Download as pdf or txt
Download as pdf or txt
You are on page 1of 197

INELASTIC DRIFT-BASED APPROACHES FOR PRELIMINARY

SEISMIC ASSESSMENT OF BUILDINGS

By
Hiu Tik CHAN

A Thesis Submitted to
The Hong Kong University of Science and Technology
in Partial Fulfilment of the Requirements for
the Degree of Doctor of Philosophy
in Civil Engineering

June 2011, Hong Kong

HKUST Library
Reproduction is prohibited without the author’s prior written consent
ACKNOWLEDGEMENT 

 
  Thank you and praise to my Lord Jesus Christ. 
 
  My sincere gratitude is to my supervisor, Professor J.S. Kuang, for his 
encouragement,  advice  and  research  support  throughout  my  doctoral 
study.    It  would  be  impossible  to  complete  the  project  without  his  help 
on any difficulties encountered throughout the duration of this project.    I 
have  enjoyed  our  discussion,  dispute  and  sharing  every  successes  and 
failures. 

 
  I  would  like  to  express  my  endless  thanks  to  my  thesis  advisory 
committee, Professor Zongjin Li and Professor Lambros S. Katafygiotis, for 
their invaluable advice throughout the course of this research work. 

 
  I  also  wish  to  thank  the  instructors  of  the  courses  I  took  during  my 
graduate  studies  at  HKUST,  namely:  Professor  J.S.  Kuang,  Professor 
Christopher  K.Y.  Leung,  Professor  Lambros  S.  Katafygiotis,  Professor  C.C. 
Chang,  Professor  Zongjin  Li,  Professor  C.M.  Chan,  Professor  Moe  M.S. 
Cheung,  Professor  X.S.  Li,  and  Professor  Thomas  T.C.  Hsu.    I  learned  a 
great deal from them all. 

 
  I  am  grateful  to  all  my  colleagues  for  encouragement  and  willing 
assistance in this research, especially Dr. K. Huang, Dr. Y.B. Ho, Ms. Gigi S.P. 
Suen, Mr. Vincent H.F. Tsoi, Mr. Terry Y.P. Yuen, Mr. Henry S.H. Luk and Mr 
Will Z. Wang. 

 
  Finally,  I  would  like  to  thank  my  family  members  for  their  love, 
support, patience and understanding.

iv
Table of Contents 

TITLE PAGE  i 

AUTHORISATION PAGE  ii 

SIGNATURE PAGE  iii 

ACKNOWLEDGMENT  iv 

TABLE OF CONTENTS  v 

LIST OF FIGURES  viii 

LIST OF TABLES  xiii 

ABSTRACT  xiv 

Chapter 1 
 
Introduction 
 
1.1  Research background  1 
1.2  Objectives  4 
1.3  Scope of thesis  5 
 
Chapter 2 
 
Literature Review 
 
2.1  Overview of current seismic assessment  6 
2.2  Modal analysis  9 
2.3  Displacement coefficient method  13 
2.4  Drift spectrum  16 
2.5  Sheikh’s method  17 
2.6  Koru’s method  18 
2.7  Lang’s method  19 
2.8  Indices based assessment  20 
2.9  Capacity spectrum method (CSM)  22 
2.10  Nonlinear static analysis  23 

v
2.11  Nonlinear response history analysis  24 
2.12  FEMA 154  26 
2.13  Seismic demand for asymmetric‐plan buildings  27 
2.14  Research significance  29 
 

Chapter 3 
 

Quick Estimate of Fundamental Period of Tall Buildings 
 
3.1  Introduction  37 
3.2  Continuum flexural‐Shear cantilever  39 
3.3  Dynamics properties of the model  40 
3.4  A model for quick estimate of fundamental period T1  41 
3.5  Prediction of degree of structural interaction  43 
3.6  Degree of structural interaction of non‐uniform building  45 
3.7  Numerical studies  47 
3.8  Concluding remarks  49 
 
Chapter 4 
 
Performance‐Based Seismic Assessment of Wall‐Frame   
Building Structures Using Inelastic Seismic Interstorey Drift Ratio 
 
4.1  Introduction  60 
4.2  Bifurcation index for elastic and inelastic behaviour  62 
4.3  Model of inelastic interstorey drift ratio  65 
4.4  Numerical investigation  69 
4.5  Concluding remarks  74 

vi
Chapter 5 
 
Modified Continuum MDOF Model for Seismic Analysis of 
Building Structures 
 
5.1  Introduction  87 
5.2  Modified continuum MDOF model  89 
5.3  Numerical investigation of the modified continuum MDOF Model  95 
5.4  General continuum representation  97 
5.5  Concluding remarks  105 
 
Chapter 6 
 
Generalised Inelastic Response Spectra 
 
6.1  Introduction  116 
6.2  Methodology  119 
6.3  Generalised inelastic response spectra  120 
6.4  Procedure of assessment using generalised inelastic response  122 
6.5  Numerical investigation  124 
6.6  Concluding remarks  127 
 
Chapter 7 
 
Conclusion 
 
7.1  Conclusion  156 
7.2  Further research  159 

References  160 

Appendices 
A.1  Continuum flexural‐shear cantilever  164 
A.2  Model under uniform lateral loading  171 
A.3  Model under inverted triangular lateral loading  176 
A.4  Model under point load at the roof  180 

vii
List of Figures 
 

Chapter 2 
Fig. 2.1  A single‐degree of freedom system  30 

Fig. 2.2  Multi‐degrees of freedom system  30 

Fig. 2.3  Shear beam model  31 

Fig. 2.4  Continuum flexural‐shear cantilever  31 

Fig. 2.5  Resultant capacity curve from superposition  32 

Fig. 2.6  Vulnerability curve of a building  32 

Fig. 2.7  Capacity spectrum method  33 

Fig. 2.8  Maximum interstorey drift versus spectral acceleration  33 

Fig. 2.9  Fractile curves  34 

Fig. 2.10  Data collection forms for   

  the three designated seismicity regions  35 

Fig. 2.11  Basic score and score modifiers  36 

Chapter 3 
Fig. 3.1  Continuum flexural‐shear model  51 

Fig. 3.2  Deformation shape of flexural‐shear cantilever  51 

Fig. 3.3  Free body diagram of flexural‐shear cantilever at time t   

  under free vibration  52 

Fig. 3.4  Free body diagram of flexural‐shear cantilever   

  under lateral loading  53 

Fig. 3.5  Index of structural interaction  54 

Fig. 3.6  Variation of α against drift ratio RΔ /Rδm  54 

viii
Fig. 3.7  The floor plans of the building  55 

Fig. 3.8  The variation of stiffness and   

  the equivalent stiffness of the building  56 

Fig. 3.9  (a) From PEER with record ID: P0714   

  (b) From PEER with record ID: P0410  57 

Fig. 3.10  Time history response at height of 207.7m  57 

Fig. 3.11  Time history response at height of 129.7m  57 

Fig. 3.12  Time history response at height of 17.5m  57 

Fig. 3.13  Floor plans of example buildings  58 

Fig. 3.14  Finite element models  59 

Chapter 4 
Fig. 4.1  Graph of SR(α)  76 

Fig. 4.2  Graph of IR(Sr, α)  76 

Fig. 4.3  Preliminary seismic assessment procedure  77 

Fig. 4.4  Floor plans of the three buildings  78 

Fig. 4.5  Models in “OpenSees”  79 

Fig. 4.6  Material properties  79 

Fig. 4.7  Displacement envelopes under four groups of ground motion  80 

Fig. 4.8  Interstorey drift ratio envelopes of four groups of ground motion 81 

Fig. 4.9  Displacement ductility of four groups of ground motion  82 

Fig. 4.10  Drift ductility of four groups of ground motions  83 

Fig. 4.11  Pseudo acceleration spectra at 5% viscous damping  84 

Fig. 4.12  Contour of Cο  84 

ix
Fig. 4.13  Comparison of maximum interstorey drift ratio   

  by proposed formula and RHA of the first building  85 

Fig.4.14  Comparison of maximum interstorey drift ratio   

  by proposed formula and RHA of the second building  85 

Fig.4.15  Comparison of maximum interstorey drift ratio   

  by proposed formula and RHA of the third building  86 

Chapter 5 
Fig. 5.1  Strain diagram of a rectangular section  106 

Fig. 5.2  (a) Continuum‐MDOF model;   

  (b) Section of the flexural cantilever  106 

Fig. 5.3  Stress strain relationship (a) Fibres in the flexural cantilever   

  (b) Shear cantilever  107 

Fig. 5.4  Capacity curve of the continuum‐MDOF model  107 

Fig. 5.5  (a) Excitation from PEER with record ID: P1121; 

  (b) The roof displacement under the excitation in (a); 

  (c) The roof displacement under the excitation two times of (a); 

  (d) The roof displacement under the excitation three times of (a)108 

Fig. 5.6  Prime and sub model  109 

Fig. 5.7  Capacity curve of building α =1.6   

  (a) Inverted triangular loading (b) Uniform loading  110 

Fig. 5.8  Capacity curve of building α =3.0 

  (a) Inverted triangular loading (b) Uniform loading  111 

x
Fig. 5.9  The capacity curve of building α =9.4 

(a) Inverted triangular loading 

(b) Uniform loading  112 

Fig. 5.10  Model with α = 1.6 (a) Roof drift; 

  (b) Interstorey drift ratio; (c) Curvature  113 

Fig. 5.11  Model with α = 3.0 (a) Roof drift; 

  (b) Interstorey drift ratio; (c) Curvature  114 

Fig. 5.12  Model with α = 9.4 (a) Roof drift; 

  (b) Interstorey drift ratio; (c) Curvature  115 

Chapter 6 
Fig. 6.1  Generalised roof drift spectra (a) and (b) (Case 1)  129 

Fig. 6.2  Generalised roof drift spectra (c) and (d) (Case 1)  130 

Fig. 6.3  Generalised roof drift spectra (e) and (f) (Case 1)  131 

Fig. 6.4  Generalised roof drift spectra (g) and (h) (Case 1)  132 

Fig. 6.5  Generalised roof drift spectra (a) and (b) (Case 2)  133 

Fig. 6.6  Generalised roof drift spectra (c) and (d) (Case 2)  134 

Fig. 6.7  Generalised roof drift spectra (e) and (f) (Case 2)  135 

Fig. 6.8  Generalised roof drift spectra (g) and (h) (Case 2)  136 

Fig. 6.9  Generalised interstorey drift ratio spectra (a) and (b) (Case 1)  137 

Fig. 6.10  Generalised interstorey drift ratio spectra (c) and (d) (Case 1)  138 

Fig. 6.11  Generalised interstorey drift ratio spectra (e) and (f) (Case 1)  139 

Fig. 6.12  Generalised interstorey drift ratio spectra (g) and (h) (Case 1)  140 

Fig. 6.13  Generalised interstorey drift ratio spectra (a) and (b) (Case 2)  142 

Fig. 6.14  Generalised interstorey drift ratio spectra (c) and (d) (Case 2)  142 

xi
Fig. 6.15  Generalised interstorey drift ratio spectra (e) and (f) (Case 2)  143 

Fig. 6.16  Generalised interstorey drift ratio spectra (g) and (h) (Case 2)  144 

Fig. 6.17  Generalised curvature spectra (a) and (b) (Case 1)  145 

Fig. 6.18  Generalised curvature spectra (c) and (d) (Case 1)  146 

Fig. 6.19  Generalised curvature spectra (e) and (f) (Case 1)  147 

Fig. 6.20  Generalised curvature spectra (g) and (h) (Case 1)  148 

Fig. 6.21  Generalised curvature spectra (a) and (b) (Case 2)  149 

Fig. 6.22  Generalised curvature spectra (c) and (d) (Case 2)  150 

Fig. 6.23  Generalised curvature spectra (e) and (f) (Case 2)  151 

Fig. 6.24  Generalised curvature spectra (g) and (h) (Case 2)  152 

Fig. 6.25  Participation factor of the 1st mode  153 

Fig. 6.26  Mean pseudo‐ acceleration spectrum   

  with 5% viscous damping of the ground motions in Table 6.1  153 

Fig. 6.27  Capacity curve of the building under inverted triangular load  154 

Fig. 6.28  Roof drift under response history analysis   

  and the prediction from the method   

  of generalised inelastic response spectra  154 

Fig. 6.29  Interstorey drift ratio under response history analysis   

  and the prediction from the method   

  of generalised inelastic response spectra  155 

Fig. 6.30  Curvature under response history analysis   

  and the prediction from the method   

  of generalised inelastic response spectra  155 

xii
Appendices 
Fig. A1  Free body diagram of flexural‐shear cantilever   

  at time t under free vibration  164 

Fig. A2  Free body diagram of flexural‐shear cantilever  171 

Fig. A3  Flexural‐shear cantilever under inverted triangular loading  176 

Fig. A4  Flexural‐shear cantilever under point load at the roof  180 

List of Tables 
 

Table 2.1  Assessment level  30 

Table 3.1  Dimensions of Structural Members  49 

Table 3.2  Structural parameters and applied lateral load  49 

Table 3.3  Comparison of fundamental periods T1 predicted   

  by proposed formula and finite‐element analysis  49 

Table 4.1  Deformation Limits suggested in ATC‐40  75 

Table 4.2  Relationship between Structural Damage and Story Drift  75 

Table 4.3  Dimensions of the members in the representative buildings  75 

Table 4.4  Ground motion records  75 

Table 6.1  Ground Motion records  128 

xiii
INELASTIC DRIFT‐BASED APPROACHES FOR PRELIMINARY 
SEISMIC ASSESSMENT OF BUILDINGS 

By Hiu Tik CHAN

Department of Civil and Environmental Engineering 
The Hong Kong University of Science and Technology 

Abstract
 
There is no provision for consideration of seismic resistance in building design codes of
practice in low-to-moderate seismic regions in this world. This could cause a significant
loss of life and economic hardship under a moderate earthquake. This is due to the high
population density and the extensive infrastructure in the urban areas. Thus there is an
urgent need to implement a seismic assessment scheme for building structures composed of a
preliminary assessment and a detailed assessment. In this thesis the methods are proposed
for the preliminary seismic assessment which is a crucial step and a screening procedure for
identifying deficient building structures from a pool of buildings for more detailed
assessment.

Existing preliminary seismic evaluation procedures are mainly for low to medium rise
building structures, in which shear deformation is dominant. Thus an evaluation procedure
for high rise building structures is proposed with consideration of inelastic behaviour and
higher vibration mode effect. A method of analysis for estimating quickly the natural
fundamental period of vibration for tall building structures is first proposed from the continuous
flexural-shear cantilever. Then an inelastic interstorey drift ratio model is derived from the
principle of capacity spectrum analysis and modified modal pushover analysis for the assessed
buildings. Numerical investigations on the representative tall wall-frame structures show that
the seismic responses of the buildings predicted by the proposed inelastic interstorey drift
ratio model agree well with those obtained from response history analysis.

xiv
A systematic and effective assessment procedure for the preliminary assessment is
developed. The proposed procedure first divides the assessed buildings into either elastic or
inelastic one using a bifurcation index. For building structures that deform inelastically may
be further examined by the inelastic interstorey drift ratio model. Performance of building
structures is then defined based on the interstorey drift ratio by comparing those given in the
current codes of practice and the literature. The proposed model is particularly suitable for
assessing the buildings that are analysed in frequency domain.

A modified continuum MDOF model, based on continuum flexural-shear cantilever, is


derived for analysing seismic responses of tall building, which is more suitable for rapid
preliminary seismic assessment in time. It particularly suits for slender buildings that have
inelastic response range and significantly higher mode effect. Conditions of model similarity
is then defined which is named as the general continuum representation. For two similar
continuum models, their seismic responses can be related by their height ratios for the same
strength ratio based on the general continuum representation. Thus the responses of a
modified MDOF model is normalised to dimensionless responses named as the generalised
inelastic responses.

A spectrum-based method using the generalised inelastic response is proposed and


recommended for preliminary seismic assessment of tall buildings that are relatively slender
with significant higher mode effects and are normally analysed in time domain. The responses
of a building structure can be predicted using the generalised inelastic responses, which are
derived from the modified continuum-MDOF model through the general continuum
representation. The generalised inelastic response are presented in the form of spectra which
are developed by the responses of 168 modified continuum MDOF models, having 8 different
degrees of structural interaction and 21 different fundamental periods under nonlinear response
history analysis. The numerical investigation shows that the seismic responses from the
proposed method agree well with those from the response history analysis. The proposed
method provided a simple and quick, yet accurate, mean of estimating the inelastic seismic
response of building structures, thus particularly being suitable for the preliminary seismic
assessment.

xv
Chapter 1 

Introduction 

1.1 Research background

Earthquakes, unpredictable and unavoidable, have brought catastrophic losses in the economy
and precious human lives. A satisfactory explanation of earthquakes, Elastic Rebound Theory,
was proposed by Henry Fielding Reid to explain the mechanisms of an earthquake after the
event of great 1906 San Francisco earthquake. Unfortunately, knowing of earthquake
mechanisms does not help to predict the occurrence of an earthquake as reliably as the weather
forecast. Prediction of the occurrence of an earthquake is difficult, but potentially dangerous
regions can be identified through the locations of fault lines. Regions of different seismicity
are classified based on the fault’s location and the probability of earthquake occurrence. Three
main regions having low, moderate and high seismicity have been defined.

1
Regions of low to moderate seismicity have a relatively lower probability of earthquakes
than those of high seismicity, but they are not entirely free of earthquakes. Little attention has
been paid on those regions, as the probability of earthquake occurrence is very small. An
earthquake can devastate a whole city if it happens in a region where the structures are not
designed explicitly against earthquakes. This is exactly the situation faced by regions of low
to moderate seismicity. Because of the low probability of earthquake occurrence, buildings
are only designed against the combined effect of static gravity and static lateral load,
particularly wind loads. Designs for static loads are completely different to designs against
earthquakes. The dynamic properties of the buildings which govern the safety of buildings
under earthquake actions are unknown in these designs.

Buildings with seismic design may not be guaranteed to behave as expected as the design
because of uncertainties in earthquakes and invalid assumptions in design. This leads to the
question on “how safe are code-conforming building structures?” being continuously asked (El
Howary and Mehanny, 2011). Therefore buildings without seismic design are undoubtedly
full of uncertainties under earthquake action because of a totally different design philosophy.
In low to moderate seismicity regions there are thousands of non-seismically designed
buildings that are unprotected from seismic loadings and it may be very dangerous to ignore
any vulnerable building in an urban area. An unexpected earthquake will trigger this danger
because damage of buildings and infrastructures are the major source contributing to the losses
in the economy and in life. Citing the 5.6-magnitude Newcastle earthquake in 1989 as an
example, significant loss of life (13 deaths) and economic hardship (A$4 billion loss) was
caused in a region of low to moderate seismicity.

One single earthquake or aftershocks to a non-seismically designed building may bring


unexpected damage or may cause the building to be unable to function normally. Earthquake
hazards cannot be overlooked because of the small probability of occurrence, because the risk
of earthquake is always evaluated from both the hazard and probability,

Risk = Hazard x Probability

Hong Kong is an example of such a region, in which buildings are not seismically designed and
even worse, it has a densely populated environment. The large population forces both
residential and commercial buildings to occupy higher space and leads to buildings to

2
commonly being 40 storeys or above. Every non-seismically designed tall building may be
regarded as a potential tragedy because the failure of one building in a dense environment may
lead to a worst case, due to a domino effect.

A remediation method for regions that have adopted non-seismic design is a


pre-earthquake structural strengthening plan which requires a full scale seismic assessment of
existing structures. This is a very involved project with thousands of vulnerable structures
needing assessment. In order to efficiently allocate resources to existing structures, a
comprehensive assessment scheme consisting of both preliminary and detailed assessment is
necessary. A preliminary seismic assessment method can help engineers to screen out
seismically hazardous buildings in a fast and cost effective manner with satisfactory accuracy.
After this screening process, potentially hazardous buildings are assessed by a detailed method
so that any deficiencies under earthquake action are clearly revealed.

To reduce seismic risk in an effective way, a comprehensive seismic assessment scheme


is necessary. For countries located in high seismicity regions, they have already established a
comprehensive seismic assessment scheme, for instance codes and guidelines provided by
Federal Emergency Management Agency (FEMA) and Applied Technology Council (ATC) in
the U.S.A. Countries or cities located in low to moderate seismicity regions mostly do not
have a seismic assessment scheme, and the lack of thus schemes under the threat of earthquakes
is a great worry. Adopting the current assessment methods and schemes from other countries
is a possible solution, however due to the design characteristics of buildings in low to moderate
seismicity, only detailed assessment methods can be applied. Lack of a preliminary seismic
assessment method is hence the major problem in regions of low to moderate seismicity.

It is expressly inconceivable to perform a detailed seismic assessment for every building


because of tremendous amount of resources and time required. Moreover detailed assessment
involves sophisticated analysis which cannot be easily performed by engineers who do not have
the appropriate background. Therefore, a preliminary seismic assessment method for
non-seismically designed buildings is proposed in this study

3
1.2 Objectives

The primary objective of this research is to develop a comprehensive and reliable seismic
assessment method that can be implemented into the preliminary seismic assessment procedure,
in order to conduct a rapid screening on a large number of building structures so as to identify
potential hazardous building structures for detailed assessment. To accomplish the primary
objective, the following secondary objectives are needed,

z To develop a practical, yet simple, engineering model for existing building structures
which could best predict their fundamental period and vibration mode shapes.

z To propose a rapid preliminary seismic assessment procedure that identifies the


behaviour of an assessed building structure into either the elastic or inelastic range of
response and to determine the performance based on the magnitude of the interstorey
drift ratio.

z To develop methods for predicting the inelastic interstorey drift ratio of the assessed
building structures based on two different analytical approaches depending on the
resources that can be provided in the assessment and the characteristic of the
building structure.

By fulfilling the objectives, the impacts that can be made practically and academically are

z To facilitate the seismic assessment of building structures in regions of low to


moderate seismicity by effectively allocating resources in the assessment.

z To fulfil the urgent need of society in the management of earthquake risk that has not
been considered in regions of low to moderate seismicity.

z To enhance understanding of inelastic seismic behaviour of building structures


through the proposed methods of modelling and analysis.

4
1.3 Scope of thesis

This thesis is divided into seven chapters. Chapter 1 is the background of this research and it
describes the current situation of seismic assessment in regions of low to moderate seismicity.
The objectives and scope of the thesis are also addressed. In Chapter 2, some typical studies
and provisions in the codes of practice on both preliminary and detailed assessment are
reviewed.

A model for quick estimate of natural fundamental period of tall buildings is presented in
Chapter 3. It is based on the continuum modelling technique derived in detail, and a
numerical investigation is included. Chapter 4 proposes a method extending the previous
continuum modelling technique for building structures to predict the inelastic interstorey drift
ratio under ground motion based on capacity spectrum analysis. A procedure for rapid
assessment when implementing the proposed method is presented.

Chapter 5 proposes a simplified structural model, which is the modified


continuum-MDOF model, for rapid preliminary seismic evaluation of building structures with
significantly higher vibration mode effect and inelastic range of responses. Three building
structures are investigated to verify the accuracy of the proposed modified continuum-MDOF
model. Models having the defined similarities are investigated and the relationships
between the seismic responses of similar models are derived. The relationships under the
conditions of model similarity are named as the general continuum representation.

In chapter 6, the responses of a modified continuum-MDOF model is normalised to


dimensionless responses and applied to similar models based on the general continuum
representation. The dimensionless responses are called the generalised inelastic responses.
They are expressed in the form of spectra named as the generalised inelastic response spectra.
The procedure for using the spectra in assessment is illustrated and the proposed method is
verified by a numerical investigation. Finally, conclusions for the thesis are drawn in
chapter 7.

5
Chapter 2 

Literature Review 

2.1 Overview of current seismic assessment

In countries that are subjected to frequent earthquake threats like China, Japan and U.S.A.,
engineers have to consider earthquake induced forces at the stage of design. This is to ensure
that the structures would behave safely and achieve the desired performance under unexpected
ground motions at any time. Not only is a precaution policy conducted at the design stage, but
a seismic vulnerability assessment scheme is carried out to rehabilitate and strengthen historical
and unqualified buildings. Without doubt, seismic design and assessment are a must in strong
seismicity regions in order to manage earthquake induced problems.

On 28th December 1989, a magnitude of 5.6 ML earthquake occurred at Newcastle of


Australia, which is located in the region of low to moderate seismicity. Structures without
seismic design were seriously damaged causing 13 deaths, 160 people hospitalised and a loss of
A$4 billion. The performance of building structures in low-to-moderate seismicity regions,

6
which have not been designed against earthquakes, is a worrying problem. More and more
attention has been paid on the evaluation of those building structures, for instance the research
conducted by Lang (2002) to evaluate the seismic risk in Switzerland. To manage seismic risk
has also become a problem in low-to-moderate seismicity regions.

Seismic risk is closely related to the performance of buildings and hence seismic
vulnerability assessment of existing buildings is one of the major tasks in the risk management
policy. With increasing understanding in the structural performance and the increasing
computational power in personal computers, more in-depth research on seismic vulnerability
assessment of buildings is being continuously conducted. Research on seismic vulnerability
assessment in general is composed of two parts, which are the development of the engineering
models of the assessed buildings and the method of analysis.

The development of the engineering model changes from being just a simple enough
SDOF model with mathematical formulae to a more detailed MDOF model in finite-element
based software. In the meantime, the method of analysis has improved from linear static
analysis, which only predicts the maximum response, to nonlinear response history analysis,
which captures the changes of response in a period of time. In spite of the different models and
methods to evaluate the performance of buildings, the main aim of seismic assessment is to
assist practicing engineers to evaluate existing buildings in an organised, reliable, fast and cost
effective way.

A comprehensive seismic assessment scheme includes different assessment levels ranging


from a preliminary to a detailed level, depending on the assessment objectives. Detailed
assessment requires engineers who have high technical knowledge in earthquake engineering to
conduct the evaluation, which is technically complex and is, both time and cost consuming in
nature. This kind of assessment can only be limited to buildings which have been identified as
potentially dangerous under earthquake actions. It is then very important to conduct a
preliminary and simplified assessment to screen out potential hazardous buildings in an urban
area. The assessment scheme starts from a preliminary to detailed assessment so that both
resources and time for assessment are minimised. A brief description of the assessment level
is discussed in the following.

7
Preliminary assessment (Level one)

The assessment is based on the observation of buildings and relies on the experience of
engineers. It is usually done by score assignment which involves little or no calculation and is
suitable for quick evaluation. This kind of assessment is also named as a walkthrough
assessment or rapid visual screening procedure because the assessment procedure is mainly
observational. Most of the time, the assessment is developed from models derived from statics.
The purpose of the assessment is to screen out deficient buildings from a larger number of
building structures. Although this kind of assessment is efficient in both time and cost, the
results are subjected to larger uncertainties relative to other in-depth assessments

Preliminary assessment (Level two)

Assessment is based on a reasonable engineering model of particular types of buildings.


Additional information other than from observation is required to define the model such as the
floor area, lateral strength and the fundamental period of the assessed building structure. The
assessment involves robust analysis on the engineering model, but the process does not involve
complicated calculations. The results from the engineering model sometimes are calibrated
from statistical results in a database which contains the responses of similar building
structures. It is used to capture minor changes in structural forms, geometry and loading,
which are the characteristic of the assessed building.

Detailed assessment (Level three)

Detailed assessment requires a full scale model of the assessed building, which includes
modelling of all the structural members and some necessary non-structural members, as well as
the material properties obtained from in-situ inspection. Individual members are represented
by the detailed model (Ile et al., 2000; Kim et al., 2005) thus that the members behave close to
the experimental findings or the real situation. All individual members are modelled carefully
and are combined to form the structural system. The large scale structural system is then
analysed by special methods which involve heavy computational effort, such as nonlinear static
analysis and nonlinear response history analysis using finite-element based software.

8
The assessment has to be done by engineers having a seismic design background and deep
understanding of structural behaviour. Because of the demanding characteristics of this kind
of assessment, it is only applied to very important buildings and buildings which are identified
as deficient buildings, after the preliminary assessment, for future retrofit and strengthening
works. The assessment is cost demanding and time consuming, but it gives results which are
more reliable than the other assessment levels. Since there are many existing methods of
analysis in earthquake engineering, some typical and representative methods of assessment are
selected and discussed in this chapter.

2.2 Modal analysis

Modal analysis is a fundamental technique in earthquake engineering. Many methods of


assessment are developed by this principle. Although it is well-known and is familiar to many
professionals in earthquake engineering, a brief description is included for completeness of the
literature review and for people without the background of earthquake engineering to
understand more advanced methods of assessment.

Modal analysis is a linear analysis, which is used to predict the responses of a multi-degree
of freedom (MDOF) model, which behaves linearly under the ground motion. It is a technique
to decompose an MDOF system into a corresponding single-degree of freedom (SDOF) system
for further analysis.

SDOF system
The simplest approach to study the dynamic behaviour of a structure under ground motion is to
idealise the structure into an SDOF system. It has a lumped mass m supported by a zero mass
member with stiffness k and damping coefficient c, as shown in Fig. 2.1. There is only one
single response in a single degree of freedom system, which is the relative lateral displacement
u (t ) in this case. The characteristic equation of the lateral displacement of the system under
ground acceleration u&&g (t ) is written as

mu&&(t ) + cu& (t ) + ku(t ) = −mu&&g (t ) (2-1)

The characteristic equation can also be written as


u&&(t ) + 2ξωnu& (t ) + ωn u (t ) = −u&&g (t )
2
(2-2)

9
k c
where the angular frequency is ωn = and the damping ratio is ξ = . The lateral
m 2mωn
displacement solved under a particular angular frequency in Eq. (2-2) is a function of time.
The response of an SDOF system under ground motions is therefore governed by the angular
frequency of the system.

The maximum lateral displacements under different angular frequencies are predicted and
the graph contains the maximum response at the corresponding frequency and is the called
response spectrum. The relative displacement spectra contain curves of the maximum relative
displacement and its corresponding frequency at different damping ratios. Once the
fundamental period of a structure is predicted, the maximum lateral displacement can be
obtained from the relative displacement spectra by assuming the assessed building structure
behaves as an SDOF system.

MDOF system
An SDOF system is simple enough for analysis, but it is too simplified and cannot well capture
the behaviour of a building structure. It is not appropriate to assume a building behaves as an
SDOF system because buildings are more complex so that a single response at the level of the
lumped mass is not enough to model the deformation of building with its height.

A multi degree of system, which is more detailed, is then used to model the building
structure. The simplest MDOF model of a five-storey building is shown in Fig. 2.2. There
are five nodes at the floor level where the floor masses are lumped and the responses are of
concern. The concerned responses are the relative lateral displacements at the floor levels.
When a building is modelled as a multi-degree freedom system similar to Fig. 2.2, the
characteristic equation of lateral displacement of the system ground acceleration is as follows:
M&x& + Cx& + Kx = − Mιu&&g (2-3)

where M is the mass matrix, C is the damping matrix, K is the stiffness matrix of the system, x is
the vector containing the concerned responses and ι is the influence vector describing the
distribution of inertia force.

For the system in Fig 2.2, the response vector x covers the five relative lateral
displacements at the floor levels. The mass matrix, damping matrix and the stiffness matrix

10
are square matrices with size 5μ5, consistent to the numbers of responses. The system has five
degree of freedom with respect to the number of responses in the system. An MDOF system
can predict the behaviour of buildings more closely because more responses are captured in the
system, such as rotation and axial deformation. However the matrix size of the system would
be larger and hence the computation effort for solving the characteristic equation Eq. (2-3)
increases.

For the simplest MDOF model, the concerned responses are usually the relative lateral
floor displacements. The response vector x in the characteristic Eq. (2-3) is solved by
separation of variables. The vector is expressed as the superposition of N modal contributions
which are consistent to the numbers of responses:
N
x(t ) = ∑φn qn (t ) (2-4)
n =1

where φ n is the shape function of the nth mode, N is the number of degree of freedom and

qn (t ) is modal coordinate.

The shape functions are solved by Eigen analysis, and an N-degree of freedom system
would have N shape functions. The shape functions have the orthogonality property which is
φnT KφrT = φnT MφrT = 0 . Because of this property, Eq. (2-3) can be simplified by the following
procedures. The characteristic equation after substitution Eq. (2-4) is given by
N N N
M ∑φn q&&n (t ) + C ∑φn q&n (t ) + K ∑φn qn (t ) = − Mιu&&g (2-5)
n =1 n =1 n =1

Multiplying both sides of the Eq. (2-5) with the shape function of rth mode becomes
φrT Mφr q&&r +φrT Cφr q& r +φrT Kφr q r = −φrT Mιu&&g (2-6)

Dividing both sides of the Eq. (2-6) with φrT Mφr , which is non-zero, becomes

q&&r +2 ξ r ω r q& r + ω r2 q r = − Γr u&& g (2-7)

φrT Cφr φrT Kφr φrT Mι


where 2 ξ r ωr = , ω 2
= and Γr = . Eq. (2-7) is actually the
φrT Mφr φrT Mφr φrT Mφr
r

characteristic equation of a single degree of freedom system with angular frequency and
viscous damping ratio calculated as follows

φrT Kφr
ωr = (2-8)
φrT Mφr

11
1 φrT Cφr
ξr = (2-9)
2 ωr φrT Mφr
The response of an MDOF system in Eq. (2-4) hence can be found by the sum of N response
vectors of SDOF systems in Eq. (2-7).

An MDOF system with N-degrees of freedom can be decomposed into N SDOF systems.
Among the N systems, the system with the smallest angular frequency ωr has the largest
participation factor Γr in Eq. (2-4) and hence has the greatest contribution in response.
Therefore most of the preliminary assessment methods estimate the seismic performance based
on the fundamental mode instead of all the modes to reduce computational effort. However
for particular systems that have non-uniform stiffness or unevenly distributed mass and
irregular geometry, etc., the contributions in response from the higher modes will be very
significant. The influence of higher modes has to be considered.

Response spectra developed from the SDOF systems allow a fast prediction of response by
graphical means. It is a powerful tool recording the maximum response of an SDOF system
with all combinations between the fundamental period and viscous damping. It can also be
used to predict the response of an MDOF system by modal analysis in the frequency domain.

Equivalent linear system using modal analysis


Modal analysis is a very appealing method, but it is only applicable to linear systems which
deform elastically. Since building structures may be excited into the inelastic range of
response under ground motion, nonlinear systems have to be considered in seismic engineering.
To extend a linear system for predicting the inelastic response range, an equivalent linear
system that has similar responses to the corresponding nonlinear inelastic system is usually
used as an alternative for seismic assessment.

For instance, an improved linear elastic procedure has been proposed (Gunay and
Sucuoglu, 2010). In the analysis, members that are expected to behaviour inelastically are
replaced by new members with reduced stiffness. The reduced stiffness of the member is
obtained by the spectral displacement under the equal displacement rule. The original model
is modified to a new linear system with reduced stiffness thus that it can be analysed through
modal analysis. The new linear system presents the inelastic behaviour of the original system.

12
2.3 Displacement coefficient method

Buildings may deform inelastically under ground motion, so a method for prediction of
inelastic displacement under ground motion is important in assessment. A building is a
multi-degree of freedom system, which can be simplified as an SDOF system by modal analysis.
The responses of the building are usually dominated by the first few modes, which have relative
larger participation factors in Eq. (2-4). It is assumed that the building would yield according
to the dominant mode, which contributes the largest part of the response. When a structure
yields, the responses from the higher modes are assumed to be negligible. The problem thus
becomes the prediction of the inelastic response of the dominant SDOF system.

If an SDOF system behaves linearly under ground motion, there are response spectra,
which allow a fast prediction of the response by graphical mean. Because of the convenient
nature and power of the response spectrum, various methods have been proposed to make use
of it for the prediction of the inelastic response. These methods involve multiplication of the
coefficients to the response spectra, to obtain inelastic responses. This is the displacement
coefficient method.

Miranda and Ruiz (2002) compared six different displacement coefficient methods to
predict the inelastic displacement demands of an SDOF system. The six methods are
classified into two distinct approaches. The first approach is to replace the inelastic SDOF
system with an equivalent linear SDOF system with a modified stiffness and damping ratio.
The response of the inelastic SDOF system is said to be equal to that of the equivalent linear
SDOF system; this is known as the equivalent linear method. The second approach multiplies
the response of linear SDOF system with an inelastic displacement ratio to obtain the response
of the inelastic SDOF system; this is the inelastic displacement ratio method.

2.3.1 Inelastic displacement ratio method


An approximate method to predict the maximum inelastic displacement demand is to multiply
the linear displacement by an inelastic displacement ratio Cμ = Δinelastic / Δ elastic . It is a ratio of

the maximum lateral inelastic displacement demand to the maximum linear displacement
demand on the same SDOF system. Miranda (2000) determined the value of the inelastic
displacement ratio for SDOF systems with different fundamental periods. The maximum

13
lateral inelastic displacement, which is the nominator of the ratio, is determined from the
nonlinear response history analysis of an SDOF model with defined hysteretic behaviour. The
maximum linear displacement, which is the denominator, is determined from the linear
response time history analysis of the SDOF model.

A graph of the inelastic displacement ratio and the displacement ductility, which is the
ratio of the inelastic displacement to the yielding displacement, is plotted. Miranda discovered
that the graph of the inelastic displacement ratio can be described by an implicit function of
building’s period and displacement ductility, given by
−1
⎡ ⎛1 ⎞ ⎤
( )
Cμ = ⎢1 + ⎜⎜ − 1⎟⎟ exp − 12Tμ − 0.8 ⎥ (2-10)
⎣ ⎝μ ⎠ ⎦
where μ is the displacement ductility.

Eq. (2-10) is an implicit equation, in which the response of the SDOF system has to found
by iteration. The iteration is simple enough in that it can either be done by programming or
manually. Given the fundamental period T and the yielding displacement of an SDOF system,
a first guess of the displacement ductility is assumed and is input in Eq. (2-10) to calculate the
inelastic displacement ratio. If the new displacement ductility calculated from the first guess
of inelastic displacement ratio is consistent to the first guess of displacement ductility, the
solution is found. Otherwise the value of ductility is updated by the new inelastic
displacement ratio and becomes the second guess. The iteration process stops when the
inelastic displacement obtained from Eq. (2-10) is consistent to the guess with an acceptable
tolerance.

2.3.2 Equivalent Linear Method


For any inelastic SDOF model with defined hysteretic behaviour under ground motion, there
exists a linear SDOF model thus where its maximum linear deformation is equal to the
maximum inelastic deformation of the inelastic SDOF model. The linear SDOF model that
gives same response is the equivalent linear model of the inelastic model. The principle of
equivalent linear method is used to find out the equivalent SDOF model.

The methodology is started from the characteristic equation of the deflection of the inelastic
SDOF system, which is given by

14
4πξo 2π
&x& + x& + ( ) 2 f (x ) / ko = − &x&g (2-11)
To To

where ξ o is the inherent damping ratio, f ( x ) is the restoring force of the system and k o is

the initial stiffness of the system. Eq. (2-11) is compared to the characteristic equation of a
linear SDOF system, which is given as follows,
4πξ eq 2π 2
&x&eq + x&eq + ( ) xeq = − &x&g (2-12)
Teq Teq

The equivalent linear method tries to find the equivalent period and damping so that the
maximum x in Eq. (2-11) is equal to the maximum xeq in Eq. (2-12).

In the study of Lin and Miranda (2008), relationships between the inelastic system and
the equivalent linear system are established. A group of inelastic SDOF systems with
various periods is subjected to ground motion under same conditions as in response history
analysis. Then large scale trial and error is done to find out the equivalent system for each
inelastic SDOF system in the group. The combination of the equivalent period Teq and

equivalent damping ratio ξeq which gives the least discrepancy between the maximum

displacement of the equivalent system and the inelastic displacement allows the solution.

The original period and damping ratio in Eq. (2-11) are plotted against the equivalent
period and damping ratio in Eq. (2-12). Two simplified empirical formulas expressed in terms
of strength ratio R, which is the ratio of elastic lateral strength to yield lateral strength, are
derived from regression analysis.
n1
ξ eq = ξ 0 + ( R − 1) (2-13)
T0n2

Teq m1
= 1+ m2
( R1.8 − 1) (2-14)
To T0
where m and n is post yield stiffness ratio.

In the assessment, the fundamental period and the strength ratio of the SDOF model are
first found. Then the corresponding equivalent SDOF system can be defined from the Eqs
(2-13) and (2-14). Finally, the response of the equivalent linear SDOF system is obtained from
the displacement response spectra.

15
2.3.3 Conclusions
The displacement coefficient method is a convenient tool which allows a quick prediction of the
inelastic response of an SDOF system. It empowers the response spectrum of linear SDOF
systems so that it can be used to model nonlinear inelastic behaviour. Despite the benefits of
the method, the fundamentals of the method only give the response of SDOF systems and not
MDOF systems. Prediction based on the nonlinear response of an SDOF system is an indirect
approach to model the nonlinear response of an MDOF system.

2.4 Drift spectrum

Drift spectrum is a seismic assessment tool similar to the response spectrum. It is developed
either from continuum models, which usually have simple mathematical expression in
behaviour. Drift spectrum expresses the maximum interstorey drift ratio as a function of the
fundamental period. Unlike the response spectrum which just gives the maximum response of
the SDOF system, drift spectrum gives the maximum interstorey drift ratio of the model, in
which modal analysis has been carried out to give the resultant response of all the modes.

Interstorey drift ratio is the derivative of the lateral deflection of the continuum model.
When an MDOF model is used, the interstorey drift ratio is defined as the difference in
displacement between successive storeys over the interstorey height. It has been used as one
of the design criteria that have to be satisfied in tall building design. It becomes more
important in nowadays in performance-based design and assessment. The interstorey drift
ratio closely relates to the shear stress at the storey level and it has been widely used as an
indicator of damage. It is more related to damage than lateral displacement, so drift spectrum
is a more direct assessment tool than response spectrum.

The properties of the drift spectrum depend on the type of model for representing the
dominant deformation characteristic. A typical model for drift spectrum is the continuous
shear beam model in Fig. 2.3 (Iwan, 1997). It is a model dominated by shear deformation in
the fundamental mode. It enables a quick assessment of low-rise building structures, which
deform in shear as does a continuum shear beam.

16
A shear beam model would limit the use of the spectrum to low-rise buildings; thus a
continuum flexural-shear cantilever is adopted (Miranda and Akkar, 2006) as shown in Fig. 2.4.
The deformation characteristic of the model is governed by the degree of structural interaction.
By increasing the degree of structural interaction from nil to infinity, the model deflects from a
pure flexural to flexural-shear and from flexural-shear to pure shear. The spectrum developed
from this model is named the generalised drift spectrum. It is an improvement of the original
drift spectrum and it is more powerful.

Drift spectrum provides a direct approach to obtain the interstorey drift ratio by the
fundamental period. It facilitates performance-based assessment, which uses interstorey drift
ratio as the indicator. However the assessment is based on linear analysis and it cannot model
the inelastic behaviour of building structures.

2.5 Sheikh’s method

The inelastic displacement method and the equivalent linear method are tools for assessment of
buildings based on the SDOF system. A conversion process from the SDOF system to the
MDOF system is necessary. Moreover, approximate methods based on SDOF most of the
time only suit buildings dominated by the fundamental mode. Tall and slender buildings
would experience a significant influence from higher vibration modes and they are not
appropriate for evaluation by approximate methods that only consider the fundamental mode.

To consider the higher mode influence, which is not included in the SDOF system, and a
give direct estimation of the assessed buildings, Sheikh (2005) proposed an empirical formula
to predict the maximum interstorey drift ratio. It is an extended displacement coefficient
method for MDOF systems. The equation of the maximum interstorey drift ratio is given by
RSD1
θ max = λ2λ1λavg (2-15)
Hb
where RSD1 is the response spectrum displacement of SDOF system of the fundamental mode,
H b is the height of the assessed building, and λ2, λ1, λavg are the drift multipliers.

The drift multipliers are determined from a building database which contains the responses
from a group of full scale computer models in finite-element based software. The drift

17
multiplier λ2 indicates the contribution of the higher vibration modes relative to the
fundamental mode. Sheikh correlated this higher vibration mode drift multiplier λ2 to the ratio
RSD2/RSD1 and proposed the following formula
λ2 = 2.0(RSD2/RSD1) + 0.35 (2-16)
From this formula, the second mode is assumed to be the main influential parameter to the
higher modes. The other drift multipliers are used to convert the maximum lateral
displacement of the SDOF system to the maximum interstorey drift ratio of the corresponding
MDOF system.

The predicted maximum interstorey drift θmax is further decomposed into flexural-rotation
component θflmax and shear-drift component θshmax. Damage which is mainly attributed to
shear-drift component can then be evaluated. This approximate method is conceptually good
because it tries to incorporate the influence of higher modes empirically from a building
database and also identifies the shear-drift component as an indicator for performance
assessment.

Recently Tsang et al. (2009) have made use of the Sheikh’s proposed formula of
interstorey drift ratio to develop a rapid assessment procedure, which can be implemented in the
spreadsheet analysis. The procedure considers the influence of the site characteristics and the
effects of soil amplification. Numerical studies have shown that site conditions significantly
influence the seismic responses of building structures. Earthquakes may excite building
structures into the inelastic range of response, and the inelastic effect on the building is
modelled by an equivalent damping ratio.

2.6 Koru’s method

A seismic vulnerability assessment method for low-rise building structures was proposed by
Koru (2002). It was developed from a very large amount of post-earthquake data. Four
different methods were investigated with verification from post-earthquake data. Each
method involves only one indicator or criterion for defining the performance. The four criteria
are the following,
1. Roof drift
2. Base-shear strength coefficient

18
3. Shear strength coefficient
4. Priority index

It was found that the roof drift and priority index were more reliable as they successfully
identified most of the buildings damaged in the database. By modifying the assessment
method with suitable material properties and construction practice, the method was adjusted to
apply to the assessed region. Observations from statistical results, in which many structural
parameters like numbers of columns, area of columns, numbers of storeys, height of building,
were taken into consideration for establishing the relationship between the degree of damage
and the indicator.

The assessment could be done by engineers without extensive specialised training to


identify most of the vulnerable buildings in the region. Despite the simplicity and convenience
of the method, it is tailored made to a particular region, and is used to assess low-rise buildings
with construction practice and site characteristics of the region. The application of the method
is limited. Moreover the influence due to strengthening or remediation work cannot be
considered because the assessment method was developed from post earthquake data, in which
buildings were not designed against earthquake.

2.7 Lang’s method

The growing concern on the management of earthquake risk in Switzerland has initiated a
series of research studies. Part of the research is to establish a method to evaluate the seismic
vulnerability of existing buildings. A seismic vulnerability assessment method was then
proposed by Lang (2002). It was applied to masonry and reinforced concrete buildings whose
lateral resistance is mainly contributed by the walls.

In the assessment of a building, a wall component can be modelled as a single degree of


freedom system (SDOF) which has a capacity curve described by roof displacement and base
shear. The capacity curve is idealised as elasto-perfectly plastic. The initial stiffness, yield
strength and ductility capacity of the wall are calculated from section analysis. The overall
resistance of the buildings can then be estimated as the superposition of the capacity curves of
each wall. For example, a building which consists of four main walls can produce four

19
capacity curves. The four capacity curves are then superposed to obtain the resultant capacity
curve of the building, like in Fig. 2.5.

Degrees of damage are related to the base shear at different stages of the capacity curve,
for instance the yielding of the first wall will initiate the first damage and then further damage
occurs at the yielding of the second wall. The stages of damage are identified in the capacity
curve of the building. The capacity curve is then converted into a vulnerability curve (Fig. 2.6)
expressed as the stage of damage and the spectral displacement. Once the spectral
displacement is known, the expected degree of damage can be visualised.

The proposed method is comprehensive and the vulnerability curve suggested by the
method facilitates assessment and the retrofit arrangements. The method involves a
reasonable amount of calculations which are mainly the section analyses of the walls. Because
of its analytical approach, which includes section analysis and identifying the force path, the
method is flexible to model buildings with different characteristics. The method however is
only applicable when the structural plan of the building is simple and the coupling between
components is not significant. Otherwise it would be difficult to establish the capacity curve
by an analytical approach.

2.8 Indices based assessment

Damage index is one of the indices based assessment. According to Kappos (1997), a damage
index consists of damage variables which quantitatively describe the degree of damage of local
elements or substructures or even the global structures. It is proposed in seismic engineering
because of its rationality in assessment. It is a number ranging from 0 to 1, which represents a
damage stage varying from no damage to failure or collapse.

Damage indices usually relate to the amount of inelastic deformation and cumulative
damage due to cyclic loading. A representative one is the Park-Ang damage index (Park and
Ang, 1985), taking the form of

D=
θ max

∫ dE (2-18)
θu M yθ u

where θmax and θu are the maximum recorded rotations of a member and the ultimate rotation

20
capacity of the member under monotonic loading respectively. The second term is the ratio of

absorbed hysteretic energy ( ∫ dE ) to the product of moment at yielding and the ultimate

capacity of the member ( M yθ u ). β is an empirical coefficient obtained by regression analysis

based on experiments. It represents the effect of cyclic loading on the damage of a structural
member.

A damage index commonly considers both the deformation and energy as the factors
governing the degree of damage. It is a good concept that not only includes the deformation,
but also the energy. Therefore, use of damage index to evaluate damage may theoretically be
better than use of interstorey drift ratio. On the other hand, use of damage indices is more
complex. Difficulties arise in determination of the effect of cycling loading β because there
exists only a few choices of β from regression analysis of limited kinds of structural members.
It is inaccurate to apply the results from a small group of data to a large variety of members in a
realistic building (Williums and Sexsmith, 1995).

Moreover, empirical factors used in damage indices always have a large coefficient of
variation because of large uncertainties in the prediction of cumulative damage.
Accumulation of uncertainties may occur because the combination of local indices from
components to intermediate indices of a substructure or entire buildings is usually done through
the summation of weighted local damage indices (Michael et al., 1989). The model of the
probabilistic density function could be a better approximation to the model uncertainties in the
relationship between the damage stage and the index (Stephens and Yao, 1986). Weighted
local damage indices may not be sufficient to reflect the local failure mechanism which causes
progressive partial collapse of a building.

Although the accuracy of the global damage indices and reliability of the methodology is
doubtful, there is great application potential and the concept of indices is good. Further
development of damage indices may reduce the uncertainties in application and make it simpler
to be applied in assessment of buildings. Research on damage indices is continuing and
further descriptions can be found in the existing literature (Kappos, 1997; Ghobarah et al.,
1999).

21
Apart from damage indices, there are assessment indices proposed by Thermou and
Pantazopoulou (2011). It is very similar to the method of assessment proposed by Koru (2002)
and Lang (2002). The method is developed for low-rise reinforced concrete buildings which
sway in ground motion mainly due to the deformation of vertical members. From static
mechanics, the area of the vertical floor members and the reinforcement ratio are used to
develop the composite index which is the governing parameter in the assessment method to
determine the interstorey drift capacity of the building. The vulnerability curves of the
composite index and the interstorey drift capacity under different magnitudes of peak ground
acceleration are compared to the elastic demand obtained from the original response spectrum
method. Based on the results, the performance of the building is determined.

2.9 Capacity spectrum method (CSM)

The capacity spectrum method was first introduced in the 1970s and rapidly became a popular
seismic vulnerability assessment and seismic design tool. The development of the capacity
spectrum method is described by Freeman (1998, 2004). The principle of the method is to
compare the lateral resistant capacity of the building to the earthquake demand graphically and
the intersection is the response. For seismic evaluation, the building is said to be safe if the
lateral capacity exceeds the earthquake demand. For seismic design, it is necessary to build a
structure which has enough lateral resistant capacity and at the same time satisfies the
serviceability limit state design.

Fig 2.7 illustrates the principle of the capacity spectrum method. In the method,
earthquake demand is presented by spectra containing demand curves with different damping in
the format of spectral acceleration versus spectral displacement (ADRS). The amount of
damping depends on the hysteretic behaviour of the buildings and it can be related to the
equivalent ductility demand. The lateral capacity of the building is estimated by the nonlinear
static analysis, which is also called pushover analysis, to obtain the capacity curve.

The capacity curve, which is represented in roof displacement versus base shear, is then
converted to the ADRS format for comparison to the response spectrum. The capacity curve
in ADRS format is called capacity spectrum. Two spectra are superposed and the intersection
between them is the expected response, also called as the performance point. The performance

22
point has to satisfy the condition where the equivalent ductility corresponding to the demand
spectrum is equal to the ductility at the capacity spectrum.

The capacity spectrum method graphically presents the relationship between capacity and
demand allowing engineers to visualise the performance of a building. However, the use of
the damped elastic response spectrum as an equivalent inelastic response spectrum is doubtful
because of the unclear relationship between the damping and hysteretic energy (Fajfar, 1999).
It is also limited to buildings subjected to a small influence by higher vibration modes because
higher vibration modes are not considered in the analysis.

2.10 Nonlinear static analysis

Pushover analysis
Pushover analysis is a powerful and popular tool nowadays to assist in the evaluation of the
seismic performance of a structural system. Because of its reliability in capturing the
behaviour of a building, it is widely accepted and promoted by researchers and authorities as a
standard for design and evaluation (ATC 40; FEMA 310; FEMA 356). The main purpose of
the pushover is to estimate the lateral resistant capacity of a building and to present the capacity
graphically for capacity spectrum analysis.

In the analysis of a multi-storey building, which is an MDOF system, inertia force on the
building induced by the earthquake is modelled as invariant distributed lateral force acting
along the building height. The magnitude of the force is continuously increased to simulate
the intensity of the ground motion and the responses are recorded during the process. With the
advanced computer software, the internal force and displacement in the members, substructures
and the entire building are observed. Hence the weaknesses and the deficiencies of the
structural system can be revealed. It provides guidance for strengthening and retrofitting.

By monotonically increasing the magnitude of the invariant force, a capacity curve under
the specific invariant force distribution is obtained. The curve is usually presented in base
shear and roof displacement. The invariant distributed force is usually derived from the
expected deformed shape of the building in the fundamental mode, which is not valid for a
building with a significant higher vibration mode effect or a building with asymmetric plan

23
shape or discontinuity in strength and stiffness. The limitations of pushover analysis have
been described (Krawinkler and Seneviratna, 1998; Huang and Kuang, 2010). There is also
debate on the shape of the distributed force because it reflects the expected failure mechanism
of the building and results in the capacity curve (Mwafy and Elnashai, 2001). Different
distributed forces would give different capacity curves.

Model pushover analysis


Pushover analysis was further modified to Model Pushover Analysis (MPA) (Chopra and Goel,
2002; Chintanapakee and Chopra, 2003). The MPA aims to overcome the deficiency of
pushover analysis in accounting for the influence of higher modes while keeping to a
reasonable calculation effort. For buildings with N degrees of freedom, the deformed shape of
the building before yielding is the linear combination of the N independent modes. This
assumption, which is not valid after yielding, is further extended to the nonlinear behaviour of
the building because of insignificant error from experimental observations.

The influence of each mode on the building is represented by its invariant force
distribution, which has identical shape to the mode shape. Then N times of pushover analyses
are performed with N different invariant distributed forces respectively. Peak responses
resulting from respective modes under the effective earthquake force are then combined
according to the combination rule to obtain the maximum responses of the building. Pushover
analysis of invariant force distributions of higher mode however are difficult to conduct
because of great numerical instability.

An improved approach is the modified modal pushover analysis (Chopra et al., 2004). In
the analysis, modal contributions of higher modes in the responses are assumed to be linear
elastic. Therefore only nonlinear static analysis is performed on the fundamental mode.
Numerical studies also found that the modified modal pushover is more applicable and gave
results close to the results from other rigorous analysis.

2.11 Nonlinear response history analysis

Nonlinear response history analysis differs from nonlinear static analysis (pushover) in the
simulation of earthquake load. Unlike the nonlinear static analysis modelling of earthquake

24
load as a monotonic increasing distributed lateral force, earthquake load in nonlinear dynamic
analysis is completely captured by using the ground motion record, which is a function of time.
This procedure is rigorous and involves less uncertainty because of using ground motion
records and the material hysteretic model, but it is a time consuming procedure due to the step
by step evaluation of the building responses in time domain.

This method is found to be very difficult to apply because the numerical convergence is
difficult to be achieved when a complicated material hysteretic model is used. If a full-scale
model is used in the analysis, the time of analysis is extremely demanding and it does not
guarantee the obtaining of results. Even though the prediction of building responses is more
reliable, the responses are sensitive to the characteristic of the ground motion record.
Responses only reflect the performance under a single earthquake and may not be consistent in
other earthquakes. This drawback can be overcome by averaging a group of ground motions.

Incremental dynamics analysis


Vamvatsikos and Cornell (2002) proposed an evaluation tool of the incremental dynamic
analysis (IDA). For every input earthquake motion record, it is scaled to different intensities
and is used to perform nonlinear dynamic analysis. A set of responses, which corresponds to
different intensities, is recorded. The relationship between magnitude and response is hence
established under a particular ground motion. By repeating the above procedure for different
ground motions, spectra similar to those in Fig. 2.8 can be obtained. It contains thirty curves
of maximum interstorey drift ratio versus spectral acceleration resulting from different
excitations. Then from a statistical approach, three representative fractile curves are produced
in Fig. 2.9. Under specific first mode spectral acceleration, the possible magnitude of the
interstorey drift ratio can be obtained with occurrence possibility.

The power of the IDA is the linkage to the reliability analysis which accounts for
uncertainties. However it is time consuming to perform such large amount of nonlinear
response history analysis. To overcome the deficiencies in incremental dynamic analysis, the
IDA is modified (Vamvatsikos and Cornell, 2006). The modification uses nonlinear static
analysis to approximate the results of nonlinear dynamic analysis (Sang and Chopra, 2006).
The IDA curves of an SDOF model, nonlinear dynamic analysis, can be estimated from a
quadrilinear capacity curve of an SDOF model obtained by nonlinear static analysis through
empirical equations. The IDA curves can be further used for estimation of the seismic demand

25
and capacity of MDOF systems (Vamvatsikos and Cornell, 2005). This new approach reduces
the computational effort, but it is then subjected to the limitations and drawbacks of pushover
analysis. Reliability is reduced because of more estimation being required.

2.12 FEMA 154

FEMA (Federal Emergency Management Agency) 154 (2002) refers to “Rapid Visual
Screening of Buildings for Potential Seismic Hazards”. It is a standard rapid visual screening
(RVS) tool for evaluation of building’s seismic performance. It helps engineers who are
experienced in seismic design to identify potential hazardous buildings for further deep
evaluation through a rapid screening means. The RVS requires the engineer to conduct a
“sidewalk survey” of buildings based on visual observation of the buildings from the exterior,
and the interior if possible. Meanwhile, the engineer needs to complete a data collection form
and determine the structural score which is the index used for the identification of the seismic
hazard.

The RVS method is established based on past earthquake records in the United State and
the data on the performance of buildings in earthquakes. Since the seismicity of regions of the
United States is classified into high, moderate and low, based on ground motions having a 2%
probability of exceedance in 50 years, three different data collection forms as shown in Fig.
2.10 are then developed using the data from each of these regions. Thus, this method can be
applied to regions of different levels of seismicity, but it is limited to regions with similar local
construction practice. The extension of this method to other regions requires modifications by
empirical coefficients, capturing the difference in construction practice and site characteristics.

Fifteen general types of buildings are considered in the data collection form. For each
type of building, a basic structural score is assigned based on the damage factor established in
the ATC 13 (an Applied Technology Council report, Earthquake Damage Evaluation Data for
California). The basic structural score is related to the probability of major physical damage,
which is defined by damage exceeding 60 percent of the building value in ATC 13. In the
RVS procedure, the engineer first selects the type of assessed building, which is associated to
the basic structural hazard score, through identification of the structural lateral-load-resisting
system and the major structural materials. To complete the form, the engineer needs to walk

26
around the building to verify and update the building identification information such as its
occupancy, size, shape and the soil type it is constructed on. Also, sketches of the plan and
elevation, as well as a photo of the building, need to be attached to the form.

Then, the engineer has to select appropriate score modifiers shown in Fig. 2.10, which are
related to the observed performance attributes and are derived from statistics using expert’s
opinions. Finally, the score modifiers are added to or subtracted from the basic structural
hazard score to obtain the final structural score. The final score can be interpreted as having 1
event of major damage in 10final score events. The probability of major damage is hence equal to
1/10final score. For example, a final score equal to 3 implies that there is a probability of 1 in 103
(i.e. 1 in 1000) that the building will collapse if such ground motions occur. The higher the
score means the better the expected seismic performance.

Based on present seismic design criteria, a final score of 2 is suggested as a “cut-off” value.
Buildings having a final score higher than the cut-off score have acceptable performance in
earthquakes. Buildings failing to meet the cut-off score are temporarily treated as seismically
hazardous and need further detailed assessment. The cut-off score can also be raised based on
the required level of safety imposed by users or society.

FEMA 154 (2002) is comprehensive and rapid. It requires little resources and is very
useful to rank buildings according to the structural score. Because of these attributes, the
assessment is also subjected to large uncertainties. They are the subjective judgment that is
highly dependent on the experience of the engineer, unavailability of detail interior inspection
and statistical uncertainties in the method. To avoid misleading assessment, the method tends
to underestimate building performance and always stands on the safe side.

2.13 Seismic demand for asymmetric-plan buildings

It is well agreed that structural asymmetry results in torsion on buildings, which may increase
deflections and stresses in members. The worst condition is the failure of a member due to
additional stress when translation and torsion happen simultaneously under ground motion.
An uncoupled period ratio was defined by Yoon and Smith (1995 a) as the pure translational
period to pure torsional period to determine the influence of torsion and translation. The ratio

27
was derived from wall-frame structures with the corresponding engineering model represented
by a shear-flexure torsion-warping cantilever which does not consider any interaction between
the systems.

The uncoupled period ratio was an indicator for the degree of translational and torsional
coupling. If the ratio was unity, the maximum degree of translational and torsional coupling
had a high possibility of occurring. A small ratio implies that the torsional influence can be
separately considered and the torsional influence is less significant than the translation. In a
further study by Yoon and Smith (1995 b) an improved method using the coupled period ratio
was proposed. The period ratio may become an indicator to assist engineers in determining the
significance of torsional influence in assessment.

Earthquake induced energy is mainly dissipated through structural damping and hysteretic
behaviour when the building has been excited into an inelastic response range. Hysteretic
dissipation ratio is defined as the ratio of the total hysteretic energy to the total energy
dissipated by the system in an earthquake. It measures the number and duration of yield
excursions in a system and hence reflects the damage potential for the system. A system
dissipating a high proportion of seismic energy through hysteretic response is more vulnerable.
In the study of Chandler and Correnza (1996), a system with large eccentricity dissipates larger
portion energy in viscous damping action as a whole, but it results in a high portion of seismic
energy being dissipated by particular elements, which is undesirable for seismic safety as local
failure arises.

To assess an asymmetric-plan building, a modal pushover analysis procedure was


suggested by Chopra and Goel (2004). Invariant force distributions which follow the mode
shapes of the three dimensional model are used. The invariant forces, including the lateral
forces and torque, are applied to the model to obtain the load deformation curve. The modal
responses of the first few terms of the modal expansion are summed up by the completed
quadratic combination rule to predict the inelastic response. The numerical investigation
showed that the suggested method is practical, yet reliable, for some asymmetric-plan
buildings.

Torsion may be highly coupled to translation and this makes assessment more complicated.
Although there are many studies concerning torsion, practical assessment methods mainly use

28
nonlinear static analysis. They are limited to particular types of structures and have to satisfy
certain assumptions. It is believed that torsion can be revealed from the response history
analysis, but it is difficult to be performed and the results are sensitive to the input model.
Torsion response in the inelastic range would be a challenging task in seismic engineering
especially in assessment. This is the reason why, in the current code of practice, the design of
a highly asymmetric-plan building is prohibited and detailed assessment is recommended for
such buildings.

2.14 Research significance

Comprehensive and quick seismic assessment methods are necessary for minimising and
managing earthquake disaster risks because it is definitely not efficient to evaluate every
building with the same amount of effort and time. It is also not appropriate to retrofit and
strengthen every building against earthquake risk without assessment. The priority order of
strengthening deficient buildings depends on the question “how urgent is the situation in regard
to a particular building?”. Although concern on earthquake risk is rising, the answer is still in
not clear.

Existing comprehensive seismic assessment methods are limited to low-rise buildings and
elastic behaviour. Detailed analysis may be reliable, accurate and applicable for most kinds of
buildings, but they are rather demanding in computational time and professional knowledge in
regard to modelling and interpreting. Therefore, detailed analysis is not the appropriate first
step in the assessment of buildings. A comprehensive seismic assessment should first be
carried out to identify the urgency for detailed analysis.

Developing a comprehensive seismic assessment method for high-rise buildings, which


can be applied to the inelastic response range, is valuable in the allocation of limited resources
and assessment. Research in this area is necessary and would be beneficial to regions without
seismic consideration in the building design.

29
Table 2.1 Assessment level
Level one Level two Level three

Cost Increasing cost

Application All buildings Suspicious buildings Deficient buildings


Response spectrum Sheikh (2005) Damage indices
Examples Rapid visual screening method Lang (2002) Capacity spectrum method
Koru (2002) Drift spectrum Nonlinear dynamic analysis

Fig. 2.1 A single-degree of freedom system

Fig. 2.2 Multi-degrees of freedom system

30
Shear beam

Fig. 2.3 Shear beam model

Fig. 2.4 Continuum flexural-shear cantilever

31
Fig. 2.5 Resultant capacity curve from superposition (Lang, 2002)

Fig. 2.6 Vulnerability curve of a building (Lang, 2002)

32
Fig. 2.7 Capacity spectrum method

Fig. 2.8 Maximum interstorey drift versus spectral acceleration


(Vamvatsikos and Cornell, 2002)

33
Fig. 2.9 Fractile curves
(Vamvatsikos and Cornell, 2002)

34
Fig. 2.10 Data collection forms for the three designated seismicity regions

35
Fig. 2.11 Basic score and score modifiers

36
Chapter 3 

Quick Estimate of Fundamental Period 

of Tall Buildings 

3.1 Introduction

In preliminary seismic design and seismic assessment, it is important to evaluate accurately the
natural fundamental frequency of vibration of a building structure, since it is directly related to
the corresponding seismic forces and deformations. Traditionally, the fundamental frequency
required in the analysis is generally estimated from empirical equations which are normally
expressed in terms of the height of a building only. For example, the natural fundamental
period of vibration for moment-resisting frames T1 = 0.1N sec., where N is the number of
storeys, where the effects of the interaction of different structural forms in the building,
rigidities of the structure and intensities of the gravity and applied lateral loads, etc., on the
vibration characteristics are not considered in the calculation. This will clearly result in an
inaccurate estimate of natural vibration frequency or period of a building.

37
Realising the consequences of inaccurate estimate of the natural frequency, improved
formulas have been derived. Rakesh et al. (1997) calibrated the period data of both reinforced
concrete and steel moment-resisting frames to produce two different sets of empirical formulas
depending on the construction materials. Later Goel and Chopra (1998) proposed another set
of empirical period formulae, including the upper and lower bound values for concrete shear
wall buildings, in which the influence of the height of building structure and the area of the
structural walls at the ground floor are considered. To obtain a more accurate estimate of the
vibration frequency, more structural parameters, such as the mass and the degree of structural
interaction, have been implemented to the period formula based on rationale engineering
models.

Building structures can be represented by engineering models developed from the


continuum techniques for conducting simplified static and dynamic analyses. For instance
Rutenberg (1975) derived an expression for the natural frequency of coupled shear walls and
Wang et al. (1992) developed earthquake design spectra for wall frame structures based on a
flexural-shear cantilever. It has been recognised that the continuum modelling is a very
useful tool for analysis of tall buildings (Smith and Coull 1991) and has been used for the
development of various drift spectra for preliminary seismic assessment. This leads Iwan
(1997) to establish a drift spectrum demand measure for earthquake ground motions, and
Miranda and Akkar (2006) further developed the generalised interstorey drift spectrum
approach.

Despite the rapid development of modelling techniques for buildings in structural


engineering, the connection between the model and the corresponding building structure is
rarely discussed, which sometimes leads to difficulties in practical use. Instead of using
rigidity and structural interactions, which are the by-products of the system, more practical
parameters that are design variables should be used in modelling. In this chapter, a theoretical,
yet practical, model is developed for the quick prediction of the natural fundamental period of
vibration for tall building structures.

The proposed method is formulated based on continuum modelling, where a tall-building


structure is considered as a continuous flexural-shear cantilever. It is presented in a closed
form mathematical expression for calculating the natural fundamental period T1, which is a
function of the interaction of different structural forms in the building, structural height, top and

38
interstorey drifts, and the intensities of the applied loading. Numerical studies pertaining to
determining the fundamental periods of moment-resisting frames, shear walls and a wall-frame
structures show that the results from the proposed formula agree very well to the finite-element
analysis (FEA). The proposed analysis has been shown to provide a simple and quick, yet
accurate, means of determining the fundamental period of vibration for tall buildings behaving
elastically.

3.2 Continuum flexural-shear cantilever

In the analysis, a tall-building structure may generally be modelled as a two-dimensional


continuum model (Smith and Coull, 1991) by assuming that the structural properties, including
strength and stiffness, are uniform in height with negligible torsional effect. A typical model is
the continuum flexural-shear cantilever as shown in Fig. 3.1. This continuum model consists
of a flexural cantilever and a shear cantilever, where the flexural cantilever is a representation
of the flexurally dominant systems, such as shear walls, while the shear cantilever represents
shear dominant systems, such as moment-resisting frame structures. The two cantilevers are
connected by axially rigid links, which may be represented by the axially rigid continuum that
transmits horizontal forces only. Thus, the flexural and shear cantilevers will deflect
identically under the action of horizontal loading.

The deformation characteristic of the continuum flexural-shear cantilever is mainly


governed by the degree of structural interaction between the flexural and shear cantilevers,
defined by
GA
α=H (3-1)
EI
where EI and GA are the flexural and shear rigidities of the building respectively, and H is the
total height of the structure.

The deformation profile of the flexural-shear cantilever is significantly influenced by the


degree of structural interaction α. The structure may deform with a pure flexural shape when
α = 0 and a pure shear shape when α becomes infinite, as shown in Fig. 3.2. In practice, the
value of α generally ranges from 0 to 2 for pure shear-wall buildings, 5 to 20 for
moment-resisting frames, and 1.5 to 6 for wall-frame structures (Miranda and Reyes, 2002).

39
3.3 Dynamics properties of the model

A free-body diagram of the flexural-shear cantilever at time t under free vibration, with the
mass replaced by its inertia force is shown in Fig. 3.3. Consider the dynamic equilibrium of
both flexural and shear elements in the flexural-shear cantilever, so the two equilibrium
equations in horizontal direction can be developed.

For the flexural element,


∂ 2u ∂V
m f dx + V f + f dx − V f + ndx = 0 (3-2)
∂t 2
∂x

and for the shear element,


∂ 2u ∂V
ms dx + Vs + s dx − Vs − ndx = 0 (3-3)
∂t 2
∂x
where mf and ms are the distributed masses of the flexural and shear elements respectively, u is
the relative lateral displacement, n is the stress of the axially rigid continuum, and Vf and Vs
are shear forces of the flexural and shear elements respectively given by
∂M ∂ ⎛ ∂ 2u ⎞ (3-4)
Vf = = ⎜ EI 2 ⎟
∂x ∂x ⎝ ∂x ⎠

∂u
Vs = −GAχ = −GA (3-5)
∂x
in which M is the bending moment and χ is the shear strain. By combining the equations of
equilibrium of the elements represented in Eqs (3-2) and (3-3) and then substituting the shear
forces given by Eqs (3-4) and (3-5), the governing equation of motion of the continuum
flexural-shear model under free vibration can be derived and simplified as
∂ 2u ∂ 4u ∂ 2u
m 2 + EI 4 − GA 2 = 0 (3-6)
∂t ∂x ∂x
where m = mf +ms is the distributed mass.

Let the relative lateral displacement be


u( x, t ) = φ ( x ) q ( t ) (3-7)

where φ(x) is the shape function and q(t) is the modal coordinate, with different independent
variables. Eq. (3-6) can be solved by the technique of separation of variables, and thus
decomposed into two equations with variables x and t respectively,

40
q′′ ( t ) + ω 2 q ( t ) = 0 (3-8)

EI ( 4) GA
φ ( x) − φ ′′ ( x ) − ω 2φ ( x ) = 0 (3-9)
m m
where ω is the natural vibration frequency of the continuum model.

The fourth-order ordinary differential equation (3-9) can be solved by considering four
appropriate boundary conditions, resulting in equation Eq. (3-10) that contains the natural
vibration characteristics of the flexural-shear cantilever, has to be satisfied under the fixed-base
condition,
⎛ α4 ⎞ α2
2 + ⎜ 2 + 2 2 ⎟ cos γ i cosh β i + sin γ i sinh β i = 0 (3-10)
⎝ βi γ i ⎠ βi γ i
where
1
γ i2 = ⎡ −α 2 + α 2 + 4mH 4ωi2 / ( EI ) ⎤ and βi2 = α 2 + γ i2
2⎣ ⎦
The natural periods and shape functions corresponding to the characteristic equation (3-10) are
then given by

m 2π
Ti = H2 (3-11)
EI γ (γ + α
i
2
i
2 2
)
⎛ x⎞ ⎛ x⎞ ⎛ x⎞ ⎛ x⎞
φi ( z ) = Ci1 cos γ i ⎜ ⎟ + Ci 2 sin γ i ⎜ ⎟ + Ci 3 cosh β i ⎜ ⎟ + sinh βi ⎜ ⎟ (3-12)
⎝H⎠ ⎝H⎠ ⎝H⎠ ⎝H⎠
in which Ci is a coefficient composed of α, βi and γi.

3.4 A model for quick estimate of fundamental period T1

It is well known that the flexural rigidity of a building has a significant effect on the natural
fundamental frequencies of vibration. A free body diagram of a flexural-shear cantilever
subjected to an inverted triangular load with a maximum intensity of wo, which gives the
smallest practical lateral load, is shown in Fig. 3.4. The governing equation with respect to
lateral deflection can then be derived and is given in a similar form to that Eq. (3-6),
∂4 y ∂ 2 y wo x
EI − GA = (3-13)
∂x 4 ∂x 2 H
where y is the static lateral deflection. With four appropriate boundary conditions, Eq. (3-13)
can be solved and the lateral deflection under the fixed-base condition is given by

41
wo H 4 ⎡⎛ α sinh α + 2 sinh α ⎞ ⎛ αx ⎞
y ( x) = ⎢⎜ 2 cosh α − α cosh α ⎟ ⎜ cosh H − 1⎟
EIα 4 ⎣⎝ ⎠⎝ ⎠
(3-14)
αx α2 ⎛ x ⎞ ⎛α2 ⎞ x ⎤
3
⎛1 α⎞
+ ⎜ − ⎟ sinh − ⎜ ⎟ +⎜ − 1⎟ ⎥
⎝α 2 ⎠ H 6 ⎝H⎠ ⎝ 2 ⎠ H ⎥⎦

The flexural rigidity EI is directly related to the design roof drift ratio RΔ, which is the ratio
of y(H)/H, and can be obtained by substituting x = H into Eq. (3-14) for derivation,

wo H 3 1 ⎛ α 2 sinh α α sinh α + 2 ⎞ (3-15)


EI = 4 ⎜
+ − ⎟
RΔ α ⎝ 3 α cosh α 2 cosh α ⎠

Substituting Eq. (3-15) into the equation of natural periods of vibration, Eq. (3-11), and
only considering the first mode of vibration (i = 1) lead to the following formula of the natural
fundamental period

mH
T1 = f (α ) i =1
RΔ (3-16)
wo

where
2πα 2
f (α ) i =1
= (3-17)
⎛α sinh α
2
α sinh α + 2 ⎞
γ 12 (γ 12 + α 2 ) ⎜ + −
⎝ 3 α cosh α 2 cosh α ⎟⎠

f(α) is a function of the degree of structural interaction between the flexural and shear
cantilevers and considered as an index of structural interaction. The value of f(α) varies
between 5.9 and 6.8 , as shown in Fig. 3.5, if only the first mode of vibration (i = 1) is considered,
giving a maximum difference of 15%, which indicates the fundamental period T1 can be
changed up to 15% due to the structural interaction.

42
3.5 Prediction of degree of structural interaction

The degree of structural interaction between the flexural and shear cantilevers, α, may be
calculated using Eq. (3-1) by identifying the flexural and shear rigidities of structural systems.
A suggested approach was proposed by Smith and Coull in their book “Tall Building Structures:
Analysis and Design”. The flexural rigidity of a building structure can be regarded as the total
flexural rigidities of the shear walls and the related components, which is
n
EI = ∑ Ei I i (3-18)
i =1

where n is the number of walls, E is the elastic modulus of the material and I is the moment of
inertia of the section.

The shear rigidity of a building is contributed from the moment resisting frames
constructed by columns and beams. The equation of shear rigidity is
12 E
GA = (3-19)
⎛1 1⎞
h⎜ + ⎟
⎝B C⎠
N
⎛I ⎞ M
⎛I ⎞
where h is the storey height, B = ∑ ⎜ b ⎟ , C = ∑ ⎜ c ⎟ ,N is the number of beams, M is the
i =1 ⎝ L ⎠ i i =1 ⎝ h ⎠ i

number of columns, L is the length of the beam. Ib and Ic are the moment of inertia of the beam’s
section and column’s section, respectively. Normally, the exact different lateral systems in a
building are very difficult to be identified due to structural couplings in the systems. Eq. (3-1)
would not be recommended generally for determining the value of α . Deflection, which
results from the action of external forces under the resultant lateral system, could be used to
predict the degree of structural interaction.

The mode of deformation is governed by the degree of structural interaction. By


increasing the degree of structural interaction, the model deforms from a flexural mode to a
shear mode. Therefore the degree of structural interaction can be predicted from the
deformation characteristic. Under lateral loading, the model deforms and the deformation
characteristic can be changed by the degree of structural interaction. By minimising the
difference in deflection between the building and the model, the degree of structural interaction
corresponding to the building can be obtained.

43
The deflection of the continuum model under uniform load shown in the appendix A2 is

wH 4 ⎡α sinh α + 1 ⎛ αx ⎞ αx 2⎛ x x 2 ⎞⎤
y(x ) = ⎢ ⎜ cosh − 1⎟ − α sinh + α ⎜
⎜ H 2 H 2 ⎟⎟⎥
− (3-20)
EIα 4 ⎣ cosh α ⎝ H ⎠ H ⎝ ⎠⎦
where w is the magnitude of the uniform load. The roof displacement is
wH 4 ⎡ α 2 α sinh α + 1⎤
y (H ) = ⎢1 + − ⎥ (3-21)
EIα 4 ⎣ 2 cosh α ⎦

A dimensionless expression which is the ratio of deflection to the roof displacement is given by

α sinh α + 1 ⎛ αx ⎞ αx ⎛ x x2 ⎞
⎜ cosh − 1⎟ − α sinh + α 2 ⎜⎜ − ⎟
y (x ) cosh α ⎝ H ⎠ H ⎝ H 2 H 2 ⎟⎠
= (3-22)
y (H ) α 2 α sinh α + 1
1+ −
2 cosh α
Eq. (3-22) is the deflection shape of the model where the roof displacement is adjusted to be
unity. If the model is identical to the building, the difference in deflection shape between the
model and the building should be zero. This is equivalent to the following mathematical
expression
n
y ( xi ) Y ( xi )
Δ=∑ − =0 (3-23)
i =1 y (H ) Y (H )

where Δ is the sum of the absolute differences, Y is the deflection of the building and i is the
floor number.

Since it is a very strict constraint where Δ is equal to zero, a more appropriate approach is
to minimise Eq. (3-23) by changing the degree of structural interaction. Although this method
would give a degree of structural interaction that can best describe the mode of deflection of the
building, it is not possible to obtain displacement of every floor. An alternative is to obtain the
displacement from the corresponding multi-degree-of-freedom (MDOF) model of the building.
Since buildings are usually designed in a finite-element-based program, the particular
displacement can be extracted from the model for analysis.

When the corresponding MDOF model of the building is not available, a more convenient
way is to approximate the value of α from the deformation profile of the building, which is
characterised by RΔ/Rδm, a ratio of the design roof drift ratio to the design maximum interstorey
drift ratio of the building. When a building is subjected to an inverted triangular load wo, the
interstorey drift ratio can be obtained by differentiating Eq. (3-14),

44
wo H 3 ⎡⎛ α sinh α + 2 sinh α ⎞ αx
Rδ ( x ) = 3 ⎢⎜
− ⎟ sinh
EI α ⎣⎝ 2 cosh α α cosh α ⎠ H
(3-24)
α x α ⎛ x ⎞ ⎛ α 1 ⎞⎤
2
⎛1 α⎞
+ ⎜ − ⎟ cosh − ⎜ ⎟ + ⎜ − ⎟⎥
⎝α 2 ⎠ H 2 ⎝ H ⎠ ⎝ 2 α ⎠ ⎥⎦

The design maximum interstorey drift ratio Rδm is obtained by substituting xo into Eq. (3-24),
where xo is the height at which the maximum interstorey drift occurs. The variation of α
against the drift ratio RΔ/Rδm is plotted in Fig. 3.6.

At the preliminary design stage, the roof drift ratio RΔ and the maximum interstorey drift
ratio Rδm may be assumed as H/500 and h/350 respectively, where h is the storey height, and are
consistent with those recommended in design codes of practice. The value of α may then be
quickly determined using Fig. 3.6. With the given value of α, the index of structural interaction
f(α) is found in Fig. 3.5. Hence, the natural fundamental period T1 can be conveniently
calculated using Eq. (3-16).

It is seen from the equation of the natural fundamental period, Eq. (3-16), that when this
equation is used to determine the value of T1 of a building, the effects of the structural
interaction, roof and interstorey drifts, gravity and lateral loads, and height of the structure on
the fundamental period of vibration have been considered.

3.6 Degree of structural interaction in non-uniform buildings

A constant degree of structural interaction implies uniform rigidity along the height and it is an
approximation to general buildings which are not uniform. The changes of stiffness along the
height most of the time is not significant because the sizes of the structural components are
restricted to provide minimum strength in the design. Significant change in stiffness only
occurs in building structures with transfer storey. A 60-storey building with non-uniform
stiffness and a transfer storey shown in Fig 3.7 is analysed to study the influence of a constant
degree of structural interaction on the response. The building is constructed by three main
floor plans. They are the floors below the transfer storey, the transfer storey and the floors
above the transfer storey. The main differences between the three floor plans are the number of
columns, the arrangement of columns, and dimensions of the columns and beams at the
periphery.

45
Since both fundamental period and the degree of structural interaction are the key
parameters for a continuum model, they have to be similar to those of the equivalent MDOF
model, so that they are equivalent to the MDOF model. A method for constructing a
continuum model from an MDOF model is proposed. The flexural rigidity and shear rigidity,
which construct the degree of structural interaction, are estimated from the deflection
characteristic of the equivalent MDOF model. The deflection of the continuum model under a
point load is given by (Appendix A4),
PH 3 PH 2 PH 3 αx PH 3 αx
y (x ) = tanh α − x− tanh α cosh + sinh (3-25)
EIα 3
EIα 2
EIα 3
H EIα 3
H
where P is the magnitude of the point load at the roof.

Since the continuum model has to be equivalent to the MDOF model, the difference in
deflection between the two models should be zero. This is equivalent to minimising the
differences in deflection between two models by changing two variables: the flexural rigidity
(EI) and the degree of structural interaction (α).
n
Δ = ∑ y ( xi ) − Ym (xi ) (3-26)
i =1

where Ym is the deflection of the MDOF model. Since both models would have similar
fundamental frequency, Eq. (3-10) of the characteristic equation containing the fundamental
frequency should be held as a minimum, which is given by
⎛ α4 ⎞ α2
2 + ⎜⎜ 2 + 2 ⎟ cos γ cosh α 2
+ γ 2
+ sin γ 1 sinh α 2 + γ 12 = 0 (3-27)
⎝ ( )2 ⎟
α + γ1 γ1 ⎠
2 1 1
γ α + γ1
2 2

where

4mH 4ω12
−α2 + α4 +
γ 12 = EI (3-28)
2
When the flexural rigidity (EI) and the degree of structural interaction (α) are obtained, the
shear rigidity can be calculated from Eq. (3-1).

By the proposed method, the rigidities of different floor plans are estimated from uniform
models made with the respective floor plans. The rigidity of the equivalent continuum model,
which has a similar deflection and fundamental frequency, are also constructed. The variation
of rigidity in the MDOF model and the equivalent continuum model are compared in Fig 3.8. It
can be seen that there is a significant change in rigidity at the transfer storey and the change is at

46
least twice that of the other floors while the change of rigidity above the transfer storey is
gradual. The rigidity of the continuum model is close to that of the floors above the transfer
storey.

Both the MDOF model and the equivalent continuum model are then subjected to two
selected excitations in Fig 3.9 from the “PEER strong Motion Database”. The peak ground
acceleration of both ground motion records is approximately 0.3g. The deflection history of
both models is then observed at three different floor levels: 207.7 m (Fig. 3.10), 129.7 m (Fig.
3.11) and the transfer storey 17.5 m (Fig. 3.12).

The comparisons in Figs. 3.10 and 3.11 show that the deflection of the equivalent
continuum model is close to that of the MDOF model in both phase and magnitude. However
at the transfer storey, the continuum model could not capture the response of the MDOF model
as shown in Fig. 3.12. The reason for this difference is the abrupt change of the stiffness at the
transfer storey. When a non-uniform MDOF model is replaced by its corresponding
continuum model, the response of the equivalent continuum model is close to the MDOF model
at the region where the rigidity is similar, but at the region where rigidity changes abruptly, the
responses predicted by the continuum model may not be reliable.

Therefore, it is concluded that an equivalent continuum model could be used to predict the
response of general buildings which do not have significant change in rigidity. For building
structures with significant change in rigidity, the equivalent model could not give a reliable
prediction at the location where rigidity changes abruptly. Thus a constant degree of structural
interaction can be used to establish an equivalent continuum model that represents general
building structures.

3.7 Numerical studies

The fundamental periods of vibration of three reinforced concrete tall buildings, including a
10-storey moment-resisting frame, a 20-storey shear-wall structure and a 30-storey wall-frame
structure, are calculated using the proposed formula and analysed by the finite element method,
employing ETABS (Computer and Structures Inc., 2005). The three buildings have the
identical storey heights of 3 m, and the floor plans are shown in Fig. 3.13. The dimensions of

47
structural members, as in the practical design, are given in Table 3.1, while the reinforcement
ratios of the members satisfy the strength and minimum reinforcement ratio requirements,
which are 0.25% for both the beam and wall sections, and 0.4% for column sections. The
design concrete cube strength is 30 MPa with density of 2400 kg/m3, and the reinforcement
yield strength is 460 MPa.

The three example buildings are designed to resist a dead load (DL) of 4.5 kN/m2, a live
load (LL) of 8 kN/m2, and lateral loads (WL) of 2.5 kN/m2, 3.0 kN/m2 and 3.5 kN/m2 at the roof
level respectively, which decrease linearly to 0 at the base level. Three load combinations are
considered in the analysis:
(1) 1.4DL + 1.6LL,
(2) 1.2DL +1.2LL + 1.2WL, and
(3) 1.0DL + 1.0LL + 1.0WL.

To evaluate the fundamental period of vibration T1 for the buildings using the proposed
formula given by Eq. (16), the structural parameters, including the degree of structural
interaction α and its corresponding index f (α ) i =1
, design roof drift RΔ and design maximum

interstorey drift ratio Rδm, as well as the distributed mass m, are first calculated, as shown in
Table 3.2. FEM models of the buildings are shown in Fig. 3.14.

Comparison of the results of the natural fundamental periods predicted by the proposed
formula and the finite-element analysis by ETABS is made and presented in Table 3.3. It can
be seen from Table 3.3 that the two sets of the results show very good agreement.

48
3.8 Concluding remarks

In this chapter, a model is developed for the quick estimate of the natural fundamental period of
vibration for tall building structures. The model is based on the continuum modelling, where a
tall-building structure is considered as a continuous, interactive flexural-shear cantilever. The
proposed model is presented in a closed form mathematical expression for use of calculating the
natural fundamental period T1, in which the effects of the interaction of different structural
forms in the building, design roof drift and maximum interstorey drift ratios, intensities of
applied loading and height of the structure on the fundamental period of vibration have been
considered. Numerical studies pertaining to predicting the fundamental periods of the tall
moment-resisting frames, shear walls and wall-frame structures show that the results from the
proposed formula and finite-element analysis agree very well. The proposed formula has been
shown to provide a simple and quick, yet accurate, means of estimating the fundamental period
of vibration for tall building structures, in particular at the preliminary seismic design and
seismic assessment stage.

49
Table 3.1 Dimensions of Structural Members
Column (m) Beam (m) Wall (m)
10-storey 30-storey 10-storey 30-storey 20-storey 30-storey
Floor frame wall-frame frame wall-frame shear walls wall-frame
st th
21 to 30 0.3 × 0.3 0.4 × 0.25 0.4
11th to 20th 0.4 × 0.4 0.5 × 0.3 0.3 0.5
st th a a
1 to 10 0.3 × 0.3 0.5 × 0.5 0.4 × 0.25 0.6 × 0.3 0.4 0.6
a st nd
In the 10-storey frame, columns at 1 floor and central columns at 2 floor are
0.35×0.35 m.

Table 3.2 Structural parameters and applied lateral load


m H wo RΔ Rδm
Buildings (kg/m) (m) (kN/m) (10-3) (10-3) α f (α ) i =1

10-storey frame 21750 30 22500 0.95 1.34 7 6.58


20-storey shear
walls 36300 60 30000 0.32 0.39 2 6.05
30-storey
wall-frame 89160 90 52500 1.10 1.30 2 6.06

Table 3.3 Comparison of fundamental periods T1 predicted


by proposed formula and finite-element analysis
Buildings Proposed (sec) FEA (sec) Difference (%)
10-storey frame 1.09 1.15 6
20-storey shear walls 0.92 0.93 0
30-storey wall-frame 2.25 2.45 9

50
Axially-rigid links

Flexural cantilever

Shear cantilever x

Fig. 3.1 Continuum flexural-shear cantilever

α=0 α=∞
Height

Height

Deflection Deflection

Fig. 3.2 Deformation shapes of flexural-shear cantilever

51
∂M
M+ dx
∂x

∂V f
Vf + dx
∂x
ndx
∂ u
2
dx m f dx
∂t 2
Vf

(a) Flexural element

∂Vs
Vs + dx
∂x
ndx χ
∂ 2u
m s dx
∂t 2
Vs

(b) Shear element


Fig. 3.3 Free body diagram of flexural-shear cantilever at time t under free vibration

52
∂M
M+ dx
∂x

∂V f
Vf + dx
∂x
wo ( x / H ) ndx

dx Vf

(a) Flexural element

∂Vs
Vs + dx
∂x
ndx
χ

Vs

(b) Shear element


Fig. 3.4 Free body diagram of flexural-shear cantilever under lateral loading

53
6.90
6.80
6.70
6.60
6.50
f (α ) i =1
6.40
6.30
6.20
6.10
6.00
5.90
5.80
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
α

Fig. 3.5 Index of structural interaction

0.85

0.80

0.75

0.70

RΔ /Rδm 0.65

0.60

0.55

0.50

0.45

0.40
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Fig. 3.6 Variation of α against drift ratio RΔ /Rδm

54
43.5 m

207.7 m
(a) Floor 1st to2nd

(b) Transfer (d) MDOF model


storey

(c) Floor 4th to 60th

Fig. 3.7 The floor plans of the building

55
Variation of EI along the height Variation of GA along the height

EI Equivalent EI GA Equivalent GA

207.9 207.9
198 198
188.1 188.1
178.2 178.2
168.3 168.3
158.4 158.4
148.5 148.5
138.6 138.6
128.7 128.7
Height (m)

118.8
Height (m)

118.8
108.9 108.9
99 99
89.1 89.1
79.2 79.2
69.3 69.3
59.4 59.4
49.5 49.5
39.6 39.6
29.7 29.7
19.8 19.8
9.9 9.9
0 0
0 50000 100000 0.00 10.00 20.00 30.00
2
EI (GNm ) GA (GN)

Fig. 3.8 The variation of stiffness and the equivalent stiffness of the building

56
0.4 0.4
0.3 0.3
Ground acceleration (g)

Ground acceleration (g)


0.2 0.2
0.1 0.1

0 0

-0.1 -0.1
-0.2 -0.2

-0.3 -0.3

-0.4 -0.4
0 5 10 15 20 25 0 5 10 15 20 25
Time (s) Time (s)

(a) (b)
Fig. 3.9 (a) From PEER with record ID: P0714 (b) From PEER with record ID:P0410

At 207.7m At 207.7m

Continuum MDOF Continuum MDOF


0.15 0.06
Displacement (m)

Displacement (m)
0.1 0.04
0.05 0.02
0 0
-0.05 -0.02
-0.1 -0.04
-0.15 -0.06
0 5 10 15 20 25 0 5 10 15 20 25
Time (s) Time (s)

(a) (b)
Fig. 3.10 Time history response at height of 207.7 m

At 129.7m At 129.7m

Continuum MDOF Continuum MDOF


0.08 0.04
Displacement (m)

Displacement (m)

0.06 0.03
0.04 0.02
0.02 0.01
0 0
-0.02 -0.01
-0.04 -0.02
-0.06 -0.03
-0.08 -0.04
0 5 10 15 20 25 0 5 10 15 20 25

Time (s) Time (s)

(a) (b)
Fig. 3.11 Time history response at height of 129.7 m

At transfer (17.5m) At transfer (17.5m)


Continuum MDOF Continuum MDOF
0.006 0.008
Displacement (m)

Displacement (m)

0.004 0.006
0.002 0.004
0.002
0 15.5
0
-0.002
-0.002
-0.004 -0.004
-0.006 -0.006
0 5 10 15 20 25 0 5 10 15 20 25

Time (s) Time (s)

(a) (b)
Fig. 3.12 Time history response at height of 17.5 m

57
Beam

9m
Column

9m

(a) 10-storey moment-resisting frame

Shear wall

2m 10 m

1m

10 m

(b) 20-storey shear walls

Column
Beam
Shear wall

5m 15 m

15 m

(c) 30-storey wall-frame


Fig. 3.13 Floor plans of example buildings

58
(a) 10-storey moment-resisting (b) 20-storey shear walls (c) 30-storey wall-frame
frame

Fig. 3.14 Finite element models

59
Chapter 4 

Performance­based seismic assessment of 
wall frame building structures using 
inelastic interstorey drift ratio 

4.1 Introduction

Cities in regions of low to moderate seismicity, especially with non-seismically designed


building structures, are subjected to seismic risk that cannot be neglected. Earthquakes bring
tremendous loss to the counties and therefore seismic assessment of building structures is
necessary for such regions. A comprehensive and cost effective seismic assessment scheme
includes different levels of assessment ranging from a preliminary to a detailed level depending
on the objectives.

60
Preliminary assessment is usually based the methods with engineering models of building
structures, which are suitable for the assessment of a larger number of buildings. To conduct a
large scale assessment in regions or countries, preliminary assessment is a compulsory step
performed by practising engineers to prioritise the assessment order of building structures for
the detailed assessment. The detailed assessment, though more reliable, is cost demanding and
time consuming in nature. Therefore the detailed assessment is only conducted on building
structures that fail the preliminary assessment.

A preliminary assessment for larger numbers of low-rise building structures has been
proposed by Lang (2002) in Switzerland. Assessed buildings can be simplified by obtaining
the strength and ductility through study of architectural plans. Capacity spectrum analysis is
then conducted on the simplified buildings to predict the seismic responses, which can be used
as an indicator to define the performance. Representative buildings are selected for
conducting the analysis and producing the fragility curve, which expresses the possible damage
grade under the expected spectral displacement. Building structures in general are sorted to
conform to one of the representative buildings and have similar fragility curves to the
representative building. It cuts short the detailed analysis for preliminary assessment.

Recently Sheikh (2005) established fragility curves of building structures based on the
maximum interstorey drift ratio, which is determined from a model composed of several drift
multipliers. A group of building structures is first selected for analysis by constructing their
corresponding full-scale models in the finite-element-based program “ETABS” with linear
elastic material properties. The fundamental periods of the models are correlated to the height
of the building and hence are used to determine the corresponding spectral displacements. The
drift multipliers are used to modify the spectral displacement, thus the influence of the
modelling and the higher modes are included. The drift multipliers are determined and
calibrated from the responses of the full-scale finite-element models.

Existing preliminary assessments are usually developed either for low-rise building
structures based on inelastic analysis, or for high-rise building structures based on elastic
analysis. It is well known that elastic analysis is less favourable than inelastic analysis for
high-rise building structures in earthquake engineering because under seismic excitation,
buildings are to behave nonlinearly and probably inelastically. This is the reason why
nonlinear analysis with well defined material properties is preferred for seismic assessment.

61
Nonlinear analysis is, however, cost and time demanding. It is not suitable for preliminary
assessment which has to be quick and simple in nature.

A quick and simple performance-based assessment that is based on the inelastic interstorey
drift model is proposed in this study for preliminary assessment of high-rise building structures.
The assessment first systemically defines building structures into two groups based on their
seismic behaviour, either elastically or inelastically, through a bifurcation index. For building
structures that deformed inelastically, a further assessment step is conducted based on the
model of the inelastic interstorey drift ratio. A model of the inelastic interstorey drift ratio is
derived from the principle of capacity spectrum analysis and the modified modal pushover
analysis (Chopra et al., 2004) on the assessed building structures, which is modelled by the
continuum flexural-shear cantilever. Depending on the interstorey drift ratio, the performance
of building structures is defined by the current codes of practice and existing literature.

4.2 Bifurcation index for elastic and inelastic behaviour

In the preliminary assessment, the performance and intensity of damage of the building
structure are usually correlated to the magnitude of deformation. Building structures are
guaranteed safe if they are elastic under ground motion. If inelastic deformation occurs in the
building structures during the excitation, permanent deformation and damage will result. If a
building structure is expected to deform inelastically, appropriate methods of assessment,
which considers inelastic behaviour, should be used.

To distinguish the seismic behaviour of building structures is the first step in preliminary
assessment for selection of the assessment method. An index, from the principle of the
capacity spectrum analysis, is defined as the ratio of the elastic spectra displacement of the 1st
mode Sd1 to the yield displacement Sdy. When it is larger than one, the assessed building
deforms inelastically. Otherwise, the assessed building deforms elastically under the given
ground motions and it is regarded as safe, with no significant damage ensuing.

62
4.2.1 Fundamental period
To estimate the elastic spectra displacement Sd1, the fundamental period of the assessed
building has to be known. A theoretical model for estimating the natural fundamental period
of vibration of building structures has been proposed in Chapter 3. The fundamental period is
derived from the continuous flexural-shear cantilever, which has various modes of deformation,
from flexure to shear, by changing the degree of structural interaction α of the continuous
cantilever. The period is given by
mH
T1 = f (α ) i =1 RΔ (4-1)
wo

where f(α)|i=1 is the index of structural interaction at the 1st mode (Fig. 3.5), m is the distributed
mass, H is the overall height, wo is the magnitude of the design inverted triangular lateral load,
and RΔ is the design roof drift ratio.

The degree of structural interaction α can be conveniently approximated from the


deformation profile of the building, which is characterised by RΔ/Rδm, a ratio of the design roof
drift ratio to the design maximum interstorey drift ratio of the building under an inverted
triangular lateral load wo. In the preliminary assessment, the roof drift ratio RΔ and the
maximum interstorey drift ratio Rδm may be assumed as H/500 and h/350 respectively, where h
is the storey height, which are consistent with those recommended in design codes of practice.
The value of α can be quickly determined using Fig. 3.6. With the given value of α, the index
of structural interaction f (α) is found in Fig. 3.5.

4.2.2 Elastic spectral displacement of 1st mode


Given the vibration period, the pseudo-acceleration, which is related to the displacement, can
be read from the demand spectrum. The demand spectrum beyond 1 second is usually
modelled as a function inversely proportional to the natural fundamental period of vibration,
given by

Sa = (4-2)
T
where Cζ is the proportionality constant corresponding to the viscous damping ζ %. With the
defined demand spectra, the elastic spectral displacement under the demand spectrum Eq. (4-2)
is obtained as

63
Celastic
Sd 1 = × T1 (4-3)
4π 2
where Celastic is the proportionality constant of the elastic viscous damping. By substituting Eq.
(4-1) into Eq. (4-3), the spectral displacement of the 1st mode at the corresponding period is
Celastic mH
Sd1 = × f (α ) i =1 RΔ (4-4)
4π 2 wo

4.2.3 Yield displacement, overstrength and strength of building structures


The actual strength of a building structure in practice is usually higher than the design strength
because of incorporating different factors of safety in the materials and load combinations,
neglecting non-structural elements and simplification of the lateral system, etc. This
characteristic of structures, where the actual strength is higher than the design strength, is
identified as overstrength. The portion of the overstrength is quantified by a multiplier, the
overstrength coefficient a. When a building structure is modelled as linear, it would yield at
(autop, aVbase) including the effect of overstrength a, where utop is the design roof displacement
under the design base shear Vbase.

From the SEAOC Blue Book (SEAOC, 2009), it is stated that the overstrength coefficient,
a, may vary from about 2.25 to 4.5. The result was obtained from a preliminary investigation
that had been conducted on various lateral systems, namely a bearing wall system, building
frame system and a moment resisting frame system. However, it also suggested that a
conservative value of overstrength coefficient should be either 2.0 or 2.5 depending on the
systems. For the preliminary assessment without performing a nonlinear static analysis, a
suggested value of overstrength coefficient could then be selected.

Building structures in general are designed against static lateral load and hence the design
strength can be approximated from the magnitude of the lateral load. Due to different lateral
patterns, it is not uncommon to consider a pattern that can result either the lower bound or the
upper bound of strength. The lower strength bound is obtained when the design lateral load
pattern is assumed to be an inverted triangular load with a maximum magnitude wo. The
design base shear of the building structure under this load pattern is
wo H
Vbase = (4-5)
2
When a nonlinear static analysis is conducted on the building structure, the elastic region of the

64
capacity curve can be modelled as linear, and yields at (autop, aVbase), including the effect of
overstrength a, where utop is the design roof displacement under the design base shear Vbase.
The capacity spectrum corresponding to the capacity curve gives the yield point expressed in
the format of spectral coordinate (Sdy , Say),
aRΔ H
Sdy = (4-6)
PF1
awo
S ay = 1
(4-7)
2mPF1 ∫ φ1 ( z )dz
0

where PF1 is the participation factor of the 1st mode, z (= x/H) is the dimensionless height, RΔ is
1
the roof drift (= utop/H) and ∫ φ (z )dz
0
1 is the integrand of the 1st mode with respected to z.

If the elastic spectral displacement is larger than the yield displacement, the assessed
building would deform inelastically and would result in damage and permanent deformation.
Therefore a bifurcation index (BI), defined as the ratio of the elastic spectra displacement to the
yield displacement, is established. The ratio of the elastic spectra displacement Sd1 to the yield
displacement Sdy is calculated from dividing Eq. (4-4) by Eq. (4-6) and is given by
S d 1 Celastic m
BI = = j (α ) i =1 (4-8)
S dy a wo RΔ H

PF1 f (α ) i =1
where j (α ) i =1 = ≈ 0.23 for α ∈ [0,20] When the value is larger than one, the
4π 2
building is damaged with unidentified intensity, which has to be defined by further assessment.

4.3 Model of inelastic interstorey drift ratio

To evaluate the performance and the intensity of damage of a building structure, interstorey
drift ratio, which has been recognised as an important instrument for performance-based
seismic assessment, is predicted. The model is based on capacity spectrum analysis and the
principle of modified modal pushover analysis and the continuum technique on the
flexural-shear cantilever model, which may vary its mode of deformation from flexure
dominant to shear dominant. The effects of the structural height, interaction of different
structural forms in the building, top and interstorey drifts, intensities of applied design load, and
the properties of the demand spectra are considered in the inelastic interstorey drift ratio model.
65
4.3.1 Demand spectra on building structures
The demand spectra, which are plots of pseudo-acceleration under natural vibration periods, are
a series of curves representing different equivalent viscous damping of the model. When the
spectra is expressed in the format of the spectral coordinate, Eq. (4-2) becomes
Cζ2
Sa = (4-9)
4π 2 S d

With increasing equivalent damping ζ %, the magnitude of the demand decreases and so
does the proportionality constant Cζ. The relationship between equivalent damping and the
proportionality constant Cζ is modelled as
− d 2Cζ
ζ = d1e (4-10)
where d1 and d2 are the modelling parameters in the regression analysis, dependent on the given
set of ground motions.

4.3.2 Performance point in capacity spectrum analysis


The performance point (Sdo , Sao), which is the predicted inelastic displacement and the
pseudo-acceleration of the corresponding model, is the intersection of the demand spectrum
and the capacity spectrum. It is obtained by solving Eqs (4-2) and (4-7), and is given by

Co2 mPF1 ∫ φ ( z )dz


1

S do = 0
(4-11)
2π 2 aw
Sao = S ay (4-12)

where Co is the proportionality constant of the demand spectrum at the performance point. It is
determined when the equivalent damping of the demand spectrum coincides with the equivalent
damping of the model.

The equivalent damping of the model is estimated from the equation proposed by Gulkan
and Sozen (1974). It was developed from the results of shake table experiments of small-scale
reinforced concrete frames. The equation has been compared to other empirical equations by
Miranda (2000) and it is showed that the equation gave estimations close to the others,
⎛ 1 ⎞⎟
ζ = ζ elastic + 0.2⎜1 − (4-13)
⎜ μ ⎟⎠

66
where ζ and ζelastic are the equivalent damping of the model at a particular μ , which is the
ductility at the performance point, and the elastic damping at the beginning, respectively.

To obtain the proportionality constant Co at the performance point, it is required to


simultaneously solve Eq. (4-6) of the yield coordinate, Eq. (4-9) of the demand spectrum,
Eq. (4-10) of the proportionality constant and equivalent damping, and Eq. (4-13) of the
equivalent damping of the model, which gives the following characteristic equation of the
proportionality constant Co at the performance point,
2
1
⎛ 0.2 ⎞
PF ∫ φ1 (z )dz ⋅ C − 2π Bdg ⎜⎜
2 2 2
⎟⎟ = 0 (4-14)
⎝ 0.2 + ζ elastic − πd1e
1 o − d 2Co
0 ⎠
where Bdg is the index of the structural properties.
a 2 RΔ Hwo
Bdg = (4-15)
m

4.3.3 Maximum seismic interstorey drift ratio


When the performance point is determined, the inelastic interstorey drift ratio can be estimated
from modified modal pushover analysis (MMPA) (Chopra et al., 2004). In the analysis, the
response contribution of the 1st mode is nonlinear inelastic while the contributions of higher
vibration modes are linearly elastic. The inelastic interstorey drift ratio is determined from the
rule of modal combination of the first three modes, which are enough for convergence, and
hence is given by
1
IDR ≈ (PF1φ1′(z )Sdo )2 + (PF2φ2′ (z )Sd 2 )2 + (PF3φ3′ (z )Sd 3 )2 (4-16)
H
where Sd2 and Sd3 are the elastic spectral displacements of the 2nd and the 3rd modes respectively;
z (= x/H) is the dimensionless height.

To simplify Eq. (4-16), the 1st mode spectral displacement is extracted as follows
Sd 1
IDRmax ≈ (PF1φ1′(z )Sr )2 + (PF2φ2′ (z )r2 )2 + (PF3φ3′(z )r3 )2 (4-17)
H
where Sr is the inelastic displacement ratio, r2 and r3 are higher mode elastic displacement
ratios. The inelastic displacement ratio Sr is defined as the ratio of the performance point to
the elastic 1st mode spectral displacement,
S do Co2 1 m
Sr = = SR (α ) (4-18)
S d 1 Celastic a RΔ Hwo

67
1
2 PF1 ∫ φ ( z )dz
where SR (α ) = 0
(4-19)
f (α ) i =1

which is plotted in Fig. 4.1. The displacement ratio ri is the ratio of elastic spectral
displacements between the higher modes and the 1st mode, which is given by

ri =
S di
= 12 12
(
γ 2 γ 2 +α2 ) (4-20)
Sd1 (
γi γi +α 2 )

By substituting Eq. (4-4) of the elastic spectral displacement of the 1st mode and zo, where
zo is the dimensionless height at which the maximum interstorey drift ratio occurs, into Eq. (17)
of the interstorey drift ratio, the maximum value is expressed as
mRΔ
IDRmax = Celastic IR(Sr ,α ) (4-21)
wo H

where IR(Sr,α) is an index of inelastic displacement ratio and structural interaction


f (α ) i =1
IR(Sr ,α ) = (PF1φ1′(zo )Sr )2 + (PF2φ2′ (zo )r2 )2 + (PF3φ3′ (zo )r3 )2 (4-22)
4π 2

For preliminary assessment, the proportionality constant Co is determined from Eq. (4-14) for
calculating the inelastic displacement ratio Sr in Eq. (4-18). With the inelastic displacement
ratio Sr, the index of inelastic displacement ratio IR(Sr, α) at particular structural interaction α
is found from Fig. 4.2. Hence, the maximum interstorey drift ratio IDRmax can be conveniently
determined from Eq. (4-21).

4.3.4 Classification of performance


Current design codes of practice and the literature have suggested the interstorey drift ratio as
an instrument to measure damage and define the performance of assessed buildings, for
example the deformation limits suggested in ATC 40 (see Table 4.1). In the current literature,
there are experimental studies concerning the deformation limits. Priestley and Kowalsky
(1998) suggested the curvature limits for wall structures through experiments on different
reinforcement details. Li (2006) carried out a shake table test of a 1:20 scale high rise building
and presented the relationship between the structural damage and the interstorey drift ratio
based on the test observations. The relevant tools for defining the performance of building
structures are given in preliminary assessment in Table 4.2.

68
A performance-based preliminary seismic assessment serving the purpose of screening a
larger number of building structures is shown in Fig. 4.3. Building structures under the
expected ground motion intensity are first divided into two groups by the bifurcation index in
which building structures behave either elastically or inelastically. For building structures that
deform inelastically, meaning the occurrence of permanent deformation, a second step in
assessment is conducted based on the model of the inelastic interstorey drift ratio.
Performance of the building structure is then suggested from the deformation limits stated in
codes of practice or the literature.

4.4 Numerical investigation

To evaluate a larger number of building structures, building structures would undergo the
suggested procedure and is divided into two groups by the bifurcation ratio based on their
behaviour. Then a more detailed classification of performance is conducted based on the
prediction of the inelastic interstorey drift ratio by comparing to the deformation limits in codes
of practice and the literature. Three representative buildings are investigated by the suggested
procedure.

4.4.1 Example buildings


Finite element analysis using ETABS (Computer and Structures Inc. 2005) is carried out for the
design of three example building structures. They are a 30-storey wall-frame structure with
floor area of 24 × 30 m2, a 25-storey wall-frame structure with floor area of 25 × 25 m2 and a
15-storey moment resisting frame with floor area of 18 × 18 m2. The structural plans are
shown in Fig. 4.4 and the dimensions of the structural members are listed in Table 4.3. The
interstorey height of all the representative buildings is 3.5m, and the overall heights of the three
buildings are 105 m, 87.5 m and 52 m, respectively.

Non-seismic design is considered for the three buildings, which are designed to resist a
dead load (D) of 7.0 kN/m2, a live load (L) of 3.0 kN/m2, and lateral wind loads (W) of 2.88 kPa,
2.80 kPa and 2.58 kPa at the roof level respectively, which decrease gradually to 0 at the base
level. Three load combinations are considered in the design: 1.4D + 1.6L, 1.2D +1.2L + 1.2W,
and 1.0D + 1.0L + 1.0W. All the buildings satisfy the roof drift limit of H/500 in the load
combination: 1.0D + 1.0L + 1.0W. The roof drift under this load combination are 0.20 m, 0.14

69
m and 0.07 m. By the drift ratio RΔ/Rδm in Fig. 3.6, the degrees of structural interaction α of the
buildings are approximately equal to 1.6, 3 and 9.4.

The three buildings are then modelled in the nonlinear finite-element-based program
“OpenSees” for response history analysis. Because of the symmetry, the buildings are
converted into two-dimensional models shown in Fig. 4.5, which have similar fundamental
periods and elastic deflections under lateral load in the corresponding buildings, in order to
reduce computational effort. The resultant masses at the floor level of the models are 242 Mg,
274 Mg and 53.32 Mg respectively. The properties of the concrete and reinforcement are
shown in Fig. 4.6 with yield strengths o approximately at 28 MPa and 413 MPa. The
hysteretic behaviour is selected as isotropic hardening.

To study the seismic behaviour of the buildings, the ground motions are selected and given
in Table 4.4. The peak ground acceleration of the ground motion is artificially increased by
amplification scalars from 1 to 4 and it results in four groups of ground motion. The maximum
seismic responses, including the storey drift and the interstorey drift ratio under the four groups
of ground motion, are recorded and the maximum responses at each storey together produce the
seismic response envelope. Fig 4.7 and Fig 4.8 show the displacement envelopes and
interstorey drift ratio envelopes with the static response under the load combination 1.0D +
1.0L + 1.0W, respectively. The ductility demand which is defined as the ratio of the maximum
seismic response to the maximum static response in the loading combination 1.0D + 1.0L +
1.0W are plotted in Fig. 4.9 and Fig. 4.10.

The deflections and interstorey drift ratios in Fig. 4.7 and Fig. 4.8 have demonstrated the
deflection characteristics of the buildings with different degrees of structural interaction. The
three structures behave differently: the 30-storey and 25-storey wall-frame deflect flexurally,
while the 15-storey frame deflects in shear. With increasing peak ground acceleration, the
responses increase. The increments of the ductility demands shown in Fig. 4.9 and Fig. 4.10
are larger than those in the peak ground acceleration. It implies a nonlinear deflection
characteristic of the building structure under ground motion.

70
4.4.2 Performance-based assessment

Step 1 Demand spectrum

To facilitate the preliminary seismic assessment in a region, the demand spectrum is


standardised either from a group of ground motion records or from codes of practice in similar
regions. In this numerical investigation, six ground motion records with similar properties
were selected from the “PEER strong Motion Database” listed in Table 4.3 (# from 1 to 6) and
are amplified 3 times to guarantee the occurrence of inelastic deflection. The
pseudo-acceleration spectra at 5% viscous damping are plotted in Fig. 4.11. Linear regression
is carried to model the spectra as a function inversely proportional to the fundamental period
from 1s to 5s, which gives C5% = 7.420. Assuming the elastic inherent viscous damping of the
building structures is 3%, which generates a proportionality constant of C3% = 9.63 by Eq.
(4-10), the relationship between equivalent damping and the proportionality constant Cζ is
modelled by Eq. (4-14) and results in d1 = 1.505 and d2 = 0.406 with root mean square value of
0.99 from Fig. 4.12.

Step2 Bifurcation index for elastic and inelastic behaviour

The applied lateral force on the 30-storey wall frame is 2.88μ15=43.20 kN/m and the
distributed mass is 242/3.5=69.14 Mg/m. The lateral force on the 25-storey wall frame is
2.79μ12.5=34.88 kN/m and a distributed mass is 274/3.5=78.29 Mg/m. The 15-storey frame
is applied a lateral force 2.58μ9=23.22 kN/m and distributed mass of 53.32/3.5=15.23Mg/m.
Nonlinear static analysis is conducted on the models and the overstrength coefficients are
approximately equal to 3, 4 and 4 respectively. The bifurcation index of the 30-storey building
is calculated using Eq. (4-8)
9.63 242 / 3.5
BI1 = × 0.23
3 (2.88 × 15) × 0.0019 × (3.5 × 30)
= 2.09
Similarly the bifurcation index of the 25-storey building is calculated and equal to BI2 =
2.22. The 15-storey frame has bifurcation index of BI3 = 1.70. Since the bifurcation indices
of all buildings are larger than 1, they deform inelastically under the expected ground motion
and need further assessment.

71
Step 3 Inelastic interstorey drift ratio

The further assessment is conducted by predicting the inelastic interstorey drift ratios. The
corresponding indices of structural properties in Eq. (4-15) are firstly calculated. The index of
structural properties of the 30-storey wall frame building is
32 × 0.0019 × 105 × (2.88 × 15)
Bdg = = 1.12
(242 / 3.5)
From Fig. 4.12, when α = 1.6 and Bdg = 1.12, the proportionality constant Co = 6.4. From Fig.
4.2, SR(1.6) = 0.21. Hence, the inelastic ratio is given by Eq. (4-18),
6 .4 2 1 69.14
Sr = × × 0.21
9.63 3 0.0019 × 105 × 43.20
= 0.84
Then the index of inelastic displacement ratio and structural interaction given in Fig. 4.13 is
IR(0.84, 1.6) = 0.78. The maximum interstorey drift ratio determined by Eq. (4-22) is
69.14 × 0.0019
IDRmax = 9.63 × × 0.28
43.20 × 105
= 0.014
Similarly, the index of the structural properties of the 25-storey wall frame building is Bdg =
1.00 and the proportionality constant is Co = 6.5 found from Fig. 4.11. From Fig. 4.12, SR(3) =
0.216 and the inelastic ratio in Eq. (4-18) is Sr = 0.95. Therefore the index of the inelastic
displacement ratio and the structural interaction given in Fig. 4.2 is IR(0.95, 3) = 0.30, which
results the maximum interstorey drift ratio IDRmax = 0.014.

The index of the structural properties of the 15-storey frame building is Bdg = 0.43 and the
proportionality constant is Co = 6.8, found in Fig. 4.11. From Fig. 4.12, SR(9.4) = 0.223 and
the inelastic ratio in Eq. (32) is Sr = 0.82. Therefore the index of the inelastic displacement
ratio and the structural interaction given in Fig. 4.2 is IR(0.82, 9.4) = 0.28, which results the
maximum interstorey drift ratio IDRmax = 0.011.

72
Step 4 Performance

Comparisons between the maximum interstorey drift ratio predicted by the proposed inelastic
drift model and by the response history analysis (RHA) are made and presented in Figs. 4.13
and 4.14. It can be seen from the Fig. 4.13 that the proposed formula, which is presented by
the dotted line, gives slightly conservative predictions for the first building. This may be due
to the use of a smoothed demand spectrum, which is commonly used in preliminary assessment.
In Fig. 4.14, the proposed formula agrees through the results obtained from response history
analysis and also tends to give higher predictions in the second building. Similar observations
are also seen in the results of the 15-storey frame shown in Fig. 4.15. In general, the proposed
formula successfully predicts an expected conservative interstorey drift ratio of the assessed
building under the given set of ground motions.

The maximum interstorey drift ratios of the assessed buildings are compared in Table 4.1.
It can be seen that both structures belong to “Damage Control” from the suggestion of ATC 40.
A more detailed description is given in Table 4.2. Although the assessed buildings are safe,
structural components such as shear walls, are damaged with the concrete crushed and the
reinforcement exposed. If “Damage Control” is the minimum requirement for building
structures to pass the preliminary assessment, the assessed building is regarded as safe, based
on this performance, and the need for further detailed assessment is low.

73
4.5 Concluding remarks

A quick and simple performance-based assessment that is based on the inelastic interstorey drift
model has been proposed in this chapter for the preliminary seismic assessment of high-rise
building structures. The preliminary assessment first divides the assessed building structures
into either elastic or inelastic categories under the expected ground motion using the bifurcation
index, which is derived from the elastic spectrum displacement to yielded spectrum
displacement in capacity spectrum analysis. Building structures that deform inelastically are
further examined by the model of inelastic interstorey drift ratio.

The model of inelastic interstorey drift ratio is derived from the principle of capacity
spectrum analysis and the modified modal pushover analysis on the assessed building structures,
which is modelled by a combined continuous flexural-shear cantilever that can describe a wide
range of modes of deformation and hence capture the deflection characteristics more closely for
multi-storey buildings. The model considers the effects of the structural height, interaction of
different structural forms in the building, top and interstorey drifts, intensities of the applied
design load, and the demand spectra properties. Under the defined demand spectrum, the
inelastic interstorey drift ratio can be quickly calculated through a series of charts.
Performance of building structure is then defined based on the interstorey drift ratio by
comparison to the current codes of practice and the literature.

Worked examples pertaining to determining the performance of three wall-frame


structures with 30-storey, 25-storey and 15-storey, designed properly against gravity and lateral
load, show that the proposed interstorey drift ratio model successfully predicts the expected
responses of the buildings obtained from response history analysis of a given set of ground
motions, and the assessment procedure for defining the building performance is simple and
systematic. The proposed assessment procedure, which serves as a screening purpose, will be
an effective, yet reasonably reliable, means for the evaluation of a large number of existing
buildings and can identify potentially damaged structures for the further detailed assessment.

74
Table 4.1 Deformation Limits suggested in ATC-40
Performance level
Immediate Damage Structural
Interstorey Drift Limit Life Safety
Occupancy Control Stability
Maximum total drift a 0.01 0.01-0.02 0.02 0.33Vi/Pi c
b
Maximum inelastic drift 0.005 0.005-0.015 No limit No limit
a
Maximum total drift is defined as the performance point displacement
b
Maximum inelastic drift is defined as the portion of the maximum total drift beyond the effective yield point.
c
Vi is the total calculated lateral shear force in storey i and Pi is the total gravity load.

Table 4.2 Relationship between Structural Damage and Storey Drift (Li et al., 2006)
Description of structural damage Storey drift
Small cracks on columns in frames 1/1,000–1/1,300
A few number of small cracks on shear walls 1/1,100–1/1,200
Many through-cracks on shear walls 1/300–1/700
Shear walls damaged with concrete crushed and reinforcement exposed 1/80–1/200

Table 4.3 Dimensions of the members in the representative buildings


30-storey wall-frame 25-storey wall frame 15-storey frame
2 2
Column 900 × 900 mm 900 × 900 mm 900 × 900 mm2
Beam 350 × 700 mm2 400 × 700 mm2 400 × 700 mm2
Shea wall 500 mm 500 mm --

Table 4.4 Ground motion records


Magnitude PGA PGV PGD
# Earthquake Record ID Date
(Ms) (g) (cm/s) (cm)
1 Chi-Chi, Taiwan P1121 7.6 1999/09/20 0.120 25.1 13.84
2 Chi-Chi, Taiwan P1185 7.6 1999/09/20 0.102 20.9 13.42
3 Chi-Chi, Taiwan P1200 7.6 1999/09/20 0.102 22.0 13.76
4 Chi-Chi, Taiwan P1202 7.6 1999/09/20 0.118 22.0 11.63
5 Chi-Chi, Taiwan P1454 7.6 1999/09/20 0.111 23.6 13.27
6 Chi-Chi, Taiwan P1491 7.6 1999/09/20 0.136 22.7 12.58

75
0.230

0.225

0.220
SR (α )

0.215

0.210

0.205

0.200
0 2 4 6 8 10 12 14 16 18 20

α
Fig. 4.1 Graph of SR(α)

1
Sr = 2.5

0.8 Sr = 2.0
IR(Sr, α)

Sr = 1.5
0.6

Sr = 1.0
0.4

0.2
Sr = 0.2
0
0 2 4 6 8 10 12 14 16 18 20

α
Fig. 4.2 Graphs of IR(Sr, α)

76
A large group Bifurcation index
of buildings Safe
<1
>1

Method of predicting interstorey drift ratio

Gathering building
information

Substitution

Explicit equation of interstorey drift ratio

Calculation

Interstorey drift ratio

Comparison to the requirement of interstorey drift


ratio based on current code of practice or the literature.
e.g. ATC 40
Li et al. (2006)

Classification of performance

Safe Minor risk Dangerous

Fig. 4.3 Preliminary seismic assessment procedure

77
30m
Death load: 7kN/m2
Live load: 3kN/m2 10m 10m

6m

24m
12m

α =1.6
25m

10m 5m

25m
5m

5m

α =3.0
18m
6m 6m
18m
6m

α =9.4

Fig. 4.4 Floor plans of the three buildings

78
2.89kPa
Floor 30 Storey height = 3.5m
Floor 29
Floor 28

2.80kPa

….....
Floor 25
Floor 24
Floor 23

….....
….................

2.58kPa Floor 15
Floor 14
………. Floor 13

….....

α =1.6 α =3.0 α =9.4
Fig. 4.5 Models in “OpenSees”

Stress MPa Stress MPa

12.5 GPa 2.8


0.23 GPa
413 4 GPa
Strain
200 GPa
(-0.0022, 23.0)
Strain

(-0.0035, 27.6)
25 GPa
Concrete Steel
Fig. 4.6 Material properties

79
30 25

1 2 3 4 1 2 3 4
25
20

20
15

Story
Story

15

10
10

5
5

0 0
0 0.5 1 1.5 0 0.5 1 1.5
Displacement (m) Displacement (m)

(a) (b)
30-storey wall frame building 25-storey wall frame building

15
1 2 3 4 Legend

-x- Static deflection


10
1. Group 1
2. Group 2
3. Group 3
Story

4. Group 4
5

0
0 0.5 1 1.5
Displacement (m)

(c)
15-storey frame

Fig. 4.7 Displacement envelopes under four groups of ground motion

80
30 25

1 2 3 4 1 2 3 4
25
20

20
15

Story
Story

15

10
10

5
5

0 0
0 0.01 0.02 0 0.005 0.01 0.015 0.02
Interstory drift ratio Interstory drift ratio

(a) (b)
30-storey wall frame building 25-storey wall frame building

15

Legend

-x- Static deflection


10 1. Group 1
2. Group 2
3. Group 3
Story

4. Group 4
1 2 3 4
5

0
0 0.01 0.02 0.03 0.04
Interstory drift ratio

(c)
15-storey frame

Fig. 4.8 Interstorey drift ratio envelopes of four groups of ground motion

81
1 1

0.9 1 2 3 4 0.9 1 2 3 4
0.8 0.8

0.7 0.7
Relative Height

Relative Height
0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 3 4 5 6 0 2 4 6 8
Displacement Ductility Displacement Ductility

(a) (b)
30-storey wall frame building 25-storey wall frame building

0.9 Legend
0.8
1 2 3 4 -x- Static deflection
0.7
1. Group 1
Relative Height

0.6 2. Group 2
0.5 3. Group 3
4. Group 4
0.4

0.3

0.2

0.1

0
0 2 4 6 8 10 12
Displacement Ductility

(c)
15-storey frame

Fig. 4.9 Displacement ductility of four groups of ground motions

82
1 1

0.9 0.9

0.8 1 2 3 4 0.8

0.7 0.7

Relative Height
Relative Height

0.6 0.6 1 2 3 4
0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 3 4 5 6 0 2 4 6 8
Drift Ductility Drift Ductility

(a) (b)
30-storey wall frame building 25-storey wall frame building

1
Legend
0.9

0.8 -x- Static deflection


0.7 1. Group 1
2. Group 2
Relative Height

0.6
3. Group 3
0.5 4. Group 4
1 2 3 4
0.4

0.3

0.2

0.1

0
0 2 4 6 8 10 12
Drift Ductility

(c)
15-storey frame

Fig. 4.10 Drift ductility of four groups of ground motion

83
12

10

8
Sa (m/s2)

7 .420
6 Sa =
T

0
0 1 2 3 4 5
Period T (s)

Fig. 4.11 Pseudo acceleration spectra at 5% viscous damping

(For d1 = 1.505, d2 = 0.406, ζelastic = 0.03)


20
6.8
7

6.4
6.6
6.9

5.9

18
6.1
6.5
6.7

6
6.3

6.2

16

14

12
6.8
7

6.4
α

6.6
6.9

5.9

10
6.1
6.5
6.7

6
6.3

6.2

4
6.8
7

9
2 1 5.
6.4
6.6

6.
6.9

6
6.7

6.5

2
6.

6.

0 0.5 1 1.5 2 2.5 3


Bdg
Fig. 4.12 Contour of Cο

84
RHA Proposed model

0.018

0.016

0.014
Interstory drift ratio

0.012

0.010

0.008

0.006

0.004

0.002

0.000
0 1 2 3 4 5 6 7

Ground motion ID

Fig. 4.13 Comparison of maximum interstorey drift ratio by proposed formula and
RHA of the first example building

RHA Proposed model

0.018

0.016

0.014
Interstory drift ratio

0.012

0.010

0.008

0.006

0.004

0.002

0.000
0 1 2 3 4 5 6 7

Ground motion ID

Fig.4.14 Comparison of maximum interstorey drift ratio by proposed formula and


RHA of the second example building

85
RHA Proposed model

0.018

0.016

0.014
Interstory drift ratio

0.012

0.010

0.008

0.006

0.004

0.002

0.000
0 1 2 3 4 5 6 7

Ground motion ID

Fig.4.15 Comparison of maximum interstorey drift ratio by proposed formula and


RHA of the third example building

86
Chapter 5 

Modified Continuum­MDOF Model for 
Seismic Analysis of Building Structures 

5.1 Introduction

The linear continuum flexural-shear cantilever model has been widely used in seismic
engineering, as discussed in Chapter 2. An example is the model used in the development of
the generalised interstorey drift spectra (Miranda and Akkar, 2006). The drift spectrum has
provided an alternative way to the traditional response spectra, and an additional tool in
assessment, for engineers to predict the seismic response of a building structure in a fast and
simple manner. However, the drift predicted from a linear model is limited to the linear elastic
response range. Although increasing the equivalent viscous damping of the model can be used
to predict the inelastic responses, it is less favourable than response history analysis of a
nonlinear model with hysteretic behaviour. A nonlinear model is preferred for predicting
inelastic response.
87
A nonlinear simplified MDOF model for the estimating of the seismic response of regular
wall-frame buildings was developed by Huang (2009). The model is a modification of the
continuum flexural-shear cantilever model, in which the characteristic of the continuum model
is implemented into a finite-element-method-based computer programme to produce a
continuum-MDOF model. Compared to a detailed model with member-by-member
representation, the degree of freedom of the simplified model is less and hence the
computational demand for the analysis is smaller. Although the use of the continuum-MDOF
model is proposed for individual detailed assessment, it has a great potential in the
development of other types of assessment, especially for preliminary seismic assessment of
large number of buildings.

In preliminary seismic assessment, the time involved is very limited. At the same time,
the effort for analysis has to be kept at minimal as possible; thus the requirement of the
modelling would not be as strict as the modelling for the detailed assessment, for instance
referred to the noniterative equivalent linear method proposed by Lin and Miranda (2008).
Similarly, a continuum-MDOF model, which has several more advantages than the SDOF
model, can also be adopted for the preliminary seismic assessment of existing building
structures after modification. In this research, a modified continuum-MDOF model, which is
an improvement of the continuum-MDOF model for the preliminary seismic assessment, is
developed. Numerical investigations are carried out for three building structures. It is
shown that the modified model gives reasonably good predictions of the interstorey drift
ratios under response history analysis as compared to the corresponding detailed model with
member-by-member representation.

A model is said to be similar to another one if it satisfies some defined relationships.


Two similar continuum models are investigated and their responses are found to be related by
the height ratio for the same strength ratio in linear static behaviour, linear dynamic behaviour
and equivalent seismic behaviour under modified modal pushover analysis. The relationship
that describes two similar models using the defined conditions of model similarity is named as
the general continuum representation. By analysing a modified continuum-MDOF model
corresponding to a typical building structure, the behaviour of other similar building
structures can be approximated.

88
5.2 Modified continuum-MDOF model

Huang (2009) developed a model in which a continuum flexural-shear cantilever is adjusted in


a finite-element-based programme as a continuum-MDOF model. The purpose of the
development is to replace a full-scale detailed MDOF model by a simplified model, so that a
time efficient and feasible method of analysis can be carried out. Numerical investigation
(Kuang and Huang, 2011) showed that the results from the continuum-MDOF model agreed
well to the full-scale model of a regular wall-frame building. The continuum-MDOF model
consists of a flexural cantilever and a shear cantilever. The properties of the model along the
height are discrete in nature. Each cantilever has one single node with lumped mass at each
level of floor corresponding to the building structure. In the continuum model, the two
cantilevers are connected by an axially -rigid membrane along the height, while in the MDOF
model, the membrane between the two cantilevers is represented by horizontal displacement
constraints at the nodes, so that the two cantilevers would deform identically.

Wall elements in the assessed building are combined to produce the flexural cantilever
based on the area of the wall elements. The hysteretic behaviour of the flexural cantilever is
then presented from the total steel ratio of the wall elements and the arrangement of the
reinforcement in the walls. The lumped masses at the nodes are obtained from the tributary
area of the floor. An initial value of shear rigidity is assumed for the shear cantilever so that
the initial fundamental periods of the continuum-MDOF mode can be calculated. Through
an iteration process, the most appropriate shear rigidity can be obtained by minimising the
difference between the period of the detailed model and the MDOF model. Lastly, an
appropriate hysteretic behaviour of the shear beam is selected.

It can be seen that the establishment of the continuum-MDOF model requires a detailed
MDOF model to be first developed or given. It greatly depends on the detailed MDOF model
or the structural plan of the assessed building. If neither piece of information corresponding
to the building is given, it would be difficult to establish the continuum-MDOF model.
Therefore it is only applicable in the detailed assessment in which the information has been
provided. For the preliminary assessment, an alternative approach has to be used; this is
because the corresponding detailed model including information of reinforcement details is
usually not available. A new approach is hence proposed for the preliminary seismic

89
assessment. It tries to approximate the behaviour of the assessed building structure based on
accessible information such as the design roof drift and maximum interstorey drift ratios,
intensities of applied loading and the height of the structure, etc.

The new approach in the modelling constructs a modified continuum-MDOF model


which is presented in the following sections, including a section on the flexural cantilever and
the shear cantilever. The stiffness and strength corresponding to the cantilevers are
determined from the continuum technique based on the applied force and deformation
relationship such as roof drift under lateral wind loading. Thus the information of the
member sizes and reinforcement detailing is not required. The differences between the
cantilevers of the continuum model and the modified continuum-MDOF model are presented
by figures and derivations. The modified continuum-MDOF model is implemented into the
finite-element based programme “OpenSees”, but the same modelling approach can be
applied into other programmes.

5.2.1 Section of flexural cantilever


A flexural cantilever bends under external lateral loading and its cross section remains plane.
The strain diagram of the section thus varies linearly from the extreme tensile fibre to the
extreme compressive fibre. The deflection characteristic of the flexural cantilever is governed
by the flexural rigidity EI, which determines the moment of resistance of the section. The
resistant moment of a rectangular section is formulated in terms of the extreme fibre strain and
the area of the section as shown in (Fig 5.1). The moment of resistance is contributed by a
couple with magnitude F and moment arm R.
M = F×R (5-1)

The magnitude of the couple can be estimated from either the resultant force of the
compressive area or the resultant force of the tensile area, as they are identical. Consider the
resultant compressive force in the section,
Eεbd
F= (5-2)
4
where E is Young’s modulus of the material, ε is the extreme fibre strain, b is the width of the
section and d is the depth of the section. The moment arm of the couple is taken as the
distance between the centroid of the compressive area and that of the tensile area,

90
2d
R= (5-3)
3
Substituting Eq. (5-2) and Eq. (5.3) into Eq. (5.1) and simplifying gives
M = EI × φ (5-4)
where I is the moment of inertia and φ is the curvature, which are given by

bd 3
I= (5-5)
12

φ= (5-6)
d

To model the nonlinearity of the flexural cantilever, which represents a structural system, a
nonlinear moment and curvature relationship has to be defined in Eq. (5-4). However, the
flexural rigidity in Eq. (5-4) only concerns the properties of the material and the cantilever
section. To represent a structural system which behaves flexurally, a corresponding flexural
rigidity should be defined. Therefore when a flexural cantilever is modelled as a collective of
flexural systems in an MDOF model, a new flexural cantilever section is adopted.

The new section of the flexural cantilever allows easy nonlinear modelling of a flexural
system with known moment of resistance. It consists of two fibres with total area 2A, where A
is the area of the fibre as shown in Fig. 5.2(b). The centroid of two circular fibres is separated
a distance r and the centroid of the section is axially rigid, which does not allow the section to
have a net axial shortening. The expression for the moment of resistance of the new section is
different to the rectangular section defined in Eq. (5-4). The moment of resistance of the
section is given by
m = f ×r (5-7)
where f is the magnitude of force in either the compressive fibre or the tensile fibre. The
magnitude of force depends on the following
f = EεA (5-8)
where ε is the fibre strain. Substituting Eq. (5-8) into Eq. (5-7), the moment of resistance
becomes
m = EI s × φ (5-9)

where Is is the moment of inertia of the simplified section and φ is the curvature, which are
given by

91
Ar 2
Is = (5-10)
2

and φ = (5-11)
r

By using the new section of the flexural cantilever with a known moment of resistance, the
nonlinear flexural hysteretic behaviour of corresponding system can be modelled by the axial
stress-strain relationship of the fibre defined in Eq. (5-8).

5.2.2 Section of shear cantilever


A shear cantilever deforms in shear under an external lateral load and the cross sections slide
relatively to each other horizontally. The deflection characteristic of a shear cantilever is
governed by the shear rigidity GA, which determines the shear resistance of the section. The
shear resistance of the section is formulated in terms of the shear strain and the shear rigidity,
V = GA × γ (5-12)
where γ is the shear strain. In the MDOF model, a shear cantilever is built by a section
which is very stiff in flexure and deforms in shear, and follows the defined shear-shear strain
relationship. A nonlinear hysteretic behaviour of shear-shear strain can be applied to Eq.
(5-12), so that a nonlinear shear cantilever can be formed to represent the shear system in the
building structure.

5.3.3 Modelling of Building Structures into Modified Continuum-MDOF Model


When a building structure is simplified into a modified continuum-MDOF model, the
deformation characteristics, stiffness and strength of the building structure have to be input into
the model. The properties of the flexural and shear cantilevers are then characterised by the
yield moment, the yield strength and the rigidity. The required information is predicted from
the following proposed approach based on continuum flexural-shear cantilever.

The yield moment can be approximated from the magnitude and distribution of the
external force and the degree of structural interaction. For instance, when the lateral load is a
uniformly distributed load with magnitude w, the yield moment at the flexural cantilever and
the shear force along the height of the shear cantilever, which are derived and presented in
Appendix A2, are given by,

92
wH 2 ⎡ α sinh α + 1 ⎤
My = − 1⎥ (5-13)
α 2 ⎢⎣ cosh α ⎦
wH ⎡α sinh α + 1 αx αx αx ⎤
VS ( x ) = ⎢ sinh − α cosh + α − ⎥ (5-14)
α ⎣ cosh α H H H⎦
The maximum shear occurs at a particular model height which can be found conveniently in
spreadsheet analysis.

The flexural rigidity of the flexural cantilever can be approximated by,


wH 3 ⎡ α 2 α sinh α + 1⎤
EI = ⎢1 + − ⎥ (5-15)
RΔα 4 ⎣ 2 cosh α ⎦

where RΔ is the roof drift ratio under uniform load w. The shear rigidity is then calculated from
the degree of structural interaction, which has been discussed in Chapter 3.
2
⎛α ⎞
GA = ⎜ ⎟ EI (5-16)
⎝H ⎠
The internal forces under an inverted triangular distributed load and point load are shown in the
Appendix A3.

Once the required information is determined, it is implemented into the modified


continuum-MDOF model. The flexural cantilever section is defined by the following
procedure.
1. Select the area and the distance of separation of the fibre.
2. Calculate the moment of inertia of the section by Eq. (5-10).
3. Calculate the yield stress of the fibre by Eq. (5-7).
4. Calculate the Young’s modulus of the fibre by Eq. (5-10) and Eq. (5-15).
5. Calculate the yield strain of the fibre by Eq. (5-8).
The shear cantilever is then directly modelled by Eq. (5-12) with the calculated shear stress
from Eq. (5-14) and shear rigidity from Eq. (5-16).

5.4.4 Worked example


A 25-storey building with the degree of structural interaction of α = 3 and the overall height
of H = 87.5 m, with interstorey heights of 3.5 m, is subjected to a uniform external lateral load

of 70 kN/m. It deforms with roof displacement = 1 × 87.5 = 0.175 m. Assuming the


500
building yields at this magnitude of external force, the maximum moment in the flexural

93
cantilever, the maximum shear in the shear cantilever, and the stiffness of the model are
calculated as,
M y = 124129 kNm
Vmax (37 m ) = 2095 kNm
EI = 699222880 kNm 2
GA = 821943 kN

The corresponding continuum-MDOF model is then built with the above information as
shown in Fig. 5.2. Each cantilever consists of 26 nodes separated at a distance 3.5 m, which is
the interstorey height of the building. The flexural cantilever section consists of two fibres
with areas of A = 10 and separation of r = 20. The yield stress and Young’s modulus of the
fibres are calculated by Eq. (5-7), Eq. (5-10) and Eq. (5-15), given by
fy My
σy = = = 620.6kN
A Ar
EI 2 EI 2 × 699222880
E= = = = 349611.4kN/m
I Ar 2 10 × 202
The hysteretic behaviour of the fibre used in the flexural cantilever is bilinear with 2% isotropic
hardening as shown in Fig. 5.3(a). Similarly, the shear cantilever is determined as shown in
Fig. 5.3(b). Lump masses are then assigned on nodes of the continuum-MDOF model to
produce the desired fundamental period of the model, which is 2.0 s in this example.

The capacity curve of the continuum-MDOF model under a uniform point load at the
nodes is shown in Fig. 5.4. The continuum model deforms 0.175 m at the roof with base shear
70 × 87 .5 = 5985 kN ; however, the continuum MDOF model achieves the same roof
displacement with a smaller base shear 5741 kN due to the difference between the discrete and
distributed loads. A discrete loading in the continuum-MDOF model would create a larger
over-turning moment due to larger moment arm. Hence a smaller base shear results at the
target roof displacement. The difference can be minimised by increasing the number of nodes
of each cantilever, so that the discrete load is close to the uniform distributed loading, but this
will increase the computational effort.

The continuum-MDOF model is also subjected to selected excitations. The excitation


selected has peak ground acceleration equal to 0.12g as shown in Fig. 5.5(a). Its peak ground
acceleration is then magnified two times and three times the original value. It is assumed from

94
Fig. 5.5(b, c, d) that increasing the magnitude of the ground acceleration would force the model
to behave more inelastically and is reflected in the permanent roof displacement after the
excitation. The inelastic behaviour of the model depends on the selected hysteretic behaviour
of the cantilever. Appropriate hysteretic behaviour can be used in the model.

5.3 Numerical investigation of the modified continuum-MDOF model

Three modified continuum-MDOF models with bilinear isotropic hysteretic behaviour are built
to model the three representative buildings which have been discussed in Chapter 4. Their
degrees of structural interaction are 1.6, 3 and 9.4, respectively. The detailed models of the
representative buildings are then compared to their corresponding modified continuum-MDOF
models in nonlinear analysis for both the static and dynamic approaches.

In the static approach, both the detailed models and the modified continuum-MDOF
models have similar capacity curves under invariant lateral forces of the inverted triangularly
distributed loading and the uniformly distributed loading, as shown in Fig. 5.7 to Fig. 5.9. The
detailed models possess a slightly softening characteristic while the modified models have a
slightly strength hardening characteristic. This is due to the progressive yielding of the
members in the detailed model system and the softening characteristic in the materials.
Although there is a slight difference in the responses of the nonlinear static analysis, the
detailed models and the modified models are very much alike.

In the dynamic approach, both the detailed models and the modified models are analysed
by nonlinear response history approach. The maximum responses including the roof
displacement, the interstorey drift ratio and the curvature of the modified models, are plotted
against the responses of the detailed models as shown in Fig. 5.10 to Fig. 5.12. If the responses
of both models are close each another, the data will be located around the solid line which has a
slope equal to 1. If the predictions from the modified continuum-MDOF model are smaller
than those from the detailed model, the data would fall into the region above the solid line.
Otherwise, it would fall into the region below the solid line. For example, the data above the
solid line indicates an underestimated response by using the modified continuum-MDOF model.
The dotted lines, which are above the solid line, have gradient increases from 1.2 to 2
representing underestimation to 1.2 to 2 times.

95
In general, the curvature response of the modified model cannot be predicted well, which
is consistent with the findings of Huang (2009). However, it is also included in the numerical
study for the reason of understanding the behaviour of the modified continuum-MDOF model
in the nonlinear inelastic response range. Only the comparisons in the roof displacement and
the interstorey ratio are made in each modified model and the detailed model.

For the building with degree of structural interaction equal to α = 1.6, it is observed that in
the graphs of the roof displacement and the interstorey drift ratio, most of data are located
around the solid line. It indicates that the maximum responses from the modified
continuum-MDOF model under response history analysis are close to those from the detailed
model in terms of the roof displacement and the interstorey drift ratio.

For the building with degree of structural interaction equal to α = 3.0, the straight line
passing through most of the data in the graph of the roof displacement has gradient equal to 1.4.
Similarly the gradient of the straight line in the graph of the interstorey drift ratio is 1.2. It
indicates that the prediction of the modified model tends to underestimate the seismic response
of the roof displacement and the interstorey drift ratio.

Consider the data of the model with degree of structural interaction of α = 9.4. The
centroid lines of the data in the graphs of the roof displacement and the interstorey drift ratio are
below the solid line. The modified continuum-MDOF model overestimates the response of
the detailed model. The modified continuum-MDOF model is said to be conservative.

Common observations and conclusions can be drawn from the above comparisons.
Under the assumption of an isotropic hardening behaviour, a typical approximation without a
detailed investigation, the interstorey drift ratio of the detailed model does not differ from that
of the continuum-MDOF model significantly. The interstorey drift ratio is well predicted by
the modified continuum-MDOF model. In a comprehensive preliminary assessment, different
hysteretic behaviours can be selected in the analysis depending on the preference of the
engineer. Since the actual behaviour of an assessed building is always unknown before a
detailed investigation, which is both time and resource consuming, a simple but typical
behaviour like bilinear isotropic hardening behaviour can be used to give an approximate result
with a reasonable amount of effort.

96
5.4 General continuum representation

A full-scale detailed model can be constructed specifically for any assessed building in a
detailed assessment, but it is not usual in the preliminary seismic assessment because of limited
resources. Therefore a simplified model has to be used. The proposed modified
continuum-MDOF model is a good simplified model for the preliminary assessment of a large
number of building structures. A continuum-MDOF model can represent building structures
with various lateral deformation characteristics, from flexure to shear. Although the effort and
requirement for the establishment of the model and the analysis have been reduced for
individual building structures, it is still a considerable task to establish models and conduct
analyses for every building structure. It would be beneficial if a single model and its
analysis result could be applied to other building structures which have similarity to the single
model; for instance, having similar deformation characteristics and dynamic properties.

For a linear elastic SDOF model, its behaviour is governed by the fundamental frequency,
which is related to the ratio of stiffness to mass. For any two SDOF models, if their ratios of
stiffness to mass are identical though their values of stiffness and mass are not equal, they
would have identical fundamental frequency and an identical response under same ground
motion. An SDOF model is objective as its behaviour is only governed by an objective
parameter, which is the frequency. A linear elastic continuum model is subjective in contrast
to an SDOF model. This is because the responses of a continuum model depend on both the
fundamental frequencies and the height of the model. However, a dimensionless continuum
model, which has taken the influence of height into the mode shape, is objective. Any two
dimensionless continuum models with identical deformation characteristics and dynamic
properties would have identical responses.

A nonlinear inelastic model is characterised by its hysteretic behaviour and its behaviour is
governed by more parameters compared to linear elastic models. The response of a nonlinear
inelastic SDOF model depends on the fundamental frequency, the initial stiffness and the
force-displacement relationship. A nonlinear inelastic SDOF model is unique when it is
analysed in the time domain. Capacity spectrum analysis, which is a static analysis equivalent
to response history analysis, is less restrictive and a nonlinear SDOF model could be related to
a similar nonlinear SDOF model.

97
An SDOF model in capacity spectrum analysis is characterised by the capacity curve, the
demand curve and the equivalent viscous damping ratio. Since it is a graphical means of
analysis, the performance point of the model depends on the geometry of the intersection
between two curves and the viscous damping ratio. For any two SDOF models with
proportional geometry of the intersection, in capacity spectrum analysis, the two performance
points would be proportional under the same viscous damping. The two SDOF models are
similar because of geometric proportion in graphical analysis.

A continuum model, which can be decomposed into several SDOF models, can be related
to other models through the same approach as in the SDOF model. By comparison of two
different continuum flexural-shear models with known similarity, the relationship between
their responses is defined. A general continuum representation can then be established based
on the relationship, which describes how two similar models, the prime model and the sub
models, are related.

5.4.1 Conditions of model similarity


Two different continuum flexural-shear models are said to be similar if they satisfy the
following conditions:

1. They have identical degrees of structural interaction


2. They have identical fundamental periods
3. They have identical static roof drift ratios when their lateral yield strengths are reached
4. Their capacity spectra are proportional to each another under the same secant stiffness
5. Their dimensions differ from each another by a constant factor k

By considering two continuum models and two sets of parameters shown in Fig. 5.13, the
mathematical relationships between two similar models can be found under the above
conditions. The subscript notation “p” of the parameter refers to the parameters of the prime
model, while the subscript notation “s” refers to the sub model. The first four conditions are
expressed mathematically from Eqs (5-17) to (5-20), which can be found in the appendix.

1. Identical degree of structural interaction,


α p = αs

98
GAp GAs
Hp = Hs (5-17)
EI p EI s

2. Identical fundamental period,


Tp = Ts

m p H p4 2π ms H s4 2π
=
EI p γ 2 (γ 2 + α 2 ) EI s γ 2 (γ 2 + α 2 )

m p H p4 ms H s4
= (5-18)
EI p EI s

3. Identical static roof drift ratio when their lateral yield strengths are reached under lateral
load, either in uniform or inverted triangular distributions,
yp ys
=
Hp Hs

w p H 3p ws H s3
= (5-19)
EI p EI s

4. Proportional capacity spectrum,


S sa S sd
=
S sa S pd

ws H s wp H p y yp
÷ = s ÷
ms H s ∫ φdz m p H p ∫ φdz PF PF

ws m p y
⋅ = s
w p ms yp

ws m p H s
⋅ = (5-20)
w p ms H p

The above four conditions involve five pairs of variables. They are the height of the
model (H), the distributed mass (m), the magnitude of the lateral force (w), the flexural rigidity
(EI) and the shear rigidity (GA). From the fifth condition, the height of the sub model is
expressed in terms of the height of the prime model as
H s = kH p (5-21)

99
where k is an arbitrary proportional constant and is referred as the height ratio. By using Eqs
(5-17), (5-18), (5-19), (5-20) and (5-21), the parameters of the sub model are solved in as
follows,
ws = bkwp (5-22)

ms = bm p (5-23)

EI s = bk 4 EI p (5-24)

GAs = bk 2GAp (5-25)

where b is another arbitrary constant.

The mass-per-unit-floor area of different buildings is more or less the same, especially for
building structures with similar functions. For similar building structures, the difference in
mass is mainly due to the dimensions of the floor area. Therefore the difference in the
distributed mass along the height should be proportional to the floor area. From the fifth
condition, when the height is increased by a factor of k, the floor area should be increased by k2
as does the distributed mass. By assuming b = k2, Eqs (5-22), (5-23), (5-24) and (5-25)
become
ws = k 3 w p (5-26)

ms = k 2 m p (5-27)

EI s = k 6 EI p (5-28)

GAs = k 4GAp (5-29)

Two continuum flexural-shear cantilever models are said to be similar if their parameters
satisfy the above equations. By satisfying the above conditions, the relationship of the
responses between two similar continuum models are investigated.

5.4.2 Static responses of two similar continuum models


The lateral displacement, the interstorey drift ratio and the curvature of the prime model under
external uniform loading are given by
wp H p4 ⎡α sinh α + 1 2⎛ z 2 ⎞⎤
y p (z ) = ⎢ (cosh α z − 1) − α sinh α z + α ⎜
⎜ z − ⎟⎥ (5-30)
EI pα 4 ⎣ cosh α ⎝ 2 ⎟⎠⎦

wp H 3p ⎡α sinh α + 1 ⎤
IDR p ( z ) = 3 ⎢
sinh αz − α cosh αz + α (1 − z )⎥ (5-31)
EI pα ⎣ cosh α ⎦

100
wp H p2 ⎡α sinh α + 1 ⎤
curp ( z ) = 2 ⎢
cosh αz − α sinh αz − 1⎥ (5-32)
EI pα ⎣ cosh α ⎦

x
where z = is the dimensionless height of the model and z ∈ [0,1] . Based on the properties
H
defined in the Eqs (5-21) and (5-26) to (5-29), the static responses of two similar continuum
models under uniform lateral loading can be found and are related by,
ys ( z ) = k ⋅ y p ( z ) (5-33)

IDRs ( z ) = IDR p ( z ) (5-34)

1
curs ( z ) = ⋅ curp (z ) (5-35)
k
The above relationships also hold for static responses under inverted triangular loading.

5.4.3 Linear dynamic responses of two similar continuum models


The linear dynamic response of a continuum model can be found from modal analysis. The
lateral displacement, the interstorey drift ratio and the curvature of the prime model are
y p (z , t ) = ∑ φ (z )q (t )
i i (5-36)

1
IDRp ( z, t ) = ∑φi′(z )qi (t ) (5-37)
Hs
1
curp ( z, t ) = ∑φ ′′(z )i qi (t ) (5-38)
H s2
where
q&&i + 2ξωi q&i + ωi2 qi = − PFi u&&g (5-39)

Since both the prime and the sub model have identical fundamental frequencies and degrees of
structural interaction, they would have same modal coordinate in Eq. (5-39) under the same
damping. By substituting Eq. (5-21) into Eqs (5-37), (5-38) and (5-39), the linear dynamic
responses of the prime and that of the sub models are as follows,
ys ( z , t ) = y p ( z , t ) (5-40)

1
IDRs ( z, t ) = ⋅ IDRp (z, t ) (5-41)
k
1
curs ( z, t ) = ⋅ curp ( z, t ) (5-42)
k2

101
5.4.3. Performance point of two similar models under capacity spectrum analysis
Response history analysis is regarded as one of the most reliable methods for seismic
assessment, but it depends on the modelling and the numerical method which cannot be
expressed easily. An equivalent analysis, capacity spectrum analysis, is used to approximately
predict the inelastic behaviour of the prime and the sub models.

From the conditions of similar models, the capacity spectrum of the prime and the sub
model are proportional to each another under the same secant stiffness. The proportional
constant is k, which is the proportional constant of the dimension or the height ratio of the
building structures stated in Eq. (5-20) and Eq. (5-21). If the performance point of the prime
model is (Sdo , Sa) under specific excitation u&& g , the performance point of the sub model would

be (kSdo , kSa) under same strength ratio or excitation ku&&g . It is explained in the followings.

When the performance point of the prime model is (Sdo, Sao), it has to satisfy two criteria,

1. The performance point is on the demand spectra, which is the maximum response of
the corresponding SDOF model under the ground motion at a particular angular
frequency and equivalent damping. The x-coordinate of the performance point is
S do = max[qo (t )] (5-43)

where
q&&o (t ) + 2ξ oωo q&o (t ) + ωo2 qo (t ) = −u&&g (t ) (5-44)

Eq. (5-44) is the equation of motion of the corresponding SDOF model; ξ o is the

damping ratio of demand curve; ωo is the angular frequency; qo (t ) is the modal

coordinate and u&&g is the ground acceleration.

The y-coordinate is the pseudo acceleration, which is


Sao = ωo2 Sdo (5-45)

2. The performance point is on the demand spectra and is also on the capacity spectrum,
hence the equivalent damping provided by the corresponding ductility on the
capacity curve is consistent to the demand curve in Eq. (5-45). That is

102
⎛ S do ⎞⎟
ξ eq = f ⎜ μ = = ξo (5-46)
⎜ S y ⎟⎠

where ξeq = f (μ ) is the function of damping ratio at particular ductility μ .

Similarly, assuming the performance point of the sub model to be (Dk , Ak ) and would

satisfy the above two criteria.

I. The x-coordinate is
Dk = max[qk (t )] (5-47)

where
q&&k + 2ξ k ωo q& k + ωo2 qk = − ku&&g (5-48)

The y-coordinate is
Ak = ω k2 Dk (5-49)

II. The damping ratio at the performance point is given by


⎛ D ⎞
ξk = f ⎜ μ = k ⎟ (5-50)
⎜ k ⋅ S y ⎟⎠

By comparing Eq. (5-44) and Eq. (5-48) and assuming ξ o = ξ k , the modal coordinate of

the prime model is related to that of the sub model as


qk (t )
qo (t ) = (5-51)
k
Substituting Eq. (5-50) into Eq. (5-47) and Eq. (5-49), the performance point of the sub
model can be expressed in terms of the prime model as
Dk = k ⋅ S do (5-52)

Ak = k ⋅ S ao (5-53)

Checking the assumption by substituting Eq. (5-51) into Eq. (5-50), the damping ratio at
the performance point is equal to that of the prime model, so the assumption is correct.

103
5.4.4 Modified modal pushover analysis
Nonlinear inelastic behaviour of building structures is analysed by the modified modal
pushover analysis by Chopra (2004), which is a well recognised method of analysis. It argues
that the inelastic range of responses is mainly contributed by the yielding of the fundamental
mode. The influence of the higher modes could be considered as linear elastic and is
combined to the yielding of the fundamental mode, similar to that in modal analysis. Based
on this principle, the inelastic responses are predicted from the modified modal pushover
analysis.

When the prime model and the sub model have the same strength ratio, the higher mode
response of the sub model is k times of the prime model. The same conclusion can be found
in the response from the capacity spectrum analysis of the fundamental mode of the prime and
the sub model according to section 5.4.3. Therefore the lateral displacement of the sub
model is k times that of the prime model under modal combination. It is given by,
ys ( z , μ ) = k ⋅ y p ( z , μ ) (5-54)

The interstorey drift ratio is the derivative of the lateral displacement which is
∂ys ( z, μ )
IDRs ( z, μ ) = = IDRp (z, μ ) (5-55)
∂x
The curvature is the second derivative of the lateral displacement which is
1
curs ( z, μ ) = ⋅ curp ( z, μ ) (5-56)
k
When the two continuum models are similar and they have the same strength ratio, the response
of the prime model and the sub model is related by Eqs. (5-54), (5-55) and (5-56). It indicates
that two buildings with identical deformation characteristics, fundamental periods, the same
proportional capacity and dimensions, their seismic responses for the same strength ratio are
proportional to each another according to the proportionality constant in their dimensions.

5.4.5 Conclusion
A modified continuum-MDOF model has been proposed for preliminary seismic assessment
of a larger number of building structures. Although the effort and requirement for the
establishment of the model and the analysis have been reduced for individual building
structures, it is still an involved task to establish the models and to conduct analysis for each
individual building structure. It would be beneficial if a single model and analysis method
could be applied to other building structures which have similarity with the single model.

104
A general continuum representation is thus established, in which models with similarity
as in Section 5.4.1, are related in responses by the height ratio for the same strength ratio.
When there is a continuum model, its responses can be applied to other models in regard to
linear static behaviour, linear dynamic behaviour and modified modal pushover analysis.
From the general continuum representation, a modified continuum-MDOF model and its
analysis result could be applied to other similar continuum-MDOF models by the height ratio
for the same strength ratio.

5.5 Concluding remarks

Based on the investigation of the properties of the modified continuum-MDOF model and the
general continuum representation, the following conclusions can be drawn.

1. The continuum-MDOF model by Huang (2009) is an improvement of the previous


continuum model because it has almost identical linear elastic response and allows
various uses of the hysteretic behaviour in modelling cantilever beams.

2. A modified continuum-MDOF model compared to Huang’s (2009) continuum-MDOF


model is more suitable for rapid and preliminary seismic assessment because both the
required information and the effort for analysis are less. It gives a reasonable good
prediction of the interstorey drift ratio under response history analysis as compared to the
detailed model.

3. A general continuum representation is derived, in which the responses of similar


continuum models are related by the height ratio for the same strength ratio. When
there is a continuum model, its responses can also be applied to other similar models in
regard to linear static behaviour, linear dynamic behaviour and modified modal pushover
analysis. Based on the general continuum representation, a modified
continuum-MDOF model discussed in paragraph above and its analysis result are applied
to other similar continuum-MDOF models by the height ratio for the same strength ratio.

105
b

Fig. 5.1 Strain diagram of a rectangular section

Shear cantilever
87.5

Flexural cantilever Centre line

A = 10
3.5

R = 20

Fibre element

(a) (b)

Fig. 5.2 (a) Continuum-MDOF model; (b) Section of flexural cantilever

106
620.64 2095.93

E=349611.44 GA=821943.63

εy γy
(a) (b)
Fig. 5.3 Stress strain relationship (a) Fibres in the flexural cantilever; (b) Shear cantilever

8000

7000

6000

(0.175, 5741)
Base shear (kN)

5000

4000

3000

2000

1000

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8
Roof displacement (m)
Fig. 5.4 Capacity curve of continuum-MDOF model

107
0.2 0.2
Ground acceleration (g)

Roof displacement (m)


0.1 0.1

0 0

-0.1 -0.1

-0.2 -0.2
0 30 60 90 120 150 180 0 30 60 90 120 150 180
Time (s) Time (s)
(a) (b)
0.4 1
Roof displacement (m)

Roof displacement (m)

0.75
0.2
0.5

0 0.25

0
-0.2
-0.25

-0.4 -0.5
0 30 60 90 120 150 180 0 30 60 90 120 150 180
Time (s) Time (s)
(c) (d)
Fig. 5.5 (a) Excitation from PEER with record ID: P1121; (b) The roof displacement under
the excitation in (a); (c) The roof displacement under the excitation two times of (a); (d) The
roof displacement under the excitation three times of (a)

108
Prime model Sub model
ms
ws
mp
wp

EIs GAs Hs
EIp GAp Hp

Sas Sub model


Spectra acceleration

Prime model
Sa

Sd Sds
Spectra displacement

Fig. 5.6 Prime and sub model

109
10000

8000
Base shear (kN)

6000

4000

2000
continuum
detail
0
0 0.5 1 1.5 2
Roof displacement (m)

10000

8000
Base shear (kN)

6000

4000

2000
continuum
detail
0
0 0.5 1 1.5 2
Roof displacement (m)
Fig. 5.7 Capacity curve of building α =1.6. (a) Inverted triangular loading;
(b) Uniform loading

110
12000

10000
Base shear (kN)

8000

6000

4000

2000 continuum
detail
0
0 0.5 1 1.5 2
Roof displacement (m)

12000

10000
Base shear (kN)

8000

6000

4000

2000 continuum
detail
0
0 0.5 1 1.5 2
Roof displacement (m)
Fig. 5.8 Capacity curve of building α =3.0. (a) Inverted triangular loading;
(b) Uniform loading

111
10000

8000
Base shear (kN)

6000

4000

2000
continuum
detail
0
0 0.2 0.4 0.6 0.8 1
Roof displacement (m)

10000

8000
Base shear (kN)

6000

4000

2000
continuum
detail
0
0 0.2 0.4 0.6 0.8 1
Roof displacement (m)
Fig. 5.9 The capacity curve of building α =9.4. (a) Inverted triangular loading;
(b) Uniform loading

112
Roof drift (%) Interstory drift ratio (%)
2 2

1.5 1.5
Original model

Original model
1 1

0.5 0.5

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Continuum model Continuum model
(a) (b)
-1
x 10
-3
Curvature (m )
3

2.5

2
Original model

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3
Continuum model x 10
-3

(c)
Fig. 5.10 Model with α = 1.6. (a) Roof drift; (b) Interstorey drift ratio; (c) Curvature

113
Roof drift (%) Interstory drift ratio (%)
2 2

1.5 1.5
Original model

Original model
1 1

0.5 0.5

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Continuum model Continuum model
(a) (b)
-1
x 10
-3
Curvature (m )
3

2.5

2
Original model

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3
Continuum model x 10
-3

(c)
Fig. 5.11 Model with α = 3.0. (a) Roof drift; (b) Interstorey drift ratio; (c) Curvature

114
Roof drift (%) Interstory drift ratio (%)
2 2

1.8

1.6
1.5
1.4
Original model

Original model
1.2

1 1

0.8

0.6
0.5
0.4

0.2

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Continuum model Continuum model
(a) (b)
-1
x 10
-3
Curvature (m )
3

2.5

2
Original model

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3
Continuum model x 10
-3

(c)
Fig. 5.12 Model with α = 9.4. (a) Roof drift; (b) Interstorey drift ratio; (c) Curvature

115
Chapter 6 

Generalised Inelastic Response Spectra 

6.1 Introduction

In preliminary seismic assessment, response spectra developed from an SDOF model has
provided an efficient and accurate means for predicting the linear elastic response of building
structures. Based on the fundamental period which is usually estimated from empirical
equations, the corresponding response is read from the spectrum at the chosen damping ratio
and the predicted fundamental period. The use of drift spectra is an alternative mean similar to
the response spectra, in which the interstorey drift ratio is directly read from the spectra at the
fundamental period. The drift spectrum (Iwan, 1997) and the generalised interstorey drift
spectrum (Miranda and Akkar, 2006) are the examples of such spectra. They were developed
from the continuum modelling technique and linear response history analysis.

116
Although response spectra and drift spectra have been used widely in seismic assessment,
their limitations in only being applicable for linear elastic responses are also well known. To
predict the nonlinear inelastic response probably occurring in an earthquake, two different
approaches developed from the SDOF models, namely the equivalent linear method and the
nonlinear inelastic method are used. The two approaches produce the equivalent linear
response spectrum and the inelastic response spectrum.

The equivalent linear response spectrum is developed from the response of an equivalent
linear SDOF model of its corresponding nonlinear inelastic system. A nonlinear system is
represented by an equivalent linear system which has a different period and damping ratio (Lin
and Miranda, 2008). The period and the damping ratio of the equivalent linear system are
defined through a large scale comparison between different models under response history
analysis. An alternative method to define the equivalent system is to replace the nonlinear
system with a linear system of a greater damping ratio while the system’s period is not changed.
A displacement coefficient is used to convert the linear response to the inelastic response
(Miranda, 2000). In both approaches, the equivalent linear system will not behave identically
to the nonlinear system in the time domain, but is intended to capture the maximum response of
the nonlinear system.

Inelastic response spectra are developed from SDOF models which possess hysteretic
behaviour. Inelastic response spectra are of two different types depending on the independent
variables used to predict the response. The response of the first type is expressed by the
ductility ratio while the second type is expressed by the strength ratio. For preliminary seismic
assessment, the strength ratio is more convenient than the ductility ratio as the ductility ratio is
an implicit parameter for non-seismically designed buildings. Iteration is needed to determine
the ductility ratio under the selected ground motion. Therefore, for preliminary assessment, in
which the assessment effort has to be minimised, inelastic response spectra that are expressed
by the strength ratio are preferred.

The strength ratio (R) from the nonlinear static analysis of a model, which is also referred
to as the strength reduction factor, is defined as,
PF1MSa
R= (6-1)
fy

where PF1 is the participation factor of the fundamental mode; M is the mass of the system; Sa is
117
the linear pseudo-acceleration in spectral ordinates; fy is the lateral yield strength of the system
under lateral load. The numerator represents the maximum base shear induced by the ground
motion of the fundamental mode. The denominator is the base shear at yield. The ratio hence
represents the induced base shear by the earthquake to the yield strength of the system. When
the ratio is larger than unit, yielding of the system occurs. Given any system, the value of the
pseudo acceleration Sa can be obtained from the linear elastic response spectra in the design
codes by the fundamental period. The lateral yield strength of the buildings can either be
obtained from the pushover analysis or estimation from the design lateral load of the building.

High-rise building structures are complicated in behaviour and cannot be described only
by an SDOF model, which only monitors one single response. They are represented by MDOF
models for capturing different responses simultaneously. When an SDOF model is used to
assess a building structure, conversion factors or procedures have to be considered to convert
the responses of an SDOF system to the responses of an MDOF system. In the process, it is
assumed that the mode shapes exist in the nonlinear inelastic response range or they are
identical to the mode shapes in the linear elastic region. The result obtained from the SDOF
system is used to amplify the linear elastic mode so that an inelastic result of an MDOF system
can be predicted. The inelastic response however differs from the elastic mode not only by
magnitude, but also by shape. The conversion process from the nonlinear SDOF model to a
nonlinear MDOF model is unfavourable.

A tall building structure experiences higher mode effects which also cannot be represented
by an SDOF model. Several SDOF models could be combined based on modal analysis to
capture the additional force and deformation due to the higher mode effect, but this can only be
valid in the linear elastic response range. Once the building structure is excited into the
inelastic response range, methods based on the SDOF model are not accurate but can be
regarded as approximate. In situations with significantly higher mode effects and an inelastic
response range, an MDOF model is usually used to represent the building structure and is
analysed by response history analysis.

In conclusion, response spectra developed from SDOF models are not suitable for
high-rise building structures which may experience an inelastic range of responses and
significantly higher mode effects. The modified continuum-MDOF model discussed in
Chapter 5 is therefore proposed for preliminary seismic assessment. Its simplicity in

118
modelling and its generality for application to building structures make it a favoured model for
the development of a response spectrum that suits high-rise building structures having
significantly high mode effects. From the general continuum representation, the responses
and the analysis results of a modified continuum-MDOF model can be applied to other similar
continuum-MDOF models. This will help to establish a generalised inelastic response
spectrum, which is a spectrum-based method, for high-rise building structures with various
deformation characteristics for use in preliminary seismic assessment.

6.2 Methodology

From the general continuum representation, the inelastic responses obtained from a continuum
model can be used to approximate the inelastic response of other similar models using the
height ratio. If a prime model (modified continuum-MDOF model) is studied, a sub model
(assessed building) which is similar to the prime model can be approximated. The
relationships for the maximum inelastic responses between the prime model and the sub model
given by Eqs (5-53), (5-54) and (5-55) are

⎪ ys = k ⋅ y p
⎪⎪
⎨ IDRs = IDR p (6-2)

⎪ curs = 1 ⋅ curp
⎩⎪ k
where the subscript “s” refers to the sub model while “p” refers to the prime model. The value
k is the height ratio and is the ratio of the height of the sub model to that of the prime model.

By substituting the expression of the height ratio k into Eq. (6-2), the inelastic responses of
the sub model are expressed as follows,
⎧ yp
⎪ ys = × Hs
⎪ H p

⎨ IDRs = IDR p (6-3)

⎪ curs = curp × H p × 1
⎪⎩ Hs

In Eq. (6-3), any information concerning the prime model is extracted and is in dimensionless
form. The dimensionless responses given in Eq. (6-4) are referred to as the generalised
inelastic responses,

119
⎧ yp
⎪ Gy =
⎪ Hp

⎨G IDR = IDR p (6-4)

⎪Gcur = curp × H p
⎪⎩

The responses of the sub model are then given by,



⎪ y =G ×H
⎪⎪ s y s

⎨ IDRs = GIDR (6-5)


⎪ 1
⎪ curs = Gcur ×
⎪⎩ Hs

By multiplying the inelastic generalised responses to the height of the sub model, the responses
of the sub model can be predicted according to Eq. (6-5).

Eq. (6-5) is derived from the general continuum representation, which applies the
responses to similar models in nonlinear inelastic behaviour by the modified modal pushover
analysis. The modified modal pushover analysis and nonlinear response history analysis
both can reveal the nonlinear inelastic behaviour of a building structure. They are regarded
as the equivalent method of each another and their results are consistent. Therefore, the
observations in the general continuum representation made by the modified modal pushover
analysis are also applied to the nonlinear response history analysis. Nonlinear response
history analysis is then conducted on the prime model to obtain the generalised inelastic
responses.

6.3 Generalised inelastic response spectra

Generalised inelastic response spectra are developed from the modified continuum-MDOF
model. The modified continuum-MDOF gives reliable predictions of the roof displacement
and the interstorey drift ratio. On the other hand, it tends to underestimate the curvature; hence
the generalised inelastic curvature spectra are not recommended for the purpose of assessment,
but it included for a general background understanding.

120
6.3.1 Generalised inelastic response
The generalised inelastic responses of a continuum flexural-shear model depend on the degree
of structural interaction, the fundamental period and the strength ratio of the particular
hysteretic behaviour under the expected ground motion. The relationships between the
responses and the various parameters are presented in the form of response spectra that are
developed from the results of the response history analysis of 168 MDOF models under
selected ground motions.

The modified continuum-MDOF model discussed in Chapter 5 is adopted to develop the


spectra. 168 modified continuum-MDOF models including 8 different degrees of structural
interaction ranging from 1 to 8 with intervals of 1, and 21 different fundamental periods ranging
from 1s to 5s with interval of 0.2s, are constructed. The height of the modified model is
selected as 100m. For each model, four groups of ground motion are used to analyse the
model, thus 4 different strength ratios, namely 1.5, 2.0, 3.0, 4.0, are considered.

The yield strength of the model for defining the strength ratio is obtained from pushover
analysis under an inverted triangular distributed force (Case 1) and a uniform distributed force
(Case 2). Thus two different kinds of strength ratios for different applied loads are defined.
They represent two different equivalent seismic forces experienced by the model and they have
been recognised as the upper and lower bound values of the equivalent seismic force. Each
group of excitations in Table 6.1 contains 11 ground motions which are obtained from FEMA
440 (Appendix F) and are scaled to achieve the defined strength ratio.

6.3.2 Generalised roof displacement spectra


The generalised roof displacement for the strength ratio defined by an inverted triangular load
(Case 1) and a uniform load (Case 2) under different degrees of structural interaction are plotted
in Fig. 6.1 to Fig. 6.8. It can be seen that the generalised roof displacement increases with the
strength ratio. Along the axis of the fundamental period, it decreases to a minimum at a
fundamental period of about 2.25s and then increases again to a relatively stable value, in which
the variation is small in the period of 3s to 5s. The influence of the degree of structural
interaction can be seen by comparing of individual spectra. The spectra with a degree of
structural interaction of 6 give the largest response.

121
6.3.3 Generalised interstorey drift ratio spectra
The spectra of the generalised interstorey drift ratios are shown in Fig. 6.9 to Fig. 6.16. Similar
to the roof displacement spectra, the response increases with the strength ratio. The minimum
response occurs at a fundamental period of about 2.2s under the selected ground motions. The
spectra with a degree of structural interaction of 8 give the maximum response.

6.3.4 Generalised curvature spectra


The spectra of standardised curvature are plotted in Fig. 6.17 and Fig. 6.24. It can be observed
that, in general, with an increasing degree of structural interaction, the response increases. The
maximum response occurs at a spectrum with a degree of structural interaction of 6.

6.4 Procedure of assessment using generalised inelastic response

For performance-based preliminary seismic assessment, the behaviour of a building structure is


determined through its expected responses under ground motion. The responses of the assessed
building structures can then be predicted by the generalised inelastic response spectra. The
procedure of the assessment using the generalised inelastic response spectra includes three
major parts,

1. Collecting information on the assessed building structure


2. Estimating the responses of the assessed building
3. Defining the performance of the assessed building

6.4.1 Collecting information of assessed building structure


The information on the assessed building is collected so that a corresponding modified
continuum-MDOF model can be defined. The information required for defining the
corresponding continuum-MDOF model is as follows,

1. The height of the building


2. The degree of structural interaction of the building
3. The fundamental period of the building
4. The strength ratio

122
A suggested method for obtaining information on the building for establishing the
corresponding continuum-MDOF model is presented. The height of the assessed building is
approximately equal to
No. of storeys μ interstorey height
The degree of structural interaction can be estimated by the drift ratio as shown in Fig. 3.6 in
chapter 3 and the fundamental period of the building can be predicted by Eq. (3-16) also
suggested in chapter 3, and is used to determine the linear elastic spectral acceleration for
calculating the strength ratio. Moreover there are various alternatives for obtaining the
fundamental period, for instance empirical formulas suggested in codes of practice or
estimations from engineering experience.

The strength ratio presented by Eq. (6-1) requires information about the participation factor,
the mass of the building, the pseudo acceleration at the fundamental period and the yield
strength of the building. The participation factor of the fundamental mode of the flexural-shear
model can be read from Fig. 6.25 by the degree of structural interaction. The mass of the
building is approximately equal to

Design floor load μ floor area μ no. of floors / gravitational force

The pseudo acceleration is simply obtained from the corresponding pseudo-acceleration spectra
by the fundamental period at the initial elastic viscous damping ratio.

The yield strength of the building can be predicted from the design lateral strength. It is
approximated by

Overstrength coefficient μ design lateral pressure μ projected area

The value of the overstrength coefficient is an implicit design parameter which is unknown.
However, from the suggestions in the SEAOC Blue Book (SEAOC, 2009), a conservative value
of the overstrength coefficient can be assigned as either 2.0 or 2.5 depending on the system,
which has bee discussed in Section 4.2.3. Moreover, a more precise value can be defined from
pushover analysis.

123
6.4.2 Estimating responses of assessed building
To predict the response of the assessed building using the generalised inelastic response, the
strength ratio of the assessed building under the expected ground motion has to be defined.
When the strength ratio is known, the generalised inelastic responses can be found from Fig. 6.1
to Fig. 6.24. The inelastic responses of the assessed building can then be approximated
according to Eq. (6-5)

6.4.3 Defining performance of assessed building


There are standards and studies that define the relationship between the responses and the
performance. By comparison to the suggested magnitudes of response stated in the current
standards and studies, the performance of a building can be found; for instance the deformation
limits in the code of practice ATC 40 and experimental findings by Li at el. (2006). In the
preliminary assessment, in which a large number of buildings are being assessed, failure to
meet the standard indicates the need for further detailed assessment.

6.5 Numerical investigation

A representative building is selected from Chapter 4 for illustration in the proposed generalised
inelastic response spectra. The selected building is the 25-storey wall-frame building. The
height of the wall-frame building is 87.5 m with an interstorey height of 3.5 m. The design
roof drift is 0.14m under a lateral pressure of 2.80 kPa, which decreases parabolically to 0 at the
base. The degree of structural interaction (α) of the building is approximately equal to 3, using
the method of drift ratio RΔ/Rδm in Fig. 3.6. The mass at the floor level of the models is 137 Mg
with fundamental period of 2 seconds.

The building is constructed as a 2-dimensional model using the nonlinear


finite-element-based programme “OpenSees” for nonlinear response history analysis. The
selected building is subjected to the ground motion in Table 6.1. The responses under the
ground motions are compared to the predictions obtained from the generalised inelastic
response spectra method.

The information required to establish the corresponding simplified continuum-MDOF


model is summarised as follows:

124
The height of the building = 87.5 m
The degree of structural interaction = 3
The fundamental period = 2 s
The floor mass of the building = 137 Mg
The pseudo acceleration obtained from Fig. 6.26 = 2.4 ms-2
From pushover analysis with the inverted triangular load pattern in Fig. 6.27, the yield
strength of the building = 7500 kN

The strength ratio (R) in Eq. (6-1)


PF1 MS a 1.46 × 3425 × 10 3 × 2.4
= =
fy 7600 × 1000
= 1.58

From the generalised inelastic response spectra, the dimensionless responses are

From Fig. 6.2, Gy = 0.005 (Case 1)


From Fig. 6.6, Gy = 0.005 (Case 2)
From Fig. 6.10, GIDR = 0.007 (Case 1)
From Fig. 6.14,GIDR= 0.007 (Case 2)
From Fig. 6.18, Gcur = 0.10 (Case 1)
From Fig. 6.22, Gcur = 0.10 (Case 2)

From Eq. (6-5), the inelastic responses of the assessed building are

⎪ ymax = G y × H

⎨ IDRmax = GIDR
⎪ 1
⎪ curmax = Gcur ×
⎩ H
Hence,

⎪ ymax = 0.005 × 87.5

⎨ IDRmax = 0.007
⎪ 1
⎪ curmax = 0.10 ×
⎩ 87.5
then

125
⎧ ymax = 0.44 m

⎨ IDRmax = 0.007
⎪ −1
⎩ curmax = 0.0011 m

Comparisons between the results obtained from the response history analysis and the
predictions obtained from the generalised response spectra method are shown in Figs. 6.28 to
6.30. It can be seen that the predictions from the generalised response spectra, the solid line,
are more conservative than the results obtained from the response history analysis. The reason
is due to the simplification of structural system from a detailed model to the modified
continuum-MDOF model. An equivalent cantilever is used for each storey of the modified
continuum-MDOF model. The modified continuum-MDOF model with isotropic hysteretic
hardening behaviour would dissipate more energy with smaller displacement compared to the
detailed model. For both the roof displacement and the interstorey drift ratio, the proposed
generalised inelastic response spectra method gives a reasonably good prediction, as compared
to the response history analysis which requires significantly larger effort.

126
6.6 Concluding Remarks

A preliminary assessment method using the generalised inelastic response is proposed. The
generalised inelastic responses are derived from the modified continuum-MDOF model
through the general continuum representation, where the responses of a prime model can be
applied to other similar models using the height ratio. The generalised responses are expressed
in the form of spectra, which describe the relationship between various parameters, including
the fundamental period, the degree of structural interaction and the strength ratio of the building
under the expected ground motions. The generalised inelastic response spectra are developed
from the responses of 168 modified continuum-MDOF models, of heights 100 m, having 8
different degrees of structural interaction and 23 different fundamental periods using nonlinear
response history analysis.

The responses of the assessed building structure are determined from the suggested
general inelastic response spectra and the height of the assessed building, and are given by

⎪ y = Gy × H

⎨ IDR = GIDR
⎪ 1
⎪ cur = Gcur ×
⎩ H
The numerical investigation of a 25-storey wall-frame building shows that the proposed method
using the generalised inelastic response spectra has successfully predicted the seismic response
of the building and it agrees satisfactorily with the response history analysis of the
corresponding 2-dimensional finite-element based model. The proposed method has provided
a simple and quick, yet accurate, means of estimating the inelastic seismic response of a
structure, particularly suited for preliminary seismic assessment.

127
Table 6.1 Ground Motion records
Extracted from FEMA 440, Appendix F, Table F-6
Magnitude PGA PGV
# Identifier Earthquake Date Station Location Source
(Ms) (g) (cm/sec)
El Centro imp. Co. Cent
1 ICCC000 Superstitn 11-24-87 6.6 0.358 46.4 CDMG
(01335)
Canyon Country-W Lost
2 LOS000 Northridge 1-17-94 6.7 Cany 0.41 43 USC
(90057)
3 G02090 Loma Prieta 10-18-89 7.1 Gilroy Array #2 (47380) 0.322 39.1 CDMG
Chi-Chi,
4 TCU122N 9-20-99 7.6 (TCU122) 0.261 34 CWB
Taiwan
5 G0390 Loma Prieta 10-18-89 7.1 Gilroy Array #3 (47380) 0.367 44.7 CDMG
Canoga Park-Topanga Can
6 CNP196 Northridge 1-17-94 6.7 0.42 60.8 USC
(90053)
7 CHY101W Chi Chi, Taiwan 9-20-99 7.6 (CHY101) 0.353 70.6 CWB
El Centro imp. Co. Cent
8 ICC090 Superstitn 11-24-87 6.6 0.258 40.9 CDMG
(01335)
Canoga Park-Topanga Can
9 CNP106 Northridge 1-17-94 6.7 0.356 32.1 USC
(90053)
10 E02140 Imperial Valley 10-15-79 6.9 El Centro Array #2 (5115) 0.315 31.5 USGS
11 E11230 Imperial Valley 10-15-79 6.9 El Centro Array #11 (5058) 0.38 42.1 USGS
CDMG: California Division of Mines and Geology
CWD: Central Weather Bureau, Taiwan
USC: University of Southern California
USGS: U.S. Geological Survey

128
α =1
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.1 Generalised roof drift spectra (a) and (b) (Case 1)

129
α =3
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.2 Generalised roof drift spectra (c) and (d) (Case 1)

130
α =5
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
0.04

0.035

0.03

0.025
y

0.02
G

0.015 R=4
R=3
0.01
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.3 Generalised roof drift spectra (e) and (f) (Case 1)

131
α =7
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(g)
α =8
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.4 Generalised roof drift spectra (g) and (h) (Case 1)

132
α =1
0.04

0.035

0.03

0.025
y

0.02
G

0.015 R=4
R=3
0.01
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
0.04

0.035

0.03

0.025
y

0.02
G

0.015 R=4
R=3
0.01
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.5 Generalised roof drift spectra (a) and (b) (Case 2)

133
α =3
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.6 Generalised roof drift spectra (c) and (d) (Case 2)

134
α =5
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
0.04

0.035

0.03

0.025
y

0.02
G

0.015 R=4
R=3
0.01
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.7 Generalised roof drift spectra (e) and (f) (Case 2)

135
α =7
0.04

0.035

0.03

0.025
y

0.02
G

0.015
R=4
0.01 R=3
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(g)
α =8
0.04

0.035

0.03

0.025
y

0.02
G

0.015 R=4
R=3
0.01
R=2
0.005 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.8 Generalised roof drift spectra (g) and (h) (Case 2)

136
α =1
0.04

0.035

0.03

0.025
IDR

0.02 R=4
G

R=3
0.015
R=2
0.01
R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
0.04

0.035

0.03

0.025
IDR

0.02 R=4
G

R=3
0.015
R=2
0.01
R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.9 Generalised interstorey drift ratio spectra (a) and (b) (Case 1)

137
α =3
0.04

0.035

0.03

0.025
IDR

0.02 R=4
G

0.015 R=3
R=2
0.01 R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
0.04

0.035

0.03

0.025 R=4
IDR

0.02
R=3
G

0.015
R=2
0.01 R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.10 Generalised interstorey drift ratio spectra (c) and (d) (Case 1)

138
α =5
0.04

0.035

0.03

0.025 R=4
IDR

0.02
R=3
G

0.015
R=2
0.01 R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
0.04

0.035

0.03 R=4

0.025 R=3
IDR

0.02
G

0.015 R=2

0.01 R=1.5

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.11 Generalised interstorey drift ratio spectra (e) and (f) (Case 1)

139
α =7
0.04

0.035

0.03

0.025 R=4
IDR

0.02
R=3
G

0.015
R=2
0.01
R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)

(g)
α =8
0.04

0.035
R=4
0.03

0.025 R=3
IDR

0.02
G

0.015 R=2
R=1.5
0.01

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.12 Generalised interstorey drift ratio spectra (g) and (h) (Case 1)

140
α =1
0.04

0.035

0.03

0.025
R=4
IDR

0.02
G

R=3
0.015
R=2
0.01 R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
0.04

0.035

0.03

0.025
IDR

0.02 R=4
G

R=3
0.015
R=2
0.01 R=1.5

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.13 Generalised interstorey drift ratio spectra (a) and (b) (Case 2)

141
α =3
0.04

0.035

0.03

0.025
R=4
IDR

0.02
G

R=3
0.015
R=2
0.01 R=1.5

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
0.04

0.035

0.03

0.025 R=4
IDR

0.02 R=3
G

0.015
R=2
0.01 R=1.5

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.14 Generalised interstorey drift ratio spectra (c) and (d) (Case 2)

142
α =5
0.04

0.035

0.03

0.025 R=4
IDR

0.02 R=3
G

0.015
R=2
0.01 R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
0.04

0.035
R=4
0.03

0.025 R=3
IDR

0.02
G

R=2
0.015
R=1.5
0.01

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.15 Generalised interstorey drift ratio spectra (e) and (f) (Case 2)

143
α =7
0.04

0.035

0.03

0.025 R=4
IDR

0.02 R=3
G

0.015
R=2
0.01
R=1.5
0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(g)
α =8
0.04

0.035 R=4

0.03
R=3
0.025
IDR

0.02
G

R=2
0.015
R=1.5
0.01

0.005

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.16 Generalised interstorey drift ratio spectra (g) and (h) (Case 2)

144
α =1
1

0.8

0.6
cur
G

0.4
R=4
0.2 R=3
R=2
R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
1

0.8

0.6
cur
G

0.4 R=4
R=3
0.2 R=2
R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.17 Generalised curvature spectra (a) and (b) (Case 1)

145
α =3
1

0.8

0.6
cur
G

R=4
0.4
R=3
R=2
0.2
R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
1

0.8

0.6
R=4
cur
G

R=3
0.4
R=2
R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.18 Generalised curvature spectra (c) and (d) (Case 1)

146
α =5
1

0.8

0.6 R=4
cur

R=3
G

0.4
R=2
R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
1

0.8 R=4

0.6 R=3
cur
G

R=2
0.4
R=1.5

0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.19 Generalised curvature spectra (e) and (f) (Case 1)

147
α =7
1

0.8
R=4

0.6
R=3
cur
G

0.4 R=2
R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(g)
α =8
1
R=4

0.8
R=3

0.6
cur

R=2
G

0.4 R=1.5

0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.20 Generalised curvature spectra (g) and (h) (Case 1)

148
α =1
1

0.8

0.6
cur
G

0.4
R=4

0.2 R=3
R=2
R=1.5
0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(a)
α =2
1

0.8

0.6
cur
G

0.4 R=4
R=3
0.2 R=2
R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(b)
Fig. 6.21 Generalised curvature spectra (a) and (b) (Case 2)

149
α =3
1

0.8

0.6
cur

R=4
G

0.4 R=3
R=2
0.2 R=1.5

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(c)
α =4
1

0.8

0.6 R=4
cur
G

R=3
0.4
R=2
R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(d)
Fig. 6.22 Generalised curvature spectra (c) and (d) (Case 2)

150
α =5
1

0.8

0.6 R=4
cur

R=3
G

0.4
R=2
R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(e)
α =6
1

0.8 R=4

0.6 R=3
cur
G

R=2
0.4
R=1.5

0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(f)
Fig. 6.23 Generalised curvature spectra (e) and (f) (Case 2)

151
α =7
1

0.8
R=4

0.6 R=3
cur
G

0.4 R=2

R=1.5
0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(g)
α =8
1.2

R=4
1

0.8 R=3
cur

0.6 R=2
G

0.4 R=1.5

0.2

0
1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
(h)
Fig. 6.24 Generalised curvature spectra (g) and (h) (Case 2)

152
2
1.9
Participation factor of 1st mode

1.8
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1
0 2 4 6 8 10 12 14 16 18 20
Degree of structural interaction
Fig. 6.25 Participation factor of 1st mode

10
9
Spectral acceleration (ms )
-2

8
7
6
5
4
3
2
1
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Fundamental period (s)
Fig. 6.26 Mean pseudo- acceleration spectrum with 5% viscous damping of ground motions
in Table 6.1

153
10000

8000
Base shear (kN)

6000

4000

2000

0
0 0.005 0.01 0.015 0.02 0.025 0.03
Roof drift ratio

Fig. 6.27 Capacity curve of building under inverted triangular load

1
0.9
0.8
0.7
Roof drift (m)

0.6
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8 9 10 11 12
Ground motion record

Fig. 6.28 Roof drift under response history analysis and prediction from method of
generalised inelastic response spectra

154
0.016

0.014

0.012

0.01
IDR

0.008

0.006

0.004

0.002

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Ground motion record

Fig. 6.29 Interstorey drift ratio under response history analysis and prediction from the
method of generalised inelastic response spectra

0.0016

0.0014

0.0012
Curvature (m )

0.001
-1

0.0008

0.0006

0.0004

0.0002

0
0 1 2 3 4 5 6 7 8 9 10 11 12
Ground motion record

Fig. 6.30 Curvature under response history analysis and prediction from method of
generalised inelastic response spectra

155
Chapter 7 

Conclusion 

7.1 Conclusion

The principal objectives of the research into the development of quick and simple, yet
accurate, performance-based methods at the stage of the preliminary seismic vulnerability
assessment, where a large number of buildings has to be evaluated in a short period of time,
have been attained. On the basis of the theoretical studies combining with numerical
investigations, the following conclusions can be drawn.

(1) A simple and effective, yet accurate model is presented for quick estimation of the natural
fundamental period of vibration for tall building structures. It is derived based on the
continuum technique, where a tall-building structure is considered as a continuous,
interactive flexural-shear cantilever. The proposed model is presented in a closed-form
mathematical expression for calculating the natural fundamental period T1, in which the
effects of the interaction of different structural forms in the building, design roof drift ratio
and maximum interstorey drift ratio, intensities of applied loading and height of the
structure on the fundamental period of vibration have been considered.
156
(2) An inelastic interstorey drift ratio model is proposed, with the use of graphical
representation for the preliminary seismic assessment of high-rise buildings that may
deform inelastically. The inelastic interstorey drift ratio model is derived from the
principle of capacity spectrum analysis and modified modal pushover analysis for the
assessed buildings. Numerical investigations on representative tall wall-frame structures,
which are designed against gravity and lateral loads, show that the seismic responses of the
buildings predicted by the proposed inelastic interstorey drift model agree well with those
obtained from response history analysis for a given set of ground motions. The proposed
model is particularly suitable for assessing the building structures that are analysed in
frequency domain.

(3) A systematic and effective procedure for preliminary seismic assessment is developed,
which serves as a tool for screening and evaluating a large number of existing buildings in
a short period of time and for identifying potential hazardous structures. The proposed
preliminary seismic assessment procedure first divides the assessed building structures
into either elastic or inelastic ones under the expected ground motion using a bifurcation
index. Buildings that behave elastically are safe, whereas buildings that may deform
inelastically are further examined by the inelastic interstorey drift model which can be
quickly implemented through a series of aided charts. Performance of the building
structure is then defined based on the interstorey drift ratio by comparing those given in
the current codes of practice and the literature.

(4) A modified continuum-MDOF model is derived for analysing seismic responses of tall
building particularly for slender building that has significantly higher vibration modes
and inelastic response ranges, which is more suitable for rapid preliminary seismic
assessment as compared to the original continuum-MDOF model because both the
required information and the effort for analysis are lesser. The numerical investigation
shows that the proposed model gives reasonably good predictions of the interstorey drift
ratio in response history analysis, as compared to the corresponding detailed model.

157
(5) A general continuum representation is derived, in which the seismic responses of two
similar continuum models are related by the height ratio for the same strength ratio in
linear static behaviour, linear dynamic behaviour and modified modal pushover analysis.
When a modified continuum-MDOF model is analysed corresponding to a particular
fundamental period and degree of structural interaction, the seismic responses of other
building structures with similar models can be approximated through the height ratio for
the same strength ratio.

(6) A spectrum-based method using the generalised inelastic response spectra is proposed and
recommended for preliminary seismic assessment of tall buildings that are relatively
slender with significant higher mode effects and are normally analysed in time domain.
The responses of a building structure can be predicted using the generalised inelastic
responses, which are derived from the modified continuum-MDOF model through the
general continuum representation. The generalised inelastic response spectra are
developed by the responses of 168 modified continuum-MDOF models, having 8 different
degrees of structural interaction and 21 different fundamental periods under nonlinear
response history analysis. The numerical investigation shows that the seismic responses
from the proposed method agree well with those from the response history analysis of the
corresponding 2-dimensional finite-element-based model. The proposed method
provided a simple and quick, yet accurate, mean of estimating the inelastic seismic
response of building structures, thus particularly being suitable for the preliminary seismic
assessment.

158
7.2 Further research

A systematic and effective procedure using the methods of inelastic interstorey drift ratio model
and generalised inelastic response spectrum has been developed for preliminary seismic
assessment. It serves as a seismic assessment tool for screening and evaluating a large number
of existing buildings in a short period of time. Therefore the principal objective of the research
has been attained.

Following the analytical development, further calibrations on the demand spectrum and
damping function may improve the accuracy of the method. Parametric study on the building
properties may simplify the analysis and increase the efficiency of the method.

The effects of torsion due to asymmetric floor plan should be further studied by
considering models with eccentricity difference in mass and stiffness centres. Adverse seismic
effect due to vertical stiffness and strength variations may be modelled. Drift multipliers that
have included the above effects could be derived either statically or analytically. They could
be implemented in the proposed inelastic interstorey drift model and generalised inelastic
response spectrum, thus that the proposed method of analysis may be applied to more variety of
situations.

159
References 
ATC 40. (1996). Seismic Evaluation and Retrofit of Concrete Buildings, Applied Technology
Council, Redwood City, California.

Chintanapakdee, C., Chopra, A. K. (2003). “Evaluation of modal pushover analysis using


generic frames,” Earthquake Engineering and Structural Dynamics, 32(3), 417-442.

Chopra, A. K., Goel, R. K. (2002). “A modal pushover analysis procedure for estimating
seismic demands for buildings,” Earthquake Engineering and Structural Dynamics, 31(3),
561-582.

Chopra, A. K., Goel, R.. K. (2004). “A modal pushover analysis procedure to estimate seismic
demands for unsymmetric-plan buildings,” Earthquake Engineering and Structural Dynamics,
33(8), 903-927.

Chopra, A. K., Goel, R. K., Chintanapakdee, C. (2004). “Evaluation of a modified MPA


procedure assuming higher modes as elastic to estimate seismic demands,” Earthquake Spectra,
20(3), 757-778.

Chandler, A. M., Correnza, J. C., Hutchinson, G. L. (1996). “Seismic torsional provisions:


influence on element energy dissipation,” Journal of Structural Engineering, 122(5), 494-500.

Computer and Structures, Inc. (2005). ETABS Version 8: Three-Dimensional Analysis of


Building Systems, Berkeley, California.

El Howary, H. A., Mehanny, S. S. F. (2011). “Seismic vulnerability evaluation of RC moment


frame buildings in moderate seismic zones,” Earthquake Engineering and Structural Dynamics,
40(2), 215-235.

Fajfar, P. (1999). “Capacity spectrum method based in inelastic demand spectra,” Earthquake
Engineering and Structural Dynamics, 28(9), 979-993.

FEMA 154. (2002). Rapid Visual Screening of Buildings for Potential Seismic Hazards: A
Handbook (second edition), Federal Emergency Management Agency, Washington, D.C.

FEMA 273. (1997). NEHRP Guidelines for the Seismic Rehabilitation of Buildings, Federal
Emergency Management Agency, Washington, D.C.

FEMA 310. (1998). Handbook for the Seismic Evaluation of Buildings, Federal Emergency
Management Agency, Washington, D.C.

FEMA 356. (2000). Prestandard and Commentary for the Seismic Rehabilitation of Buildings,
Federal Emergency Management Agency, Washington, D.C.

FEMA 440. (2005). Improvement of Nonlinear Static Seismic Analysis Procedures, Federal
Emergency Management Agency, Washington, D.C.

160
Freeman, S. A. (1998). “The capacity spectrum method as a tool for seismic design,”
Proceedings of the 11th European Conference on Earthquake Engineering, Paris, France.

Freeman, S. A., (2004). “Review of the development of the capacity spectrum method,” ISET
Journal of Engineering Technology, paper no. 438, 41(1), 1-13.

Ghobarah, A., Abou-Elfath, H., Biddah, A. (1999). “Response-based damage assessment of


structures,” Earthquake Engineering and Structural Dynamics, 28(1), 79-104.

Goel, R. K., Chopra, A. K. (1998). “Period formulas for concrete shear wall buildings,” Journal
of Structural Engineering, 124(4), 426-433.

Gulkan, P., Sozen, M. A. (1974). “Inelastic Response of Reinforced Concrete Structures to


Earthquake Motions,” ACI Journal, December, 604-610.

Gunay, M. S., Sucuoglu, H. (2010). “An improvement to linear-elastic procedures for seismic
performance assessment,” Earthquake Engineering and Structural Dynamics, 39, 907-931.

Huang, K. (2009). Continuum MDOF Model for Seismic Analysis of Wall-frame Structures,
Ph.D. dissertation, Department of Civil and Environmental Engineering, The Hong Kong
University of Science and Technology.

Huang, K., Kuang, J. S. (2010). “On the applicability of pushover analysis for seismic
evaluation of medium- and high-rise buildings.” The Structural Design of Tall and Special
Buildings, 19(5), 573-588.

Ile, N., Reynouard, J. M. (2000). “Nonlinear analysis of reinforced concrete shear wall under
earthquake loading,” Journal of Earthquake Engineering, 4(2), 183-213.

Iwan, W. D. (1997). “Drift spectrum measure of demand for earthquake ground motions,”
Journal of Structural Engineering, 124(4), 397-404.

Kappos, A. J. (1997). “Seismic damage indices for RC buildings: evaluation of concepts and
procedures,” Progress in Structural Engineering and Materials, 1(1), 78–87.

Kim, T. W., Foutch, Douglas A., LaFave, James M. (2005). “A practical model for seismic
analysis of reinforced concrete shear wall buildings,” Journal of Earthquake Engineering, 9(3),
393-417.

Koru, Z. B. (2002). Seismic Vulnerability Assessment of Low-rise Reinforced Concrete


buildings, Ph.D. dissertation, Purdue University.

Krawinkler, H., Seneviratna, GDPK. (1998). “Pros and cons of a pushover analysis of seismic
performance evaluation,” Engineering Structures, 20(4-6), 452-464.

Kuang, J. S., Huang, K. (2011). “Simplified multi-degree-of-freedom model for estimation of


seismic response of regular wall-frame structures,” The Structural Design of Tall and Special
Buildings, 20(3), 418-432.

161
Lang, K. (2002). Seismic Vulnerability of Existing Building, IBK Report No. 273, Swiss
Federal, Institute of Technology (ETH), Zurich, Switzerland.

Li, C. S., Lam, S. S., Zhang, M. Z. (2006). “Shaking table test of a 1:20 scale high-rise
building with a transfer plate system,” Journal of Structural Engineering, 132(11),
1732-1744.

Lin, Y. Y., Miranda, E. (2008). “Noniterative equivalent linear method for evaluation of
existing structures,” Journal of Structural Engineering, 134(11), 1685-1695.

Mazzoni S., McKenna F., Scott M. H., Fenves G. L. (2006). OpenSees Manual,
http://opensees.berkeley.edu/OpenSees/manuals/usermanual/ (last visited in August 2010).

Michael, J. S., Andrei, M. R., Young, J. P. (1989). “Seismic damageability assessment of R/C
buildings in Eastern U.S,” Journal of Structural Engineering, 115(9), 2184-2203.

Miranda, E. (2000). “Inelastic displacement ratios for structures on firm sites,” Journal of
Structural Engineering, ASCE, 126(10), 1150-1159.

Miranda, E., Ruiz, G. J. (2002). “Evaluation of approximate methods to estimate maximum


inelastic displacement demands,” Earthquake Engineering and Structural Dynamics, 31(3),
539-560.

Miranda, E., Reyes, C. J. (2002). “Approximate lateral drift demands in multistorey buildings
with nonuniform stiffness,” Journal of Structural Engineering, ASCE, 128(7), 840-849.

Miranda, E., Akkar, S. D. (2006). “Generalised interstorey drift spectrum,” Journal of


Structural Engineering, ASCE, 132(6), 840-852.

Mwafy, A. M., Elnashai, A. S. (2001). “Static Pushover versus dynamic collapse analysis of RC
Buildings,” Engineering structures, 23(5), 407-424.

Pacific Earthquake Engineering Research Centre. (2000). PEER Strong Motion Database,
http://peer.berkeley.edu/smcat/ (last visited in August 2010).

Park, Y. J., Ang, AHS. (1985). “Mechanistic seismic damage model for reinforced concrete,”
Journal of Structural Engineering, 111(4), 722-739.

Priestley, M. J. N., Kowalsky, M. J. (1998). “Aspects of drift and ductility capacity of


rectangular cantilever structural walls,” Bulletin of the New Zealand National Society for
Earthquake Engineering, 31(2), 73-85.

Rakesh, Goel, R. K., Chopra, A. K. (1997). “Period formulas for moment-resisting frame
buildings,” Journal of Structural Engineering, 123(11), 1454-1461.

Rutenberg, A. (1975). “Approximate natural frequencies for coupled shear walls,” Earthquake
Engineering & Structural Dynamics, 4(1), 95-100.

162
Sang, W. H., Chopra, A. K. (2006). “Approximate incremental dynamic analysis using the
modal pushover analysis procedure,” Earthquake Engineering and Structural Dynamics,
35(15), 1853-1873.

SEAOC (2009). SEAOC Blue Book: Seismic Design Recommendations, Structural Engineers
Association of California.

Sheikh, M. N. (2005). Seismic Assessment of Buildings in Hong Kong with Special Emphasis
on Displacement Based Approaches, Ph.D. dissertation, Department of Civil Engineering, The
University of Hong Kong, Hong Kong.

Smith, S. B., Coull, A. (1991). Tall Building Structures: Analysis and Design, Wiley, New York.

Stephens, J. E., Yao, J. T. P. (1987). “Damage assessment using response measurements,”


Journal of Structural Engineering, 113(4), 787-801.

Thermou, G. E., Pantazopoulou, S. J. (2011). “Assessment indices for the seismic vulnerability
of existing R.C. buildings,” Earthquake Engineering and Structural Dynamics, 40(3), 293-313.

Tsang, H. H., Ray Su, K. K., Nelson Lam, T. K., Lo, S. H. (2009). “Rapid assessment of
seismic demand in existing building structures,” The Structural Design of Tall and Special
Buildings, 18, 427-439.

Vamvatsikos, D., Cornell, C. A. (2002). “Incremental Dynamic Analysis,” Earthquake


Engineering and Structural Dynamics, 31(3), 491-514.

Vamvatsikos, D., Cornell, C. A. (2005). “Direct estimation of the seismic demand and capacity
of multidegree-of-freedom systems through incremental dynamic analysis of single-degree-of
freedom approximation,” Journal of Structural Engineering, 131(4), 589-599.

Vamvatsikos, D., Cornell, C. A. (2006). “Direct estimation of the seismic demand and capacity
of oscillators with multi-linear static pushovers through incremental dynamic analysis,”
Earthquake Engineering and Structural Dynamics, 35(9), 1097-1117.

Wang, Y. P., Reinhorn, A. M., Soong, T. T. (1992). “Development of design spectra for actively
controlled wall frame buildings,” Journal of Engineering Mechanics, ASCE, 118(6),
1201-1220.

Williams, M. S., Sexsmith, R. G., (1995). “Seismic damage indices for concrete structures: a
state of the art review,” Earthquake Spectra, 11(2), 319-349.

Yoon, Y. S., Smith, S. B. (1995)(a). “Estimating period ratio for predicting torsional coupling,”
Engineering Structure, 17(1), 52-56.

Yoon, Y. S., Smith, S. B. (1995)(b). “Assessment of translational-torsional coupling in


asymmetric uniform wall-frame structures,” Journal of Structural Engineering, 121(10),
1488-1496.

163
Appendices
A.1 Continuum flexural-shear cantilever
q

EI GA EI GA

•• •• ••
ug ug ug

Flexural element Shear element

∂M
M + dx
∂x

∂V f ∂Vs
Vf + dx Vs + dx
∂x ∂x
χ
dx ∂u
2
ndx ∂ 2u
m f dx 2 ms dx 2
∂t ∂t
Vf Vs

Fig. A1 Free body diagram of flexural-shear cantilever at time t under free vibration

164
A free-body diagram of the flexural-shear cantilever at time t under free vibration with the
mass replaced by its inertia force is shown in Fig. A1. Consider the dynamic equilibrium of
both flexural and shear elements in the flexural-shear cantilever, and two equilibrium
equations in horizontal direction can be developed.

For the flexural element,


∂ 2u ∂V
m f dx + V f + f dx − V f + ndx = 0 (1)
∂t 2
∂x

and for the shear element,


∂ 2u ∂V
ms dx + Vs + s dx − Vs − ndx = 0 (2)
∂t 2
∂x
where mf and ms are the distributed masses of flexural and shear elements respectively, u is the
relative lateral displacement, n is the stress of the axially rigid continuum, and Vf and Vs are
shear forces of the flexural and shear elements respectively and given by
∂M ∂ ⎛ ∂ 2u ⎞ (3)
Vf = = ⎜ EI ⎟
∂x ∂x ⎝ ∂x 2 ⎠

∂u
Vs = −GAχ = −GA (4)
∂x
in which M is the bending moment and χ is the shear strain. By combining the equations of
equilibrium of the elements presented by Eq. (1) and Eq. (2) and then substituting the shear
forces given by Eqs (3) and (4), the governing equation of motion of the continuum
flexural-shear model under free vibration can be derived and simplified as
∂ 2u ∂ 4u ∂ 2u
m + EI − GA =0 (6)
∂t 2 ∂x 4 ∂x 2
where m = mf + ms is the distributed mass.

Let the relative lateral displacement be


u( x, t ) = φ ( x ) q ( t ) (7)

where φ(x) is the shape function with respected to coordinate x along the height and q(t) is the
modal coordinate, with different independent variables. Eq. (6) can be solved by the technique
of separation of variables and becomes

165
EI (4 ) GA ''
φ (x ) − φ (x )
− q (t ) m
''
m
= (8)
q(t ) φ (x )
Functions at the right and left hand side have augment independent of each another. Since
equation (8) holds for any x and t, equation (8) is forced to equal to a constant, say ω 2 . Thus
Eq. (8) is decomposed into two equations with variables x and t respectively,
q′′ ( t ) + ω 2 q ( t ) = 0 (9)

EI ( 4) GA
φ ( x) − φ ′′ ( x ) − ω 2φ ( x ) = 0 (10)
m m
where ω is the natural vibration frequency of the continuum model. To solve for the shape
function, let φ ( x ) = Ae ax and substitute it into equation (10). A quadratic equation of a2 is
formed as follows and the values of a2 are solved.
GA 2 mω 2
a4 − a − =0 (11)
EI EI
where a12 and a22 are the roots of Eq. (11)
2
⎛ GA ⎞ 4mω
2
GA
− ⎜ ⎟ +
EI ⎝ EI ⎠ EI
a12 = or (12)
2
2
⎛ GA ⎞ 4mω
2
GA
+ ⎜ ⎟ +
EI ⎝ EI ⎠ EI
a22 = (13)
2
From Eqs (12) and (13), four solutions of a can be obtained a = ±iγ o and a = ± β o , where

2
GA ⎛ GA ⎞ 4mω 2
− + ⎜ ⎟ +
EI ⎝ EI ⎠ EI
γo = (14)
2
2
GA ⎛ GA ⎞ 4mH 4ω 2
+ ⎜ ⎟ +
EI ⎝ EI ⎠ EI
βo = (15)
2
Therefore the shape function is expressed as
φ (x ) = A1e β x + A2e − β x A3eiγ x + A4e −iγ
o o o ox
(16)

From Euler’s Formula, Eq. (16) is converted into trigonometric functions as follows
φ (x ) = C1 cos γ o x + C2 sin γ o x + C3 cosh β ox + C4 sinh β o x (17)

166
The derivatives of the shape function with respected to x are
φ ′(x ) = −C1γ o sin γ o x + C2γ o cos γ o x + C3 β o sinh β o x + C4 β o cosh β o x (18)

φ '' ( x ) = −C1γ o2 cos γ o x − C2γ o2 sin γ o x + C3βo2 cosh βo x + C4 βo2 sinh βo x (19)

φ (3) (x ) = C1γ o3 sin γ o x − C2γ o3 cos γ o x + C3 β o3 sinh β o x + C4 β o3 cosh β o x (20)

The shape function in Eq. (17) has four unknown constants Ci , which have to be solved from

the boundary conditions of the model.


1. Zero displacement at the fixed support
u (0, t ) = 0
φ (0)q(t ) = 0
φ (0) = 0
C1 cos 0 + C2 sin 0 + C3 cosh 0 + C4 sinh 0 = 0
C1 + C3 = 0

Therefore C1 = −C3 (21)

2. Zero slope at the fixed support

u' (0, t ) = 0
φ ' (0)q(t ) = 0
φ' (0) = 0
− C1γ o sin0 + C2γ o cos0 + C3βo sinh0 + C4βo cosh0 = 0
γ oC2 + βoC4 = 0
βo
Therefore C2 = − C4 (22)
γo
3. Zero moment at the roof
M (H ) = 0
∂ 2u ( H , t )
EI =0
∂x 2
φ '' (H )q(t ) = 0
φ ' ' (H ) = 0
− C1γ o2 cos γ o H − C2γ o2 sin γ o H + C3 β o2 cosh β o H + C4 β o2 sinh β o H = 0
Substituting Eq. (21) and (22) into the above equation will obtain the follows
C3γ o2 cos γ o H + C4 β oγ o2 sin γ o H + C3 β o2 cosh β o H + C4 β o2 sinh β o H = 0
( ) (
C3 γ o2 cos γ o H + β o2 cosh β o H + C4 β oγ o sin γ o H + β o2 sinh β o H = 0 )

167
C4 (β oγ o sin γ o H + β o2 sinh β o H )
Therefore C3 = − (23)
γ o2 cos γ o H + β o2 cosh β o H
From Eqs (21), (22) and (23), it can be observed that the constants C1, C2 and C3 depend on C4
which can be arbitrary values. The magnitude of the shape function hence can be selected as
any arbitrary value C4. The shape function is
φ (x ) = C1 cos γ o x + C2 sin γ o x + C3 cosh β ox + C4 sinh β o x
where
C4 (β oγ o sin γ o H + β o2 sinh β o H )
C1 =
γ o2 cos γ o H + β o2 cosh β o H
βo
C2 = − C4
γo
C4 (β oγ o sin γ o H + β o2 sinh β o H )
C3 = −
γ o2 cos γ o H + β o2 cosh β o H
4. Zero resultant shear force at the top,
∂ 3u (H , t ) ∂u (H , t )
EI − GA =0
∂x 3
∂x
EIφ (3 ) (H ) − GAφ ' (H ) = 0
GA '
φ (3 ) ( H ) − φ (H ) = 0
EI
C1γ o3 sin γ o H − C2γ o3 cos γ o H
+ C3 β o3 sinh β o H + C4 β o3 cosh β o H
GA ⎛ − C1γ o sin γ o H + C2γ o cos γ o H ⎞
− ⎜⎜ ⎟=0
EI ⎝ + C3 β o sinh β o H + C4 β o cosh β o H ⎟⎠

⎛ GA ⎞
C1 ⎜ γ o3 sin γ o H + γ o sin γ o H ⎟
⎝ EI ⎠
⎛ GA ⎞
− C2 ⎜ γ o3 cos γ o H + γ o cos γ o H ⎟
⎝ EI ⎠
⎛ GA ⎞
+ C3 ⎜ β o3 sinh β o H − β o sinh β o H ⎟
⎝ EI ⎠
⎛ GA ⎞
+ C4 ⎜ β o3 cosh β o H − β o cosh β o H ⎟ = 0
⎝ EI ⎠
Substituting Eqs (21), (22) and (23) into the above equations lead to,

168
(
C4 β oγ o sin γ o H + β o2 sinh β o H ⎛ 3 )
⎜ γ o sin γ o H +
GA ⎞
γ o sin γ o H ⎟
γ o cos γ o H + β o cosh β o H ⎝
2 2
EI ⎠
β ⎛ GA ⎞
+ o C4 ⎜ γ o3 cos γ o H + γ o cos γ o H ⎟
γo ⎝ EI ⎠


(
C4 β oγ o sin γ o H + β o2 sinh β o H ⎛ 3 )
⎜ β o sinh β o H −
GA ⎞
β o sinh β o H ⎟
γ o cos γ o H + β o cosh β o H ⎝
2 2
EI ⎠
⎛ GA ⎞
+ C4 ⎜ β o3 cosh β o H − β o cosh β o H ⎟ = 0
⎝ EI ⎠
GA 2
By assigning α 2 = H , β = β o H , γ = γ o H , C4 = 1 , the above equation is simplified,
EI
βγ sin γ + β 2 sinh β 3
γ 2 cos γ + β 2 cosh β
(γ sin γ + α 2γ sin γ )
(
+ βγ 2 cos γ + α 2 β cos γ )
βγ sin γ + β 2 sinh β 3

γ 2 cos γ + β 2 cosh β
(β sinh β − α 2 β sinh β )
(
+ β 3 cosh β − α 2 β cosh β = 0 )
βγ sin γ + β 2 sinh β 3
γ cos γ + β cosh β
2 2
(γ sin γ + α 2γ sin γ )

(
+ βγ 2 cos γ + α 2 β cos γ )
βγ sin γ + β 2 sinh β 3

γ 2 cos γ + β 2 cosh β
(β sinh β − α 2 β sinh β )
(
+ β 3 cosh β − α 2 β cosh β = 0 )
β 3 cos γ + βγ 2 cosh β
βγ sin γ + β 2 sinh β 2
+
γ 2 cos γ + β 2 cosh β
(β γ sin γ − βγ 2 sinh β ) = 0
β 2γ 2 (cos 2 γ + sin 2 γ ) + (β 4 + γ 4 )cos γ cosh β
( ) ( )
+ β 2γ 2 cosh 2 β − sinh 2 β + βγ β 2 − γ 2 sin γ sinh β = 0

(( )2
)
2 β 2γ 2 + β 2 − γ 2 + 2 β 2γ 2 cos γ cosh β + βγα 2 sin γ sinh β = 0

⎛ α4 ⎞ α2
Therefore 2 + ⎜⎜ 2 + 2 2 ⎟⎟ cos γ cosh β + sin γ sinh β = 0 (24)
⎝ β γ ⎠ βγ

Since β 2 = α 2 + γ 2 , Eq. (24) can be expressed in tern of γ and is given by,

⎛ α4 ⎞ α2
2 + ⎜⎜ 2 + 2 ⎟ cos γ cosh α 2
+ γ 2
+ sin γ sinh α 2 + γ 2 = 0 (25)
⎝ ( α +γ γ ⎠
2 2 ⎟
) γ α +γ
2 2

where

169
GA
α= H
EI
4mH 4ω 2
−α + α +
2 4

γ = EI
2
γ is a function of the natural frequency ω , so Eq. (25) is actually the governing equation of
the angular frequencies of the model.

170
A.2 Model under uniform lateral loading

n
w

EI GA EI GA

Flexural element Shear element

∂M
M+ dx
∂x

∂V f ∂Vs
Vf + dx Vs + dx
∂x ∂x
χ
w dx ndx

Vf
Vs

Fig. A2 Free body diagram of flexural-shear cantilever

171
A free-body diagram of the flexural-shear cantilever under uniform lateral load is shown in
Fig. A2. Consider the equilibrium of both flexural and shear elements in the flexural-shear
cantilever, and two equilibrium equations in horizontal direction can be developed.
For the flexural element,
∂V f
Vf + dx − V f + ndx − wdx = 0 (26)
∂x
and for the shear element,
∂Vs
Vs + dx − Vs − ndx = 0 (27)
∂x
By combining the equations of equilibrium of the elements presented by Eqs (26) and (27) and
then substituting the shear forces given by Eqs (3) and (4), the governing equation of deflection
of the continuum flexural-shear model under uniform lateral load can be derived and simplified
as
∂4 y ∂2 y
EI 4 − GA 2 = w (28)
∂x ∂x
GA 2
By assigning α 2 = H , Eq. (28) is expressed as
EI
2
(4 ) ⎛α ⎞ w
y − ⎜ ⎟ y '' = (29)
⎝H⎠ EI
Eq. (29) is an ordinary differential equation and the corresponding homogeneous solution can
be solved by letting yh = Aesx in Eq. (29)
2
⎛α ⎞
As e − ⎜ ⎟ As 2e sx = 0
4 sx

⎝H⎠
2
⎛α ⎞
s −⎜ ⎟ = 0
2

⎝H ⎠
α
s=±
H
Therefore the homogenous solution is
αx αx
− αx αx
yh = A1e H + A2e H
= C3 cosh + C4 sinh (30)
H H
The particular solution of Eq. (29) can be solved by considering,
wH 2
y =−''
p
EIα 2
wH 2 x
y 'p = − + C2
EIα 2
172
Therefore the particular solution is

wH 2 x 2
yp = − + C2 x + C1 (31)
2 EIα 2
Then the general solution to Eq. (29) is obtained from summation of Eqs (30) and (31)
αx αx wH 2 x 2
y = C1 + C2 x + C3 cosh + C4 sinh − (32)
H H 2 EIα 2
The derivatives of the deflection with respected to x are
⎛α ⎞ αx ⎛α ⎞ αx wH 2 x
y′( x ) = C2 + C3 ⎜ ⎟ sinh + C4 ⎜ ⎟ cosh − (33)
⎝H⎠ H ⎝H⎠ H EIα 2
2 2
⎛α ⎞ αx ⎛α ⎞ αx wH 2
y′′(x ) = C3 ⎜ ⎟ sinh + C4 ⎜ ⎟ cosh − (34)
⎝H⎠ H ⎝H⎠ H EIα 2
3 3
⎛α ⎞ αx ⎛α ⎞ αx
y′′′( x ) = C3 ⎜ ⎟ cosh + C4 ⎜ ⎟ sinh (35)
⎝H⎠ H ⎝H⎠ H
The unknown constants Ci in Eq. (32) are solved from the boundary conditions of the model.

1. Zero displacement at the fixed support,


y (0) = 0
C1 + C3 = 0

Therefore C1 = −C3 (36)

2. Zero rotation at the fixed support,


y′(0 ) = 0
⎛α ⎞ ⎛α ⎞
C2 + C3 ⎜ ⎟ sinh 0 + C4 ⎜ ⎟ cosh 0 = 0
⎝H ⎠ ⎝H⎠
C2 + C4α = 0

⎛α ⎞
Therefore C2 = −C4 ⎜ ⎟ (37)
⎝H⎠
3. Zero moment at the roof,
M (H ) = 0
EIy′′(H ) = 0
2 2
⎛α ⎞ ⎛α ⎞ wH 2
C3 ⎜ ⎟ cosh α + C4 ⎜ ⎟ sinh α − =0
⎝H⎠ ⎝H⎠ EIα 2

wH 4 C sinh α
Therefore, C3 = − 4 (38)
EIα cosh α
4
cosh α

173
4. Zero resultant shear force at the roof,
V f (H ) + Vs (H ) = 0
EIy′′′(H ) − GAy′(H ) = 0
2
⎛α ⎞
y′′′(H ) − ⎜ ⎟ y′(H ) = 0
⎝H⎠
3 3
⎛α ⎞ ⎛α ⎞
C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α
⎝H⎠ ⎝H⎠
2
⎛α ⎞ ⎛ ⎛α ⎞ ⎛α ⎞ wH 3 ⎞
−⎜ ⎟ ⎜⎜ C2 + C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α − ⎟=0
⎝H⎠ ⎝ ⎝H⎠ ⎝H⎠ 2 EIα 2 ⎟⎠

Substituting Eqs (36), (38) and (39) into the above equation gives,
3
⎛ α ⎞ wH
C4 ⎜ ⎟ + =0
⎝H⎠ EI

wH 4
Therefore C4 = − (40)
EIα 3
The coefficients Ci in Eq. (32) are given by
wH 4 wH 4 sinh α
C1 = − −
EIα 4 cosh α EIα 3 cosh α
wH 3
C2 =
EIα 2
wH 4 wH 4 sinh α
C3 = +
EIα 4 cosh α EIα 3 cosh α
wH 4
C4 = −
EIα 3
Substituting the above coefficients into equation (32), leads to
⎡ wH 4 wH 4 sinh α wH 3 x ⎤
⎢ − − + ⎥
⎢ EIα 4
cosh α EIα 3
cosh α EIα 2 ⎥
⎢ ⎛ wH 4 wH 4 sinh α ⎞ αx ⎥
y (x ) = ⎢+ ⎜⎜ + ⎟⎟ cosh ⎥
⎢ ⎝ EIα cosh α EIα cosh α ⎠
4 3
H⎥
⎢ wH 4 αx wH 2 x 2 ⎥
⎢− sinh − ⎥
⎣ EIα 3 H 2 EIα 2 ⎦
⎡ 1 α sinh α α 2 x ⎤
4 ⎢
− − + ⎥
wH ⎢ cosh α cosh α H ⎥
=
EIα 4 ⎢ ⎛ 1 α sinh α ⎞ αx αx α 2 x 2 ⎥
⎢+ ⎜ cosh α + cosh α ⎟ cosh H − α sinh H − 2 H 2 ⎥
⎣ ⎝ ⎠ ⎦

174
Therefore the equation of deflection of the model under uniform loading is

wH 4 ⎡α sinh α + 1 ⎛ αx ⎞ αx 2⎛ x x 2 ⎞⎤
y(x ) = ⎢ ⎜ cosh − 1⎟ − α sinh + α ⎜ H 2 H 2 ⎟⎟⎥
⎜ − (41)
EIα 4 ⎣ cosh α ⎝ H ⎠ H ⎝ ⎠⎦
The roof displacement is then given by from Eq. (41)
wH 4 ⎡ α 2 α sinh α + 1⎤
y (H ) = ⎢1 + − ⎥ (42)
EIα 4 ⎣ 2 cosh α ⎦

The roof drift ratio from Eq. (42) is


y (H ) wH 3 ⎡ α 2 α sinh α + 1⎤
RΔ = = ⎢1 + − ⎥ (43)
H EIα 4 ⎣ 2 cosh α ⎦

The interstorey drift ratio from Eq. (41) is


wH 3 ⎡α sinh α + 1 αx αx ⎛ x ⎞⎤
Rδ ( x ) = y′( x ) = 3 ⎢
sinh − α cosh + α ⎜1 − ⎟ ⎥ (44)
EIα ⎣ cosh α H H ⎝ H ⎠⎦
The curvature from Eq. (42) is
wH 2 ⎡ α sinh α + 1 αx αx ⎤
cur ( x ) = 2 ⎢
cosh − α sinh − 1⎥ (45)
EIα ⎣ cosh α H H ⎦
The moment in flexural cantilever is
wH 2 ⎡ α sinh α + 1 αx αx ⎤
M ( x ) = EIy (x ) =
''
2 ⎢
cosh − α sinh − 1⎥ (46)
α ⎣ cosh α H H ⎦
The shear in flexural cantilever is
wH ⎡α sinh α + 1 αx αx ⎤
V f ( x ) = EIy ''' ( x ) = ⎢ sinh − α cosh ⎥ (47)
α ⎣ cosh α H H⎦
The shear in shear cantilever is
wH ⎡α sinh α + 1 αx αx αx ⎤
VS (x ) = GAy ' (x ) = ⎢ sinh − α cosh + α − ⎥ (48)
α ⎣ cosh α H H H⎦

175
A.3 Model under inverted triangular lateral loading

q
w

EI GA EI GA

Fig. A3 Flexural-shear cantilever under inverted triangular loading

Similar to the derivation of Appendix A.2 Eq. (29), the governing equation of deflection is
2
(4 ) ⎛α ⎞ x w
y − ⎜ ⎟ y '' = (49)
⎝H⎠ H EI
The particular solution of Eq. (49) can be solved by considering the follows,
wHx
y′p′ = −
EIα 2
wHx 2

yp = − + C2
2 EIα 2
Therefore the particular solution is
wHx 3
yp = − + C 2 x + C1 (50)
6 EIα 2
Then the general solution to Eq. (49) is obtained from summation of Eqs (30) and (50)
wHx 3
y = C1 + C 2 x + C 3 cosh αx + C 4 sinh αx − (51)
6 EIα 2
The derivatives of the deflection with respected to x are
⎛α ⎞ ⎛α ⎞ wHx 2
y ′( x ) = C2 + C3 ⎜ ⎟ sinh αx + C4 ⎜ ⎟ cosh αx − (52)
⎝H⎠ ⎝H ⎠ 2 EIα 2

176
2 2
⎛α ⎞ αx ⎛α ⎞ αx wHx
y ( x ) = C3 ⎜ ⎟ cosh + C4 ⎜ ⎟ sinh
''
− (53)
⎝H⎠ H ⎝H⎠ H EIα 2
3 3
⎛α ⎞ αx ⎛α ⎞ αx wH
y′′′( x ) = C3 ⎜ ⎟ sinh + C4 ⎜ ⎟ cosh − (54)
⎝H⎠ H ⎝H ⎠ H EIα 2
The unknown constants Ci in Eq. (32) are solved from the boundary conditions of the model.
1. Zero displacement at the fixed support,
y (0) = 0
C1 + C3 = 0

Therefore C1 = −C3 (55)


2. Zero rotation at the fixed support,
y′(0 ) = 0
⎛α ⎞ ⎛α ⎞
C2 + C3 ⎜ ⎟ sinh 0 + C4 ⎜ ⎟ cosh 0 = 0
⎝H ⎠ ⎝H⎠
C2 + C4α = 0

⎛α ⎞
Therefore C2 = −C4 ⎜ ⎟ (56)
⎝H⎠
3. Zero moment at the roof
M (H ) = 0
EIy′′(H ) = 0
2 2
⎛α ⎞ ⎛α ⎞ wH 2
C3 ⎜ ⎟ cosh α + C4 ⎜ ⎟ sinh α − =0
⎝H⎠ ⎝H⎠ EIα 2

wH 4 C sinh α
Therefore C3 = − 4 (57)
EIα cosh α
4
cosh α
4. Zero resultant shear force at the roof
V f (H ) + Vs (H ) = 0
EIy′′′(H ) − GAy′(H ) = 0
2
⎛α ⎞
y′′′(H ) − ⎜ ⎟ y′(H ) = 0
⎝H⎠
3 3
⎛α ⎞ ⎛α ⎞ wH
C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α −
⎝H⎠ ⎝H⎠ EIα 2
2
⎛α ⎞ ⎛ ⎛α ⎞ ⎛α ⎞ wH 3 ⎞
−⎜ ⎟ ⎜⎜ C2 + C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α − ⎟=0
⎝H⎠ ⎝ ⎝H⎠ ⎝H⎠ 2 EIα 2 ⎟⎠

177
Substituting Eqs (55), (56) and (57) into the above equation will obtain the follows,
3
⎛ α ⎞ wH wH
C4 ⎜ ⎟ + − =0
⎝ H ⎠ 2 EI EIα
2

wH 4 wH 4
Therefore C4 = − + (58)
2 EIα 3 EIα 5
The coefficients Ci in Eq. (51) are given by
wH 4 wH 4 sinh α wH 4 sinh α
C1 = − − +
EIα 4 cosh α 2 EIα 3 cosh α EIα 5 cosh α
wH 3 wH 3
C2 = −
2 EIα 2 EIα 4
wH 4 wH 4 sinh α wH 4 sinh α
C3 = + −
EIα 4 cosh α 2 EIα 3 cosh α EIα 5 cosh α
wH 4 wH 4
C4 = − +
2 EIα 3 EIα 5
Substituting the above coefficients into equation (51), leads to
⎡⎛ α sinh α + 2 sinh α ⎞⎛ αx ⎞ ⎤
⎢ ⎜ − ⎟⎜ cosh ⎟ ⎥
wH 4 ⎢⎝ 2 cosh α α cosh α ⎠⎝ H⎠ ⎥
y(x ) = (59)
EIα 4 ⎢ ⎛ 1 α ⎞ ⎞ x⎥
3
αx α ⎛ x ⎞ ⎛ α
2 2
⎢+ ⎜ − ⎟ sinh − ⎜ ⎟ + ⎜⎜ − 1⎟⎟ ⎥
⎢⎣ ⎝ α 2 ⎠ H 6 ⎝H⎠ ⎝ 2 ⎠ H ⎥⎦
The roof displacement is then given by from Eq. (59),
wH 4 ⎡ sinh α α 2 α sinh α + 2 ⎤
y (H ) = ⎢ + − ⎥ (60)
EIα 4 ⎣ α cosh α 3 2 cosh α ⎦

The roof drift ratio from Eq. (59) is


y (H ) wH 3 ⎡ sinh α α 2 α sinh α + 2 ⎤
RΔ = = ⎢ + − ⎥ (61)
H EIα 4 ⎣ α cosh α 3 2 cosh α ⎦

The interstorey drift ratio from Eq. (59) is


⎡⎛ α sinh α + 2 sinh α ⎞ αx ⎤
⎜ − ⎟ sinh
3 ⎢ ⎥
wH ⎢⎝ 2 cosh α α cosh α ⎠ H
Rδ ( x ) = 3 ⎢
⎥ (62)
EIα αx α ⎛ x ⎞ ⎛ α 1 ⎞ ⎥
2
⎛1 α⎞
⎢+ ⎜ − ⎟ cosh − ⎜ ⎟ + ⎜ − ⎟⎥
⎢⎣ ⎝ α 2 ⎠ H 2 ⎝ H ⎠ ⎝ 2 α ⎠⎥⎦
The curvature from Eq. (59) is
⎡⎛ α sinh α + 2 sinh α ⎞ αx ⎤
2 ⎢⎜ 2 cosh α − α cosh α ⎟ cosh H ⎥
wH ⎢⎝ ⎠
cur ( x ) = ⎥ (63)
EIα 2
⎢ ⎛1 α⎞ αx x ⎥
⎢+ ⎜ − ⎟ sinh − ⎥
⎣ ⎝α 2 ⎠ H H ⎦

178
The moment in flexural cantilever is
⎡⎛ α sinh α + 2 sinh α ⎞ αx ⎤
2⎢⎜ 2 cosh α − α cosh α ⎟ cosh H ⎥
wH ⎝ ⎠
M ( x ) = EIy′′( x ) = 2 ⎢ ⎥ (64)
α ⎢ ⎛1 α⎞ αx x ⎥
⎢+ ⎜ − ⎟ sinh − ⎥
⎣ ⎝α 2 ⎠ H H ⎦
The shear in flexural cantilever is
⎡⎛ α sinh α + 2 sinh α ⎞ αx ⎤
⎢⎜ 2 cosh α − α cosh α ⎟ sinh H ⎥
wH ⎢⎝ ⎠
V f (x ) = EIy′′′(x ) = ⎥ (65)
α ⎢ ⎛1 α⎞ αx 1 ⎥
⎢+ ⎜ − ⎟ cosh − ⎥
⎣ ⎝α 2 ⎠ H α ⎦
The shear in shear cantilever is
⎡⎛ α sinh α + 2 sinh α ⎞ αx ⎤
⎢⎜ 2 cosh α − α cosh α ⎟ sinh H ⎥
wH ⎢⎝ ⎠
Vs ( x ) = GAy′( x ) = ⎥ (66)
α ⎢ ⎛1 α⎞ αx α ⎛ x ⎞ ⎛ α 1 ⎞⎥
2

⎢+ ⎜ − ⎟ cosh − ⎜ ⎟ + ⎜ − ⎟⎥
⎢⎣ ⎝ α 2 ⎠ H 2 ⎝ H ⎠ ⎝ 2 α ⎠⎥⎦

179
A.4 Model under point load at the roof

P q

EI GA EI GA

Fig. A4 Flexural-shear cantilever under point load at the roof

Similar to the derivation of Appendix A.2 Eq. (29), the governing equation of deflection is
2
⎛α ⎞
y (4 ) − ⎜ ⎟ y '' = 0 (67)
⎝H⎠
The particular solution of Eq. (49) can be solved by considering,
y ′p′ = 0
y ′p = C 2
Therefore the particular solution is
y p = C2 x + C1 (68)
Then the general solution to Eq. (49) is obtained from summation of Eqs (30) and (50),
αx αx
y = C1 + C2 x + C3 cosh + C4 sinh (69)
H H
The derivatives of the deflection with respected to x are
α αx α αx
y′(x ) = C2 + C3 sinh + C4 cosh (70)
H H H H
2 2
⎛α ⎞ αx ⎛α ⎞ αx
y′′(x ) = C3 ⎜ ⎟ cosh + C4 ⎜ ⎟ sinh (71)
⎝H ⎠ H ⎝H⎠ H
3 3
⎛α ⎞ αx ⎛α ⎞ αx
y′′′( x ) = C3 ⎜ ⎟ sinh + C4 ⎜ ⎟ cosh (72)
⎝H⎠ H ⎝H ⎠ H

180
The unknown constants Ci in Eq. (32) are solved from the boundary conditions of the model.
1. Zero displacement at the fixed support,
y (0) = 0
C1 + C3 = 0

Therefore C1 = −C3 (73)


2. Zero rotation at the fixed support,
y′(0 ) = 0
⎛α ⎞ ⎛α ⎞
C2 + C3 ⎜ ⎟ sinh 0 + C4 ⎜ ⎟ cosh 0 = 0
⎝H ⎠ ⎝H⎠
C2 + C4α = 0

⎛α ⎞
Therefore C2 = −C4 ⎜ ⎟ (74)
⎝H⎠
5. Zero moment at the roof,
M (H ) = 0
EIy′′(H ) = 0
2 2
⎛α ⎞ ⎛α ⎞
C3 ⎜ ⎟ cosh α + C4 ⎜ ⎟ sinh α = 0
⎝H⎠ ⎝H⎠
Therefore C3 = −C4 tan α (75)
6. Zero resultant shear force at the roof,
V f (H ) + Vs (H ) = 0
EIy′′′(H ) − GAy′(H ) = 0
2
⎛α ⎞
y′′′(H ) − ⎜ ⎟ y′(H ) = 0
⎝H⎠
3 3
⎛α ⎞ ⎛α ⎞
C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α
⎝H⎠ ⎝H⎠
2
⎛α ⎞ ⎛ ⎛α ⎞ ⎛α ⎞ ⎞ P
−⎜ ⎟ ⎜⎜ C2 + C3 ⎜ ⎟ sinh α + C4 ⎜ ⎟ cosh α ⎟⎟ =
⎝H⎠ ⎝ ⎝H⎠ ⎝H⎠ ⎠ EI
Substituting Eqs (73), (74) and (75) into the above equation gives,
3
⎛α ⎞ P
C4 ⎜ ⎟ =
⎝H ⎠ EI

PH 3
Therefore C4 = (76)
EIα 3

181
The coefficients Ci in Eq. (69) are given by
PH 3 tan α
C1 =
EIα 3
PH 2
C2 = −
EIα 2
PH 3 tan α
C3 = −
EIα 3
PH 3
C4 =
EIα 3
Substituting the above coefficients into equation (51), get
PH 3 PH 2 PH 3 αx PH 3 αx
y (x ) = tanh α − x− tanh α cosh + sinh (77)
EIα 3
EIα 2
EIα 3
H EIα 3
H
The roof displacement is then given by from Eq. (77)
PH 3 PH 3
y (H ) = tanh α − (78)
EIα 3 EIα 2
The roof drift ratio from Eq. (77) is
y (H ) PH 2 PH 2
RΔ = = tanh α − (79)
H EIα 3 EIα 2
The interstorey drift ratio from Eq. (77) is
PH 2 PH 2 αx PH 2 αx
Rδ ( x ) = − − tanh α sinh + cosh (80)
EIα 2
EIα 2
H EIα 2
H
The curvature from Eq. (77) is
PH αx PH αx
cur ( x ) = − tanh α cosh + sinh (81)
EIα H EIα H
The moment in flexural cantilever is
PH ⎛ αx αx ⎞
M (x ) = EIy′′(x ) = ⎜ − tanh α cosh + sinh ⎟ (82)
α ⎝ H H⎠
The shear in flexural cantilever is
⎛ αx αx ⎞
V f (x ) = EIy′′′(x ) = p⎜ − tanh α sinh + cosh ⎟ (83)
⎝ H H⎠
The shear in shear cantilever is
⎛ αx αx ⎞
Vs ( x ) = GAy′(x ) = p⎜ − 1 − tanh α sinh + cosh ⎟ (84)
⎝ H H⎠

182

You might also like