ratio h/d = 2 and Υx = 0.78, results shown for (a) Plan S1 (b/d = 1,
Υy = 0.78), (b) Plan S2 (b/d = 2, Υy = 1.00), and (c) Plan S3 (b/d = 3, Υy = 1.07).

ings, over the ranges Tb from 2 to 4.5 seconds, η from 0.25 to 1.5, building slenderness ratio

h/d from 0.5 to 2.0, and vertical frequency ratio Ωz = 30.4, which leads to rocking properties

Υx , Υy and Ωθz (Fig. 8.2, Table 8.1). Because the deformations and axial forces are essentially

independent of Ωz (Sec. 7.5.2), the design equations are also applicable to other values of Ωz .

Using methods that have been fine-tuned in previous chapters, the following equation

was developed to estimate the normalized peak deformation ucorn :

ūcorn = 2.30Tb−1.82 η −1.43 (8.18)

and is multiplied by u∗y (Eq. 3.6) – a function of Tb , η and median or design ground velocity

u̇gyo – to give the peak deformation ucorn in any direction:

7.01 0.18 −0.43


ucorn = T η u̇gyo (8.19)
4π 2 b

The regression coefficients for Tb and η determined from bidirectional response analysis were

found to be nearly identical to those determined earlier from unidirectional analysis (Sec. 7.6),

and were thus constrained to be the same. As a result, Eq. 8.19 varies from its one-directional

counterpart (Eq. 7.21) by a constant factor of about 1.13, which is identical to the result ignoring

rocking (Sec. 3.7). Despite the complexities of the system model and the possible variations in

parameters, the peak deformation can be described by a relatively simple equation (Eq. 8.19),

206
and the values for bidirectional and unidirectional analysis vary by not only a constant, but the

same constant as when rocking of the system was neglected entirely! Demonstrated for three

different plan layouts (Fig. 8.8a-c), Eq. 8.19 is a good approximation to the median deformation

ucorn .

Equations to estimate both the positive and negative force increments ∆P+ /Pst and

∆P− /Pst (deviation of force extrema from the static force, Eq. 7.18), which were not found to

be statistically distinct when averaged over many ground motions, were developed by methods

presented previously. That is, force increments rather than force extrema were used for statis-

tical analysis (Sec. 7.4.2); and equations were fit to the normalized force increment ∆P̄ /Pst ,

and multiplied by u∗y (Sec. 7.4.1) to obtain the actual force increment.

Two equations to estimate the force increment ∆P/Pst are presented. The first is given

as:
∆P 0.54
= f (Tb , η)Υx−1.66 Υy−0.54 u̇gyo (8.20)
Pst 4π 2
where
(−1.23+0.59 ln Tb )
f (Tb , η) = Tb η (0.05+0.24 ln η+0.65 ln Tb ) (8.21)

and the axial force extrema P+ /Pst and P− /Pst are computed by:

P+ /Pst = 1 + ∆P/Pst (8.22a)

P− /Pst = 1 − ∆P/Pst (8.22b)

Equation 8.20, substituted into Eq. 8.22, is adequate to estimate the force extrema in symmetric

systems, as demonstrated in Fig. 8.7 for three different values of Υx and Υy . If this were the

final goal, no further improvement would be necessary. However, Eq. 8.20, when extended

to include the effect of eccentricity, was found to give poor estimates of the force extrema in

asymmetric-plan systems. For this reason the following improved equation is presented for

symmetric-plan systems:
 0.48  0.28
∆P 0.46 ymax dyz xmax dxz
= f (Tb , η)Υx−1.42 Υy−0.68 u̇gyo (8.23)
Pst 4π 2 rx rx ry ry
where f (Tb , η) is given in Eq. 8.21, xmax and ymax are the x and y-direction components of

the distance from the CM to the outlying bearing; similarly dyz = (h/2)2 + e2by and dxz =

207

(h/2)2 + e2bx are the components in the y-z and x-z planes, respectively, of the distance from

the CM to the CR of the system. Recall also that h is the height of the building, and rx and

ry are radii of gyration about the x and y-axes. For symmetric-plan systems, the improvement

in Eq. 8.23 over Eq. 8.20 is marginal, as demonstrated in Figs. 8.8 and 8.7, which depict the

accuracy of the respective design equations by comparison with the exact median.
   
dyz
8.4.6 Interpretation of Regression Parameters ymax rx rx and xmax dxz
ry ry

The improvement in Eq. 8.23 for asymmetric-plan systems relative to Eq. 8.20 is due to the

additional regressors (explanatory variables): (ymax /rx · dyz /rx ) and (xmax /ry · dxz /ry ). These

“rocking” regressors can be shown to be analogous to the torsional regressor (1 + c/rz · eb /rz ),

which was applied to account for the increase in bearing deformation in an asymmetric-plan

system relative to a symmetric system (Sec. 4.5, Appendix 4.1). From a simplistic point-of-

view, the above regressor in the torsional problem related to motion or rotation about the

z-axis; hence similar regressors relating to rotation about the x or y-axes, could be expected

to contribute in predicting the rocking response. Because bearing axial force, the quantity to

be predicted, is in general proportional to axial deformation, and bearing axial deformation is

a component of the rotation about either the x or y-axis, any parameter that can help predict

the rotation of the system about these axes will have some value in predicting bearing axial

forces.

An obvious difference between the two rocking regressors and the torsional regressor is

the additive factor of 1 in the torsional regressor. This factor is not needed for rocking; while

the torsional regressor relates to the increase in deformation contributed by torsion, rocking is

assumed to be the only contributor to the axial force increments. We now select one of the

rocking regressors (ymax /rx · dyz /rx ) and demonstrate its analogy to c/rz · eb /rz . First, eb is the

well-known eccentricity or distance from the CM to CR in the x-y plane, and dyz represents the

same distance in the y-z plane, because the structure CM is vertically h/2 from the isolation

system CR. Second, c is the absolute distance from the CM to the outlying bearing in the x-y

plane, while ymax is the y-component of this distance in the y-z plane. (The vertical component

208
(a) (b) (c)
20 20 20

15 15 η=.25 15
η=.25 η=.25
ucorn (cm)

100.5 10 0.5 10 0.5


0.75 0.75 0.75
1.0 1.0 1.0
5 1.5 5 1.5 5 1.5

0 0 0
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

2.5 2.5 2.5


Exact median
Design equation
2 2 2
P+ /Pst

1.5 1.5
1.5 1.0 1.0
1.0 0.75 0.75
1.5 0.75 1.5 0.5 1.5 0.5
0.5 η=.25 η=.25
η=.25
1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

1 1 1
η=.25
0.5 η=.25 η=.25
0.5 0.75 0.5 0.5 0.5 0.5
P− /Pst

1.0 0.75
1.5 1.0 0.75
1.5 1.0
0 0 0 1.5

−0.5 −0.5 −0.5


1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.7: Design equation for bearing deformation ucorn (Eq. 8.19) and axial forces P+ /Pst
and P− /Pst (Eqs. 8.20 and 8.22), compared to the exact median by RHA to the LMSR ensemble,
for: (a) Plan S1, Υx = Υy = 0.78, (b) Plan S5, Υx = 0.63, Υy = 0.87 and (c) Plan S9, Υx = 0.58,
Υy = 0.82; all based on slenderness ratio h/d = 2.

209
(a) (b) (c)
2.5 2.5 2.5
Exact median
Design equation
2 2 2
P+ /Pst

1.5 1.5
1.5 1.0 1.0
1.0 0.75 0.75
1.5 0.75 1.5 0.5 1.5 0.5
0.5 η=.25 η=.25
η=.25
1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

1 1 1
η=.25
0.5 η=.25 η=.25
0.5 0.75 0.5 0.5 0.5 0.5
P− /Pst

1.0 0.75
1.5 1.0 0.75
1.5 1.0
0 0 0 1.5

−0.5 −0.5 −0.5


1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.8: Improved design equation for axial forces P+ /Pst and P− /Pst (Eqs. 8.23 and 8.22),
compared to the exact median by RHA to the LMSR ensemble, for (a) Plan S1, Υx = Υy = 0.78,
(b) Plan S5, Υx = 0.63, Υy = 0.87 and (c) Plan S9, Υx = 0.58, Υy = 0.82 based on slenderness
ratio h/d = 2.

is dropped from this factor since it is parallel to the axial deformation that is relevant.) Finally,

since both eb and c in the torsional regressor are normalized by the applicable radius of gyration

rz , normalizations by the comparable radius rx are applied to the rocking regressor. The same

analogy extends to the second regressor (xmax /ry · h/2ry ) for rocking about the y-axis.

For symmetric buildings, it can be shown that the rocking regressors vary over a fairly

narrow range, i.e., with values ranging from 1.2 to 1.5. Furthermore, their values are not

independent of, but closely related to, the frequency ratios Υx and Υy , respectively (Eq. 8.17).

For these reasons, the additional regressors were not identified in Sec. 8.3.3 as parameters that

influence the response of the system. However, as already mentioned, their inclusion in Eq. 8.23

led to a substantial improvement in the accuracy of estimated axial forces for asymmetric-plan

buildings.

210
8.5 Asymmetric-Plan Systems

Next, response trends are studied and design equations are developed for asymmetric-plan sys-

tems. Of particular interest is the response of asymmetric-plan systems relative to comparable

symmetric systems. For this purpose, the corresponding symmetric system, first introduced

in Sec. 4.2.2, has properties Tb , η, Υx , Υy , and Ωθz identical to the asymmetric system, but

eccentricity eb /rz = 0. This definition does not uniquely determine a corresponding symmetric

system, which for a general asymmetric-plan building may be difficult to define.

Recall that plan-asymmetry is studied here by varying the stiffnesses and strengths of in-

dividual bearings to introduce eccentricity into a rectangular-plan system; this idealized system

approach was implemented in Plans S4-S9. The corresponding symmetric system to an ideal-

ized asymmetric system has the same plan layout without the variation of individual bearing

properties that induced the eccentricity; this simple means of comparison can be seen as an

advantage to the idealized system approach.

The median peak corner deformation ucorn and force extrema (P+ /Pst and P− /Pst ) over

all bearings are computed as described in Sec. 8.4 for symmetric systems. For comparison

of asymmetric-plan and symmetric systems, the deformation and force-increment ratios are

defined as:

(ucorn )A (∆P+ /Pst )A (∆P− /Pst )A


ûcorn = ∆P̂+ = ∆P̂− = (8.24)
(ucorn )S (∆P+ /Pst )S (∆P− /Pst )S

where the subscripts A and S refer to the asymmetric and corresponding symmetric systems, re-

spectively. Consistent with Sec. 7.4.2, force-increment ratios were found to be more meaningful

than ratios of the actual forces.

8.5.1 Trends for Deformation and Force-Increment Ratios

Median values of both deformation and force-increment ratios appear to increase linearly with

increasing eccentricity ebx /rz (Fig. 8.9). The slope of increase in these ratios varies widely, as

observed in the spreading of data in Fig. 8.9 with increasing eccentricity, but this variation was

found to be unsystematic with respect to other parameters (Tb , η, Υx , Υy ). Thus, eccentricity is

the only parameter that has been linked to increases in both deformations and force extrema in

211
(a) (b) (c)
1.4 1.4 1.4

1.3 1.3 1.3

1.2 1.2 1.2


ûcorn

∆P̂+

∆P̂−
1.1 1.1 1.1

1 1 1

0.9 0.9 0.9


0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
ebx /rz ebx /rz ebx /rz

Figure 8.9: (a) Deformation ratio ûcorn and (b,c) force increment-ratios ∆P̂+ and ∆P̂− with
varying eccentricity ebx /rz . Not meant to be distinguishable, the lines represent different values
of Tb , η, Υx , and Υy applied to Plans S4-S9.

asymmetric-plan systems. This linear increasing relation breaks down slightly for the negative

force-increment ratio ∆P̂− (Fig. 8.9c), as values of this ratio can fall below unity in the small

eccentricity range (eb /rz  0.05) before resuming the increasing trend.

Actual peak deformations and force extrema, rather than their ratios, are compared for

a symmetric system (Plan S5) and idealized systems with several different values of ebx /rz in

Fig. 8.10. As depicted here, the increase in the deformation or force extrema for the largest

eccentricity considered, typically 10-20% (Fig. 8.9), does not seem that significant. Adjusting

the normalized strength η, for instance, from 0.5 to 1.0 has much greater consequence for the

response of the isolation system (Fig. 8.10). Nevertheless, plan asymmetry provides no benefit

to the system, and should be minimized if xspossible.

8.5.2 Design Equations to Estimate Peak Response

Since deformation and force-increment ratios were found to vary only with eccentricity (Fig. 8.9),

Eqs. 8.19 and 8.23, which estimate lateral deformations and axial force increments in symmetric

systems, can be extended to asymmetric-plan systems by multiplying by a factor that accounts

for eccentricity.

The largest lateral deformations in asymmetric-plan systems occur in exterior or corner

bearings due to torsion. The normalized distance c/rz from the CM to the outlying bearing

212
(a) (b) (c)
15 2 1

η=0.5 1.8 0.8


η=1.0
10
η=1.0
ucorn (cm)

1.6 0.6

P+ /Pst

P− /Pst
η=0.5
η=0.5
1.4 0.4
ebx /rz = 0
5 η=1.0
0.025
0.05
0.075 1.2 0.2
0.1
0.15
0 1 0
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.10: For a symmetric system and several different eccentricities ebx /rz , median re-
sponse spectra for (a) peak corner deformation ucorn , and axial forces (b) P+ /Pst and (c)
P− /Pst ; Plan 5 (Υx = 0.63, Υy = 0.87, Ωθz = 1.27) based on slenderness h/d = 2.

varies in actual asymmetric-plan systems and should be accounted for in the design equation.

When lateral-torsional coupling alone was considered, the peak deformation maximized over a

circle of radius c was shown to be a good upper bound estimate to the peak deformation at

some exterior bearing of distance c from the CM (Sec. 4.5.2), and a similar assumption is made

here to extend to systems with rocking. The peak deformation sampled at varying distances

from the CM in the idealized system serves as data to develop a design equation applicable to

general asymmetric plans. To this end, equations were fit to the deformation ratio ûco :

uco
ûco = (8.25)
(ucorn )S

which is the ratio of peak deformations in an asymmetric-plan relative to a corresponding

symmetric system – maximized over a circle of radius c – and an upper bound to the peak

deformation ratio of any set of bearings located along that circle.

Regression analysis of data from the idealized systems (Plans S4-S9) and corresponding

symmetric systems led to an equation for ûco and hence the following design equation for the

213
total peak deformation in an asymmetric-plan system:

uco = (ucorn )S ûco


 
c eb 0.7
= (ucorn )S 1 + (8.26)
rz rz

with u(corn )S computed from Eq. 8.19. The regression model in Eq. 8.26 was chosen to be the

same as when rocking was omitted (Sec. 4.5.2), and as required, reduces to the deformation in

a symmetric system when eb /rz = 0. Even with all the improvements to the model, Eq. 8.26

differs from the original estimate of deformation in asymmetric-plan systems excluding rocking

and vertical deformations (Eq. 4.25) only slightly. The largest discrepancy is in the coefficient

for the eccentricity term, 0.7 here compared to 0.88 in Eq. 4.25, which has only a small effect

on the total estimated deformation.

Because the distance to outlying bearings has already been accounted for in symmetric

systems (Eq. 8.23, Sec. 8.4.6), regression analysis of the forces in the idealized asymmetric

and corresponding symmetric systems led to the following design equation for the peak force

increment in an asymmetric system:


   
∆P ∆P
= ∆P̂
Pst Pst S
A
   
∆P eb
= 1+ (8.27)
Pst S rz

with (∆P/Pst )S defined in Eq. 8.23. The regression coefficient for (1 + eb /rz ) was rounded from

1.01 to 1, which is consistent with the observed linear variation of force-increment ratios with

eb /rz (Fig. 8.9). As required, the force increments (Eq. 8.27) reduce to those of a symmetric

system when eb /rz = 0. (Actual forces are obtained by Eq. 8.22 with Eq. 8.27 for ∆P/Pst .)

Although analysis of the data for the force increment ratio ∆P̂ resulted in statistically distinct

coefficients for ∆P̂+ and ∆P̂− , separate coefficients were not adopted, as they would have

resulted in only inconsequential changes to the total estimated force increments (Eq. 8.27).

Also, possible correlation of ∆P̂ and other parameters (Tb , η, Υx , Υy and Ωθz ) was considered

by testing different regression models, but the influence of these additional parameters was

found to be negligible.

214
Table 8.2: Properties of Asymmetric-Plan Systems

Plan No. A1 A2 A3
ebx/rz 0.049 0.108 0
eby /rz 0.049 0 0.057
Ωθz 1.180 1.280 1.172
h 3t 6t d 2d 0.5d d
Υx 0.95 0.66 0.80 0.47 1.09 0.79
Υy 0.95 0.66 1.08 0.80 1.05 0.90

For asymmetric-plan systems, Eqs. 8.26 and 8.27 represent the best estimates of the

peak bearing deformation and axial-force increments (maximum or minimum) that consider

all modeling complexities introduced in later chapters, including rocking with bidirectional

excitation and axial-load effects. Because Eq. 8.26 for peak lateral deformation differs from

the equation that neglects rocking entirely (Eq. 4.25) only slightly, it is typically unnecessary

to include rocking and axial-load effects in the isolated-structure model to obtain a reasonable

estimate of the peak bearing deformation. On the other hand, rocking must be included to

obtain any estimate of axial forces.

8.5.3 Application to Three Asymmetric-Plan Systems

The accuracy of these design equations (Eqs. 8.26 and 8.27), developed from response data

for idealized systems, is investigated for three asymmetric-plan systems, defined formerly in

Sec. 4.3 (Fig. 4.5a), whose properties are listed in Table 8.2. The eccentricities (ebx /rz , eby /rz )

and torsional frequency ratios (Ωθz ) of these plans are repeated here. The values listed for Υx

and Υy are based on two different assumed building heights h for each plan, which correspond

approximately to slenderness ratios h/d of 1.0 and 2.0.

For strength coefficients µ = 0.05 and 0.1, each system in Table 8.2 was subjected to the

LMSR ensemble with median PGV of 35 cm/s and ωd = 3.05, which led to normalized strengths

η = 0.46 and 0.92 (Sec. 4.5.3). The deformation of each corner bearing was computed two ways:

(1) as the median (determined by nonlinear RHA) over the LMSR ensemble and (2) estimated

from Eq. 8.26; and then maximized over all corners for both. Forces were treated similarly,

215
except since the system can rock about any axis, bearings located around the perimeter were

tested for occurrence of the largest axial forces. For each asymmetric plan, the exact median

deformation and maximum axial force are compared to the corresponding design equation in

Fig. 8.11. The minimum axial force, which has been omitted in Fig. 8.11, showed similar trends.

Equation 8.26, which differs for each plan only due to differing values of eccentricity, is

a good estimate of the exact (median) peak deformation over all corners (Fig 8.11a). Equa-

tion 8.26, based on the deformation maximized over a circle of radius rz and hence an upper

bound to the corner deformation, is not necessarily conservative as it was for systems without

rocking (Table 4.1.) Nevertheless, the discrepancy between the exact median and the design

equation is obviously small.

On the other hand, Eq. 8.27 is not always a good estimate of the exact axial forces

(Fig. 8.11b,c). While this design equation can provides reasonable force estimates (Fig. 8.11b),

for smaller values of Υx and Υy typical of more slender buildings (Fig. 8.11c) it noticeably

underestimates the positive axial force for two of the three plans (A1 and A3). This lack-of-fit

also seems to be worse for η = 0.92 than for η = 0.46. The visual impression of Fig. 8.11b,c is

validated by Fig. 8.12, where the percent discrepancy in Eq. 8.27 relative to the exact median

is plotted for the positive force increment. This discrepancy ranges in value from -25% to 25%,

and is often greater than 10% in magnitude. Also, increases in Υx , Υy do not necessarily lead

to reduced percent error; Fig. 8.11 may be deceptive since the force increments corresponding

to the larger Υ are small in magnitude.

Unfortunately, errors as large as 20%, especially on the unconservative side, imply that

Eq. 8.27 has limited value in estimating the axial forces in bearings. Goodness of fit measures

and error analysis (Fig. 8.13) together indicate that the design equations can reasonably predict

forces in the idealized systems by which they were developed. Specifically, the design equations

appear to estimate the maximum axial force pretty accurately for a range of idealized systems

(Fig. 8.13a-b), and percent errors are typically held within 10% (Fig. 8.13c). Thus, it would

appear that the difficulties mentioned in Sec. 8.3.1 prevent successful application of the idealized

system approach; and a good fit cannot be found because idealized systems are no longer

216
(a)
15 15 15
Plan A1 Plan A2 Plan A3
η=0.46 η=0.46 η=0.46
10 10 10
uco (cm)

η=0.92 η=0.92 η=0.92

5 5 5
Υx =0.95 Υx =0.80 Υx =1.09
Υx =0.66 Υx =0.47 Υx =0.79
Design equation Design equation Design equation
0 0 0
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
(b)
2.5 2.5 2.5
Plan A1, Υx =0.95 Plan A2, Υx =0.80 Plan A3, Υx =1.09

2 η=0.46 2 2
P+ /Pst

η=0.92

1.5 1.5 1.5

1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
(c)
2.5 2.5 2.5
Plan A1, Υx =0.66 Plan A2, Υx =0.47 Plan A3, Υx =0.79
η=0.46
2 2 2 η=0.92
P+ /Pst

1.5 1.5 1.5

1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.11: For asymmetric plans A1-A3 (Table 8.2), comparison of exact median and design
equation for (a) peak bearing deformation uco (Eq. 8.26); and (b,c) maximum axial forces
(Eq. 8.27), separated for different Υx . In each case, a dotted line for the design equation is
shown next to a solid line for the exact median.

217
(a) Plan A1 (b) Plan A2 (c) Plan A3
30 30 30
Υx =0.95, η=0.46 Υx =1.09, η=0.46
Υx =0.95, η=0.92 Υx =1.09, η=0.92
20 20 20
Υx =0.66, η=0.46 Υx =0.79, η=0.46
% Discrepancy in ∆P+

Υx =0.66, η=0.92 Υx =0.79, η=0.92


10 10 10

0 0 0

−10 −10 −10


Υx =0.80, η=0.46
Υx =0.80, η=0.92
−20 −20 −20
Υx =0.47, η=0.46
Υx =0.47, η=0.92
−30 −30 −30
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.12: Percent discrepancy in the design equation (Eq. 8.27) for positive axial force
increment ∆P/ Pst compared to its exact median for asymmetric plans (a) A1, (b) A2, and (c)
A3.

fully representative of general asymmetric-plan systems when rocking and axial-load effects are

included in the analysis. These results seem to suggest that the best way to interpret rocking

effects in asymmetric-plan systems is through detailed modeling and analysis on a case-by-case

basis, and applicability of any simplified code procedure should reflect this.

8.6 Conclusions

To accurately determine the controlling response of the isolation system, ground excitation in

both lateral directions and associated three-dimensional effects (lateral-rocking-torsion interac-

tion) of the system should be considered. For this purpose, a previously developed model for

lead-rubber bearings that includes axial-load effects, or lateral vertical coupling, was enhanced

to accommodate motion in two lateral directions. The improved bearing model was incorpo-

rated into symmetric and asymmetric-plan base-isolated building models, and relevant peak

responses were determined by response history analysis to the LMSR strong ground motion en-

semble. Design equations were developed to estimate both peak lateral deformation (Eqs. 8.19

and 8.26) and peak axial forces (Eqs. 8.23 and 8.27) in the bearings for both symmetric and

asymmetric-plan systems, respectively. In an attempt to obtain results widely applicable to gen-

eral asymmetric-plan systems, an idealized asymmetric system was analyzed, where eccentricity

218
(a)
2.5 2.5 2.5
η=.25 Plan S5 Plan S7 Plan S9
0.5
2 0.75 2 2
P+ /Pst

1.0
1.5 1.5 1.5 1.5

1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
(b)
2.5 2.5 2.5
Plan S5 Plan S9
2 2 2
P+ /Pst

1.5 1.5 1.5


Plan S7
1 1 1
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

(c)
% Discrepancy in ∆P+

30 30 30
20 Plan S5 Plan S7 20
Plan S9
20
10 10 10
0 0 0
−10 −10 −10
−20 −20 −20
−30 −30 −30
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Tb (sec) Tb (sec) Tb (sec)

Figure 8.13: Design equation for positive axial forces P+ /Pst (Eqs. 8.27 and 8.22) compared
to the exact median by RHA to the LMSR ensemble, for idealized plans S5, S7 and S9, based
on: (a) h/d=1 and (b) h/d=2; (c) percent discrepancy in the design equation (Eq. 8.27) for the
above cases, with h/d=1 (solid lines) and h/d=2 (dashed lines).

was introduced by varying the stiffnesses and strengths of individual bearings in a rectangular

plan. This investigation, the culmination of work in previous chapters, supports the following

conclusions:

1. Axial-load effects in isolation bearings could theoretically lead to accidental torsion in

symmetric buildings. Rocking of the structure causes variation of the axial loads, which

in turn cause the stiffnesses and strengths of individual bearings to vary and can induce

a time-varying eccentricity. However, accidental torsion in the isolation system from this

source has been found to be insignificant.

219
2. The peak isolator deformation can generally be determined to sufficient accuracy using

a single degree-of-freedom model of the system, as this deformation was only modestly

influenced by refinements to the global model of the structure or individual bearing mod-

els. It was already concluded in Chapter 7 that even if rocking is neglected entirely,

response spectra for lateral deformation are within 10% of those when rocking and axial-

load effects are included. In extending these results to consider bidirectional excitation

and plan-asymmetry, remarkable parallels are drawn between systems with and without

rocking. In both cases, the design equation for deformation is amplified by 13% for bidi-

rectional excitation. Similar functions of eccentricity were used to account for effects of

plan-asymmetry in systems with and without rocking. This ability to simply and ac-

curately predict the isolator deformation may be lost if structural flexibility, neglected

here, should happen to significant; however, given estimates of deformation should be

conservative.

3. Unlike lateral deformation, the variation of axial forces in the isolators is directly related

to rocking, and can be closely correlated to the rocking-to-vertical frequency ratios Υx

and Υy for rocking about the x and y-axes. The maximum and minimum axial forces

in symmetric systems are influenced by these frequency ratios, but not independently by

other rocking-related parameters such as the building slenderness ratio, the plan aspect

ratio, or distribution of bearings over the plan. The increase in axial force extrema for

bidirectional excitation is largest in square plans and negligible in narrow plans, and is

indirectly accounted for by Υy .

4. An anology was drawn between rocking and torsion, where in both cases, rotation about

the center of mass axis contributes to the response of interest, which can be computed

as a function of radial distance from this center of mass. Using this comparison to tor-

sion, additional regression parameters were identified for variation of axial forces due to

rocking, and inclusion of these parameters led to some improvement in the design equa-

tions, particularly for asymmetric-plan buildings. However, the majority of the torsional

220
effects were accounted for by the eccentricity while the majority of the rocking effects

were accounted for by the rocking frequency ratios.

5. The design equations for axial forces are complex with many factors and coefficients that

some of the attraction of the method as originally applied to deformations has been lost.

Furthermore, while the design equations to estimate axial forces are accurate for sym-

metric systems and for the idealized asymmetric systems for which they were developed,

they are shown to err by up to 25% when applied to actual asymmetric-plan systems.

Thus, the idealized system approach seems to have reached its limit of application, i.e.,

idealized systems are not sufficiently representative of general asymmetric-plan systems

when rocking and axial-load effects are included. Alternatively, design axial forces in

asymmetric-plan buildings should be evaluated case-by-case.

221
Appendix 8A: Notation

b maximum dimension or length of symmetric-plan building

b/d length-to-width ratio, or plan aspect ratio

c distance from CM to outlying corner

C damping matrix of isolated block

C damping matrix of system using uncoupled nonlinear bearings

Cz , Cz damping submatrix for rocking and vertical deformations

cbzi viscous damping coefficient of single bearing in z-direction

d minimum dimension or width of symmetric-plan building

ebx , eby stiffness eccentricities in x and y-directions

eb total magnitude of stiffness eccentricity: (e2bx + e2by )1/2

ebxy additional eccentricity representative of coupling between the two

rocking deformations

fbxi , fbyi lateral forces of ith bearing in x and y-directions

fDi vertical damping force of ith bearing

h building height

hb total bearing height including steel and rubber layers

h/d building slenderness (height-to-width) ratio

K stiffness matrix of system using uncoupled nonlinear bearings

kb lateral postyield stiffness of isolation system, summed over all bear-

ings

kbi nominal postyield stiffness of single bearing: GAs /hb

kbz vertical stiffness of isolation system, summed over all bearings

kbzi nominal vertical stiffness of single bearing: Ec As /hb

kbθx , kbθy rocking stiffnesses for rotations about x and y-axes

kbθz torsional stiffness for rotation about z-axis

m mass of isolated block

nb total number of bearings

222
nbx , nby number of bearings across the plan in the x and y-directions, for

symmetric systems

PE Euler buckling load, neglects shear stiffness of the bearing: π 2 EIs /h2b

Pi axial or vertical force in ith bearing (positive in compression)

P+ maximum axial force observed in any bearing

P− minimum axial force observed in any bearing (negative if tensile)

∆P+ , ∆P− maximum and minimum axial force increments; maximum force de-

viation in either direction from static force

∆P̂+ , ∆P̂− ratio of maximum and minimum axial force increments in

asymmetric-plan relative to symmetric-plan system

Po value of strength used to calibrate variable-strength model, typically

about 1/3 of the bearing static load

Pst static load, gravity load, or bearing design axial load

Q lateral characteristic strength of isolation system, summed over all

bearings

Qi nominal characteristic strength of a single isolator

rx , ry , rz radii of gyration about x, y, and z-axes

sxi , syi deformations of shear spring (in two-spring model of ith bearing) in

x and y-directions

S bearing shape factor; D/4t for a single bearing

Tb isolation period, corresponds to nominal postyield stiffness of isola-

tion system

ubxi , ubyi , ubzi lateral deformation of ith bearing in x and y-directions

ubo peak lateral deformation at CM in any direction

ubyo peak lateral deformation in the y-direction in a symmetric system

ucorn peak deformation at outlying corner in any direction

ûcorn peak deformation ratio at outlying corner: (ucorn )A /(ucorn )S

uco peak deformation in any direction maximized at a circle of radius c

from CM; an upper bound to ucorn


223
ûco peak deformation ratio at circle of radius c: uco /(ucorn )S

ügx , ügy x and y-components of ground acceleration

ux , uy , uz deformations of the CM in the x, y and z-directions

u∗y Q/kb = µg/ωb2

vi vertical deformation of the ith bearing due to axial flexibility

xmax , ymax distance from CM to outlying bearing in x and y-directions

zxi , zyi x and y-components of yield function, or dimensionless plastic vari-

able governing yielding of the ith bearing

γi ratio of nominal strength in single bearing to total strength: Qi /Q

ζθx , ζθy damping ratios for rocking about x and y-axes

ζz damping ratio for vertical motion

η strength of isolation system normalized by peak ground velocity:

µg/(ωd u̇gyo )

θx , θy , θz rotations of isolated block about x, y and z-axes

θxi , θyi rotation of bearing about x and y-axes in two-spring model

µ characteristic strength ratio of isolated block: Q/(mg)

Υx , Υy rocking-to-vertical frequency ratios: Ωθx /Ωz , Ωθy /Ωz

ωb lateral frequency of isolation system, corresponding to postyield stiff-

ness

ωbz vertical frequency of isolation system

ωbθx , ωbθy rocking frequencies about x and y-axes

ωbθz torsional frequency of isolation system for rotation about z-axis

Ωθx , Ωθy rocking-to-lateral frequency ratios: ωbθx /ωb , ωbθy /ωb

Ωz vertical-to-lateral frequency ratio: ωbz /ωb

Ωθz torsional-to-lateral frequency ratio: ωbθz /ωb

ωd corner frequency separating velocity and displacement-sensitive spec-

tral regions

224
REFERENCES

[1] International Code Council, Falls Church, VA. International Building Code, 2000.

[2] Naeim F and Kelly JM. Design of Seismic Isolated Structures: From Theory to Practice.

John Wiley & Sons, 1999.

[3] Skinner RI, Robinson WH, and McVerry GH. An Introduction to Seismic Isolation. John

Wiley & Sons, 1993.

[4] Kawamura S et.al. Seismic isolation retrofit in Japan. Proc., 12th World Conference on

Earthquake Engineering, Paper No. 2523. New Zealand Society for Earthquake Engineer-

ing, Upper Hutt, New Zealand, 2000.

[5] Saito H et.al. Structural design of base-isolated high-rise buildings with high-damping

laminated rubber bearings. Proc., 11th World Conference on Earthquake Engineering,

Paper No. 1330. Elsevier Science Ltd., Oxford, England, 1996.

[6] Griffith MC, Aiken ID, and Kelly JM. Displacement control and uplift restraint for base-

isolated structures. ASCE Journal of Structural Engineering, 116(4):1135–1148, 1990.

[7] Logiadis I, Zilch K, and Meskouris K. Prestressed bearings in the seismic isolation of

structures. Proc., 11th World Conference on Earthquake Engineering, Paper No. 1895.

Elsevier Science Ltd., Oxford, England, 1996.

[8] Nagarajaiah S, Reinhorn AM, and Constantinou MC. Experimental study of slid-

ing isolated structures with uplift restraint. ASCE Journal of Structural Engineering,

118(6):1666–1683, 1992.

[9] Wang X-F and Gould PL. Dynamics of structures with uplift and sliding. Earthquake

Engineering and Structural Dynamics, 22(12):1085–1095, 1993.

225
[10] Kelly JM. Dynamic and Failure Characteristics of Bridgestone Isolation Bearings, Rep.

No. UCB/EERC-91/04. Earthquake Engineering Research Center, University of California,

Berkeley, CA, 1991.

[11] Clark PW, Aiken ID, and Kelly JM. Experimental Studies of the Ultimate Behavior of

Seismically-Isolated Structures, Rep. No. UCB/EERC-97/18. Earthquake Engineering Re-

search Center, University of California, Berkeley, CA, 1997.

[12] Youssef N et.al. Building case study: Los Angeles City Hall. The Structural Design of Tall

Buildings, 9(1):3–24, 2000.

[13] Kelly JM. Tension buckling in multilayer elastomeric bearings. ASCE Journal of Engi-

neering Mechanics, 129(12):1363–1368, 2003.

[14] Griffith MC et.al. Experimental Evaluation of Medium-Rise Structures Subject to Uplift,

Rep. No. UCB/EERC-88/02. Earthquake Engineering Research Center, University of

California, Berkeley, CA, 1988.

[15] Kelly JM, Buckle IG, and Koh C-G. Mechanical Characteristics of Base Isolation Bearings

for a Bridge Deck Model Test, Rep. No. UCB/EERC-86/11. Earthquake Engineering

Research Center, University of California, Berkeley, CA, 1987.

[16] Aiken ID, Kelly JM, and Tajirian FF. Mechanics of Low Shape Factor Elastomeric Seismic

Isolation Bearings, Rep. No. UCB/EERC-89/13. Earthquake Engineering Research Center,

University of California, Berkeley, CA, November 1989.

[17] Tyler RG and Robinson WH. High-strain tests on lead-rubber bearings for earthquake load-

ings. Bulletin of the New Zealand National Society for Earthquake Engineering, 17(2):90–

105, 1984.

[18] Hwang J-S and Hsu T-Y. Experimental study of isolated building under triaxial ground

excitations. ASCE Journal of Structural Engineering, 126(8):879–886, 2000.

[19] PEER Strong Motion Database [Online]. http://peer.berkeley.edu/smcat, 2000.

226
[20] Krawinkler H. Personal Communication, 2001.

[21] Chopra AK. Dynamics of Structures: Theory and Applications to Earthquake Engineering,

2nd edition. Prentice Hall, Upper Saddle River, NJ, 2001.

[22] Huang W-H. Bi-directional Testing, Modeling, and System Response of Seismically Isolated

Bridges, PhD thesis. University of California, Berkeley, CA, 2002.

[23] Makris N and Chang S-P. Effect of viscous, viscoplastic, and friction damping on the

response of seismic isolated structures. Earthquake Engineering and Structural Dynamics,

29(1):85–107, 2000.

[24] Mahin SA and Lin J. Construction of Inelastic Response Spectra for Single-Degree-of-

Freedom Systems: Computer Programs and Applications, Rep. No. UCB/EERC-83/17.

Earthquake Engineering Research Center, University of California, Berkeley, CA, 1983.

[25] Sehayek S. Effect of Ductility on Response Spectra for Elasto-Plastic Systems, Pub. R76-42.

Department of Materials Science and Engineering, Massachusetts Institute of Technology,

Cambridge, Massachusetts, 1976.

[26] Murakami M and Penzien J. Nonlinear Response Spectra for Probabilistic Seismic Design

and Damage Assessment of Reinforced Concrete Structures, Rep. No. UCB/EERC-75/38.

Earthquake Engineering Research Center, University of California, Berkeley, CA, 1975.

[27] Fenves GL et.al. Analysis and testing of seismically isolated bridges under bi-axial excita-

tion. Proc., Fifth Caltrans Seismic Research Workshop. California Dept. of Transportation,

Sacramento, CA, 1998.

[28] Chopra AK and Goel RK. Evaluation of NSP to estimate seismic deformation: SDF

systems. ASCE Journal of Structural Engineering, 126(4):482–490, 2000.

[29] Anderson E and Mahin SA. Displacement-based design of seismically isolated bridges.

Proc., Sixth US National Conference on Earthquake Engineering. Earthquake Engineering

Research Institute, Oakland, California, 1998.

227
[30] Fajfar P. Capacity spectrum method based on inelastic demand spectra. Earthquake

Engineering and Structural Dynamics, 28(9):979–993, 1999.

[31] Wadsworth HM, editor. Handbook of Statistical Methods for Engineers and Scientists, 2nd

edition. McGraw-Hill, 1998.

[32] Montgomery DC, Runger GC, and Hubele NF. Engineering Statistics. John Wiley & Sons,

1998.

[33] Metcalfe AV. Statistics in Civil Engineering. Arnold Applications of Statistics Series.

Arnold of Hodder Headline Group, London, England, 1997.

[34] Mosqueda G, Whittaker AS, and Fenves GL. Characterization and modeling of friction

pendulum bearings subjected to multiple components of excitation. ASCE Journal of

Structural Engineering, 130(3):433–442, 2004.

[35] Pan TC and Kelly JM. Seismic response of torsionally coupled base-isolated structures.

Earthquake Engineering and Structural Dynamics, 11(6):749–770, 1983.

[36] Kelly JM. Earthquake-Resistant Design with Rubber. Springer, 1997.

[37] Jangid RS and Kelly JM. Torsional displacements in base-isolated buildings. Earthquake

Spectra, 16(2):443–454, 2000.

[38] Lee DM. Base-isolation for torsion reduction in asymmetric structures under earthquake

loading. Earthquake Engineering and Structural Dynamics, 8(4):349–359, 1980.

[39] Eisenberger M and Rutenberg J. Seismic base isolation of asymmetric shear buildings.

Engineering Structures, 8(1):2–8, 1986.

[40] Nagarajaiah S, Reinhorn AM, and Constantinou MC. Torsion in base-isolated structures

with elastomeric isolation systems. ASCE Journal of Structural Engineering, 119(10):2932–

2951, 1993.

228
[41] Jangid RS and Datta TK. Nonlinear response of torsionally coupled base isolated structure.

ASCE Journal of Structural Engineering, 120(1):1–22, 1994.

[42] Scheller J and Constantinou MC. Response History Analysis of Structures with Seismic

Isolation and Energy Dissipation Systems: Verification Examples for Program SAP2000,

Rep. No. MCEER 99-02. Multidisciplanary Center for Earthquake Engineering Research,

Buffalo, NY, 1999.

[43] Constantinou MC, Mokha A, and Reinhorn AM. Teflon bearings in base isolation.

II:Modeling. ASCE Journal of Structural Engineering, 116(2):455–474, 1990.

[44] Bondonet G and Filiatrault A. Frictional response of PTFE sliding bearings at high

frequencies. ASCE Journal of Bridge Engineering, 2(4):139–148, 1997.

[45] Constantinou MC, Tsopelas P, Kasalanati A, and Wolff ED. Property Modification Factors

for Seismic Isolation Bearings, Rep. No. MCEER 99-12. Multidisciplanary Center for

Earthquake Engineering Research, Buffalo, NY, 1999.

[46] Morgan TA. Characterization and Seismic Performance of High-Damping Rubber Isolation

Bearings, Master’s thesis. University of California, Berkeley, 2000.

[47] Koh C-G and Kelly JM. Effects of Axial Load on Elastomeric Isolation Bearings, Rep.

No. UCB/EERC-86/12. Earthquake Engineering Research Center, University of California,

Berkeley, CA, 1987.

[48] Chalhoub MS and Kelly JM. Reduction of the Stiffness of Rubber Bearings due to Com-

pressibility, Rep. No. UCB/SESM-86/06. Structural Engineering and Structural Mechan-

ics, University of California, Berkeley, CA, 1986.

[49] Aiken ID et.al. Experimental studies of the mechanical characteristics of three types

of seismic isolation bearings. Proc., 10th World Conference on Earthquake Engineering,

Vol. 4, 2281–2286. AA Balkema, Rotterdam, 1992.

[50] Simo JC and Hughes TJR. Computational Inelasticity. Springer, New York, NY, 1998.

229
[51] Pan TC and Kelly JM. Seismic response of base-isolated structures with vertical-rocking

coupling. Earthquake Engineering and Structural Dynamics, 12(5):681–702, 1984.

[52] Lobo RF, Naeim F, and Hart GC. 3D-nonlinear analysis of multistory base isolated build-

ings with significant uplift. Proc., Sixth US National Conference on Earthquake Engineer-

ing. Earthquake Engineering Research Institute, Oakland, California, 1998.

[53] Ryan K and Hall JF. Aspects of building response to near-source ground motions. Proc.,

Structural Engineering World Wide, Paper T162–5. Oxford, England, Elsevier Science

Ltd, 1998.

[54] Hall JF and Ryan KL. Isolated buildings and the 1997 UBC near-source factors. Earthquake

Spectra, 16(2):393–412, 2000.

[55] Koh C-G and Balendra T. Seismic response of base-isolated buildings including P -δ effects

of isolation bearings. Earthquake Engineering and Structural Dynamics, 18(4):461–473,

1989.

[56] Goel RK and Chopra AK. Vibration Properties of Buildings Determined from Recorded

Earthquake Motions, Rep. No. UCB/EERC-97/14. Earthquake Engineering Research Cen-

ter, University of California, Berkeley, CA, 1997.

[57] Almazan JL and de la Llera JC. Lateral torsional coupling in structures isolated with the

frictional pendulum system. Proc., 12th World Conference on Earthquake Engineering,

Paper No. 1534/6/A. New Zealand Society for Earthquake Engineering, Upper Hut, New

Zealand, 2000.

[58] Almazan JL and de la Llera JC. Analytical model of structures with frictional pendulum

isolators. Earthquake Engineering and Structural Dynamics, 31(2):305–332, 2002.

[59] Almazan JL and de la Llera JC. Physical model for dynamic analysis of structures with

FPS isolators. Earthquake Engineering and Structural Dynamics, 32(8):1157–1184, 2003.

230
[60] Almazan JL and de la Llera JC. Accidental torsion due to overturning in nominally sym-

metric structures isolated with the FPS. Earthquake Engineering and Structural Dynamics,

32(6):919–948, 2003.

231

You might also like