Sunteți pe pagina 1din 460

Study of Irregular RC Buildings under Seismic effect

A
THESIS
SUBMITTED TO
DEPARTMENT OF CIVIL ENGINEERING
NATIONAL INSTITUTE OF TECHNOLOGY KURUKSHETRA
FOR THE AWARD OF DEGREE OF

DOCTOR OF PHILOSOPHY
IN
CIVIL ENGINEERING

BY

S.Varadharajan
2K 10 NITK/ 1273/10
UNDER THE SUPERVISION OF
Dr. V.K.SEHGAL
PROFESSOR
&
Dr. BABITA SAINI
ASSOCIATE PROFESSOR
DEPARTMENT OF CIVIL ENGINEERING, NATIONAL INSTITUTE OF TECHNOLOGY
KURUKSHETRA-136119, INDIA
JANUARY, 2014

CANDIDATES DECLARATION

I hereby certify that the work which is being presented in the thesis, entitled Study of
Irregular RC Buildings Under Seismic Effect for the award of the degree of Doctor of
Philosophy submitted in the department of Civil Engineering, National Institute of
Technology, Kurukshetra, is an authentic record of my own work carried out under the
supervision of Dr. V. K. Sehgal (Professor) and Dr. Babita Saini (Associate Professor),
Department of Civil Engineering, National Institute of Technology, Kurukshetra, India. The
matter presented in this thesis has not been submitted in part or in full for the award of
degree/diploma of this or any other University/Institute.

S.Varadharajan
(2K -10, NITK 1273/10)
Date

CERTIFICATE
Certified that the thesis entitled Study of irregular RC Buildings under seismic
effect submitted by S.Varadharajan in fulfillment of the requirements for the award
of the degree of Doctor of Philosophy in Civil Engineering, is the candidates own
work carried out by him under our supervision and guidance. The work presented in
the thesis has not been submitted for the award of any other degree of this or any
other University/Institute.

Supervisor:

Co-Supervisor:

DR. V.K.SEHGAL

DR. BABITA SAINI

PROFESSOR

ASSOCIATE PROFESSOR

DEPARTMENT OF CIVIL ENGINEERING


NATIONAL INSTITUTE OF TECHNOLOGY, KURUKSHETRA
(INSTITUTE OF NATIONAL IMPORTANCE)
KURUKSHETRA (HR)-136119

ii

ACKNOWLEDGEMENTS

I express my immense sense of gratitude, inner sentiments to my esteemed and


worthy guides Dr. V.K.Sehgal and Dr. Babita Saini, Department of Civil Engineering,
National Institute of Technology Kurukshetra. Their precious and prompt remarks during this
research work and thesis preparation are especially appreciated. I also thank them for their
support and motivation during this research work. I sincerely acknowledge my grateful
thanks to Director NIT Kurukshetra, Head, Faculty members, Non-teaching staff for
providing me the research facilities and work oriented environment. I would like to thank
Dr. L.M.Saini, Department of Electrical Engineering for his motivation and support during
this dissertation work.
I would also like to thank my dear friends and fellow research scholars Summit
Jalota, Anupam Bhandari and Konark Sharma for their support and encouragement
throughout this research work. Iam deeply indebted to my parents (Mr. P.K.Srinivasan and
Mrs Hemam Bujam) and to my elder brother (Mr.S.Devarajan), without their strong support
it would have been very difficult for me to undertake this research work. I would like to
acknowledge Ministry of Human Resources and Development (MHRD), Govt. of India for
providing me research fellowship and providing other necessary facilities for carrying out the
research work. Finally, I would like to acknowledge the authors and publishers who have
permitted me to use pictures from their research work.

(S.Varadharajan)

iii

ABSTRACT
A structure can be classified as irregular if it contains irregular distributions of mass, stiffness
and strength or due to irregular geometrical configurations. Different codes prescribe
different limits for these irregularities like as per IS 1893:2002, a storey in a building is said
to contain mass irregularity if its mass exceeds 200% than that of the adjacent storey. If
stiffness of a storey is less than 60% of the adjacent storey; in such a case the storey is
termed as weak storey, and if stiffness is less than 70 % of the storey above, then the
storey is termed as soft storey. In reality, many existing buildings contain irregularity due to
functional and aesthetic requirements. However, past earthquake records show the poor
seismic performance of these structures. This is due to ignorance of the irregularity aspect in
formulating the seismic design methodologies by the seismic codes (IS 1893:2002,
EC8:2004, UBC 1997, NBCC 1989, NBCC 2005 etc.).
The review of seismic design codes and reported research studies show that the
irregularity has been quantified in terms of magnitude ignoring the effect of location of
irregularity. However, location of irregularity has a significant influence on the seismic
response (Al Ali and Krawinkler 1998; Nassar and Krawinkler 1991; Das and Nau 2003).
The cracking of structural members is a realistic phenomenon during seismic excitation, and
this results in change of stiffness and strength (Paulay and Priestley 1992; Priestley 2003).
This aspect has been ignored by the previous research studies. In the present study, the
above limitations are addressed and a new index has been proposed to represent the
effects of structural irregularity and cracking. The proposed approach has been compared
with code approaches and previous research studies to evaluate the efficiency of the
proposed index.
The estimation of fundamental time period of vibration is a critical step in seismic design
and analysis of the building structures as it is a represents the global seismic demands of
the structure. The period of the building mainly depends upon building properties like mass,
stiffness, seismic excitation, storey height, number of storeys, cracking etc. In reality, the
building models often encounter different forms of structural irregularity and cracking which
result in change of strength and stiffness of the building. However, these aspects have been
ignored in code proposed empirical expressions to estimate the fundamental time period.
These expressions are unsuitable to estimate the realistic seismic demands of the structure.
Therefore, there is a necessity of modified equations to estimate the fundamental time
period. The review of previous approaches show the preference of eigen value analysis in
estimating the fundamental time period. However, this method has certain limitations and it

iv

yields over - conservative estimate of the seismic response. In addition, this method ignores
the effect of ground motion which results in unrealistic estimation of the fundamental time
period and seismic demand (Masi and Vona 2010). The present research work overcomes
these limitations to propose modified equations to estimate the fundamental time period
incorporating the aspects of structural irregularity and cracking. In contrast to previous
approaches, which represent the fundamental time period in terms of building height, the
present research works expresses fundamental time period as a function of irregularity
index. Finally, the application of the proposed fundamental time period equations in
estimation of seismic demands and in seismic vulnerability assessment has been briefly
discussed.
The estimation of seismic demands is a critical step in seismic design process. The
seismic design codes like EC8:2004 have suggested the procedures to estimate seismic
demands but the aspect of irregularity and cracking has been ignored in formulating these
procedures. In other words, the seismic demands of irregular and regular structures are
estimated using same approach. Moreover, code equations are proposed considering
SDOF system and elastic analysis which is unrealistic. Therefore, there is a need to modify
the code rules to estimate seismic demands of irregular structures. In the present study, the
inelastic seismic demands of the irregular building models are determined at first. Then a
mean relation factor irr correlating the elastic and inelastic seismic demands has been
proposed. Secondly, based on regression study conducted on results of inelastic dynamic
analysis, modified equations to estimate the seismic demands have been proposed .These
equations have been compared with the equal displacement rule (proposed by EC 8:2004)
to estimate the seismic demands. Finally, applicability of the proposed mean relation factors
and in seismic design methodologies (Performance based design and displacement based
design) has been briefly discussed.
The reinforced concrete (RC) frame structures mainly rely on beam and column elements
to resist both seismic and gravity loads. The inability of these elements to carry the load
leads to collapse of the structure. Collapse capacity of structures under seismic excitation
has always been a crucial aspect in determining seismic performance of the structure.
Accurate prediction of the collapse capacity is very important as the structural collapse is a
source of life and monetary losses. Few seismic design codes (IBC 2003) and some
reported studies (Jones and Farzin 2010; Jones 2012; Lignos and Krawinkler 2009; Kim and
Lee 2010; Kim and Choi 2011) have discussed methods to determine the collapse capacity
of the building structures. The majority of these approaches are based on SDOF system and

static method of analysis. However, these approaches ignore the effect of structural
irregularity and cracking. In the present research work, collapse capacity of the irregular
building models has been determined using inelastic dynamic analysis and simple equations
to estimate the collapse capacity and probability of collapse has been proposed in terms of
the author proposed irregularity index. Moreover, a detailed review of damage indices have
been presented and a new damage index based on inelastic seismic demands have been
proposed and compared with code approach and previous reported study (Park and Ang
1985a) to demonstrate the simplicity and efficiency of the proposed approach. Finally, the
appendix section deals with some additional results from this research work.

vi

LIST OF PUBLICATIONS
(A) INTERNATIONAL PUBLICATIONS
1.Varadharajan, S., Sehgal, V.K., and Saini, B.(2012), Seismic response of multistory
reinforced concrete frame with vertical mass and stiffness irregularities, The structural
design of tall and special buildings, Wiley publications , DOI: 10.1002/tal.1045.
2. Varadharajan, S., Sehgal, V.K., and Saini, B. (2013), Determination of Inelastic deformation
demands of RC Moment resisting setback frames, Archives of civil and mechanical
engineering, Elsevier Publications, Vol.13, No.3,pp.370-393.
3. Varadharajan, S., Sehgal, V.K., and Saini, B. (2013), Seismic response of multistory
reinforced concrete frame with vertical setback irregularities, The structural design of tall and
special buildings, Wiley publications, DOI: 10.1002/tal.1147
4. Varadharajan, S., Sehgal V.K., Saini, B. (2013), Seismic response of multistory reinforced
concrete frame with vertical irregularities, ASCE Journal of structural engineering,
(communicated for publication).
(B) NATIONAL PUBLICATIONS
1. Varadharajan, S., Sehgal, V.K., and Saini, B. (2011), Seismic Response of Building Frames
with Vertical Setback and Stiffness Irregularity, IUP Journal of structural engineering,
Hyderabad, India, Vol.i, pp.52-61.
2. Varadharajan, S., Sehgal, V.K., and Saini, B. (2012),

Review of different structural

irregularities in buildings, Journal of Structural Engineering, Structural engineering research


center (Govt. of India Enterprise) Chennai India, Vol. 39, No. 5, December 2012 - January
2013 pp. 393-418

vii

TABLE OF CONTENTS
Page
No.
Candidates declaration

(i)

Certificate

(ii)

Acknowledgement

(iii)

Abstract

(iv)

List of publications

(vii)

Table of contents

(viii)

List of Figures

(xiv)

List of Tables

(xxxv)

Symbols and Notations

(xxxviii)

CHAPTER 1 : INTRODUCTION
1.1 Structural irregularity in buildings

1.2 Failures occurred during the past earthquakes

12

1.2.1 Failure due to vertical irregularities


1.2.1.1 Failure of setback structures
1.2.2 Failure of plan irregular buildings

12
19
29

1.3 Brief discussion on strong column weak beam design philosophy

37

1.4 Review of research works regarding plan irregularities

38

1.4.1 Single storey building models

38

1.4.2 Multistorey plan asymmetric structures

47

1.5 Review of research works regarding vertical irregularities

51

1.6 Brief review of Building Models used in previous research works

60

1.7 Discussions on Literature review

65

1.8 Organization of the Dissertation work

67

1.9 Brief summary and conclusions

69

CHAPTER 2 : STRUCTURAL MODELING AND ANALYSIS

56

2.1 Introduction

71

2.2.1 Definition of the base case

72

2.2.2 Definition of building models with plan irregularities

72

viii

2.2.3 Definition of building models with vertical irregularities


2.2.4 Definition of building models with vertical setback irregularity
2.3 Details of ground motions used

72
73
79

2.3.1 Scaling of the ground motion records


2.4 Review of analysis methods

79
81

2.4.1 Linear approach

81

2.4.1.1 Linear static analysis

82

2.4.1.2 Linear dynamic analysis

82

2.4.1.3 Time history method

84

2.4.2 Nonlinear methods

84

2.4.2.1 Non - Linear static analysis

84

2.4.2.2 Non - Linear dynamic analysis

85

2.5 Modeling of the inelastic behavior

86

2.5.1 Clough and Johnson model (1)

86

2.5.2 Takeda model (2)

87

2.5.3 Bouc-Wen model and its modifications (3)

88

2.5.4 Ramberg-Osgood model (4)

90

2.5.5 Kunnath et al. (1990) deterioration model (5)

91

2.5.6 Sivaselvan and Reinhorn model (6)

92

2.5.7 Yield Line Plastic Hinge models (7)

93

2.5.8 FEMA 356 Component model (8)

94

2.5.9 Song and Pincheira model (9)

95

2.5.10 Dutta and Das (2002) model (10)

97

2.5.11 Ibarras model (11)

97

2.5.12 Selection of hysteretic model to represent the inelastic behavior

97

2.6 Modeling of cracking behavior of structural elements

98

2.6.1 Strength - dependent stiffness reduction factors

98

2.6.2 Member initial stiffness using analytical study

100

2.7 Quantification of irregularity

102

2.7.1 Sensitivity analysis

102

2.7.2 Evaluation of the proposed irregularity indices for different

106

irregular building models


(a) Building models with mass, stiffness and strength irregularity

ix

106

(b) Building models with setback irregularity

107

(c) Building models with plan irregularity

111

(d) Evaluation of proposed irregularity index to capture effects of

111

cracking
2.8 Brief summary and conclusions

113

CHAPTER 3: DETERMINATION OF FUNDAMENTAL TIME - PERIOD OF


IRREGULAR BUILDINGS
3.1 Introduction

115

3.2 Building period - height relationship proposed by seismic design codes and

115

previous research works


3.2.1 Brief discussion on literature review

124

3.3 Modified equations for estimation of fundamental time period of irregular

124

buildings
3.31 Eigen-value analysis

124

3.3.1.1 Eigen-value analysis incorporating the cracking effects

124

3.3.2 Inelastic dynamic analysis

134

3.3.3 Alternate method of representing the fundamental time period

141

3.4 Generalized variation of fundamental time period

141

3.5 Range of applicability of proposed relations

142

3.6 Application of proposed period height relations in seismic design methods

145

3.6.1 Brief description of spectrum based displacement based design (DBD)

145

3.6.2 Application of the proposed fundamental time period equations in design

147

methodologies to estimate the seismic response


3.7 Seismic vulnerability assessment

153

3.7.1 Review of methods for seismic vulnerability assessment

153

3.7.2 Application of proposed fundamental time period equations in seismic

155

vulnerability assessment Method proposed by Glaister and Pinho (2003)


3.7.2.1 Method proposed by Glaister and Pinho (2003)
3.8 Brief summary and conclusions

162

CHAPTER 4 : DETERMINATION OF INELASTIC DEFORMATION

DEMANDS OF IRREGULAR BUILDING FRAMES


4.1 Introduction

165

4.2 Brief discussion on code provisions and previous literature works pertaining

165

to deformation demands
(a) Effect of mass irregularities on inelastic deformation demands

166

(b) Effect of stiffness and strength irregularity on deformation demands

166

(c) Effect of setback irregularities

185

(d) Effect of setback irregularities

191

4.2.2 Relation between elastic and inelastic deformation demands


4.3 Determination of behavior factor of irregular buildings

197
198

4.3.1 Brief review of different European codes regarding behavior factor

199

4.3.2 Review of literature with respect to behavior factor

199

4.3.3 Simplified relation between proposed irregularity index and

203

behavior factor
4.4 Estimation of deformation demands

206

4.4.1 Equal displacement rule proposed by EC 8 to estimate seismic

206

demands
4.4.2 Disadvantage of equal displacement rule

211

4.4.3 Equations to calculate global demands

211

4.4.4 Equations to calculate storey level demands in terms of

212

maximum interstorey drift ratio (Ir)


4.4.5 Equations to calculate local demands in terms of maximum

213

plastic hinge rotation (i)


4.5 Effect of cracking on deformation demands

217

4.6 Application of the proposed irregularity index and proposed mean relation

217

factors in DBD and PBD


4.7 Brief summary and main conclusions

222

CHAPTER 5: PREDICTION OF SEISMIC PERFORMANCE OF IRREGULAR


BUILDINGS
5.1 Introduction

225

5.2 Review of research works regarding collapse capacity of buildings

225

xi

5.2.1 Review of collapse assessment methods


(a) Single degree of freedom systems

226
226

(b) Non-linear static analysis

229

(c) Finite Element analysis

230

(d) Incremental Dynamic analysis

235

(e) Shake table experiments to determine the collapse capacity

238

5.2.2 Main conclusions derived from the literature review

242

5.3 Performance based earthquake engineering for collapse assessment

244

5.4 Modeling of RC frames

245

5.4.1 Element deterioration modes

245

5.4.2 Local and global collapse

246

5.5 Modeling for collapse assessment

259

5.5.1 Calibration procedure

260

(a) Effective stiffness and flexural strength

260

(b) Plastic-Rotation capacity (cap pl)

262

(c) Post-yield hardening strength ratio

263

(d) Pre-capping rotation capacity (pc)

263

(e) Cyclic energy dissipation capacity ()

263

5.6 Modeling uncertainties for collapse assessment


5.6.1 Modeling approach and collapse capacity assessment

263
264

methodology
5.7 Variation of collapse capacity for different building models
5.7.1 Equations to estimate the probability of collapse and collapse

271
274

capacity (Method 3)
5.8 Estimation of damage indices

276

5.8.1 Literature review pertaining to damage index

276

5.8.2 Brief discussion on damage indices

277

5.8.3 Park and Ang damage index (Method 4)

281

5.8.4 Seismic damage indices proposed by the Author (Method 5)

283

5.8.5 Comparison of probability of collapse using different approaches

284

5.9 Brief summary and main conclusions

288

xii

CHAPTER 6 :SUMMARY AND CONCLUSIONS


6.1 Summary

291

6.2 Main conclusions

292

REFERENCES

297

APPENDIX A : DETAILS OF ADDITIONAL GROUND MOTION ADOPTED

349

APPENDIX B : ADDITIONAL SEISMIC RESPONSE PARAMETERS

351

B.1 Equations to compute strength demands in members

351

B.2 Stress demands

351

APPENDIX C : SHORT PERIOD STRUCTURES

357

C.1 Introduction

357

C.2 Irregularity index for short period structures

357

APPENDIX D : ADDITIONAL RESULTS FOR SHORT PERIOD

365

STRUCTURES
APPENDIX E: MEAN VALUES OF SEISMIC RESPONSE PARAMETERS FOR

373

LONG PERIOD STRUCTURES


APPENDIX F: MEAN VALUES OF SEISMIC RESPONSE PARAMETERS

388

FOR LONG PERIOD STRUCTURES


APPENDIX G: ESTIMATION OF MEAN RELATION FACTORS

xiii

403

LIST OF FIGURES
Page
No.
3

Figure 1.1

Classification of different types of structural irregularities

Figure 1.2

The New Yorker hotel New York City, USA

Figure 1.3.

Manhattan city showing different irregular buildings

(a) View 1
(b) View 2
Figure 1.4

Pictures of existing irregular buildings

(a) Setback building in New Delhi (India)


(b) Empire state building
(c) 500 fifth (http://wikipedia.org)
Figure 1.5

Re-entrant corner irregularity

11

Figure 1.6

Mass irregularity

11

Figure 1.7

Irregular distribution of stiffness in the building system

11

Figure 1.8

Vertical setback irregularity

12

Figure 1.9

Failure of olive view medical state center (Nisee Berkeley)

13

Figure 1.10

Bending of columns and failure of parking structure during

15

Northridge earthquake
Figure 1.11

Failure of buildings due to soft storey in Turkey during Bingol

15

earthquake 2003 (Dogangun 2004)


Figure 1.12

Failure of buildings due to weak storey

17

(a) Garage formation at base floor (Durrani et al. 2008;


Steinbrugge http://nisee.berkeley)
(b) Weak storey failure during Kashmir earthquake (Kirac et al.
2011)
(c) Failure of weak storey with overhang (Kirac et al. 2011)
Figure 1.13

Failure of buildings due to soft storey in Turkey during Sumatra

19

earthquake 2004 earthquake (Kirac et al. 2011)


Figure 1.14

Failure of buildings due to soft storey in Turkey during Izmit

19

1999 earthquake (Kirac et al. 2011)


Figure 1.15

Overview of building TO-9 9 (Westenenk 2013)


(a) West side view
(b) East side view

xiv

23

(c) South side view and close-up of deformed axis J


(d) Plan view
(e) Details of the main collapse at story 12
(f) Damage in short columns in axis 1A
(g) Overall condition of the 21st storey
Figure 1.16

Internal damage of building TO-9: (a) Structural plan of the 12th

25

storey, (b) Partial collapse of the 16th storey, (c) Partial collapse
of the 20th storey, (d) damaged beams of the 12th storey, (e)
Column in axes G-5 of the 12th storey, (f) Column in axes G-4
(Westenenk 2013)
Figure 1.17

Internal damage of building TO-9: (a) Slab deformation in storey

25

16, (b) Overview of damage in storey 18, (c) Severely damaged


column in storey 21 (Westenenk 2013)
Figure 1.18

Failure due to overhangs (Kirac et al. 2011)

27

Figure 1.19

Failure due to overhangs (Dogangun et al.2003)

27

Figure 1.20

Failure of balconies at Shizugawa hospital during Tsunami 2011

27

(Fraser et al. 2013)


Figure 1.21

Damage from irregularity during 1978 Miyagi Ken Oki Japan

31

earthquake (Ellingwood 1980)


(a) Overall view of damage of three storey reinforced concrete
buildings
(b) Damage to column along periphery along the wall
Figure 1.22

Damage from Irregularity during 2010 Haiti earthquake

33

(a) Overall view of damage to Ministry of Culture building


(b) Damage on flexible side c) Damage on stiff side
(Mid America Earthquake Engineering Center, University of
Illiniosis at Urbana Champaign
Figure 1.23

Damage due to irregularity during Guatemala earthquake 1976

35

(a) Overall view of a hotel terminal building in Guatemala


(b) Collapse of second storey due to shear failure of column
(c) Close of one of the collapsed column (Bertero, 1997)
Figure 1.24

First storey column damage due to weak column strong beam


philosophy (Nisee Berkley)

xv

37

Figure 1.25

Strong beam weak column structural system of the Celtikuyu

37

school building and dormitory building(Dog angun 2004)


Figure 1.26

Strong beam weak column structural system of the Panta

38

parking school at Banta Aceh


Figure 1.27

Types of eccentricity

39

(a) Mass eccentricity


(b) Stiffness eccentricity
(c) Strength eccentricity
Figure 1.28

Structural model considered by De-La colina (1999)

44

Figure 1.29

Second Hysteretic model proposed by Dutta and Das (2002)

44

Figure 1.30

Models considered by Aziminejad and Moghadam (2010)

46

Figure 1.31

Different types of vertically irregular building models used by

56

Das and Nau (2003)


Figure 1.32

Hysteresis models

62

(a) Elasto-plastic model


(b) Bi-linear model
Figure 1.33

(a) Cloughs degrading stiffness model

62

(b) Takedas hysteresis model


Figure 2.1

Period - height relationships for building models adopted in the

74

analytical study complying with Goel and Chopra (1997)


Figure 2.2

Setback building models varying from 6 to 18 storeys adopted

75

for the analytical study


(a) 6 storey
(b) 9 storey
(c) 12 storey
(d) 15 storey
Figure 2.3

Setback building models varying from 18 to 24 storeys adopted

76

for the analytical study


(a) 18 storey
(b) 21 storey
(c) 24 storey
Figure 2.4

Setback building models varying from 27 storeys to 30 storeys


adopted for the analytical study

xvi

77

(a) 27 storey
(b) 30 storey
Figure 2.5

Setback building models varying from 33 to 36 storeys adopted

78

for the analytical study


(a) 33 storey
(b) 36 storey
Figure 2.6

Mean of ground motions 1 Scaled to EC 8 spectrum (PEER)

79

Figure 2.7

EC 8 spectrum (Target spectrum) to which ground motions have

81

been scaled (PEER)


Figure 2.8

Peak-Oriented degrading stiffness model proposed by Clough

86

and Johnston 1966


Figure 2.9

Hysteric model proposed by Sucuoglu and Erberik (2004)

87

Figure 2.10

Bouc-Wen model, separation of linear restoring force component

89

from hysteric restoring force component (Foliente 1995)


Figure 2.11

Effect of parameter

90

Figure 2.12

Degradation in system properties: (a) Strength degradation, (b)

90

Stiffness degradation (Foliante et al. 1995)a)


Figure 2.13

Hysteretic pinching effect: Baber and Nooris (1986) PinchIng-

90

Function (Foliante et al. 1995)


Figure 2.14

Force deformation relationships in Ramberg-Osgood model

91

(Carr 2003)
Figure 2.15

Three parameter model (Kunnath et al. 1990)

92

Figure 2.16

Sivaselvan model to represent hysteric behavior proposed by

93

Sivaselvan and Reinhorn (2000)


Figure 2.17

Yield line model proposed by Lee and Stojadinovic (2000)

94

Figure 2.18

General force deformation behavior of structural components

95

(FEMA 356, 2000)


Figure 2.19

Song Pincheira Model; (a) Backbone curve, (b) Hysteresis

96

rules for cycles of increasing deformation amplitude, (c)


Hysteresis rules for small amplitude or internal cycles
Figure2.20

(a) Moment Curvature (M ) relationship assuming constant


stiffness (Priestley, 2003)
(b) Moment-Curvature relationship under realistic condition i.e.

xvii

99

considering strength and stiffness to be interdependent


(Priestley 2003)
Figure 2.21

Moment curvature for a rectangular column (Priestley 2003)

101

Figure 2.22

Effective stiffness of a large rectangular column (Priestley 2003)

101

Figure 2.23

Variation of irregularity index with magnitude of mass, strength

104

and stiffness irregularity


Figure 2.24

Variation of irregularity index with location of mass, strength and

104

stiffness irregularity
Figure 2.25

Variation of irregularity index with mass, strength and stiffness

105

irregularity considering the cracking effects


Figure 2.26

Variation of irregularity index with plan irregularity (with and

105

without cracking affects)


Figure 2.27

Variation of irregularity index with setback irregularity with and

106

without cracking affects cracking


Figure 2.28

Plan irregular Building models for study of variation of proposed

107

irregularity index
Figure 2.29

Code limits of setback irregularity

109

(a) IS 1893:2002
(b) ASCE 7.2005
(c) EC 8:2004
Figure2.30

Frame geometry for definition of irregularity indices proposed by

110

Karavasilis et al. (2008a)


Figure 2.31

Setback Frames considered for the comparison of the

110

irregularity index
Figure 3.1

Results of RC MRF frames using eigen - value analysis by Goel

117

and Chopra (1997)


Figure 3.2

Tunnel form construction technique used in Turkey (Balkaya and

119

Kalkan 2003)
Period - height relationship for different irregular models based
Figure 3.3
Figure 3.4

126

on gross stiffness using eigen - value (EV) analysis


Mean percentage difference (on comparison with regular
building model) in fundamental time period of different irregular
building models

xviii

126

Figure 3.5

Period - height relationship for different irregular models based

127

on cracked stiffness (Priestley 2003 approach)


Figure 3.6

Mean percentage difference in fundamental time period

(on

127

comparison with regular building model) for different irregular


building models considering cracking effect
Figure 3.7

Mean percentage difference between periods obtained using

128

different cracking approaches (Paulay and Priestley 1992 and


Priestley 2003)
Figure 3.8

Period - height relationship for different building models using

130

eigen value analysis considering gross stiffness for


(a) Mass irregularity
(b) Plan irregularity
(c) Strength irregularity
(d) Setback irregularity
(e) Stiffness irregularity
Figure 3.9

Period - height relationship for different building models using

132

eigen value analysis considering cracked stiffness (Priestley


2003)
(a) Mass irregularity
(b) Stiffness irregularity
(c) Strength irregularity
(d) Setback irregularity
(e) Plan irregularity
Figure 3.10

Period - height relationship for different irregular models based

135

on gross stiffness using inelastic dynamic analysis


Figure 3.11

Period - height relationship for different irregular models based

135

on cracked stiffness (Priestley 2003 approach) using inelastic


dynamics analysis
Figure 3.12

Period - height relationship for different building models using


IDA analysis considering for (Gross stiffness consideration)
(a) Mass irregularity
(b) Plan irregularity
(c) Stiffness irregularity

xix

136

(d) Strength irregularity


(e) Setback irregularity
Figure 3.13

Period - height relationship for different building models using

138

IDA analysis considering cracked stiffness (Priestley 2003


consideration) for:
(a) Mass irregularity
(b) Plan irregularity
(c) Stiffness irregularity
(d) Strength irregularity
(e) Setback irregularity
Figure 3.14

Comparison between proposed equation and dynamic analysis

142

for irregular buildings


Figure 3.15

Effect of magnitude of mass, stiffness and strength irregularity

143

on fundamental time period


Figure 3.16

Effect of location of mass, stiffness and strength irregularity on

143

fundamental time period


Figure 3.17

(a) Effect of setback irregularity on fundamental time period with

144

and without cracking effects


(b) Effect of plan irregularity with and without cracking effects on
fundamental time period
Figure 3.18

Effect of cracking on fundamental time period

145

Figure 3.19

Displacement spectrum generated for use in spectrum based

146

DBD
Figure 3.20

Comparison of proposed equation by different methods for

148

estimating period (by eigen - value analysis and IDA analysis


including cracking effects for building models with mass
irregularity
Figure 3.21

Comparison of proposed equation by different methods for

148

estimating period (by eigen - value analysis and IDA) including


cracking effects for building models with stiffness irregularity
Figure 3.22

Comparison of proposed equation by different methods for


estimating period (by eigen - value analysis and Inelastic
dynamic analysis) including cracking effects for strength

xx

149

irregular buildings
Figure 3.23

Comparison of proposed equation by different methods for

149

estimating period (by eigen - value analysis and Inelastic


dynamic analysis) including cracking effects for setback irregular
buildings
Figure 3.24

Comparison of proposed equation by different methods for

150

estimating period (eigen - value analysis and IDA) including


cracking effects for plan irregular buildings
Figure 3.25

Percentage increase (with reference to regular building model)

150

of IDR for building models with mass irregularity by different


methods adopted (M 200 M1000 represents mass irregularity
with a magnitude ranging from 200 % to 1000 %)
Figure 3,26

Mean Percentage increase of IDR (with reference to regular

151

building model) for plan irregular building models by different


methods adopted
Figure 3.27

Mean percentage increase of IDR (with reference to regular

151

building model) for building models with stiffness irregularity by


different methods adopted (S 25 S100 represents stiffness
irregularity with a magnitude ranging from 25 % to 100 %)
Figure 3.28

Mean percentage increase of IDR (with reference to regular

152

building model) for building models with strength irregularity by


different methods adopted (ST 25 ST100 represents strength
irregularity with a magnitude ranging from 25% to 100%)
Figure 3.29

Mean percentage increase of IDR (with reference to regular

152

building models) for building models with setback irregularity by


different methods adopted
Figure 3.30

Seismic Vulnerability assessment using deformation based

156

approach (Glaister and Pinho 2003) where HLsi is the equivalent


building height, TLsi is the building period corresponding to
demand, PLsi is probability of failure corresponding to demand,
LS1, LS2 and LS3 are the corresponding performance levels
Figure 3.31

Glaister and Pinho (2003) methodology for Seismic vulnerability


assessment of 6 storey irregular building (DC is the demand

xxi

158

curve, CC is the capacity curve, CP is the seismic capacity)


Figure 3.32

Mean derived capacity periods for building models considered in

158

the analytical study


Figure 3.33

Mean probability of collapse for mass irregular building models

159

based on Glaister and Pinho (2003) methodology


Figure 3.34

Mean probability of collapse for stiffness irregular building

159

models based on Glaister and Pinho (2003) methodology


Figure 3.35

Mean probability of collapse for strength irregular building

160

models based on Glaister and Pinho (2003) methodology


Figure 3.36

Mean probability of collapse for setback irregular building

160

models based on Glaister and Pinho (2003) methodology


Figure 3.37

Mean probability of collapse for plan irregular building models

161

based on Glaister and Pinho (2003) methodology


Figure 3.38

Mean probability of collapse of buildings using Glaister and


Pinho (2003) methodology [ Method 1)

161

Hazus methodology

[Method 2) [GPG and GPC Glaister and Pinho (2003) method


with gross and cracked stiffness, HG,HC represent Hazus
methodology with gross and cracked stiffness)
Figure 3.39

Mean probability of collapse for buildings of buildings using

162

different cracking approaches (GPC1 represents Priestley 2003


approach and GPC2 represents Paulay and Priestley 1992
approach)
Figure 4.1

Variation of global deformation demands along building height

169

for mass irregular buildings


Figure 4.2

Variation of mean storey deformation demand in terms of

169

maximum IDR (Ird) along building height for mass irregular


buildings
Figure 4.3

Variation of mean local deformation demand along building

170

height for mass irregular buildings


Figure 4.4

Mean Percentage increase on comparison with regular building


model (Mean of all height categories) of deformation parameters
(Mean of rd, Ird, i). with variation of magnitude and location of
mass irregularity (Model category 1 to 3 and 4 to 6 denotes

xxii

170

irregularity at bottom, middle and top storeys with and without


cracking effects)
Figure 4.5

Failure of building due to heavy mass at top during Bhuj

171

earthquake (Humar et al. 2001)


Figure 4.6

Variation of global deformation demands along building height

174

for stiffness irregular buildings (BSS,MSS,TSS denotes stiffness


irregularity at bottom one third, middle one third and top one
third storeys)
Figure 4.7

Variation of storey deformation demands along building height

174

for stiffness irregular buildings


Figure 4.8

Variation of local deformation demands along building height for

175

stiffness irregular buildings


Figure 4.9

Mean percentage increase (on comparison with regular building

175

model) of deformation parameters (Mean of rd, Ird, i) with


variation of magnitude of stiffness irregularity in building models
Figure 4.10

Variation of global deformation demands along building height

176

for strength irregular buildings (BSST,MSST,TSST denotes


stiffness irregularity at bottom one third, middle one third and top
one third storeys)
Figure 4.11

Variation of storey deformation demands along building height

176

for strength irregular buildings


Figure 4.12

Variation of local deformation demands along building height for

177

strength irregular buildings


Figure 4.13

Mean percentage increase (on comparison with regular building

177

model) of deformation parameters with reference to regular


building for strength irregularity building models
Figure 4.14

Failure of buildings due to strength and stiffness irregularity at

179

bottom storeys (Taskin et al. 2013)


Figure 4.15

Failure of buildings due to strength and stiffness irregularity at

179

bottom storeys (Romeo et al. 2013)


Figure 4.16

Failure of tall buildings with stiffness and strength irregularity

181

during Bhuj earthquake (Humar et al. 2001)


Figure 4.17

Failure of tall buildings due to strength and stiffness irregularity

xxiii

183

at bottom storeys during Chi Chi earthquake 1999 (Tsai et al.


2000)
Figure 4.18

Failure of buildings with stiffness and strength irregularity at

183

middle stories during 1995 Kobe (Japan) earthquake, (Elnashai


and Di Sarno, 2008)
Figure 4.19

Failure of buildings with stiffness and strength irregularity at

185

middle stories during Bhuj earthquake (Paul and Dogan, as


reoprted in Kirac et al. 2013)
Figure 4.20

Variation of global deformation demands along building height

187

for setback irregular buildings


Figure 4.21

Variation of storey deformation demands along relative building

187

height for setback irregular buildings

Figure 4.22

Variation of local deformation demands along relativebuilding

188

height for setback irregular buildings


Figure 4.23

Failure of Torre o Higgins building near setback (at 12th storey)

188

during Chile earthquake 2010 (Kato et al. 2010)


Figure 4.24

Variation of global demands along building height for setback

189

irregular buildings
Figure 4.25

Variation of storey deformation demands along building height

189

for setback irregular buildings


Figure 4.26

Variation of local deformation demands along building height for

190

setback irregular buildings


Figure 4.27

Mean percentage increase in deformation demands (on

190

comparison with regular building model) for setback irregular


buildings with reference to regular buildings (Model 1 Gross
stiffness, Model 2 Cracked stiffness)
Figure 4.28

Comparison of global demands of plan irregular structures

193

designed using different seismic design codes


Figure 4.29

Comparison of storey demands of plan irregular structures

194

designed using different seismic design codes


Figure 4.30

Comparison of local demands of plan irregular structures

194

designed using different seismic design codes


Figure 4.31

Effect of plan irregularity on seismic deformation demands

xxiv

195

(a) Global demands


(b) Storey demands
(c) Local demands
Figure 4.32

Percentage increase in mean deformation demands for plan

196

irregular buildings
Figure 4.33

Building collapse due to torsion during L Aquilla earthquake

196

2009 (Verdane et al. 2010)


Figure 4.34

Relation between irregularity index and behavior factor for gross

204

stiffness
Figure 4.35

Relation between irregularity index and behavior factor for

204

cracked stiffness
Figure 4.36

Relation between irregularity index and behavior factor for gross

205

stiffness
Figure 4.37

Relation between irregularity index and behavior factor for

205

cracked stiffness
Figure 4.38

Pictorial representation of equal displacement rule ( EC 8 2004)

207

Figure 4.39

Comparison of EC 8 equation with dynamic analysis for rd

212

Figure 4.40

Comparison of the proposed equation with dynamic analysis for

212

rd
Figure 4.41

Comparison of equation using the proposed equation and

214

dynamic analysis for Ir using


(a) First approach
(b) Second approach
Figure 4.42

Comparison between EC 8 and dynamic analysis for storey

215

demands of irregular buildings


Figure 4.43

Local demands of irregular buildings

215

(a) Comparison between the proposed equation and EC 8


procedure
(b) Comparison between the proposed equation and dynamic
analysis
Figure 4.44

Design details of columns of the irregular and regular building

220

models studied
Figure 4.45

Design details of beams of the irregular and regular building

xxv

221

models studied
Figure 4.46

Reinforcement weights for building models

221

Figure 4.49

Serviceability criteria check based on inelastic analysis

222

a) Interstorey drift
b) Chord rotation (rad)
Figure 5.1

Flexural hinging at top and bottom of columns at a Mosque in

248

Banda Aceh during Turkey earthquake (Ghobarah et al. 2006)


Figure 5.2

Punching shear failure of flat plate construction

(Arslan and

248

Column shear failure in five storey office building (Ghobarah et

250

Korkmaz 2007)
Figure 5.3

al. 2006)
Figure 5.4

Shear failure of column (Arslan and Korkmaz 2007)

250

Figure 5.5

Column shear failure due to buckling of longitudinal bars during

250

Turkish earthquakes (Arslan and Korkmaz 2007)


Figure 5.6

Column shear failure due to wide spacing of lateral ties during

252

Turkish earthquakes (Arslan and Korkmaz 2007)


Figure 5.7

Shear failure of rectangular column and square column during L

252

Aquilla earthquake 2009 (Ricci 2011b)


Figure 5.8

Shear failure of beam column joint duringTurkish earthquakes

252

(Arslan and Korkmaz 2007


Figure 5.9

Failure of beam column joint and undamaged joint during

254

Turkish earthquakes (Arslan and Korkmaz 2007)


Figure 5.10

Anchorage pull out failure of lapped reinforcement (Kam et al.

254

2010)
Figure 5.11

(a) Inadequate lap splice

254

(b) Beam column joint failure due to lack of transverse


reinforcement at a office building in Banda Aceh (Arsalan and
Korkmaz 2007)
Figure 5.12

Joint failure with evident (a) longitudinal bar buckling, (b)

256

diagonal cracking failure in concrete joint (c,d) Failure


mechanisms in joint column interface (Ricci et al.2011b)
Figure 5.13

Shear

- axial failure of column during Darefield earthquake

(Kam et al.2010)

xxvi

256

Figure 5.14

Different modes of collapse (NISEE, UCSD)

258

Figure 5.15

Ibarras nonlinear hinge model

260

Figure 5.16

Basic layout of the Building model

265

Figure 5.17

(a) Results of incremental dynamic analysis

267

(b) Collapse fragility curve with adjustment for different errors


Figure 5.18

Variation of collapse capacity with parameter Epsilon ()

268

Figure 5.19

Effect of vertical collapse capacity

268

Figure 5.20

Pyne group corporation building New - Zealand photographed

269

from North East elevation after Darrfield earthquake (Kam et


al. 2010)
Figure 5.21

(a) Overall failure of PGC building due to detachment of 1st and

269

2nd storey columns from its base due to which they lost their
vertical load carrying capacity
(b) Zoom A detail
(c) Zoom b detail
(d) Zoom c detail
(e, f) detail of damage to beam column joints (Kam et al. 2010)
Figure 5.22

Variation of collapse capacity with post capping rotation ratio

272

Figure 5.23

Variation of collapse capacity with the bay width

272

Figure 5.24

Variation of collapse capacity with location of mass, stiffness

272

and strength irregularity


Figure 5.25

Variation of collapse capacity with magnitude of mass, stiffness

273

and strength irregularity


Figure 5.26

Variation of collapse capacity with magnitude of setback

273

irregularity
Figure 5.27

Variation of collapse capacity with magnitude of plan irregularity

273

Figure 5.28

Variation of different parameters with cracking for building

274

models with mass, stiffness and strength irregularity


Figure 5.29

Comparison of collapse capacity evaluated using the proposed

275

method with the dynamic analysis


for the irregular buildings considered in the analytical study
Figure 5.30

Comparison of probability of collapse using the proposed


method with the dynamic analysis for the irregular buildings

xxvii

275

considered in the analytical study


Figure 5.31

Comparison of damage index (Park and Ang 1985 a)

283

determined using proposed equation and dynamic analysis for


irregular buildings considered in the analytical study (DIPA
denotes Damage index proposed by Park and Ang 1985a)
Figure 5.32

Comparison of Park and Ang (1985a) index determined using

284

proposed equation and dynamic analysis for irregular buildings


Figure 5.33

Mean probability of collapse for buildings with mass irregularity

285

Figure 5.34

Mean probability of collapse for buildings with stiffness

285

irregularity
Figure 5.35

Mean probability of collapse for buildings with strength

286

irregularity
Figure 5.36

Mean probability of collapse for buildings with setback

286

irregularity
Figure 5.37

Mean probability of collapse for buildings with plan irregularity

287

Figure 5.38

Mean percentage difference in probability of collapse evaluated

287

by different methods
Figure A.1

Response spectrums adopted by different codes of practice

350

Figure A.2

Scaling of ground motions in Table A.1 for long period structures

350

(PEER)
Figure B.1

Comparison of propose equation with the dynamic analysis for

352

(a) Strength demand in terms of over-strength factor ()


Figure B.2

Stress demand for irregular building models considered in the

352

analytical study
Figure C1

Different geometries considered in the analytical study of short

358

period structures
Figure C2

Mean scaled spectrum for ground motions in Table C.2 for short

360

period structures
Figure C.3

Mean scaled spectrum for ground motions in Table C.3 for short

360

period structures
Figure C.4

Relation between proposed irregularity index and behavior factor

361

for short period structures considering gross stiffness


Figure C.5

Relation between proposed irregularity index and behavior factor

xxviii

362

for short period structures considering

cracked stiffness

(Priestley 2003)
Figure C.6

Comparison of behavior factor for short period structures using

362

proposed equation and dynamic analysis for gross stiffness


Figure C.7

Comparison of behavior factor for short period structures using

363

proposed equation and dynamic analysis


Figure E.1

Mean values of irregularity index for building models with mass

373

irregularity
Figure E.2

Mean values of behavior factor for building models with different

373

height categories with presence of mass irregularity


Figure E.3

Mean values of roof displacement for building models with mass

373

irregularity
Figure E.4

Mean values of interstorey drift for building models with mass

374

irregularity
Figure E.5

Mean values of rotation for building models with mass

374

irregularity
Figure E.6

Mean values of overstrength factor for building models with

374

mass irregularity
Figure E.7

Mean values of stress demands for building models with mass

375

irregularity
Figure E.8

Mean values of collapse capacity for building models with mass

375

irregularity
Figure E.9

Mean values of fundamental time period for building models

375

with mass irregularity


Figure E.10

Mean values of irregularity index for building models with

376

stiffness irregularity
Figure E.11

Mean values of behavior factor for building models with stiffness

376

irregularity
Figure E.12

Mean values of roof displacement

for building models with

376

Mean values of interstorey drift for building models with stiffness

377

stiffness irregularity
Figure E.13

irregularity
Figure E.14

Mean values of rotation for building models with stiffness

xxix

377

irregularity
Figure E.15

Mean values of overstrength factor building models with stiffness

377

irregularity
Figure E.16

Mean values of stress demand for building models with stiffness

378

irregularity
Figure E.17

Mean values of collapse capacity

for building models with

378

Mean values of fundamental time period for building models

378

stiffness irregularity
Figure E.18

with stiffness irregularity


Figure E.19

Mean values of irregularity index for building models with

379

strength irregularity
Figure E.20

Mean values of behavior factor for building models with strength

379

irregularity
Figure E.21

Mean values of maximum roof displacement for building models

379

with strength irregularity


Figure E.22

Mean values of interstorey drift for building models with strength

380

irregularity
Figure E.23. Mean values of rotation for building models with strength

380

irregularity
Figure E.24

Mean values of overstrength factor for building models with

380

strength irregularity
Figure E.25

Mean values of stress for building models with strength

381

irregularity
Figure E.26

Mean values of collapse capacity for building models with

381

strength irregularity
Figure E.27

Mean values of fundamental time period for building models

381

with strength irregularity


Figure E.28

Mean values of irregularity index for building models with

382

setback irregularity
Figure E.29

Mean values of behavior factor for building models with setback

382

irregularity
Figure E.30

Mean values of roof displacement


setback irregularity

xxx

for building models with

382

Figure E.31

Mean values of interstorey drift for building models with setback

383

irregularity
Figure E.32

Mean values of rotation for building models with setback

383

irregularity
Figure E.33

Mean values of overstrength factor for building models with

383

setback irregularity
Figure E.34

Mean values of stress demand for building models with setback

384

irregularity
Figure E.35

Mean values of collapse capacity

for building models with

384

Mean values of fundamental time period for building models

384

setback irregularity
Figure E.36

with setback irregularity


Figure E.37

Mean values of irregularity index for building models with plan

385

irregularity
Figure E.38

Mean values of behavior factor for building models with plan

385

irregularity
Figure E.39

Mean values of roof displacement for building models with plan

385

irregularity
Figure E.40

Mean values of interstorey drift for building models with plan

386

irregularity
Figure E.41

Mean values of rotation for building models with plan irregularity

386

Figure E.42

Mean values of overstrength factor for building models with plan

386

irregularity
Figure E.43

Mean values of stress for building models with plan irregularity

387

Figure E.44

Mean values of collapse capacity for building models with plan

387

irregularity
Figure E.45

Mean values of fundamental time period for building models with

387

plan irregularity
Figure F.1

Mean values of irregularity index for building models with mass

388

irregularity
Figure F.2

Mean values of behavior factor for building models with different

388

height categories with presence of mass irregularity


Figure F.3

Mean values of maximum roof displacement for building models

xxxi

388

with mass irregularity


Figure F.4

Mean values of interstorey drift for building models with mass

389

irregularity
Figure F.5

Mean values of rotation for building models with mass

389

irregularity
Figure F.6

Mean values of overstrength factor for building models with

389

mass irregularity
Figure F.7

Mean values of stress demands for building models with mass

390

irregularity
Figure F.8

Mean values of collapse capacity for building models with mass

390

irregularity
Figure F.9

Mean values of fundamental time period for building models

390

with mass irregularity


Figure F.10

Mean values of irregularity index for building models with

391

stiffness irregularity
Figure F.11

Mean values of behavior factor for building models with stiffness

391

irregularity
Figure F.12

Mean values of roof displacement

for building models with

391

Mean values of interstorey drift for building models with stiffness

392

stiffness irregularity
Figure F.13

irregularity
Figure F.14.

Mean values of rotation for building models with stiffness

392

irregularity
Figure F.15

Mean values of overstrength factor for building models with

392

stiffness irregularity
Figure F.16

Mean values of stress for building models with stiffness

393

irregularity
Figure F.17

Mean values of collapse capacity

for building models with

393

Mean values of fundamental time period for building models

393

stiffness irregularity
Figure F.18

with stiffness irregularity


Figure F.19

Mean values of irregularity index for building models with


strength irregularity

xxxii

394

Figure F.20

Mean values of behavior factor for building models with strength

394

irregularity
Figure F.21

Mean values of roof displacement

for building models with

394

Mean values of interstorey drift for building models with strength

395

strength irregularity
Figure F.22

irregularity
Figure F.23

Mean values of rotation for building models with strength

395

irregularity
Figure F.24

Mean values of overstrength factor for building models with


strength irregularity

Figure F.25

395

Mean values of stress demand for building models with strength

396

irregularity
Figure F.26

Mean values of collapse capacity

for building models with

396

Mean values of fundamental time period for building models

396

strength irregularity
Figure F.27

with strength irregularity


Figure F.28

Mean values of irregularity index for building models with

397

setback irregularity
Figure F.29

Mean values of behavior factor for building models with setback

397

irregularity
Figure F.30

Mean values of roof displacement

for building models with

397

Mean values of interstorey drift for building models with setback

398

setback irregularity
Figure F.31

irregularity
Figure F.32

Mean values of rotation for building models with setback


irregularity

Figure F.33

398

Mean values of overstrength factor for building models with

398

setback irregularity
Figure F.34

Mean values of stress demand for building models with setback

399

irregularity
Figure F.35

Mean values of collapse capacity

for building models with

setback irregularity
Figure F.36

Mean values of fundamental time period for building models

xxxiii

399

with setback irregularity


Figure F.37

399

Mean values of irregularity index for building models with plan

400

irregularity
Figure F.38

Mean values of behavior factor for building models with plan

400

irregularity
Figure F.39

Mean values of maximum roof displacement for building models

400

with plan irregularity


Figure F.40

Mean values of interstorey drift for building models with plan

401

irregularity
Figure F.41

Mean values of rotation for building models with plan irregularity

401

Figure F.42

Mean values of overstrength factor for building models with plan

401

irregularity
Figure F.43

Mean values of stress for building models with plan irregularity

402

Figure F.44

Mean values of collapse capacity for building models with plan

402

irregularity
Figure F.45

Mean values of fundamental time period for building models


with plan irregularity

402

xxxiv

LIST OF TABLES

Table 1.1

Irregularity limits prescribed by IS 1893:2002, EC8:2004,

Page No.
9

UBC 97, NBCC 2005


Table 1.2

Irregularity limits prescribed by IBC 2003, TEC 2007 and

10

ASCE 7.05
Table 1.3.

Descriptions of different models adopted by Tso and

40

Sadek (1989)
Table 1.4

Code evaluation results Chandler and Hutchinson (1992)

41

Table 1.5

Results obtained considering different codes (Chandler

41

and Hutchinson 1992)


Table 1.6

Different position of centers of mass, stiffness, strength

46

and displacement for different values of .


Table 1.7

Different position of centers of mass, stiffness, strength

49

and displacement for different values of


Table 1.8

Different model configurations proposed

50

Table 1.9

Building models used by Nasser and Krawlinker (1991)

55

Table 1.10

Results of analytical study obtained by Esteva (1992)

55

Table 1.11

First system of comparison

63

Table 1.12

Second system of comparison

64

Table 1.13

Third system of comparison

65

Table 2.1

Detailed location of irregularities

74

Table 2.2

Details of ground motion used

80

Table 2.3

Comparison of hysteresis models

97

Table 2.4

Reduction factors proposed by Paulay and Priestley

100

(1992)
Table 2.5

Effective member stiffness as per FEMA 356

100

Table 2.6

Expressions to determine the yield curvature by Paulay

100

and Priestley (2003)


Table 2.7

Sensitivity analysis results

103

Table 2.8

Comparison of irregularity indices of seismic design

107

codes and proposed method for building models with


mass, stiffness and strength irregularity (gross stiffness)

xxxv

Table 2.9

Code prescribed limits of setback irregularity

109

Table 2.10

Comparison of the irregularity indices for setback

110

structures
Table 2.11

Comparison of the irregularity indices for plan irregular

111

structures
Table 2.12

Evaluation of the proposed irregularity indices for mass,

112

stiffness and strength with cracking considerations


Table 2.13

Comparison of the irregularity indices for setback

112

structures with cracking considerations


Table 2.14

Comparison of the irregularity indices for plan irregular

112

structures cracking considerations


Table 3.1

Empirical equations to predict fundamental time period of

119

tunnel form buildings


Table 3.2

Summary of time period expressions proposed by other

123

researchers
Table 3.3

Detailed values of parameter h and

h for different

128

irregular building models based on eigen - value analysis


Table 3.4

Detailed values of parameter h and

h for different

140

irregular building models based on inelastic dynamic


analysis
Table 4.1

Limits on deformation demands by some of the seismic

166

design codes
Table 4.2

Building failures due to plan asymmetry during different

192

earthquakes
Table 4.3

Eccentricities proposed by different seismic design codes

193

Table 4.4

Mean values of parameter irr for irregular building

197

models
Table 4.5

Performance of Equal displacement rule

210

Table 4.6

Equations to estimate the seismic response parameters

217

for irregular building models


Table 5.1

Different mode of collapse of structural members

245

Table 5.2

Collapse types of different types of frames

246

Table 5.3

Summary of local damage indices (cumulative) proposed

278

xxxvi

by different researchers
Table 5.4

Summary of other local damage indices proposed by

279

different researchers
Table 5.5

Summary of global damage indices proposed by different

280

authors
Table 5.6

Interpretation of Park and Ang global damage index

282

Table A.1

Ground motion data pertaining to ATC 3-06 (1978)

349

[NEHRP 1994, UBC 97, IS1893:2002, EC8:2004; ,IBC


2003] for long period structures (PEER)
Table A.2

Ground motion data pertaining to ATC 3-06 (1978)

319

[NEHRP1994, UBC 97, IS1893:2002, EC8 : 2004, IBC


2003] for long period structures
Table C.1

System of ground motions used for short period

359

structures
Table C.2

Ground motion data pertaining to ATC 3-06 (1978)

359

[NEHRP 1994 , UBC 97, IS1893:2002, EC8:2004, ,IBC


2003] for short period structures (PEER)
Table C.3

Detailed locations of irregularities

359

Table C.4

Sensitivity analysis results

360

Table G.1

Mean relation factor hm for long period structures

411

Table G.2

Mean relation factor hm for short period structures

412

xxxvii

SYMBOLS AND NOTATIONS


(Ae)i

Area of the storey

parameter representing stiffness degradation

A+, A-

Hysteresis loops

A1 to A6

Hysteresis load paths

A11

Hysteric shape parameters

A22

Point in Hysteresis loop

ABS

Absolute square rule of modal combination

AC

Accelerating stiffness deterioration

(Ae)i+1

Area of the storey above ith storey

Ag

Gross area of column

As

Area of reinforcement

Asw

Area of coupled shear wall

Aw

Area of shear wall

a,a

Constant

a11

Constant

ag

Acceleration due to ground motion

aKo

Yield stiffness

as1

Parameter indicating slip

B+, B-

Hysteresis loops

B1 to B6

Hysteresis load paths

b1, b2 and b3

Constants

B11

Number of bays

xxxviii

b11 to b16

Constants

B22

Point in hysteresis loop

b22

Constant

BAU

Bauschinger effect

BS

irregularity in the bottom storey

BAD

Basic strength deterioration

BSM1 BSM2

Mass irregularity of magnitude M1 and M2 at bottom storey

BSS1 BSS2

Stiffness irregularity of magnitude S1 and S2 at bottom storey

BSST1 BSST2

Strength irregularity of magnitude ST11 and ST2 at bottom storey

bw

Width of the building

Code approach

C 11, C 12

Constants

C1 to C6

Hysteresis load paths

C22, C22

Points in Hysteresis loop

Cas

Column drift ratio corresponding to shear failure

CC

Capacity curve

CI

Cumulative index

Clvss

Column drift ratio corresponding to loss of vertical carrying capacity

CM

Center of mass

CME

Collapse mechanism

CO

Collapse type

CP

Capacity point

CQC

Complete quadratic combination rule

xxxix

CR

Center of rigidity

CS

Center of stiffness

Ct

Constant

CV

Center of Strength

CY

Cyclic hardening

Constant

Displacement

D*y

Yield displacement in equation

Total displacement

D11

Diameter of column

D22,D22

Points in hysteresis loop

DAF

Dynamic amplification factor

DB

Deformation based

DBD

DBD is displacement based design

DC

Demand curve

De

Distance from epicenter

Df

Damage due to final softening

Di

denotes the deformation at ith cycle

Dpl

Damage due to plastic softening

Dstorey

Damage in a storey

DU

Ductile frame

Du

Ultimate deformation

Depth of the section

xl

d*

Total inter - storey drift

d*y

Yield inter - storey drift in equation

d1

Diameter of the column

d11

Distance between center of strength and rigidity

da1

Column depth

davg

Average of drifts computed at both sides of a structure

dc

Centerline distance between ties

dm

Maximum displacement

dmax

Maximum drift computed at a particular storey level

dy

Yield displacement

Energy

EB

Energy based damage index

Ef and Ei

Energies at failure and at ith cycle

Ef+ and Ef-

Failure energies in positive and negative deformation cycles

EI

Flexural rigidity

EIg

Effective stiffness ratio (gross)

EIy

Effective stiffness at yield

Eiy/Eig

Effective stiffness ratio

ELF

Equivalent lateral force procedure

Ep

Modulus of elasticity of shear wall

Ep+ an Ep-

Potential energies in positive and negative deformation cycles

Es

Modulus of elasticity of reinforcement

EV

Eigen value analysis

xli

EVCS ,EVGS

Eigen value analysis with gross and cracked stiffness

es, ev, em, e2

Stiffness, strength, mass eccentricity and dynamic eccentricity

eox,eoy

Eccentricities along longer and shorter plan width

F, F11

Force, Frame type with value of 1, 2 and 3 for masonry infill frame,
open first floor and bare frame

FBD

Force based design

FDR

Flexural damage ratio

Fe

Maximum force

Fh

Hysteresis force restoring component

FHC

Follower half cycle

Fi

Peak force at ith cycle

Fk

Force at stiffness k

Fkk

Linear force restoring component

FRF

Force reduction factor

Ft

Restoring force

Fu

Peak Strength/Yield force corresponding to ultimate deformation

Fy

Yield force

Fyt

Yield strength of tensile steel

fc

Concrete compressive strength

fy

Steel yield strength

GPC

Glaister and Pinho (2003) method for cracked stiffness

GPG

Glaister and Pinho ( 2003) method for gross stiffness

Acceleration due to gravity

xlii

Building height

Storey height

h(z)

Pinching function

HA

Hazus methodology

hb

Height of the beam

HC

Hazus method for cracked stiffness

HG

Hazus method for gross stiffness

HLsi

Equivalent building height at the corresponding to a specific


performance level

HO

High

hsw

Height of shear wall

HY

Hysteresis based

IDA

Inelastic dynamic analysis

IDGS, IDCS

Inelastic dynamic analysis results with gross and cracked stiffness

IDGS1,IDCS1

using method 1 and method 2

IDR

Interstorey drift

If

Importance factor

Ibb

Moment of inertia of beam

Ibb

Moment of inertia of column

Ig

Gross moment of inertia of column

Ir

Interstorey drift ratio

Ird

Maximum interstorey drift

JPDF

Joint probability density function

xliii

Stiffness matrix/ Stiffness

KC

Post-capping stiffness for use in Ibarra element model

Kdam

Damaged stiffness

Ke1

Secant stiffness through the yield point, for use in Ibarra element model

Ki

Initial stiffness

Area ratio of infill wall to total panel

Area ratio of infill panels to shear wall

Area ratio of infill walls to total panels

KD22 and KE22

Stiffness at the hysteresis load paths

ke,ki,k2

Elastic stiffness, post yield and post peak stiffness

Km

Maximum stiffness

Ko

Initial stiffness

KS

Hardening stiffness for use in Ibarra element model, i.e. stiffness


between y and cap,pl

Ku

Unloading stiffness

Kund

Undamaged stiffness

Building width

Lf

Flange buckling wavelengths and web buckling wavelengths

LO

Low ductile frames

ls

Radius of gyration

Ls/H

Shear span ratio

LS1, LS2, LS3

Performance levels

Lw

Web bucking lengths

xliv

Lws

Length of shear wall,

Mass /Mass matrix

M 200 M1000

Mass irregularity with a magnitude ranging from 200 % to 1000 %

M1, M2

Mass irregularity from 200% and 400 % and from 600 %, 800 % and
1000 %

MC

Mass irregular building model wih cracking effects

Mac

Actual (deteriorated) value of the yield moment (force)

Mc

Moment at capping point

Mi, Ma

Mass of Ith storey and the storey adjacent to ith storey

Mk

Modal mass

Mm

Maximum bending moment

Mmm

Total modal mass

Mp

Moment at pivot

Mk

Modal mass of Kth floor

My

Moment at yield

Msy

Flexural strength

Mt

Total modal mass of the building

My0

Moment at the yield point in the theoretical skeleton curve

Mu

Magnitude of earthquake

MFDR

Maximum flexural damage ratio

MO

Medium

MT

Model name

MS

Irregularity at the middle storey

xlv

MSM1,MSS1,MST1

Mass, stiffness and strength irregularity at middle storey

M.O.I

Moment of inertia

Parameter to account for variation in ground motion spectral shape

Number of storeys

NCI

Non cumulative index

NCR

Normalized cumulative rotation

NDU

Non -ductile frame

Nu

Axial load in Figure 2.21

Number of cycles

nI

Number of cycle at ith iteration

nf

Number of cycles at failure

ns

Number of stories in the building model, and nb represents the number


of bays in the first storey of the building model

Oa

Open area in diaphragm and diaphragm stiffness,

OMRF

Ordinary moment resisting frames

Plan irregularity

Pa

Axial load

Pc

Mean probability of collapse

P1, P2, P3

Models with plan irregularity

P1C, P2C,P3C

Plan irregular building models with cracking effects

PO

Post capping deterioration

Pc

Probability of collapse

Py and pc

Yield strength and strength at crack

xlvi

Pk

Modal participation factor at ith mode

Pi and Pr

Combinations of participation factor from jth to the kth mode for irregular
and regular buildings for long period structures

PBD

Performance based design

PH

Plastic hinge

PHC

Primary half cycle

PGA

Peak ground acceleration

PGV

Peak ground velocity

PEER

Pacific earthquake engineering research center

PLsi

Probability of failure corresponding to demand

Behaviour factor for gross stiffness

qc

Behavior factor for cracked stiffness

Regular building model

R2

Correlation coefficient

RC

Reinforced concrete

Ri

Re-entrant corner projection limit

Rr

Response reduction factor

RSD

Residual strength

RTR

Record to record variability

Constant/beam column strength ratio in Appendix section

Ratio of positive to negative displacements in Chapter 5

r1

Parameter that controls abruptness of stiffness

r1

Beam column strength ratio

xlvii

r22

Constant

rki

Modified stiffness

rx , ry

Torsional radius in x and y direction.

Strength capping

SC

Stiffness irregular building model with cracking consideration

S 25 S100

Stiffness irregularity with a magnitude ranging from 25 % to 100 %

S1 and ST1

Stiffness and strength irregularity of magnitude 25 %, 50 % respectively

S11

Ratio in percentage of shear walls to total floor area

S11

Ratio of percentage of shear wall to total floor area

S2 and ST2

Stiffness and strength irregularity of magnitude 75 % and 100 %)

Sa1

Stirrup spacing

SB

Shear beam model

SBa

Setback irregularity in adjacent storey

SBi

Setback irregularity at ith storey

SCA

Strength capping

Sdd

Spectral displacement at ith mode of vibration

Sdg

Global damage index

Sdl

Local damage index

Sds

Storey damage index

Sdst

Diaphragm stiffness

Se/Ag

Spectral acceleration

SE

Setback irregularity

SEA,SEB,SEC

Building model with setback originating from bottom one third, top one

xlviii

third and top one third storeys, SEAC, SEBC, SECC denotes the
respective models with cracking consideration

SF

Scale factor

Si, Si+1,

Stiffness of storey I and of the storey above

SMRF

Special moment resisting frames

SRSS

Square root of sum of square rule for modal combination

SS

Single storey models

ST 25 ST100

Strength irregularity with a magnitude ranging from 25 % to 100 %

STC

Strength irregular building models with cracking consideration

Constant defined by Stephens and Yao (1987)

sn

Rebar buckling coefficient

Fundamental time period

Tb

Building time period

Tc

Critical time period

Tdam

Damaged fundamental time period

Te

Elastic period

TLsi

Building period corresponding to demand

TS

Top storey

Tundam

Undamaged fundamental time period

Tw

Thickness of shear wall

Ty

Yield period

Buckling amplitude of fiber

UO

Unloading strength deterioration

xlix

uig

Ground motion intensity

Reduction factor proposed by EC 8

Vb

Base shear

Seismic weight

Weighing parameter in Chapter 5

Unit floor weight including tributary shear wall weight

wi

Weight function

Wk

Weight of kth floor

Xi

Damping ratio in ith mode of vibration and is = ci*/2Mi*i

Represents that the adopted property has ben adopted in the model
and N represents the reverse case

ya

Maximum web buckling amplitude occurs,

Location of the center of rotation of the plastic hinge mechanism

Total displacement in Equation

(s)

Strength degradation parameter


Total displacement in BWBN model

Z
n

n 1

Z, Z

Zu

Displacement in BWBN model in n and n -1 cycles


Ultimate displacement in Bouc and Wen model

Greek symbols

Parameter representing stiffness degradation

Rate of strength deterioration

11, 13, 14 23

Constant

Parameter depending upon stiffness and strength deterioration

Constant

Constant

Parameter representing frequency content of ground motion/Ratio of


building time period and critical time period in Appendix C.

Parameter depending upon stiffness and strength deterioration

Constant

22

Hysteric shape parameter

Irregularity index for long period structures

Constant

Constant

Irregularity index for short period structures

22

Parameter representing hysteric shape

Parameter representing pinching effects and yield base shear


coefficient in Chapter 5

Parameter depending upon stiffness and strength deterioration

(i)ort

Average reduced storey drift of ith storey of building

+, -

Deformation at positive and negative half cycles

pu py

Deformation as described in Figure 2.19

li

Yield - line mechanism rotation corresponding to beam plastic hinge


rotation

Lf I

Flange displacement

Lw

Displacement at location

Deformation during positive cycle

11

Correction factor to fundamental time period

11

Correction factor to fundamental time period

Deformation at failure

Deformation at ith cycle

Maximum deformation

Deformation in BWBN model

Ultimate deformation

Deformation at yield

Maximum deformation under earthquake load

Strain in the column

Reinforcement ratio

Ratio of area of longitudinal reinforcement in compression to that in


tension

11min

Ratio of minimum shear wall area to the total floor area

al11

Ratio of short side shear wall area to the total floor area

as11

Ratio of short side shear wall area to the total floor area

Ratio of shear wall area to floor plan area for walls aligned in the direction
in which period is calculated

lii

Stress demand

Model

Standard deviation due to modeling uncertainty

RTR

Standard deviation due to record to record variability

Total

Standard deviation (total)

(s)

Stiffness degradation parameter

Constant

Constant

22

Parameter representing strength degradation

ci

Stiffness Irregularity factor defined at ith storey of building

ki

Strength Irregularity factor defined at ith storey of building

s and b

Parameters representing number of storeys and number of bays as


defined by Karavasilis et al. (2008a)

Rotation

+p FHC

Plastic rotation (positive) at follower half cycle half cycle

+p FHC

Plastic rotation (negative) at follower half cycle half cycle

+p PHC

Plastic rotation (positive) at primary half cycle

Initial rotation

ca

Capping rotation

cappl

Plastic rotation capacity post capping rotation capacity

Maximum rotation

-p phc

Plastic rotation (positive) at primary half cycle

p+ and p-

Potential rotation at positive and negative deformation cycles

pc

Post capping rotation capacity

liii

ua

Ultimate rotation

Yield rotation

y, u,

Curvature, yield curvature, unloading curvature, normalized curvature

y1, n1 and n2

The angles of inclined yield lines with respect to the cross section on
which axial deformation is applied

s , b

Indices proposed by Karavasilis et al. (2008a)

Cyclic energy dissipation capacity

hm

Mean relation factor correlating Ibarra et al. (2005) and Dutta and Das
(2002) model

1, 11

Constant

irr

Mean relation factor correlating elastic and inelastic seismic response

Ductility ratio

13

Constant

Maximum rotational demand

Maximum ductility ratio

Initial ductility ratio

Initial ductility ratio

rd

Maximum roof displacement

Over - strength factor

Natural frequency of building, regular building model, irregular building


model

ir,, irr,,

Modal combinations of frequency of vibration of the irregular and regular


building frames from mode 1to mode k.

liv

Axial load ratio and is equal to Pa/Agfc

v(s)

Stiffness degradation parameter

Strength degradation parameter

lv

CHAPTER 1
INTRODUCTION

1.1 Structural irregularity in buildings


The component of the building, which resists the seismic forces, is known as lateral force
resisting system (L.F.R.S). The L.F.R.S of the building may be of different types. The
most common forms of these systems in a structure are special moment resisting
frames, shear walls and frame-shear wall dual systems. The damage in a structure
generally initiates at location of the structural weak planes present in the building
systems. These weaknesses trigger further structural deterioration which leads to the
structural collapse. These weaknesses often occur due to presence of the structural
irregularities in stiffness, strength and mass in a building system. The structural
irregularity can be broadly classified as plan and vertical irregularities.
A structure can be classified as vertically irregular if it contains irregular distribution
of mass, strength and stiffness along the building height. As per IS 1893:2002, a storey
in a building is said to contain mass irregularity if its mass exceeds 200% than that of the
adjacent storey. If stiffness of a storey is less than 60% of the adjacent storey, then a
storey is termed as weak storey. If stiffness of a storey is less than 70% or above as
compared to the adjacent storey, then the storey is termed as soft storey. In reality,
many existing buildings contain irregularity, and some of them have been designed
initially to be irregular to fulfill different functions e.g. basements

for commercial

purposes created by eliminating central columns. Also, reduction of size of beams and
columns in the upper storeys to fulfill functional requirements and for other commercial
purposes like storing heavy mechanical appliances etc. This difference in usage of a
specific floor with respect to the adjacent floors results in irregular distributions of mass,
stiffness and strength along the building height. In addition, many other buildings are
accidentally rendered irregular due to variety of reasons like non-uniformity in
construction practices and material used. The building can have irregular distributions of
mass, strength and stiffness along plan also. In such a case it can be said that the
building has a horizontal irregularity. The detailed classification of structural irregularity is
presented in Figure 1.1 and code limits have been shown in Table 1.1 and Table 1.2.
From review of code limits it can be clearly said that majority of the codes prescribe
similar guidelines for the irregularities based on magnitude ignoring the aspect of
irregularity location which is unrealistic (discussed in detail in the next chapter).

This page is intentionally left blank

Although irregular buildings are preferred due to their functional and aesthetic
considerations is evident from examples of realistic existing irregular buildings is shown
in Figures 1.2 to 1.4. The past earthquake records show poor seismic performance of
these structures during earthquakes as discussed in the next section. The different types
of irregularities are presented from Figures 1.5 to 1.8.

Figure 1.1 Classification of different types of structural irregularities

Figure 1.2 The New Yorker hotel New York City, USA (wikipedia.org, Daniel
Schwen)

This page is intentionally left blank

(a)

(b)
Figure 1.3 Manhattan city showing different irregular buildings: (a) View 1
wikepedia.org, Martin Durrscanabel, (b) View 2 (Wikemedia, Hakilon)

This page is intentionally left blank

(a)

(b)

(c)
Figure 1.4 Pictures of existing irregular buildings: (a) Setback building in New
Delhi (India) (/http://www.quakesafedelhi.net, (b) Empire state building (wikepdia,
RadioFan BigMac, (c) 500 fifth (http://wikipedia.org, Gryffindor)

This page is intentionally left blank

Table 1.1 Irregularity Limits prescribed by IS 1893:2002, EC8:2004, UBC 97, NBCC
2005
Type of
Irregularity

IS 1893:2002

EC8 2004

UBC 97

NBCC 2005

Ri 5%

Ri 15%

dmax 1.2 d avg

rx > 3.33 eox


ry > 3.33 eoy
rx and ry > ls,

dmax 1.2 d avg

dmax 1.7 d avg

Oa > 50%
Sdst > 50%

rx2 > ls2 + eox2


ry2 > ls2 + eoy2

Od > 50%
Sdst > 50%

Horizontal
Re-entrant
Corners
(Figure1.5)

Torsional
irregularity

Diaphragm
Discontinuity

Ri 15%
(Figure 1.5 )

Vertical
Mass
(Figure 1.6)

Mi < 2 Ma

Should not
reduce abruptly

Mi < 1.5 Ma

Mi < 1.5 Ma

Stiffness
(Figure 1.7)

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

(a) Soft Storey

(b)Weak Storey

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

Si < 0.8Si+1

Si < 0.7Si+1
or
Si < 0.8 (Si+1 +
Si+2 + Si+3)

Si < Si+1

Si < 0.8Si+1

SBi < 1.3 SBa

SBi < 1.3 SBa

Setback
irregularity
(Figure 1.8)

SBi < 1.5 SBa

SBi < 0.3. plan


dimension of
the first storey

Table 1.2 Irregularity Limits prescribed by IBC 2003, TEC 2007 and ASCE 7.05
Type of Irregularity
Horizontal

IBC 2003

TEC 2007

ASCE 7.05

Re-entrant
Corners

Ri 20%

Ri 15%

Torsional
irregularity

dmax 1.2 d avg

dmax 1.2 d avg


dmax 1.4 d avg

Diaphragm
Discontinuity

Oa > 33%

Oa > 50%
Sdst > 50%

Mass

Mi < 1.5 Ma

Mi < 1.5 Ma

Stiffness

Si < 0.7Si+1
or
Si < 0.8 (Si+1 + Si+2
+ Si+3)

(a) Soft Storey

Si < 0.7Si+1
or
Si < 0.8 (Si+1 + Si+2
+ Si+3)

Vertical

(b) Weak Storey

Si < Si+1

Setback
irregularity

SBi < 1.3 SBa

ki = (i / hi) avr /
(i+1 / hi +1) avr
> 2.0 or

ci = (Ae)i / < 0.80

Si < 0.7Si+1
or
Si < 0.8 (Si+1 + Si+2 +
Si+3)

Si < 0.7Si+1
or
Si < 0.8 (Si+1 + Si+2 +
Si+3)
Si < 0.6Si+1
or
Si < 0.7 (Si+1 + Si+2 +
Si+3)
SBi < 1.3 SBa

where ( Ae ) i - Area of the storey, (Ae)i+1 - Area of the storey above ith storey, Si, Si+1,

Si+2- Stiffness of ith, i +1th and i + 2th storey,

SBi -Setback irregularity limits, Oa, Sdst th

Open area in diaphragm and diaphragm stiffness, Mi,Ma - Mass of i storey and the
storey adjacent to ith storey, dmax,davg - Maximum drift computed at a particular storey
level, Average of drifts computed at both sides of a structure, ls- Radius of gyration, ci,

ki - Stiffness irregularity factor defined at ith storey of building and stiffness irregularity
factor defined at ith storey of building.

10

rx,ry - Torsional radius in x and y direction, i, (i)ort - Reduced storey drift of ith storey of
building and average reduced storey drift of ith storey of building, Ri - Re-entrant corner
projection limit.

Figure 1.5 Re-entrant corner irregularity

Figure 1.6 Mass irregularity

Figure 1.7 Irregular distribution of stiffness in the building system

11

(a)

(b)

(c)

Figure 1.8 Vertical setback irregularity

1.2 Failures occurred during the past earthquakes


A building structure may collapse or suffer severe damage under the action of seismic
forces due to sudden change in mass, stiffness and strength along vertical or a
horizontal plane. As discussed in the previous section, presence of structural
irregularities triggers the structural collapse. The next subsections give examples of
building failures that have occurred during the past earthquakes due to presence of
different types of structural irregularities.

1.2.1 Failure due to vertical irregularities


The vertical irregularities can be sub - classified into mass, stiffness, strength and
setback irregularity. The Olive medical center was one of the building which failed during
San Fernado earthquake in 1971. It was a six-storeyed building with mass irregularity in
the form of excess earth fill at the first storey. Furthermore, structural walls were present
at the second floor level which resulted in stiffness and strength irregularities in the
second storey. In addition several columns in the ground storey contained inadequate
lateral confinement. Therefore, the first two storeys of the building which were critical for
the building stability contained irregularities of mass, strength and stiffness. After
occurrence of earthquake it was observed that the first two storeys which supported the
whole building incurred heavy damage, and in contrary the upper four floors sustained a
very less damage (Figure 1.9). A similar collapse of a four storey building was observed
due to open first storey during Northridge earthquake (1994) as shown in Figure 1.10.
The soft-storey effect was the main reason for collapse of many multi-storey R/C
buildings during the earthquakes that occurred in Turkey during the last decade (Adalier
and Aydingun 1998; Durumus et al. 1999; Huang and Skokan 2002; Sezen et al. 2003;
Eyidogan et al. 2003; http:// bingol.meb.gov.tr/index.html).. The majority of the residential
and commercial buildings built in Turkey had soft storeys at the first-floor level which

12

were often used for commercial purposes. These storeys were generally enclosed with
glass windows instead of brick infill walls so as to be used as showrooms. The heavy
masonry infills starting immediately above the soft storey which created a large variation
of mass, stiffness and strength in the bottom storeys. The previous earthquake damages
and results of analytical studies showed that the structural systems with a soft storey led
to serious problems during severe earthquake ground shaking. During the occurrence of
an earthquake, the presence of a soft storey increased the deformation demands
significantly and the first-storey columns were expected to dissipate the whole seismic
energy. Many failures and collapses can be attributed to the increased deformation
demands in conjunction with poorly designed columns. Figures 1.10 to 1.14 shows such
failures and damages during different earthquakes. The soft storey has been one of the
major reasons of damage throughout the world during earthquakes as evident from
seismic reports (EQE,1995; Youd et al. 2000; Yoshimura 2003). Therefore, it is
prescribed to avoid sudden change of mass, stiffness and strength along the building
height especially at the bottom storey.

Figure 1.9 Failure of Olive View Medical State Center (Moehle and Mahin, Nisee
Berkeley)

13

This page is intentionally left blank

14

Figure 1.10 Bending of columns and failure of parking structure during Northridge
earthquake (http://www.ngde.noaa.gov/seg/hazard/slideset/earthquake)

Figure 1.11 Failure of buildings due to soft storey in Turkey during Bingol
earthquake 2003 (Dogangun 2004)

15

This page is intentionally left blank

16

(a)

(b)

Figure 1.12 Failure due to weak storey: (a) Garage formation at base floor
(Durrani et al. 2008; Steinbrugge http://nisee.berkeley), (b) Weak storey failure
during Kashmir earthquake (Kirac et al. 2011), (c) Failure of weak storey with
overhang (Kirac et al. 2011)

17

This page is intentionally left blank

18

Figure 1.13 Failure of buildings due to soft storey in Turkey during Sumatra
earthquake 2004 earthquake (Kirac et al. 2011)

Figure 1.14 Failure of buildings due to soft storey in Turkey during Izmit 1999
earthquake (Kirac et al. 2011)

1.2.1.1 Failure of setback structures


The setback structures are quite common in modern buildings as evident from
existing realistic buildings. The failiure of structures during previous earthquakes have
been reported by researchers (Penelis et al. 1978; Spyropoulos 1979; Kam et al. 2010;
Fraser et al. 2013) as shown from Figures 1.15 - 1.20. This section describes the
particular behavior of TO 9 building during Chile earthquake 2010.

19

This page is intentionally left blank

20

The internal inspection of the building did not show significant damage throughout the
building height. Building TO-9 suffered its main collapse at the twelfth storey, where its
east section collapsed and dragged the upper storeys downwards (Figure 1.16 b, c).The
north and west facades (Figure.1.15 a) experienced very little damage as compared to
the east and south facades (Figure. 1.15b, c). In addition to the main collapse which
occurred in storey 12, two other partial collapses were observed in storeys 16 and 20,as
evident from the east elevation (Figure. 1.15b). A close up of the main collapse at storey
12 (axis J) is shown in Figure 1.15e (axes in story 12 is defined in Figure. 1.16 (a). The
collapsed region shown is between axes 4 and 7, and G and J. The slabs crushed on to
the ceiling of the 11th storey. Figure1.16e shows that the slab between axis 4 and 7
dropped one complete storey downwards. In comparison, the slab between axis 9 and
10 dropped downwards by half storey. Columns were observed to be completely
collapsed as shown in Figure. 1.16 e and 1.16f. All other structural elements were
deformed as observed from Figure. 1.16d. Moreover, the wall thickness in axis J
changed from 25 to 20 cm in the 12th storey (Westenenk et al. 2013). At axis J, upper
storeys displaced towards west by 13 cm (Figure. 1.16c). Storeys above the 12th storey
experienced a partial collapse due to presence of stiffer walls as shown by blue lines in
Figure. 1.16a. These walls observed light damage. The west elevation (axis C) had
negligible damage as evident from Figure. 1.16a. Tensile cracks due to torsion were also
observed in the wall located along axis C (refer to the close-up figure to the top left of
Figure. 1.16). The slabs played an important role in collapse prevention as they held the
collapsed regions and core walls together which prevented a larger collapse. Figure.
1.15g shows the partial collapse of storeys 20 and 21 from which it could be seen that
columns farther from the core of shear walls were severely damaged. In addition, the
slab partially collapsed due to loss of vertical support along axes J and G. In general,
building damage was observed to increase with height. This fact contradicts the capacity
design philosophy propagated by EC8:2004 in which plastic hinges should occur at the
first storeys. Little to no damage was observed between storeys 2 and 4; severe damage
occurred in a single wall in the 5th storey; and storeys 6 to 12 showed increasing damage
in height; with a main collapse in the 12th storey. Local damage to structural elements
such as slabs and columns increases in upper storeys; damage of storeys 16, 18 and 21
is shown in Figure. 1.17 a, b and c, respectively. Based on judgment all these types of
irregularities and setbacks in the building played a crucial role in its collapse (Westenenk
2013). Moreover, failure of 21 storeyed building at midheight due to presence of setback
was observed during Chile earthquake 2010 (Figure 1.18).The setback presence in the
form of projected balconies are also one of the major reasons of failure during
earthquakes (Figure 1.18 to 1.20).

21

This page is intentionally left blank

22

Figure 1.15 Overview of building TO-9: (a) West side view, (b) East side view, (c)
South side view and close-up of deformed axis J, (d) Plan view, (e) Details of the
main collapse at story 12, (f) Damage in short columns in axis 1A, (g) Overall
condition of the 21st storey (Westenenk 2013)

23

This page is intentionally left blank

24

Figure 1.16 Internal damage of building TO-9: (a) Structural plan of the 12th storey,
(b) Partial collapse of the 16th storey, (c) Partial collapse of the 20th storey, (d)
damaged beams of the 12th storey, (e) Column in axes G-5 of the 12th storey, (f)
Column in axes G-4 (Westenenk 2013)

(a)

(b)

(c)

Figure 1.17 Internal damage of building TO-9: (a) Slab deformation in storey 16, (b)
Overview of damage in storey 18, (c) Severely damaged column in storey 21
(Westenenk 2013)

25

This page is intentionally left blank

26

Figure 1.18 Failure due to overhangs (Kirac et al. 2011)

Figure 1.19 Failure due to overhangs (Dogangun 2004)

Figure 1.20 Failure of balconies at Shizugawa Hospital during Tsunami 2011


(Fraser et al. 2013)

27

This page is intentionally left blank

28

1.2.2 Failure of plan irregular buildings


Damage to irregular structures caused by asymmetry in plan has been observed during
many major and minor earthquakes during the past. The non-coincident centers of mass
and stiffness in a structure generate plan asymmetry which causes torsional vibration
resulting in severe damage to structural components in the more laterally flexible regions
of the structure. The building in Figure 1.21 shows failure of a three-storey reinforced
concrete building from the Miyagi-Ken-Oki (Japan) earthquake in 1978 due to torsion.
Due to presense of a stiff wall, the center of stiffness shifted towards the wall. This
resulted in twisting of building with respect to the center of stiffness. This was due to
occurrence of torsion generated by the eccentricity between the centers of mass and
stiffness. The torsion resulted in severe damage of columns along the periphery away
from the wall. Figure 1.22 shows the Ministry of Culture building which was damaged
due to torsion during the Haiti earthquake in 2010. The presence of stiff core area on
one side of the building resulted in torsion which led to damage of lateral load-resisting
members away from the center of stiffness. Due to failure of these members, the whole
storey experienced a downward pull which led to the total collapse of the building.
Likewise, the six-storey reinforced concrete hotel in Guatemala City (Figure 1.23) had
torsional irregularity due to the eccentric location of a rigid service core area. This
building experienced a severe damage during the Guatemala earthquake in 1976. The
columns of the building located on the flexible side failed due to their incapability to
resist the shear force increase due to the torsion. This resulted in second storey collapse
in the building (indicated by blue arrows in the figure). Moreover, severe damage and
collapse of the structures due to symmetric layouts of lateral load-resisting members
because of torsional effects were reported in previous literature works (Wyllie et al. 1986;
Anderson 1987; Elnashai at el. 2010).
As discussed, the irregular structures rendered poor seismic performance, and this
is mainly due to ignorance of the irregularity aspect in formulating the seismic design
methodologies by the seismic codes (IS 1893:2002, EC8:2004, UBC 1997, NBCC 1995,
NBCC 2005 etc.). Nevertheless, ignoring the irregularity aspect may induce noticeable
errors in estimation of the seismic response parameters. Moreover, a structural engineer
requires a good understanding regarding behavior of these irregular systems. However,
most of the past research works mainly focus on the regular systems. Therefore, there
is a need of a comprehensive evaluation of effects of different types of irregularity on the
seismic response parameters to formulate improved design philosophy for these
structures.

29

This page is intentionally left blank

30

(a)

(b)
Figure 1.21 Damage from irregularity during 1978 Miyagi Ken Oki Japan
earthquake (Ellingwood 1980): (a) Overall view of damage of three storey
reinforced concrete buildings, (b) Damage to column along periphery along the
wall (Ellingwood 1980)

31

This page is intentionally left blank

32

(a)

(b)

(c)

Figure 1.22 Damage from Irregularity during 2010 Haiti earthquake: (a) Overall view
of damage to Ministry of Culture building, (b) Damage on flexible side (c) Damage
on stiff side (Mid America Earthquake Engineering Center, University of Illiniosis
at Urbana Champaign)

33

This page is intentionally left blank

34

(a)

(b)

(c)
Figure 1.23 Damage due to irregularity during Guatemala earthquake 1976: (a)
Overall view of a Hotel terminal building in Guatemala, (b) Collapse of second
storey due to shear failure of column, (c) Close - up of one of the collapsed
column (Bertero 1997)

35

This page is intentionally left blank

36

1.3 Brief discussion on strong column weak beam design


philosophy
This philosophy advocates the use of stronger columns as compared to the beams.
From the review of building statistics in developing countries like India it was observed
that relatively deep beams with flexible columns were used frequently. When frame
system buildings have strong beams and weak columns, the beams remain elastic and
the columns fail due to shear failure or compression crushing during the seismic
excitation. This kind of failure was very common in the recent strong or moderate
earthquakes during Northridge, Bingol and Bhuj earthquake (Jain et al. 2002; Goel and
Chopra 2005) as shown from Figures 1.24 to 1.26. To avoid such failures, and adequate
ductile detailing measures have been prescribed by different seismic design codes.

Figure 1.24 First storey column damage due to weak column strong beam
philosophy (Moehle and Mahin, Nisee Berkeley ACI SP -127)

(a) Dormitory building

(b) School Building

Figure 1.25 Strong beam weak column structural system of the Celtikuyu school
and dormitory: (a) Dormitory building, (b) School building (Dog angun 2004)

37

Figure 1.26 Strong beam weak column structural system of the Panta parking
school at Banta Aceh (Dog angun 2004)

1.4 Review of research works regarding plan irregularities


Assessment of the performance of building structures during past earthquakes suggests
that plan irregularities are one of the important causes of damage during occurrence of
an earthquake. Plan irregularity may occur due to irregular distribution of mass, stiffness
and strength along the plan. In past years lot of research effort has been done to study
the behavior of plan asymmetric buildings during seismic excitation (Tso and Myslimaj
2003; Tso, and Bozorgnia 1986; Tso and Sadek 1985; Tso and Sadek1989)

1.4.1 Single storey building models


Earlier studies investigated the torsional effects on plan irregular building systems with
single storey building models. One of the main reasons for adopting single storey models
was their simplicity. These models were used to determine the influence of torsion on
seismic response parameters and these results were further used to formulate design
methodologies for plan irregular building systems. However in recent years multistorey
building models are used to determine the realistic inelastic torsional response of plan
irregular building systems. But due to complexities, the use of multistorey building
models is limited and it is one of the major reasons that single storey building models are
still preferred by many researchers (Ladinovic 2008; Lignos and Gantes 2005; Luchinni
et al. 2011). Previous researchers on plan irregularities using single storey models
mainly focused on variation of positions of CM (Center of mass) or CS (Center of
stiffness) with respect to each other to create eccentricity. The Main aim was to
determine the torsional response of building systems due to eccentricity. To create
eccentricity some researchers varied position of CS or CR keeping position of CM

38

constant, the eccentricity generated in this case was called as stiffness eccentricity (es).
Some researchers varied position of CM keeping position of

CS as constant, the

eccentricity generated in this case was called as mass eccentricity (em) [Tso and
Myslimaj 2003]. Differing from earlier approaches some researchers created difference
in strengths of resisting elements to vary position of center of strength (CV) with respect
to CM, and the eccentricity generated was known as strength eccentricity (ev) . The
definitions of eccentricity have been described pictorially in Figure 1.27.

(a)

(b)

(c)
Figure 1.27 Types of eccentricity: (a) Mass eccentricity, (b) Stiffness eccentricity,
(c) Strength eccentricity
Research works on plan irregular building systems started in early 1980s with Tso
and Sadek (1985) who determined the variation in ductility demand by performing
inelastic seismic response of simple one storey mass eccentric model with stiffness
degradation using Cloughs stiffness degradation model and bi-linear hysteric model.
Results of analytical study showed that the time period had predominant effect on the
ductility demand after the elastic range. The comparison of results showed a 20 %
difference in the results obtained between Cloughs and bilinear model. Irregular
distributions of strength and stiffness are one of the major causes of failures during the
earthquakes. Both of these irregularities are interdependent and to study the effect of
these irregularities on seismic response, researchers like Tso and Bozorgnia (1986)
determined the inelastic seismic response of plan asymmetric building models (as

39

described in Table 1.3) with strength and stiffness eccentricity using curves proposed by
Tso and Dempsey (1990). Results of analytical study showed the effectiveness of the
curves proposed by Tso and Dempsey (1990) except for torsionally stiff structures with
low yield strength.
Sadek and Tso (1989) performed inelastic analysis of mono-symmetric building
systems with strength eccentricity as described in Table 1.3. The center of strength was
defined in terms of yield strength of resisting elements. From analytical studies it was
found that the code defined eccentricities based on stiffness criteria were useful in
predicting the elastic seismic response. However, in inelastic range parameter of
strength eccentricity was found to be useful in determining seismic response.
Pekau and Guimond (1990) checked the adequacy of accidental eccentricity to
account for the torsion induced due to the variation of strength and stiffness of the
resisting

elements

which

was

achieved

using

elasto-plastic

force-deformation

relationship. Results of analytical study showed occurrence of torsional amplification due


to strength and stiffness variation. Finally, the code prescribed provision of 5% for
accidental eccentricity was found to be inadequate.
Table 1.3 Descriptions of different models adopted by Tso and Sadek (1989)
S.No

Model

Description

Name
1

Me

Mass eccentric model with all three resistant elements having


equal yield deformation

Se1

Stiffness eccentric Model with identical yield strength.

Se2

Stiffness eccentric Model with identical yield deformation.

Duan and Chandler (1991) based on their analytical studies on plan irregular
buildings, systems the change in design eccentricity

in Mexico code 87 was

recommended as 1.5es + b and 0.5es - 0.1b as compared to the earlier value of es 0.1b
and es 0.05b.
Chandler and Hutchinson (1992) determined the effects of torsional coupling on one
storey stiffness eccentric building systems. From the results of analytical studies, a
strong dependence of torsional coupling effects on natural time period of the structure
was observed. Also, the effectiveness of torsional design provisions as prescribed by
different codes of practice (ATC 3-06, NEHRP, NBCC 90, and EC8:1989) was
determined by conducting elastic and inelastic analysis on one storey stiffness eccentric

40

building systems. The code evaluation results obtained for asymmetric building system
as per different codes have been shown in Table 1.4 and Table 1.5. Results of analytical
study showed greater displacement of flexible edge as compared to stiff edge.
Table 1.4 Code Evaluation results by Chandler and Hutchinson (1992)
S.No
1

Code
NEHRP

Results
Inadequate for building systems with small and moderate eccentricity.
Satisfactory results for building systems with large eccentricity.

ATC

Same as NEHRP.

NBCC

Inadequate for buildings with low time periods ( T 0.5S )


Over-conservative for higher time periods at all eccentricities.

EC8

Conservative for small eccentricity.


Over conservative for medium to large eccentric buildings system with
higher time periods.

Table 1.5 Results obtained by Chandler and Hutchinson (1992)


S.No

Code Name

Results

NZS

Conservative Estimate of displacement

UBC

Conservative Estimate of displacement for DAF / FRF 1

NBCC

Conservative Estimate of displacement for DAF / FRF 0.6 1.0

Chandler et al. (1995) verified the torsional provisions prescribed by different codes
of practice. For analytical study, two types of building models namely torsionally
balanced (TB) and torsionally unbalanced (TU) were considered. The torsional
unbalance in the building model was created by varying position of center of stiffness
inducing stiffness eccentricity equal to 0.05b. The torsionally unbalanced building models
were further divided into two types namely A1 and A2 with moderate and low torsional
stiffness. Results of analytical studies showed the variation in seismic response in
models A1 and A2 with flexible edge experiencing greater deformation as compared to
the stiff edge. The stiff edge of building systems with small time period (T < 1 Sec)
designed according to NZS 4203 and EC8:1989 experienced least additional ductility
demand. However the additional ductility demand was found to be largest for building

41

systems with T > 1 Sec. In case of TU systems designed according to EC 8 -1989 the
ductility demand exceeded by 2.5 % as compared to the TB system.
Ferhi and Truman (1996) determined seismic response of building systems with the
presence of stiffness and strength eccentricity. Both elastic and inelastic seismic
behavior were studied. From analytical study of the building systems it was observed
that the seismic response showed greater dependence on stiffness eccentricity in elastic
range. However, the effect of strength eccentricity on seismic response was observed to
be in the inelastic range.
Chandler and Duan (1997) developed an optimized procedure for determining the
seismic response of torsionally balanced and unbalanced structures. The parameters
like eccentricity (e), normalized radius of gyration (Pk), force reduction factor (R) and
uncoupled lateral period (Ty) were included in the proposed optimization procedure. The
authors proposed design eccentricity expression and over strength factor expressions
and compared it with code defined expressions. The codes used in the study were
UBC 94, EC8-94 and NBCC-95 . The analytical study was conducted both on torsionally
balanced (TB) and torsionally unbalanced (TU) models. Results of analytical study
showed that the over strength factor was found to be substantially lower as compared to
UBC-94 and NBCC-95 but higher than EC8 for entire range of Pk. However, the results
of proposed procedure are comparable to code defined procedures for torsionally
unbalanced structures (TU). The parameters e, pk, R, Ty considered in the design
procedure were found to influence the seismic response. Finally the procedure was
found to be applicable to single storey and multistorey torsionally unbalanced structures.
De-La-Colina (1999) studied the effects of torsion on simple torsionally unbalanced
building systems considering the earthquake components in two perpendicular
directions. The effects of following parameters were studied, (a) seismic force reduction
factor, (b) design eccentricity, (c) natural time period. The structural model used for the
analytical study is shown in Figure 1.28. Based on the results of analytical study it was
concluded that, with increase in the force reduction factor, the ductility demand reduced
for flexible element. Regarding the effect of time period it was found that for torsionally
unbalanced stiff elements the ductility demand increased with time period and vice versa
for torsionally unbalanced flexible elements. The increase in value of stiffness
eccentricities reduced the normalized ductility demand. Based on these results it was
concluded that strength eccentricity had greater effect on seismic response as compared
to stiffness eccentricity.
Ghersi and Rossi (2001) determined the influence of bi - directional seismic
excitation on seismic response of stiffness eccentric one storey building systems using
elastic and inelastic analysis. The seismic response from the inelastic analysis was

42

compared with the results of elastic analysis. Results of analysis showed that the
consideration of effects of
in

seismic

response.

bi - directional seismic excitation resulted in minor variation


Elastic

analysis

using

unidirectional

seismic

excitation

overestimated the seismic response.


De Stefano and Pintuchhi (2002) considered the phenomenon of inelastic
interaction between axial forces and horizontal forces in modeling of plan irregular
stiffness asymmetric building systems. Based on results of analytical study it was
concluded that consideration of interaction phenomenon between axial force and
horizontal force resulted in reduction of floor rotation by 20%.
Dutta and Das (2002) studied the seismic response of

single storey plan

asymmetric structures subjected to bi - directional seismic excitation. From analytical


study the authors proposed two hysteresis models as represented in Figure 1.29 (a, b).
These models accounted for strength and stiffness deterioration of RC structural
elements subjected to cyclic loading. From the results of analytical study it was found
that local deformation demands both at stiff and flexible edge showed variation when
strength deterioration was considered. The consideration of unidirectional seismic
excitation resulted in lower values of deformation demands at both flexible and stiff edge.
These results were found similar to Tso and Myslimaj (2002).
Tso and Myslimaj (2003) proposed a new approach called yield distribution based
approach for strength and stiffness distribution. For analytical study the authors modeled
a single storey structure with a rigid rectangular deck supported by two resisting
elements in X and five resisting elements in Y direction. The resisting elements were
modeled using elasto-plastic, the bilinear and Cloughs hysteresis models for force
deformation relationship. The authors proposed a design parameter on which location
of center of mass (CM), rigidity (CR), strength (CV) and yield displacement (CV) depend.
The models were subjected to dynamic analysis to determine the balanced CV-CR
location. From the results of analytical study it was found that the structure satisfied
balanced CV-CR location and had low torsional response when value of b lies between
zero and unity.
Fujii et al. (2004) suggested a simplified non-linear analysis procedure for plan
asymmetric structures with stiffness eccentricity modeled as SDOFs and MDOFs.
Results of analytical study showed that the torsionally stiff building systems experienced
greater oscillations in first mode as compared to the torsionally flexible building systems.
On comparison of responses of MDOF and SDOF models for TS and TF building
systems it was found that SDOF models were found to be applicable to torsionally stiff
building systems only. Finally, the proposed analysis procedure was found to be efficient
in determining the seismic response of TS building systems

43

Figure 1.28 Structural model considered by De-La colina (1999)

(a)

(b)
Figure 1.29 (a, b) Hysteresis model proposed by Dutta and Das (2002)
Moghadam and Aziminejad (2005) performed PBD (Performance based design) of
asymmetric structures. The researchers evaluated the seismic response of single storey
structures (code designed) with irregular configuration for optimizing mass, stiffness and
strength center configurations corresponding to different levels of plastic hinge
formations. The authors adopted the concept of balanced CV - CR location proposed by
Tso and Myslimaj (2003) to evaluate best performance level of the structure. Based on
the analytical study it was concluded that the best location of CV - CR (Center of

44

stiffness and Center of rigidity) depended upon the required performance level of the
structure and also on damage indices as shown in Table 1.6.
Shakib and Ghasemi (2007) have determined the effect of consideration of near fault
and far fault excitations on seismic response of different type of plan asymmetric
structures with stiffness asymmetry. Following Tso and Myslimaj (2003) who suggested
balanced CV-CR location to minimize rotational deformation, the authors suggested a
new approach to minimize rotational deformation. In the proposed approach in which the
strength distribution pattern is made equal to yield displacement distribution modified by
a parameter . From results of analytical study it was found that in case of near fault
motions when > 0, the displacement demand on stiff edges is greater as compared to
the flexible edges. In case of far fault motions when < 0, the displacement demands
were greater on flexible edges as compared to stiff edges.
Jarernprasert et al. (2008) determined the inelastic torsional response of single storey
plan asymmetric systems with stiffness eccentricity designed in accordance with IBC
2006 and Mexico city building code 2004. For analysis of this building model modal
analysis procedure was adopted. The effect of seismic excitation on following
parameters was studied, (a) ratio of uncoupled torsional to transitional frequencies, (b)
design target ductility, (c) elastic natural time period and normalized static eccentricity.
The researchers also proposed new reduction and amplification factor for these
parameters (a,b,c). From results of analytical study it was found that these parameters
(a,b,c) had large influence on the inelastic behavior of the building system. Regarding
the comparison of codes it was found that IBC 2006 code overestimated the design
forces at both flexible and stiff edge of building system whereas the Mexico city building
code overestimated the design forces at flexible side. The use of reduction and
amplification parameters leads to the ductility demands closer to target ductility demands
but the displacements computed were nearly four times to that of equivalent symmetric
structure.
Ladinovic (2008) represented inelastic seismic response of plan asymmetric
structures with stiffness and strength eccentricity in the form of base shear torque
surface (BST). The factors influencing BST surface were strength eccentricity, lateral
capacity, torsional capacity and distribution of strength along plan.
Aziminezad and Moghadam (2010) determined the effects of strength distribution
and configuration of strength, rigidity and mass on seismic response of one storey plan
asymmetric building system subjected to near field and far field ground motions. Models
with different values of yield displacement, strength and stiffness eccentricity were
considered as shown in Figure.1.30. The models were analyzed by dynamic nonlinear

45

analysis and from results of analytical study it was found that for torsionally flexible
building systems, the strength distribution had minor effect both for near field and far field
excitations. But seismic response of torsionally stiff building systems was largely
influenced by strength distribution. Regarding the modal periods it was found that modal
periods along X-axis had the maximum value as compared to other two modal periods
and ratio of lateral to torsional frequency was found to be greater in y direction. Further it
was concluded that the torsionally stiff building systems with balanced CV-CR location
performed better than other building models both in case of near and far field excitation.
Table 1.6 Different positions of centers of mass, stiffness, strength and
displacement for different values of
S.NO

Positions of C.M, C.V, C.D

Position of CV coincides with CD, strength distribution takes same


shape as yield displacement

0-1

Value of ev decreases position of CV starts shifting from CD towards CM.

Position of CV coincides with CM and position of CR is shifted towards


left of C.M at a distance equal to ed

<1

CR and CV shift towards left of CM.

Figure. 1.30 Models considered by Aziminejad and Moghadam (2010)

Luchinni et al. (2011) determined the nonlinear seismic response of single storey
building models with eccentricities in both directions with base shear torque procedure
and verified this approach using IDA analysis. For analytical study four types of building

46

models namely S1, S2, R1, and R2

were modeled. The S1 model was a one way

asymmetric system with es = 0.1bw.The model S2 was a two way asymmetric system with

es = 0.05bw in both directions. The model R1 contained uniform strength distribution in xdirection only whereas model R2 contained uniform strength distributions in both
directions. The results of analytical study showed that base shear torque surface was
efficient in predicting the location of center of rigidity. The seismic response predicted
was comparable with that of IDA analysis.

1.4.2 Multistorey plan asymmetric structures


In previous analytical studies on plan irregular structures, single storey models were
widely used due to their simplicity and their ability to clearly depict the effect of different
seismic response parameters. Most of the design criteria were formulated on the basis of
results obtained in single storey models. But several researchers proved that single
storey models resulted in inaccurate prediction of torsional response. The development
of powerful software tools has made modeling and analysis of multi-storey building
models much simpler. The multi-storey building models give realistic prediction of
torsional response. Although studies on plan irregular building models started in 1990s
(Killar and Fajfar 1997,2002; Moghadam 1998; but, Fajfar et al. (2002) were one of the
major researcher in this field who proposed a new method which was an extension of
N2 method. The proposed method was applicable to the realistic 3D building models. For
analytical study a eight storey R.C. building with structural walls was modeled. The mass
eccentricity was introduced in the building model by displacing center of mass in both
horizontal directions by 5% and 15% respectively. The results of proposed procedure
were compared with that of non-linear dynamic analysis. From comparison of results the
ability of proposed method to predict the seismic response of torsionally stiff structure
was justified. However, the method did not include the effects of lateral torsional coupling
and was found to be un-conservative as compared to the N2 method.
De-la-Colina (2003) made assessments of several code specified procedures
regarding analysis of procedures for multistorey building systems with mass and stiffness
irregularity subjected to bi - directional seismic excitation (EI Centro earthquake).
Analytical studies were carried out on several 5 storey buildings having mass and
stiffness eccentricity. Shear beam models were used by researchers to represent
resisting elements. Based on the code defined procedures the authors had found out the
optimal values of storey eccentricity.
Chopra and Goel (2004) proposed a new method based on extension of their earlier
method (Chopra and Goel 2002). In the proposed method the torsional amplification of

47

the structure was accounted for by application of the lateral forces in combination with
the torsional moments at each floor of the structure. The lateral forces and torsional
moments were obtained from the modal analysis of the structure. A comparison between
the results of the proposed method and non-linear dynamic analysis was made for
building systems with different uncoupled lateral to torsional vibration periods. From the
results of analytical study the accuracy of proposed procedure for symmetric structures
was verified. However the accuracy of proposed procedure decreases with the increase
in magnitude of torsional coupling which is due to the use of complete quadratic
combination (CQC) rule for modal combination.
Correlating with his earlier studies, Fajfar et al. (2005) again proposed a new
method based on N2 method. In the proposed method combination of modal responses
obtained from pushover analysis of 3D structures were made with the results obtained
from linear dynamic analysis. In the proposed procedure the displacements and
deformation distributions along height were controlled by N2 method and the magnitude
of torsional amplification was defined by the linear dynamic analysis.
Stathopoulos and Anagnostopoulos (2005) were one of the few researchers who had
made attempt to evaluate torsional response of realistic 3D structures by nonlinear
analysis (Both as per EC8:2004 and UBC 97). The authors conducted analytical studies
on realistic 3 storeyed and 5 storeyed RC framed buildings (with flexible and stiff edges)
subjected to bi - directional excitations. From the results obtained (multistorey structures)
it was found that the inelastic displacement was greater at flexible side as compared to
the stiff side. However, the results obtained in case of single storey structures were
contradictory to the results obtained in case of multistorey structures with mass
irregularity under the action of

bi - directional seismic excitation. It was observed that

the torsionally stiff building systems undergo less plastic deformation as compared to the
torsionally flexible building systems. These findings contradict the results obtained from
single storey models.
Penelis and Kappos (2005) proposed a method to determine the inelastic torsional
response of plan asymmetric single storey and multistorey structures. The models used
for analytical studies were single degree of freedom (SDOF) systems and incorporated
the effects of torsional and translational modes. In the proposed method the spectral
load vectors were obtained from the elastic spectral analysis and these load vectors
were applied on the structure to carryout 3D pushover analysis. The results of the
proposed procedure were compared with that of non-linear dynamic analysis. It was
found that the inelastic seismic response obtained by both methods vary by 10% in case
of single storey structures and by 20 % in case of multistorey structures.

48

Marusic and Fajfar (2005) determined the elastic and inelastic seismic response of five
storey steel framed structure with mass eccentricity. The eccentricities were taken as
5%, 10% and 15% of the plan dimensions. For analytical study, three types of building
models were adopted as described in Table 1.7. For the building model the first storey
height was kept as 4m and other storey heights as 3.5m. The multistorey structure was
subjected to the bi - directional seismic excitation. The results obtained at flexible edges
were almost comparable with Perus and Fajfar (2005). However, the results of both
papers did not correlate in case of stiff edges of torsionally stiff and flexible building
systems.

Table 1.7 Description of models used by Marusic and Fajfar (2005)


Model Name

Description
Torsionally stiff building

model with moment resistant beam column

connections (All beam-column connections).

F1

Building Model with torsional stiffness equal to Model S with moment


resistant beam column connections (Corner beams only)

F2

Building Model with torsional stiffness less than Model S and F1.

Stefano et al. (2006) determined the difference between the inelastic seismic
response of one storey and multistorey plan asymmetric structures. For analytical study
a single storey and a six storey steel frame with mass applied at 0.15 b of the geometric
structure causing mass eccentricity was created in the building model. The effect of overstrength of resisting elements was also evaluated. Analytical studies showed the
influence of over-strength on ductility demand of the building systems and this influence
showed variation for single and multistorey building systems. Finally it was found that
seismic response obtained from single storey model was different from that obtained
from multi-storey models. From results of analytical study it was found that for
eccentricity ratio of less than 0.5, number of resistant planes in direction of seismic
response had no influence on seismic response and the lateral displacements decreased
with increase in ductility demand. The parameters like degree of torsional coupling,
uncoupled lateral time period and eccentricity had larger influence on seismic response.
Ghersi et al. (2006) determined the effectiveness of modal analysis procedure in
evaluating the inelastic seismic response of multistorey plan asymmetric structure. In a
six storey asymmetric steel framed building, asymmetry was introduced by variation of
applying load at 0.15L away from geometric center causing mass eccentricity. Results of
modal analysis were compared with that of static analysis and Chandler procedure to

49

check accuracy of the latter. The proposed method yielded good seismic performance of
buildings as compared to other methods of analysis. However, the strength distribution
along plan given by the proposed method is comparable with method suggested by
Ghersi and Rossi but it was simpler in application as compared to the latter method.
Aziminejad and Moghadam (2009) determined seismic performance of eight 5 storey
plan asymmetric (Stiffness and strength) building systems with different strength
distributions. The eight different building systems in location of position of center of
rigidity and strength (Table 1.8) were considered. These building models were analyzed
using nonlinear dynamic analysis using OPENSEES software. From results of analytical
study it was concluded that building systems with strength eccentricity equal to one
fourth of distance between positions of strength and stiffness performed better on
rotation and drift criteria.
Table 1.8 Different model configurations considered
S.No

Ratio of stiffness to yield

Model Name

displacement eccentricity
1

Symmetric

Stiffness Symmetric

Balance (0.75 CV CR)

0.75

Balance (0.5 CV CR)

0.5

Balance (0.25 CV CR)

0.25

Strength Symmetric

De-Stefano (0.25 CM-CR)

-0.33

De-Stefano (0.5 CM-CR)

-1

Stahopoulos and Anangnopoulos (2010) evaluated the effectiveness of accidental


eccentricity provisions. For analytical study the authors created four types of building
models. The first and second models were one storey shear beam with stiffness
eccentricity and one storey frame models with mass eccentricity respectively. The third
model and fourth model were three storey and five storey frame type building with
combination of mass and stiffness asymmetry along plan. The shear beam models were
modeled considering a bilinear force-displacement behavior with magnitude of strain
hardening taken equal to 0.05. For idealization of frame members, plastic hinge and

50

Takedas moment-rotation relationships were used in creating the plastic hinge model.
The one storey and three storey building models were subjected to the accidental
eccentricities from 0 to 0.05L, whereas the five storey building model was subjected to
an additional eccentricity of 0.1L in addition to earlier mentioned eccentricities. Results of
analytical study suggested that in case of one storey shear beam models, the
consideration of accidental design eccentricity (ADE) resulted in reduction of ductility
demands of edge elements in case of building systems with larger time period (T y). For
Ty > 0.5s the ductility demand reduces by 10 % for ADE = 0.05L and by 10-20% for ADE
= 0.10L.
Anangnostopoulos et al. (2010) determined inelastic torsional response of single
storey and multi-storey building models with mass and stiffness eccentricity. The building
models were designed in accordance with EC8 and IBC code provisions. The inelasticity
in the building models was introduced by assuming Takedas moment-rotation
relationship and strain hardening ratio was taken as 0.05. The inelastic plastic hinge
models were further subdivided into three categories namely SIMP1, SIMP2 and SIMP3
The building models were analyzed using time history analysis using ANSR software
programs. From results of analytical study, it was found that for models SIMP1 and
SIMP2 the flexible edges of building were found to be the critical elements which
correlates with results obtained for single storey models by previous researchers. The
seismic response of SIMP3 model was found to be strongly dependent on seismic
loading and in this case critical elements were stiff edges which contradicts with results
obtained for single storey models.

1.5 Review of research works regarding vertical irregularities


Irregularities of mass, stiffness, strength and geometry along building height may be
termed as vertical irregularity. These irregularities may be present singly or in
combination. Different types of vertical irregularities have different effects on seismic
response. The effect of these irregularities should be considered and incorporated in
current seismic design codes. The research works concerned with vertical irregularities
started in early 1970s with Chopra and Kan (1973) who studied the seismic response of
series of eight storey shear buildings subjected to the earthquake motion data. The main
objective was to determine the effect of yielding of first storey on upper storeys. From the
results of analytical study it was found that an ideal plastic mechanism and a low yield
force are required in the first storey for safety of higher floors of the structure. The
irregularities of mass, stiffness and strength were represented by parameters of mass

51

ratio (Mr), stiffness ratio (Sr), strength ratio (STr) which may be defined as the ratio of
mass, stiffness and strength of storey under consideration to that of the adjacent storey.
Humar and Wright (1977) studied the seismic response of multistorey steel building
frames with and without setback irregularity using one ground motion data. Based on
analytical study it was concluded that, in case of building frames with setbacks, the
storey drift was found to be greater at upper portion of setback and smaller in the base
portion. Also, the drift of building frames with setbacks was found to be less as compared
to their regular counterparts. This approach was extended by Aranda (1984) who
determined the seismic response of structure with and without setback irregularity
founded on soft soil. From the results of analytical studies it was confirmed that the
ductility demand and its increase in upper portion of setback was higher as compared to
the base portion and structures with setbacks experienced higher ductility demand as
compared to their regular counterparts.
Fernandez (1983) determined the elastic and inelastic seismic response of
multistorey building frames with irregular distribution of mass and stiffness. Reduction in
storey stiffness resulted in increased storey drift and structures with constant variation of
mass and stiffness in vertical direction showed better seismic performance as compared
to the structures with abrupt variations.
Moelhe (1984) determined the seismic response of R.C structures with irregularities.
For analytical study, nine storey building frames with 3 bays with structural walls were
modeled. The irregularity in building models was created by discontinuation of structural
walls at different storey heights. Based on the analytical results it was found that the
seismic response not only depended on extent of structural irregularities but also on the
location of irregularities. Experimental studies are necessary to verify the accuracy of
analytical results and researchers like Moehle and Alarcon (1986) performed
experimental tests on two small prototype R.C. building frames subjected to the ground
motion data. The tests were performed using shake table. The two small scale R.C.
building models used for the study were named as FFW and FSW. The FFW model
had two frames of nine storey having 3 bays each and the third frame was also of 9
storey but had prismatic wall, this model represented the building systems without any
irregularity. The vertical irregularities were introduced in the building models by
discontinuation of shear wall at first storey and this building model was designated as
FSW Rest of the features in both FFW and FSW were same. The displacements of
top floor were computed for all these building models using elastic and inelastic dynamic
analysis. From the analytical study it was concluded that in case of FSW ductility
demand increased abruptly at the vicinity of discontinuity of shear wall and this increase
was found to be 4 to 5 times higher as compared to the FFW models. Further the

52

inelastic dynamic analysis was found to be more efficient as compared to the elastic
analysis in determining the effect of structural discontinuities.
Barialoa and Brokken (1991) determined the effect of strength and stiffness
variation on nonlinear seismic response of multistorey building frames. For analytical
study 8 storey buildings with 5 bays were modeled. The building frames were subjected
to three different category of time periods namely low, medium and high. Each building
category was further subdivided into two more categories based on base shear namely
weak and strong. In the weak building the base shear was 15 % of total seismic weight
whereas in strong building the base shear was 30 % of total weight of the structure. The
results of analytical study showed that the time period of structure increased during
seismic excitation and this increase was more pronounced in case of weaker structures.
Ruiz and Diederich (1989) conducted analytical studies on five and twelve storey
building models with strength irregularity. The strength irregularity in the building model
was created by modeling first storey of the structure as the weak storey in the first case.
In the second case the infill walls in top storey were modeled as brittle and in the third
case the infill walls were modeled as ductile. From results of analytical study it was found
that the yielding, failure and formation of plastic hinges in infill walls was greatly
influenced by time period of seismic excitation.
Shahrooz and Moehle (1990) determined the seismic response of building systems
with vertical setbacks. The authors conducted both experimental and analytical tests to
improve previous design methodologies for design of setback buildings. For performing
the experimental study a six storey RC frame having setback at mid-height was
prepared. From results of experimental study it was found that there was no abrupt
variation in the displacement along the building height. The interstorey drifts were found
to be largest with increased damage and abrupt reduction in lateral force at location of
setbacks. The variation of lateral displacement and force along building height suggested
that the translational seismic response of the building parallel to direction of setback was
influenced by fundamental mode of vibration. For performing analytical study six storey
building frames with six different patterns of setbacks were modeled and designed in
accordance with UBC code of practice. For all of these frames the floor plan dimensions
and mass ratios were varied from 3 to 9 times as suggested by UBC 1988 code of
practice which differentiated symmetric and setback structures on the basis of plan
dimensions and mass ratios. The analyses of these frames were carried out by modal
analysis procedure as prescribed by UBC 1988 code of practice. From results of
analytical study it was concluded that all these frames experienced similar magnitude
and distribution of ductility demand. The frames with similar mass ratios and floor plan

53

dimensions but with different setback heights experienced different amount of damage
which contradicted the approach of UBC 1988 code.
Nassar and Krawinkler (1991) conducted parametric study on multistorey (3,
5,10,20,30, 40 storey) SDOF and MDOF systems (with strength irregularity) with
different periods of seismic excitation ranging from 0.217s 2.051s. The models used
are described in Table 1.9. In case of SDOF models the strength demand was
represented in terms of strength reduction factor which represents the reduction in
strength of structural elements. In case of MDOF systems it was found that strength
demand and displacement demand depend on failure mechanisms developed and the
presence of weak first storey increased the ductility demand and overturning moments.
Esteva (1992) evaluated the seismic response of building frames with soft first
storey by using non-linear analysis. For simplification of analytical study the shear beam
model was used to represent the building systems. The first main purpose of analytical
study was to observe the bilinear hysteric behavior of the building systems with and
without consideration of P-Delta effects. The second main purpose of the analytical study
was to determine the effect of influence ratio r (which was defined as the ratio of average
value of lateral shear safety factor for upper storeys to the bottom storeys) on ductility
demand. The results of analytical study are shown in Table 1.10. Wood (1992) observed
that presence of setbacks did not affect the dynamic seismic response which was more
or less similar for symmetrical structures.
Wong and Tso (1994) used elastic response spectrum analysis to determine seismic
response of structures with setback irregularity and it was observed that buildings with
setback irregularity had higher modal masses causing different seismic load distribution
as compared to the static code procedure.
Duan and Chandler (1995) conducted analytical studies on building systems with
setback irregularity using both static and modal spectral analysis and based on the
results of analytical studies, it was concluded that both

static and modal analysis

procedures were inefficient in predicting the concentration of damage

in structural

members near level of setbacks.


Valmudsson and Nau (1997) evaluated seismic response of multistorey buildings
with vertical irregularities. For analytical study two dimensional shear beam building
models with five, ten and twenty storeys were prepared. The structural irregularities were
introduced in the building models by varying the mass, stiffness and strength. From
analytical studies it was found that introduction of mass and stiffness irregularity resulted
in minor variation in the seismic response. The storey drifts were increased in range of
20% - 40 % for 30 % decrease in the stiffness of the first storey, with constant strength.
The strength reduction of 20 % doubled the ductility demand.

54

Al-Ali and Krawinkler (1998) evaluated the effect of mass, stiffness and strength
and their combinations on seismic response of a 10 storey structure. Elastic and inelastic
dynamic analyses were used for the analytical study. Based on the results of analytical
study it was observed that, when irregularities were considered separately; the strength
irregularity had the maximum impact on roof displacement and mass irregularity had the
minimum impact on the roof displacement. When combination of irregularities was
considered, the combination of stiffness and strength irregularity had the maximum
impact on roof displacement
Table 1.9 Building Models used by Nassar and Krawinkler (1991)
MT (Model

N (No. of

Model Description

Name)

storeys)

Beam Hinge

3,10,20,30,40

Plastic hinges form in beam only

Column Hinge

3,10,20,30,40

Plastic hinges form in column only

Model 3

3,10,20,30,40

Plastic hinges form in first storey columns

Table 1.10 Results of Analytical Study obtained by Esteva (1992)


Time period

Influence ratio

Ductility Demand

Low

Increase from 1.0 to 3.0

Increase by 30 %

Medium

No impact

No impact

High

Increase from 1.0 to 3.0

Increase from 50 % - 100%

Kappos and Scott (1998) made comparison between static and dynamic methods of
analysis for evaluating the seismic response of R.C frames with setback irregularity. On
comparison between results of both methods it was concluded that dynamic analysis
yielded results different from that of static analysis. In the analytical study, other forms of
irregularities like mass, stiffness and strength irregularity were not included.
Magliulo et al. (2002) conducted parametric studies on multistorey RC frames (5, 9
storey) with mass, stiffness and strength irregularity designed for low ductility class as
per EC 8 provisions. The authors evaluated the seismic response of the irregular frames
and compared it with the seismic response of building frames without any irregularity.
From the analytical study it was found that mass irregularity does not affect plastic
demands. In case of strength irregularity, irregular distribution of strength in beams
increased the seismic demand. However, seismic demands were not affected due to
irregular strength distribution in columns. Finally the authors concluded that the

55

parameter of storey strength (define) as prescribed by EC8 and IBC codes was
ineffective in predicting strength irregularity.
Das and Nau (2003) evaluated the effects of stiffness, strength and mass irregularity
on inelastic seismic response of large number of multistorey structures. For analytical
study a large number of buildings with three bays in direction of seismic action and with
number of storeys ranging from 5-20 were modeled as shown in Figure 1.31. The
structural irregularities in these building models were introduced by variation of mass
ratio, stiffness ratio, storey strength and by considering the effect of masonry infills.
These frames were designed as special moment resisting frames (S.M.R.F) based on
strong column weak beam design philosophy in accordance with different codes of
practice namely ACI 1999 and UBC 97. The forces on these S.M.R.F frames were
computed using ELF (Equivalent lateral force) procedure as prescribed in ACI 99 and
UBC 97 code. From results of analytical study it was concluded that the seismic
response parameters like first mode shape and fundamental time period as computed by
ELF procedure were similar for symmetrical and unsymmetrical structure. The storey drift
computed for five storey and ten storey structures with combination of mass, strength
and stiffness irregularities at bottom storey showed an abrupt increase over code
prescribed limit of 2%. The ductility demands showed an abrupt increase near the
location of irregularity but this increase never exceeded the designed ductility capacity of
the members. Finally, the mass irregularity had least impact on the structural damage
index and for all the building models analyzed it was found to be less than 0.40.

E1 E2

E3 E6

(a) TYPE A,B,C Taller first, intermediate and top storey, (b) TYPE t, m, b - Irregular mass
distributions, (c) E1-E2 Open ground floor, E3 E6 Partial infill

Figure 1.31 Different types of vertically irregular building models, Das and Nau
(2003)
Chintanpakdee and Chopra (2004) evaluated the effects of strength, stiffness and
combination of strength and stiffness irregularity on seismic response of multistorey
frames. For analytical study, different 12 storey frames were modeled based on strong
column weak beam theory. The irregularity in strength and stiffness were introduced at
different locations along height of the building models. The building models were
analyzed using time history analysis by subjecting the building model to 20 different

56

ground motion data. From analytical study it was concluded that irregularities in strength
and stiffness when present in combination had the maximum effect on the seismic
response. Further maximum variation in the displacement response along height was
observed when irregularities were present on the lower storeys.
Tremblay and Poncet (2005) evaluated the seismic response of building frames with
vertical mass irregularity designed according to NBCC provisions by static and dynamic
analysis. Based on the analytical study it was concluded that both static and dynamic
method of analysis (as prescribed by NBCC provisions) resulted in similar values of
storey drifts and hence they were ineffective in predicting the effects of mass irregularity.
Fragiadakis et al. (2005) determined the seismic response of building systems with
irregular distribution of strength and stiffness in vertical direction. After conducting the
analytical study it was concluded that seismic performance of the structure depended on
type and location of irregularity and on intensity of seismic excitation. Modal pushover
analysis (MPA) procedure is an important analytical tool to evaluate the seismic
performance and several researchers like Lignos and Gantes (2005) investigated the
effectiveness of Modal pushover analysis procedure (MPA) in determination of the
seismic response pf multistorey steel braced frame (4, 9 storey) with stiffness
irregularities. Based on the results of analytical study it was concluded that MPA
procedure was incapable of predicting failure mechanism and collapse of the structure.
Khoury et al. (2005) designed a 9 storey steel framed structure with setback
irregularity as per Israeli steel code SI 1225(1998). The height and locations of setback
were varied for the analytical study. Results of analytical studies confirmed that higher
torsional response was obtained in tower portion of setbacks.
Tremblay and Poncet (2005) who conducted extensive study on multistorey building
frames with mass irregularity as per NBCC code, Ayidin (2007) evaluated the seismic
response of buildings with mass irregularity by ELF procedure (as prescribed by Turkish
code of practice) and by time history analysis. The researcher had modeled multistorey
structures ranging from 5 to 20 storey height. The mass irregularity was created by
variation in mass of one storey with constant mass at other storeys. Based on the
analytical study author concluded that the mass irregularity affects the shear in the
storey below and ELF procedure overestimates the seismic response of the building
systems as compared to the time history analysis. Some researchers preferred dynamic
analysis over MPA procedure to evaluate seismic response due to its accuracy.
Fragiadakis et al. (2006) proposed an IDA (Incremental dynamic analysis) procedure for
estimating seismic response of multistorey frame (9 storeys) with stiffness and strength
irregularity contrary to Lignos and Gantes (2005), Alba et al. (2005) who used MPA
procedure to evaluate the seismic response of building frames with stiffness irregularity.

57

Based on the analytical results the authors concluded that the proposed method was
effective in predicting effects of irregularity in building frames. Finally, the authors
concluded that the effect of irregularity is influenced by location and type of irregularity
and building systems subjected to unidirectional seismic excitation underestimate the
seismic demand significantly. Basu and Gopalakrishnan (2007) developed a simplified
method of analysis for determination of seismic response of structures with setbacks and
torsional irregularity. The assessment by the proposed method was made by applying it
on four building models. In case of building models with scattered positions of CM the
proposed method evaluates seismic response considering average value of position of
CM whereas perturbation analysis considered the exact location of CM at different floor
levels to evaluate the seismic response. Results of analytical study showed that for
building systems with vertically aligned CM. The frequencies obtained by proposed
procedure and perturbation analysis were observed to be in close agreement. However,
the results of frame shear forces differed by 7%. In case of second example, the modal
response obtained by proposed method and perturbation analysis was similar, but
difference in frame shear force was found to be 4% for upper storeys and 1% for base
storeys. In case of third building model, the frequencies obtained by proposed procedure
and perturbation analysis were in close agreement, but difference of results in case of
frame shear forces as 10 % at ground storey level and 4% at first storey level. In case of
fourth example the difference of results in estimation of frame shear forces as high as
50%, Thus, it was concluded that the proposed position is not applicable to the building
models where the prescribed limit of scattering of CM was exceeded.
Karavasilis et.al. (2008a) performed extensive parametric study on steel frames with
different types of setback irregularity designed as per European seismic and structural
codes (EC 8 :2004). From analysis, the databank of output parameters corresponding to
number of storeys, beam to column strength ratio, geometrical irregularity etc. was
created. Based on the deformation demands four performance levels were identified and
these were (a) occurrence of first plastic hinge, (b) Maximum inter-storey drift ratio
(IDRmax) equal to 1.8 % ; (c) IDRmax equal to 3.2%, (d) IDRmax equal to 4.0%. The results
for different types of setback structure were expressed in terms of these performance
levels. From analytical study it was concluded that interstorey drift ratio (IDR) increased
with increase in storey height and tower portion of setback experienced maximum
deformation as compared to the base portion.
Athanassiadou (2008) made the assessment of seismic capacity of the RC structures
irregular in elevation. The author modeled three multistorey frames. Out of these three
frames ,two ten storey plane frames were modeled with two and four large setbacks in
their upper floors and the third frame was regular in elevation. These three frames were

58

subjected to 30 different ground motions and designed as DCH and DCM frames
(designed for high ductility and medium ductility) as per Euro code 8. Then non linear
dynamic analysis of the frames was carried out by subjecting the frame to the ground
motion data of the earthquake and parameters of rotation, base shear and interstorey
drift were evaluated. Based on the analytical study it was found that the performance of
both DCM and DCH frames was found to be satisfactory as per guidelines of EC 8.
Karavasilis et al. (2008b) evaluated the seismic response of family of 135 plane steel
moment resisting frames with vertical mass irregularities and created databank of
analytical results. The authors used regression analysis technique to derive simple
formulae to evaluate seismic response parameters using the analysis databank. Results
of analytical studies suggested that the mass ratio had no influence on deformation
demand. The results obtained from proposed formulae were found to be comparable
with results of dynamic analysis.
Sadasiva et al. (2008) evaluated the effect of location of vertical mass irregularity on
seismic response of the structure. A 9 storey regular and irregular (with vertical
irregularity) frame was analyzed and designed as per New Zealand code of practice in
two ways. Firstly, it was designed to have maximum interstorey drift at all levels
(represented as CDCSIR). Secondly, it was designed to have a constant stiffness
(represented by CS) at all levels. To make clear distinction between regular and irregular
structure, a special notation form was used by the authors of form NS-M-L-(A), where Nno.of storeys, S-Shear beam, M- Type of model [i.e. S(Shear beam) or SFB (Shear
Flexure beam), (A) Mass ratio].The deformation was represented in the form of graphs.
For the study Los Angeles earthquake records had been used and inelastic time history
analysis of the structure was performed using Ruamoko software. Based on this analysis
it was concluded that in case of both CS and CISDR model the interstorey drift produced
is maximum when mass irregularity is present at topmost storey and irregularity
increases the interstorey drift of the structure. However, this magnitude varied for both
CS and CISDR type of models.
Ambrisi et al. (2009) proposed a modified pushover analysis method for determining
the seismic response of building structures, and found the comparable results by both
pushover and inelastic dynamic analysis for setback frames.
Dinh Van Thuat (2011) determined the storey strength demands of irregular buildings
under strong earthquakes. The strength irregularity in the building models was
introduced in terms of storey strength factor which represents the relative reserve
strength of the storey against failure. A large number of analysis of building models
ranging from 7 storeys to 19 storeys were conducted. The analysis results indicate the
variation in seismic demands due to introduction of irregularity.

59

Kappos and Stefanidou (2010) proposed a new deformation design method based
on inelastic analysis for the setback frames. From analysis results, adequate seismic
performance of the setback frames designed as per the proposed method was observed.
Kim and Hong (2011) determined the collapse resisting capacity of the building
models with stiffness and strength irregularity. The irregularity in the building models was
created by removal of column in the intermediate storey. However, analysis results
suggested minor variation in the collapse potentials of regular and irregular structures.
Lu et al. (2012) performed non-linear time history analysis of the tall setback building
and found excessive damage concentration in storeys adjacent to setbacks.

1.6 Brief review of Building Models used in previous research works


The models used by previous researchers can be broadly categorized by two systems of
classifications. As per first system of classification these models can be broadly
categorized into three types namely shear beam (SB), plastic hinge (PH) and 3D frame
models. In the first model, the building system is assumed to consist of a rigid
rectangular deck of mass m supported by lateral load resisting elements represented by
a shear beam. This type of building model is used to represent single and multistorey
building systems with lesser degrees of freedom. But use of this model to represent
multistorey systems is questionable due to various reasons as discussed in later
sections of this chapter. This model is used by a large number of researchers due to its
simplicity and easy representation. The shear beam model (SB) is not suitable for
representing multistorey building systems and hence this model does not represent the
actual building systems. The predictions given by shear beam models are less accurate.
In the second type of models the plastic hinges are modeled at the ends of beams and
columns to evaluate the nonlinear response of building systems. These models are
closer to reality as compared to first type of models but still do not represent the actual
building systems. The application of first two models is more frequent in case of 2D
plane frames than 3D building frames due to complex geometry of 3D building frames.
The third type of models can be termed as 3D frame models and these models have
been developed by recent researchers. These models are quite complex and involve
large number of degree of freedom systems and are prepared with the help of complex
software programs. These models are very close to the actual building systems and yield
accurate results. The second classification of building systems is based on force
displacement hysteretic relationship of resisting elements of buildings. The resisting
elements can have different type of force-deformation represented by models (Figures
1.32 and 1.33) namely

60

(a) Elasto plastic and bilinear hysteric model


(b) Cloughs model,
(c) Takedas model
(a) Elasto- plastic and bilinear hysteretic models
The elasto-plastic hysteretic model has been used by many researchers due to its
simplicity. The maximum displacement of a building system with elasto-plastic force
deformation relationship was found same as for elastic force deformation relationship
for building systems with initial time period greater than 0.5 s. To account for the strain
hardening effect a positive slope was assigned to post yield stiffness. This model was
called as bilinear model. However, this model does not represent the stiffness
degradation appropriately. Hence, this model is not suited for non-linear analysis of RC
structures.
(b) Cloughs model of stiffness degradation
A qualitative model incorporating the stiffness degradation in conventional elasto
plastic model was developed by Clough and was called Cloughs stiffness degrading
model. In this model the main response point during the loading cycle shifted towards the
maximum response point but the slope during unloading remained same as the initial
elastic slope. By virtue of this modification, Cloughs model was able to represent the
flexural behavior of reinforced concrete. From the analysis of series of SDOF system
using this model, Clough arrived at the following conclusion
(i) For building systems with higher initial time period both Cloughs and elasto-plastic
model yielded same results in terms of ductility demand
(ii) Cloughs model yielded larger ductility demand as compared to elasto-plastic model
for short period structures.
(iii) Response waveforms of both models were different.
The main advantage of this model is that it is simple and can be used for non-linear
analysis using strain hardening characteristics.
(c) Takedas hysteretic model
A more refined and complex model for representing the stiffness degradation was
prepared by Takeda in 1970 based on his experimental observations. The proposed
model included the stiffness changes due to flexural cracking, yielding and strain
hardening. In Takedas model the stiffness during unloading cycle was reduced as the
fraction of the initial deformation. Takeda also prepared set of guidelines for the load
reversals within the outermost hysteresis loop, which were major improvement over

61

Cloughs model. The main disadvantage of this model was that extensive damage
caused by shear and bond deterioration was not considered. Comparison of models
used by different researchers has been presented in Table 1.11 to 1.13.

(a)

(b)

Figure 1.32 Hysteresis models: (a) Elasto-plastic model, (b) Bi-linear model

(a)

(b)
Figure 1.33 Hysteresis models: (a) Cloughs degrading stiffness model, (b)
Takedas model

62

Table 1.11 First system of classification of models used by different researchers


MT

SB

Reference
Tso

and

Sadek

Advantages

1985;

Tso

and Easy

Disadvantages
Does not represent

Bozorgnia 1986;Tso and Sadek 1989; idealization and the actual structure.
Duan and Chandler 1991; De-La Colina formulation

Not

designed

for

1999; Tso and Myslimaj 2002; Tso and

gravity loads.

Myslimaj 2003; De-LaColina 2003;

Interdependence of

Chopra and Goel 2004; Stathopoulos

stiffness and stiffness

and Anagnostopoulos 2003,2005,2010;

was assumed.

Shakib and Ghasemi 2007;Aziminejad


and Moghadam 2009; Sadasiva et al.
2008;Ladinovic2008;
Anagnostopoulos et al.2010.

PH De - La Colina 2003; Stathopoulos The Non-linear More


and

Anagnostopoulos

Aziminejad

and

2003,

Moghadam

2010; analysis
2009; required.

Anagnostopoulos 2010;

complex

and

is difficult to model.
The location of plastic
effects

the

seismic

response.

3D

Fernandez 1983; Wong and Tso1994; Closer to actual Complex and difficult
Moehle

1984;Moehle

and

Alercon buildings.

1986; Bariola and Brokken 1991; Ruiz


and Diederich 1989; Shahrooz and
Moelhle 1990; Wood 1992;Vamudsson
and Nau 1997; Kappos and Scott
1998;Das and Nau 2003;Chopra and
Goel 2004; Lignos and Gantes 2005;
Penelis

and

Kappos

2005;

Sthathopoulos and Anagnostopoulos


2005; Tremblay and Poncet 2005;
Shakib and Ghasemi 2007; Karavasilis
et al. 2008 a, b;Kappos and Stefanidou
2010;Lu et al.2012.

63

to formulate.

Table 1.12 Second system of comparison


MT

SS

Reference

Advantages

Tso and Sadek 1985, 1989; Tso and Simple.

Disadvantages

Does not represent

Bozorgnia 1986; Duan and Chandler Easy idealization and the actual structure
1991; Hall et al. 1995; Chandler et al. formulation.

with large degree

1995; Bugeja and Thambiratinam 1999;

of freedom.

De-la colina 1999; Ghersi and Rossi


2001;Dutta and Das 2002; Tso and
Myslimaj 2003; De - La colina 2003;
Chopra and Goel 2004;
Fajfar

2005;

Perus and

Sthathopoulos

and

Anagnostopoulos 2005; Ghersi and


Rossi 2006; Jarernprasert et al. 2008
Ladinovic

2008;

Stathopoulos

and

Anagnostopoulos 2010; Luchhini et al.


2011.
MS

Chopra and Goel 2004; Choi 2004; Represents the actual More complex and
Sthathopoulos and Anagnostopoulos structure and seismic difficult to model as
2005; Fajfar et al. 2005; Lignos and response obtained is compared
Gantes 2005; Marusic and Fajfar 2005; much closer to reality.

to

SS

models.

Fragiadakis et al. 2005; Karavasilis et Can involve large no. Requirement of the
al. 2008; Aziminejad and Moghadam of degree of freedom.

complex softwares.

2009;Aziminejad and Moghadam 2010;


Sthathopoulos and Anagnostopoulos
2010; Anagnostopoulos et al. 2010;
Luchinni et al. 2011.

where SS and MS represent single storey and multistory building models

64

Table 1.13 Third system of classification of models used by other researchers

MT
1

Reference
Tso

and

Sadek

Advantages

Disadvantages

1989; Simple

Less accurate for building

Ghersi and Rossi 2001;

systems with T > 0.5s.

Dutta and Das 2002;


Tso and Myslimaj 2003;
Gehrsi and Rossi 2006;
Valmudsson and Nau 1997.
2

Tso and Myslimaj

2003; Strain hardening effect Does

not

account

for

Bugeja and Thambiratnam has been included

stiffness change due to

1999; Duan and Chandler

increase in displacement

1997; Sthathapoulos and

amplitude reversal.

Anagnostopoulos
2003;

2003,

Jarenpraresrt

et

al.2008.
3

Tso and Myslimaj 2003;

Used

for

nonlinear Larger ductility demand as

Tso and Sadek 1985;

analysis includes strain compared

Tso and Sadek 1989;

hardening effect.

to

elasto

plastic elements.

De-La-Colina 1999, 2003.

Comparable values with


model 1 for high period
structures.

Das

and

Nau

2003; Includes

effects

of

Excessive damage caused

Fragiadakis et al. 2005; flexural, cracking and by shear and bond not
Stathopoulos

and strain hardening.

considered.

Anagnostopoulos 2010

In Table 1.13, MT 1 - 4 represents different hysteresis models namely, (a) Elastoplastic


model, (b) Bilinear model, (c) Cloughs model, (d) Takeda model

1.7 Discussions on Literature review


The presence of structural irregularity changes the seismic response and the change in
the seismic response depends upon type of structural irregularities. As mentioned
previously, structural irregularities may be classified into horizontal and vertical
irregularities. On comparing research works regarding plan and vertical irregularity, it
was found that large numbers of research works were conducted on plan and vertical

65

irregularities which shows that previous researchers have given equal importance to both
types of irregularities.
For modeling plan irregularities, some researchers used single storey models and
others used multistorey building models. Single storey models were largely used by
previous researchers with majority of recent researchers preferring multistorey building
models in comparison to the multistorey building models with few exceptions
(Stathopoulos and Anagnostopoulos 2010; Anagnostopoulos 2010). The expressions
proposed in seismic design codes (regarding plan and torsional irregularity) is based on
single storey SDOF systems and elastic analysis (Moghadam 1998). Thus, the code
provisions are not valid for multistorey building models. Therefore, there is a necessity to
revise these provisions. Moreover, the modified expressions needs to be generalized to
have wider applicability.
Different centers of buildings like CM, CV and CR have a huge impact on seismic
response of building systems as torsion generated depends upon positions of these
centers with respect to each other. Several researchers (Tso and Myslimaj 2002; Tso
and Myslimaj 2003; Aziminejad, A. and Moghadam 2010) have proposed the concept of
balanced CV-CR location to generate minimum torsional response. One of the main
issue in this concept is that the previous researchers have not been able to find a CV-CR
location which gives optimum values for all the seismic response parameters like drift,
ductility and rotation etc. In general, it was found that if some position of CV, CR, CM
reduces drift and ductility, then other portion reduces rotation i.e. no particular position of
CV, CR and CM results in optimum values of seismic response parameters (Tso and
Myslimaj 2003). Thus, selection of seismic response parameters for optimization is up to
the priority of the designer and it may vary according to user requirements and building
specifications. From result of researches it was found that different locations of CV, CR
and CM yielded different results and the effects of these locations were different on
different seismic response parameters. Optimum position of CV - CR was found to be
highly dependent on type and period of seismic excitation (Tso and Myslimaj 2003).
Research works regarding vertical irregularities are fewer in number as compared to
plan irregularities. Also, the main focus of research works was to vary either mass,
stiffness and strength ratios to study the effect of this variation in seismic response (Ruiz
and Diederich 1989; Ayidin 2007; Karavasilis et al. 2008; Sadasiva et al. 2008). One of
the main conclusions was that effect of irregularity depended on extent and location of
irregularity and variation in seismic response parameters was found at the vicinity of
irregularity. Some of the vertical irregularities like strength and stiffness were found to be
interdependent and their relation was evaluated by some of the researchers (Dutta and
Das 2002; Priestley 2003). Many researchers created stiffness and strength irregularity

66

by discontinuation of shear walls at particular storey height, this method of introducing


irregularity was adopted for building models of different storey heights. The height and
location of discontinuation of shear wall was also varied and in every such case it was
observed that there was a large variation in ductility demand in the vicinity of
discontinuation of irregularity (Fernandez 1983; Moehle 1984; Bariola and Brokken 1991;
Maruic and Fajfar 2005).
Regarding vertical setback irregularity, the top portion of setback was found to have
greater deformation as compared to the base. Regarding the method of analysis used by
various researchers, inelastic dynamic analysis and pushover analysis were used by
majority of researchers. The researchers especially like Chopra and Goel (Chopra and
Goel 2002; Chopra and Goel 2004) have done extensive work in improvement of
pushover analysis to match the accuracy of dynamic analysis, but it was found that
pushover analysis even after a large improvement was found short of dynamic analysis
and was applicable only for certain types of loadings and building systems. Regarding
the analysis method used the main focus was on improving the Modal Pushover analysis
procedure to match results of dynamic analysis. With the development of advance
softwares, the dynamic analysis of different types of complex multi-storey structures can
be easily performed. Therefore, use of dynamic analysis is more preferable and because
MPA procedure are less accurate as compared to dynamic analysis. This, justifies the
use of dynamic analysis.

1.8 Organization of the Dissertation work


The first Chapter presents a basic introduction to different types of plan and vertical
irregularities that exist in the building systems. The restrictions of code prescribed design
philosophies has been discussed in brief, and a need to propose new seismic design
philosophies is justified. Some of the actual structural failures that have occurred due to
the presence of irregularities during past earthquakes have been briefly discussed. The
last portion of this chapter presents a detailed literature review of the past research
works concerned with different types of irregularity. In the final section of this chapter, a
brief discussion on the literature review has been presented.
The second Chapter deals with the structural modeling and analysis aspect. In this
chapter, different types of structural models adopted for the analytical studies have been
described first. Firstly, the details of building models with plan and vertical irregularities
have been described with details of different ground motions adopted. Secondly, the
different inelastic modeling and cracking approaches have been reviewed. Based on
their relative merits and demerits appropriate approaches have been adopted. Thirdly,

67

overcoming the code limitation (which captures magnitude of irregularity alone), an


irregularity index is proposed based on dynamic characteristics of building to efficiently
capture the aspects of irregularity (magnitude and location) and structural cracking.
Finally, the proposed irregularity index is compared with the code and previous approach
for irregular building models with cracking effects.
The third Chapter aims at proposing the empirical expressions for period - height
relationship of different irregular building models. The period - height relationships of
these building models have been evaluated by considering gross stiffness and cracked
stiffness of the members as separate cases. The cracked stiffness using two different
approaches. The period - height relationships are evaluated using both eigen - value
analysis and inelastic time history analysis. A brief comparison has been made between
the period - height relationships evaluated by different methods and for different
stiffnesses. Finally, the applications of the proposed period - height relationships in
evaluation of seismic response and in seismic vulnerability assessment has been briefly
discussed.
The fourth Chapter deals with the study and estimation of seismic demands for the
irregular buildings. The variation of the inelastic seismic demands with location and
magnitude of irregularities has been discussed at first. Secondly, a parameter irr
correlating the inelastic and elastic seismic demands have been proposed. Thirdly, a
detailed review of behavior factor has been conducted and a new approach to estimate
behavior factor for irregular buildings (considering gross and cracked stiffness) has been
proposed. Finally, based on the regression analysis on results of dynamic analysis
generated, simple equations to estimate the inelastic seismic demands have been
proposed. The applicability of the proposed equations in current methods of seismic
design (PBD and DBD) have been briefly discussed.
The fifth chapter deals with the aspect of prediction of seismic performance of the
building models considered for the analytical study. The seismic performance of the
building models has been predicted in terms of collapse capacity and seismic damage
index. The first part of this chapter deals with the aspect of collapse capacity. In the first
part, a brief overview of collapse assessment methodology has been presented. The
different aspects concerned with determination of collapse capacity and probability of
collapse are discussed in detail. The collapse capacity and probability of collapse for the
building models considered in the analytical study have been evaluated. Based on
regression analysis conducted on the building models, simple equations to estimate
these parameters have been proposed. The second part of this chapter deals with
evaluation of seismic performance in terms of damage indices. In this part, a brief review
of different damage indices are discussed at first. Finally, a new damage index to

68

represent the seismic performance has been proposed based on inelastic seismic
response of the irregular buildings. The proposed index is compared with Park and Ang
(1985) index to determine its effectiveness. The sixth and the last Chapter summarizes
the work performed in this dissertation and gives details of the key findings of this
research work.
The appendix section in the last portion of this dissertation deals with details of some
of the important additional aspects regarding this research work.

1.9 Summary and main conclusions


This Chapter mainly deals with the introduction to the aspect of structural irregularity. In
this Chapter different forms of structural irregularity and code provisions (of different
countries) regarding irregularities have been briefly described. Secondly, a detailed
literature review regarding different aspects of structural irregularity has been presented.
Finally, a brief organization of this dissertation work has been presented. Based on the
above study following conclusions are arrived at
(a) The code provisions of different countries prescribed similar limits of structural
irregularity based on magnitude of irregularity only ignoring the location aspect which
makes them unrealistic.
(b) Majority of the previous research works (Tso and Bozorgnia, 1986;Tso and Sadek,
1989 etc.) have adopted single storey building models and SDOF systems (due to
simplicity) to determine the seismic response which is unrealistic. However, the recent
researchers have adopted multistorey building models (Karavasilis et al. 2008;
Athanssiadou 2008; Kappos and Stefanidou 2010; Lu et al. 2012) for seismic analysis
which is more closer to reality.
(c) The vertical irregularities are one of the main reason of failure of these structures
during past earthquakes. However, with advancement of time, the aspect of vertical
irregularity grabbed more attention. Many researchers (Tso and Bozorgnia 1986;Tso and
Sadek 1989) assumed strength and stiffness to be independent which was later
contradicted by other researchers (Dutta and Das 2002; Priestley 2003; Crowley 2003).
The seismic performance of setback structures is unclear case of setback irregularities,
the view of the researchers is quite unclear with some indicating the adequate seismic
performance and others opposing this view. However, majority of the researchers have
observed the aggravated deformation demands near tower portion of the setback.
(d) The past researchers have mainly adopted linear elastic analysis due to simplicity.
Howevr, this method was observed to be over - conservative (Al-Ali and Krawinkler
1998; Das and Nau 2003). With advancement of time, pushover method became quite

69

popular and yielded more accurate estimate of the seismic response as compared to the
previous one. However, these methods were preferable when higher mode contributions
were insignificant. Many researchers have tried to improve its accuracy by including
higher mode contributions but the accuracy of non linear dynamic analysis could not be
achieved (Chopra and Goel 2002; Goel and Chopra 2004). Although, computationally
complex and tedious, non linear time history analysis was observed to yield accurate
results (Karavasilis et al. 2008a,b; Athanassioudu 2008).

70

CHAPTER 2
STRUCTURAL ANALYSIS AND MODELING

2.1 Introduction
The presence of structural irregularity in a building has a significant impact on its seismic
response. The structural irregularity aspect has not been adequately addressed by the
codes in formulating the seismic design methodologies. The past earthquake records
show that the irregular buildings exhibit a poor seismic performance which shows
inadequacy of the seismic design codes based on which these buildings were designed.
Therefore, structural irregularity aspect needs to be incorporated in formulating the
seismic design methodologies. Moreover, the code procedures of seismic design are
based on elastic analysis and single degree of freedom system (SDOF) which is
unrealistic. In this Chapter, the building models with different types, magnitude and
location of irregularity have been described at first. Secondly, different analysis methods
available to obtain the seismic response have been discussed and based on review of
analysis methods a suitable method has been adopted for analysis of irregular building
models.
As discussed in the previous Chapter, the code approaches and previous literature
works have formulated the seismic design methodologies based on elastic analysis and
SDOF system. Therefore, it is very essential to perform inelastic analysis to obtain
realistic estimate of seismic demands. To achieve this purpose, different inelastic
modeling approaches have been reviewed and a brief comparison between these
approaches has been presented in tabular form. Based on this comparison a suitable
inelastic modeling approach has been adopted to determine inelastic behavior of the
building models.
As discussed previously, the structural irregularities have a significant impact on the
seismic response. Although, most of the seismic design codes have quantified the
structural irregularity; but, these approaches are based on magnitude of irregularity only
ignoring the irregularity location. However, both magnitude and location of irregularity
have a significant impact on the seismic response (Nassar and Krawinkler 1991; Al Ali
and Krawinkler 1998; Das and Nau 2003). Therefore, the present research work
presents a new approach in the form of an irregularity index to quantify different types of
irregularity incorporating both type, magnitude and irregularity location. Finally, the
proposed index is compared with the code defined approaches for different irregular
building models to illustrate the effectiveness of the proposed approach in representing
the irregularity.

71

2.2 Definition of the building models


The present study adopts building models with different type, magnitude and location of
irregularity. The seismic responses of these building models have been compared with
that of the regular building model. These building models have been briefly described in
following subsections.

2.2.1 Definition of the base case


The base case has been represented by the building models without any irregularity in
mass, stiffness and strength distribution and may be called as a regular building model.
The different forms of structural irregularities have been introduced in the regular building
model to generate irregular building models. These irregularities have been introduced
by variation in mass ratio, stiffness ratio, strength ratio, geometry and eccentricity in
accordance with previous research works (Al-Ali and Krawinkler 1998; Moghadam,
1998).

2.2.2 Definition of building models with plan irregularity


The structural models used in the present study are of regular plan shapes with storey
height ranging from 6 36 storeys. The base plan size has been kept as 4m x 4m with
number of bays varying from 1 to 5 in both X and Z directions. The bay width has been
adopted as 4m in accordance with the current seismic design codes (IS 1893:2002, EC
8:2004). To create horizontal irregularity, the center frame and corner frames have been
moved with respect to each other from their initial position to a distance varying from

0.05bw (bw is longer plan width) to 0.30 bw at intervals of 0.05 bw. This results in shift of
position of center of strength and center of stiffness which generates torsion. This
methodology to create horizontal irregularity has been adopted in accordance with
Moghadam (1998). The self-weight of the frame is taken as the dead load and the soil
class has been assumed as the hard soil. The imposed load is assumed as 3 kN/m2 in
accordance with IS:875. The expected earthquake ground motion has been defined by
the EC8 design spectrum with a PGA equal to 0.5g. The load combinations have been
adopted in accordance with EC 8:2004 and IS 1893:2002. The compressive and tensile
strengths

of

concrete

and

steel

have

been

assumed

as

25N/mm2

and

415 N/mm respectively.


2.2.3 Definition of building models with vertical irregularities
The vertical irregularity can be subdivided into irregularities due to mass, stiffness,
strength and setback irregularity. The details of the building models are similar to as

72

described in the previous section except the aspect of irregularity. In accordance with the
previous research works (Nassar and Krawinkler 1991; Al-Ali and Krawinkler 1998; Das
and Nau 2003; Karavasilis et al. 2008 a), the irregularity has been generated in terms of
variation of mass, stiffness, strength and setback along the building height. The mass
irregularity has been varied from 0% to 1000%, and stiffness and strength irregularities
have been varied from 0% to 100%. The mass variation has been introduced by
decreasing the imposed load in a storey to create mass irregularity). The review of
previous research works pertaining show that previous research works have generated
strength and stiffness variation by (a) Variation in size of

structural members

(Karavasilis et al.2008a,2008b). (b) Variation in storey height (Al Ali and Krawinkler
1998; Das and Nau 2003), Nevertheless, majority of the research works have adopted
the second method to generate stiffness and strength variation. However, it is worthwhile
to compare both these approaches. To achieve this purpose, the present study adopts
both these methods to generate stiffness and strength variation in a building. Moreover,
both these methods result in variation of strength and stiffness (due to interdependence
of strength and stiffness). To avoid confusion, the former method (beam column size
variation) has been named as stiffness irregular (S) building model and the latter
(variation in storey height) has been named as strength irregular (ST) building model.
Furthermore, the total structure has been divided into three parts; (a) Bottom one - third
storey, (b) Middle one - third storey, (c) Top one - third storey. In the first case the
irregularities have been kept in the bottom one- third storeys (BS) keeping the middle
one third (MS) and the top one third storeys (TS) devoid of irregularities. In the
second and third case, irregularities have been kept in the middle one - third storeys and
top one - third storeys respectively, keeping other two building parts devoid of
irregularities (Table 2.1). This is done to determine the effect of irregularity location on
the seismic response. In Table 2.1, the highlighted cases generated maximum impact on
the seismic response. Thus, these cases have been named as the main cases.

2.2.4 Definition of building models with vertical setback irregularity


The setback generally represents the simultaneous reduction of mass and stiffness. The
present study is based on frames that are plane and orthogonal with bay widths and
storey heights equal to 4m and 3m respectively. Moreover, the fundamental time period
of the buildings considered in the analytical study has been kept within limits proposed
by Goel and Chopra (1997) to ensure that these frames are representative of general
moment resisting RC frames as per definitions of Goel and Chopra (1997) (Figure 2.1).
The setback frames (Figure. 2.2 to Figure 2.5) adopted are low to high rise frames with
number of storeys varying from 6 to 36 storeys. These frames represent different

73

geometrical configurations of setbacks. The building models adopted in the analytical


study (Figure.2.1) can be categorized into following types (Figure 2.2 a); (a) Regular
frames (geometry No. 39 etc.), (b) Frames with smaller setbacks (geometry No. 1,2 etc.
(c) Frames with larger setbacks (i.e. geometry No. 10, 11, 12 etc.), (d) Frames with
central tower ( i.e. geometry No. 35, 36, 37 etc), (e) Frames with setbacks at various
levels i.e. geometry No. 25, 26, 28 etc). It is worthwhile to note that the frames adopted
for analytical study have a wide range of beam sizes varying from 0.20m x 0.30m to 0.45
x 0.55m, and the corresponding column sizes vary from 0.30m x 0.45m to 0.55m x
0.65m for determination of collapse response parameters. For modeling and analysis E
Tabs (2009) software has been used in accordance with previous research studies
(Balkaya and Kalkan 2003; Sextos et al. 2011)
108

H (m)

90
T

72

1.4 T
54

Tb

36
18
0.6

2.3

T (Sec)

Figure 2.1 Period - height relationships for building models adopted in the
analytical study complying with Goel and Chopra (1997)

Table 2.1 Detailed locations of irregularities

Total cases

BS (No. of storeys in which


irregularity is present)

MS

TS

2,1

same as BS

same as BS

3,2,1

same as BS

same as BS

12

4,3,2,1

same as BS

same as BS

12

15

5,4,3,2,1

same as BS

same as BS

15

18

6, 5,4,3,2,1

same as BS

same as BS

18

21

7, 6, 5,4,3,2,1

same as BS

same as BS

21

24

8, 7, 6, 5,4,3,2,1

same as BS

same as BS

24

27

9, 8, 7, 6, 5,4,3,2,1

same as BS

same as BS

27

30

10, 9, 8, 7, 6, 5,4,3,2,1

same as BS

same as BS

30

33

11, 10, 9, 8, 7, 6, 5,4,3,2,1

same as BS

same as BS

33

36

12, 11, 10, 9, 8, 7, 6, 5,4,3,2,1

same as BS

same as BS

36

74

21

22

23

24

25

26

27

28

10

29

30

11

12

13

31

32

33

14

15

34

16

35

17

36

18

37

19

38

20

39

(a)

21

22

23

24

25

26

27

28

10

29

11

30

12

31

13

32

14

33

15

34

16

35

17

36

18

37

19

20

38

39

(b)

21

22

23

24

25

26

10

27

28

29

30

11

12

13

14

15

16

17

18

19

31

32

33

34

35

36

37

38

13

14

15

16

17

18

19

35

36

37

38

39

20

40

(c)

20

21

22

23

24

25

26

27

28

10

29

11

30

12

31

32

33

34

(d)

Figure 2.2 Setback building models varying from 6 to 18 storeys adopted for the
analytical study: (a) 6 storey, (b) 9 storey, (c) 12 storey, (d) 15 storey

75

10

21

22

23

24

25

26

27

28

29

30

11

12

13

14

15

16

17

18

19

20

31

32

33

34

35

36

37

38

39

40

14

15

16

17

18

19

20

37

38

39

(a)

21

22

23

24

25

26

27

28

10

29

11

30

12

31

13

32

33

34

35

36

(b)

20

21

22

23

24

25

26

27

28

10

29

11

30

12

31

13

32

14

33

15

34

16

35

36

17

37

18

38

19

39

(c)

Figure 2.3 Setback building models varying from 18 to 24 storeys adopted for the
analytical study: (a) 18 storey, (b) 21 storey, (c) 24 storey

76

21

22

23

24

25

26

27

28

29

10

11

30

31

12

13

32

14

15

16

17

18

19

20

33

34

35

36

37

38

39

40

13

14

15

16

17

18

19

20

(a)

21

22

23

24

25

26

27

28

29

10

11

30

31

12

32

33

34

35

36

37

38

39

40

(b)

Figure 2.4 Setback building models varying from 27 storeys to 30 storeys adopted
for the analytical study: (a) 27 storey, (b) 30 storey

77

10

11

12

13

14

15

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

11

12

16

17

18

19

20

36

37

38

39

40

16

17

18

19

20

36

37

38

39

40

(a)

21

22

23

24

25

26

27

28

29

10

30

31

32

13

33

14

34

15

35

(b)

Figure 2.5 Setback building models varying from 33 to 36 storeys adopted for the
analytical study: (a) 33 storey, (b) 36 storey

78

2.3 Details of ground motions used


The main aim of this study is to determine the effect of structural irregularities on seismic
demands. The different ground motions adopted for analytical study have been obtained
from PEER database as indicated in Table 2.2. The critical time period for these ground
motions has been computed using Newmark and Riddle algorithm (Riddel and Newmark
1976) that divides the total response spectrum into constant acceleration, constant
velocity and constant displacement regions.

2.3.1 Scaling of the ground motion records


The ground motion records have been scaled to bring these ground motions to common
intensity level which reduces the dispersion of the ground motion records. Moreover
scaling enables combination of different ground motion records, and seismic response of
the structure for this combination can be determined. The details of different aspects of
scaling are available in previous research studies (Miranda 1993; Vidic 1994; Kalkan and
Chopra 2011). For the present analytical study, the ground motion records have been
scaled to EC 8 spectrum of PGA = 0.5 g. The ensemble of scaled accelograms of ground
motions are shown in Figure 2.6 and the target spectrum to which these accelograms
have been scaled are shown in Figure 2.7.

Figure 2.6 Mean of ground motions 1 scaled to EC 8 spectrum (PEER)

79

Table 2.2 Details of ground motions used

Name of Earthquake and Station

Date

Mu

De

PGA

Tc

SF

43

1.74

0.334 18.84

29

2.63

0.507 6.288

Kern Country (Taft)

21/07/1952 7.7

San Fernado (Castaic)

09/02/1971 6.6

Imperial Valley(Calexico)

15/10/1979 6.6

15

2.70

0.253 6.176

Imperial Valley(Delta)

15/10/1979 6.5

44

3.44

0.601 3.778

Coalinga (Cantua Creek School)

02/05/1983 6.4

26

2.75

0.601 4.598

Loma Prieta (Gilroy Array #4)

18/10/1989 6.9

16

4.09

0.402 4.254

Loma Prieta (Coyote Lake Dam)

28/06/1992 6.9

22

4.75

0.503 6.734

Loma Prieta (SF Intern Airport)

17/01/1994 6.9

64

3.23

0.701 26.63

Landers (Desert Hot Springs)

22/01/1994 7.3

23

1.68

0.352 7.365

Northridge (LA-Centinela St)


Northridge (Castaic Old Ridge
Route)
Northridge (Hollywood Wiloughby
Ave)
Northridge (LA N Faring Rd)

17/01/1994 6.7
17/01/1994 6.7

31
23

3.15
5.58

0.551 4.636
0.401 2.460

17/01/1994 6.7

26

2.41

0.901 5.652

17/01/1994 6.7

24

2.68

0.601 6.075

Northridge (LA-Hollywood Stor FF)

24/01/1994 6.7

26

3.52

0.351 4.442

Northridge (Glendale-Las Palmas)

18/01/1994 6.7

25

2.02

0.401 9.927

Northridge (LA-Chalon Road)

22/01/1994 6.7

24

2.21

0.201 5.880

Northridge (Moorpark Fire Sta)

17/01/1994 6.7

28

2.86

0.251 6.017

Northridge (LA-S Grand Ave)

17/01/1994 6.7

37

2.84

0.601 6.343

Northridge (Mt Wilson-CIT Seis Sta)

17/01/1994 6.7

46

2.29

0.451 17.19

Northridge (San Gabriel-E.Grand


Ave.)
Northridge (Canoga Park-Topanga
Can)

17/01/1994 6.7

42

2.51

0.301 9.456

17/01/1994 6.7

16

4.12

0.251 2.942

Northridge (LA-Century City CC North) 17/01/1994 6.7

26

2.51

0.451 4.989

Northridge (LA-Obregon Park)

17/01/1994 6.7

37

3.10

0.501 6.162

Northridge (LA-Baldwain Hills)

17/01/1994 6.7

38

3.48

0.401 6.295

Northridge (LA-Wonderland Ave.)

17/01/1994 6.7

31

1.65

0.501 11.56

Northridge (Psadena-N Sierra Madre)

23/01/1994 6.7

23

0.17

0.301 10.20

Northridge (Leona Valley#3)

17/01/1994 6.7

39

2.40

0.352 15.73

(Km) (m/s2)

where Mu is magnitude of earthquake, De is distance of epicenter, PGA is peak ground


acceleration, Tc is the critical time period, SF is the scale factor of the accelogram

80

Figure 2.7 EC 8 spectrum (Target spectrum) to which ground motions have been
scaled (PEER) (Sdd is the spectral displacement, TB, TC and TD are periods at
different points of acceleration spectrum)

2.4 Review of analysis methods


The seismic response of the building systems shows a large dependence on the type of
analysis method adopted. In past years, the analysis methods were confined to linear
static approach due to its simplicity. Although these methods yielded safe design; but
were observed to be over conservative. The development of sophisticated computers
and analysis programmes enabled the researchers to move forward towards a more
rational approach by stimulating the actual earthquakes on the building models to obtain
the realistic seismic response; these methods were categorized under dynamic analysis.
Both the static and dynamic analysis were further sub-classified into linear and nonlinear
methods depending upon the force deformation relationship of the structural members.
These methods have been briefly described in this section.

2.4.1 Linear approach In the linear approach, the force is assumed to be constant
with time. This approach can be further sub-classified into linear static and linear
dynamic approach as discussed in following subsections.

81

2.4.1.1 Linear static approach


This approach has been prescribed by most of the design codes for building structures
with smaller storey height. As per this approach, the seismic response can be computed
by applying set of lateral forces to the structure. The linear static approach is a force
based and the design parameters mainly depend on computation of base shear which in
turn depends upon fundamental time period and seismic weight.
The factor Rris the response reduction factor and it intends to account for both
damping and ductility element in the structural system at yield and ultimate
displacement. Therefore, for a system with light damping made of brittle material, the
parameter Rr would be closer to unity, and for a heavily damped ductile frame it would
range from 2 to 5 (UBC 1997). However, the parameter Rr assumes the value of 3 and 5
for OMRF and SMRF frames (IS 1893:2002). Likewise, different seismic design codes
prescribe different value of parameter Rr. Nevertheless, the importance factor If
depends on the required seismic performance of the structure, and assumes different
values as per different seismic design codes.
The fundamental time period is estimated by code expressions which are slightly
different from each other. These equations are derived based on Rayleighs method and
this aspect has been discussed in detail in the next Chapter.

2.4.1.2 Linear dynamic approach


The linear dynamic approach is similar to the linear static approach and uses the
structural model linearly elastic in nature. However, this analysis adopts the dynamic
forces contrary to the linear static approach which employs the static forces. The
dynamic forces in this method are applied in the form of the code specified response
spectrum to the structure. Therefore, it provides a greater insight into the structural
response as compared to the linear static approach. Furthermore, the representative
ground motion is not reduced by the response modification factor Rr (Chopra 1973).
This method requires an eigen-value analysis of the building analytical model to
determine the natural frequencies and the mode shapes. By use of the mathematical
procedures and a response spectrum corresponding to the specified damping, the modal
frequencies and shapes are further used to compute the spectral demands. These
spectral demands are used to calculate the member forces, displacements, storey
shears, base reactions etc. These modal forces are then combined using an established

rule (SRSS, ABS, CQC) to calculate the total response quantity to achieve better
accuracy. The equation of dynamic equilibrium of a structure with N degrees of freedom
under seismic excitation as

82

d 4t
d 2t
d 4t
dt

M
k

4
4
2
dx
dx
d x
dx

(2.1)

The equation 2.1 can be written in the form of modal displacements as

d 4m
d 2m
d 4t
dm

C
*

K
*

M
**
k

4
2
4
dx
dx
d x
dx

(2.2)

d 4m
d 2m
d 4t
dm

M
**
k

4
4
2
dx
dx
d x
dx

(2.3)

M *

M *

using property of orthognality of mode shapes, equation 2.3 can be rewritten as

d 4m
dm2
d 4t
dm
**
M i* 4 Ci* 2 Ki*

M
k

4
i
dx
dx
dx
dx

(2.4)

where i = 1,2,to N,d4t/dt4 and dm4/dx4 are acceleration vector and modal acceleration
vector. Dividing equation 2.4 with parameter Mi* we get

M i* d 4t
d 4m
dm2
2 dm
4 2 X ii 2 i ** 4
dx
dx
dx
M i dx

(2.5)

where Xi is the damping ratio in ith mode of vibration and is = ci*/2Mi*i


th

i = Angular frequency in i mode of vibration


After determination of each modal displacement vector, the relative displacements are
computed from

dx
dm

dt
dt

(2.6)

Moreover, in the response spectrum analysis, the maximum modal amplitude is equal to
the spectral displacement corresponding to the structural damping and modal frequency
at the i th mode. Therefore, for building system with N degrees of freedom subjected to
the seismic excitation, the absolute modal maximum is given by

YiMax Pk Sdd

(2.7)

where Pk is the modal participation factor at ith mode and Sdd is the spectral displacement
at ith mode of vibration. The maximum physical displacement can be computed as

U I max yI max i Pk Sdd i

(2.8)

The force vector producing the above displacements is presented as

f si K (Umax ) Pk Sdd K i

(2.9)

Also, equation 2.9 can be rewritten as

f si K (Umax ) Pk Sdd M i
where Sdd is the spectral displacement at ith mode

83

(2.10)

The above equation results in the maximum seismic response quantities. Furthermore,
the seismic response at each mode is combined individually to get the total seismic
response using different modal combination rules (SRSS, ABS and CQC).
2.4.1.3 Time history method
Time-History analysis is a step-by-step procedure where the loading and the response
history are evaluated at successive time increments. During each step the response is
evaluated from the initial conditions existing at the beginning of the step (displacements
and velocities and the loading history in the interval). In this method, the non-linear
behavior may be easily considered by changing the structural properties (e.g. stiffness,
k) from one step to the another. Therefore, this method is very effective to determine the
non-linear response,. However, in linear time history analysis, the structural properties
are assumed to remain constant and a linear behavior of structure is assumed during the
entire loading history. As a consequence the mode superposition method as already
discussed in previous section is applicable.

2.4.2 Nonlinear methods


The nonlinear analysis methods perform better than linear analysis methods. These
methods provide realistic estimate of seismic demands of the structural components
deforming in the inelastic range. Furthermore, these methods are very effective to obtain
the rational estimates of structural components of strength and stiffness deterioration in
the inelastic range. Through employment of these techniques, it is possible to identify the
areas of large deformation demands and these areas can be provided with special
detailing to obtain the desired behavior. This approach has been further classified into
nonlinear static and nonlinear dynamic analysis as discussed below:
2.4.2.1 Non-Linear static analysis
Nonlinear static procedure starts with the definition of control node in a structure. The
previous research works (Moghadam 1998; Fajfar et al. 2002; Fajfar et al. 2005)
pertaining to non-linear static analysis have considered the control node at center of
mass of roof of the building. In this procedure, the mathematical model of the structure
using this approach is prepared incorporating the aspect of material and geometrical
nonlinearities. Then the modeled structure is subjected to monotonically increasing loads
resulting in the increased displacement. This process is repeated until the occurrence of
structure collapse. Since, the mathematical model directly incorporates the effects of
inelastic material response; it results in fairly accurate estimate of the seismic response.
Before initiating the static analysis procedure, the gravity loads are applied to the

84

structure. Then the lateral load profiles of the building model are selected approximately
to represent the distribution of the inertia forces during an earthquake. These forces vary
in a complex manner during the seismic excitation. In elastic range, the inertia forces
mainly depend upon factors like ground motion characteristics and mode shapes of the
building. If the building response is in the non-linear range, then the distribution of these
forces is influenced by localized yielding of the structural components. For performing
seismic analysis and design, simplified procedures are required that can capture the
worst possible scenarios of the building. The different patterns of force distribution
methods are discussed in detail in FEMA 273 and FEMA 274. The non-linear static
procedure although more accurate than linear elastic analysis fails to give an exact
estimate of the seismic response. The main disadvantage with this procedure is that it
does not account for variation in the dynamic response and inertial load patterns which
vary with degrading strength and stiffness. Furthermore, it ignores the effect of higher
modal contributions. Therefore, a more rational nonlinear approach needs to be adopted
to get realistic estimate of seismic demands.

2.4.2.2 Non-Linear dynamic analysis


In this method, the seismic response of the structure is evaluated using step-by-step
time history analysis. The main methodology of this procedure is almost similar to the
static method of analysis. However, this approach differs in the concept that the design
displacements are not established using the target displacement; but, are estimated
through dynamic analysis by subjecting the building model to an ensemble of the ground
motions. The calculated seismic response is very sensitive to the ground motion
characteristics, and the analysis is carried out for more than one ground motion record.
To perform the non-linear dynamic analysis, the equation prescribed by the Newmarks
method (Chopra 2001; Cook 1988 and Humar 1990) can be suitably extended. Based on
review of analytical methods, the non-linear dynamic analysis method is adopted for the
analytical study due to its accuracy and efficiency in determining the inelastic seismic
response of a system subjected to the ground motion data. The review of previous
research works show that the past research works have adopted static methods in
majority for simplicity. However, the present research works in majority have adopted
dynamic analysis (especially non-linear dynamic analysis) to achieve better accuracy to
estimate the realistic seismic demands. Moreover, different seismic design codes
prescribe dynamic analysis for medium and tall structures and it has been used by
recent researchers as well (Karavasilis et al. 2008 a,b; Panda and Ramachandra 2010).
Therefore, non-linear dynamic analysis method has been adopted in the present study to
determine the seismic response of the building models.

85

2.5 Modeling of the inelastic behavior


To model the inelastic behavior different hysteresis models are used to represent the
force deformation behavior of the structural elements. This section starts with the brief
review of different hysteresis models, then a brief comparison is made between these
models in tabular form and an appropriate hysteresis model is selected to represent the
inelastic behaviour.

2.5.1 Clough and Johnson model (1)


The bilinear elastic-plastic hysteresis models were used in past years due to their
simplicity. The bilinear models were first proposed by Clough and Johnston (1966) who
proposed a bilinear hysteric model with reloading stiffness. The proposed model was a
bilinear approximation of the tri-linear model developed by Hisada et al. (1962) to
represent the hysteretic behavior of reinforced concrete structures.
In Clough and Johnston model, the degradation of the reloading stiffness was based
on the maximum displacement in the direction of the loading path due to which it was
referred as the peak-oriented model. The Clough and Johnston model is shown in
Figure 2.8 a and 2.8 b. Mahin and Bertero (1972) pointed out a deficiency in this model
that the original Cloughs model assumed that the reloading path is directed towards the
previous maximum response point after unloading process (Figure 2.8 a). To overcome
this deficiency, the model was modified such that the reloading occurs elastically until the
condition that the immediately preceding unloading point A is reached before the
response is directed towards the previous peak point C (see Figure 2.8 b). In the
modified Clough model the amount of stiffness degradation is a function of the peak
deformation.

(a)

(b)

Figure 2.8 Peak-Oriented degrading stiffness model: (a) Clough and Johnston
1966), (b) Modified model

86

Also, other researchers [Nielsen and Imbeault (1971), Anagnostopoulos (1972) and Iwan
(1973)] developed linear hysteretic models incorporating the stiffness degradation.
Sucuoglu and Erberik (2004) developed a simple linear hysteretic model based on the
stiffness - degrading model by Clough and Johnston (1966) but, the proposed model
incorporated the energy based strength degradation rule. The degradation rules for
asymmetric behavior incorporated in the deterioration model has been shown in Figure
2.9

Figure 2.9 Hysteretic model proposed by Sucuoglu and Erberik (2004)

where KD22 and KE22 are the respective stiffness at the hysteresis load paths, F denotes
force and D denotes displacement, A22,B22,C22,C22,D22,D22 denote the points in
hysteresis load paths, Ko the initial stiffness, Ku is the unloading stiffness, aKo is the yield
stiffness, Fy is the yield force, A1 to A6, B1 to B6 and C1 to C6 are constants and terms
1,2,3 represent hysteresis load paths

2.5.2 Takeda Model (2)


Takeda (1970) conducted experimental study to determine the behavior of a number of
medium-size reinforced concrete components under lateral load reversals for light to
medium levels of axial load. The Takeda model included (a) stiffness change at flexural
cracking and yielding, (b) hysteresis rules for inner hysteresis loops inside the outer loop,
and (c) unloading stiffness degradation with deformation. This model can be modified to
represent bilinear behavior by choosing the cracking point to be the origin. This modified
model is called as bilinear Takeda model (as shown in the last chapter). This model is
similar to the Clough model except that the bilinear Takeda model has more hysteresis
rules for inner hysteresis loops (Otani and Sozen, 1972; Otani 2002). Fukada (1969)

87

proposed a degrading trilinear hysteresis model that simulated the flexural stiffness
characteristics of RC structural members.

2.5.3 Bouc-Wen model and its modifications (3)


Bouc (1967) proposed a smoothly varying hysteresis model for single degree of freedom
(SDOF) systems subjected to the forced vibration. Wen (1980) generalized Boucs
hysteresis constitutive law and determined a procedure to obtain an approximate
solution for random vibration analysis. The procedure suggested by Wen (1980) was
based on the method of equivalent (or statistical) linearization.
Baber and Wen (1981) incorporated stiffness and/or strength degradation in Bouc
(1967) model. These parameters were adopted as a function of hysteretic energy
dissipation and were applied to multiple degree of freedom (MDOF) systems. This model
was further modified by incorporating the pinching effects and compatibility with Baber
and Wen and Wens equivalent linearization solution was maintained [Baber and Noori
(1985, 1986)]. The seismic response results obtained by the equivalent linearization
technique were comparable with the results obtained by Monte Carlo simulation. The
final model, proposed by Baber and Nooris (1986) was a single-element pinching model
and was known as the Bouc-Wen-Baber-Noori (BWBN) model as shown in Figure 2.10.
The modified Bouc (1967) model was used by many researchers for the analysis of
hysteretic systems subjected to the random vibrations. The improvements were made in
this model (Baber and Wen 1981; Baber and Noori 1986), and this advanced model was
able to predict the hysteretic behavior of structural elements with better accuracy. The
proposed model was based on separation of a linear restoring force component

(Fkk=(1-aKu) and a hysteretic restoring force component (Fh=(1-Ks) as shown in


Figure 2.10 (b,c), Therefore, the total restoring force can be represented as

Ft = a ' ku (1 a)kz

(2.11)

In the BWBN model the hysteretic displacement and the total displacement are related
by the following first order differential equation
g
g n 1
g

A
.
u

v
(

u
z
z

11
22
22
g
Z h( z )

22

n
z

(2.12)

where A11, 22,22, 22 are hysteretic shape parameters (if n22 = , the elasto-plastic
hysteretic case is obtained); and are strength and stiffness-degradation parameters
respectively (if 22/22 = 1.0, the model does undergo strength and stiffness degradation),
h(z)is the pinching function introduced by Baber and Noori (1986) [if h(z)=1.0, the model

88

does not pinch],

n 1

Z, Z

are respective displacements at n and n-1 cycles Denoting the

hysteretic energy dissipated at time t ,the strength and stiffness degradation parameters
can be defined by equation.

v( s ) 1 s

(2.13)

( s) 1 s

(2.14)

The effect of parameter was shown in Figure 2.11. The pinching effect can take
different forms which depends upon and several other parameters. An example of this
affect can be seen from dz/du and Z/Zu ( Ratio of displacement at a cycle and its ultimate
value) plot as shown in Figure 2.12 and 2.13.

There are as many as 13 parameters in

this model with very few of them related to engineering quantities. The values for these
parameters have been obtained for specific cases through identification of different types
of hysteretic systems. This model is very versatile and almost any hysteretic behavior
can be modeled, but it is complex and lacks relations with engineering design
parameters. Thyagarajan and Iwan (1990) showed that this model does not obey the
basic principles as it exhibited a higher drift at small value of post-yield stiffness because
of this reason it is rarely used except for theoretical studies.
Clarke (2005) developed an alternate model to the BoucWen model. This model
used the hyperbolic sine function as the driving function as opposed to the Fn function
(used by Bouc and many researchers subsequently). The proposed model was stable
and less costly and was preferred for nonlinear optimization process.

a)

b)

c)

Figure 2.10 Bouc-Wen model, separation of linear restoring force component from
hysteretic restoring force component: (a) Schematic model, (b) Non damping
linear restoring force component, (c) Hysteretic restoring force component
(Foliente 1995)

89

Figure 2.11 Effect of parameter (Foliente 1995)

Figure 2.12 Degradation in system properties: (a) Strength degradation, (b)


Stiffness degradation (Foliante et al. 1995)

Figure 2.13 Hysteretic pinching effect: Baber and Nooris (1986)

PinchIng-

Function (Foliante et al. 1995)

2.5.4 Ramberg-Osgood Model (4)


The Ramberg-Osgood proposed model (1943) is shown in Figure 2.14. It is a nondegrading smooth model, and describes loading and unloading/reloading curve in the
form of a mathematical equation. In this model, the value of the exponent r1dictates the
shape of the curve (r1=1) implies elastic behavior, and (r1=) implies elastic-perfectly
plastic

behavior).

cyclic

hardening

factor

can

be

easily

applied

to

the

unloading/reloading curve, but its value is highly dependent on history of the ground
motion, hardening parameters, strength and stiffness degradation parameters. This

90

model can be easily implemented in the Ruaumoko analysis platform (Carr 2003), which
also includes a large list of other hysteretic models (Carr 2003). In general, this model is
very rarely used in US, except for special conditions (Kukreti and Abolmaali 1999) and in
the auto and air/space industry. Another smooth non-degrading hysteretic model is
proposed by the Menegotto and Pinto (1973) which is available in OpenSees analysis
platform (McKenna 1997), the Bouc-Wen model (Bouc 1967; Wen 1976).

Figure 2.14 Force deformation relationships in Ramberg-Osgood model (Carr


2003)
th

where F and Fi is the force at i cycle and at first yield, K0 is the initial elastic stiffness
and is the deformation and i and y are the deformation at ith cycle and at yield.

2.5.5 Kunnath et al. (1990) deterioration Model (5)


The

inelastic deformations in reinforced concrete members not only concentrate in

critical sections but, spread across a finite region known as the plastic hinge length.
Kunnath et al. (1990) proposed a new deterioration model based on distributed plasticity
approach. This model was basically an extension and simplification of Takizawa (1975)
model. The Kunnath et al. (1990) model has been incorporated in versions of IDARC
(Valles et al. 1996). The model is simple and versatile and is based on integration of the
curvature diagram that results in a basic incremental moment-rotation relationship. From
this relationship, element stiffness matrix can be obtained in a closed form and can be
directly used in a computer program. The basic envelope of this model has been shown
in Figure 2.15.
Kunnath et al. (1990) incorporated shear deformation in this model and proposed an
improved version in which nonlinear distribution was developed for tapered elements
(Kunnath et al. 1992), and a multi-linear distribution with non-symmetric properties was

91

developed for the three dimensional model for reinforced concrete structures (Lobo
1994).The three model parameter proposed by Kunnath has been shown in Figure 2.15.

Figure 2.15 Three parameter model (Kunnath et al. 1990)

where Py and Pc are the yield strength and strength at crack, d is the depth of the
section, m is the maximum deformation under earthquake load, a, are the parameters
representing stiffness degradation, strength degradation and pinching effects

2.5.6 Sivaselvan and Reinhorn Model (6)


The Bouc (1967) model was modified by several researchers (Wen 1976; Baber and
Noori 1985; Casciati 1989; Reinhorn et al. 1995), and by Sivaselvan and Reinhorn
(2000) into a versatile hysteretic model (SHM) with pinching, stiffness and strength
degradation characteristics derived from the inelastic material behavior. Sivaselvan and
Reinhorn (2000) model was conceptually based on a general category of non-degrading
and degrading models developed by Iwan (1966) and an extended model developed by
Mostaghel (1999) to include smooth curvilinear segments. The Sivaselvan and Reinhorn
(2000) model presents (Figure 2.16) modeling of 1D inelastic material, and is derived on
the basis of the theory of visco-plasticity and resembles endochronic constitutive theory.
In addition, this model permits the use of a curved or multi-linear backbone curve. From
the review of previous literature it can be said that this model and its previous versions
do not consider cyclic (isotropic) hardening, and inclusion of strength cap and postcapping behavior considered by the model are not clear. Moreover, as compared to the

92

Bouc-Wen models the Sivaselvan and Reinhorn (2000) model has much closer relation
to physical quantities that are under the control of the designers.

a) Strength Deterioration

b) Stiffness Deterioration

Figure 2.16 Sivaselvan model to represent hysteretic behavior: (a) Strength


degradation, (b) Stiffness degradation (Sivaselvan and Reinhorn 2000)

where is the parameter representing stiffness degradation, K0 is the initial stiffness,


Kuis the unloading stiffness, My is moment at yield, is curvature, yis yield curvature, u
is unloading curvature, Mp = moment at pivot, My = My* (Mcur,fMy1) is on the right side of
the initial elastic branch and vice versa.

2.5.7 Yield Line Plastic Hinge Models (7)


The yield-line concept was introduced by Gioncu and Petcu (1997), Anastasiadis et al.
(2000), and Mller et al. (1997) to model the buckled shapes of structural members
observed in experiments. The yield-line models were constructed using plastic
mechanism to determine rotation capacity of European H-section beams. Lee and
Stojadinovic (2004) proposed a new cyclic yield-line plastic hinge model to determine the
connection rotation capacity based on a similar concept (Figure 2.17) considering the
post-peak connection strength degradation due to local and lateral - torsional buckling.
The limit state functions were used to establish a connection to predict crack initiation at
local buckles in the plastic hinge. In addition, low-cycle fatigue crack initiation based on a
cumulative local strain concept at the critical yield line were also considered. The primary
purpose of the Lee and Stojadinovic (2004) model is to evaluate new connections
analytically before the required proof-tests.

L f
h0
ya
2

93

Lw
ya y

(2.15)

U 2f L2f L f L f

U w2 L2w Lw Lw

(2.16)

(2.17)

where Lf and Lw are flange and web buckling wavelengths, is the length of flange, is
the yield - line mechanism rotation corresponding to beam plastic hinge rotation, Lf1 is
flange displacement, Lfw is a displacement at location ya where maximum web buckling
amplitude occurs, y0 defines the location of the center of rotation of the plastic hinge
mechanism, U is the buckling amplitude of fiber y1, n1, n2 are the angles of inclined
yield lines with respect to the cross section on which axial deformation is applied.

Figure 2.17 Yield line model proposed by Lee and Stojadinovic (2004)

2.5.8 FEMA 356 Component Model (8)


This model was first reported in the ATC-33 project, which led to the publication of the
FEMA 273 report and later to the FEMA 356 report. Structural components with reliable
ductility were modeled as shown in Figure 2.18. This model has a strength cap at point
C, and beyond this point, the strength drops rapidly (from C to D) to a residual value until
the deformation reaches a point (point E), where the strength drops to zero. This model
was intended to be used for static pushover analysis, and does not include any
hysteretic rules for cyclic deterioration. Even in the static analysis the steep cliff causes
many problems of numerical stability and leads to pushover curves with strange looking
saw-tooth behavior.

94

Figure 2.18 General force deformation behavior of structural components (FEMA


356, 2000)

2.5.9 Song and Pincheira Model (9)


This model incorporates a strength cap and post-capping behavior but ignores the cyclic
strength deterioration. The model deterioration rules are described in detail in Song and
Pincheira (2000) as shown in Figure 2.19. The backbone curve of the model is shown in
Figure 2.19(a). As seen, the model is capable to capture the strength cap after the
system exceeds the deformation, u which corresponds to the ultimate strength Fu. A
residual strength path is also incorporated in the model. Figure 2.19 (b) illustrates the
hysteretic behavior for loading cycles of increasing displacement amplitude. Stiffness
reductions upon unloading are considered after yielding and are controlled by a
parameter . Strength decay is introduced in this model only after the peak strength Fu
was reached. If the system unloads after the peak strength has been reached, reloading
is directed toward the mirror image of the point at which unloading occurred, i.e., points

A+ and B+ as shown in Figure 2.19 (c). However, additional strength decay may be
considered by increasing the deformation of the reloading target point by a parameter
(points A- and B-). The reloading target points represent the maximum attainable
resistance that the system can develop in subsequent cycles once the maximum
strength, Fu has been reached. The pinching of the hysteretic loops is considered in the
model by modifying the reloading stiffness of the system and is controlled by a
parameter as shown schematically in Figure 2.19 (c).
Strength degradation upon reloading is gradually reduced in subsequent cycles by a
parameter at a rate of n-1 per cycle, where n is the number of internal cycles.The
main disadvantage of the Song and Pincheira model is that it is not capable to capture
basic strength and post - capping strength deterioration unless if an approximate

95

procedure developed by Pincheira et al. (1999) is used to account for in-cycle


degradation. The procedure consists of using very small time increments together with
an arbitrary positive stiffness to compute a force unbalance, which is applied in the
subsequent time step.

(a)

(b)

(c)
Figure 2.19 Song Pincheira Model; (a) Backbone curve, (b) Hysteresis rules for
cycles of increasing deformation amplitude, (c) Hysteresis rules for small
amplitude or internal cycles

96

where A , B , A , B are hysteresis loops, and ,, are the parameters which


depend upon rate of strength and stiffness deterioration, Fu, Fy, Fr, Fnp are the forces at
respective points as mentioned in Figure 2.19 (a), + and - denote deformation in
positive and negative half cycle, pu and py are described in Figure 2.19.

2.5.10 Dutta and Das Model (10)


Dutta and Das (2002) proposed a model based on stiffness and strength deterioration
characteristics.

The

stiffness

and

strength

irregularities

were

proved

to

be

interdependent. This model has been discussed in detail in the previous Chapter.

2.5.11 Ibarras model (11) - This model is discussed in detail in the fifth Chapter.
The detailed comparison of models has been presented in Table 2.3.
Table 2.3 Comparison of hysteresis models
MT

CY

BAU

SCA

RSD

1
2
3
4
5
6
7
8
9
10
11

N
N
Y
Y
Y
Y
Y
Y
N
N
N

N
N
Y
Y
Y
Y
N
Y
N
N
N

N
N
N
N
N
N
Y
Y
Y
Y
Y

N
N
N
N
N
N
Y
Y
Y
Y
Y

Cyclic Deterioration
BAD
N
N
Y
Y
Y
Y
N
Y
N
Y
Y

PO
N
N
N
N
N
N
N
Y
N
N
Y

UO
Y
Y
Y
Y
Y
Y
N
Y
Y
Y
Y

AC
Y
Y
Y
Y
N
Y
Y
Y
Y
Y
Y

CY Cyclic hardening, BAU Bauschinger effect, SCA Strength capping, RSD


residual strength, BAD Basic strength deterioration, PO - Post capping strength
deterioration, UO Unloading strength deterioration, AC Accelerating stiffness
deterioration, Y represents that the above property has been adopted in the model and N
represents the reverse case

2.5.12 Selection of hysteretic model to represent the inelastic behavior


Based on review of the hysteresis model, the comparison between different hysteresis
models is presented in Table 2.3. From review of the models it was concluded that
majority of the hysteresis models were able to represent cyclic stiffness and strength

97

deterioration. However, the hysteresis model proposed by Ibarra et al. (2005) has been
adopted in the present study for inelastic modeling of the building frames. This is due to
he fact that calibration of parameters of inelastic model by Ibarra are based on large
number of experimental tests on structural members (Haselton and Deierlein 2007).
Moreover, Ibarraa model captures majority of failure modes of RC structural members
encountered in reality (Haselton and Deierlein 2007; Liel 2008; Haselton et al. 2011a,b).
Therefore, owing to its simplicity and efficiency in representing the inelastic behavior,
Ibarras model has been adopted for inelastic modeling.

2.6 Modeling of cracking behavior of structural elements


The elastic stiffness of the building frame is an important property which is required at
initial stages of seismic design to evaluate the lateral strength. However, the gross
stiffness obtained in the previous section yields deformation demands lower than the
actual demands even under the moderate seismic action. The seismic excitations
subject the structural members in RC moment resisting frames to moment reversals
along their length resulting in flexural cracking at ends with central un-cracked regions.
This phenomenon results in variation of moment of inertia (M.O.I) along the member
length. This M.O.I is influenced by magnitude and sign of moment, section geometry,
axial load and reinforcement. Also, the phenomenon of diagonal and shear cracking of
structural members which occurs due to shear and cyclic loading affects the stiffness
and M.O.I considerably (Paulay and Priestley 1992). However, majority of the seismic
design codes consider the gross stiffness of the structural members in estimating the
period and seismic demands which is inappropriate, as cracking is a realistic
phenomenon during the seismic excitation (Priestley 2003). Therefore, ignorance of
cracking aspect would induce errors in the seismic response and to get an realistic
estimate of period height relationship, effect of cracking should be considered. This
aspect has been ignored by majority of research works with few exceptions (Crowley
2003;Tan and Balendra 2007). The different approaches to model the cracking behavior
of structural elements have been discussed in next subsections (2.6.1 and 2.6.2).

2.6.1 Strength - dependent stiffness reduction factors


The stiffness reduction factors as described in the previous section do not account for
flexural strength of the member hence strength and stiffness of the flexural members are
assumed to be interdependent in proposing these reduction factors (Table 2.4 and Table
2.5). This assumption is very useful as it simplifies the conventional design process and
assuming the gross section properties, the stiffness and time period can be predicted at

98

initial stages of design process. This can be easily implied in force based design in which
the design spectral acceleration and required strength can be readily calculated using
these assumptions (Priestley, 2003). The result of the assumption that strength is
independent of stiffness is shown in Figure 2.22. However, many research works (Dutta
and Das 2002, Priestley 2003) based on their experimental and analytical studies
conclude these two properties (Strength and stiffness) to be interdependent. The
research work of Priestley 2003 is of specific interest regarding this aspect. Priestley
(2003) prescribes the expressions to determine yield curvature of structural members in
terms of geometrical properties of the members as given below in Table 2.6.
These equations can be accepted considering the assumption that the section is
considered to yield when both ends of the reinforcement yield. Due to this the yield
curvature of the section can be taken twice of the yield strain (Crowley 2004). Therefore,
if the yield curvature is assumed to be constant (Priestley 2003), then stiffness and
strength of the members are found to be interdependent as shown in Figure 2.20 a.
Therefore, considering this justification, elastic periods of the structure cannot be
accurately determined before the determination of final strength of the members. This
suggests that seismic design is an iterative process in which the member stiffness
should be updated at each iteration as shown in Figure 2.20 b (Priestley 2003).
Therefore, in evaluating the stiffness reduction factors, strength and stiffness should be
considered interdependent.

(a)

(b)

Figure 2.20 Moment Curvature (M ) relationship: (a) Assuming constant


stiffness (Priestley 2003), (b) Moment-Curvature relationship under realistic
condition i.e. considering strength and stiffness to be interdependent (Priestley
2003)

99

Table 2.4 Reduction factors proposed by Paulay and Priestley (1992)


Member

Range of Ig

Recommended Ig

Rectangular beams

0.30 0.50 I g

0.40 I g

T and L shaped beams

0.25 0.45I g

0.35I g

Columns

Pa 0.5 f c Ag

0.70 0.90 I g

0.80 I g

Pa 0.2 f c Ag

0.50 0.70I g

0.60 I g

Pa 0.05 f c Ag

0.30 0.50 I g

0.40 I g

Table 2.5 Effective member stiffness as per FEMA 356 (2000)


Flexural

Shear

Axial

Rigidity

Rigidity

Rigidity

Beams (Non-prestressed)

0.50 Ec I g

0.40Ec Aw

Beams (Prestressed)

Ec I g

0.40Ec Aw

Columns with compressions due to design


gravity load greater than equal to 0.5Agfc
Columns with compressions due to design
gravity load less than equal to 0.3Agfc

0.70 Ec I g

0.40Ec Aw

Ec I g

0.50 Ec I g

0.40Ec Aw

Es As

Component

Table 2.6 Expressions to determine the yield curvature (y) by Paulay and
Priestley (1992)
S.No

Component

Circular Columns

Rectangular columns

2.25 Y / d1
2.10 Y / d1

Rectangular cantilever walls

2.00Y / Lw

T-Section beams

1.70 Y / hb

where fc is concrete compressive strength and Ag,Ig is gross area and gross MOI of
column, D is the diameter of column, hc is the height of column, y or yield strain in
the column, hb is the height of the beam, Lws is length of shear wall, As and Es are
modulus of elasticity and area of reinforcement.

2.6.2 Member initial stiffness using analytical study


In this approach, the initial stiffness of the structural members in a analytical frame can
be determined from analytical formulations proposed by Priestley (2003). Priestley
(2003) determined yield stiffness and moment curvature relationship for a large square

100

column for different reinforcement, and for different levels of axial load ratio was
determined (Figure 2.23). From these curves, following observations were drawn at
(a) Yield curvature is insensitive to variation in the axial load ratio and in the
reinforcement ratio.
(b) Moment-curvature relationship of column has a strong influence of axial load ratio
and reinforcement ratio on moment capacity that can be clearly seen from the results of
the bilinear moment curvature. Priestley (2003) plotted curves between stiffness ratio
and axial load ratio for different reinforcement ratios as shown in Figure 2.21. The
stiffness ratio was calculated as EI/EIgross. from Figure 2.21 as reported by Priestley
(2003). Also, the dimensionless results presented in Figure 2.22 can be applied to
structures with material strength and column sizes other than the adopted in the present
study provided that EIgross has been used to calculate the effective stiffness of the
members.

Figure 2.21 Moment curvature for a rectangular column (Priestley 2003)

Figure 2.22 Effective stiffness of a large rectangular column (Priestley 2003)

101

where Nu is the axial load, fc is the concrete compressive and Ag is the gross cross
sectional area of column

2.7 Quantification of irregularity


As discussed in the earlier section, the code-defined approaches quantify the irregularity
limits in terms of magnitude only and the effect of location of irregularity is ignored.
Therefore, the first main aim of the present study is to quantify the irregularity in terms of
its magnitude and location. The irregularity can be effectively captured by the dynamic
response parameters like mass participation factors and frequency of vibration.
Therefore, to ascertain the dominance of these parameters on seismic behavior of
buildings, sensitivity analysis has been conducted on building models as described in
previous sections of this chapter.

2.7.1 Sensitivity analysis


Sensitivity analysis has been one of the important tools which is often used to determine
the effect of input parameter on the output parameter. The sensitivity index can be
defined as the ratio of the standard deviation of the input parameter to that of the output
parameter in accordance with Kose et al. 2008. To determine the effect of dynamic
response parameters on seismic deformation demands, sensitivity analysis was
conducted and the results have been expressed in Table 2.7. In this study effect of the
dynamic response parameters have been studied on seismic deformation demands,
since these are the critical parameters which represent the seismic performance of the
structure. From the results presented in Table 2.7, it can be clearly seen that the
participation factor dominates the frequency of vibration for the irregular building frames
considered in the analytical study. Therefore, based on the results of sensitivity analysis
the irregularity index is proposed as
k

c
j

Pi
Pr

(2.18)

where c, Pi , Pr are the proposed irregularity index, combinations of participation factor


from jth to the kth mode for irregular and regular buildings for long period structures.
From equation.2.18, it can be observed that in representing the irregularity in a
building system, higher modal contributions are included, as nonlinear behavior of the
structures is sensitive to the higher modal contributions (Jones 2012). Moreover,
ignorance of higher modes would result in inaccurate representation of the seismic
response parameters (Chopra and Goel 2002; Chopra ad Goel 2004).

102

The participation factor for k th floor ( Pk ) can be obtained from the equation as
suggested by Rayeleigh as

Pk

Mk
M mm

(2.19)

The modal mass (Mk) for each mode can be obtained as


2

Wkik

M k n1

g Wk (ik ) 2
1

(2.20)

and frequency of vibration is obtained by

K M 2 0

(2.21)

where Wk and Mk are seismic weight and modal mass of k th floor and Mmm is the total
modal mass of the building and ik is the mode shape coefficient for that floor.

Table 2.7 Sensitivity analysis results


S.No

Parameter

Mass
participation
Frequency
of
factor
vibration

Sensitivity index with respect to the deformation


demand parameter
IDR

rd

0.53
0.39

0.64
0.52

0.58
0.46

The proposed irregularity index assumes a value between 0.43 1 for building models
considered in the analytical study. The proposed irregularity index varies with building
properties and irregularity. It is worthwhile to study the variation of the irregularity index
with these properties. The mean irregularity index is plotted with building properties from
which it could be observed that
(a) The proposed irregularity index decreased with increase in the magnitude of mass,
stiffness and strength irregularity. However, this decrease was more pronounced in case
of stiffness, strength and setback irregularities (Figure 2.23).
(b) The proposed irregularity index is affected by location of irregularity and it assumed
minimum value for the case when mass irregularity is at bottom storey and it increased
with height of location of mass irregularity. This pattern reversed in case of strength,
stiffness and setback irregularity (Figure 2.24). proposed irregularity index decreased
with increase in eccentricity and for greater reduction in setback dimensions (Figures
2.25 and 2.26). In addition, consideration of cracking affects decreased the irregularity
index (Figure 2.25 to 2.27).

103

108
90

ST1
ST2

H (m)

72

S1
S2

54

M1
M2

36
18
0.78

0.89

Mean c
Figure 2.23 Variation of irregularity index with magnitude of mass, strength and
stiffness irregularity (M1 = Mean of mass irregularity of magnitude 200%, 400%,
600%, M2 = mean of mass irregularity of magnitude 800 % and 1000 %, S1 and ST1
represent stiffness and strength irregularity (mean) of magnitude 25% and 50%
respectively, S2 and ST2 represent stiffness and strength irregularity (mean) of
magnitude 75% and 100%)

108
90

ST1BS
ST1MS
ST1TS
S1BS
S1MS
S1TS
M1BS
M1MS
M1TS

H (m)

72
54
36
18
0.79

0.88

0.97

Mean c
Figure 2.24 Variation of irregularity index with location of mass, strength and
stiffness irregularity (BS, MS and TS denotes irregularity location at bottom one
third, middle one third and top one third storeys)

104

108
ST1

90

ST1C

H (m)

72

S1
S1C

54

M1
36

M1C

18
0.8

0.9

Mean c
Figure 2.25 Variation of irregularity index with mass, strength and stiffness
irregularity considering the cracking effects (M1C,S1C,ST1C denotes M1,S1,ST1
with cracking effects)

108

P1C

90

P2C
P3C

H (m)

72

P1
54

P2
P3

36

18
0.88

0.94

Mean c
Figure 2.26 Variation of irregularity index with plan irregularity with and without
cracking affects (P1,P2,P3 denotes building models with eccentricity varying from
0.05 bw and 0.10 bw, 0.15b and 0.20 bw, 0.25 bw and 0.30 bw; PIC, P2C and P3C
models represent P1,P2,P3 models with cracking effects)

105

108

90

SEA
C
SEA

H (m)

72

SEB
54

SEB
C
SEC

36

SEC
C
18
0.46

0.68

0.9

Mean c

Figure 2.27 Variation of irregularity index with setback irregularity with and
without cracking effects (SEA,SEB,SEC denotes building model with setback
originating from bottom one third, middle one third and top one third storeys,
SEAC,SEBC,SECC denotes the respective models with cracking consideration)

2.7.2 Evaluation of the proposed irregularity indices for different irregular


building models
The proposed irregularity index as discussed in the previous section is based on
dynamic response parameters which is a more efficient approach as compared to
previous approaches for quantification of irregularity. But, it is worthwhile to evaluate the
irregularity limits as per proposed approach and the other approaches to illustrate the
effectiveness of the proposed approach.
(a) Building models with mass, stiffness and strength irregularity
To illustrate the effectiveness of the proposed approach, the irregularity indices for the
irregular building models with a grid plan of 16m x 20m (Figure 2.28) have been
evaluated using both code approach and the proposed irregularity index. The irregularity
limits considered are as follows (a) Mass irregularity limit 600 % with irregularities placed
at bottom, middle and top one third storeys (A,B,C), (b) The stiffness and strength
irregularity limits considered as, 50 % with different location of irregularities (A, B, C)

106

respectively. The rectangular beam column sizes have been considered for analytical
study (main beam and column sizes considered are 0.30m x 0.45m, and 0.4 x 0.55 with
loading, material and ground motion details as described in previous Chapter).Table 2.9
shows the irregularity indices evaluated using code proposed indices (ASCE 7: 2005; IS
1893:2002) and the proposed index. From Table 2.8, it can be concluded that although
the building frames contain irregularities at different locations, still the code defined
approaches prescribe the same limits of irregularity. However, the proposed approach
was observed to be more efficient as compared to the code approach.

(a) Plan

(b) Elevation

Figure 2.28 Plan irregular Building models for study of variation of proposed
irregularity index

Table 2.8 Comparison of irregularity indices of seismic design codes and


proposed method for building models with mass, stiffness and strength
irregularity
Mass irregularity Models
c

Stiffness irregularity
c

Code

(%)

ASCE
limit IS

600

0.97 D

50

1.5

1.5 0.58 G

50

1.5

1.5 0.51

600

0.95 E

50

1.5

1.5 0.66 H

50

1.5

1.5 0.60

600

0.93 F

50

1.5

1.5 0.73 I

50

1.5

1.5 0.68

MT

MT

Code
models
(%) ASCE
LimitIS

MT

Code
Models
LimitIS
(%) ASCE

Strength irregularity
ST

(b) Building models with setback irregularity


The setback irregularity has been identified by several seismic design codes. As per IS
1893:2002, a building is said to be setback irregular when the lateral dimension of the

107

building frame in any storey is greater than 150% of the lateral dimension of the adjacent
storey. As per other seismic design codes, the above prescribed limit is 130%. The
setback limit as per different codes has been shown in Table 2.9.
The pictorial representation of setback limit as per IS 1893:2002, ASCE 7:05,
EC8:2004 is shown in Figure 2.29 from which it can be observed that the codes consider
the ratio of lateral dimension of one storey to that of the adjacent storey to define the
setback irregularity. This definition ignored the presence of setbacks in other floors and
gradual variation of the setback irregularity was ignored. This results in inaccurate
representation of the setback irregularity. To address the above issue, several
researchers have proposed the irregularity index to define the irregularity. Karavasilis et
al. 2008 a (following the approach of Mazzolini and Piluso 1996) proposed a pair of
irregularity indices to represent the setback irregularity. The parameters proposed by
Karavasilis et al. (2008a) have been represented in equations 2.22 and 2.23, and the
definition of terminologies in these equations has been expressed pictorially in Figure.
2.30.

1 ns 1 Li

ns 1 1 Li1

1 b Hi

nb 1 1 Hi1

(2.22)

n 1

(2.23)

where n s represents the number of storeys in the building model, and nb represents the
number of bays in the first storey of the building model. Li and H i are the width and the
height of the i th storey.
The parameters s and b as proposed by Karavasilis et al. (2008 a) vary from 1 to
1.4, and from 1 to 2.39 respectively. From equations 2.22 and 2.23 it can be clearly
seen that the indices proposed by Karavasilis et al. (2008 a) depend upon physical
geometry of the setback alone and it is inconvenient to use two indices to represent the
same setback frame. Furthermore, the Karavasilis indices were based on the
assumption that the beam and column sizes are uniform throughout their length and
building mass is uniformly distributed throughout height and length of the frame. These
assumptions are unrealistic from practical considerations; therefore, there is a need of a
simple parameter that describes all aspects of setback irregularity effectively.
The author proposed index effectively addresses these shortcomings. However, It is
worthwhile to compare it with the irregularity index proposed with the code approach and
Karavasilis et al. (2008a) approach. To achieve this purpose, the irregularity indices by
the different approaches were evaluated for the building frame shown in Figure. 2.31
(subjected to its self-weight as the dead load). The building frame consisted of eight

108

storeys with five bays (with a bay width of 4m) in both transverse directions (X and
Z).The results of irregularity indices for frames considered in Figure 2.31 has been
shown in Table 2.10 (with input details similar to those adopted in case of mass, stiffness
and strength irregularity). From the results presented in Table 2.11 it can be seen that
although irregular building frames A,B and C had different configurations of setback still,
the code-defined approaches codes prescribe the same irregularity limit. Therefore, code
approaches are ineffective in capturing the variation of setback irregularity. Apart from its
limitations (as discussed earlier), the Karavasilis et al. (2008a) approach also yielded the
same value of irregularity indices for frames A, B and C (Figure.2.31) in spite of different
setback geometries of these frames. Therefore, it can be said that both code approach
and Karavasilis approach exhibit a similar performance in quantifying the setback
irregularity of frames A, B and C. However, the proposed approach effectively captures
the variation in the setback irregularity as evident from Table 2.10.

Table 2.9 Code prescribed limits of setback irregularity

S.No

Code Name

Setback Irregularity Limit

IS 1893:2002

150 %

EC8:2004

130%

UBC 97

130%

NBCC 2005

130%

IBC 2003

130%

TEC 2007

ASCE 7.05

130%

(a)

(b)

(c)

Figure. 2.29 Code limits of setback irregularity: (a) IS 1893:2002, (b) ASCE 7.2005
(c) EC 8:2004

109

Figure 2.30 Frame Geometry for definition of irregularity indices proposed by


Karavasilis et al. (2008a)

Figure 2.31 Setback Frames considered for the comparison of the irregularity
index
Table 2.10 Comparison of the irregularity indices for setback structures
Karavasilis
approach
s
b

MT

IS 1893:2002

ASCE
7.05

0.8

0.5

1.309

1.75

0.54

0.8

0.5

1.309

1.75

0.46

0.8

0.5

1.309

1.75

0.61

1.75

EC8:2004

110

(c) Building models with plan irregularity


Plan irregularity has been identified by several seismic design codes and researchers as
discussed in previous chapter. To evaluate the effectiveness of the proposed index in
capturing the variation of plan irregularity, a building frame as shown in Figure 2.28 was
considered (with input details similar to those adopted for mass, stiffness and strength
irregular structures). The plan irregularity was created by varying the center and corner
frame by a distance of 0.15bw as separate cases. The code approach of eccentricity
failed to differentiate between the movement of center frame and corner frames (Model A
Mean of cases in which center frame is moved to left and right; Model B Right Corner
frame is moves towards center frame; Model C Left corner frame is moved towards
center frame). However, in reality this aspect does affect the seismic response
(Moghadam 1998). Therefore, it can be said that the code approach is inefficient in
capturing the effect of plan irregularity. Moreover, on comparing the values of proposed
approach for different frames (Table 2.11) it can be said that the proposed approach is
more efficient in representing the plan irregularity. This approach is simple and can be
easily incorporated to evaluate the inelastic seismic response. This aspect has been
discussed in detail in the coming chapters.
Table 2.11 Comparison of the irregularity indices for plan irregular structures
MT

IS 1893:2002

ASCE 7.05

EC8:2004

A
B
C

0.15
0.15
0.15

0.15
0.15
0.15

0.15
0.15
0.15

0.935
0.903
0.865

(d) Evaluation of proposed index to capture the effects of cracking of structural


members
As discussed in the previous section the cracking aspect effects the seismic response
significantly. The review of seismic design codes and previous literature shows that there
is absence of a parameter to effectively represent the cracking effects. As the proposed
index is based on dynamic response of buildings it can effectively capture the cracking
effects. However, it is worthwhile to check the applicability of the proposed index
regarding the cracking aspect. To achieve this purpose the building models as described
in sections (with input detail similar to previous section) are considered and the stiffness
reduction factors as per Priestley (2003) have been used. The modeling of cracking
phenomenon is discussed in the previous section. The irregularity limits for cracked and
un-cracked frames have been presented in Table 2.12 to 2.14 from which it can be

111

clearly seen that the proposed approach is effective in capturing the effects of cracking
as well.

Table 2.12 Evaluation of the proposed irregularity indices for mass, stiffness and
strength with cracking considerations
Mass irregularity models

MT

Code

(%)

limit
ASCE

Stiffness irregularity

Strength irregularity

models

models

Code

(%)

Limit

MT

IS

ASCE

IS

MT ST

Code

(%)

Limit

ASCE

IS

600 6

0.93 D

50

1.5

1.5

0.54 G

50

1.5

1.5 0.47

600 6

0.90 E

50

1.5

1.5

0.63 H

50

1.5

1.5 0.54

600 6

0.87 F

50

1.5

1.5

0.69 I

50

1.5

1.5 0.63

Table 2.13 Comparison of the irregularity indices for setback structures

with

cracking considerations
Karavasilis
approach
s
b

MT

IS 1893:2002

ASCE
7.05

0.6

0.5

1.309

1.75

0.51

0.6

0.5

1.309

1.75

0.42

0.6

0.5

1.309

1.75

0.57

1.75

EC8:2004

Table 2.14 Comparison of the irregularity indices for plan irregular structures with
cracking considerations
Frame type

IS 1893:2002

ASCE 7.05

EC8:2004

0.15

0.15

0.15

0.90

0.15

0.15

0.15

0.88

0.15

0.15

0.15

0.86

112

2.8 Brief summary and main conclusions


This Chapter begins with a brief introduction of structural irregularity aspect and its
impact on seismic response of buildings. Then a detailed description of building models
used in the analytical study has been presented with description of different types,
magnitude and location of irregularities. The analysis method has a significant impact on
accuracy of the seismic response obtained. In this Chapter an effort has been made to
review the analysis methods available with respect to their efficiency and accuracy.
Based on this review, appropriate analysis method has been adopted for analytical
study. As discussed earlier in this chapter, the code procedures to estimate seismic
demands are mainly based on SDOF system and elastic analysis which is unrealistic.
Therefore, to estimate the realistic seismic demands, inelastic modeling and analysis
needs to be adopted. To achieve this purpose, different models (as proposed in previous
research works) to incorporate inelasticity have been reviewed with respect to their
efficiency, merits and demerits in representing different modes of deterioration. An
appropriate model has been chosen to represent the inelastic behavior of the irregular
building frames. Finally, an irregularity index which overcomes the limitations of the
previous indices to represent irregularity has been proposed. The proposed index has
been evaluated for building models with different types of irregularity and compared with
code approaches and previous research works. The main conclusions from this Chapter
are
(a) From the analysis methods available, time history method of analysis has been
adopted from the analytical study (in spite of its complexity) to achieve accurate estimate
of the seismic response. In the present analytical study, inelastic seismic response has
been evaluated to get a realistic estimate of the seismic response.
(b) The review of different inelastic models showed that majority of the models were
based on SDOF systems which is unrealistic. Moreover, these approaches were
complex and tedious to evaluate. In comparison, hysteresis model proposed by Ibarra
was simple and captured important modes of failure of structural members. Moreover,
calibrations of this model were based on rigorous experimental tests on RC structural
members (Haselton and Deierlein 2007). Therefore, this model has been adopted to
determine the inelastic seismic response.
(c) The review of the seismic design codes and literature works indicated that the aspect
of irregularity location has always been ignored in quantification of irregularity. This
approach was unrealistic as irregularity location has a significant impact on the seismic
response (Nassar and Krawinkler 1991; Al Ali and Krawinkler 1998; Das and Nau
2003). In the present research work, these limitations are overcome and an improved

113

irregularity index which captures both magnitude and location of irregularity has been
proposed by the author. The results of analytical study showed that the code approach
and previous research work (Karavasilis et al. 2008 a) were observed to be inefficient in
capturing all aspects of structural irregularity and cracking. However, the proposed
irregularity index achieved this purpose (as evident from Table 2.8, Tables 2.10 to 2.11).
Thus, the proposed index could be preferred over other approaches to capture the
aspect of irregularity effectively. Moreover, the author proposed has been applied to
estimate seismic response parameters (fundamental time period, deformation demands,
collapse capacity etc.) as discussed in the next Chapters
(d) Cracking of structural members is a realistic phenomenon during seismic excitation.
However, this aspect has been ignored by majority of seismic design codes and
research works in formulating the seismic design philosophies. Nevertheless, few
researchers (Paulay and Priestley 1992; Priestley 2003) have addressed this aspect by
proposing appropriate stiffness reduction factors. The comparison of different
approaches to incorporate cracking showed that the Priestley (2003) approach was most
suitable as it considered the interdependency of strength and stiffness which is closer to
reality and consistent with previous research works (Crowley 2003;Priestley 2003;Dutta
and Das 2002). Therefore, this methodology has been adopted to determine the model
the cracking behavior. Finally, the results of analytical study showed that the proposed
index was efficient in effects of cracking as well. (Tables 2.12 to 2.14).

114

CHAPTER 3
DETERMINATION OF FUNDAMENTAL TIME-PERIOD OF IRREGULAR
BUILDINGS

3.1 Introduction
The estimation of fundamental time period of vibration is a critical step in seismic design
and analysis of the building structures as it is a representative of global seismic demands
of the structure. The period of the building mainly depends upon building properties like
mass, stiffness, seismic excitation, storey height, number of storeys, cracking etc. In
reality, the building models often encounter different forms of structural irregularity and
cracking. The structural irregularity and cracking both affect the fundamental properties
like mass and stiffness and hence alter the fundamental time period. However, these
aspects have been ignored in code proposed empirical expressions to estimate the
fundamental time period. The seismic design codes generally relate the fundamental
time period and height of the structure in the form of simple empirical expressions. These
relationships have been idealized for force based design and hence yield conservative
estimates of the seismic response. However, in reality deformation demands are the key
parameters which represent the seismic performance of the structure (Crowley and
Pinho 2006). Therefore, a relationship which accurately predicts the period height
relationship is very essential. In this Chapter, an effort has been made to propose
realistic period - height relationship at first considering the phenomenon of cracking and
structural irregularity using different methodologies. Thereafter, a brief comparison has
been made between results evaluated using these methodologies. Finally, the
applicability of the proposed period - height relationships in seismic design
methodologies and in seismic vulnerability assessment has been briefly discussed.

3.2 Building period - height relationship proposed by seismic design


codes and previous research works
As discussed in the previous section, the period height relationships are very essential
in determining the seismic demands. The code proposed period height relationship
was initially obtained by conducting the regression analysis on time period data set
consisting of experimentally determined periods of buildings. It can be observed from the
previous research works (ATC 1978; Goel and Chopra 1991) that these data sets were
very few in number because of unpredictable nature of earthquakes and very few
buildings were equipped with period measuring instruments. These databases were
collected by researchers and based on regression analysis conducted on these

115

databases simple empirical expressions to estimate period were proposed in terms of


height due to following reasons
(a) Height of the structure was found to be the parameter which had major influence on
the period (Goel and Chopra 1991; Hong and Hwang 2000; Dym and Williams 2008).
(b) These equations can be readily used to estimate the seismic design parameters like
base shear and moment without prior knowledge of cross sectional dimensions of the
structural members.
(c) These expressions can be readily incorporated in initial stages of spectrum based
displacement based design (DBD), and later stages of performance based design (PBD).
The empirical formulae to estimate the fundamental time period appeared first in U.S.
building code ATC3-06 (ATC 1978) as

T Ct H 0.75

(3.1)

where Ct was assumed as 0.03 for reinforced concrete moment resisting frames and
height H is expressed in feet. The numerical value Ct was obtained from regression
analysis of buildings during the 1971 San Fernado earthquake. Simplifying equation 3.1,
other expression to estimate the fundamental time period was proposed in NEHRP
(1994) as

T 0.1N

(3.2)

However, the proposed expression was restricted to buildings with twelve storey height
with a minimum storey height of 3.05m (10ft). The European seismic design code (EC
8:2004) proposed period - height relationship as

T 0.075H 0.75

(3.3)

As previously mentioned this empirical relationship was based on Rayleighs


assumptions. The height H is expressed in meters. Therefore, the coefficient 0.03 is
converted to 0.075. This formula was originally proposed based on experimental study of
buildings in California. Moreover, due to differences in soil condition, climate and
difference in construction practices, this formula should be modified for European region.
Goel and Chopra (1997) proposed an improvement of period expression in ATC
3.06. The experimental period values for eight earthquakes occurred in California
(beginning with 1971 San Fernado earthquake and ending on 1994 Northridge
earthquake) were collected. The period data was collected for 42 steel moment resisting
frames (S.M.R.F), 27 RC frames and 16 SW frames. The periods were measured in both
the orthogonal directions. In addition, Goel and Chopra (1997) evaluated the periods for
these buildings based on analytical expression proposed in ATC 1978, and observed
that the predicted periods were close to the lower bound value of ATC 1978 expression

116

for buildings up-to 160 ft height. However, for building ranging from 160 ft - 225 ft, the
code equation underestimated the period values as compared to the experimental
observation.
Therefore, based on best fit obtained from regression analysis, Goel and Chopra
(1997) proposed the generalized equation to estimate fundamental time period as

T h H h

(3.4)

The authors conducted different types of regression analysis with and without constraint
with higher scatter obtained in unconstrained regression analysis with value of = 0.92.
However, the scatter was observed to be less in case of constrained analysis with =

0.90.
Figure.3.1 shows the best fit + standard deviation, and best fit - standard deviation
for RC MRF proposed by Goel and Chopra (1997). In Figure. 3.1 a difference of periods
was obtained for different cases of ground motion (uig < 0.15g and uig > 0.15g). This
difference is attributed to the cracking phenomenon during seismic excitation.

Figure 3.1 Results of RC MRF frames using eigen - value analysis (Goel and
Chopra 1997)

Hong and Hwang (2000) showed that the difference in construction practices do
effect the fundamental time period significantly. Hong and Hwang (2000) experimentally
determined the fundamental time period of RC moment resisting frames located in
Taiwan through vibration measuring instruments. Based on the experimental results, the

117

empirical relationship between building period and height was derived. However, the
obtained relation was different from that of U.S. building code formula. On comparing the
numerical values of fundamental time period it was observed that the formula based on
data pertinent to Taiwanse buildings under predicted the period value as compared to
U.S. Code proposed formulae. This implies that Taiwanese buildings are stiffer than the
Californian buildings. However, the code proposed expressions have a major
disadvantage that these expressions are based on Rayleighs analysis which is based on
following assumptions as follows
(a) Lateral forces are distributed linearly over height of the structure.
(b) Base shear is proportional to (1 / T 3 ) .
(c) Weight of the building is distributed linearly over height.
(d) Deflected shape of building under application of lateral forces is linear over its height.
These assumptions are unrealistic and contradictory to reported research works (
Karavasilis et al. 2008 a; Athanassiadou 2008; Ricci et al. 2011 a). In addition, these
expressions were based on the data of building located in a certain region and subjected
to a certain ground motion. Therefore, it could not be applied universally to all buildings.
Lee et al. (2000) aimed at evaluating the reliability of fundamental time period
expressions proposed by different seismic design codes of practice (Korean code UBC
97; NBCC 1995) for shear wall dominant building systems situated in South east Asia.
To achieve this purpose, the fundamental time period of 50 apartment buildings with
shear wall dominant systems was experimentally determined. A large database was
generated comprising of time periods of these buildings. Simplified expressions were
proposed to estimate the fundamental time period based on regression analysis
conducted on this database as
T b1

1
H b 2 b3
Lsw

(3.5)

where T is the fundamental time period, Lsw is the length of shear wall, H is the building
height, b1.b2,b3 are constants. The proposed equations showed a close agreement with
dynamic analysis and a large difference on comparison with the code proposed
equation.
Like wise, Balkaya and Kalkan (2003) proposed empirical equations based on
analysis of 3D finite element modeling techniques (using E Tabs software) for buildings
situated in Turkey. Balkaya and Kalkan (2003) generated a database consisting of time
periods of 80 buildings constructed using tunnel form construction techniques prevailing
in Turkey (Figure 3.2) Based on regression analysis conducted on this database, the
equation to estimate fundamental time period was proposed as

118

b15
T C11 H b1111b12 asb1311alb141111min
J b16

(3.6)

where T is the fundamental time period of the building, H is the total height of the
building, as11 and al11 are the ratio of short and long side shear wall area to the total
floor area. 11min is the ratio of minimum shear wall area to the total floor area, C11, b11 to

b16 are the constants as specified by the Balkaya and Kalkan (2003) based on the
regression analysis (Table 3.1). The comparison of results of proposed equations with
the dynamic analysis results showed a close agreement between both these methods.

Figure 3.2 Tunnel form construction technique used in Turkey (Balkaya and
Kalkan 2003)
Table 3.1 Empirical equations to predict fundamental time period of tunnel form
buildings
Plan
S
R

C11

b11

b12

b13

b14

b15

0.158
0.001

1.400
1.455

0.972
0.170

0.812
-0.485

1.165
-0.195

-0.719
0.170

b16
0.130
-0.094

Navarro and Oliveira (2004) experimentally determined time periods of 235 buildings.
Furthermore, regression analysis conducted on the analysis results the empirical
expression to estimate the time period was proposed as

T 0.045N
Where N is number of storeys

119

(3.7)

Ghrib and Mamedov (2004) presented a new approach to estimate the fundamental
time period and incorporated the effect of flexible foundation system in determination of
fundamental time period of shear wall dominant systems. The fundamental time period of
set of buildings was determined experimentally and based on the regression analysis the
empirical expression to estimate the fundamental time period was proposed as

2
where ,m,EI

are

EI 114
m(1 11112 )

(3.8)

frequency, mass and flexural rigidity of building, 11,11 are the

constants depending on the building properties.


The proposed equation was observed to be in good agreement with experimental
results. However, code approaches (UBC 97; NBCC 2005) showed a large difference
with experimental results that showed the inefficiency of the code approach.
Wallace and Moehle (1992) proposed empirical expression of fundamental time
period for building systems with coupled shear wall as
1/ 2

wh sw
H

T 6.2 N
2 Lsw C12 gE p

(3.9)

where H, hsw N, Lsw are height of the building, height fo the shear wall, number of
storeys, Lsw and shear wall length respectively, C12 is a constant, Ep,w,w are modulus of
elasticity of shear wall, unit floor weight including tributary shear wall weight, ratio of
shear wall area to floor plan area for walls (aligned in the direction in which period is
calculated) respectively.
Wang and Wang (2005) extending the approach of Wallace and Moehle (1992)
presented a new analytical formulation to determine the fundamental time period of
buildings with coupled shear wall. The analytical formulation was mainly based on
Sturm-Liouville differential equation and involved constants like h and . The proposed
expression to estimate parameters 13 and 13 are presented as

142 k i3132 ( EI )1

(3.10)

13 1 I ( Aswa112 ) 1

(3.11)

where a11, 13 , 14 , 13 are constants, Asw is the area of the coupled shear wall
The proposed expression was observed to be more simpler as compared to the previous
approaches and yielded comparable results with code equations and previous
approaches. Crowley (2003) and Crowley and Pinho (2006) observed that cracking of
RC structural members has a significant impact on the fundamental time period of
buildings. The un-cracked and yield fundamental time period of European RC buildings

120

using eigen - value analysis were determined. Based on the regression analysis
conducted on the analysis results the equation to estimate the fundamental time period
was proposed as

T 0.055H

(3.12)

where H is the height of the building


Kose (2008) determined the fundamental time period of building frames with
masonry infills. In this study, the effects of parameters like building time period, number
of bays, shear walls, infill panels on fundamental time period of buildings were evaluated.
To determine the fundamental time period of buildings, 189 building models with
selected parameters using SAP software were modeled. To incorporate non-linear
behavior of structures, iterative modal analysis was employed. Based on regression
analysis conducted, the equation to estimate the period was proposed as

T 0.0935 0.0301H 0.0156B11 0.039F11 0.1656S11 0.0232l

(3.13)

where B11 is Number of bays, F11 denotes frame type with value of 1, 2 and 3 for
masonry infill frame, open first floor and bare frames, S11 is the ratio in percentage of
shear wall area to total floor area, I is the area ratio of infill walls to total panels.
The analytical results showed that the presence of infillls and shear walls reduced the
fundamental time period. The proposed equation was compared with the code equation
and dynamic analysis which showed that the code equations under predicted the
fundamental time period, and proposed equations yielded better estimate of fundamental
time period as compared to the dynamic analysis.
Guler et al. (2008) computed the fundamental periods of some RC buildings,
considering the effect of infill walls using ambient vibration tests and elastic numerical
analyses. A period-height relationship relevant to Turkish RC moment-resisting frames
was derived for a fully elastic condition and was proposed as

T 0.026H 0.90

(3.14)

Furthermore, results showed that the Goel and Chopra (1973) expression and UBC code
expression overestimated the seismic response which in turn led to overestimation of the
seismic demands. However, the proposed expression (equation 3.14) were observed to
be in close agreement with the experimental results.
Some existing RC buildings typical of the Egyptian building stock were studied
through ambient vibration measurements by Sobai et al. (2008). Period values from in
situ tests were compared with the Egyptian code formula for RC buildings, as well as
with the results of numerical simulations performed considering different values of some
structural parameters (mechanical properties of concrete and of masonry infills). Results
confirmed that the characteristics and properties of infills strongly influence the period

121

values of RC framed buildings. Also, the Egyptian code proposed formula (i.e. T = 0.1N)
was observed to overestimate the period values as compared to the

experimental

results of period. However, the experimental values were in close agreement with those
provided in other studies based on ambient vibration measurements carried out on
buildings from other countries (Gallipoli et al. 2009)]. The elastic (Te) and yield (Ty)
values of the period, period elongation (stiffness degradation) during and after strong
ground shaking is also an issue of great interest and it significantly influences the period
height relationship (Calvi et al. 2006).
Oliveira and Navarro (2010) experimentally and analytically determined the
fundamental time periods of existing building frames and expressed it as a function of
building height. The building models adopted had different types of irregularities. The
results of analytical study showed a close agreement between both the results.
Masi and Vona (2010) experimentally determined the fundamental time period of a
realistic building subjected to Malese earthquake including the aspect of masonry infills
and cracking. The building structures studied suffered moderate to heavy damage under
seismic actions. The results showed that the fundamental time period is strongly
influenced by the input ground motion and seismic damage and it increased with
intensity of ground motion and with increase in the seismic damage.
Ricci et al. (2011) through analytical study confirmed the effect of masonry infills on
the fundamental time period and the time periods of a set of buildings in Portugal
incorporating the effect of masonry infills by conducting the eigen - value analysis were
determined. Based on regression analysis conducted on the fundamental time period
results were proposed as

Tx 0.012H

(3.15)

Ty 0.016H

(3.16)

Meziane et al. (2012) experimentally determined the fundamental time period of


realistic buildings and compared it with Algerian Seismic Code (RPA 99) for seismic
vulnerability assessment. The results obtained were extrapolated to buildings of the
same topology built during the 1949 to 1954 period in the northern part of Algeria. The
analysis results showed that the Algerian code proposed empirical formulas consider
only the geometrical dimension (length, width and height) and the structural design of the
buildings. The fundamental periods of vibration of twenty-two buildings, located in Algiers
were calculated using the empirical formulas given in the RPA 99 and were observed
lower to the experimental results. Based on experimental results, the fundamental time
period in longitudinal and transverse directions for seismic vulnerability assessment were
prescribed as

122

Fi 0.1683.

H
1.76880
bw

(3.17)

Fi 0.35716.

H
2.39782
bw

(3.18)

where bw is width of the building


Carrillo and Alcocer (2013) based on their experimental studies on realistic low rise RC
buildings proposed the empirical expressions as

T 0.09

H
Tw

(3.19)

where Tw is the thickness of the shear wall


The proposed expression was compared this expression with expression proposed by
Goel and Chopra (1997) and observed higher values of time period obtained by the
former A brief summary of fundamental time period expressions proposed by other
researchers have been presented in Table 3.2.

Table 3.2 Summary of time period expressions proposed by other researchers

Author

Place

Kobayashi et al. (1987)

Mexico city

20

0.105N

Midorikawa (1990)

Santiago de chile

107

0.049N

Midorikawa (1990)

Villa Del Mar (Chile)

21

0.049N

Lagomarsino (1993)

182RC + SW

0.051N

Kobayashi et al. (1996)

Granda (Spain)

21

0.050N

Enomoto et al. (1999)

Almeria (Spain)

34

0.089N +0.032

Espinoza (1999)

Barcelona (Spain)

25

0.060N

Enomoto et al. (2000)

Caracas (Venezuala)

57

0.049N

Sanchez et al. 2002

Adra (Spain)

39

0.049N

Navarro et al. 2002

Granada (Spain)

89

0.049N

Messele and Tadese2002

Addis Abbabba (Ethopia)

28

0.057N,0.018H

Satake et al. 2003

Japan

205RC + SW

0.015H

Dunand et al. 2002

Grenoble (France)

26

0.015H

Oliveira 2004

Lisbon (Portugal)

193

0.042N

Navarro and Oliveira 2004

Lisbon (Portugal)

37

0.045N

Gallipoli et al. 2009

Potenza Senigallia (italy)

65

0.016H

123

3.2.1 Brief discussion on literature review


The review of the past literature with respect to fundamental time period of buildings
showed that the fundamental time period has been considered as a critical parameter in
seismic design of buildings and in seismic vulnerability assessment. As discussed in
literature review section, many researchers have conducted the experimental study on
realistic buildings. Based on regression analyses conducted on experimental results,
simple equations were proposed to estimate fundamental time period in terms of building
height. This is because the building height is the major parameter on which the
fundamental time period depends, and these relations simplify the seismic design
process as explained earlier. However, these expressions were based on buildings with
certain characteristics located in a certain region. Therefore, these equations are not
universally applicable. Moreover, these equations ignore the aspect of structural
irregularity and cracking. Furthermore, many researchers have determined the
fundamental time period using eigen - value analysis (Kose 2008; Ricci et al. 2011a etc).
However, eigen value analysis is based on Rayleighs assumptions and need to be
tested for irregular buildings. Therefore, there is a necessity of a new formula to estimate
the fundamental time period of irregular buildings (including cracking effects). These
formulae should be incorporated in the seismic design process to estimate realistic
seismic demands. In the present study an effort has been made to address these
aspects.

3.3 Modified equations for estimation of fundamental time period


As discussed earlier, the estimation of the fundamental time period is a critical step in
seismic design process. However, code approach mainly depends on the experimental
data set of certain buildings located in a certain region, and ignores the aspect of
structural irregularity and cracking. Moreover, on studying these building configurations it
was observed that these buildings do not represent all forms of irregularities. Therefore,
this section aims to propose modified period height relationships overcoming these
deficiencies. To estimate the modified period height relationship for irregular buildings,
two different approaches have been adopted as:
(a) Eigen - value analysis (b) Inelastic dynamic analysis

3.3.1 Eigen-value analysis (EV) - Eigen - value analysis of a structure is an elastic


analysis which mainly depends upon properties of the structure. The equation to
estimate fundamental period of vibration can be expressed as

[K 2M ] 0

124

(3.20)

and time period can be obtained from equation (3.20) as

2
(3.21)

where K is the stiffness matrix, and M is the mass matrix of the structure
The eigen-value analysis mainly depends upon mass and stiffness of the structure.
Equations 3.20 and 3.21 are relatively simple and are based on Rayleighs method.. To
determine the effectiveness of eigen-value analysis, it is worthwhile to determine the
period - height relationship of the irregular building models adopted in the present study.
To achieve this purpose eigen-value analysis has been conducted on the irregular
building models (described in previous chapter), and the mean period height
relationship for these building models have been plotted considering gross stiffness
criteria as presented in Figure 3.3. Figure 3.3 shows that the code approach failed to
differentiate between the irregular building models and yielded similar value of
fundamental time period for all the building models irrespective of type, magnitude and
location of irregularity. However, eigen - value analysis clearly differentiated between
different types of irregularity. Eigen-value analysis results showed that the fundamental
time period increased with building height for all irregular buildings. On comparison with
regular building model, a greater percentage of increase was observed in case of
stiffness, strength and setback irregular building models (6.12%, 7.25%, 9.05%) as
compared to building models with mass and plan irregularity (1.34%, 3.09%) [Figure 3.4].
Furthermore, consideration of cracking aspect increased the fundamental time period
(2.13%, 3.24%, 7.05%, 7.93%, 10.12% for mass, plan, stiffness, strength and setback
structures respectively) [Figure 3.5 and Figure 3.6].
Among, the cracking approaches considered, Paulay and Priestley (1992) approach
yielded higher period due to higher stiffness reduction factors. The mean difference
between both the cracking approaches (Paulay and Priestley 1992; Priestley 2003) has
been observed as1.33%, 2.33%, ,3.012%, 3.212% and 3.632 % for mass, plan, stiffness,
strength and setback buildings respectively (Figure 3.7). As per previous research works
(Masi and Vona 2010), the increase in fundamental time period increases the seismic
demand. Therefore, a special care should be taken in case of tall structures with
irregularities, and

appropriate design measures adopted should ensure safety and

serviceability of these structures. The period-height relationships for different irregular


building models for gross stiffness and cracking stiffness evaluated using eigen value
analysis has been presented in Figures 3.8 to 3.9 and in Table 3.3.

125

108

H (m)

90

C
M

72

S
ST

54

SE
P

36
18
0.6

1.25

1.9

2.55

3.2

T (Sec)
Figure 3.3 Period - height relationship for different irregular models based on
gross stiffness using eigen - value (EV) analysis

Mean Percentage ( %)

10

7.5

2.5

0
1

Irregularity Category
Figure 3.4 Mean percentage difference (on comparison with regular building
model) in fundamental time period of different irregular building models
(Irregularity category 1, 2,3,4,5 denotes building models with mass, plan, stiffness,
strength and setback irregularity)

126

108

90

H (m)

C
MC

72

PC
SC

54

STC
SEC

36

18
0.6

1.95

3.3

T (Sec)
Figure 3.5 Period - height relationship for different irregular models based on
cracked stiffness (Priestley 2003 approach)

Mean Percentage ( %)

11

8.5

3.5

1
1

Irregularity Category
Figure 3.6 Mean percentage difference in fundamental time period (on comparison
with regular building model) for different irregular building models considering
cracking effect

127

Mean Percentage ( %)

5.5
4.4

M
3.3

P
S

2.2

ST
SE

1.1
0
1

10

11

Building Category
Figure 3.7 Mean percentage difference between periods obtained using different
cracking approaches (Paulay and Priestley 1992 and Priestley 2003) [Building
category

to

11

represent

buildings

with

storey

heights

as

6,9,12,15,18,21,24,27,30,33 and 36 respectively)


Table 3.3 Detailed values of parameters h and h for different irregular building
models based on eigen - value analysis

Model
M200
M400
M600
M800
M1000
S25
S50
S75
S100
ST25
ST50
ST75
ST100
SEC
SEB
SEA
P1
P2
P3

Gross stiffness
h
h
0.0646
0.0635
0.0645
0.0663
0.0650
0.0691
0.0715
0.0741
0.0722
0.0740
0.0763
0.0786
0.0744
0.0786
0.0743
0.0739
0.0637
0.0624
0.0629

0.8131
0.8173
0.8145
0.8092
0.8144
0.8053
0.8039
0.8003
0.8122
0.7946
0.7971
0.7946
0.8112
0.7932
0.8155
0.8276
0.8132
0.8286
0.819

128

Priestley 2003
h
h
0.0763
0.0736
0.0721
0.0731
0.0772
0.0717
0.0776
0.0768
0.0743
0.0753
0.0755
0.0764
0.0781
0.0713
0.0701
0.0683
0.0650
0.0623
0.0654

0.7721
0.7861
0.7911
0.7902
0.7767
0.8132
0.7986
0.8062
0.8208
0.8120
0.8201
0.8209
0.8186
0.8147
0.8292
0.8721
0.8216
0.8357
0.8140

This page is intentionally left blank

129

108

108

T = 0.06478H

90

0.8137

72
54

T = 0.0630H0

54
36

36

18

18
0.6

1.1

1.6

2.1

0.7

2.6

1.2

1.7

2.2

2.7

T (Sec)

T (Sec)

(a)

(b)

108

108

90

T = 0.0717H

90

0.805

72

H (m)

H (m)

.8201

72

H (m)

H (m)

90

54
36

T = 0.0758H

0.799

72
54
36

18

18
0.6

1.2

1.8

2.4

0.6

1.225

T (Sec)

1.85

2.475

3.1

T (Sec)

(c)

(d)
108

T = 0.0756H

H (m)

90

0.812

72
54
36
18
0.6

1.25

1.9

2.55

3.2

T (Sec)

(e)

Figure 3.8 Period - height relationship for different building models using eigen
value analysis considering gross stiffness for: (a) Mass irregularity, (b) Plan
irregularity, (c) Strength irregularity, (d) Setback irregularity, (e) Stiffness
irregularity

130

This page is intentionally left blank

131

108

108

T = 0.0744H

0.783

T = 0.0642H

90

72

H (m)

H (m)

90

54

0.823

72
54
36

36

18

18
0.6

1.1

1.6

2.1

0.7

2.6

1.2

1.7

2.7

T (Sec)

T (Sec)

(a)

(b)

108

108

90

T = 0.0751H

90

0.809

72

H (m)

H (m)

2.2

54

T = 0.0763H

0.817

72
54
36

36

18

18
0.6

1.62

0.6

2.64

1.95

3.3

T (Sec)

T (Sec)

(c)

(d)
108

H (m)

90

T = 0.0699H

0.838

72
54
36
18
0.6

1.3

2.7

3.4

T (Sec)

(e)

Figure 3.9 Period - height relationship for different building models using eigen value analysis considering cracked stiffness (Priestley 2003 consideration): (a)
Mass irregularity, (b) Plan irregularity, (c) Strength irregularity, (d) Setback
irregularity, (e) Stiffness irregularity

132

This page is intentionally left blank

133

3.3.2 Inelastic dynamic analysis


In the previous sections, the estimation of fundamental time period using eigen - value
analysis has been discussed. The eigen - value analysis is preferred by most of the
seismic design codes and previous research works pertaining to the fundamental time
period (Kose 2008; Ricci et al. 2011a) due to its simplicity. However, the eigen - value
analysis mainly depends on building properties and is based on Rayleighs analysis
which has certain limitations as described in the previous sections.
The fundamental time period is an important property and is affected by ground
motion characteristics (Guler et al. 2008; Masi and Vona 2010). Furthermore, the
building period increases due to stiffness degradation occurring during seismic excitation
as observed by previous research works (Calvi et al.2006). Therefore, a realistic
approach involving the ground motion aspect needs to be adopted to determine the
fundamental time period. In the present study, an effort has been made to address these
shortcomings. To achieve this purpose, non linear dynamic analysis has been adopted
to determine the realistic estimate of the period height relationships for different
irregular building models (described in Chapter 2). Moreover, incorporation of aspect of
ground motion results in close agreement between the periods obtained using analytical
and experimental approach (Gallipoli et al. 2009; Masi and Vona 2010). The time history
records as described in previous Chapter have been adopted to conduct inelastic
dynamic analysis. Based on best fit obtained, the generalized equations to estimate the
fundamental time period has been presented in Figure 3.10 and 3.11. From these
figures, it could be observed that pattern of variation of fundamental time period is similar
to pattern obtained from eigen value analysis. However, the percentage variation of
fundamental time period is different, with inelastic dynamic analysis yielding lower
periods as compared to eigen value analysis (0.587%, 0.812%, 1.54%, 1.983%, 2.38
% for mass, plan, stiffness, strength and setback irregular structures respectively). This
shows the over - conservativeness of eigen value

analysis. The period - height

relationships for different irregular building models with detailed equations using inelastic
dynamic analysis have been presented in Figure 3.12 to 3.13 and Table 3.4. These
equations can be readily used to compute fundamental time periods without performing
the complex inelastic dynamic analysis. Furthermore, these equations can be used to
estimate the seismic response and to determine the seismic vulnerability of different
classes of buildings as discussed in later sections of this Chapter.

134

108
90

H (m)

72

S
ST

54

SE
P

36
18
0.6

1.2

1.8

2.4

T (Sec)

Figure 3.10 Period - height relationship for different irregular models based on
gross stiffness using inelastic dynamic analysis

108

90

H (m)

MC
72

PC
SC

54

STC
SEC

36

18
0.6

1.25

1.9

2.55

3.2

T (Sec)

Figure 3.11 Period - height relationship for different irregular models based on
cracked stiffness (Priestley 2003 approach) using inelastic dynamic analysis

135

108

108

T = 0.06045H

0.815

T = 0.05880H

90

72
54

54

36

36

18

18
0.6

1.5

2.4

0.6

1.55

T (Sec)

2.5

T (Sec)

(a)

(b)
108

108
90

T = 0.0729H

90

0.798

72

H (m)

H (m)

0.835

72

H (m)

H (m)

90

54

T = 0.07160H

0.805

72
54
36

36
18

18
0.6

1.62

2.64

0.6

T (Sec)

1.75

2.9

T (Sec)

(c)

(d)
108

H (m)

90

T = 0.0747H

0.808

72
54
36
18
0.6

1.7

2.8

T (Sec)

(e)

Figure 3.12 Period - height relationship for different building models using IDA
analysis considering for: (a) Mass irregularity, (b) Plan irregularity, (c) Stiffness
irregularity, (d) Strength irregularity, (e) Setback irregularity (Gross stiffness
consideration)

136

This page is intentionally left blank

137

108

108
T = 0.0637H

.820

T = 0.0640H0

90

72

H (m)

H (m)

90

0.819

54
36

72
54
36

18

18

0.5

1.5

2.5

0.6

1.65

T (Sec)

T (Sec)

(a)

(b)

108

108
T = 0.06808H

0.834

90

72

H (m)

H (m)

90

2.7

54
36

0.829

72
54
36

18
0.62

T = 0.0709H

18
1.17

1.72

2.27

2.82

0.65

1.825

T (Sec)

T (Sec)

(c)

(d)
108

T = 0.0689H0.850

H (m)

90
72
54
36
18
0.6

1.3

2.7

3.4

T (Sec)

(e)

Figure 3.13 Period - height relationship for different building models using IDA
analysis considering cracked stiffness (Priestley 2003 consideration) for (a) Mass
irregularity, (b) Plan irregularity, (c) Stiffness irregularity, (d) Strength irregularity
(e) Setback irregularity

138

This page is intentionally left blank

139

Table 3.4 Detailed values of parameters h and h for different irregular building
models based on inelastic dynamic analysis

Gross stiffness

Priestley 2003

Model

M200

0.0623

0.8173

0.0654

0.8102

M400

0.0645

0.8119

0.0637

0.8200

M600

0.0663

0.8093

0.06122

0.8221

M800

0.0641

0.8163

0.06287

0.8268

M1000

0.0632

0.8203

0.06571

0.8179

S25

0.0723

0.7892

0.06211

0.8432

S50

0.0751

0.7865

0.06921

0.8250

S75

0.0731

0.7991

0.07120

0.8318

S100

0.0714

0.8094

0.06983

0.8392

ST25

0.0721

0.7963

0.07123

0.8189

ST50

0.0713

0.8086

0.06923

0.8354

ST75

0.0722

0.8117

0.07112

0.8323

ST100

0.0708

0.8182

0.07217

0.8322

SEA

0.0781

0.7891

0.06612

0.8551

SEB

0.0729

0.8150

0.06928

0.8506

SEC

0.0731

0.8221

0.07139

0.8508

P1

0.0630

0.8125

0.06241

0.8232

P2

0.0576

0.8401

0.06571

0.8131

P3

0.0558

0.8531

0.06413

0.8239

140

3.3.3 Alternate method of representation of fundamental time period


The previous sections deals with results of eigen - value analysis and inelastic dynamic
analysis to estimate the fundamental time period. In the proposed expressions, the time
period was expressed as function of building height for reasons as discussed earlier.
This method of representation has inherent disadvantages as two buildings with similar
height would result in a similar time period which in turn would result in a similar seismic
response. In the present section, the above shortcoming has been addressed by
proposing a modified period - height relationships for irregular buildings of different
height categories. However, it is quite difficult to represent large variations in magnitude
and location of irregularity using constants h and h as it resulted in large number of
equations as evident from Table 3.3 and Table 3.4. Therefore, it is convenient to
represent the fundamental time period in terms of a index which is based on dynamic
response. To achieve this purpose, the building models with irregular building models as
described in previous chapter have been analyzed using inelastic dynamic analysis and
a correction factor for time period equation for the building models in terms of the author
proposed irregularity index has been expressed as
T 11 (0.075H 0.75 )

(3.22)

where 11 is the correction factor proposed by the author as


11 2.334 105 c5 1.610 104 c4 7.230 103 c3 0.943 102 c2 0.5512c 1.553 (3.23)

It is worthwhile to note that in Figure 3.14, the maximum value (of the fundamental
time period) is normalized to unity for effective representation of comparison between the
fundamental time period evaluated by proposed method and dynamic analysis. The
range of period values for building models considered in analytical study is presented in
the appendix section. The evaluated values of fundamental time period by both the
methods were found to be in close agreement with a correlation coefficient of 0.9729.

3.4 Generalized variation of fundamental time period


This section summarizes the main conclusions of previous section which aimed at
determined the fundamental time period variation incorporating the effects of structural
irregularity and cracking. The variation of mean time period with different building
properties and structural irregularities have been plotted as shown from Figure 3.15 to
3.18 from which following generalized conclusions have been be drawn as
(a) Fundamental time period increased with height of the structure (Figure 3.15).
(b) Fundamental time period is insensitive to variation in the magnitude of mass
irregularity. However, stiffness, strength and setback irregularity showed a significant

141

effect on fundamental time period. The fundamental time period increased with increase
in the magnitude of stiffness, strength irregularity (Figure 3.15).
(c) The fundamental time period increased marginally with increase in the position of
mass irregularity i.e the building frames with mass irregularity at the top floors will have
the greater time period but a reverse trend was observed in case of stiffness and
strength irregularity (Figure 3.16). Fundamental time period increased for setbacks
originating from bottom storeys. This is due to greater reduction of floor area which in
turn reduces mass, stiffness and strength which results in variation of fundamental time
period (Figure 3.17a).
(d) Fundamental time period increased with increase in magnitude of plan irregularity
(Figure 3.17b). Fundamental time period increased with consideration of cracking
phenomenon for all types of irregular buildings (Figure 3.15 to 3.18).

3.5 Range of applicability of the proposed relations


The author proposed relations are developed based on the assumptions as follows (a)
Hard soil, (b) SMRF frames, (c) Building height from 6 - 36 storeys, (d) Irregularity index
ranging between 0.40 1, (e) For RC building frames with time period within range
proposed by Goel and Chopra (1997), (f) Tb > Tc where Tb is the building time period and

Tc is the characteristic period (g) Mass ratios from 0 % - 1000%, (h) Stiffness and
strength ratios from 0% - 100%, (i) Bay width ranging from 4m - 6m as prescribed in
Indian and European codes of practice.

T (Proposed Equation)

R2 = 0.9729

0.825

0.65

0.475

0.3
0.3

0.475

0.65

0.825

T (Dynamic Analysis)

Figure 3.14 Comparison between proposed equation and dynamic analysis for
irregular buildings

142

108

H (m)

90

M1
M2
S1
S2
ST1

72
54

ST2

36
18
0.6

1.4025

2.205

3.0075

T (Sec)
Figure 3.15 Effect of magnitude of mass, stiffness and strength irregularity on
fundamental time period

108

H (m)

90

M1TS
M1MS
M1BS
S1BS
S1MS
S1TS
ST1BS
ST1MS
ST1TS

72
54
36
18
0.6

1.3

2.7

3.4

T (Sec)
Figure 3.16 Effect of location of mass, stiffness and strength irregularity on
fundamental time period

143

108

H (m)

90

SEC
SECC

72

SEB
SEBC

54

SEA
SEAC

36
18
0.75

2.175

3.6

T (Sec)
(a)

108

H (m)

90

P1
P1C

72

P2
P2C

54

P3
P3C

36
18
0.65

1.3

1.95

2.6

3.25

T (Sec)
(b)
Figure 3.17 (a) Effect of setback irregularity on fundamental time period with and
without cracking effects, (b) Effect of plan irregularity with and without cracking
effects on fundamental time period

144

108
M1C

90

H (m)

M2C
72

S1C
S2C

54

ST1C
36

ST2C

18
0.7

1.35

2.65

3.3

T (Sec)
Figure 3.18 Effect of cracking on fundamental time period

3.6 Application of proposed period height relations in seismic


design methods
The derived period height relationships can be effectively used in the seismic design
methodologies. The seismic design methodologies can be broadly categorized into two
main basic categories namely; (a) Force based design (FBD) (b) Displacement based
design (DBD). These two methods vary in their development and application. The force
based design method is a traditional method in which design forces are obtained by
multiplying the yield forces with force reduction factor. Although FBD yields conservative
estimates of seismic demands, it is still preferred due to its simplicity despite having
certain drawbacks as discussed below
(a) Interdependency of strength and stiffness was neglected.
(b) The time period evaluated was based on Rayleighs assumptions.
(c) The expressions to estimate the ductility capacity and force reduction factors
proposed by EC 8:2004 seismic design code is based on equal displacement rule which
is unrealistic and over - conservative. This aspect has been discussed in detail in the
next Chapter. Due to these shortcomings the FBD method has not been adopted to
evaluate the applicability of proposed period equations in design methodologies.

3.6.1 Brief description of spectrum based DBD (Displacement based


design)
DBD method has been widely used by researchers like Bommer and Pinho (2006),
Akkar and Bommar (2007; Priestley et al. (2007). The use of displacement spectrum is

145

more advantageous as compared to acceleration spectrum as it requires less calculation


(Priestley 2003), and generation of displacement spectrum can be initiated with
knowledge of period, Therefore, accurate estimation of the fundamental time period is
important (in this method), as it determines the seismic response based on which
sections of structural members are decided. The proposed expressions (to estimate
fundamental time period) have been applied to the spectrum based DBD, in which the
time acceleration spectrum was converted to the equivalent displacement spectrum by
multiplying it with factor r11 (as per Priestley et al. 2007) as
r11

T2
4 2

(3.24)

The parameter r11 is evaluated using the approximate value of the fundamental time
period obtained from the author proposed equations.
The converted spectrum (Figure 3.19) has been applied to the irregular building
models (as discussed in Chapter 2) to obtain the section dimensions of the members.
The, application of the proposed period - height relationships may improve the accuracy
in estimation of seismic demands of irregular structures. To demonstrate the
effectiveness of the proposed approach, an example of building models with different
forms of irregularity have been considered. The inelastic displacement spectrum for the
building models has been generated considering ductility factor equal to unity (Figure
3.19).
0.1

0.075

SE

D (m)

ST
S

0.05

P
M
C

0.025

0
0

2.5

7.5

10

T (Sec)
Figure 3.19 Displacement spectrum generated for use in spectrum based DBD

146

3.6.2 Application of proposed fundamental time period equations in the


design methodologies to estimate the seismic response
The first aim of this section is to study the effect of the method of determination of
fundamental time period on the seismic response. The IDR is the parameter chosen, as
it is the most critical parameter to determine safety and stability of the structure. This is
evident from the provisions of most of the seismic design codes (EC 8:2004; IS
1893:2002) which aim at limiting IDR. To achieve this purpose, the IDR for all the
irregular building models have been evaluated using the code approach, eigen - value
analysis and inelastic dynamic analysis considering gross stiffness and cracked stiffness
(EVGS, EVCS, IDGS1, IDCS1, IDGS2, IDCS2 denote method category from 1 to 6) and
the mean percentage variation due to irregularity and cracking effects have been studied
(in comparison to regular building model). The pattern of variation of IDR values for all
irregular models have been shown in Figure 3.20 to 3.24. From these figures it could be
clearly observed that the IDR value increased with storey height for all types of irregular
building models. However, the mean IDR values were insensitive to variation in the
magnitude of mass irregularity with marginal increase of 2.54 % from 6 to 36 storey
frames (Figure 3.25). The plan irregular structures exhibited greater sensitiveness to the
IDR increase with storey height with a mean percentage increase of 5.23 % ( Figure
3.26). However, the strength, stiffness and setback structures observed maximum
increase in IDR with magnitude along the building height (mean percentage increase of
14.12%, 16.42% and 22.98% for stiffness, strength and setback irregular buildings as
shown in Figures 3.25 to 3.29). This is one of main reasons of failure of these structures
as evident from past earthquakes records (as discussed in Chapter 1). The cracking
phenomenon aggravated the deformation demands and is more critical to building
frames with stiffness, strength and setback irregularity (6.85%, 8.12%, 10.23% increase
in IDR as compared to gross stiffness consideration) as compared to building frames
with mass and plan irregularity (3.3%. 4.56%) as evident from Figures 3.25 to 3.29.
These results are in correlation with the fundamental time period results obtained in
previous sections.
The comparison of methodologies to evaluate fundamental time periods suggests
that the code procedures are unable to differentiate between different irregular buildings
and overestimate the seismic demands as compared to eigen - value analysis and
inelastic dynamic analysis. Eigen - value analysis performed better than the code
procedures and clearly differentiated between different types of irregular buildings but
this procedure has inherent disadvantages as discussed earlier. But inelastic dynamic
analysis is more credible as it includes the aspect of different ground motion parameters

147

(Chopra and Goel 2002; 2004; Athanassiadou 2008; Karavasilis 2008a). Eigenvalue
analysis overestimates the seismic response as compared to inelastic dynamic analysis
(1%, 2.14%, 3.17%, 4.34% and 5.41% for mass, plan, stiffness, strength and setback
irregular buildings respectively). The method of representation of time period has a
marginal influence on the seismic response with second method (IDGS2, IDCS2)
yielding a lower period with a mean difference of 1.14% for irregular buildings considered
in the analytical study.

108
IDGS1
IDGC1
IDGS2
IDCS2
EVGS
EVCS
C

H (m)

90
72

54
36
18
0.08

0.135

0.19

0.245

0.3

Mean IDR (cm)

Figure 3.20 Comparison of proposed equation by different methods for estimating


period (by eigen - value analysis and IDA analysis including cracking effects for
building models with mass irregularity (C denotes the code approach IDGS1,IDGS2
denote inelastic dynamic analysis results and proposed method based on regression
conducted on inelastic analysis)

Figure 3.21 Comparison of proposed equation by different methods for estimating


period (eigen - value analysis and IDA) including cracking effects for building models
with stiffness irregularity

148

108

H(m)

90

IDGS1
IDCS1
IDGS2
IDCS2
EVGS
EVCS
C

72
54
36
18
0.21

0.26

0.31

0.36

0.41

Mean IDR(cm)
Figure 3.22 Comparison of proposed equation by different methods for estimating
period (by eigen - value analysis and Inelastic dynamic analysis) including
cracking effects for strength irregular buildings

108

H (m)

90

IDGS1
IDCS1
IDGS2
IDCS2
EVGS
EVCS
C

72

54

36

18
0.2

0.26

0.32

0.38

0.44

Mean IDR(cm)
Figure 3.23 Comparison of proposed equation by different methods for estimating
period (by eigen - value analysis and IDA) including cracking effects for setback
irregular buildings

149

108

H (m)

90
IDGS1
IDCS1
IDGS2
IDCS2
EVGS
EVCS
C

72

54

36

18
0.12

0.185

0.25

0.315

0.38

Mean IDR(cm)
Figure 3.24 Comparison of proposed equation by different methods for estimating
period (eigen - value analysis and IDA) including cracking effects for plan irregular
buildings

Mean Percentage ( %)

3.75

M200
M400
2.5

M600
M800
M1000

1.25

0
1

Method category
Figure 3.25 Percentage increase (with reference to regular building model) of IDR
for building models with mass irregularity by different methods adopted (M 200
M1000 represents mass irregularity with a magnitude ranging from 200 % to 1000
%)

150

Mean Percentage ( %)

10

7.5

P1
5

P2
P3

2.5

0
1

Method category
Figure 3.26 Mean Percentage increase of IDR (with reference to regular building
model) for plan irregular building models by different methods adopted

Mean Percentage ( %)

34

25.5

S25
S50
17

S75
S100

8.5

0
1

Method category
Figure 3.27 Mean percentage increase of IDR (with reference to regular building
model) for building models with stiffness irregularity by different methods adopted
(S 25 S100 represents stiffness irregularity with a magnitude ranging from 25 %
to 100 %)

151

Mean Percentage ( %)

44

33
ST25
ST50

22

ST75
ST100

11

0
1

Method category
Figure 3.28 Mean percentage increase of IDR (with reference to regular building
model) for building models with strength irregularity by different methods adopted
(ST 25 ST100 represents strength irregularity with a magnitude ranging from
25% to 100%)

Mean Percentage ( %)

48

36

SEC
SEB

24

SEA
12

0
1

Method category
Figure 3.29 Mean percentage increase of IDR (with reference to regular building
models) for building models with setback irregularity by different methods
adopted

152

3.7 Seismic Vulnerability assessment


The damage caused by the recent earthquakes in both the social and economic aspect
has created an increased awareness amongst populations, industry, commerce and
governments Both private and government organizations especially insurance agencies
are investing a lot of time, money and energy regarding this research aspect. Such
efforts have resulted in development of softwares like HAZUS (FEMA, 1999; Whitman et
al. 1997) which provided a detailed methodology to determine the potential damage
caused by an earthquake. However, development of such programmes and methodology
requires large set of experimental data of buildings subjected to the seismic excitation.
These data were very difficult to obtain due to variety of reasons as discussed earlier.
The review of previous research studies showed that the seismic demands had often
been expressed in terms of PGA, PGV or ground motion intensity. However, new
developments in the area of displacement-based design (Moehle 1992; Calvi and
Pavese 1995; Kowalsky et al. 1995) owing to logical reasoning and its importance has
made the deformation demands as major parameters in seismic capacity assessment
(Calvi 1999).
The demand spectrum (a curve of spectral acceleration versus spectral
displacement) in conjunction with pushover curve has often been used to represent the
structural capacity of the system (Freeman et al. 1975; Freeman 1998; Fajfar 1999). One
of the important parts of this procedure involving the derivation of these demand and
capacity curves and the identification of their intersection (known as the performance
point, SD), that is especially demanding in terms of time, computing power, and required
input data. The method of estimation of seismic capacity should be more sensitive to
parameters like building properties and ground motion rather than the method of
calculation. The present study aims to determine the seismic vulnerability assessment
using capacity spectrum concept and using simplified method proposed by Glaister and
Pinho (2003).

3.7.1 Review of methods for seismic vulnerability assessment


The features of a model to estimate the loss due to seismic excitation depends on its
purpose of estimation for emergency planning in which the earthquake is considered as
a single event. The repetition of an earthquake is generally used to estimate the annual
loss due to earthquake considering all feasible earthquake scenarios have been
considered and the resultant losses were ranked according to their probability of
occurrence (McGuire, 2001). A complete model to determine the earthquake loss must
include all the elements of hazard including aspects like fault rupture, tsunami,
liquefaction and landslides.

153

However, the present research work is focused on the seismic vulnerability assessment
of the building inventory to the effects of strong ground shaking which is the main cause
of earthquake damage resulting in a direct economic loss. Macroseismic intensity scales,
such as the Modified Mercalli (MM), MSK or MCS are to represent the ground shaking.
Intensity for seismic vulnerability assessment is in direct relation to damage in different
classes of buildings (Musson 2000). However, its application is limited because
prediction of intensity values for future earthquakes, especially considering soil effects
requires these discrete index values to be assumed as continuous variables. This can be
achieved by the use of instrumental parameters of the ground motion such as peak
ground acceleration (PGA) which is widely used as a base for loss estimation studies
(King et al. 1997). However, PGA has a very poor correlation with structural damage
during earthquakes. Peak ground velocity (PGV), which is related to the energy in the
ground motion is a better indicator of the damage potential than PGA and has been
widely used as the basis for some more recent earthquake loss functions (Miyakoshi et
al, 1997; Yamazaki and Murm 2000). Nevertheless, single parameters such as PGA,
PGV and macro-seismic intensity do not entirely represent the frequency content of the
ground motion. These parameters ignore the influence of the natural and effective period
of vibration of buildings in determining loading level (to be experienced) during the
seismic excitation. This parameter can be effectively represented by complete
descriptions of the .ground motion such as response spectra, because of this the
response spectra has been used in previous research works for loss estimation
(Scawthorn et al.1981; Shinozuka et al.1997).
In recent years the seismic vulnerability assessments are more focused towards the
deformation parameters. The review of previous literature reveals

the use of

displacement response spectrum (or acceleration-displacement response spectrum) to


represent the damage capacity of the ground motion (e.g. Calvi, 1999; Faccioli et al,
1999). In methodology proposed by previous researchers [Kircher et al. 1997; Whitman
et al.1997; FEMA, 1999), the seismic performance has been obtained by the
performance point as the intersection of capacity (spectrum showing lateral displacement
of the structure) and demand spectrum (spectral acceleration vs spectral displacement).
The fragility curves were constructed on these basis to read the expected probability in
each damage state (I, II, III IV). This method considers period-height relationships and
inelastic deformation (ductility) in the estimation of damage states. In addition, a
consideration to construction practices and effect of retrofitting has also been considered
in this method which makes it time and cost consuming. This method is inapplicable to
cases where the available data on the soil conditions and building stock is limited in
addition to constraints of time and cost that necessitate use of a simplified approach.

154

However, the adopted simplified approach should retain the elements of the effective
period of vibration of structures and structural displacements as the major variables in
determination of seismic capacity of structures. The Hazus methodology is very popular
and has been widely used by previous and current researchers for estimation of seismic
capacity (Bommer et al. 2002; Spence et al. 2002; Ayala et al. 2005; Calvi et al. 2006;
Hueste and Bai 2006; Polese et al. 2008; Lu and Shi 2012; Baltzopoulou 2012). Calvi
(1999) made such an attempt and derived seismic capacity of beam and column sway
frames based on a displacement-based method. Calvi (1999) extends this method for
seismic vulnerability assessment for different classes of buildings. To deal with variation
of the building stock, Calvi (1999) established upper and lower limits of the material and
geometric variables of the building stock. The corresponding time period were
determined using these maximum and minimum limits of possible structural capacities.
These capacity limits define the areas intersected by a given response spectrum chosen
to represent the seismic demand.. The probability of attaining the associated limit state is
then determined by integrating the volume of a joint probability density function (JPDF)
between capacity and period above and below the spectral line and within the defined
ranges. The JPDF was assumed as constant in the previous research work, most likely
for achieving simplicity. However, in this research work ;(a) The variation of ductility with
effective height was neglected, (b) The ductility - height relationship was given utmost
importance neglecting the relationship between displacement capacity, material
properties, building height and period.

3.7.2 Application of proposed fundamental time period equations in seismic


vulnerability assessment
This section aims at discussing the applicability of proposed fundamental time period
equations in seismic vulnerability assessment. To achieve this purpose the Glaister and
Pinho (2003) method has been adopted. The method proposed by Glaister and Pinho
(2003) is an extension of method proposed by Calvi (1999) .This method has been
conceptually described in Figure 3.30 (a). In this method, the capacity curves of the
structure for different limit states were plotted using: (a) Analytical relationships between
displacement capacity and height, b) Empirical relationships between height and elastic
period. The seismic performance is obtained by intersections between any given
displacement spectrum (scaled by a damping factor with respect to the ductility demand),
and its corresponding capacity curves which indicate the periods (TLsi) that mark the
boundaries of the various limit states (Mi). The time period at intersection of demand and
capacity curve indicate the capacity period at which capacity (defined by a given limit

155

state) exactly matches the demand. Therefore, all periods to the left of this limit
represent buildings with capacity failing below demand, and vice versa for periods to
right of this period. These building periods are then transformed into their equivalent
heights (HLsi), using the previously mentioned period-height relationships, and plotted as
a curve of cumulative distribution function (CDF) versus height. This determines the
proportions of the buildings failing in each limit state (Figure 3.30 b). It is worthwhile to
note that the inherent dispersion in the vulnerability of any individual class of building is
neglected. Therefore, this approach is an alternative to the capacity spectrum method to
estimate directly the distribution of seismic capacity across a particular class of buildings
at a specific location and for any given earthquake ground motion. The detailed capacity
height equations and their derivation are available in Glaister and Pinho (2003). The
Glaister and Pinho (method has been used in previous research works as described
below (Crowley et al.2004;Crowley et al. 2005; Su and Shi 2012). To apply this
methodology, the demand and capacity curves for the building models were evaluated.

(a)

(b)

Figure. 3.30 Seismic Vulnerability assessment using deformation based approach


(Glaister and Pinho 2003) where HLsi is the equivalent building height, TLsi is the
building period corresponding to demand, PLsi is probability of failure
corresponding to demand, LS1, LS2 and LS3 are the corresponding perormance
levels

156

The demand curve has been idealized by the displacement spectrum generated from
ensemble of 27 records. The time Acceleration spectra was converted into
displacement spectrum using the following relation

T2
ag
4 2

(3.24)

The capacity curve has been generated as per EC 8:2004 criteria. The building
models adopted in the present study were subjected to the seismic vulnerability
assessment by above two methods (Hazus (FEMA 1999); Glaister and Pinho 2003). The
brief description of Glaister and Pinho (2003) methodology is presented in Figure 3.31
(using an example of a 6 storey irregular building model). Likewise, mean capacity
periods for the building models were evaluated by the author as shown in Figure 3.32.
Based on this method, the seismic vulnerability assessment has been presented (in
terms of probability of collapse) for irregular building models is presented from Figure
3.33 to 3.38. From these figures it has been observed that
(a) The building models with vertical irregularities (except mass irregularity) exhibited
greater seismic vulnerability as compared to building models with plan irregularity (Figure
3.33 to 3.37)
(b) The probability of collapse obtained considering the gross stiffness has been
observed to be much lower as compared to consideration of cracked section. This is due
to the fact that the consideration of gross stiffness results in increased stiffness which
reduces the deformation demands. Therefore, lesser probability of collapse is obtained.
Among the cracked section considerations, the analytical formulations yield lesser
probability due to assumptions which resulted in unrealistic stiffness reduction factors.
However, Priestley (2003) reduction factors result in realistic estimate of collapse
probability (Figure 3.38).
(b) The Glaister and Pinho (2003) yielded greater probability of collapse as compared to
Hazus methodology (FEMA 1999) for the building models considered in the analytical
study (Figure 3.37). This is due to use of static method of analysis (in Hazus
methodology). Nevertheless, the former method can be a alternate to the latter (which is
tedious and time consuming) in evaluating the probability of collapse. This would result
in simplicity and economy in seismic vulnerability assessment. The code proposed
equations have a distinct disadvantage that they cannot be used in conjunction with
Glaister and Pinho (2003) method to evaluate seismic vulnerability of irregular buildings.
This is because, code equations ignored the aspect of structural irregularity and cracking
(Figure 3.39).

157

0.12

D (m)

CP (T = 0.786 S)

DC
CC

0.06

0
0

T (Sec)
Figure 3.31 Glaister and Pinho (2003) methodology for Seismic vulnerability
assessment of 6 storey irregular building (DC is the demand curve, CC is the
capacity curve, CP is the seismic capacity)

108

H (m)

90
72
54
36
18
0.7

1.3

1.9

2.5

3.1

Mean Capacity Period (Sec)


Figure 3.32 Mean derived capacity periods for building models considered in the
analytical study

158

108
M 200

90

H (m)

M 400
M 600

72

M 800
M 1000

54
36
18
0.12

0.155

0.19

0.225

0.26

Mean Pc (%)
Figure 3.33 Mean probability of collapse for mass irregular building models based
on Glaister and Pinho (2003) methodology

108

H (m)

90

S 25
S 50

72

S 75
S 100

54

36
18
0.08

0.145

0.21

0.275

0.34

Mean Pc (%)
Figure 3.34 Mean probability of collapse for stiffness irregular building models
based on Glaister and Pinho (2003) methodology

159

108
90

ST 25

H (m)

ST 50

72

ST 75
ST 100

54

36
18
0.08

0.24

0.4

Mean Pc (%)
Figure 3.35 Mean probability of collapse for strength irregular building models
based on Glaister and Pinho (2003) methodology

108
90
SEC

H (m)

72

SEB
SEA

54

36
18
0.08

0.185

0.29

0.395

0.5

Mean Pc (%)
Figure 3.36 Mean probability of collapse for setback irregular building models
based on Glaister and Pinho (2003) methodology

160

108
90

H (m)

P1

72

P2
P3

54

36
18
0.08

0.13

0.18

0.23

0.28

Mean Pc (%)
Figure 3.37 Mean probability of collapse for plan irregular building models based
on Glaister and Pinho (2003) methodology

108
90

H (m)

GPG

72

HG
HC

54

GPC

36
18
0.18

0.255

0.33

Mean Pc (%)
Figure 3.38 Mean probability of collapse of buildings using Glaister and Pinho
(2003) methodology [ Method 1) Hazus methodology [Method 2) [GPG and GPC
Glaister and Pinho (2003) method with gross and cracked stiffness, HG,HC
represent Hazus methodology with gross and cracked stiffness)

161

108

H (m)

90
72

GPC2
GPC1

54
36
18
0.21

0.27

0.33

Mean Pc (%)
Figure 3.39 Mean probability of collapse for buildings of buildings using different
cracking approaches (GPC1 represents Priestley 2003 approach and GPC2
represents Paulay and Priestley 1992 approach)

3.8 Summary and main conclusions


This Chapter deals with fundamental time period of buildings and has been divided into
three sections. In the first section, a brief literature review and its shortcomings
pertaining to the aspect of fundamental time period has been discussed. In the second
section, the fundamental time period is evaluated using eigen - value analysis and
inelastic dynamic analysis. These methods have been compared with each other.
Thirdly, simple period - height relationships based on eigen - value analysis and inelastic
dynamic analysis have been proposed. In addition, the fundamental time period is also
expressed as a function of irregularity index. These approaches are compared at end of
this section. The third section deals with applicability of the proposed period - height
relationships in estimation of seismic response parameters and in seismic vulnerability
assessment procedure proposed by Glaister and Pinho (2003). Finally, collapse
probability of the irregular building models evaluated using the proposed time period
equations by Glaister and Pinho (2003) method have been compared with Hazus (FEMA
1999) method are compared. The main conclusions observed from this Chapter are
(a) The code proposed equations were based on experimentally determined periods of
certain set of buildings. Moreover, these equations ignore the aspect of structural
irregularity and cracking which is unrealistic. However structural irregularity and cracking
affect the seismic demands. Therefore, these equations are not universally applicable.
The analysis results obtained by conducting eigen value analysis and inelastic dynamic

162

analysis on irregular building models showed that, (i) The fundamental time period
increased with increase in irregularity magnitude with marginal increase for mass and
plan irregular structures, (ii) The presence of irregularities in bottom storeys aggravated
the time period as compared to its location in middle and top storeys (except for mass
irregular building models for which a reverse trend was observed).
(b) Eigen - value analysis clearly differentiated between period - height relationships of
different building models was observed to be more accurate as compared to the code
equations in estimating the fundamental time period. However, this method has its
inherent disadvantages as it is based on Rayleighs assumptions and effect of ground
motion has been ignored. This limitation is overcome in inelastic dynamic analysis.
Therefore, in the present research work, the modified equations to estimate the
fundamental time periods (incorporating the structural irregularity and cracking have
been proposed) have been proposed in terms of building height. This has been done as
(i) Height is the main parameter on which time period depends, (ii) It simplifies the design
process and prior knowledge of section dimensions is not required. Although, the
proposed period height relationships effectively capture structural irregularity and
cracking but large number of equations are required to capture every aspect of
irregularity and cracking which is inconvenient. Therefore, a different approach has been
adopted and the fundamental time period is expressed in terms of irregularity index in
form of a single equation which makes it simple. The comparison of proposed equations
showed that eigen - value analysis yielded higher periods as compared to inelastic
dynamic results but exhibited much better performance as compared to the code
approach. However, the proposed equations yielded comparable results with dynamic
analysis hence are preferable.
(c) The author proposed equations are very useful and easily applicable in seismic
design methodologies. The observations suggest that, application of code proposed
period equations in these methodologies yielded a conservative estimate of seismic
response as compared to dynamic analysis. In comparison, the author proposed
equations were simple and effectively fitted into the framework of design methodologies,
and yielded accurate estimate of the seismic response.
(d) The proposed equations were observed to be quite useful in seismic vulnerability
assessment method proposed by Glaister and Pinho (2003). From seismic vulnerability
assessment, it was observed that presence of irregularity has a significant impact on the
seismic response. The seismic vulnerability assessment methods by Calvi (1999) and
Hazus (FEMA 1999) method yielded similar results. Since, Hazus methodology is quite
complex and time consuming; Glaister and Pinho (2003) methodology can be
conveniently adopted. Therefore, the author proposed equations in conjunction with

163

Glaister and Pinho (2003) method can be effectively used for seismic vulnerability
assessment which would save time, money and energy.

164

CHAPTER 4
DETERMINATION OF INELASTIC DEFORMATION DEMANDS OF
IRREGULAR BUILDING FRAMES

4.1 Introduction
The presence of structural irregularities has been a common feature in modern buildings
and these are preferred mainly due to functional and aesthetic requirements. The
irregularities can be broadly classified into plan and vertical irregularities. These
irregularities along with their sub-classifications have been discussed in detail in Chapter
1. The presence of irregularity in a building system leads to variation of the seismic
response in the building which depends upon type, magnitude and location of
irregularity.
The seismic design codes have suggested the procedures to estimate deformation
demands but the aspect of irregularity has been ignored in formulating these procedures.
In other words, the deformation demands of irregular and regular structures are
estimated using the same rules. Moreover, these rules are proposed considering SDOF
system and elastic analysis which is unrealistic. Therefore, there is a need to modify the
code rules to estimate deformation demands of irregular structures.
In this Chapter, the variation of inelastic deformation demands with different type,
magnitude and location of irregularity have been studied first. Secondly, a brief review of
procedures prescribed by various seismic codes has been presented. Thirdly, the
inelastic deformation demands for building models with different types of irregularity as
described in Chapter 2 have been determined. Based on non-linear regression analysis
on the seismic response results, simple equations to estimate the seismic response
parameters have been proposed by the author in terms of irregularity index (as proposed
by author in Chapter 2). Finally, applicability of these equations in displacement based
and performance based design has been briefly discussed.

4.2 Brief discussion on code provisions and previous literature works


pertaining to deformation demands
The previous section presents a brief introduction to the aspect of deformation demands.
This section presents the detailed comparison of inelastic seismic response of different
irregular building models. At the end of this section, the relation between elastic and
inelastic seismic response has been evaluated in form a parameter.
The deformation demands are the critical parameters which play a vital role in seismic
design and analysis of buildings and strongly influence the seismic performance of

165

buildings. Therefore, majority of the previous research works are mainly focused on
estimation of these demands. The limits on the deformation demands by some of the
seismic design codes have been presented in Table 4.1 from which it can be clearly
seen that the codes specify the maximum deformation limit in terms of storey height only.
These limits have been fixed based on elastic analysis and the aspect of structural
irregularity and cracking have been clearly ignored. Therefore, these limits are unrealistic
and are unsuitable for design of structures. Moreover, the detailed literature review
pertaining to deformation demands has been presented in Chapter 1 which showed that
(a) Majority of previous research works pertaining to deformation demands are mainly
based on elastic analysis and SDOF systems which is unrealistic. Moreover, code
provisions are formulated based on the abovementioned unrealistic assumptions.
Therefore, there is a necessity to revise code provisions to propose a new rule to
estimate the realistic seismic response.

Table 4.1 Limits on deformation demands by some of the seismic design codes

Code

Storey drift Limit

IS 1893:2002

0.004h

EC 8 2004

0.075h/v

UBC 1997

0.025 h (T< 0.7), 0.020 (T > 0.020)

NBCC 2005

0.01h,0.02h,0.025h

(b) The majority of the previous research works were confined to determination of
deformation demands for low rise and mid rise buildings (0 18) (Athanassiadou et al.
2008; Karavasilis et al. 2008a; Kappos and Stefanidou 2010). However, review of
statistics of different building heights in different countries (Wikepedia) show that the
existing buildings and newly constructed buildings exceed the height of those studied in
the previous research works. The deformation demands of the structures vary with
building height. Therefore, there is a need of estimation of seismic response of tall
buildings and a revised design methodology needs to be proposed.

166

(c) As discussed earlier, that majority of the previous research works have adopted
elastic analysis and SDOF systems with a exception of few researchers who have
adopted inelastic modeling approach. These approaches have adopted simple inelastic
models capturing failures due to strength and stiffness deterioration (Cloughs model;
Bouc and Wen model). These inelastic models were based on SDOF system. This raises
a question of applicability of these models to estimate the realistic seismic response.
Moreover, these works fail to modify the current seismic design procedures and propose
a new procedure to incorporate the aspects of structural irregularity and cracking.
In the present research work, an effort has been made by the author to address these
shortcomings.

4.2.1 Inelastic deformation demands of irregular buildings


As discussed in the previous section, the code procedure generally prescribes the elastic
analysis of the structure which is a conservative approximation. However, in reality the
structures exhibit inelastic behavior. The inelastic behavior can be determined by
modeling the hysteretic behavior incorporating the non - linear force deformation
relationship in structural members of the frame. The brief overview of hysteretic behavior
of models proposed by different authors has been discussed in detail in Chapter 2. In the
present study, the inelastic behavior has been modeled using the Ibarra et al. (2005) due
to its simplicity and efficiency in representing the cyclic behavior. The main aim of this
section is to study the inelastic seismic response of the irregular building structures (as
described in Chapter 2), and to establish a relationship between inelastic and elastic
seismic response. These relationships can be effectively used in preliminary stages of
PBD and DBD as discussed in detail in last section of this Chapter.
(a) Effect of mass irregularities on inelastic deformation demands
In this section, the effect of mass irregularities on height-wise distribution of deformations
demands (inter-storey drift ratio, maximum roof displacement and maximum rotational
demand) has been studied. To determine the effects of irregularity, the seismic
responses of the irregular building models have been compared with the regular building
model (which has been considered as the base case). As discussed in the second
Chapter, the irregular building models have been generated by introducing the different
types of irregularity of varying magnitudes at different locations. In this section, the effect
of mass irregularity has been studied on inelastic deformation demands namely;
(a) Maximum roof displacement (rd)
(b) Maximum inter - storey drift (Ird)

167

(c) Maximum rotational demand (i)


The variation of inelastic deformation demands along the building height has been
presented form Figure 4.1 to 4.4 for different magnitudes of mass irregularity (M1 =
200%, 400%, 600%; M2 = 800%, 1000%). From these figures it can be observed that
deformation demands increased with increase in position of mass irregularity from
bottom of the structure. The deformation demands increased by 1.34%, 1.45%, 1.39%
(rd, Ird, i) for a mass ratio M1 when the irregularity moves from bottom to top storey.
This percentage increased marginally with a mean percentage of 2.05 %, 3.19, 2.64 %
for a mass ratio of M2 (Figure 4.4). Furthermore, following generalized conclusions have
been drawn as
(i) The variation of peak values of deformation demands (rd, Ird, i) followed a similar
pattern with similar percentage of increase which was marginally higher for mass
irregularity present at the top storey. This is a very important observation which leads to
the conclusion that the occurrences of mass irregularity are critical at the top storey. This
is evident from building failure due to heavy roofs at the top storeys during June 27, 1998
AdanaCeyhan (Turkey) earthquake (one of the major reasons of building failures) as
reported by Adaliar and Aydingun 2001. The effect of mass irregularity presence is
depicted clearly in Figure 4.5 which shows failure of a building due to mass irregularity at
top storey during Bhuj earthquake 2001. However, many buildings in reality (Burj Khalifa,
Empire state etc as shown in Chapter 1) have mass irregularities at top storeys due to
functional requirements. But it is preferable to avoid these irregularities to avoid damage
during seismic excitation.
(ii) The lateral deformation demands increased marginally with the increase in magnitude
of mass irregularity. This observation is very useful as the lateral deformation demands
were considered to be critical parameters in design of buildings. Therefore, the effect of
mass irregularity was generally neglected in formulating the seismic design
methodologies. However, the effect of mass irregularities should not be underestimated
as mass irregularity in a storey leads to an increase in large additional moments and
shears in the adjacent storey which results in instability of the structure (Das and Nau
2003). Therefore, presence of mass irregularities should be given a serious
consideration in seismic design of buildings. Moreover, presence of additional mass in a
building portion alone may instigate the torsion in the buildings which is one of the main
reasons of failure of buildings during past earthquakes (Chapter 1).
(iii) The cracking phenomenon aggravated the deformation demands of mass irregular
buildings by a mean percentage of 4.34 %.

168

108
BSM1

90

MSM1

H (m)

72

TST1
BSM2

54

MSM2
TST2

36

18
0.9

1.95

Mean rd (cm)
Figure 4.1 Variation of mean global deformation demand along building height for
mass irregular buildings (BSM,MSM,TSM denotes mass irregularity at bottom one
third, middle one third and top one third storeys, R denotes a regular building
model)

108
BSM1

90

H (m)

MSM1

72

TST1
BSM2

54

MSM2
TST2

36
18
0.08

0.19

0.3

Mean Ird (cm)


Figure 4.2 Variation of mean storey deformation demand in terms of maximum IDR
(Ird) along building height for mass irregular buildings

169

108
BSM1
MSM1
TSM1
BSM2
MSM2
TSM2
R

90

H (m)

72
54
36
18
0.018

0.026

0.034

Mean i (rad)
Figure 4.3 Variation of mean local deformation demand along building height for
mass irregular buildings

Model category

6
5
M1000

M800
M600

M400
M200

2
1
0

Mean percentage ( % )
Figure 4.4 Mean Percentage increase on comparison with regular building model
(Mean of all height categories) of deformation parameters (Mean of rd, Ird, i). with
variation of magnitude and location of mass irregularity (Model category 1 to 3
and 4 to 6 denotes irregularity at bottom, middle and top storeys with and without
cracking effects)

170

Figure 4.5 Failure of building due to heavy mass at top during Bhuj earthquake
(Humar et al. 2001)

(b) Effect of stiffness and strength irregularity on deformation demands


The stiffness and strength variation are very commonly encountered in the modern
buildings due to functional and aesthetic considerations e.g. reduction of beam-column
sizes (for economy) and increase in height of a particular storey (especially to create
basements, car-parking etc.). Likewise, variation of reinforcement and concrete grade
also results in stiffness and strength irregularities. The introduction of these irregularities
in the building results in variation of the seismic response. The variations on inelastic
deformation demands for stiffness irregular buildings (S1 = 25%, 50%; S2 = 75% and
100%) have been plotted from Figure 4.6 to 4.8. The mean percentage increase of
inelastic deformation demands (in comparison to the regular building model) with
building height (6 to 36 storeys) for the case when irregularity is at the bottom storey has
been observed as 13.34%, 15.34%, 14.36% for parameters rd Ird and i respectively.
This percentage increase of deformation demands reduced to 11.23%, 13.34% and
12.75% (rd, Ir and i) for stiffness irregularity in middle storeys, and to 9.13%, 11.85%
and 10.45% (rd, Ird and i) for stiffness irregularities in the top storeys respectively
(Figure 4.9).

171

This page is intentionally left blank

172

Moreover, the deformation demands increased with the magnitude of stiffness


irregularity with a mean percentage of 13.23%, 15.23%, 14.23% and 22.34%, 24.35%,
23.67% for irregularity magnitudes of S1 and S2.
The effect of strength irregularity on the seismic response has been shown in Figure
4.10 to 4.12. From these figures it can be observed that the strength irregular buildings
showed a similar pattern of variation of deformation demands (with magnitude and
location of irregularity) as exhibited by stiffness irregular building models. The mean
deformation demands decreased by 15.12%, 17.23% and 19.23% for location of
irregularities (From bottom to top storey). Also, the mean deformation demands
observed an increase of 21.12%, 24.12% and 22.43% for magnitude of strength variation
from 0% to 100% (Figure 4.13) which is greater as compared to the case of stiffness
irregular buildings. Moreover, following generalized conclusions have been drawn as
(i) The stiffness and strength irregularity followed a reverse trend as compared to mass
irregularity with maximum increase in deformation demands observed when these
irregularities were present in bottom storeys. In addition, a substantial increase of
deformation demands has been observed when these irregularities were present at the
middle storeys. These observations are consistent with building failures during past
earthquakes (Figure 4.14 to Figure 4.19). However, these irregularities are still preferred
in modern buildings due to functional requirements. Therefore, appropriate design
measures should be adopted to ensure safety and serviceability of these structures.
(ii) The previous research works (pertaining to structural irregularity) were confined to
mid rise buildings ignoring the tall structures. However, due to increasing population and
space congestion especially in developing countries, tall structures are being
increasingly preferred. This is evident from recent statistics on building height of recent
buildings (Wikepedia) which shows emergence of tall structures. The analysis results
showed that the presence of stiffness and strength irregularities in case of tall buildings
had a huge impact on inelastic deformation demands with highest percentage of
increase being observed for 36 storeyed structure (21.13 %, 25.45 % and 23.23 % for

rd, Ird, i respectively). The estimated seismic responses show that occurrences of
irregularity in tall structures is more critical. Therefore, a special attention should be paid
to analysis and design of these structures. (iii) Presence of cracking aggravates the
deformation demands by a mean percentage of 6.13 % and 7.86 % in case of stiffness
and strength irregular building models, (iv) Contrary to the case of mass irregularity, the
seismic response parameters showed a large sensitivity to the stiffness and strength
irregularities with location of these irregularities being more critical at the bottom storeys.

173

108

BSS1

90

H (m)

MSS1
TSS1

72

BSS2
54

MSS2
TSS2

36

18

2.4

3.175

3.95

4.725

5.5

Mean rd (cm)

Figure 4.6 Variation of global deformation demands along building height for
stiffness irregular buildings (BSS,MSS,TSS denotes stiffness irregularity at bottom
one third, middle one third and top one third storeys)

H (m)

108
90

BSS1
MSS1

72

TSS1

54

BSS2
MSS2
TSS2

36

18
0.08

0.25

0.42

Mean Ird (cm)

Figure 4.7 Variation of storey deformation demands along building height for
stiffness irregular buildings

174

108

BSS1
MSS1
TSS1
BSS2
MSS2
TSS2
R

90

H (m)

72
54
36
18
0.018

0.031

0.044

Mean i (rad)
Figure 4.8 Variation of local deformation demands along building height for
stiffness irregular buildings

Model category

6
5

S100
S75
S50
S25

4
3
2
1
0

14

21

28

35

Mean Percentage ( % )

Figure 4.9 Mean percentage increase (on comparison with regular building model)
of deformation parameters (Mean of rd, Ird, i) with variation of magnitude of
stiffness irregularity in building models

175

108

H (m)

90

BSST1
MSST1
TSST1
BSST2
MSST2
TSST2
R

72
54
36
18
2.4

3.3

4.2

5.1

Mean rd (cm)
Figure 4.10 Variation of global deformation demands along building height for
strength irregular buildings (BSST,MSST,TSST denotes stiffness irregularity at
bottom one third, middle one third and top one third storeys)

108

BSST1
MSST1
TSST1
BSST2
MSST2
TSST2
R

H (m)

90
72
54
36
18
0.05

0.15

0.25

0.35

0.45

Mean Ird (cm)


Figure 4.11 Variation of storey deformation demands along building height for
strength irregular buildings

176

108

BSST1
MSST1
TSST1
BSST2
MSST2
TSST2
R

H (m)

90
72
54
36
18
0.018

0.032

0.046

Mean i (rad)
Figure 4.12 Variation of local deformation demands along building height for
strength irregular buildings

Model category

6
5
ST100
ST75
ST50
ST25

4
3
2
1
0

18

27

36

45

Mean Percentage ( % )
Figure 4.13 Mean percentage increase (on comparison with regular building
model) of deformation parameters with reference to regular building for strength
irregularity building models

177

This page is intentionally left blank

178

Figure 4.14 Failure of buildings due to strength and stiffness irregularity at bottom
storeys (Taskin et al. 2013)

Figure 4.15 Failure of buildings due to strength and stiffness irregularity at bottom
storeys (Romao et al. 2013)

179

This page is intentionally left blank

180

(a) Collapse of a building apartment in Fengyunan district in Taiwan during Chi Chi
earthquake 1999

(b) Collapse of a high rise apartment in Dali during Chi Chi earthquake 1999
Figure 4.16 Failure of tall buildings due to strength and stiffness irregularity at
bottom storeys during Chi Chi earthquake 1999 (Tsai et al. 2000)

181

This page is intentionally left blank

182

Figure 4.17 Failure of tall buildings with stiffness and strength irregularity during
Bhuj earthquake (Humar et al. 2001)

Figure 4.18 Failure of buildings with stiffness and strength irregularity at middle
storeys during 1995 Kobe (Japan) earthquake, (Elnashai and Di Sarno 2008))

183

This page is intentionally left blank

184

Figure 4.19 Failure of buildings with stiffness and strength irregularity at middle
storeys during Bhuj earthquake (Paul and Dogan, as reported in Kirac et al. 2013)

(c) Effect of setback irregularities


The setback irregularity is one of the most common types of irregularity in the modern
buildings. The functional and aesthetic requirements are the main reasons for preference
of these structures. The setback buildings are quite useful in urban areas where there is
a space constraint and closer proximity of buildings is required. In such areas, these
buildings provide adequate sunlight and ventilation for the bottom storeys. In addition,
these buildings comply with the building byelaw restrictions of Floor area ratio as per
building code of India.
The presence of setbacks in a building results in abrupt reductions of the floor area
which in turn results in change of mass, stiffness and strength along the building height.
This results in variation of dynamic characteristics of these buildings as compared to the
regular buildings. This aspect has been ignored by the seismic design codes in
formulating the seismic design methodologies. This is reflected in poor seismic
performance of setback structures during the past earthquakes (Penelis et al. 1978;
Spyropoulos 1982; Penelis and Kappos 1997). In addition the setbacks in the form of
projected balconies also experienced a severe failure during earthquakes (Chapter 1).
Therefore, there is a need to study the variation of deformation demands due to setback
irregularity.
This section at first aims to estimate the deformation demands at vicinity of setback.
To achieve this purpose, the base building models are used. The deformation demands

185

for these building models have been plotted in Figure 4.20 to 4.22. From these figures it
could be clearly observed that, deformation demands show an abrupt increase (with a
mean percentage of 13.45%) near presence of setbacks which shows the criticality of
this aspect. This observation is consistent with observed building failures during past
earthquakes (Figure 4.23). Figure 4.23 shows failures of a 21 storeyed building at its 12th
storey (due to presence of setback) during Chile 2010 earthquake. Moreover, projections
in form of balconies have also experienced severe failures during past earthquakes
(Chapter 1). Therefore, storeys adjacent to setback should be given a special design
consideration due to avoid failure.
The reported literature review regarding the setback structures (Chapter1) shows that
the seismic performance of the setback structures is quite unclear with some
researchers indicating their adequate seismic performance (Wood 1992; Duan and
Chandler 1995; Mazzolini and Piluso 1996; Romeo et al. 2004; Tena Colunga
2005;Athanassiadou 2008; Kappos and Stefanidou 2010) and other suggesting the
opposite view (Humar and Wright 1977; Pekau and Green 1997; Shahrooz and Moehle
1990; Chen et al. 2000; Khourey et al. 2005; Lu et al. 2012). Therefore, it is necessary to
ascertain the seismic response of setback structures. To achieve this purpose, non
linear dynamic analysis has been conducted on setback building models and the
variation of inelastic deformation demands with setbacks has been presented in Figures
4.24 to 4.26. From these figures, it can be observed that the deformation demands in
case of setback buildings are maximum when setbacks originate from the bottom storeys
(as it results in greater reduction of floor area which reduces the mass, strength and
stiffness drastically). The percentage variation in deformation demands for setback
irregularity originating from bottom one - third storeys has been observed to be 34.45%,
37.23%, 35.23% (rd, Ir, i) and this percentage reduced to 25.23%, 23.12% and
24.12% (rd, Ird, i) for setbacks originating from top one third storeys (Figure 4.27). The
percentage increased with building height with maximum value for 36 storeyed buildings
41.34%, 45.23% and 43.12% (rd, Ird, i). Moreover, it has been observed from previous
earthquake failures that the failure of buildings occur near vicinity of setback (Refer
Figure 1.15 to 1.18, Chapter 1). This implies that presence of setbacks in a building
aggravates the deformation demands. Therefore, setback structures exhibit poor seismic
performance as compared to their regular counterparts and this seismic performance
degrades further with increase in the building height especially in case of tall structures
(18 storeys to 36 storeys).Moreover, presence of cracking has the maximum impact on
setback buildings with 13.23 % (mean percentage) increase of deformation demands.

186

0.75

Hr

SEBS
SEMS

0.5

SETS
R

0.25

0
2.4

3.3

4.2

5.1

Mean rd (cm)
Figure 4.20 Variation of global deformation demands along relative building height
for setback irregular buildings (SEBS, SEMS,SETS denotes building models with
setbacks originating from bottom one third, middle one third and top one third
storeys, Hr is the effective building height obtained by normalizing the maximum
building height to unity)

0.75

Hr

SEBS
SEMS

0.5

SETS
R

0.25

0
0.08

0.3

0.52

Mean Ird (cm)


Figure 4.21 Variation of storey deformation demands with relative building height
for setback irregular buildings

187

0.75

Hr

SEBS
SEMS
0.5

SETS
R

0.25

0
0.018

0.036

0.054

Mean i (rad)
Figure 4.22 Variation of local deformation demands along building height for
setback irregular buildings

(a)

(b)

Figure 4.23 Failure of Torre o Higgins building near setback (at 12th storey) during
Chile earthquake 2010 (Kato et al. 2010)

188

108

H (m)

90

SEA

72

SEB
SEC

54

36
18
2.4

4.05

5.7

Mean rd (cm)

Figure 4.24 Variation of global deformation demands along building height for
setback irregular buildings

108

H (m)

90

SEA

72

SEB
SEC

54

36
18
0.08

0.18

0.28

0.38

0.48

Mean Ird (cm)


Figure 4.25 Variation of storey deformation demands along building height for
setback irregular buildings

189

108

H (m)

90

SEA

72

SEB
SEC

54

36
18
0.018

0.034

0.05

Mean i (rad)
Figure 4.26 Variation of local deformation demands along building height for

Model category

setback irregular buildings

2
SEA
SEB
SEC

12

24

36

48

Mean Percentage ( % )
Figure 4.27 Mean percentage increase in deformation demands (on comparison
with regular building model) for setback irregular buildings with reference to
regular buildings (Model 1 Gross stiffness, Model 2 Cracked stiffness)

190

(d) Plan irregularity


The review of previous literature works suggests that the aspect of plan asymmetry has
been given utmost importance. The presence of the plan asymmetry makes it more
vulnerable to damage during occurrence of an earthquake which is evident from
excessive damage of buildings due to plan asymmetry during previous earthquakes as
listed in Table 4.2. The plan asymmetry results in torsion in a building which is due to
non-coincidence of center of mass and center of rigidity. The distance between these
two points is called as eccentricity (as discussed in detail in Chapter 1). The eccentricity
generated due to other factors like non-uniform distribution of load and difference
between the actual and computed stiffness/strength yield of elements. These factors
lead to coupling between lateral and torsional motions in the building system which
results in non-uniform distribution of the floor displacement and uneven seismic
demands in lateral resisting elements at different locations of the building systems which
eventually leads to its failure. The torsion originating from these effects is called as
accidental torsion. Due to damage caused by torsional irregularities during past
earthquakes, the seismic torsional response of irregular structures has been extensively
investigated by many researchers throughout the last few decades as discussed in
Chapter 1. The eccentricity can be easily estimated using code proposed equations as
given in Table 4.3) and by expressions proposed by other researchers.
The review of torsional provisions of different seismic design codes show similarity.
To verify the seismic design provisions the base building models as described in Chapter
2 are designed as per different torsional provisions for eccentricity ratio of P2. The
deformation demands are shown in Figures 4.28 to 4.30 from which similarity of seismic
design provisions could be clearly observed. Moreover, review of previous research
works and codes pertaining to plan irregularity shows that many research works have
been performed on SDOF systems using elastic analysis, but inelastic analysis of
multistorey building systems was studied by very few researchers. In most of the studies,
the researchers have used a simple shear beam model (Stathopoulos, and
Anagnostopoulos 2003; 2005; 2010; Anagnostopoulos 2010) and analyzed these
building models by subjecting them to limited number of ground motions. The shear
building models are not good representative of framed buildings in a seismic zone as
shear building model comprises of strong beams which causes plastic hinges to occur in
columns. This is in contradiction with strong beam weak column philosophy advocated
by seismic design codes (EC 8 : 2004; UBC 97) and previous research works (Tso
1994). Moreover, shear beam model may lead to unreliable estimate of seismic
response as modeling of a shear beam involves assigning a very high stiffness to beams
which may change fundamental period of vibration and mode of failure (Moghadam and

191

Tso 1996b). These shortcomings have been addressed in the present study and effect of
variation of plan irregularity has been evaluated for building models as described in
Chapter 2 as shown in Figure 4.31. From Figure 4.31, it could be observed that the
deformation demand increased with increase in eccentricity (varied from 0.15bw to
0.30bw). The mean percentage of increase in deformation demands with variation of
eccentricity (0.15bw to 0.30 bw) has been observed to be 3.23%, 5.12%, 4.43% for
parameters rd, Ird, i (Figure 4.32). This observation is consistent with failures of
building during observed during previous earthquakes (Figure 4.33) and reported
experimental (Negro et al. 2004; Jeong and Elnashai 2005; De - La colina 2007;
Mohammed et al. 2008) and analytical studies (Moghadam 1998; Perus and Fajfar 2005;
Sthathopoulos 2003; 2005; 2010) concerned with plan irregularity. The review of
previous literature reveals that majority of the researchers have ignored the variation of
deformation demands with building height for plan irregular building models. However,
from present research work, it could be observed that the deformation demands
observed a mean percentage increase of 4.23 % with building height (from 6 storey to 36
storey). However, this percentage of increase was smaller as compared to vertical
irregularities except for mass irregularity (Figure 4.4, 4.9, 4.13). Moreover, presence of
cracking aggravates the deformation demands by a mean percentage of 4.53 %.
Nevertheless, it is very necessary to propose new effective procedures considering
inelastic response of plan irregular building systems.
Table 4.2 Building failures due to plan asymmetry during different earthquakes
Earthquake and Year

Reference

Michanocan earthquake (1985)

Esteva 1987; Esteva and Ruiz 1989

Hyogoken Nanbu earthquake in Kobe Nakashima et al. 1998


(1995)
Loma Prieta earthquake (1999)

Mitchel et al. 1990

Turkish earthquake (Izmit earthquake 1999 , Dogangun 2004


Bingol earthquake 2003)
Wan chuan earthquake 2008

Zhao et al. 2009; Leiping et al. 2010

L Aquilla earthquake 2009

Ricci et al. 2011 b; Verdane et al. 2011

Darefield earthquake 2010

Kam et al. 2010

Great East Japan earthquake 2011

Fraser et al. 2013

192

Table 4.3 Eccentricities proposed by different seismic design codes


Design code

Primary design eccentricity

ASCE7.05;IBC

es 0.05 Axb

2009
EC 8:2003
KBCS 2005

NBCC1995

IS 1893

Secondary design eccentricity


es 0.05 Ax b

e2
)es 0.05bw
es

es 0.05bw

es 0.05bw

es 0.05bw

1.5es 0.10bw

1.5es 0.10bw

0.5es 0.10bw

0.5es 0.10bw

0.05bw

1.5es 0.05bw

(1

where es,e2,bw are static eccentricity, dynamic eccentricity and building width

108

90

H (m)

ASCE
72

EC8
KBCS
NBCC

54

IS
36

18
2.86

3.66

4.46

Mean rd (cm)
Figure 4.28 Comparison of global demands of plan irregular structures designed
using different seismic design codes

193

108
90

H (m)

ASCE
72

EC 8
KBCS

54

NBCC
IS

36
18
0.12

0.17

0.22

0.27

0.32

Mean Ird (cm)


Figure 4.29 Comparison of storey demands of plan irregular structures designed
using different seismic design codes
108

90

H (m)

ASCE
72

EC 8
KBCS
NBCC

54

IS
36

18
0.024

0.031

0.038

Mean i (rad)

Figure 4.30 Comparison of local demands of plan irregular structures designed


using different seismic design codes

194

108
90

H (m)

P1
72

P2

54

P3
R

36
18
2.4

3.025

3.65

4.275

4.9

Mean rd (cm)
(a)
108
90

H (m)

P1
72

P2

54

P3
R

36
18
0.08

0.205

0.33

Mean Ird (cm)


(b)
108
90

H (m)

P1

72

P2

54

P3
R

36
18
0.018

0.028

0.038

Mean i (rad)
(c)
Figure 4.31 Effect of plan irregularity on seismic deformation demands: (a) Global
demands, (b) Storey demands, (c) Local demands

195

Model category

P1
P2
P3

2.5

7.5

10

Mean Percentage ( % )

Figure 4.32 Mean Percentage increase in mean deformation demands (on


comparison with regular building model) for plan irregular buildings with
reference to regular buildings (Model 1 Gross stiffness, Model 2 Cracked
stiffness)

Figure 4.33 Building collapse due to torsion during L Aquilla earthquake 2009
(Verdane et al. 2010)

196

4.2.2 Relation between elastic and inelastic deformation demands


The relation between the inelastic and elastic seismic response has been determined in
terms of a parameter called irr. This parameter is defined as the ratio of maximum
inelastic response to the maximum elastic response. The parameter irr has been
determined and presented for different irregular building models as shown in Table 4.4.
This parameter can be used which is used in PBD and DBD methodologies as discussed
in detail in last section of this chapter.
Table 4.4 Mean values of parameter irr for irregular building models
Mean relation factor (irr)
N
M

ST

SE

1.012

1.0453

1.065

1.11

1.034

1.024

1.0723

1.092

1.14

1.056

12

1.037

1.092

1.112

1.227

1.087

15

1.045

1.128

1.153

1.252

1.112

18

1.054

1.1786

1.191

1.282

1.145

21

1.072

1.217

1.243

1.314

1.180

24

1.079

1.245

1.276

1.352

1.269

27

1.085

1.292

1.312

1.384

1.305

30

1.098

1.312

1.343

1.412

1.342

33

1.102

1.343

1.382

1.439

1.379

36

1.118

1.389

1.421

1.468

1.413

4.3 Determination of behavior factor for irregular buildings


The seismic design procedures formulated by seismic design codes are usually based
on the results of elastic analyses under the application of static forces. As per Eurocode
8 (EC8: 2004), the static force F equivalent to seismic action can be evaluated as

197

F 2.5

21Se / Ag (T )W
q

(4.1)

EC8:2004 gives a set of q-factors (where 21 is a constant and, Se /Ag (T )is spectral
acceleration at first mode period) related to the different types of structures, with their
potential capacity of energy dissipation conventionally evaluated on the base of the
ultimate failure mechanism.

4.3.1 Brief review of different European codes regarding behavior factor


This section investigates in detail the provisions of different European codes regarding
the aspect of behaviour factor. The EC8:2004 prescribes similar behavior factors for
precast and cast in situ structures provided these structures have seismic connections
as defined in EC 8 clause 5.11.2.1. In the old ENV version of EC8 (ENV 1998-1:1994)
precast frames were prescribed a lower q factor with respect to cast-in-situ frames. This
resulted in incorrect estimation of the structural ductility. As per EC 8 Overall ductile
behaviour is ensured if the ductility demand is spread over a large number of elements
and locations in the structure. As per this statement, the cast-in-situ monolithic frame
should be characterized by better seismic performance than the precast hinged frames.
In contrast similar seismic performance of both these frames were observed (Biondini
and Toniolo 2000). This point of view is also reflected by the new statement included in
the last version of EC8 (2004): Overall ductile behavior is ensured if ductility demand
involves globally a large volume of the structure spread to different elements and
location of all its storeys. Therefore, the amount of energy dissipated mainly depends
upon global volume involved in dissipation rather than the number of plastic hinges. The
Italian code (Technical rules for constructions DM 14.01.2008) prescribes similar
values of behavior factors as adopted by EC 8:2004 and gives similar values for precast
and cast in situ structures and hinged frames. For a brief period (20032008), a
provisional code for seismic design (Ordinance PCM3274) was applied which had rules
similar to EC8:2004.
However, Portugese and Yugoslavian code (YU 1981) prescribed similar behaviour
factors as EC 8 code (Biondini and Tonilo 2009). Greece has two codes namely a) the
seismic code EAK, b) the code for reinforced concrete EKOS which had provisions
similar to EC8 and EC2 codes respectively. For both these types of structures the qfactors are lower than those given by EC8:2004.
The FBD is a basic procedure for seismic design of building structures adopted in
earthquake-resistant design practice (EC8, 2004). In this method, the design base shear
is calculated by dividing the base shear (in the elastic range) by a behavior (or strength

198

reduction) factor q. The selection of appropriate values of behaviour factor was the most
controversial issue during the development stage of seismic design provisions for
building structures (Uang 1991). Nowadays, seismic design codes prescribe constant
values of q for various types of structures to restrict inelastic deformation at a level
necessary for protection of human life and economic losses in case of seismic excitation.
Although, the design lateral forces are computed with respect to the above-mentioned
limit state, majority of the seismic codes prescribe following additional limit states
(a) Resistance to minor earthquakes without damage of structural and non structural
elements
(b) Global collapse of the building to be avoided in case of the most severe seismic
event. The former can be called as an evaluation procedure after the detailing of the
structure and is not directly involved in the process of design. However, the second
criteria should always be satisfied as a result of the ductile design of the structural
elements FBD pertains to the performance-based design philosophy (Bozorgnia and
Bertero 2004) as different objectives are verified or assumed to be satisfied for different
levels of seismic intensity. Nevertheless, it would be more appropriate if these objectives
are directly involved as input parameters in the design process rather than being verified
after the detailing of the structure. The modified approach would permit the calculation of
the design base shears in accordance with various limit states (defined by the pair of
target deformation and intensity of seismic motion). Therefore, a new behavior factor is
required to convert FBD to a more rational direct performance-based design
methodology. Moreover, the q factor should be independent of the mechanical properties
of the frame to be used at the first steps of the design process where the sections are
not known yet but it should be a function of the target deformation to be applicable to any
limit state.

4.3.2 Review of literature with respect to behavior factor


The behavior or strength reduction factor has been the subject of intense research
studies since the 1960s. It is defined as the ratio of elastic base shear to the base shear
at occurrence of first plastic hinge. In general, the methods to evaluate the behaviour
factor can be broadly classified into two main categories based on degree of freedom as
: (a) methods based on SDOF systems; and (b) methods based on MDOF systems.
Furthermore, the behavior factor can also be classified based on the damage index used
in the definition of the limit state (which includes q factor also). The damage index
accounts for several parameters like
(a) maximum deformation

199

(b) cumulative damage


(c) maximum deformation and cumulative damage.
The research works using SDOF systems can be characterized by a broad range of
properties, such as the natural period of vibration, the force-deformation relationship, and
the level of viscous damping. The review of previous literatures show that SDOF
systems were subjected to a large number of physical accelerograms or by pulse
waveforms. Moreover, it was concluded that the parameters like the period of vibration of
the SDOF system and the ductility ratio had a significant impact on the behaviour factor.
Therefore, these research works have proposed reduction factors in a common q--T
format (behavior factor ductility ratio time period). The detailed review of these
studies is available in Miranda and Bertero (1994). Cuesta and Aschheim (2001) applied
a reduction factor (evaluated based on pulse waveforms) to the elastic spectra of ground
motions to determine the corresponding inelastic spectrum. This procedure to generate
the inelastic spectra was repeated by using the reduction factors proposed by other
researchers. A comparison of the produced inelastic spectra with the real ones showed
that the pulse reduction factor and the Vidic et al. (1994) proposed reduction factor
resulted in accurate estimation of the inelastic spectra. The proposed behaviour factors
have been incorporated in FEMA 273 (1997) and EC8 (2004) in terms of the nonlinear
static (pushover) analysis. The above-mentioned behavior factors resulted in a seismic
design which assumed that structural damage occurs only due to maximum response
deformation and hence the aspect of cumulative damage was ignored. However, a more
rational definition of the behavior factor should incorporate both maximum deformation
and energy dissipation. Miranda (1993) determined the effect of soil conditions on
behaviour factor. The authors proposed behaviour factors for different q--T relations for
firm, alluvium and soft soils.
Cosenza et al. (1993) determined the collapse spectra for a ductility ratio of 4 using
four accelograms. The collapse limit state was assumed as one for the selected damage
function. Chai et al. (1998) formulated the duration-dependent inelastic seismic design
spectrum using the strength reduction/ behaviour factor. Furthermore, the lateral strength
demand corresponding to the plastic energy capacity was observed to be equal to the
portion of input seismic energy contributing to the damage.
Manfredi (2001) proposed an energy spectrum (using behavior factor) in which the
hysteretic energy is calculated as a function of parameters like ductility, peak pseudovelocity, and a seismic index that considered ground motion duration as well. Malholtra
(2002) developed a cyclic-demand spectrum and concluded that the structural capacity
(strength, stiffness, and energy-dissipation) depends on the number of inelastic load

200

cycles. Decanini et al. (2002) proposed a strength reduction factor with reference to the
limit state of collapse. The results of analytical study confirmed the dependence of
collapse behavior factor on parameters like inelastic modeling rule, the soil class, the
ultimate ductility, and the period of vibration.
On comparison, majority of the researchers have proposed behaviour factor
considering SDOF system. Also, majority of the researchers have

correlated the

required maximum storey ductility of the MDOF system to be equal to the ductility factor
of the SDOF system by correlating the yield base shear of a MDOF system with the yield
strength of a SDOF system of the same period (Nassar and Krawinkler 1991;
Seneviratna and Krawinkler 1997).
Santa-Ana (2004) conducted an extensive study on the behaviour factor considering
MDOF systems. The authors adopted nonlinear dynamic analyses of steel buildings for
the ground motions for different soil conditions. The ratio of the SDOF base shear versus
the MDOF base shear was observed to be highly dependent on the number of storeys of
the frame, the soil conditions (stiff or soft) and ductility. Mohammadi and Naggar (2004)
proposed a relation for this ratio which was observed to be highly dependent on ductility.
The above mentioned research works have evaluated the behavior factor of MDOF
structures based on the reduction of the force of SDOF systems. Moreover, some
researchers [Elnashai and Broderick 1996; Kappos 1999; Mwafy and Elnashai 2002;
Grecea and Dubina 2003]

have adopted a different approach and evaluated the

behavior factor based on the reduction of the base shear of the actual MDOF building
structure. These research works define the limit state in terms of maximum deformation
damage indices such as the interstorey drift ratio (IDR) or the rotational demand. The
definition of limit states of MDOF building structures (based on damage indices that
account both for maximum deformation and energy dissipation) has been the object of
many research efforts as presented in detail in Ghobarah et al. (1999).
Khashaee (2005) and Sorace (1998) observed that the limited experimental results
are one of the major roadblocks in damage assessment of buildings. It was observed
that, the behaviour factor (considering MDOF system) directly related to the limit states
(described by energy based damage indices) was yet to be proposed. However, many
previous research works report correlation between maximum deformation indices and
cumulative damage indices. Thus, the behaviour factors as described in conjunction with
the above-mentioned correlations can be effectively used in the cumulative damage
based seismic design procedure. The proposed behaviour factor depended upon the
limit state under consideration and basic structural characteristics, such as the number of
storeys and the joint capacity design factor which the designer intends to adopt at the
phase of the selection of the sections of the frame. Any limit state can be explicitly

201

defined via the selection of the allowable value of a dimensionless deformation damage
index, i.e. the maximum storey ductility (). A correlation study between the maximum
and the Park-Ang damage index (Park and Ang 1985a) provided the tool for the
damage-based interpretation of the limit states under consideration.
Dolsek and Fajfar (2008) proposed a new behaviour factor as the ratio of elastic and
inelastic spectral acceleration. The behavior factor was in conjunction with equal
displacement rule and N2 method for determining seismic response of SDOF system.
The analysis results showed that the proposed procedure to yielded comparable results
with dynamic analysis. Moreover, using the proposed procedure, satisfactory results
were obtained for MI-RC frames. Similarly, adopting this procedure and EC8:2004
prescribed behaviour factor a adequate seismic performance of 4 storey RC frame was
observed by Dolsek (2010).
Athanassiadou (2008) used the EC 8 proposed behaviour factor for seismic design of
RC setback structures (DCM and DCH class). The inelastic dynamic analysis conducted
on the building models showed their satisfactory seismic performance. Moreover, the
ductility class was observed to have the least impact on the seismic performance which
was quite an important observation as DCM class will be more preferred as it is
economical and yields comparable performance with DCH class.
Krlik and Krlik Jr (2009) derived the behaviour factor using non-linear analysis as
per methodology suggested by Chopra (2001). The proposed behavior factor was
compared with EC 8 code proposed behaviour factor in determining the seismic
performance of irregular steel concrete composite frame (with stiffness and strength
irregularity). The analysis results showed the inefficiency of code proposed behaviour
factor in determining the effects of irregularity. However, the derived factor achieved this
purpose and hence was prescribed for assessment of irregular structures.
Tomaevic and PolonaWeiss (2010) through their experimental studies on masonry
buildings confirmed the adequacy of behaviour factors proposed by Eurocode 8 in
conjunction with pushover analysis in calculating the global ductility demand.
Hatzigeorgiou Liolios (2010) prescribed appropriate reduction in behaviour factor for
vertically irregular structures. The adopted behaviour factors showed a close agreement
with the non-linear dynamic analysis in estimating the ductility demand.
Ricci et al. (2011a) reported the poor seismic performance of Italian buildings (during
6th April 2009 LAquila earthquake) designed adopting the EC 8 behaviour factor. This
showed the inadequacies of EC 8:2004 provisions. In contrast, Magliulo et al. (2012)
observed EC 8:2004 proposed behavior factors to yield over-conservative estimate of
the seismic response for plan irregular structures. Kumar et al. (2013a) observed
behaviour factor to strongly influence the seismic response. In addition, the EC 8

202

provisions and EC 8 prescribed behaviour factor were observed to yield unconservative


estimate of seismic demands.
The review of literature showed that behaviour factor has a significant influence on
seismic response and design. However, the performance of the structures designed as
per code proposed behavior is quite unclear with some researchers finding the adequate
seismic performance

and others finding it to underestimate and overestimate the

seismic response. Moreover, the implication of behaviour factor in case of irregular


buildings is missing. Therefore, the present study aims to estimate the behaviour factor
for irregular buildings in terms of the proposed irregularity index.

4.3.3 Simplified relation between proposed irregularity index and behavior


factor
The previous section deals with review of code provisions and previous research works
pertaining to behavior factor. From the literature review it was observed that the code
proposed q factors were defined on the basis of empirical choices, not supported by a
rigorous investigation of sufficient reliability (Karavasilis et al. 2006). Moreover,
behaviour factors are assumed independent from the type of ground motion, effects of
material and construction practices, structural irregularity and cracking which is
unrealistic. Therefore, this section aims to evaluate a behavior factor addressing these
shortcomings. To account for building properties, ground motion parameters and
structural irregularity, the behaviour factor is evaluated as the scale factor of the
acceleogram that drives the structure to a particular performance level to that at
occurrence of the first plastic hinge (Elnashai and Broderick 1996). The different
performance levels can be obtained from SEAOC manuals (SEAOC 1999; SEAOC
2000). In the present study the behavior factor is evaluated as the ratio of scale factor of
the accelogram which at which the structure experiences maximum drift to the scale
factor of the accelogram at occurrence of the first plastic hinge. To evaluate the
proposed behavior factor non-linear static and non linear dynamic analysis has been
used. As the irregularity index proposed by the author (Chapter 2) is efficient in capturing
all forms of irregularities and cracking. Hence, the behavior factor in terms of
proposed irregularity index (Figure 4.34 and Figure 4.35).

the

Based on best fit the

equations for the plot between irregularity index and behavior factor, the author has
proposed equation to estimate the behaviour factor as
q 0.0305c3 0.758c2 8.123c 11.35 (for gross stiffness)
qc 0.0928c3 0.6644c2 9.023c 12.132 (for cracked stiffness)

203

(4.2)
(4.3)

7.2

R2 = 0.9929

5.85

4.5

3.15

1.8
0.4

0.7

c
Figure 4.34 Relation between irregularity index and behavior factor for gross
stiffness
7.2

R2 = 0.9859

q (Cracked)

5.9

4.6

3.3

2
0.4

0.7

c(Cracked)
Figure 4.35 Relation between irregularity index and behavior factor for cracked
stiffness (Priestley 2003)
On comparing the proposed behavior factors evaluated using the proposed
equations and dynamic analysis, a close agreement between both these results have
been observed (Figure 4.36 and 4.37). Moreover, using the irregularity index, behavior
factor for irregular buildings can be obtained. The proposed behavior factor can be
effectively used to generate response spectrums for a prescribed ductility level which
can further be used in seismic design of buildings (Chopra 2001; Fajfar 1999;
Karakostasa et al.2007). This aspect has been discussed in Chapter 3. Also, the
proposed behavior factor can be effectively applied for seismic analysis and design as

204

per EC 8: 2004 guidelines as discussed in the next section. It is worthwhile to note that
for comparison of behavior factors evaluated using proposed equation and dynamic
analysis, the maximum value of behavior factor is normalized to unity to achieve better
clarity of representation.
1

q (Proposed equation)

R2 = 0.9846

0.7

0.4
0.4

0.7

q (Dynamic analysis)

Figure 4.36 Relation between irregularity index and behavior factor for gross
stiffness

qc (Proposed equation)

R2 = 0.9907

0.7

0.4
0.4

0.7

qc (Dynamic analysis)
Figure 4.37 Relation between irregularity index and behavior factor for cracked
stiffness

205

4.4 Estimation of deformation demands


The estimation of deformation demands is a critical step in seismic design of buildings.
There are several methods to ascertain the deformation demands. However, equal
displacement rule has become quite popular due to its simplicity and ease of application
in determining the seismic capacity. This rule has been discussed in detail in the
following subsection.

4.4.1 Equal displacement rule proposed by EC 8 to estimate deformation


demands
The structural design of buildings in elastic range will be uneconomical as force
demands become very large. An economical approach can be adopted by accepting
some level of damage and consequently ductility of the structure can be effectively used
to reduce force demands of the structure to these acceptable levels. In general, ductility
of the structure can be defined as ability of the structure to withstand large deformations
beyond yield point without fracture. The ductility demand is the maximum ductility
experienced by the structure during an earthquake and it depends upon the structural
and earthquake characteristics. The ductility supply is the maximum ductility that a
structure can withstand without fracture. The design provisions must ensure a ductile
failure of structure rather than a brittle failure to ensure life safety. To achieve this
purpose, the ductility supply should exceed the maximum ductility demands and this is
known as capacity design principle. The capacity design provisions have been briefly
presented below
(a) The plastic hinge should form in beams before columns.
(b) An adequate confinement to concrete shall be provided using closely spaced hoops.
(c) The structural irregularities shall be avoided as far as possible.
(d) Adequate flexural and shear strengths of structural members.
The ductility can be easily defined in terms of displacement. For an SDOF system
with clear yield point, the displacement ductility is defined as the ratio of maximum
displacement and displacement at first yield as presented in equation 4.4.

xmax
xy

(4.4)

As per EC 8 : 2004, the force reduction or behavior factor is expressed as

Fe
Fy

(4.5)

where Fe is the peak force developed in the SDOF system if the structure were to deform
elastically and Fy is the yield load of the system.

206

As per EC 8: 2004, at long periods the elastic and yielding structures roughly experience
the same peak displacement and hence force reduction factor can be assumed equal to
ductility and this is known as equal displacement rule (Figure 4.38).

Figure 4.38 Pictorial representation of equal displacement rule ( EC 8 2004)


Lestuzzi and Badoux (2003) observed that the equal displacement rule was accurate
and independent of the hysteresis model (for both real and synthetic earthquakes) for
structures with initial natural frequencies below a frequency limit (generally between 1.5
and 2 Hz.
Pinto (2005) reviewed the EC 8: 2004 rule for estimating the deformation demands
and observed EC 8 provisions to be effective in designing the structures.
Karavasilis et al. (2006) determined the maximum displacement profiles of regular
multi - storeyed steel frames by conducting inelastic dynamic analysis on the building
frames. The influence of specific parameters such as the number of storeys, number of
bays, level of inelastic deformation induced (by the seismic excitation and joint capacity
design factor) on the inelastic deformation demands were studied. A series of large
number of non linear dynamic analysis were conducted to propose simple equations to
estimate the maximum displacement profile of multi - storey steel frames. The analysis
results indicated that among the parameters studied, the number of storeys had the
maximum impact on the displacement profiles of the frames. Finally, the proposed
equations were compared with EC 8:2004 proposed procedure and inelastic dynamic
analysis. The analysis results showed that proposed equations yielded comparable
results with dynamic analysis. However, EC 8:2004 procedure overestimated the
deformation demands.
Dolsek and Fajfar (2007) proposed a simple approach for probabilistic assessment of
plan asymmetric structures. In their analysis, the IDA

method was replaced by N2

method which was used in conjunction with equal displacement rule (used to define the
near collapse limit state) for seismic capacity assessment. For conducting the analytical

207

study the authors have adopted 20 ground motions form European seismic database
and the inelastic spectra was developed in accordance with Dolsek and Fajfar (2004).
Furthermore, this approach was compared with the IDA results for a realistic 3D building
(SPEAR building) modeled using FEM approach. The analysis results showed close
agreement between the proposed procedure and IDA analysis for the tested building.
This approach was further observed to be efficient for masonry infilled frames as well
(Dolsek and Fajfar 2008). This observation is contrary to findings of Karavasilis et al.
(2006) (based on regular structures), Karavasilis et al. (2008b); Karavasilis et al. (2008a)
who observed equal displacement rule to overestimate the seismic demands.
Biondini et al. (2008) proposed a simplified method using equal displacement rule in
conjunction with pushover analysis method for capacity based design of multistorey precast concrete frames with hinged beams. The structures were classified as ordinary and
less flexible structures as these structures had a fundamental time period less than 2.0
sec. The researchers included a partial safety factor of 1.20 to make allowance for
modeling uncertainty. The analysis results showed that the safe values of displacement
ductility were obtained for a behavior factor of 4.
Hatzigeorgiou and Beskos (2009) determined the inelastic displacement ratio for
SDOF system subjected to repeated earthquakes. The analysis results showed that the
inelastic deformation demands mainly depend upon several parameters like period of
vibration,the viscous damping ratio, the strain-hardening ratio, the force reduction factor
and the soil class. Therefore, based on regression analysis conducted on large number
of experimental results proposed simple equations were proposed to estimate the
inelastic deformation demands in terms of these parameters. The comparison of
proposed equation and EC 8:2004 equation showed that the EC 8:2004 obtained
demands were unrealistic.
Like Dolsek and Fajfar (2008), Alderighi, and Salvatore (2009) used equal
displacement rule in conjunction with N2 method for steel composite frames and
observed results of experimental and analytical studies in close agreement.
Perus et al.(2008) determined the target displacement for SDOF systems using a
quadra linear pushover curve in conjunction with equal displacement rule. The results of
proposed procedure and dynamic analysis were comparable for lower PGA. However,
differing from previous findings the equal displacement rule was observed to
underestimate the deformation demands.
Gunay and Sucuolu (2009) developed an equivalent linearization procedure based on
equivalent displacement rule and SDOF system. The procedure employed response
spectrum analysis and consisted of construction of equivalent linear system by making
appropriate reduction in size of the structural members. The maximum modal

208

displacement demands of an equivalent linear system were determined by equal


displacement rule and by using independent non-linear time history analysis. The
analysis results indicated that for the linear elastic spectrum, the proposed procedure in
conjunction with equal displacement rule was effective. However, use of inelastic
spectrum reduced the efficiency of this procedure.
Kreslin and fajfar (2010) have developed a procedure based on N2 method and equal
displacement rule. This procedure was applied to structures with plan and vertical
irregularity. The comparison of proposed procedure with non-linear dynamic analysis
showed that the proposed procedure provided a rough estimate of the seismic response
due to modeling uncertainty. The inclusion of higher modal contributions improved the
accuracy of proposed procedure. Still, it was observed to be inaccurate as compared to
dynamic analysis for irregular structures. Contrary to these findings, Mehanny and
Howary (2010) observed that equal displacement rule to accurately predict the seismic
response of multistorey building systems. Oropeza et al. (2010) observed this rule to
under-predict the seismic response in case of buildings frames with nonstructural
components.
Karavasilis and Seo (2011) determined the seismic performance of SDOF systems
with supplemental damping

designed as per EC 8 2004 employing the equal

displacement rule. The analysis results showed that the studied systems failed in the
global criteria which implies the under - prediction of seismic response by equal
displacement rule. This was later confirmed by Chen et al. (2012) and Karavasilis et
al.2012).
Sullivan (2011) proposed an energy factor method employing equal displacement
rule. It was observed that the proposed method yielded results comparable with dynamic
analysis for RC wall structures. Lumantarana (2011) determined seismic response of
structures (designed as per FBD consideration) subjected to stiffness and strength
deterioration. The authors used equal displacement rule to estimate the deformation
demand and compared it with the results of dynamic analysis. The analysis results
showed close agreement between the results of equal displacement rule and dynamic
analysis for displacement sensitive region of the response spectrum. However, equal
displacement rule underestimated the seismic response as compared to dynamic
analysis for the acceleration and velocity sensitive regions of the response spectrum.
Bhatt and Bento (2011) proposed a new procedure by adopting N2 method and
equal displacement rule for estimating the seismic demands of multistorey buildings. The
authors designed the buildings using EC8:2004 code; Italian code and the proposed
procedure. The analysis results indicated better seismic performance of structures
designed as per proposed procedure which showed the over-conservativeness of equal

209

displacement rule. This was later observed by Kumar et al. (2013b) for multi storey steel
frames.
Franchin and pinto (2012) proposed an approximate method for seismic design of
reinforced concrete (RC) structures to meet multiple structural performance requirements
as (1) the closed-form solution for the mean annual rate of exceedance of a limit state
and (2) the empirical equal-displacement rule. These design objectives were
incorporated in a gradient-based search algorithm. The analysis of results showed that
the proposed method yielded comparable results with inelastic time history analysis. The
brief summary of literature review pertaining to behaviour factor has been presented in
Table 4.5.
The above literature review reveals the difference in observations of researchers with
some observing adequate seismic performance of this rule whereas the others having a
different view (some researchers observing the equal displacement to under - predict the
seismic response and vice versa).

Table 4.5 Performance of equal displacement rule


Name of researcher

Year

Performance of equal displacement

Lestuzzi and Badoux

2003

Karavasilis et al.

2006

Accurate
rule
Overestimate

Dolsek and Fajfar

2007

Accurate

Dolsek and Fajfar

2008

Accurate

Karavasilis et al.

2008

Overestimate

Karavasilis et al.

2008

Overestimate

Perus et al.

2008

Underestimate

Hatzigeogrgiu and Beskos

2009

Unrealistic

Adergi and Salvitore

2009

Accurate

Gunay and Lu

2009

Accurate for linear elastic spectrum

Kreslin and Fajfar

2010

overestimate

Mehanny and Howary

2010

Accurate

Oropeza et al.

2010

Underestimate

Karavasilis and Seo

2011

Underestimate

Sulivian

2011

Accurate

Lumantarna

2011

Underestimate

Franchin and Pinto

2012

Accurate

Bhatt and Bento

2012

Overestimate

Kumar et al.

2013a,b Overestimate

210

4.4.2 Disadvantage of equal displacement rule


The equal displacement rule is widely prescribed by EC 8 : 2004 and is based on
following assumptions as
D Dy q

(4.6)

d dy q

(4.7)

In equations 4.6 and 4.7, uniform profiles of parameters Dy and dy were assumed
during the seismic excitation which was unrealistic and contradictory to previous
research works pertaining to the irregular structures (Athanassiadou 2008; Karavasilis et
al. 2008 a,b). These assumptions are still to be validated for RC buildings with
irregularities and an alternative rule to equal displacement rule is yet to be proposed for
irregular RC buildings.

4.4.3 Equations to calculate global demands


The global demand is generally expressed in terms of the maximum roof displacement.
As discussed earlier, EC 8:2004 employs equal displacement rule which is based on an
unrealistic assumption that displacement ductility and behavior factor are equal.
Therefore, there is a strong need of a new rule that gives a realistic estimate of the
deformation demands.
From the previous studies it has been observed that the parameters like irregularity,
behavior factor and building height affect the global demands significantly. Thus, based
on regression analysis conducted on the analysis results, the equation to estimate
maximum roof displacement for irregular buildings has been proposed as

rd 0.00621 c 0.00000862H 0.00713q

(4.8)

where H is in m and the maximum ductility demand is expressed in m.


The proposed equation can be used to compute the global demand for a given value
of c, q and H, and vice versa. These calculations are the preliminary requirements of the
performance based design (Li and Li 2010; Jones and Farzin 2010; Asgarian and
Nojoumi 2012), and displacement based design methods (Leiva 1996; Medhakar 2000;
Sullivan 2012; Tsai 2012).
On comparing the proposed equation with the EC 8: 2004 prescribed equation for the
total seismic response database, it can be said that the EC 8: 2004 equation
overestimated the seismic response (Figure 4.39). This fact is evident from higher
central value of 1.32( rd EC 8/ rd dyn ratio) as compared to the value of 1.09 (rd Eq/ rd dyn
ratio) for irregular building models. However, the proposed equation yielded results
closer to the dynamic analysis with a correlation coefficient of 0.9835 (Figure. 4.40). As

211

dynamic analysis is considered to predict the accurate results (EC 8: 2004; ASCE 7:05,
Tremblay and Poncet 2005; Chopra and Goel 2002 etc.). Therefore, it can be said that
the code results were uneconomical. It is worthwhile to note that in comparison of
seismic response parameters evaluated using proposed equation, EC 8:2004 rule and
dynamic analysis, the maximum value of these parameters has been normalized to unity
for better clarity of representation.

rd (EC 8)

0.725
2

R = 0.4382

0.45
0.45

0.725

rd(Dynamic analysis)

Figure 4.39 Comparison of EC 8 equation with dynamic analysis for rd


rd (Proposed equation)

R2 = 0.9835

0.725

0.45
0.45

0.725

rd(Dynamic analysis)

Figure 4.40 Comparison of proposed equation with dynamic analysis for rd

4.4.4 Equations to calculate Storey Level Demands in terms of maximum


interstorey drift ratio (Ir)
The seismic response database results suggested that the inter-storey drift ratio
exhibited a strong dependence on the root drift ratio (rd/H) and the irregularity index.

212

Therefore, based on correlation studies between these parameters, the equation to


evaluate the inter-storey drift ratio (Ir) for irregular building frames has been proposed as
I r 0.065c 0.012

rd
H

0.0095q 0.562

(4.9)

The proposed equation (first approach) represents a simple relationship and can be
used to compute Ir for a given value of rd, and

conversely the maximum roof

displacement can also be computed for any particular value of Ir. The comparison
between proposed equation, dynamic analysis and code equation showed that the code
equation overestimated the Ir value with a central value of 1.369 (code) as compared to
the central value of 1.08 (Proposed equation) as compared to the dynamic analysis.
Also, a close agreement between dynamic analyses and proposed equation has been
observed with a correlation coefficient of 0.9831 (Figure. 4.41 a).
A different approach (second approach) could be adopted to compute the maximum
inter-storey drift ratio and it is worthwhile to predict Ir on the basis of rd using equation
4.8 4.9. By this way the uncertainties of both these equations can be combined. On
comparing the values obtained by dynamic analysis and the proposed equation in this
case, the correlation coefficient was observed to be 0.9632 (Figure. 4.41 b), which is
somewhat lower than the earlier case (R2 = 0.9708). Although, the EC 8:2004 approach
is safe, but it is quite unreasonable (Figure 4.42). The method proposed by the author is
more preferable because it is simple and ensures both safety and economy in the
seismic design.

4.4.5 Equations to calculate Local demands in terms of maximum rotational


demand (i)
As per definitions of Kunnath and Kalkan (2004), the local demand of a structure
corresponds to the maximum inelastic rotations at the ends of beams and the columns. It
may be critical to stability of the structure and has been the reason for failure of structural
members in the past (Arslan and Korkmaz 2007). The evaluated maximum rotational
demand (i) was correlated with the strength reduction factor (q). Based on the
regression analysis conducted on the seismic response databank results of local
demands obtained for the building models considered, the equation for obtaining the
local demand for frames with irregularities of different magnitudes placed at different
locations has been proposed by the author as

i 0.005230 c 0.003712 q 0.01298

(4.10)

where q is the behavior factor and i is the maximum rotational demand computed in
radians. The equations 4.7 to 4.8 are simple and showed the influence of the proposed

213

irregularity index on the maximum rotational demand. The results obtained using the
proposed equations were compared with that of dynamic analysis and these
comparisons were plotted in the form of a graph (Figure. 4.43 a). The correlation
coefficient between both the methods was found to be 0.9926 which showed the
accuracy of the proposed approach in computing the local behavior of the irregular
structure. However, EC 8:2004 overestimated the local demands as observed in case of
global and storey demands (Figure 4.43b).

Ir (Proposed equation)

R2 = 0.9831

0.62

0.24
0.24

0.62

Ir (Dynamic analysis)
(a)
R2 = 0.9632

Ir (Proposed equation)

0.62

0.24
0.24

0.62

Ir (Dynamic analysis)
(b)
Figure 4.41 Comparison of equation using proposed equation and dynamic
analysis for Ir: (a) First approach, (b) Second approach

214

Ir (EC 8)

0.62
2

R = 0.4117
0.24
0.24

0.62

Ir (Dynamic analysis)

Figure 4.42 Comparison between EC 8 and dynamic analysis for storey demands

i (Proposed equation)

of irregular buildings
1

R2 = 0.9926

0.775

0.55
0.55

0.775

i(Dynamic Analysis)

(a)

i (EC 8)

R2 = 0.4096

0.775

0.55
0.55

0.775

i(Dynamic Analysis)

(b)
Figure 4.43 Local demands of irregular buildings: (a) Comparison between
proposed equation and EC 8:2004 procedure, (b) Comparison between proposed
equation and dynamic analysis

215

This page is intentionally left blank

216

4.5 Effect of cracking on deformation demands


As already discussed in the previous chapter that the phenomenon of cracking affects
the seismic response. Cracking increases the deformation demands due to stiffness
reduction and this aspect has been ignored by the design codes (IS 1893:2002; EC 8
2004, UBC 97 etc) and previous research works pertaining to the structural irregularities
(Valmudsson and Nau 1997 Al-Ali and Krawinkler 1998; Das and Nau 2003; Karavasilis
et al. 2008 a,b). However, some research works (Paulay and Priestley 1992; Priestley
2003 etc) have considered this aspect and have proposed the appropriate reduction
factors to account for cracking. The different stiffness reduction factors proposed by
different authors to account for cracking have been already discussed in the previous
Chapter. Among the different stiffness reduction factors proposed, the Priestley (2003)
has been adopted due to reasons explained in Chapter 2. Therefore, accounting for the
cracking consideration a new set of seismic response data has been obtained by nonlinear dynamic analysis. Based on the regression analysis conducted on this seismic
response database, the equations to estimate the deformation demands of the irregular
building frames are proposed (Table 4.6). The range of applicability of proposed relations
have been previously mentioned in Section 3.5 (Chapter 3).
Table 4.6 Equations to estimate the Seismic response parameters for irregular
building models
Equations to estimate Seismic response parameters

R2

rd

rd 0.06321c 0.0000921H 0.654q

0.9823

Ir

0.000834q 0.01348c 4.73

rd
0.02136
0.0086H

0.9912

Parameter

i 0.006129c 0.0082162q 0.01118

0.9823

4.6 Application of the proposed irregularity index and proposed mean


relation factors in DBD and PBD
The main aim of this section is to discuss the applicability of the author proposed index,
proposed equations and proposed mean relation factors in DBD and PBD respectively.
In this section, the applicability of the proposed mean relation factors and proposed
irregularity index has been shown in DBD methodology adopted by Kappos and
Stefanidou (2010). This methodology is specifically developed for design of irregular

217

buildings and hence is used to demonstrate the applicability of the present research
work. The main steps of this method are
(a) Elastic design of the structure
In this step, Kappos and Stefanidou (2010) conducted elastic analysis to determine the
section dimensions of the structural members, and the maximum elastic deformation
demands for these members were determined consequently. The corresponding value
for the maximum inelastic deformation demands are obtained using the relations
proposed by Panagiotakos and Fardis 2001. Finally, using the value of maximum
inelastic deformation demands, the sections are redesigned to obtain the modified cross
section values and reinforcements. It is worthwhile to note that the relations proposed by
Panagiotakos and Fardis (2001) are based on buildings with open first storey. Therefore,
these relations ignore the aspect of other forms of irregularity. Therefore, these relations
are inapplicable to building models with other forms of irregularity. Moreover, these
relations are based on plastic hinge rotation ratio which is less critical in determining the
seismic performance as compared to inter-storey drift ratio which is the most important
parameter representing stability of structure as indicated by seismic design codes (IS
1893:2002, EC 8:2004; UBC 97). However, the author proposed relations were
specifically based on buildings with different types of irregularities and IDR is considered
in proposing these relations.
As compared to previous approach of Kappos and Stefanidou (2010.) The authors in
their approach have fixed the maximum elastic storey drift as per EC 8:2004, which limits
the maximum relative drift between the storeys equal to 0.0075h/v.
where h is the storey height and v is a reduction factor equal to 0.5 for class iii and
class iv
Therefore, for this structure the maximum relative drift is limited to 45 mm [(0.0075 x
3000)/0.5], and using the relation factor the corresponding inelastic relative drift is
obtained using Table 4.2 as 55.215s mm (45 x 1.227). Using this deformation value the
corresponding section dimensions and reinforcements are obtained.. The description
and comparison of seismic performance of these models used for design are shown in
Figures 4.44 and 4.45. The building models A, B, C are regular frames designed in
accordance with method suggested by EC 8:2004, Kappos and Stefanidou (2010), and
author. Similarly, the corresponding irregular frames are named as D, E and F.
(b) Selection of seismic action
An ensemble of 27 ground motions were applied to the selected building models and
Ibarra et al. (2005) model has been selected to model the inelastic behavior.

218

(c) Serviceability verifications


The serviceability verification of a structure is a critical step in seismic design as the
displacement based design is mainly based on limiting a certain deformation parameter
(Inter-storey drift ratio in this case) ignoring other parameters. Therefore, other
deformation parameters (maximum roof displacement and maximum plastic hinge
rotation) may violate the code prescribed limit for the adopted member sections (This
aspect has been described by Kappos and Stefanidou 2010 as one of the main
weakness of the displacement based design). Therefore, serviceability check necessarily
should be performed to ensure that the structure satisfies all the safety criteria. To verify
the revised design, serviceability verifications have been performed on the refined
building models by subjecting the building models to an ensemble of 27 ground motions.
The anticipated damage level correspond to easy and repairable damage preferably to
the non structural elements. Moreover, the longitudinal reinforcement has been already
defined by limiting the inelastic deformations which excludes the possibility of occurrence
of local and global failure. The main aim of this step was to check the seismic
performance for the selected ground motions. The performance objectives are checked
for inelastic inter-storey drifts and plastic hinge rotations. To determine the inelastic
seismic response, Kappos and Stefanidou (2010) performed the complex inelastic
modeling and this inelastic model was subjected to the nonlinear dynamic analysis. The
overall design procedure becomes quite complex and time consuming due to this step,
but this step was quite essential and cannot be avoided. However, using the author
proposed relations; the serviceability verification results can be easily determined without
performing the complex inelastic modeling and non-linear dynamic analysis. The ground
motion adopted are same as for 2D and 3D building models. The value of seismic
response parameters with code permissible limits for all building models using both
inelastic dynamic analysis and author proposed equations have been presented in
Figures 4.46 and 4.47. From these figures. it can be clearly observed that all the building
models considered exhibited a satisfactory performance on serviceability criteria and the
regular building models (Model A Model C) exhibited a similar seismic performance.
However, for irregular building models (Model D Model F) considerable difference in
seismic response was observed with Model F (designed using author proposed
relations) performed exceptionally well on both serviceability and economy criteria
(Figure 4.46 and Figure 4.47) as compared to the other irregular building model (Model
D and Model E). In a similar manner, the proposed mean relation factors and the
irregularity index can be applied to performance-based design (PBD) as proposed by
Kappos and Panagoupoulos (2004).

219

Figure 4.44 Design details of columns of the regular and irregular building models
studied

220

Figure 4.45 Design details of beams of the regular and irregular building models studied
(All dimensions in mm)

Figure 4.46 Reinforcement weights for building models

221

(a)

(b)
Figure 4.47 Serviceability criteria check based on inelastic analysis: (a) Interstorey
drift (cm), (b) Rotation (radians)

4.7 Brief summary and main conclusions


This Chapter mainly deals with the study of the effect of different types of irregularity on
height-wise distribution of inelastic deformation demands. This chapter has been divided
into four sections. In the first section, the code guide lines regarding the deformation
demands with their limitations. The shortcomings of previous literature works pertaining
to the deformation demands have been discussed.
In the second section, elastic and inelastic deformation demands have been
evaluated and variation of these demands along the building height in correlation with
reality (failure of irregular buildings during previous earthquakes) has been presented in
brief. Secondly, the mean ratio of inelastic to elastic deformation demands for different
building models are presented in the form of a parameter irr.
In the third section, a detailed literature review regarding the aspect of behaviour
factor has been presented. Secondly, a different approach to evaluate the behavior
factor in accordance with Elnashai and Broderick (1996) has been adopted to evaluate
behavior factor of irregular buildings with gross and cracking consideration. Thirdly, the
applicability of behaviour factor is briefly presented and the proposed behaviour factor is

222

used in equations to estimate the deformation demands as discussed in later sections of


this chapter. In the fourth section, aspect of equal displacement rule has been discussed
in detail (literature review and its shortcomings, disadvantages of equal displacement
rule). Secondly, based on regression analysis conducted on analysis results, simple
equations have been presented to estimate the

deformation demands. In these

equations, deformation demands have been presented as a function of irregularity index,


behavior factor (evaluated using the proposed equations. Finally, applicability of the
proposed equations and parameter ir (evaluated in previous sections of this Chapter)
has been discussed in current methods of seismic design (DBD and PBD). The main
conclusions derived from this Chapter are as follows
(a) The review of code procedures shows that the limits of drifts are based on elastic
analysis and are defined in terms of storey height. These drift limits are observed to be
overconservative and uneconomical. The magnitude and location of irregularities had a
strong influence on the inelastic seismic response. As compared to mass and plan
irregularity, the variation in magnitude of stiffness, strength and setback irregularity had a
more pronounced effect on deformation demands. Regarding the aspect of irregularity
location, the presence of stiffness, strength and setback irregularity in bottom storeys
were observed to be more critical. However, a reverse was observed for mass irregular
building models.
(b) The review of previous literature works showed that majority of the research works
are confined to low rise and mid rise buildings varying from 0 18 storeys
(Athanassiadou 2008; Karavasilis 2006; Karavasilis 2008 a,b; Kappos and Stefanidou
2010; Kreslin and Fajfar 2012). However, current statistics of building heights
(Wikepedia.org) in different countries indicate that majority of buildings with greater
heights (with storey height from 18 to 36 storeys). These structures are preferred due to
space constraint and functional requirements. Therefore, estimation of deformation
demands of these buildings is essential. The analysis results showed a larger variation of
inelastic deformation demands (due to variation in magnitude and location of
irregularities) for tall buildings (18 36 storey) as compared to mid - rise and low rise
frames. These analysis results were compiled to generate and regression analysis
conducted on seismic analysis results to propose simple equations to estimate seismic
response parameters like maximum roof displacement, interstorey drift, rotation,
overstrength, stress etc.
The proposed equations have been compared with dynamic analysis and EC 8
equations. The comparison showed close agreement between proposed equation and
dynamic analysis with a high correlation coefficient. However, equal displacement rule
EC 8 procedures overestimated the seismic response. This implies that EC 8 design

223

provisions are overconservative and uneconomical. Moreover, the proposed equations


incorporated the effect of critical parameters like type, magnitude and irregularity location
along with structural cracking which was missing in previous research works and code
procedures to estimate seismic demands. Finally, it was observed that proposed
parameter irr effectively fitted into current methods of seismic design (PBD and DBD)
and is observed to yield safe and economical design as compared to EC 8 : 2004 and
Kappos and Stefanidou (2010) approach. Hence this parameter can be used as an
effective alternate to equal displacement rule (proposed by EC 8: 2004) in conjunction
with PBD and DBD in seismic design of structures.

224

CHAPTER 5
PREDICTION OF SEISMIC PERFORMANCE OF IRREGULAR BUILDINGS

5.1 Introduction
The reinforced concrete (RC) frame structures mainly rely on beam and column elements to
resist both seismic and gravity loads. The inability of these elements to carry the load leads
to collapse of the structure. Collapse capacity of structures under seismic excitation has
always been a crucial aspect in determining seismic performance of the structure. Accurate
prediction of the collapse capacity is very important as the structural collapse is a source of
life and monetary losses. Few seismic design codes (IBC 2003) and some reported studies
(Jones and Farzin 2010; Jones 2012; Lignos and Krawinkler 2009; Kim and Lee 2010; Kim
and Choi 2011) have discussed methods to determine the collapse capacity of the building
structures. Some researchers (Adam et al. 2004; Bernal et al.1987; Miranda and Akkar
2003; Takizawa and Jennings 1980) have used SDOF systems using FEM analysis and
other researchers (Vamvatsikos and Cornell 2002; Medina and Krawinkler 2003; Ibarra et al.
2005; Zarein and Krawinkler 2009) have preferred incremental dynamic analysis. This
approach predicts the median collapse capacity in terms of maximum storey drift. Since the
frames used in the analytical study are MDOF systems, the used approach is different from
methods employed for SDOF systems. Furthermore, researchers like Ibarra and Krawinkler
(2004) have proposed a new approach for modeling of beam-column connections for
seismic hazard analysis and collapse capacity assessment.

5.2 Review of research works regarding collapse capacity of buildings


Many of the buildings as per current provisions of seismic design codes have failed during
the past earthquake. This shows the inefficiency in design provisions of current seismic
codes. The seismic design codes are based on parameters like strong column weak beam
philosophy and storey drift limitation. These parameters just give broad guide lines and have
never been demonstrated by analytical or experimental or field case studies (Liel 2008).
Jennings and Husid (1968) observed lateral instability collapse of structures due to
increasing level of deformations in the inelastic range. This effect is not incorporated in the
modern seismic codes. Similarly, Bernal (1987) found that code provisions account for the
P-Delta effects, but this consideration was based on static analysis of the structures. Bernal

225

1992; Willliamson 2003 observed that the inelastic dynamic collapse of the structure cannot
be avoided by limiting the magnitude of elastic storey drifts.
Challa and Hall (1994) observed flexural collapse of columns of 20 storey building
designed as per code provisions due to formation of plastic hinges leading to flexural failure
of columns. Similarly, Martin and Villaverde (1996); Medina and Krawinkler (2005) observed
large amount of plastic hinging in columns of regular frames (designed as per strong column
weak beam design philosophy of current seismic code provisions). Roeder et al. 1993;
Schnieder 1993 observed inelastic storey drifts exceeding the prescribed limits in a structure
designed as per 1988 UBC provisions.
These collapses of the modern buildings indicated the inadequacy of the current
seismic design code provisions. Therefore, prevention of collapse oriented approach of
design is the primary objective of the present study.
Bernal et al 1987; Esteva 1992; Hamburger 1997; Astaneh-Asl 1998; Li and Jirsa
1998; Griffith et al. 2002 observed that there are no other measures to estimate collapse
capacity apart from code recommended measures. The code methodology for estimation of
collapse capacity is not reliable as collapse of a structure involves complex mechanisms like
excessive deformations, P-delta effects, material degradation, local buckling, yielding and
cracking. In addition, the code suggested methodology ignores the behaviour of the
structure under the action of dynamic loading. However, Araki and Hjelmstad (2000)
observed that the unloading stage immediately after loading cycle might restore structural
stability.

5.2.1 Review of collapse assessment methods


The review of seismic design codes and previous research studies showed that the
collapse assessment methods were mainly based on the code prescribed drift limits ignoring
the effect of seismic excitation. Majority of the past research works pertaining to collapse
capacity assessment were based on SDOF systems with few researchers preferring MDOF
system. These approaches have been briefly discussed below:

(a) Single degree of freedom systems


The collapse assessment methods can be developed considering either SDOF or MDOF
system. Takizawa and Jennings (1980) used SDOF systems to determine the ultimate
seismic capacity of the building system subjected to the combined action of gravity loads
and seismic excitation. The proposed model consisted of two rotational springs at the end of

226

beam column joints and exhibited a trilinear force deformation behavior. The building
model was subjected to seismic excitation which led to its collapse. From the analytical
study it was observed that
(i) The collapse of a structure is significantly influenced by the duration of the ground motion
and it increased for ground motion of longer duration and vice versa.
(ii) The increase in stiffness of a structure created a significant difference between the
collapse and damage states of the building.
(iii) A single parameter is incapable of predicting the destructive potential of the ground
motions. This observation was contradictory to previous research works (Ibarra and
Krawinkler 2004; Hall 1994; Krawinkler et al. 2003).
In a similar study, Bernal (1987) determined the influence of the gravity loads by
considering the P - Delta effects in terms of an empirical formula consisting of ductility factor
and stability coefficient. Furthermore, a non-degrading SDOF system and an ensemble of
four record motions were used for generation of this spectra. The proposed formula was
derived from an amplification spectra which was generated by dividing the inelastic
spectrum ordinates obtained under the influence of gravity effects to those ordinates
obtained by ignoring the gravity effects. The results of analytical study showed that the
amplification factors as proposed by the seismic design codes were underpredictive, which
increased with increase in the ductility factor.
Bernal (1992) developed a method to determine the safety of two dimensional buildings
against dynamic instability. The multistorey structure was idealized as equivalent SDOF
system. Based on statistical regression, the collapse safety margin was proposed as the
ratio of actual base shear capacity of the structure divided by the statistically determined
minimum base shear. Based on results of the analytical study for series of multistorey
structures, it was concluded that the collapse safety against dynamic instability strongly
depended upon shape of the failure mechanism. For the regular structures, the collapse
mechanism was identified by the static analysis by assuming lateral forces proportional to
the storey weights.
Extending Bernals approach MacRae (1994) proposed a method which considered
second order effects on oscillators with different force-deformation behavior. The results of
analytical study showed that the inelastic deformation demands and structural stability had a
strong influence on post-elastic to elastic stiffness ratio.
Williamson (2003) determined the effect of damage accumulation and second order
effects on inelastic seismic response of the building systems. The model used in the

227

analytical study consisted of a rigid column with a concentrated mass at top with a rotational
spring. The rotational spring was modeled with a bilinear degrading moment - rotational
relationship at the base. It was observed that the damage accumulation and second order
effects had a significant impact on behavior of the studied system even for the small values
of the axial force.
Miranda and Akkar (2003) proposed an empirical relationship for SDOF systems to
determine the minimum lateral strength of the structure required to avoid its failure against
the dynamic instability. The proposed relations were based on non-linear regression
conducted on large number of SDOF systems (founded on stiff soil subjected to an
ensemble of 72 ground motions). The collapse strength was expressed as the function of
natural period and post-yield stiffness of the system as these two parameters affect the
system collapse significantly. The studied system was a SDOF system with bilinear force
deformation relationship as indicated by the post-yield branch with a negative slope. The
force deformation behavior represented the effect of geometric nonlinearity and strength
degradation. The system collapsed by degrading force-deformation relationship when the
restoring force in the system reduces to zero. Finally, Adam et al. (2004) proposed a
procedure to estimate the collapse capacity of MDOF system considering the PDelta
effects. However, for simplicity the authors had idealized the building systems used as
SDOF systems. The properties of the SDOF system were based on the results of pushover
analysis. The pushover analysis assumed that the post-yielding global stiffness obtained
represented the global or local mechanism involved, when the structure approached the
state of dynamic instability. The collapse capacity was determined through a series of
dynamic analysis on SDOF system. The collapse was assumed to occur when a small
increment in the ground motion intensity resulted in large deformations. The accuracy of the
proposed procedure was confirmed by conducting the analytical studies on a two single bay
SDOF systems subjected to an ensemble of 40 ground motions.
Han et al. (2010) developed a modal pushover analysis (MPA) based approximate
procedure for quantification of collapse potential of structural systems. The computationally
complex incremental dynamic analysis (IDA) procedure was avoided This procedure was
replaced by MPA of the structure in conjunction with empirical equations to determine
collapse strength ratio for the first-mode of SDOF systems. The analysis results showed that
higher modes of vibration played essentially no role in estimating the ground motion
intensity required to cause collapse of the structure which

was contradictory to the

observations of previous research studies(Chopra and Goel 2002; Chopra and Goel 2004).

228

The collapse fragility curves were evaluated for 6, 9, and 20 storey regular special momentresisting steel frames computed by the MPA and IDA analysis, The analysis results showed
that the MPA based approximate procedure required only a small fraction (1% in one
example) of the computational effort as compared to IDA analysis, and still highly accurate
results were achieved.

(b) Non-linear static analysis


In recent years non-linear static analysis (pushover analysis) method has become quite
popular and is widely used for determination of collapse capacity of the structures
(described in detail in Chapter 1). The pushover method was introduced in FEMA
Publication No. 273 (1997) and was updated in FEMA Publication No. 356 (2000). This
procedure is applicable in cases where higher modal contributions are insignificant. If these
higher mode effects are significant, then non - linear dynamic analysis of the structure needs
to be performed along with pushover analysis. As per procedure suggested by FEMA-356,
the structure is modeled considering the nonlinear force deformation behavior of structural
elements. Then, base shear - roof displacement relationship is established by subjecting the
building model to monotonically increasing lateral forces along the building height. This
procedure is applied till the roof displacement exceeds the target displacement. The target
displacement is intended to represent the maximum displacement likely to be experienced
by the structure under a selected seismic hazard level.
The seismic demands at the prescribed target displacement element forces, storey drifts
or plastic hinge rotations are compared with the code prescribed criteria of acceptability
which depends on material properties like the construction material , structural elements and
prescribed performance level. The structural collapse is assumed when the base shear roof displacement curve exceeds the point of zero slope and follows a negative slope due to
P - Delta effects. At this point, the structure is incapable to resist gravity loads. This
procedure is based on assumptions that
(i) The nonlinear response of a structure can be related to the response of an equivalent
SDOF system.
(ii) Lateral distribution of forces assume a uniform profile over entire height of the structure
during the seismic excitation and phenomenon of cyclic effects is ignored. In addition, the
progressive changes in the dynamic properties (that take place in a structure as it
experiences yielding and unloading during an earthquake) are neglected. However, in
reality, the nonlinear structural behavior is load-path dependent and deformation demands

229

depend on ground motion characteristics. It may lead to underestimation of storey drifts and
inaccurate estimation of the location of formed plastic hinges. This is particularly true for
structures deforming in the inelastic range.
(iii) Krawinkler and Seneviratna 1998; Chi et al. 1998; Kim and DAmore 1999; Gupta and
Kunnath 2000; Goel and Chopra 2004; Chopra and Goel 2004; Maison and Hale 2004 have
found that the above non - linear procedure did not provide an accurate assessment of
behavior of structures during Northridge Earthquake 1994.
To overcome these shortcomings, an improved nonlinear static procedures considering
the fundamental mode was proposed by researchers like Bracci et al. 1997; Sasani et al.
1998; Gupta and Kunnath 2000. Some other research works considered the contribution of
higher modes (Sasani et al. 1998; Chopra and Goel 2002; Goel and Chopra 2005;
Elghazouli 2008; Han et al. 2010; Shafei et al.2011; Kreslin and Fajfar 2011; Kreslin and
fajfar 2012) and advocated the use of use a time-variant distribution of the equivalent lateral
forces. On observing the performance of these procedures, it was found that these
procedures led to better prediction in some cases (Han et al. 2010; Shafei et al. 2011), but
none of these procedures were universally applicable. Thus, nonlinear static methods are
unreliable in predicting the collapse capacity of structures and margin of safety against a
global collapse.
(c) Finite Element analysis
In determining the seismic collapse capacity, many of the researchers differing from the
previous approach of using the nonlinear static analysis have adopted a step-by-step finiteelement analysis. The review of previous literature works show that the assessment of the
collapse capacity of a structure (to resist an earthquake induced collapse) is tedious. This is
due to dependence of collapse capacity on large number of parameters as
(i) The characteristics of the ground motion e.g., intensity, frequency content, and
duration of seismic excitation
(ii) Dynamic properties of the structure (Fundamental time period of structure and mass
participation factor)
(iii) Structural geometry.
(iv) Post-elastic and post-buckling behavior of structural members.
(v) Strength and stiffness of these components and their degradation due to cyclic
loading.
(vi) Due to interaction between vertical loads and lateral drifts.

230

(vii) Due to the interaction of the structure with its nonstructural components e.g., the effect
of components such as stairways and cladding on structural stiffness and strength.
(viii) Soil-structure interaction effects.
(ix) Residual stresses and initial imperfections e.g., the influence of fabrication residual
stresses and member out-of straightness on member stiffness.
Therefore, finite element analysis is very essential so that all these parameters are
included which will result in accurate prediction of structural collapse capacity. This analysis
consists of following steps
(1) The deformed configurations of the structure are used to define the equations of motion
and consequently these configuration are updated at each step.
(2) To improve accuracy of this approach, nonlinear large-deformation elements were used.
(3) Different aspects like plasticity, local instabilities, and crack formation are also included
to obtain a realistic estimate of collapse capacity.
(4) The structural analysis (with different ground motions and characteristics to obtain better
accuracy) is carried out and seismic response of the structure is monitored for any abrupt
variation.
Ger et al. (1993) determined the factors that initiated the collapse of a 22 storeyed steel
building during the 1985 earthquake in Mexico City. The structure consisted of open-web
girders, welded box columns, and H-shaped diagonal braces. Firstly, the non-linear hysteric
relationships were developed for the components mentioned. Then, using

three-

dimensional finite-element model (including the geometric stiffness of its structural


elements), the building model was analyzed under the action of three components of the
ground motion recorded at a station near the buildings site during the earthquake. The
analysis results indicated that the ductility demands exceeded the prescribed limits in case
of longitudinal girders. Due to failure of these girders a redistribution of forces and moment
occurred. This resulted in local buckling of the columns of the second and fourth storeys.
This local buckling, in turn, resulted in loss of load carrying capacity of the columns which
generated torsion and P-Delta effects which eventually led to the collapse of the structure.
The observed results were comparable with an identical building adjacent to the collapsed
one. Therefore, it can be concluded that collapse of a structure could be accurately
predicted using the force - deformation relationships.
Likewise, Challa and Hall (1994) evaluated the nonlinear response and collapse
capacity of moment-resisting plane steel frames under the action of seismic excitation. To
achieve the above aim, a 20-storyed frame designed in accordance with zone 4

231

requirements of the 1991 Uniform Building Code (UBC 1991) was analysed under two types
of ground motions. The first ground motion used was of the oscillatory type with long period
components (a scaled-up version of a record from the 1971 San Fernando earthquake), and
the second one was of the impulsive type (simulated displacement pulses intended to
represent a near-field ground motion). The beam-column joints were modeled using shear
panel elements. In modeling of the elements, the geometric stiffness of the elements and
realistic stress-strain relationships for both the element fibers and the panel zones were
considered. The deformed shape of the structure was studied at each step interval of the
seismic excitation. As such the phenomenon like, strain-hardening, axial-flexural yield
interaction, residual stresses, spread of yielding, P Delta effects and column buckling were
accounted for in the analysis but the effect of stiffness and strength degradation was
neglected. It was observed that during the collapse of the structure, the frame reaches a
state in which its lateral displacement increased with each cycle of response and then a
unsuitability in a structure occurred due to P-Delta effects. Finally, it was concluded that
ignorance of strength and stiffness deterioration would just simply delay the structural
collapse but would not avoid it under the action of ground motion of less severe intensity.
This observation was contrary to Ibarra et al. (2005).
Differing from previous approaches to estimate collapse capacity, Martin and Villaverde
(1996) developed a methodology for assessment of partial or total collapse of a structure to
identify the structural members that would fail first and would subsequently lead to collapse
of the structure. The proposed method consisted of a step-by-step nonlinear finite-element
analysis and subsequently the effective stiffness matrix of the structure was updated at each
step of the analysis. The partial or total collapse of the structure was assumed at a point
where the effective stiffness matrix became zero or negative. The part of the structure at
which the collapse was initiated at first was located by identifying the structural nodes that
corresponded to the zero or negative stiffness matrix. This methodology was applied to a
steel cantilever beam and a two storey, three-dimensional steel moment-resisting frame and
accuracy of this method was validated.
Mehanny and Deierlein (2001) proposed a methodology to evaluate the structural
collapse under the seismic action using a component damage index developed by them.
The building models were modeled using finite element approach and subjected to the
inelastic time-history analysis for calculation of the damage index for structural members. A
global stability index was further proposed as the ratio of the gravity load at which the

232

structure reached its global stability limit to the actual gravity load. This index correlated the
collapse capacity to ground motion intensity.
Hanganu (2002) adopted finite element modeling to determine the seismic performance
of RC buildings. A numerical procedure for the prediction of local and global damage in civil
engineering structures using the finite element method and a continuum damage model
were presented by the authors. The proposed method was observed to be adequate for the
computation of the limit load in reinforced concrete (RC) structures and for the prediction of
the failure mechanisms. Finally, the applicability and efficiency of the proposed method was
demonstrated with the aid of some numerical examples.
Sabelli et al. (2003) determined collapse capacity of steel braced frames with buckling
restrained braces. The finite element modeling approach was adopted to achieve better
accuracy in collapse capacity assessment. The analytical results showed improved seismic
performance of buckling restrained frames as compared to their regular counterparts.
Haselton and Deierlein (2007) conducted collapse capacity assessment of a four-storey
office building using incremental dynamic analysis. The structure was modeled using FEM
approach. The analysis results indicated that the seismic performance of the building was
appropriate with the expectations of current building codes, having a life-safety performance
level (ASCE 356) of 2% in 50 year event. Furthermore, sensitivity analysis showed that the
record-to-record variability was the single most important contributor to the total variability in
structural response. However, the combined affect of other modeling and design
uncertainties had a significant impact on collapse capacity assessment resulting in the total
dispersion being roughly double as compared to the dispersion caused due to record-torecord variability. To further investigate modeling uncertainties, the researchers compared
the responses predicted by using a force-based fiber model and a lumped plasticity model.
These comparisons showed that the concrete tensile strength and tension stiffening effect
had a significant impact on collapse capacity assessment.
Khandelwal et al. (2007) developed finite element based macro-models of buildings to
determine the progressive collapse of steel framed buildings. The developed models were
calibrated to account for different failure modes.. The models were used to determine the
progressive collapse of a 10 storey steel frame designed for moderate and high seismic risk.
The analysis results showed better performance of system designed for high risk as
compared to that designed for moderate risk. The better performance was attributed to a
large extent to layout and system strength rather than the influence of improved ductile
detailing. However, the proposed method does not provide information about the reserve

233

capacity of the system and so its results should be carefully evaluated. The authors
extended their study to eccentrically braced steel frames and observed similar results (as in
their previous study) for these frames as well.
Izzuddin et al. (2008) proposed a new methodology to determine the collapse capacity of
building systems considering the effect of sudden column loss. The finite element modeling
of building frames was adopted to achieve this purpose. The proposed method was capable
of incorporating both simplified as well as complex modeling approaches. Moreover, the
proposed assessment framework employed three stages namely (i) determination of the
nonlinear static response, (ii) simplified dynamic assessment, and (iii) ductility assessment.
The analysis results showed that parameters advocated in previous research works like
energy absorption capacity, redundancy and ductility were not individually suitable as
measures of structural robustness. However, the system pseudo-static capacity was
identified as a new and rational measure of building robustness under sudden column loss
scenarios as compared to other parameters. Finally, the proposed methodology was
observed to be efficient in determining the collapse capacity of multistorey building frames.
Feng fu (2009) performed experimental and analytical study to estimate the collapse
capacity of multistorey generic frames. For analytical study, the authors adopted the finite
element modeling for determining the collapse capacity of a 20 storey frame. The finite
element modeling incorporated the material characteristics and non-linear geometric
behavior. In determining the collapse capacity, the structural behavior of the building under
the sudden loss of columns for different structural systems and for different scenarios of
column removal were included. The analysis results showed that the adopted finite element
models accurately displayed the overall behavior of the 20 storey buildings under the
sudden loss of columns which provided important information for the additional design
guidance on progressive collapse. The analytical results were observed to be in close
agreement with the experimental results.
Li et al. (2011) proposed improved tie force based method to determine the collapse
capacity of the building structures. In the present study, the progressive collapse was
referred as the local damage due to occasional and abnormal loads which in turn leads to
the development of a chain reaction mechanism leading to progressive and catastrophic
failure. The tie force (TF) method is a design techniques for resisting progressive collapse in
which a statically indeterminate structure is idealized in form of a locally simplified
determinate structure by assumed failure mode. A numerical study on two reinforced
concrete (RC) frame structures (modeled using FEM approach) was performed to show that

234

the current TF method was inadequate in increasing the progressive collapse resistance.
The previous TF method ignored important factors as load redistribution in three
dimensions, dynamic effect, and internal force correction. As such, an improved TF method
was proposed in this study. The applicability and reliability of the proposed method was
verified through numerical design examples.
Baradaran Shoraka et al. (2012) proposed new analytical models and performed collapse
analysis using experimental and analytical study. The researchers performed analytical
modeling using FEM approach of a seven-storey non-ductile reinforced concrete frame
building located in Los Angeles. The results of the study showed close agreement between
analytical and experimental formulations.
(d) Incremental Dynamic analysis
In recent years, the incremental dynamic analysis has emerged as a powerful tool to study
the overall seismic behavior of structures and in determining the overall global collapse of
the structure (FEMA 356). In the process of incremental dynamic analysis, a series of
nonlinear dynamic analyses were performed in which the intensity of the ground motion
selected for the collapse determination was incrementally increased until the global collapse
capacity of the structure has been reached. Also, ground motion intensity is plotted (spectral
acceleration at the fundamental natural period of the structure) against a seismic response
parameter (maximum roof displacement, maximum inter-storey drift ratio), and such a plot is
called as collapse fragility curve. The structure was assumed to collapse globally when the
curve becomes flat. It is the stage when a large increase in the structural response was
generated by a small increase in the ground motion intensity. As different ground motions
may have different frequency contents, durations, intensity; this process was repeated under
different ground motions to obtain a realistic average of the collapse capacity measure.
Incremental dynamic analyses started in early 1977 with Bertero (1980) who extensively
studied this method and investigated later in other research studies (Vamvatsikos and
Cornell 2002). Vamvatsikos and Cornell 2002 adopted incremental dynamic analysis to
determine the intensity-response curves for a 5 storey braced steel frame, a 3 storey
moment resisting frame and a 20 storey moment-resisting steel frames. The researchers
proposed simplified techniques to efficiently carryout an incremental dynamic analysis and
emphasized on formation of a response database which contained the information obtained
from the different curves (obtained by applying different ground motions). The authors
observed that incremental dynamic analysis was a valuable tool that simultaneously

235

considered both seismic demands and their global capacities. Also, some important
properties of the fragility curves like non-monotonic behavior, discontinuities, multiple
collapse capacities, and their extreme variability from ground motion to ground motion were
considered in the analytical study. The analysis results showed that a complete incremental
dynamic analysis required an intense computational effort. Improving their previous
approach, the authors (Vamvatsikos and Cornell 2004) proposed a simplified method to
perform IDA efficiently. The applicability of the proposed procedure (with aid of a detailed
example of 9-storey steel moment-resisting frame) was demonstrated by the researchers.
The relationship between an incremental dynamic analysis and a static pushover analysis
was explored. Moreover, the researchers developed (Vamvatsikos and Cornell, 2005) a
simplified method to evaluate the seismic response and collapse capacities of MDOF
systems using SDOF systems.
Lee and Foutch (2002) evaluated the performance of 20 steel frame buildings designed
as per 1997 NEHRP provisions (FEMA 1998). This building model was subjected to an
ensemble of 20 earthquake ground motions using non-linear time history analysis for
comparison of collapse probabilities. To determine the collapse capacity, the ground
motions were scaled such that they represent 2% probability of exceedence in 50 years.
Several important aspects like beam-column joint ductility, panel zone deformation,
influence of interior gravity frames were considered in modeling the frames. The beamcolumn joints were characterized by a gradual strength degradation after a rotation of 0.03
radians The collapse probability was estimated in terms of maximum storey drift demands
and both local and global drift capacities were considered in this comparison. The local drift
capacities used in the analytical study were defined in terms of maximum angular drift that
the beam-column joints can sustain before loss of gravity load carrying capacity. The global
drift capacities were obtained by performing an incremental dynamic analysis for each of the
buildings and plotting the corresponding seismic response parameter (maximum storey drift
ratio) versus spectral acceleration curves. As discussed

previously, the building was

considered to collapse when the collapse fragility curve becomes flat or at a point at which
the maximum storey drift ratio at which this curve reached a slope equal to 20% of the
slope in the elastic region of the curve. However, if the prescribed slope limit was not
violated before a particular limit (before a storey drift ratio of 0.10 is attained), then the
global storey drift capacity was assumed equal to 0.10. The results of the analytical study
showed the satisfactory collapse performance of the buildings.

236

Ibarra and Krawinkler (2004) proposed a simplified methodology to estimate the global
collapse capacity of deteriorating framed building structures subjected to the ground motion
excitation. The proposed methodology advocated the use of a relative intensity measure
[Sa (T1)/g].
However, the proposed intensity measure was observed to be equivalent to the
reduction factor used in building codes for the analysis of yielding structures and for
structures without any over-strength. Similar to previous approaches (Vamvatsikos and
Cornell 2004), the deteriorating hysteresis models were used in the proposed methodology
to incorporate the cyclic behavior of the structural components with large inelastic
deformations. These deterioration models represented majority of the deterioration modes
as observed in the experimental tests. To determine the seismic collapse capacity, the
intensity measure was increased until the point at which the collapse fragility curve (intensity
measure versus normalized maximum roof drift curve) became flat. The relative intensity
measure at this point represented the collapse capacity. In the proposed methodology, the
factors like the uncertainties in the frequency content of ground motions and the
deterioration characteristics of the structural elements were considered. The computer
program Drain -2DX was used by the authors to determine the collapse capacity.
Following the same footsteps, Prakash et al. (1993), Ibarra and Krawinkler (2005) used
the proposed methodology to determine the effect of several parameters on seismic
collapse capacity for different types of frame structures. Collapse fragility curves were
developed to determine the mean annual collapse capacity. The building models considered
in the parametric study were framed structures with stiff and flexible single-bay frames with
3, 6, 9, 12, 15, and 18 storeys with formation of plastic hinges at beam ends and the base of
the columns. The results of the analytical study showed that, the slope of the post-yield
softening branch (moment rotation relationship of the yielding members) and the
displacement (at which this softening begins) to be the dominating factors in determining the
seismic collapse capacity. In addition, the cyclic deterioration and ground motion duration
factors were observed to be less dominant factors in the collapse of structures. This
observation was contrary to findings of Takizawa and Jennings (1980), who observed
significant contribution of ground motion duration on seismic collapse capacity.
Adopting a similar approach, Ayoub et al. (2004) evaluated the influence of stiffness and
strength degradation on the seismic collapse capacity. The authors conducted incremental
dynamic analysis on a SDOF structure (with T = 1.0 s). For the analytical study, three
degrading constitutive hysteresis models (a bilinear model, a modified Clough model, a

237

pinching model) accounting for the possibility of collapse were considered and consequently
collapse fragility curves for this building system were plotted. The collapse was assumed at
a point at which the building systems strength reduced to zero. The energy criterion was
defined on the basis of strength softening and stiffness and strength degradation. For
analytical study the building models were subjected to an ensemble of 80 ground motions.
Results of analytical study showed that collapse probability increased with increase in
stiffness and strength degradation. Therefore, it is necessary to adopt hysteresis models for
collapse capacity assessment.
Ibarra et al. (2005) evaluated the influence of stiffness and strength degradation on the
seismic demands of structures at vicinity of collapse. Ibarra et al. (2005) developed simple
hysteresis models to represent stiffness and strength deterioration properties, and calibrated
the modes of strength and stiffness deterioration by utilizing experimental data from tests on
steel, plywood, and reinforced concrete components. The proposed models and some
previous models were used to determine the seismic response of a SDOF system (natural
period of 0.9 s and a damping ratio of 5%) subjected to an ensemble of 40 ground motions.
The ground motions were scaled to various intensity levels and demand - intensity curves
were developed to determine the collapse capacity of the system. The authors based on
their analytical study concluded that the strength and stiffness deterioration had a least
influence on seismic response near collapse of the structure.
Haselton et al. (2011 a) with aid on incremental dynamic analysis determined the seismic
collapse capacity of multi-storey frames without structural irregularity by considering the
sidesway collapse mechanism using dynamic analysis and observed collapse capacity to
decrease with the storey height. Moreover, it was observed sidesway collapse was one of
the major mechanisms contributing towards collapse. In continuation with their previous
studies, the authors included vertical collapse mechanisms along with side-sway
mechanisms in determining the collapse capacity (Haselton et al. 2011b). The observations
suggested that the consideration of vertical collapse mechanisms reduced the overall
collapse capacity of the structure which was an important observation as these mechanisms
were neglected in majority of the previous research works (Haselton and Deierlein 2007;
Haselton et al. 2011a).
(e) Shake table experiments to determine the collapse capacity
The review of previous literature works reveals that a very few experiments were
conducted in which structural models were tested until the collapse stage. Kato et al. (1973)

238

conducted shake table tests on simple models until progression of collapse state by using
the shake table. A comparative study between experimental and analytical results
considering strain hardening and P - Delta effects was conducted. The structural models
consisted of 15 cm high H steel columns fixed at both ends with a concentrated mass on top
of the columns. On basis of their studies, it was concluded that the test results are
accurately predicted by the analytical studies except for some softening of the hysteresis
loops due to Bauschinger effect.
Vian and Bruneau (2003) conducted experimental study on 15 simple specimens built
with four steel columns which were connected to a rigid mass. The building models were
subjected to the ground motion intensity of the 1940 El Centro ground acceleration record.
The ground motion intensity was progressively increased till the collapse fragility curve
became flat. From the analytical study results it was observed that the inelastic behavior of
the specimens showed a high dependence on the traditional stability factor defined for a
single storey structure as P//K0H. Specimens with a stability factor of less than 0.1 sustained
the ground motions with larger ductility demands and accumulated drifts than those with a
stability factor of more than 0.1. Moreover, the analytical study exhibited a good correlation
with the experimental data.
Kanvinde (2003) conducted experimental study on 19 simple structures with a storey
height of 254 m, a bay width of 610 mm and a floor plan of 305 mm x 610 mm up-to the
collapse state. The structures were built with four steel flat columns connected to a rigid
1.425 kN 320 lb steel mass. They were tested using shake table under two of the ground
motions recorded during the 1994 Northridge earthquake: Obregon Park and Pacoima Dam.
The specimens collapsed after the formation of plastic hinges were formed at the top and
bottom of the columns and consequently a storey mechanism was formed. Nonlinear
dynamic analysis of the specimens were conducted to evaluate the ability of analytical tools
to predict their response under very large displacements was conducted. The hysteresis
model used was characterized by a yield envelope and a nonlinear hardening exponential
law. The parameters for this model were determined from monotonic and cyclic tests of the
columns used to build the experimental models. By comparing the displacement time
histories obtained analytically and experimentally it was observed that the analytical
simulations predicted results with an average error of about 15% as compared to the results
obtained from the shake table tests.

239

Similarly, Elwood and Moehle (2003) conducted shake table tests of two one-half
scale reinforced concrete plane frames to investigate the mechanisms that led to the
seismic collapse of reinforced concrete frames with non-ductile or low-ductility columns.
The objective of these tests was to determine the loss of column axial capacity due to
column shear failure. The two test specimens used in the study consisted of three columns
fixed at their base and interconnected by a beam at their upper end. The only difference
between the two specimens was the magnitude of the axial force applied to the center
column. This center column was designed with widely spaced transverse reinforcement to
make it vulnerable to a shear failure and a subsequent axial failure during the tests. A
scaled version of a ground motion recorded during the 1985 Chile earthquake was used as
input. The measured response of the test specimens were compared with results from a
nonlinear dynamic analysis of an analytical model of the shake table specimens. This
analytical model was formulated with a proposed column element incorporating empirical
models that predicted the column drifts at the time of shear and axial load failures. These
empirical models were developed to evaluate the influence of such failures in a building
frame analysis. It was observed that he analytical model provided a good estimate of the
measured drifts up to the point of shear failure but failed to capture the large displacements
that occurred after that. Because of the underestimation of the column drifts, it also failed to
detect the axial failure of the center column.
Although not precisely a shake table experiment, it is also worthwhile describing the
full-scale experiment conducted by Nakashima et al. (2006). A full scale steel moment
resisting frame up-to the collapse state was tested. The frame had three storeys, two bays in
the longitudinal direction, and one bay in the transverse direction. The plan dimensions were
12m by 8.25m and the total height was 8.5 m. The columns and beams were made with
cold-formed square tubes and hot-rolled wide flange sections, respectively. The floors were
formed with metal deck sheets and cast-in-situ reinforced concrete. The horizontal loading
was applied by two jacks placed at the center of the third floor along the frames longitudinal
direction. The loading protocol consists of displacement amplitudes that produce overall drift
angles (horizontal displacements at the loading point over frame height of 1/ 200, 1/ 100, 1/
75, 1/ 50, 1/ 20, and 1/15 radians, with each amplitude repeated two or three times). In this
test, the authors observed that in the last portion of the loading the first-storey shear force
resistance decreased significantly owing to column local buckling, plastic elongation of the
anchor bolts at the base of the columns and crushing of the concrete underneath the column

240

base plates. Finally, local buckling occurred at the top of the columns of this storey which
led to development of a storey collapse mechanism.
The series of numerical simulations were performed with an analytical model of the
tested frame to investigate the correlation between analytical and experimental study. The
analytical model consisted of the beams and columns represented with elastic beam
elements and bilinear rotational springs at the ends of the elements. A rotational spring was
inserted at the bottom of the first-storey columns to account for the flexibility of the column
base plates and anchors. P - Delta effects and the deformation of the panel zones at the
beam-column joints were considered. The authors observed that with the proper adjustment
of the basic material properties, strain hardening after yielding, and the composite action
between the steel beams and the reinforced concrete floor slabs, the numerical simulations
were capable of accurately predicting the observed cyclic behavior of the frame up to a drift
angle of 1/25 radians. Because the simple analytical model used cannot account for the
deterioration that takes place in the structural elements at large drifts, the numerical
simulations fail to reproduce the experimental results in the last portion of the loading.
Donatello Cardone (2007) conducted shake table tests on a 3-storey, 2-bay, RC frame.
The shake table tests were compared with the pushover analysis methods. Four types of
pushover analysis were conducted namely
a) Capacity Spectrum Method (CSM).
(b)Displacement Coefficient Method (DCM), presented in FEMA - 273 and further developed
in FEMA 356.
(c) N2 Method, implemented in the EC 8:2004.
The pushover analyses are conducted using DRAIN -2DX software using four different
lateral force distributions (uniform, triangular, modal-proportional, and multimodal fully
adaptive). In the numerical model used, RC members were modeled as fiber elements. The
analytical results were compared with the experimental results predictions for two similar
1:3.3-scale structural models, with and without infilled masonry panels, respectively. The
comparison was made in terms of maximum storey displacements, interstorey drifts, and
shear forces. The results showed that all non linear static method results were found to be
in close agreement with experimental results. In addition, the lateral load pattern was found
to slightly affect the accuracy of the results for the three-storey model considered, even if
collapse occurs with a soft storey mechanism.
Elwood and Moehle (2008) determined the behavior of two half-scale, one-storey frames
with axial loads . The frames were subjected to unidirectional simulated earthquake motion

241

which was applied at the base of the structure. The observation results suggested that the
shear failures of an interior column led to axial-load failure and redistribution of internal
forces to adjacent framing components.
Su et al. (2009) conducted shake table tests on a 1/35 prototype building model of a
supertalll building (260m in height) designed as per Chinese code. The building was a
complex structure system and consisted of a steel reinforced concrete core, a belt truss
storey and a number of SRC columns to resist vertical and lateral loads. In addition, the
tower had two setbacks in elevation and two sets of inclined columns which created
irregularity and complexity in the building system. The prototype structure was analysed for
dynamic properties, acceleration and displacement responses of the model structure, the
failure mode and the dynamic responses of the modelstructure and the prototype structure
were investigated. The shake table results showed adequate seismic performance of the
building system.
Lu et al. (2012) determined the seismic behaviour of a 53-storey tower with the height of
250m. This supertall building was composed of a reinforced core, two trusses (a
strengthened storey at 20th 21st floor and a high-level transfer storey at 37th 38th floor) and
eight composite (steel-encased concrete) mega-columns below the high-level transfer storey
in the exterior perimeter of the building. To conduct the experimental study, a prototype 1/30
scaled model of the building was prepared. The main aim was to study its dynamic
characteristics and to evaluate its earthquake resistant capacity. The model test results
indicated satisfactory seismic performance of the model under the action of earthquake of
intensity 7. In addition, a 3D finite element analysis of model structure was performed and
close agreement between experimental and analytical results was observed.

5.2.2 Main conclusions derived from the literature review


The review of previous literature works showed that the currently available methods to
assess the collapse capacity of a structure exhibited a poor performance in estimating the
collapse capacity. Many researches used SDOF models and pushover analysis to estimate
the collapse capacity. This approach was observed to be unreliable as collapse capacity of
a structure strongly depended on the assumed shape of the failure mechanism which
cannot be accurately predicted with the help of pushover analysis [Bernal (1992, 1998)].
Nonlinear static methods were also unreliable as they lack a theoretical foundation and were
based on unrealistic assumptions and neglected the dynamic response of building systems.
The methods based on finite-element models were tedious and required consideration of

242

large number of factors as discussed earlier. Likewise, the methods based on an


incremental dynamic analysis were computationally demanding but were reliable. However,
consideration of a large number of ground motions to obtain the better statistical averages
with a high level of confidence is quite necessary. Krawinkler et al. (2003) searched for
alternative intensity measures to reduce such dispersion. In case of methods based on
incremental dynamic analysis, the results may also depend on the selection of the
parameter used to measure the ground motion intensity and spectral acceleration at the
fundamental period of the structure.
In nutshell; it was observed that
(1) The reliability of collapse assessment methods is affected by the accuracy and
convergence problems which are likely to occur due to the nonlinear behavior and large
displacements that a structure may experience near collapse state
(2) The accuracy of the available collapse assessment methods has not been verified
through experimental or field studies
(3) Very few shake table experiments have been carried out to study the collapse of
structures and verify the adequacy of collapse assessment methods
(4) Very few researchers tested the models of real structures for collapse assessment under
earthquake excitations
(5) A very little knowledge regarding the inherent safety margin against collapse in codedesigned structures (Maison and Bonowitz 2004) was obtained
(6) Aspect of structural irregularity and cracking phenomenon have been ignored by the
previous research works in analytical as well as experimental study. Therefore, new
methodology to estimate collapse capacity for these structures is required.
Thus, further research is required before the collapse capacities of structures and the
associated safety margin against collapse may be evaluated with confidence. There is a
need for experiments with realistic full-scale specimens, subassemblies and whole
structures in which the specimens are tested all the way to collapse. These experiments are
needed to
(i) Improve the understanding of the conditions that lead to a collapse in real, full-scale,
three-dimensional structures.
(ii) Evaluate the capability of existing numerical analysis techniques to predict building
behavior at the deformation levels involved when a structure is near collapse.
(iii) Assess the adequacy of current collapse assessment methods.

243

(iv) Generate data on the hysteretic behavior of structural components as they approach the
failure state.
(v) Calibrate simplified analysis models.
There is also a need to evaluate the inherent safety margin in current seismic provisions
against a structural collapse and establish whether or not this safety margin is adequate
enough. Ultimately, the great challenge for the profession is the development of simplified
techniques that correlate well with experiments and advanced methodologies to estimate in
a reliable way the aforementioned collapse capacities and safety margin. These techniques
are urgently needed to facilitate
(i) Scrutiny of present and future code provisions in regard to their ability to provide an
adequate safety margin against collapse
(ii) Development of design procedures would explicitly make structures resist a collapse with
a specified safety margin
(iii) Identification and strengthening of weak structural components that may compromise the
collapse capacity of a structure. Undoubtedly, the availability of such techniques could help
to improve the seismic performance of building structures and minimize the number of
catastrophic failures during earthquakes.

5.3 Performance based earthquake engineering for collapse assessment


In recent years the performance based design methodology has developed rapidly and its
implementation has been extended to estimation of collapse capacity of the buildings. This
methodology is developed by PEER (Pacific earthquake engineering research center) and is
presented in detail in previous research works (Calvi et al. 2002; Miranda and Akkar 2003;
Haselton and Deierlein 2007; Sasani and Kropelneki 2007; Liel et al. 2008; Perus et al.
2008; Zarein and Krawinkler 2009 etc.). Performance based earthquake engineering
predicts the seismic damage probabilities through combination of ground shaking hazard
and seismic response obtained through non-linear dynamic analysis. The ground motion
hazard can be typically represented by a hazard curve that represents exceedence of
ground motion intensity above a particular limit at a particular site. The structural response is
predicted in terms of the engineering demand parameters like maximum roof displacement,
inter-storey drift ratio, rotation ratio etc. These parameters can be used to compute the
seismic damage index or decision variables that describe the magnitude of repairs and
rehabilitation required after occurrence of an earthquake.

244

5.4 Modeling of RC frames


Analytical model used to capture the key collapse modes requires the accurate modeling of
deterioration associated with seismic loading of the constituent elements such that all
possible local and global collapse modes are captured. The main modes of element
deterioration are concerned with the design and detailing requirements used in the structural
design. After identification of all collapse modes, the main likely collapse modes are
identified using the experimental tests, engineering judgment, past earthquake damage
experience. Then on the basis of the identified collapse modes, appropriate damage models
and simulation were selected to predict collapse.

5.4.1 Element deterioration modes


The main modes of element deterioration of RC frame components were identified from
review of the past literature works. The deterioration modes were basically classified into
five categories based on the type of structural element and physical behavior associated
with deterioration ( Figures 5.1 to Figure 5.14 and i Table 5.1 and 5.2.)

Table 5.1 Different modes of collapse of structural members

Reasons for collapse

CO

CME

DU

NDU

SD

S1

HO

HO

Beam and column flexural hinging

S2

MO

HO

Column hinging resulting in a soft storey mechanism

S3

LO

HO

Beam and column flexure shear failure

S4

LO

HO

Joint shear failure due to hinging of beams and columns

S5

LO

HO

Pulling out of reinforcing bars or splicing of columns

V1

LO

MO

Axial collapse of column due to column shear failure

V2

LO

HO

Axial collapse of column due to column flexure shear failure

V3

LO

MO

Punching shear failure of columns leading to slab collapse

V4

LO

MO

Column instability leading to diaphragm failure

V5

LOMO

MO-HO

VE

Overturning moments leading to column axial collapse

where CO - Collapse type, CME Collapse mechanism, DU Ductile frame, NDU Non
ductile frame, HO - High, MO Medium LO Low, SD Sidesway, VE Vertical

245

5.4.2 Local and global collapse


Structural collapse occurs when the seismic excitation results in ground shaking which in
turn results in element deterioration modes that combine to form a collapse mechanism. For
RC frame structures, the possible collapse scenarios and contributing element modes are
identified and shown in Table 5.2. For RC frame structures, the design and detailing
requirements mainly govern the mode of collapse, and different building codes (ACI 2002;
ASCE 2002; IBC 2003) suggest provisions that prefer ductile collapse of the structure. The
strong column weak beam philosophy as propagated by different seismic design codes
recommends the formation of flexural hinges first in the beams and then in the columns.
However, it does not eliminate the plastic deformations in the columns completely (Ibarra
2003). Likewise, the shear strength design provisions adopted by design codes are provided
to avoid shear failure in beam-column elements. However, the older RC structures contain
minimum transverse reinforcement which makes them vulnerable to wider range of the
collapse modes (Aycardi et al., 1994; Kurama et al., 1994; Kunnath et al. 1995; El-Attar et
al. 1997; Filiatrault et al. 1998). The structures reported in the research works have shown a
tendency to fail in soft storey due to column hinging mechanisms. Moreover shear failure of
columns mainly depend upon columns design and gravity loading, and columns may fail
due to lap splice failure, pull out of bottom reinforcing bars due to less stringent detailing
requirements. Nevertheless, seismic performance of the structure can also be enhanced by
use of seismic retrofitting measures and supplemental damping devices (Calvi 1997; Tan
and Balendra 2007; Jangid 2000; Kori and Jangid 2008; Patil and Jangid 2011a,b; Satish
Kumar et al. 2002, 2012; Ramaraju et al. 2013).These element deterioration modes are
effectively captured by Ibarras model as discussed in the next section.
Table 5.2 Collapse types of different types of frames
Type of collapse
A

mechanism

Reasons for collapse


Combination of collapse modes S1, S2, S3, S4, S5 and V2

Due to V5 collapse mode

Due to S3,V1 and V2 collapse mode.

Due to S4 collapse mode

Due to S5 collapse mode

Due to V3 collapse mode

246

This page is intentionally left blank

247

Figure 5.1 Flexural hinging at top and bottom of columns at a Mosque in Banda Aceh
during Turkey earthquake (Ghobarah et al. 2006)

Figure 5.2 Punching shear failure of flat plate construction (Arslan and Korkmaz 2007)

248

This page is intentionally left blank

249

Figure 5.3 Column shear failure in five storey office building (Ghobarah et al. 2006)

Figure 5.4 Shear failure of column (Arslan and Korkmaz 2007)

Figure 5.5 Column shear failure due to buckling of longitudinal bars during Turkish
earthquakes (Arslan and Korkmaz 2007)

250

This page is intentionally left blank

251

Figure 5.6 Column shear failure due to Wide spacing of lateral ties Turkish
earthquakes (Arslan and Korkmaz 2007)

Figure 5.7 Shear failure of rectangular column and square column during L Aquilla
earthquake 2009 (Ricci et al. 2011b)

Figure 5.8 (a) Shear failure of beam column joint during Turkish earthquakes (Arslan
and Korkmaz 2007)

252

This page is intentionally left blank

253

Figure 5.9 Failure of beam column joint and undamaged joint Turkish earthquakes
(Arslan and Korkmaz 2007)

Figure 5.10 Anchorage pull out failure of lapped reinforcement (Kam et al. 2010)

a)

b)

Figure 5.11: (a) Inadequate lap splice, (b) Beam column joint failure due to lack of
transverse reinforcement at a office building in Banda Aceh (Arsalan and Korkmaz
2007)

254

This page is intentionally left blank

255

Figure 5.12 Joint failure with evident (a) longitudinal bar buckling, (b) diagonal
cracking failure in concrete joint (c,d) Failure mechanisms in joint column interface
(Ricci et al.2011b)

Figure 5.13 Shear - axial failure of column during Darefield earthquake ( Kam et al.
2010)

256

This page is intentionally left blank

257

(a)

(b)

(c)

(d)

(e)

(f)

(a) Flexural hinging of beam column elements Managua earthquake, (b) Column
compressive failure (Chi chi earthquake 1999) (c) Beam column shear failure Miyagi
earthquake 2003, (d) Joint shear failure Chi Chi earthquake 1999, (e) Pull out and
rebar connections (Chi Chi earthquake 1999), (f) Slab column connection failure
during Northridge earthquake 1994
Figure 5.14 Different modes of collapse [NISEE, UCSD Structural Engineering and
Earthquake Disaster-Research Laboratory (Tohoku Univ.)]

258

5.5 Modeling for collapse assessment


The collapse capacity of the structure is determined by inelastic modeling of beam-column
elements in RC frame structures. The inelastic behavior is modeled in the form of plastic
hinges in beam-column elements that utilize a hysteresis force deformation relationship.
Review of previous research work shows that the researchers like Ibarra et al. (2005),
Medina and Krawinkler (2005) have proposed hysteresis models to estimate the collapse
capacity. The hysteresis model proposed by Ibarra (Figure 5.15) is composed of a trilinear
monotonic backbone with a set of hysteresis rule that capture most of the failure modes
encountered in the experimental studies. This model considers important aspects like
concrete crushing, rebar buckling and fracture, bond failure. The cyclic deterioration adopted
in the model was based on two parameters namely normalized energy dissipation capacity
and cyclic deterioration parameter which depicts rate of cyclic deterioration. Therefore,
cyclic deterioration is based on several parameters to control cyclic and monotonic loading,
and these factors are shown in Figure 5.15. The negative stiffness branch of post-peak
response is an important aspect of this model and it enables modeling of strain-softening
behavior associated with phenomenon like rebar buckling and fracture, concrete crushing,
bond failure etc. The Ibarras model also captures five basic modes associated with cyclic
deterioration as mentioned below
(a) Strength deterioration of the post-peak strain-softening branch
(b) Strength deterioration of the inelastic strain-hardening branch
(c) Unloading stiffness deterioration,
(d) Accelerated reloading stiffness deterioration
(e) Additional reloading stiffness deterioration is automatically incorporated through the
peak-oriented cyclic response rules.
This element model proposed by Ibarra requires the specification of seven
parameters to control both cyclic and monotonic behavior of the model: My, y, Ks, cap, and

Kc, , and c2. However, this model should be calibrated before application to determine the
collapse capacity, and to empirically determine stiffness, capping (peak) point, post-peak
unloading stiffness and hysteretic stiffness/strength deterioration for reinforced concrete
beam-column elements to be used in collapse simulation of RC frames. To determine the
values for these parameters, Ibarra et al. (2005) conducted 255 experimental tests on RC
columns and based on the regression analysis conducted on the analysis results, empirical
equations for these plasticity models were proposed by Ibarra et al.(2005). The detailed
report on calibration study is available in Haselton and Deierlein (2007).

259

5.5.1 Calibration procedure


The parameters representing Ibarras collapse model are calibrated using experimental
results of rectangular columns included in PEER Database (Berry et al. 2004). The
database includes RC columns with ductile and non-ductile detailing with variation in levels
of axial loads and geometries. Haselton and Deierlein (2007) carried out carried out
calibration of Ibarras model by performing experimental study on 255 columns to cover
different failure modes as discussed in previous sections of this Chapter. Based on this
experimental study, the researchers have proposed set of equations to estimate the
parameters associated with Ibarras model. The proposed equations provided a link
between the column design parameters such as axial load, modeling parameters,
transverse reinforcement etc. The model parameters were predicted by these equations it in
terms of its mean value. These parameters with their definition and equation predicting them
(proposed by Haselton and Deierlein 2007) have been discussed in the next section.

Figure 5.15 Ibarras nonlinear hinge model (Ibarra et al.2005)


where Kc, Ke1, KS are Post-capping stiffness for use in Ibarra element model , Secant
stiffness through the yield point, for use in Ibarra element model , Hardening stiffness for
use in Ibarra element model, i.e. stiffness between y and cap,pl..
(a) Effective stiffness and flexural strength
In general, the stiffness of the RC members changes after cracking and it depended on the
load and the deformation level. Two types of effective stiffness are defined in this section.

260

(i) Secant stiffness at yield point of the component


(ii) Secant stiffness at 40% yield point of the component
The stiffness was presented as fraction of gross stiffness of the section and accounts for
major modes of deformation including flexure, shear and bond slip. The equation for secant
stiffness to yield depends upon axial load ratio and shear span ratio and was expressed as
P
L
0.07 0.59 a ' 0.07 s
EI g
H
Ag f c
EI y

0.35

EI y
EI g

0.8 0.2

EI y
EI g

0.6

(5.1)

(5.2)

where Pa is the axial load, Ag is the cross sectional area of column, Ls/H is the shear span
ratio, fc is the concrete compressive strength
The uncertainty in prediction is given by logarithmic standard deviation (LN = 0.28), and the
correlation coefficient is observed as 0.80. The lower and upper limit of stiffness is specified
because there is a limited data available on the column analysis.
As per the second definition, the effective stiffness was defined as the secant stiffness
corresponding to the 40 % of the yield force of a RC column was predicted by equation 5.3
(proposed by Ibarra et al. 2005).
EI y
P
L
0.8 .
0.02 0.98 a ' 0.09 s , 0.35
EI g
EI g
H
Ag f c
EI y

(5.3)

From comparison of both equations, it was observed that equation 5.3 yielded over conservative results of stiffness as compared to equation 5.2 (Liel 2008).
FEMA 356 permits the use of the prescribed equations based on the level of axial load
in columns: 0.5EIg when v < 0.3 and 0.7 EIg when v > 0.3 (ASCE 2000).The stiffness values
predicted by FEMA 356 were observed to be higher as compared to the values predicted by
equations used in the present study. This is due to the fact that FEMA 356 only includes
flexural deformation neglecting bond slip deformation. Elwood and Moehle (2006) proposed
equation for effective stiffness that includes all components of deformation (flexure, shear,
bond slip) where effective stiffness is defined as secant stiffness at yield point of component.
Elwood and Moehle (2007) proposed improved equations by adopting the Ls/H term
(Proposed by Liel 2008) to reduce the prediction uncertainty. The flexural strength adopted
in Ibarras model was calculated using the relation proposed by Panagiotakos and Fardis
(2001) showed good agreement with the experimental data as reported by Haselton and
Deierlein (2007).

261

(b) Plastic-Rotation capacity (cap pl)


The plastic rotation capacity of a structural member mainly depends upon the axial load ratio
(v = Pa/Agfc) and area of transverse reinforcement. The parameters like concrete
compressive strength (fc),area of longitudinal reinforcement ratio in tension and compression
(p and p), area ratio of transverse reinforcement (psh), as1 (accounts for slip and is taken
equal to unity), and rebar buckling coefficient also have a significant impact on the plastic
rotation capacity. The parameter

cap

indicated whether reinforcing bar slip was included in

the equation to compute the plastic rotation capacity. This parameter assumed a value of
unity for all the columns. However, this parameter was included in the equation because
bond slip was absent in some of the experimental configurations. The equation to compute
plastic rotation ratio was proposed by Ibarra et al.(2005) as

cap, pl 0.12(1 0.55as1 )(0.16)v (0.02 40sh )0.43 (0.54)0.01 f ' c (0.66)0.1s (2.27)10
n

(5.4)

where as1 is the parameter indicating slip, v is the axial load ratio, sh is the transverse
reinforcement ratio in reinforced concrete column, Sn is the rebar buckling coefficient and is
calculated by expression Sn = (s/db)(fy/100)0.5, s is stirrup spacing and dc is depth of column
section, measured as centerline-to-centerline distance between stirrups.
Most of the input parameter in equation 5.4 were unit-less except fc which was expressed in
Mpa. For the structural members with reinforcement asymmetrically arranged the plastic
rotation capacity in equation 5.4. should be multiplied by a term proposed by Fardis and
Biskinis (2003). This term accounted for ratio of areas between compressive and tensile
steels and was expressed as

' fy

max(0.01
fc

fy

max(0.01
fc

0.225

(5.5)

where p is the ratio of longitudinal reinforcement in tension to that in compression. Equation


5.5 did not differentiate between seismic (135 degrees) and non seismic hooks (90 degrees)
in transverse reinforcement. Since structures with stirrups have non seismic hooks more
often and this effect is included in psh term in equation 5.4. Non ductile structural members
are expected to have non-seismic hooks.
It is worthwhile to compare the predicted rotation capacity and ultimate rotation capacity
predicted by Fardis were based on study of 700 different RC elements including beams,
columns and walls (Panagiotakos and Fardis 2001b; Fardis and Biskinis 2003). The

262

equation 5.5 determined the capping point but the Fardis equation was based on the
ultimate point (Point at which 20 % strength loss is assumed). By calculating the plastic
rotation capacity using equation proposed by Fardis and comparing with equation 5.4
showed that plastic rotation capacity was under-predicted . This correlates with the results of
FEMA 356 and ASCE 41 which over predicted the plastic rotation capacity as compared to
the equation 5.5.
(c) Post-yield hardening strength ratio
Regression analysis conducted on the experimental test results of columns showed that the
parameters like axial load ratio and concrete strength have an insignificant impact on the
post yield hardening strength ratio and a constant value was recommended for this factor.

(d) Pre-capping rotation capacity (pc)


The equation proposed for estimating pre-capping rotation capacity was presented as

pc 0.76(0.031)v (0.02 40sh )1.02 0.10

(5.6)

where v is the axial load level, psh is the area of transverse reinforcement. This equation
yielded better results for non-ductile columns since it was proposed based on experimental
tests conducted on non-ductile columns.
(e) Cyclic energy dissipation capacity ()
The analysis results carried by Haselton and Deierlein 2007 showed that the cyclic energy
deterioration factor depended upon several parameters like axial load ratio (v), spacing of
transverse reinforcement (sa1/da1) Based on the regression analysis conducted by the
authors, the equation to compute cyclic deterioration parameters is proposed as

(170.7)(0.27)v (0.10)sa / da1

(5.7)

where Sa1 is the stirrup spacing and da1 is the column depth.

5.6 Modeling uncertainties for collapse assessment


To determine the collapse capacity and collapse probability of a structure, a robust
analytical model that captures the non-linear behavior and takes into account different
sources of uncertainty was required. From previous research works, it could be clearly
observed that ground motion was the major source of uncertainty. The uncertainty in the

263

ground motion is represented by site specific hazard which related the spectral intensity to
exceedence frequency. In addition to the ground motion uncertainty, other uncertainty
issues related to stimulation of structural response (concerned with the analysis method
adopted) and the extent to which the real structure was represented by the analytical model.
In cases where non-linear time history analysis was used the uncertainty in the prediction
capacity of the equations proposed was affected by parameters like strength, stiffness,
deformation capacity and energy dissipation capacity. Moreover, the uncertainty was also
related to type of model used (2D or a 3D model).
The present study involves probabilistic assessment of collapse risk through non-linear
time history analysis which incorporated uncertainties associated with ground motion and
structural modeling. The past research works indicated that the modeling uncertainties
related to parameters like damping, material strength and mass have relatively small effect
on the overall uncertainty in collapse predictions. However, these studies were confined to
determinations of pre - collapse analysis. Contrary to this approach, Ibarra et al. (2005) has
shown that uncertainty related to the modeling deformation capacity and post-peak
softening response of component model had a significant influence on the collapse
assessment of the structure. However, Liel (2008) has effectively dealt with these issues
and quantified the modeling uncertainties in the form of a parameter. The present work
adopts the approach similar to Liel (2008) and Haselton and Deierlein (2007) in quantifying
the modeling uncertainties.

5.6.1 Modeling approach and collapse capacity assessment methodology


The modeling approach for assessment of collapse capacity assessment has been
previously discussed and the two-dimensional models as described in Chapter 2 have been
used in the analytical study. The beam-column elements have been modeled using a nonlinear hinge model developed by Ibarra et al. (2005) as shown in Figure 5.15 .The collapse
assessment methodology is described through an example of a 18 storeyed building (Figure
5.16) . The collapse assessment methodology has been briefly described as follows
a) Nonlinear time history analysis was performed to evaluate the collapse capacity of the
building models subjected to the 27 different ground motions as described in Chapter 2
(Vamvatsikos and Cornell 2002).
To determine the seismic collapse capacity, amplitude of each horizontal ground motion
has been scaled with respect to the spectral acceleration at the first mode period Se/Ag (T1).
These scaled accelograms were applied individually to the building models using an IDA

264

approach (Figure 5.17 a). The scale factors were continuously increased until collapse of
the structure occurred. In the present approach collapse of a building has been defined as
the point at which the maximum storey drift exceeded the code prescribed limit (side-sway
collapse).
(a) The collapse fragility curves were constructed, these curves represented probability of
collapse as a function of ground motion intensity (Figure 5.17 b, Curve R 1).
(b) To account for the modeling uncertainties an additional dispersion ( total = Model (0.5) +

RTR

(0.32) = 0.593) was introduced as suggested by Haselton and Deierlein (2007, Liel

2008). The Curve R2 shows the adjusted collapse capacity accounting for modeling
uncertainties.

(a) Plan

(b) Elevation

Figure 5.16 Basic Layout of the Building model


(c) The median of the collapse fragility curve was increased to account for ground motion
spectral shape as shown in Figure 5.17 b (curve R 3).
(d) As per the previous research works, the ground motion characteristics played an
important role in determining the structural response. Therefore, the spectral acceleration is
considered as an effective parameter in determining the intensity measure of the ground
motion (Baker and Cornell 2005). However, considerable difference in seismic response of
the structure subjected to the ground motion scaled to a fixed spectral acceleration was
observed. Hence, there is a necessity to consider other parameters apart from intensity
measure to reduce the error due to record-to-record variability. To account for this error,
Boore et al. (1997) introduced a parameter , The parameter can be determined by
equations proposed by Boore and Atkinson (2006) The effect of the parameter on the
structural characteristics has been presented in Figure 5.18 which shows the variation of
parameter with ground motion characteristics. Therefore, the parameter epsilon should be

265

considered in determining the collapse capacity. To consider the effects of threedimensional ground motions, the lower collapse capacity from each ground motion was
recorded as the building collapse capacity.
(e) The modified collapse capacity accounting for vertical collapse modes as obtained
(Figure 5.19).
In general, the seismic design codes are mainly focused on limiting the side-sway
collapse which is evident from the fact that most of the codes consider the side-sway
mechanism as the predominant collapse mechanism and prescribe the limits for the lateral
drift. However, in reality the building models may collapse due to side-sway collapse, vertical
collapse also or combination of both. This aspect has been ignored by most of the seismic
codes and majority of the research works pertaining to the collapse capacity assessment.
To avoid such failures, Aslani (2005) and Liel (2008) have discussed the importance of
vertical collapse modes in their research works. Aslani (2005) based on his experimental
studies on RC columns (92 cyclic tests) developed fragility functions to compute the vertical
collapse capacity. In his studies, the author observed four damage states for the flexureshear critical failure of columns. The first state started with light cracking followed by the
second state in which severe cracking of the columns occurred. The shear failure of the
column occurred in the third state which was characterized by cross cracking and yielding of
transverse reinforcement. The final damage state signified that the column had lost its
capacity to resist vertical loads which is most dangerous as it indicates complete collapse of
the structure. The fragility functions proposed by Aslani 2005 predicted the last two critical
damage states (third and fourth) as a function of column drift ratio. These fragility functions
are described as

Cas

Clvcc

1
pa
0.26
Ag f c
sh

25.4

1
100

pa

25.4
0.26

dc
Ag f y s sh
a1

(5.8)

1
10

(5.9)

where Cas and Clvcc are column drift ratios corresponding to shear failure and loss of vertical
carrying capacity, Pa is the axial load on the column, Ag is columns gross cross sectional

266

area, sh is area of transverse steel provided in the column, fy is yield strength of tensile steel
and dc is centerline distance between ties and Sa1 is the stirrup spacing

Se/Ag (T) [g]

0
0

0.04

0.08

IDR Max
(a)
1

0.75

Pc

R1

0.5

R2
R3

0.25

0
0

Se/Ag (T) [g]


(b)
Figure 5.17 (a) Results of incremental dynamic analysis, (b) Collapse fragility curve
with adjustment for different errors
The effect of considering the vertical collapse modes along with the side-sway collapse
modes has been presented in Figure 5.19. From Figure 5.19 it can be observed that
probability of collapse was maximum when both vertical and lateral modes were considered
(A2) as compared to the case where sway mechanism (A1) and vertical collapse mode (A3)
were considered separately. This observation is in correlation with building failure observed

267

during Darfield earthquake 2010 (Figures 5.20 and 5.21). Therefore, the vertical collapse
should necessarily be considered in determining the collapse to get the realistic estimate of
seismic collapse capacity of the building. In the present study the column drift ratio has been
computed from the dynamic analysis results by subtracting the joint shear deformation and
beam rotation from the inter-storey drift as suggested by Liel 2008.
(e) Probability of collapse for all the building models has been computed.
(f) The regression analysis is conducted on the results obtained for the building models
considered, and equations to estimate the collapse capacity and probability of collapse in
terms of different parameters were proposed.
6

cp(g)

0
-2

-1

Epsilon ()
Figure 5.18 Variation of collapse capacity (cp) with parameter epsilon ()
1
0.75

Pc

A1

0.5

A2
A3

0.25
0
0

Se/Ag (T) [g]


Figure 5.19 Effect of vertical collapse capacity

268

Figure 5.20 Pyne group corporation building New - Zealand photographed from North
East elevation after Darfield earthquake (Kam et al. 2010)

(a)

(c)

(b)

(d)

(e)

(f)

Figure 5.21 (a) Overall failure of PGC building (during Darfield earthquake 2010) due
to detachment of 1st and 2nd storey columns from its base due to which they lost their
vertical load carrying capacity, (b) Zoom A detail, (c) Zoom B detail, (d) Zoom C detail,
(e, f) Detail of damage to beam column joints (Kam et al.2010)

269

This page is intentionally left blank

270

5.7 Variation of collapse capacity for different building models


The previous section deal with the methodology of determining the collapse capacity. This
section aims to study the variation of collapse capacity with structural irregularity. The
collapse capacity of a building is an important aspect as it indicates probability of failure of
the structure. Moreover, collapse capacity of a structure varies with type, magnitude and
location of irregularity. Therefore, it is very important to study the variation of collapse
capacity with these aspects. To achieve this purpose, the collapse capacity of the building
models as discussed in Chapter 2 are determined. The variations of median collapse
capacity with different types of irregularity have been plotted in Figure. 5.22 to 5.26 from
which following observations are drawn at
(a) Mean collapse capacity increased with the increase in pc/ pl ratio for large values of
parameter . (Figure. 5.22).
(b) Mean collapse capacity (cp) increased linearly with yield base shear coefficient and
cyclic deterioration parameter () (Figure. 5.22).
(c) Increase in the maximum roof displacement increased the mean collapse capacity and
variation in the bay width had a least impact on the collapse capacity (Figure 5.23).
(d) Collapse capacity of the buildings was influenced by location of the irregularity. For mass
irregular buildings, the collapse capacity was greatest for the case when mass irregularity at
bottom position and least for mass irregularity at the top storey which is marginal. This trend
was reverse for building frames with stiffness and strength irregularity i.e. building frames
with strength and stiffness irregularity at the bottommost position had the least collapse
capacity as compared to the case when irregularities were placed at top storeys (Figure.
5.24).
(e) Mean collapse capacity (cp) decreased with increase in the magnitude of mass, stiffness
and strength irregularity. This increase was marginal for mass irregular building models but
this was higher for stiffness and strength irregularity (Figure. 5.25).
(f) The mean collapse capacity was observed to be lowest for the case when setbacks
originated from the bottom storey as compared to the case when setbacks originate from top
storeys) [Figure 5.26]. The collapse capacity has been observed to decrease with increase
in eccentricity[Figure 5.26].
(g) The mean collapse capacity reduced with cracking consideration and this percentage
was greater for setback irregularity. [Figures 5.26 to 5.28].

271

= 0.095
= 0.19
= 0.38

Mean cp (g)

6
4
2
0
0

10

15

20

25

Mean pc/ pl
Figure 5.22 Variation of collapse capacity with post capping rotation ratio

Mean cp (g)

Bay 1
Bay 2
Bay 3

0
1.5

2.5

3.5

Mean rd (cm)
Figure 5.23 Variation of collapse capacity with the bay width

Mean cp (g)

6
M1
M2
S1

3.5

S2
ST 1
ST 2

1
0

21.25

42.5

63.75

85

Mean
Figure 5.24 Variation of collapse capacity with location of mass, stiffness and
strength irregularity

272

M1BS

Mean cp (g)

5.4

M1MS
M1TS
S1BS
S1MS

3.7

S1TS
ST1BS
ST1MS

2
0

22

44

66

ST1TS

Mean
Figure 5.25 Variation of collapse capacity with magnitude of mass, stiffness and
strength irregularity
5.5

Mean cp(g)

SEC
5

SECC
SEB

4.5

SEBC
SEA

SEAC
3.5
0

25

50

75

Mean
Figure 5.26 Variation of collapse capacity with magnitude of setback irregularity
5.5

Mean cp(g)

P1
P1C

P2
4.5

P2C
P3

P3C
3.5
0

25

50

75

Mean
Figure 5.27 Variation of collapse capacity with magnitude of plan irregularity

273

Mean cp (g)

M1
M1C
S1

3.5

S1C
ST 1
ST1C

1
0

21.25

42.5

63.75

85

Mean
Figure 5.28 Variation of different parameters with cracking for building models with
mass, stiffness and strength irregularity

5.7.1 Equations to estimate the probability of collapse and collapse capacity


(Method 3)
The methodology of determining the collapse capacity has been briefly described in the
previous section. The previous research works (Liel 2008;; Shafei et al. 2011) suggested
that the parameters like fundamental time period, yield base shear coefficient (), irregularity
index, inter-storey drifts, plastic rotation capacity (pl) and post capping rotation capacity
(pc) etc. have a significant impact on the collapse capacity of the building. These
parameters (1, pl , pc) have been determined by the method as suggested by Liel 2008.
Furthermore, based on the regression analysis conducted on the seismic response
databank consisting of parameters with high sensitivity the equation to predict collapse risk
has been proposed by the author as

cp 1.631 0.289T 0.3348 C 0.632 4.45

pc
0.87m' 0.1213C s 0.074Clvcc
pl

Pc 0.2112 0.1312T 0.0143 C 0.0385 0.436

pc
0.0714m' 0.1246C s 0.069Clvcc
pl

(5.10)

(5.11)

The collapse capacity (cp) and probability of collapse (Pc) are determined by proposed
equation 5.10 and equation 5.11) and compared with the dynamic analysis results for the
building models and found to be in close agreement with the correlation coefficients of
0.9573 and 0.9813 (Figure 5.29 and Figure 5.30). In Figure 5.19, the maximum value of

274

collapse capacity is normalized to unity for effective presentation of comparison between the
proposed method and dynamic analysis.

cp(Proposed equation)

R2 = 0.9573

0.775

0.55
0.55

0.775

cp(Dynamic Analysis)
Figure 5.29 Comparison of collapse capacity evaluated using the proposed method
with the dynamic analysis

Pc(Proposed equation)

R = 0.9819

0.775

0.55
0.55

0.775

Pc(Dynamic Analysis)
Figure 5.30 Comparison of probability of collapse using the proposed method with
the dynamic analysis for the irregular buildings considered in the analytical study

275

5.8 Estimation of Damage indices


The damage analysis (for application in the seismic assessment and rehabilitation of
existing buildings and performance based approaches) has been a subject of intense
research during the past decade with many researchers contributing towards this aspect
(Sreekala et al. 2011; Kamaris et al. 2013 etc). The concept of describing the damage state
of the structure by one number or in the form of a damage index is attractive because of its
simplicity. However, development of such index is complex as it should be applicable to
various structural systems at advanced stages of inelastic deformation and up to the
collapse state. The damage state of a structure can be defined in different ways:
(a) a binary damage state (failure/no failure)
(b) a discrete valued damage state using qualitative indicators such as none, minor,
reparable, severe and failure states.
The empirical and theoretical approaches have been applied to determine various estimates
of structural damage. The empirical damage models are based on damage statistics
observed due to the structural damage occurred during previous earthquakes. These
models provide useful qualitative information on the overall seismic performance of
structural systems. However, the empirical evaluation is not efficient in predicting the
strength reserve and response characteristics of a structure with a specified degree of
damage because:
(i) It completely ignores the aspect of inelastic cyclic deformation.
(ii) The future earthquakes may have different characteristics as compared to the present
ones e.g. intensities, duration, and frequency content.
(iii)The buildings in other locations and recently built structures which are designed as per
current codes may differ from the structures which were used to develop the damage
statistics.
(iv) The dynamic characteristics of the buildings included in the statistical analysis may have
changed due to repairs and damage accumulation from previous earthquakes.

5.8.1 Literature review pertaining to damage index


Damage in a structure under loading can be defined as the degradation or deterioration of
its integrity resulting in reduction of its load capacity. In earthquake-resistant design of
structures, some degree of damage in the structural members is generally accepted. This is
done because the cost of a structure designed to remain elastic during a severe earthquake
would be very large. Thus, existing seismic codes, e.g., EC8:2004, and more recent

276

performance-based seismic design methods (Fajfar and Krawinkler 1997; FEMA 1997) have
adopted the concept of damage to establish structural performance levels corresponding to
increasing levels of the seismic actions. These performance levels mainly describe the
damage of a structure through damage indices such as the inter-storey drift ratio (IDR),
member plastic rotations etc. Several methods to determine damage indices as functions of
certain response parameters have been presented in the literature. In general, these
methods can be classified into non-cumulative or cumulative. The most commonly used
parameter is ductility which relates damage only to the maximum deformation and is still
considered as a critical design parameter by codes. To account for the effects of cyclic
loading, simple hysteresis rules of stiffness and strength degradation were included in
various non - cumulative indices in various research studies (Banon and Veneziano 1982;
Roufaiel and Meyer 1987a; Cosenza et al. 1993). The non cumulative indices have been
mainly used for reinforced concrete members. Cumulative-type indices can be classified into
deformation based (Stephens and Yao 1987) and hysteresis based (McCabe and Hall 1989;
Bracci et al. 1989) indices and methods which consider the effective distribution of inelastic
cycles and low-cycle fatigue of metals through a concept of linear damage accumulation
(Krawinkler and Zohrei 1983). Sucuoglu and Erberik (2004) developed low-cycle fatigue
damage models for deteriorating systems on the basis of experimental and analytical
results. Kamaris et al. (2013) proposed a new damage model exhibiting strength and
stiffness degradation which considered the phenomenon of low-cycle fatigue and the
interaction between axial force and bending moment at a section of a beam-column steel
member. (Park and Ang 1985) proposed cumulative damage indices as combinations of
deformation and energy dissipation. In these methods damage was expressed as a linear
combination of the damage caused by excessive deformation and that due to repeated
cyclic loading effects (Park and Ang 1985). An extensive review pertaining to the aspect of
damage indices can be found in Powell and Allahabadi (1988). Finally, (Lemaitre 1992)
employed the concept of continuum damage mechanics

in conjunction with the finite

element method for damage analysis of steel and reinforced concrete structures
(Hatzigeorgiou and Beskos 2007; Kamaris et al. 2009). The damage indices used in the
previous studies have been presented in Tables 5.3 to 5.5

5.8.2 Brief discussion on damage indices


In general, the number of load cycles have a dominant impact on the damage experienced
by the structure during the seismic excitation. Therefore, non-cumulative indices can just

277

give approximate measure of damage but have wide applicability as they are easy to
evaluate. Many cumulative damage indices reflect the mechanism of seismic damage such
as fatigue equations or energy equations. However, the major problem with these
formulations is the estimation of weighing factors or exponents which must be either derived
from the experimental data or from arbitrary values. Moreover, these parameters vary with
structural configurations. Thus, large number of experimental data is required to calibrate
them accurately. Regarding the deformation based models both Wang and Shah (1987) and
Chung et al. (1987) use arbitrary exponents but Jeong and Iwan (1987) have calibrated their
model against the experimental tests on beam column joints. In addition , the application of
inner s rule to represent non-linear cyclic behaviour is questionable.
The use of weighted measures to estimate the global damage has following limitations
(a) The reliability of global index in turn depends upon effectiveness of local index on basis
of which it was derived.
(b) There is no method to exactly determine the weights that should be used to different
structural elements at different levels of structural damage. The softening damage indices
have a distinct disadvantage that they do not indicate the distribution of damage in a
system. Final softening is a measure of stiffness degradation and is related to flexural
damage ratios. The maximum softening index gives better indication of damage but it is
difficult to apply to the structure not equipped with strong motion instruments.
Table 5.3 Summary of local damage indices (cumulative) proposed by different
researchers
Name and year

Type of index

Niu and Ren (1996)

CDI

Park and Ang (1985)

CDI

Kunnath et al. (1992)

CDI

Bracci et al. (1989)

CDI

Columbo and Negro (2005)

CDI

Damage Index
E

DI m
23
u
Eu

DI
DI

m
dE
e
u
Fy u

m y
dE
e
u y
M yu

D Dm D Dm D , Dm
DI 1

278

F 11


m
, D m y
My
f y

M ac
M yo

Table 5.4 Summary of other local damage indices proposed by different researchers

Type of
Name and year

Damage Index
index
F
DI i i

Gosain et al.(1977)

EB

Banon et al.(1981)

CI (DB)

Banon et al.(1981)

NCI (FDR)

Roufaiel and Meyer


(1987a)

NCI (FDR)

DI

DB

D (wi

Chung et al. (1987)

i Fy y

NCR

DI

Ko
Km

k f k k
m o
km k f ko

ni
ni

w
)
i
nf ,i
nf ,i


1 b r 22
DI
22
f

Stephens and Yao (1987)

DB

Wang and Shah (1987)

HY

Jeong and Iwan (1988)

DB

n
DI i
n
i f

Powell and Allahabadi


(1988)

NCI

DI

m 1
o 1

Kratzig (1989)

EB

p ,i

Kunnath et al.(1992)

HY

DI

exp( slb) 1
exp( sl ) 1

Ei

E f Ei

M m (i 1) M mi(1 e

dE )

M yu

Mehanny and Deierlein


(2001)

DB

( p Currentphc ) p phc
i 1


( pu
) p FNC
i 1

279

Table 5.5 Summary of global damage indices proposed by different authors


Type of

Name and year

Damage Index

index
Park , Ang, wen 1985,
1987,Chung et al. 1989,
1990, Kunnah 1990, 1991

Weighted
index

Bracci (1989)

Weighted
index

Roufaiel and Meyer 1987 (b)

DStorey

Mork (1992)

DStorey

w D
w D
i

Softening
indices

DGlobal

b 1
i
b
i

m y 14.2 y ( fund / f dam ) 1

f y
f y
2
Tund
Dpl 1 2
Tm

Plastic
softening
Di Pasquale and Cakmak
1989, Di Pasquale 1990

D E
E

Final
softening

Df 1

Final
softening

D1 1

b22 is a constant, CI Cumulative index, Df, Dpl

2
Tund
2
Tdam

k1,m
k
, D2 1 2,m
k1,und
k2,und

Deformation at failure and at plastic

softening, Dstorey - Damage in a storey, Ef+, Ef-, - Energy at failure during positive and
th

negative half cycle, DB - Damage based, Ei - Energy at i iteration, EB - Energy based ,

Ep+, Ep-, Potential energy at positive and negative, F11 - Constant, Fi and Fy Force in ith
th

cycle and at yield, FDR - Flexural damage ratio, HY Hysteresis based i cycle, ko,kf, km Initial stiffness and maximum stiffness, Mac - Actual (deteriorated) value of the yield moment
(force), My0 - Value characterizing the yield point in the theoretical skeleton curve, Mm, Mm i
and Mm (i+1) - Maximum moment, Maximum moment at ith and i + 1th cycle, mm and my Maximum and yield moment, MFDR
cumulative index, r

22

- Maximum flexural damage ratio, NCI - Non

- constant, Tundam - Undamaged fundamental time period, Tdam -

Damaged fundamental time wi - Weighing factor, 23 - Constant, m, o, - Maximum


ductility ratio, initial ductility ratio and ductility ratio at curvature , a - Constant, m and y Maximum and yield rotation, e - constant, f, +- Deformation at failure and deformation
at positive half cycle, b,c - Locally defined constants, sl - Slip parameter, nf and nf - Number
th

of cycles till i iteration, number of cycle at failure, i and y - Deformation in ith cycle and at

280

yield, half cycle, u - Ultimate displacement, , m, y - Normalized curvature, maximum


curvature, yield curvature, p+PHC, p+FHC - Positive rotation in primary half cycle and follower
half cycle, p

+
PHC,

p+FHC - Negative rotation in primary half cycle and follower half cycle, u

Ultimate rotation,

5.8.3 Park and Ang (1985 a) damage index (Method 4)


Park and Ang (1985a) damage index is best known and most widely used index. It can be
expressed as a simple linear combination of normalized deformation and energy dissipation
as

DI

m
dE
e
u
Fy u

(5.12)

The first term in this expression is pseudo static displacement measure which considers the
cumulative damage. The main advantages of this model are that
(a) It is simple.
(b) It has been calibrated against significant amount of observed damage including shear
and bond failures.
(c) It captures the effect of cyclic degradation effectively.
Park et al. (1985 b) classified the structural damage into five categories and prescribed
the damage index associated with these conditions as shown in Table 5.6. Park et al. (1985
b) suggested that for practical purposes the structure was repairable if the damage index is
less than 0.4, and the damage is un-repairable if the damage index exceeded 0.4, and a
value of 1.0 or more indicates the total collapse.
The Park and Ang (1985a) index was implemented in original version of IDARC (Park,
Reinhorn, Kunnath 1987) and has been used in large number of seismic vulnerability
studies (Ref).
Kunnath et al.(1992) presented a slightly modified version of Park and Ang (1985 a) index
and adopted Moment and curvature in his damage index which was represented as

DI

m y
dE
e
u y
M yu

(5.13)

This index was later adopted by Stone and Taylor (1993) who proposed a damage
classification based on 82 experimental tests conducted on circular columns. A number of
difficulties were observed in using equations 5.13. A major difficulty was encountered in
computation of ultimate deformation u or u and strength deterioration parameter e. Park

281

and Ang (1985a) proposed regression equations for two variables in terms of parameters
like shear span ratio, axial load, longitudinal and confining reinforcement ratio and material
strengths. However, the term e yielded very small values and therefore made the energy
term negligible and hence cycling loading effect was ignored. The other researcher like
Kunnath et al. (1990) and Stone and Taylor (1993) have proposed regressions to determine
parameter e (0.1 0.5) which yielded substantial energy terms. Ciampoli (1989) used a
realistic approach and assumed e as 0.27 with coefficient of variation as 0.6. Another main
problem arises with non - linearity of damage scale with a damage index of 0.4 implying
severe damage. From results of analytical studies, it was found that the damage index
showed large dependence on the proposed irregularity index and the behavior factor, and
no separate distinctions were made between short and long period structures (results for
short period structures have been presented in appendix C). However, in reality seismic
damage experienced would depend largely upon the time period of ground motion and the
structure. Therefore, based on the regression analysis conducted on the results of the
damage index (proposed by Park and Ang 1985a) obtained for the building models
considered in the analytical study. The simplified equations to estimate Park and Ang
(1985a) has been proposed by the authors as (Method 4)

DI 0.0991q 0.0326c

(5.14)

On comparison between the proposed equation and dynamic analysis, both results were
found to be in close agreement with a correlation coefficient of 0.9964 as shown in Figure
5.31.
Table 5.6 Interpretation of Park and Ang (1985a) global damage index
Damage

Physical appearance of the structure

DI

classification
Slight

Occurrence of cracking throughout

Minor

Occurrence of minor cracks throughout the structure < 0.2

Moderate

along with crushing of concrete columns


Appearance of cracking followed by spalling of concrete < 0.5

< 0.1

in weaker elements.
Severe

Severe crushing of columns followed with buckling of

< 0.9

reinforcement.
Collapse

Total or partial structural collapse.

282

> 1.0

DIPA (Proposed equation)

0.54
R2 = 0.9964

0.34

0.14
0.14

0.235

0.33

0.425

0.52

DIPA (Dynamic analysis)

Figure 5.31 Comparison of damage index (Park and Ang 1985 a) determined using
proposed equation and dynamic analysis for irregular buildings considered in the
analytical study (DIPA denotes Damage index proposed by Park and Ang 1985a)

5.8.4 Seismic damage indices proposed by the Author (Method 5)


On observing the previous damage indices used it could be clearly observed that majority of
the damage indices fail to capture the cyclic response. Park and Ang (1985 a) and Kunnath
et al. (1992) proposed index captured the damage due to cyclic deterioration but involves
too many parameters which made calculation of this index tedious. Although, author has
tried to represent the Park and Ang (1985) index calculation much simpler by proposing the
simple equations to estimate these damage indices. However, a new approach could be
adopted regarding this aspect. From the review of previous research works it has been
observed that inelastic seismic demands could be effectively used to represent the cyclic
deterioration (Kunnath and Kalkan 2004). Therefore, in the present study, the author has
proposed global (Sdg), storey (Sds) and local damage indices (Sdl) as ratio of corresponding
inelastic seismic demand to its prescribed code limit.

Sdg

I (irr )
rd (irr )
(irr )
Sds rd
Sdl i
,
,
I rd (ref )
rd (ref )
i (ref )

(5.15)

where rd(irr), Ird(irr), i(irr) are the value of respective parameters for irregular building
model, rd(ref), Ird(ref), i(ref) are the limit value of respective parameters considered by
author (as H/500 as per ASCE 7.05, 0.004h as per IS 1893:2002, and i(ref) has been

283

evaluated for maximum drift). The proposed index ranges from 0 to 1 for no failure and for
complete collapse respectively. As observed from equation 5.15 that it was very simple and
can be easily estimated. The increase in the damage index represents increase in the level
of seismic damage. The proposed equation for global damage index has been observed to
be in close agreement with Park and Ang 1985 a index (Figure 5.32).
0.54

DIPA

R2 = 0.9888

0.34

0.14
0.14

0.235

0.33

0.425

0.52

D I (Proposed Index)

Figure 5.32 Comparison of Park and Ang (1985a) index determined using proposed
equation and dynamic analysis for irregular buildings

5.8.5 Comparison of probability of collapse using different approaches


This section aims to compare the probability of collapse estimated using four different
approaches
(a) Method 1 - Probability of collapse evaluated using GlaIster and Pinho (2003) method
(b) Method 2 - Probability of collapse using Hazus methodology
(c) Method 3 - Probability of collapse using Ibarras non linear hinge model
(d) Method 4 -Park and Ang (1985a) damage index using author proposed simplified
expressions.
(e) Method 5 - Proposed damage index. The probability of collapse of irregular building
models studied have already been plotted in Chapter 3. In this section, the probability of
collapse using method 3, 4, 5 have been computed (Figure 5.33 to 5.37) and percentage
difference between all five methods have been plotted (Figure 5.38). From these figures it
could be clearly observed that

284

(i) Method 3 yielded highest mean probability of collapse with a percentage difference of
1.52%, 3.12% as compared to Method 1 and Method 2. However, this percentage reduced
to 1.23% and 1.41% for Method 4 and Method 5.
(ii) Method 4 and Method 5 yielded very close results with a mean difference of 0.18%
indicating the effectiveness of the proposed seismic damage index.
(iii) Method 2 yielded lowest collapse probability especially for structures with stiffness,
strength and setback irregularity.
108

H (m)

90
72

METHOD 3
METHOD 4

54

METHOD 5

36
18
0.12

0.19

0.26

Mean Pc (%)
Figure 5.33 Mean probability of collapse for buildings with mass irregularity

108

H (m)

90
72

METHOD 3
METHOD 4

54
METHOD 5

36
18
0.16

0.22

0.28

Mean Pc (%)

Figure 5.34 Mean probability of collapse for buildings with stiffness irregularity

285

108

H (m)

90
METHOD 3

72
METHOD 4
METHOD 5

54
36
18
0.25

0.34

0.43

Mean Pc (%)

Figure 5.35 Mean probability of collapse for buildings with strength irregularity

108

H (m)

90

72

METHOD 3
METHOD 4

54
METHOD 5

36

18
0.28

0.37

0.46

Mean Pc (%)

Figure 5.36 Mean probability of collapse for buildings with setback irregularity

286

108

H (m)

90
72

METHOD 3
METHOD 4

54
METHOD 5

36
18
0.16

0.22

0.28

Mean Pc (%)

Percentage difference in Pc

Figure 5.37 Mean probability of collapse for buildings with plan irregularity

3.4

2.55

M
S
ST

1.7

SE
P

0.85

0
1

Method of estimating probability of collpase

Figure 5.38 Percentage difference with respect to method 2 in estimating probability


of collapse evaluated by different methods

287

5.9 Brief summary and main conclusions


This Chapter has been divided into three sections. The first section proceeds with a brief
introduction to the concept of collapse response followed by a detailed literature review
regarding different aspects of collapse capacity of buildings. At the end of this section, main
conclusions and shortcoming derived from previous literature works have been presented.
In the second section, the aspect of modeling to determine the collapse capacity has been
discussed at first. The adopted model (Ibarras model) with details of its calibration with the
equations to predict different parameters associated this model have been briefly described.
The collapse assessment procedure is described in detail and demonstrated for a example
building. Using this procedure, collapse capacity and probability of collapse for all the
building models (as described in Chapter 2) has been determined. Based on regression
studies on analysis results, simple equations to estimate collapse capacity and probability of
collapse have been proposed.
In the third section, a brief review of damage indices used has been presented with their
merits and demerits. A special consideration has been given to Park and Ang (1985) index.
This index has been evaluated for building models adopted in the present study and it has
been expressed in terms of the irregularity index (as proposed in Chapter 2). Moreover,
three new damage indices have been proposed and compared with the Park and Ang
(1985a) index for the irregular building models considered.
Finally, the probability of collapse has been evaluated using Park and Ang (1985) Index
and the proposed index. These collapse probabilities have been compared with those
obtained using Glaister and Pinho (2003) and Hazus (FEMA 1999) methodology. The main
conclusions of this chapter can be summarized as
(a) The review of previous literature works pertaining to the collapse capacity showed that
the SDOF models in conjunction with the pushover analysis were inefficient in predicting the
realistic collapse capacity. However, the FEM approach with IDA method as a analytical tool
yielded accurate estimate of the collapse capacity. However, this method was
computationally complex and tedious. Nevertheless, the aspect of structural irregularity and
cracking have been ignored in majority of the research works pertaining to collapse
capacity.
(b) The analytical study showed that the type, magnitude and location of irregularities had
strong influence on collapse capacity. The collapse capacity has been observed to be least
for building models with strength, stiffness and setback irregularity especially when these
irregularities were present at the bottom storey. This observation is of practical significance

288

as these irregularities are often encountered in the realistic buildings at bottom storeys (due
to functional requirements) and appropriate consideration should be given while designing
these buildings. This trend of variation of collapse capacity was observed to be reverse for
mass irregular buildings. For mass irregular buildings, the maximum impact on collapse
response was observed for the case when mass irregularity was present at the top storey.
The plan irregular building models showed less sensitivity to collapse response as
compared to the vertical irregularities (except mass irregularity). This may be due to least
sensitivity of seismic response to plan irregularity. In addition, it implies effectiveness of
dynamic torsional provisions as prescribed by the seismic design codes. The consideration
of cracking phenomenon reduced the collapse capacity.
(c) As discussed in earlier sections, FEM method in conjunction with IDA method is an
effective tool to determine the collapse capacity. However, due to its computational
complexities and time consumed, many researchers have adopted pushover analysis to
estimate the collapse capacity (Han et al.2010; Shafei et al. 2011). In this Chapter an effort
has been made to determine the collapse capacity accurately using IDA method. The
Ibarras nonlinear model has been adopted to achieve this purpose. This model captures all
major collapse modes observed during the experimental studies. The calibration parameters
for this model and the collapse assessment methodology has been adopted from Haselton
and Deierlien 2007. The seismic design codes (EC 8:2004; IS 1893:2002; UBC 97) have
considered sidesway collapse mechanism to be predominant collapse mechanism. This
approach has been adopted in previous literature works as well (Han et al.2010; Shafei et
al. 2011; Haselton et al. 2011a). However, vertical collapse mechanisms also influence the
collapse capacity and has been the cause of building failures during past earthquakes
(Figure 1.19 and Figure 5.20). Thus, ignoring this aspect may induce noticeable errors in
estimation of collapse capacity (Liel 2008; Haselton et al. 2011b).This is evident from failure
of buildings due to vertical collapse mechanisms during previous earthquakes (Figure 5.1 to
5.14). Therefore, the present research work makes an effort to incorporate both these
mechanisms (sidesway and vertical collapse mechanisms) along with aspects of structural
irregularity and cracking. Based on the regression analysis conducted on the analysis
results obtained, simple equations to estimate the collapse capacity and probability of
collapse for irregular buildings have been proposed in terms of irregularity index proposed
by the author. This has been done to incorporate effects of structural irregularity and
cracking. The comparison between the proposed equation and dynamic analysis showed a
close agreement of results obtained by both these methods.

289

(d) A brief review of literature works concerned with seismic damage indices showed that
majority of the seismic damage indices used are unable to capture the cyclic deterioration
effects of structural members during the seismic excitation. Park and Ang 1985a) made an
effort and proposed an index to overcome this shortcoming. However, the Park and Ang
(1985a) proposed index was observed to be complex and difficult to evaluate. In the present
study an effort has been made to simplify the evaluation of Park and Ang (1985 a)index. In
this effort, the Park and Ang (2005a) index has been evaluated for all the irregular models
and using regression analysis it has been expressed in terms of irregularity index in form of
a simple equation. The proposed equation showed comparable results with dynamic
analysis. Thus, the evaluation of this index was made simpler in the present study.
Moreover, an simplified damage index has been proposed in the present study (based on
inelastic seismic response of buildings) to effective capture the effects of deterioration. The
simplified index is proposed as the ratio of inelastic inter-storey drift of irregular frame to that
of code limit, as inelastic seismic response is capable of representing the effects of cyclic
deterioration (Kunnath and Kalkan 2004).The proposed equation showed close agreement
with results of dynamic analysis with a high correlation coefficient. Finally, comparisons
between the collapse probabilities evaluated have been performed. The analysis showed a
closer agreement between Method 1, Method 3, Method 4 and Method 5 as compared to
the second method which yielded lower collapse probability.

290

CHAPTER 6
SUMMARY AND CONCLUSIONS

6.1 Summary
A structure can be classified as irregular if it contains irregular distributions of mass,
strength and stiffness. The structural irregularity can be further classified as horizontal
and vertical irregularity. In reality, many existing buildings contain irregularity, and some
of them have been designed initially to be irregular to fulfill different functions e.g.
basements for commercial purposes created by eliminating central columns, and
reduction of sizes of beams and columns in the upper storeys to fulfill functional
requirements and for other commercial purposes like storing heavy mechanical
appliances etc. This difference in usage of a specific floor with respect to the adjacent
floors results in irregular distributions of mass, stiffness and strength in the building. Also,
many other buildings are accidentally rendered irregular due to variety of reasons like
non-uniformity in construction practices and material used. However, these irregular
structures (designed as per code provisions) exhibit poor seismic performance as
evident from the past records. The different seismic design codes prescribe different
limits of irregularity as discussed in Chapter 1. The review of previous seismic design
codes showed that the irregularities have been classified in terms of magnitude only
ignoring the irregularity location. However, review of previous literature (Al Ali and
Krawinkler 1998; Das and Nau 2003;Karavasilis et al. 2008a,b; Athanassiadou 2008)
show that irregularity location has a significant impact on the seismic response.
Therefore, code measures are inappropriate in quantifying the irregularity. In the present
study, an irregularity index has been proposed (based on dynamic response of buildings)
to capture both magnitude and location of irregularity. The proposed approach has been
compared with different code approaches in quantification of irregularity.
The fundamental time period is an integral parameter in seismic design
methodologies (FBD and DBD). Therefore, accurate estimation of this parameter is very
essential as it affects the seismic design of structural members. The review of previous
literature shows that the code proposed expressions were based on regression analysis
conducted on the data set consisting of experimentally determined period of few
buildings located in a certain region. Therefore, these expressions cannot be applied
universally to all regions and to all categories of buildings. On observing the previous
literature works it could be observed that majority of the time period expressions are
evaluated using eigen value analysis (based on Rayleighs analysis) which has certain
inherent limitations as discussed in Chapter 3. Nevertheless, majority of the seismic
design codes and previous literature works have ignored the aspect of structural

291

irregularity and cracking in estimating the fundamental time period, which in turn would
result in inappropriate design of structural members. Therefore, there is a necessity of
improved empirical expressions to estimate the fundamental time period for irregular
buildings. The present study overcomes above discussed limitations and proposes
modified expressions to estimate fundamental time period based on and inelastic
dynamic analysis. Finally, the application of estimated fundamental time period in
seismic design methods and in seismic vulnerability assessment has been briefly
discussed by the author.
The deformation demands are the key parameters which determine the behavior of
the structure. The design codes like FEMA 356 (2000) and EC8:2004 have formulated
the equations for estimating the deformations based on SDOF systems and elastic
analysis which is not realistic. Therefore, the procedures prescribed by these codes are
unsuitable for design of new structures. The present study aimed to propose the simple
equations to estimate the realistic seismic demands of irregular buildings based on
inelastic dynamic analysis. The aspect of collapse response is very crucial as it decides
safety and stability of structures. A large number of research studies were devoted to
evolve the collapse assessment methodology using SDOF and MDOF system. The
research studies regarding this aspect were confined to regular building models.
Therefore, the present study aimed to extend this approach to irregular building models
as well. Moreover, the collapse assessment methodologies (as reported in previous
studies) were complex and tedious to evaluate. Hence, simple empirical expressions
have been proposed (in the present study) to estimate the collapse capacity and
probability of collapse. Finally, on review of literature pertaining to damage indices like
Park and Ang (1985a) index captured the cyclic deterioration of buildings under seismic
excitation. However, this index was tedious and complex to evaluate. Therefore, the
present research study proposes a simplified expression to estimate this index.
Moreover, a new simplified approach was also proposed and compared with Park and
Ang (1985a) index for seismic vulnerability assessment of irregular buildings.

6.2 Main conclusions


(a) The review of previous literature works pertaining to analysis methods showed that
linear static analysis was simple and easily applicable to structures. However, this
method was over-conservative. The pushover analysis method was more accurate as
compared to the previous method and represented the nonlinear response of buildings;
but these methods are applicable where higher modal contributions are insignificant. The

292

time history analysis was observed to be the most accurate method among the analysis
methods available; but this method was complex and time consuming.
(b) On comparison of inelastic models, the hysteric model (proposed by Ibarra et
al.2005) was observed to be simple and efficient in representing the inelastic behavior.
Moreover, the calibrations for this model provided by Haselton and Deierlein (2007) were
based on experimental studies on large number of RC structural elements and covered
majority of failure modes of RC members observed in reality. Therefore, this model was
observed to be most suitable for the analytical study.
(c) The review of the seismic design codes and literature works indicated that the
magnitude of irregularity being given a greater preference in representing the irregularity.
However, location of irregularity has a significant impact on the seismic response
(Nassar and Krawinkler 1991; Al-Ali and Krawinkler 1998; Das and Nau 2003;
Karavasilis 2008b). Therefore, both magnitude and location of irregularity should be
captured effectively to represent the irregularity. The review of seismic design codes and
previous research studies indicate absence of such an index. To overcome these
limitations and to represent the aspect of irregularity effectively; an irregularity index has
been proposed by the author based on dynamic response parameters (Participation
factor and frequency of vibration). The comparison of the proposed index with previous
approaches clearly indicated the effectiveness of the proposed approach in capturing the
effects of different types of irregularity and structural cracking. In addition, the proposed
index easily fitted into the framework of current design methodologies and can be used
for seismic vulnerability assessment.
(d) The code proposed equations to estimate the fundamental time period ignored the
aspect of structural irregularity and cracking which is unrealistic. The eigen - value
analysis was more accurate (and clearly differentiated between different forms of
irregularity) as compared to the code equations in estimating the fundamental time
period. However, these equations mainly depend on mass and stiffness of structure and
ignored the aspect of ground motion parameters. Moreover, eigen - value analysis is
based on Rayleighs assumption which has certain limitations due to which they under
predict the time period as compared to the experimental results and dynamic analysis
results. The author has adressed these limitations and conducted inelastic dynamic
analysis on irregular building models incorporating the cracking effects (as described in
Chapter 2). Based on regression analysis conducted on time period results obtained,
simple equations to estimate fundamental time period of irregular buildings (incorporating
cracking effects) have been proposed by the author. The proposed equations showed
close correlation with dynamic analysis as compared to the code approach. Moreover,
the author proposed equations were very useful and easily applicable in current design

293

methodologies (PBD and DBD). The application of code equations in these


methodologies yielded conservative estimates of the seismic response. However, the
author proposed equations were simple and effectively fitted into the framework of these
design methodologies. In addition, the author proposed equations exhibited better
performance as compared to code approach, and yielded both safe and economical
estimate of the seismic response. Thus, the author proposed expressions can be
effectively adopted for seismic design of irregular buildings. Nevertheless, the proposed
equations were observed to be quite useful in seismic vulnerability assessment method
proposed by Calvi (1999) and extended by Glaister and Pinho (2003). From the seismic
vulnerability assessment of irregular buildings, it was observed that type and location of
irregularity and cracking has a significant impact on the seismic response. The seismic
vulnerability assessment methods by Glaister and Pinho (2003) (in conjunction with
proposed equations) was simple and easily to evaluate as compared to Hazus (FEMA
1999) method. Hazus method has a distinct disadvantage that it is based on static
method of analysis which is unrealistic. However, the author proposed equations
overcomes this limitation. The Hazus method underestimated the collapse probability as
compared to Glaister and Pinho (2003) method. Therefore, the author proposed
equations in conjunction with Glaister and Pinho (2003) method can be effectively used
for seismic vulnerability assessment which would save excess time and energy to be
invested in Hazus (FEMA 1999) methodology.
(e) The review of code procedures showed that the limits of drifts were based on SDOF
system and elastic analysis. Therefore, these equations yielded overconservative
estimate of the seismic response (Karavasilis 2008a; Karavasilis 2008b). The review of
previous literature works showed that the tall structures (18 storey to 36 storey) have
been ignored in majority of reported literature works pertaining to structural irregularity
and cracking. However, from the results of present research works, an abrupt increase in
seismic demands for tall structures was observed. As tall structures are being
increasingly preferred nowadays (Wikepedia), findings of present research work are
significant in formulating the improved design methodologies for these structures. The
study of effect of structural irregularities showed that presence of strength, setback and
stiffness irregularities had a greater impact on the drift as compared to plan and mass
irregularity. Moreover, aspect of irregularity location (ignored by the seismic design
codes) had a significant impact on the seismic performance of buildings. The location of
mass irregularity was observed to be critical in top storeys, and a reverse trend was
observed for other forms of vertical irregularities (stiffness, strength and setback
irregularity). The review of previous literature showed that the performance of code
proposed behavior factor and equal displacement in estimating the seismic demands

294

was not clear with some researchers advocating the better performance and others
suggesting the opposite view. However, with respect to estimation of seismic
performance of irregular structures, the code proposed behavior factor and equal
displacement rule was observed to yield over-conservative estimate of seismic demands
which is uneconomical. Therefore, a necessity of a modified rule (to overcome these
limitations) was justified. The seismic analysis results it has been concluded that the
parameters like irregularity index, behavior factor and structural irregularities have a
significant impact on the seismic response. Based on regression analysis conducted on
seismic analysis results, simple equations to estimate the inelastic seismic demands in
terms of the irregularity index (proposed by the author in second Chapter) have been
proposed. On comparison with dynamic analysis, the author proposed equations to
estimate deformation demands yielded accurate results; but EC 8 equations were
observed to be over - conservative. Furthermore, the proposed equations effectively
fitted into framework of current design methodologies and yielded safe and economical
design as compared to code and previous research work (Kappos and Stefanidou 2010).
The proposed research work also proposed a direct empirical relation between the
behaviour factor and irregularity index incorporating the aspects of structural irregularity
and cracking. The proposed behavior factor can be easily applied to generate the
inelastic ductility spectrums (methodology described in Chopra 2001; Medhakar and
Kennady 2000 a,b). which can be further used in seismic analysis and design of
buildings The proposed behaviour factor has been effectively used to estimate the
seismic demands as discussed earlier.
(f) The review of previous literature works showed that the SDOF models in conjunction
with the pushover analysis were incapable of predicting the realistic collapse capacity.
However, in spite of being tedious and computationally complex, the FEM approach with
IDA as a analytical tool yielded accurate estimate of the collapse capacity. It is quite
important to note that the aspect of structural irregularity was clearly ignored in majority
of the research works pertaining to the collapse capacity. The analytical study showed
that the collapse capacity to be significantly influenced by type, magnitude and location
of irregularity and cracking. In correlation with deformation demands, the collapse
capacity was observed to be least for building models with strength, stiffness and
setback irregularity. Moreover, these irregularities when present at bottom reduced the
collapse capacity drastically. However, these irregularities had a significantly lower
impact for the case when they were present in the middle and in the top storeys. This
trend of variation of collapse capacity was observed to be reverse for mass irregular
buildings. For mass irregular buildings, the maximum impact on collapse response was
observed mass irregularity was present at the top storey. The plan irregular building

295

models showed less sensitivity to collapse response as compared to the vertical


irregularities (except mass irregularity). This may be due to effectiveness of dynamic
torsional provisions as prescribed by the seismic design codes. In this chapter an effort
has been made to quantify the collapse capacity and to propose simple equations to
estimate the collapse capacity of irregular buildings. The Ibarras nonlinear hinge model
(Ibarra et al. 2005) has been adopted to achieve this purpose due to reasons as
discussed earlier. Based on non-linear analysis on collapse response database, simple
equations have been proposed to estimate collapse probability and collapse capacity.
The comparison between the proposed equation and dynamic analysis showed a close
agreement between collapse capacity obtained by both methods. Finally, a brief review
of literature works concerned with seismic damage indices showed that majority of the
seismic damage indices used were unable to capture the cyclic deterioration effects of
structural members during the seismic excitation. Park and Ang (1985a) made an effort
and proposed an index to overcome this shortcoming. However, the Park and Ang
(1985a) proposed index was observed to be complex and difficult to evaluate. In the
present study, an effort has been made to simplify the evaluation of Park and Ang
(1985a) index. In this effort, the Park and Ang (1985a) index was evaluated for all the
irregular models and using regression analysis it has been expressed in terms of
irregularity index in form of a simple equation (in terms of irregularity index proposed by
author). The proposed equation showed comparable results with dynamic analysis.
Thus, the evaluation of this index was made simpler in the present study. Moreover, a
new damage index has been proposed by author in the present study to capture the
effects of deterioration due to cyclic loading. The simplified index has been proposed as
the ratio of inelastic inter-storey drift of irregular frame to that of regular frame as inelastic
inter-storey drift is capable of representing the effects of cyclic deterioration (Kunnnath
and Kalkan 2004). The proposed equation showed close agreement with results of
dynamic analysis with a high correlation coefficient. Thus, the author proposed index can
be effectively used for assessment of seismic damage. The seismic vulnerability
assessment of irregular buildings estimated by five methods has been compared. The
comparison showed method 1,3,4,5 (details of methods are discussed in Chapter 3 and
Chapter 5) to be in close agreement. However, method 2 yielded lowest probability of
collapse as it is based on static method of analysis. The effects of irregularity can be
reduced by adopting ductile detailing provisions as per IS 13920. Finally, the range of
applicability of the proposed relations has been shown in Chapter 3, these ranges can be
exceeded as a scope for future study.

296

REFERENCES
ACI (2002), Building code requirements for structural concrete, ACI 318, American
concrete institute, U.S.A.
Adaliar, K., and Aydingun, O. (2001), Structural engineering aspects of the June 27,
1998 AdanaCeyhan (Turkey) earthquake, Engineering structures, Vol.23, pp. 343
355.
Adam, C., Ibarra, L.F., and Krawinkler, H. (2004), Evaluation of P-delta effects in nondeteriorating MDOF structures from equivalent SDOF systems, In Proceedings of 13th
World Conference on Earthquake Engineering, Vancouver, B.C., Canada, Paper No.
3407.
Akkar, S., and Bommer, J.J. (2007), Prediction of elastic displacement response spectra
in Europe and the Middle East, Earthquake Engineering and Structural Dynamics,
Vol.36, N0.10,pp.1275 1301.
Al-Ali, A.A.K., and Krawinkler, H. (1998), Effects of Vertical Irregularities on Seismic
Behavior of Building Structures, Report No. 130, 1998, The John A. Blume Earthquake
Engineering Center, Department of Civil and Environmental Engineering, Stanford
University, Stanford, U.S.A.
Alba, F., Ayala, A.G., and Bento, R. (2005), Seismic performance evaluation of plane
frames regular and irregular in elevation, In Proceedings of the 4th European workshop
on the seismic behavior of irregular and complex structures, CD ROM, Thessaloniki,
August 2005.
Alderighi, E., and Salvatore, W. (2009), Structural fire performance of earthquakeresistant composite steelconcrete frames, Vol.31, No.4, pp. 894 909.
Allahabadi, R., and Powell, G.H., (1988), DRAIN 2DX user guide, Report UCB/EERC88/06, Earthquake engineering Research center, University of California, Berkeley CA.
Ambrisi, A.D., Stefano, M.D., and Tanganelli, M. (2009), Use of Pushover Analysis for
Predicting Seismic Response of Irregular Buildings: A Case Study, Journal of
earthquake engineering, Vol.13, pp.10891100.

297

Anagnostopoulos, S.A. (1972), Nonlinear dynamic response and ductility requirements


of building structures subjected to earthquakes, Report No. R72-54, Department of Civil
Engineering, Massachusetts Institute of Technology, Cambridge, MA.
Anagnostopoulos, S.A., Alexopoulou, C., and Stathopoulos, K.G. (2010), An answer to
an important controversy and the need for caution when using simple models to predict
inelastic earthquake response of buildings with torsion, Earthquake Engineering and
Structural Dynamics, Vol.39, pp 521-540.
Anastasiadis, A., Gioncu, V., and Mazzolanni, F.M. (2000), New trends in the evaluation
of available ductility of steel members, In Behavior and Ductility of Steel Structures,
SDSS 99, Timisoara, 0-11 September 1999, Journal of Constructional Steel Research,
Vol.55, pp. 211-227.
Anderson, R.W. (1987). "The San Salvador Earthquake of October 10, 1986 Review of
Building Damage," Earthquake Spectra, Vol. 3, No.3, 497-541.
Araki, Y., and Hjelmstad, K.D. (2000), Criteria for assessing dynamic collapse of
elastoplastic structural systems, Earthquake Engineering and Structural Dynamics, Vol.
28, No. 8, pp. 1177 1198.
Aranda, G.R. (1994). Ductility Demands for R/C Frames Irregular in Elevation, In
Proceedings of the eighth World Conference on Earthquake Engineering, San Francisco,
U.S.A., Vol. 4, pp. 559-566.
Arslan, M.H., and Korkmaz, H.H. (2007), What is to be learned from damage and failure
of reinforced concrete structures during recent earthquakes in Turkey?, Engineering
Failure Analysis, Vol.14, pp.122
ASCE (2000), Prestandard and commentary for rehabilitation of buildings ( FEMA
356), Washington DC, Federal emergency management agency.
ASCE (2002), Minimum Design Loads for Building and Other Structures (ASCE/SEI 702), American Society of Civil Engineers, New York, U.S.A.
ASCE (2005), Minimum Design Loads for Building and Other Structures (ASCE/SEI 705), American Society of Civil Engineers, New York, U.S.A.
Aschheim, M., and Black, E.F. (2000), Yield point spectra for seismic design and
rehabilitation, Earthquake Spectra, Vol.16, No.2, pp.317 336.

298

Asgarian, B., Nojoumi, R.M., and Alanjari, P. (2012), Performance-based evaluation of


tall buildings using advanced intensity measures (case study: 30-story steel structure
with framed-tube system, The Structural Design of Tall and Special Buildings,
DOI: 10.1002/tal.1023.
Aslani, H. (2005), Probalistic Earthquake loss estimation and Loss Disaggregation in
buildings, Doctoral Dissertation (2005), Stanford university, U.S.A.
Astaneh-Asl, A., Modjtahedi, D., and McMullin, K. (1998), Stability of damaged steel
moment frames in Los Angeles, Engineering structures, Vol. 20, pp.433446.
ATC (1978), Tentative positions for development of seismic regulations of buildings
(ATC 3-06), Applied Technology council, Structural Engineers association of California,
US department of commerce, Washington D.C.
Athanassiadou, C.J. (2008), Seismic performance of R/C plane frames irregular in
elevation, Engineering Structures, Vol.30, No.5, pp. 1250-1261.
Ayala, D.D. (2005), Force and displacement based vulnerability assessment for
traditional buildings Bulletin of earthquake engineering, Vol. 3, pp. 235 265.
Aycardi, L.E. Mander, J., and Reinhorn, A. (1994), Seismic resistance of reinforced
concrete structures designed only for gravity loads: Experimental performance of
subassemblages, ACI Structural journal, Vol.91, No.5, pp.552 563.
Aydan, O. (1997), The seismic characteristics and the occurrence pattern of Turkish
earthquakes, Turkish Earthquake Foundation. TDV/ TR 97-007. January 1997.
Ayidin, K. (2007). Evaluation of Turkish seismic code

for mass irregular buildings,

Indian journal of engineering and material sciences, Vol.14, pp.220-234.


Ayoub, A., Mijo, C., and Chenouda, M. (2004), Seismic fragility analysis of degrading
structural systems, In proceedings of 13th world conference on Earthquake engineering,
Vancouver B.C., Canada, Paper No. 2617.
Aziminejad, A. and Moghadam, A.S. (2010), Fragility-Based performance evaluation of
asymmetric single-story buildings in near field and Far field earthquakes, Journal of
Earthquake Engineering, Vol.14, pp.789-816.

299

Aziminejad, A., and Moghadam, A.S. (2009), Performance of Asymmetric Multistory


Shear Buuldings with Different Strength Distributions, Journal of Applied Sciences,
Vol.9, pp.1082-1089.
Baber, T., and Noori, M.N. (1985), Random vibration of degrading, pinching systems,
Journal of Engineering Mechanics, ASCE, 111, No.8, pp. 1010-1026.
Baber, T., and Noori, M.N. (1986), Modeling general hysteresis behavior and random
vibration application, Journal of Vibration, Acoustics, Stress and Reliability in Design,
Vol. 108, pp. 411-420.
Baber, T., and Wen, Y.K. (1981), Random vibration of hysteretic degrading systems,
Journal Engineering Mechanics., ASCE, Vol.107, pp. 1069- 1089.
Baker, J.W., Cornell, C.A. (2005), A vector-valued ground motion intensity measure
consisting of spectral acceleration and epsilon Earthquake Engineering and Structural
Dynamics, Vol.34, No.10, pp.1193217.
Balkaya, A., and Kalkan, E. (2003) Estimation of fundamental periods of shear-wall
dominant building structures, Earthquake Engineering and Structural dynamics, Vol.32,
No.7, pp.985-998.
Baltzopoulou, A.D., Eleftheriadou, A.K., and Karabinis, A.I. (2012) Seismic Vulnerability
and Risk Assessment of the Building Stock of Attica (Greece) and Correlation to the
Actual Repair Cost 15th world conference on earthquake engineering, LEBOA 2012.
Banon , H., Biggs, J.M., and Irvine, H.M. (1981), Seismic damage in reinforced concrete
frames,Journal of structural engineering, ASCE, Vol.107, No.9, pp.1713 1729.
Banon, H., and Veneziano, D. (1982), Seismic safety of reinforced concrete members
and structures, Earthquake Engineering and Structural dynamics, Vol.10, pp.179193
Baradaran Shoraka, M., and Yang, T.Y., and Elwood, K.J. (2012), Seismic loss
estimation of non-ductile reinforced concrete buildings, Earthquake Engineering and
Structural dynamics, Earthquake Engineering & Structural Dynamics, Vol. 42, No.2, pp.
297310.
Bariola, V., and Brokken, S. (1991). Influence of strength and stiffness on seismic
structural behavior, Bulletin of Seismology and Earthquake Engineering, Vol.23, pp.427434.

300

Basu, D., and Gopalakrishnan, N. (2007). Analysis for preliminary design of a class of
torsionally coupled buildings with horizontal setbacks, Engineering Structures, Vol.30,
No.5, pp. 1272-1291.
Bernal, D. (1987), Amplification factors for inelastic p-Delta effects in earthquake
analysis, Earthquake Engineering and Structural Dynamics, Vol. 15, No.5, pp.635651.
Bernal, D. (1992), Instability of buildings subjected to earthquakes, Journal of structural
engineering, ASCE, Vol.118, No. 8, pp. 2239 2260.
Bernal, D. (1998), Instability of buildings during seismic response, Engineering
structures, Vol. 20, pp. 496 502.
Berry, M., and Ebrehard, M. (2003), Performance models for structural damage in
reinforced concrete columns, Pacific earthquake engineering database, U.S.A.
Bertero V.V. (1997), An Illustrated Introduction to Earthquake Engineering Principles
Simplicity, Symmetry, and Regularity, Web version by Vivian Isaradharm, National
Information Service for Earthquake Engineering University of California, Berkeley,
http://nisee.berkeley.edu.
Bertero, V. V. (1980), Strength and deformation capacities of buildings under extreme
environments, Structural engineering and structural mechanics, A Volume Honoring
Edgar P. Popov, K. Pister, ed., Prentice-Hall, Englewood Cliffs, N.J, pp. 188237.
Bertero, V.V., Anderson, J.C., and Krawinkler,H. (1994), Performance of steel building
structures during the Northridge Earthquake, Rep. No. UCB/EERC-94/09, Earthquake
Engineering Research Center, University of California, Berkeley, California, U.S.A.
Bhatt, C., and

Bento, R. (2011), Assessing the seismic response of existing RC

buildings using the extended N2 method, Bullitten of earthquake engineering, Vol.9,


pp.1183 1201.
Biondini, F. and Toniolo, G. (2000), Comparative analysis of the seismic response of
precast and cast-in-situ frames, Studies and Researches, Graduate School for
Concrete Structures, Politecnico di Milano, 21, 117.
Biondini, F., and Toniolo, G.. (2009), Probabilistic Calibration and Experimental
Validation of the Seismic Design Criteria for One-Story Concrete Frames, Journal of
earthquake engineering, Vol.13, pp.426 462.

301

Biondini, F., Toniolo, G., and Tsionis, G. (2008), Seismic design criteria for multi-storey
precast structures, In proceedings of 14th World conference on earthquake engineering,
October 12-17, 2008, Beijing, China
Bommer, J., Spence, R., Erdik, M., Tabuchi, S., Aydinoglu, N., Booth, E., del Re, D.. and
Peterken, 0. (2002) "Development of an earthquake loss model for Turkish catastrophe
insurance", Journal of seismological research letters, Vol. 6, No.3, pp.431 446.
Bommer, J.J., and Pinho, R. (2006), Adapting earthquake actions in Eurocode 8 for
performance-based seismic design, Earthqauke engineering and structural dynamics,
Vol.35, No.1, pp. 39 55.
Boore, D. M., and Atkinson, G.M. (2007), Boore-Atkinson NGA Ground Motion Relations
for Geometric Mean Horizontal Component of Peak and Spectral Ground Motion
Parameters,, Pacific Earthquake Engineering Research Center 2007/01, University of
California at Berkeley.
Boore, D.M., Joyner, W.B., and Fumal, T.E. (1997), Equations for estimating horizontal
response spectra and peak acceleration from western North American earthquakes: a
summary of recent work, Seismological Research letters, Vol.68, No.1, pp.128-153.
Bouc, R. (1967), Forced vibration of mechanical systems with hysteresis, In proceeding
of 4th conference on Nonlinear oscillation.
Bozorgnia, Y., and Bertero, V.V. (2004), Earthquake Engineering. From Engineering
Seismology to Performance-Based Engineering, CRC Press, Boca Raton, Florida.
Bracci , J.M., Reinhorn, A.M., Mander, J.B., and Kunnath, S.K. (1989), Deterministic
model for seismic damage evaluation of RC structures, Technical Report NCEER 89 0033.
Bracci, J.M., Kunnath, S.K., and Reinhorn, A.M. (1997), Seismic performance and
retrofit evaluation of reinforced concrete structures. Journal of structural engineering,
Vol. 123, No. 1, pp. 3 -10.
Bugeja, M.N., Thambiratnam, D.P., and Brameld, G.H. (1999), The influence of stiffness
and strength eccentricities on the inelastic earthquake response of asymmetric
structures, Engineering Structures, Vol.21, No.9, pp.856-863.

302

Cakmak, A.S., Rodriguez Gomez, S., Dipasquale, E. (1991), Seismic damage


assessment for reinforced concrete structures, In proceedings of ist national conference
on Soil dynamics and Earthquake Engineering, Karlsruhe, Germany, pp.515 -544.
Calvi, G. M. and Pavese, A. (I995), "Displacement based design of building structures,"
In proceedings of fifth SECED conference on European Seismic design practice,
Chester, United Kingdom, pp. 127 132.
Calvi, G.M. (1999), A displacement-based approach for vulnerability evaluation of
classes of buildings," Journal of Earthquake engineering, Vol. 3, No. 3, pp. 411 438.
Calvi, G.M. and Pavese, A. (1997) Conceptual design of isolation systems for bridge
structures, Journal of Earthquake Engineering Vol. 1, pp.193-218.
Calvi, G.M., Magenes, G., and Pampanin, S. (2002), Relevance of Beam-Column Joints
Damage and Collapse in R.C. Frame Assessment, Journal of Earthquake Engineering
Vol. 6, pp.75-100.
Calvi, G.M., Pinho, R., Crowley, H (2006), State-of the-knowledge on the period
elongation of RC buildings during strong ground shaking, In proceedings of first
European Conference on Earthquake Engineering and Seismology, Geneva, 38
September 2006, Paper Number 1535
Calvi, G.M., Pinho, R., Magenes, G., Bommer, J.J., Restrepo-Vlez, L.F., and Crowley,
H. (2006), The development of seismic vulnerability assessment methodologies for
variable geographical scales over the past 30 years, ISET Journal of Earthquake
Engineering Technology, Vol. 43, 75-104.
Cardone, D. (2007), Nonlinear Static Methods vs Experimental Shaking Table Test
Results, Journal of earthquake engineering, Vol.11, pp.847 875.
Carr, A. J. (2003), Ruaumoko user manual, University of Canterbury, New Zealand.
Carrillo, J., and Alcocer, S.M. (2013), Simplified equation for estimating periods of
vibration of concrete wall housing, Engineering Structures, Vol.52, pp.446 454.
Casciati, F. (1989), Stochastic dynamics of hysteretic media, Structural Safety, Vol. 6
(2-4), pp. 259269.

303

Chai, Y. H., Fajfar, P., and Romstad, K. M. (1998) Formulation of duration-dependent


seismic design spectrum, Journal of structural engineering, ASCE, Vol.124, No.8,
pp.913 921.
Challa, V.R.M., and Hall, J.F. (1994), Earthquake collapse analysis of steel frames,
Earthquake Engineering and Structural Dynamics, Vol. 23, No. 11, pp.11991218.
Chandler, A. M., and Duan, X. N. (1997) Performance of asymmetric codedesigned
buildings for serviceability and ultimate limit states, Earthquake engineering & structural
dynamics, Vol.26, No.7, pp. 717 735.
Chandler, A. M., and Hutchinson, G. L. (1992), Effect of structural period and ground
motion parameters on the earthquake response of asymmetric buildings, Engineering
Structures, Vol.14, No.6, pp.354 360.
Chandler, A.M., Correnza, J.C., and Hutchinson, G.L. (1995). Influence of accidental
eccentricity on inelastic seismic torsional effects in buildings, Engineering structures,
Vol.17, No.3, pp.167-178.
Chen, C., Lam, N.T.K., and Mendis, P. (2000), The bifurcation behavior of vertically
irregular buildings in low seismicity regions, In proceedings of 12th world conference.
On earthquake engineering, New Zealand (2000).
Chen, C., Ricles, J.M., Karavasilis, T.L., Chae,Y., and Sause, R. (2012), Evaluation of a
real hybrid simulation system for performance evaluation of structures with rate
dependent devices subjected to seismic loading, Engineering Structures, Vol. 35, pp. 71
- 82.
Chi, W.M., El-Tawil, S., Deierlein, G.G., and Abel, J.F. (1998), Inelastic analysis of a 17story steel framed building damaged during Northridge, Engineering structures, Vol.20,
pp.481 495.
Chintanapakdee, C., and Chopra, A.K. (2004), Seismic Response of Vertically Irregular
Frames: Response History and Modal Pushover Analyses, Journal of Structural
Engineering, ASCE, Vol. 130, No. 8, pp. 1177-1185.
Choi, B.J. (2004). Hysteretic energy response of steel moment-resisting frames with
vertical mass irregularities, The Structural Design of Tall and Special Buildings, Vol. 13,
pp.123 -144.

304

Chopra (2001), Dynamics of Structures: Theory and Application in Earthquake


Engineering, 2nd Ed., Prentice Hall, Englewood cliffs, NJ, 2001.
Chopra, A. K., and Kan, C. (1973). Effects of stiffness degradation on ductility
requirements for multistorey buildings. Earthquake Engineering and Structural Dynamics,
Vol.2, No.1, pp. 35 45.
Chopra, A.K. and Goel, R.K. (1973), Theory and Application in Earthquake Engineering,
2nd Ed., Prentice Hall, Englewood cliffs, NJ, 1973.
Chopra, A.K. and Goel, R.K. (2000). Building period formulas for estimating seismic
displacements, Technical Note, Earthquake spectra , 16 (2), 533-536.
Chopra, A.K., and Goel, G.K. (2002), A modal push over analysis procedure for
estimating seismic demands for buildings, Earthquake Engineering and Structural
Dynamics, Vol. 31, pp.561582.
Chopra, A.K., and Goel, G.K. (2004), A Modal pushover analysis procedure to estimate
to estimate seismic demands for unsymmetric-plan buildings, Earthquake Engineering
and Structural Dynamics, Vol.33, pp.903-927.
Chopra, A.K., and Goel, R.K. (2004), Evaluation of modal and FEMA pushover
analyses: Vertically regular and irregular generic frames, Earthquake spectra, Vol.20,
No.1, pp. 255 271.
Chou, J.H. (2006), Simplified methods to predict earthquake induced sidesway collapse
in modern reinforced concrete special moment frames, M.Sc. thesis, Department of Civil
and Environmental Engineering, University of California; 2006.
Chung, Y.S., Meyer, C., Shinozuka , M. (1987) , Seismic damage assessment of RC
members, Technical Report NCEER 87 022, National Center for earthquake
Engineering Research, State university of New York, Buffalo, NY.
Chung, Y.S., Meyer, C., Shinozuka , M. (1989 a), Modeling of concrete damage,
Structural journal, American Concrete Institute, Vol.86, No.3, pp.259 271.
Chung, Y.S., Meyer, C., Shinozuka , M. (1989 b) , Automated damage controlled
design of RC buildings, In proceedings of 5tn international conference on structural
safety and reliability (ICOSSAR 89), Vol.II, pp.1121 1128.

305

Chung, Y.S., Meyer, C., Shinozuka , M. (1990) , Automated sesmic design of reinforced
concrete building frames, Structural journal, American Concrete Institute, Vol.87, No.3,
pp.326 340.
Ciampoli, M.,Giannini, R., Nuti, C.,Pinto, p.e. (1989) , Seismic reliability of non linear
structures with stochastic parameters by directional stimulation, In proceedings of 5th
national conference on structural safety and reliability (ICOSSAR 89), San Francisco,
CA, Vol.II , pp.1121 1128.
Clarke, R. (2005), Non-Bouc degrading hysteresis model for nonlinear dynamic
procedure seismic design, Journal of Structural Engineering, ASCE, Vol.131, No.2, pp.
287 291.
Clough, R.W., and Johnston, S.B. (1966), Effect of stiffness degradation on earthquake
ductility requirements, In proceedings of Second Japan National Conference on
Earthquake Engineering, 1966, pp. 227-232.
Colombo, A. and Negro, P. (2005). A damage index of generalised applicability.
Engineering Structures, Vol.27, No.8,pp.1164-1174.
Comartin, C. D., Greene, M., and Tubbesing, S. K.(1995), The Hyogo-Ken Nanbu
earthquake, January 17, 1995, Preliminary Reconnaissance Report., Earthquake
Engineering Research Institute, Oakland, California.
Comartin, C.D., (1995), Guam earthquake of August 8, 1993 Reconnaissance Report
Earthquake spectra, Supplement B to Vol. 11.
Cook, R.D., Malkus, D,S., and Plesha, M.E. (1988), Concepts and Application of finite
element analysis, 3 edition, John Wiley and sons.
Cosenza E, Manfredi G, Ramasco R (1993) The use of damage functionals in
earthquake engineering: A comparison between different methods, Earthquake
Engineering and Strucural Dynamics, Vol.22, No.10, pp.855 868.
Cosenza, E., Manfredi, G., and Ramasco, R. (1993) The use of damage functionals in
earthquake engineering: A comparison between different methods, Earthquake
Engineering and Structural Dynamics, Vol.22, No. 10, pp. 855 868.

306

Crowley, H. (2003), Periods of vibration for displacement based assessment of


buildings, Phd dissertation submitted to rose school of earthquake engineering, IUSS
Pavia italy.
Crowley, H., and Pinho, R. (2006) Simplified equations for estimating the period of
vibration of existing buildings, In proceedings of first European conference on
earthquake engineering and seismology, paper no. 1122, Geneva Switzerland.
Crowley, H., Pinho, R., and Bommer, J. (2004) A probabilistic displacement based
Vulnerability assessment procedure for earthquake loss estimation, Bulitten of
earthquake engineering, Vol. 2, pp. 173 219.
Crowley, H., Pinho, R., and Bommer, J., and Bird, J. (2005), The impact of epistemic
uncertainty on an earthquake loss mode Earthquake Engineering and Structural
Dynamics Vol.34, No. 4, pp. 1653 1685.
Cuesta, I. and Aschheim, M. A. (2001) Inelastic response spectra using conventional
and pulse R-factors, Journal of Structural Engineering, ASCE, Vol.127, No.9, pp. 1013
1020.
Das, S., and Nau, J.M. (2003). Seismic Design Aspects of Vertically Irregular
Reinforced Concrete Buildings, Earthquake Spectra, Vol. 19, No. 3, pp. 455-477.
De- Stefano, M. and Pintucchi, B, (2002), A model for analyzing inelastic seismic
response of planirregular building structures, In proceedings of 15th ASCE Engg.
Mech. Conf. June 2-5, 2002, Columbia University, New York, NY.
Decanini, I. D., Bruno, S., and Mollaioli, F. (2002) Earthquake damage estimation using
response indices, In proceedings of 4th international conference on Structural dynamics
EURODYN 2002, Grundmann, H. and Schueller, G. I. eds., Munich, Germany, A. A.
Balkema, Rotterdam, 2002, Vol.2, pp.1335:1340.
De-la-Colina, J., Acua, Q., Hernndez, A. and Valds, J. (2007). Laboratory Tests of
Steel Simple Torsionally Unbalanced Models, Earthquake Engineering & Structural
Dynamics, Vol. 36, No.7, pp. 887 907.
De-la-Colina, J. (1999), Effects of torsion factors on simple non linear systems using
fully bidirectional analysis, Earthquake Engineering and Structural Dynamics, Vol.28,
pp. 691-706.

307

De-la-Colina, J. (2003), Assessment of design recommendations for torsionally


unbalanced multistory buildings, Earthquake Spectra, Vol. 19, pp. 4766.
Delhi earthquake safety initiative for lifeline buildings: a project for model seismic
evaluation

and

retrofitting

of

five

life-line

buildings

in

Delhi,

/http://www.quakesafedelhi.net/S, Jan. 23,2007.


deteriorating systems, Earthquake Engineering and Structural dynamics, Vol.33, No.1,
pp. 69 88.
Di Pasquale , E., and Cakmak, A.S. (1987), Detection and assessment of seismic
structural damage, Technical Report NCEER 87 0015, National Center for
earthquake engineering Research , State University of New York, Buffalo NY.
Di Pasquale , E., and Cakmak, A.S. (1988), Identification of servicibility limit state and
detection of seismic structural damage, Technical Report NCEER 87 0022, National
Center for earthquake engineering Research , State University of New York, Buffalo NY.
Di Pasquale , E., and Cakmak, A.S. (1990), Relationships between global damage
indices and local stiffness degradation, Journal of Structural Engineering, ASCE,
Vol.116, No.5, pp.1440 1456.
Dinh Van Thuat. (2011), Story strength demands of irregular frame buildings under
strong

earthquakes,

The

Structural

Design

of

Tall

and

Special

Buildings

DOI.10.1002/tal.713.
DM (2008), Technical rules for constructions Ministry of Transport and infrastructures,
Italy
Dogangun, A. (2004) Performance of reinforced concrete buildings during the May 1,
2003 Bingo l Earthquake in Turkey, Engineering Structures, Vol.26, pp. 841 856.
Dogan, M. Earthquake analysis of buildings. Eskisehir Osmangazi University. Publication
no: 143. ISBN: 978-975-7936-52-7.
Doangn, A., & Livaolu, R. (2006), A comparative study of the design spectra defined
by Eurocode 8, UBC, IBC and Turkish Earthquake Code on R/C sample buildings,
Journal of Seismology, Vol.10, No.3, pp.335 351.

308

Dolsek, M., and Fajfar, P., (2004) Inelastic spectra for infilled reinforced concrete
frames, Earthquake Engineering and Structural Dynamics, Vol. 33, No.15, pp. 1395
1416.
Dolsek, M., and Fajfar, P., (2005), Simplified non-linear seismic analysis of infilled
reinforced concrete frames Earthquake Engineering and Structural dynamics, Vol. 34,
pp.4966
Dolek, M., and Fajfar, P. (2007) Simplified probabilistic seismic performance
assessment of plan-asymmetric buildings, Earthquake Engineering and Structural
Dynamics, Vol.36, No.13, pp. 2021-2041.
Dolsek, M., and Fajfar, P. (2008), The effect of masonry infills on the seismic response
of a four-storey reinforced concrete frame - a deterministic assessment, Engineering
Structures, Vol. 30, No.7, pp. 1991-2001
Dolek, M., and Fajfar, P. (2008a) The effect of masonry infills on the seismic response
of a four-storey reinforced concrete frame - a deterministic assessment, Engineering
Structures, Vol.30, No.7, pp. 1991 2001.
Dolsek, M. (2010), Development of computing environment for the seismic performance
assessment of reinforced concrete frames by using simplified nonlinear models. Bulletin
of Earthquake Engineering, Vol.8, No.6, pp. 1309-1329.
Duan, X.N., and Chandler, A.M. (1991). Seismic torsional response and design
procedures for a class of setback frame buildings, Earthquake Engineering and
Structural Dynamics, Vol.24, pp.761777.
Duan, X.N., and Chandler, A.M. (1995), Seismic torsional response and design
procedures for a class of setback frame buildings, Earthquake Engineering and
Structural Dynamics , Vol.24, pp. 761777.
Dunand F, Bard P-Y, Chatelain JL, Guguen P, Vassail T, Farsi MN (2002) Damping
and frequency from random method applied to in situ measurements of ambient
vibrations. Evidence for effective soil structure interaction. In proceedings of 12 th
European conference on earthquake engineering, paper 869
Durmus.A, Dogangun., A, Husem M., Pul, S. (1999), Preliminary report on August
17 1999 Kocaeli (Izmit) earthquake, Turkish Chamber of Civil Engineers-Bureau of
Trabzon. September. 1999. (in Turkish).

309

Durrani, A.J., Elnashai, A.S., and Hashash Youssef., MA, Kim Sung Jig., Masud, A.
(2008), The Kashmir earthquake of October 8, A Quick Look Report Mid-America
Earthquake Center, University of Illinois at Urbana-Champaign; 2008. (Chapter 1)
Dutta, S.C., and Das, P.K. (2002). Validity and applicability of two simple hysteresis
models to asses progressive seismic damage in R/C asymmetric buildings, Journal of
Sound and Vibration, Vol.257, No.4, pp. 753 777.
Dym, O.L., and Williams, H.E. (2003) Estimating fundamental

frequencies of tall

buildings, Journal of Structural Engineering, ASCE, Vol.133, No.10, pp. 1-5.


EAK(2000), Greek Seismic Code - Organization of Seismic Planning and Protection,
Ministry of Public works Greece , Athens, Grece
EC2 (1992). Eurocode 2, Design of concrete structures, Part 1.1: General Rules for
Buildings, European Prestandard ENV 199311/1992, European Committee for
Standardization (CEN), Brussels
EC3 (1992). Design of steel structures, Part 1.1: General Rules for Buildings, European
Prestandard

ENV

199311/1992,

Eurocode

3,

European

Committee

for

Standardization (CEN), Brussels.


EC3 (1992). Eurocode 3, Design of steel structures, Part 1.1: General Rules for
Buildings, European Prestandard ENV 199311/1992, European Committee for
Standardization (CEN), Brussels
EC8 (1989), Design for structures in seismic regions, Part 1. General and building
report, (EUR12266EN), Commission of European committee, Brussels.
EC8 (1993), Design for structures in seismic regions, Part 1. General and building report
(EUR12266EN), Commission of European committee, Brussels.
EC8 (1994),Design provisions for earthquake resistance of structures, Part 2,, European
Standard EN 19981, Stage 51 Draft, . Eurocode 8, European Committee for
Standardization (CEN), Brussels.
EC8 (2004),Design of structures for earthquake resistance, Part 1: General rules,
seismic actions and rules for buildings, European Standard EN 19981, Stage 51 Draft,
. Eurocode 8, European Committee for Standardization (CEN), Brussels.

310

EC8 (2004),Design of structures for earthquake resistance. General rules seismic


actions and rules for buildings (EN 1998-1:2004), European committee for
Standardization, Brussels,
EKOS (2000), Greek code for design of R/C structures, Ministry of Public works Greece
, Athens, Grece
El-Attar, A. G., White, R.N., and Gergely, P. (1997). "Behavior of Gravity Load Designed
Reinforced Concrete Buildings Subjected to Earthquakes." ACI Structural Journal,
Vol.94, No.2, pp. 133 - 145.
Elghazouli, A.Y., Castro, J.M., and Izzuddin, B.A. (2008), Seismic performance of
composite moment-resisting frames, Engineering Structures, Vol.30, No.7, pp.1802
1819.
Ellingwood, B.R. (1980). An Investigation of the Miyagi-ken-oki, Japan Earthquake of
June 12, 1978, U.S. Department of Commerce/National Bureau of Standards, NBS
Special Publication 592.
Elnashai, A. S., and Broderick, B. M. [1996] Seismic response of composite frames-II.
Calculation of behaviour factors, Engineering Structures,Vol.18, No.9 ,pp. 707723.
Elnashai, A.S. and Di Sarno, L. (2008). Introduction to Earthquake Engineering, Wiley,
Chichester, United Kingdom.
Elnashai, A.S., and Broderick, B.M. (1996). Seismic response of composite frames-II.
Calculation of behavior factors, Engineering Structures, Vol.18, No.9, pp.707-723.
Elnashai, A.S., and Mwafy, A.M. (2002). Over strength and force reduction factors of
multistorey reinforced-concrete buildings, The Structural Design of Tall and Special
Buildings, Vol. 11, No. 5, pp.329351.
Elnashai, A.S., Gencturk, B., Kwon, O.S., Al-Qadi, I.L., Hashash, Y., Roesler, J.R., Kim,
S.J., Jeong, S.H., Dukes, J. and Valdivia, A. (2010), The Maule (Chile) Earthquake of
February 27, 2010: Consequence Assessment and Case Studies, Mid-America
Earthquake (MAE) Center Report, CD-Release 10-04, Department of Civil and
Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, IL, USA.
Elwood, K. J., and Eberhard, M.O. (2006), Effective stiffness of reinforced concrete
columns, PEER Research digest (2006 -1 ): pp. 1-5.

311

Elwood, K. J., and Moehle, J.P. (2003), Shake table tests and analytical studies on the
gravity load collapse of reinforced concrete frames. Rep. 2003/01, Pacific Earthquake
Engineering Research Center, Univ. of California, Berkeley, California.
Elwood, K. J., and Moehle, J.P.(2008), Dynamic collapse analysis for a reinforced
concrete frame sustaining shear and axial failures, Vol.37,No.7, pp.991 1012. pp. 1-5.
Elwood, K. J., Matamoros, A.B., Wallace, W.J., Lehman, D.E., Heintz, J.A., Mitchell,
A.D.,Moore, M.A.,Valley, M.T.,Lowes, L.N., Comartin, C.D, and Moehle, J.P. (2007),
Update to ASCE/SEI 41 Concrete provisions, Earthquake spectra, Vol.21, No.1, pp. 71
89.
Enomoto, T., Navarro, M., Snchez, F.J., Vidal, F., Seo, K., Luzn, F., Garca, J.M.,
Martn, J., Romacho, M.D. (1999), Evaluacin del comportamiento de los edificios en
Almera mediante el anlisis del ruido ambiental 1a Asamblea Hispano-Lusa.
Aguadulce (Almera, Spain), 9-13 / Febrero de 1998. CD-ROM. ISBN, 84-95172-10-0
Enomoto, T., Schmitz, M., Abeki, N., Masaki, K., Navarro, M., Rocavado, V., Sanchez, A.
(2000), Seismic risk assessment using soil dynamics in Caracas, In proceedings of 12 th
world conference on earthquake engineering, Venezuela.CD-ROM
EQE. (1995), An summary report of the January 17, 1995 Kobe Earthquake,
hhttp://www.eqe.com/ publications/kobe/kobe.htmi, April 1995.
Espinoza, F. (1999), Determinacin de las caractersticas dinmicas de estruturas,
Tesis Doctoral, Universidad Politcnica de Catalunya
Esteva L., and Ruiz , S.E. (1989), Seismic failiure rate of multistory frames, Journal of
Structural Engineering, ASCE, Vol.115, No.2, pp.268 283.
Esteva, L. (1987), Summary of October 1, 1987 Whittier Califronia earthquake: A EQE
Quick look report, EQE.
Esteva, L. (1992). Nonlinear Seismic response of Soft- First Story Buildings Subjected
to Narrow Band Accelerograms, Earthquake Spectra, Vol. 8, pp.373-389.
E-Tabs (2009), Integrated software for structural analysis and design. Version
9.0.Berkeley (California).Computers & Structures, Inc.; 2007.

312

Eyidogan, H., Tuysuz, O., Akyuz S, O., keler, A. (2003), May 1, 2003 Bingo l
Earthquake.

ITU

Eurasia

Institute

of

Earth

Sciences,

hhttp://

www.eies.itu.edu.tr/Deprem/Bingol.PDFi, May 12, 2003 (in Turkish). Chapter 1


Faccioli, E., Pessina, V., Calvi, G. M. and Borzi, B. [I9991 "A study on damage scenarios
for residential buildings in Catania," Journal of seismological research letters, Vol.3, pp.
327 343.

Fajfar P, Krawinkler H (eds) (1997) Seismic designmethodologies for the next generation
of codes. Balkema, Rotterdam
Fajfar, P. (1999) "Capacity spectrum method based on inelastic demand spectra",
Earthquake Engineering and Structural dynamics, Vol. 28, No. 9, pp. 975 993.
Fajfar, P., Magliulo, G., Maruic, D. and Peru, I. (2002), Simplified non-linear analysis
of asymmetric buildings, In Proceedings of the third European workshop on the seismic
behaviour of irregular and complex structures, CD ROM, Florence, September 2002.
Fajfar, P., Marusic, D., and Perus, I. (2005), Torsional effects in the pushover-based
seismic analysis of buildings, Journal of Earthquake engineering, Vol.9, No.6, pp.831854.
Fardis, M.N., and Biskinis, D.E. (2003), Deformation capacity of RC members, as
controlled by flexure or shear, Otani symposiam.
FEMA (1988). NEHRP recommended provisions for development of seismic regulations
of New buildings, 1988 edition, Earthquake hazards Reduction Series 17 , Federal
Emergency management Agency (FEMA), Building seismic safety council. Washington
D.C. Part 1: Provisions, 1988.
FEMA (2003), HAZUS MH MRI Technical manual, Washington DC, FEMA.
FEMA (Federal Emergency Management Agency)

(I999) Earthquake Loss Estimation

Methodology: User's Manual, Federal Emergency Management Agency, Washington


DC, U.S.A.
FEMA (Federal Emergency Management Agency) (1997), Building Seismic Safety
Council. NEHRP guidelines for the seismic rehabilitation of buildings, Report No.
273,Washington (DC): Federal Emergency Management Agency.

313

FEMA (Federal Emergency Management Agency) (1997), Building Seismic Safety


Council. NEHRP guidelines for the seismic rehabilitation of buildings, Report No.
274,Washington (DC): Federal Emergency Management Agency.
FEMA (Federal Emergency Management Agency) (2000 a), Recommended seismic
design criteria for new steel moment frame buildings. Report No. 350, Washington, D.C.
FEMA (Federal Emergency Management Agency) (2000 b) Prestandard and
commentary for the seismic rehabilitation of buildings. Report No. 356, prepared by the
American Society of Civil Engineers for the Federal Emergency Management Agency,
Washington, D.C.
FEMA (Federal Emergency Management Agency) (2000),Prestandard and commentary
for the seismic rehabilitation of buildings, Report FEMA 356. Washington (DC). 2000.
FEMA. (1973), NEHRP guidelines for the seismic rehabilitation of buildings, Report No.
273, prepared by the Applied Technology Council for the Building Seismic Safety
Council, Washington, D.C.
FEMA. (Federal Emergency Management Agency). (1998), NEHRP recommended
provisions for seismic regulations for new buildings and other structures: Part Iprovisions. Report No. 302, 1997 Ed., prepared by the Building Seismic Safety Council
for the Federal Emergency Management Agency, Washington, D.C.
Feng Fu (2009), Progressive collapse analysis of high-rise building with 3-D finite
element modeling method, Journal of constructional steel research, Vol.65, No.6,
pp.1269 1278.
Ferhi, A., & Truman, K. Z. (1996), Behaviour of asymmetric building systems under a
monotonic loadI,Engineering structures, Vol. 18, No.2,pp. 133-141.
Fernandez, J. (1983), Earthquake Response Analysis of Buildings Considering the
effects of Structural Configuration, Bulletin of the International Institute of Seismology
and Earthquake Engineering (Tokyo, Japan, Nov 1983), Vol.19, pp.203-215.
Filiatrault, A., Lachapelle, E., and Lamontagne, P. (1998). "Seismic performance of
ductile and nominally ductile reinforced concrete moment resisting frames II. Analytical
Study." Canadian Journal of Civil Engineering, Vol.25, No.2, pp.: 342-352.

314

Foliente, G. (1995), Hysteresis modeling of wood joints and structural systems, Journal
of Structural Engineering, ASCE, No.121, pp. 1013-1022.
Fragiadakis, M., Vamvatsikos, D., and Papadrakakis, M. (2005), Evaluation of the
influence of vertical Stiffness irregularities on the seismic response of a 9-story steel
frame, In Proceedings of the 4th European workshop on the seismic behavior of
irregular and complex structures, CD ROM. Thessaloniki, August 2005.
Fragiadakis, M., Vamvatsikos, D., and Papadrakakis, M. (2006), Evaluation of the
Influence of Vertical Irregularities on the Seismic Performance of a Nine-Storey Steel
Frame, Earthquake Engineering & Structural Dynamics, Vol. 35, No. 12, pp. 1489-1509.
Franchin, P. and Pinto, P. (2012). Method for Probabilistic Displacement-Based Design
of RC Structures. Journal of Structural Engineering, ASCE, Vol. 138, No.5, pp. 585
591.
Fraser, S., Raby, A., Pomonis, A., Goda, K., Chian, S.C., and Macabuag, J. (2013),
Tsunami damage to coastal defences and buildings in the March 11th 2011 Mw9.0
Great East Japan earthquake and tsunami, Bulletin of earthquake engineering, Vol.11,
pp. 205 239.
Freeman, S. A. (I998), "The capacity spectrum method as a tool for seismic design," In
Proceedings of the Eleventh European Conference on Earthquake Engineering, A. A.
Balkema, Rotterdam, 1998.
Freeman, S.A., Nicoletti, J.P., and Tyrrell,J.V. (1975), Evaluation of existing buildings for
seismic risk - A case study of Puget Sound Naval Shipyard, Bremerton, Washington," In
Proceedings of the U.S. National Conference on Earthquake Engineering, Earthquake
Engineering Research Inst., Oakland, California, pp. 113-122.
Fujii, K., Nakano, Y., and Snada, Y. (2004), A simplified nonlinear analysis procedure
for single-story Asymmetric buildings, Journal of Japan Association for Earthquake
Engineering, Vol.4, No.2, pp.1-20.
Fukada, Y. (1969), Study on the restoring force characteristics of reinforced concrete
buildings (in Japanese), In Proceedings, Kanto Branch Symposium, Architectural
Institute of Japan, No. 40, 1969, pp.121-124.

315

Gallipoli, M.R., Mucciarelli, M., and Vona, M. (2009), Empirical estimate of fundamental
frequencies and damping for Italian buildings, Earthquake Engineering and Structural
Dynamics, Vol.38, pp. 973 988.
Ger, J.F., Cheng, F.Y., and Lu., L.W. (1993), Collapse behavior of Pino Suarez Building
during 1985 Mexico City earthquake, Journal of Structural Engineering, ASCE, Vol.119,
No.3, pp. 852 870.
Ghersi, A., and Rossi, P.P. (2001), Influence of bidirectional seismic excitation on
inelastic response of single storey plan irregular systems, Vol.23, No.6, pp.579-591.
Ghersi, A., and Rossi, P.P. (2006), Influence of Design procedures on Bi-eccentric Plan
asymmetric systems, Structural design of Tall and special buildings, Vol.15, pp.467-480.
Ghobarah, A., Abou-Elfath, H., and Biddah, A. (1999), Response-based damage
assessment of structures, Earthquake Engineering and Structural dynamics, Vol.28,
No.1, pp. 79 104.
Ghobarah, A., Saatcioglu, M., and Nistor, I. (2006), The impact of the 26 December
2004 earthquake and tsunami on structures and infrastructure, Engineering structures,
Vol.28, No.2, pp.312 326.
Ghrib, F., and Mamedov, H. (2004) Period formulas of shear wall buildings with
flexible bases, Earthquake Engineering and Structural dynamics, Vol. 33, No.3, pp. 295
314.
Gioncu, V., and Petcu, D. (1997), Available rotation capacity of wide-flange beams and
beam columns: part 1. Theoretical approaches, Journal of Construction Steel Research,
43, pp.161217.
Glaister, S., and Pinho R. (2003) Development of simplified deformation based method
for seismic vulnerability assessment, Journal of Earthquake engineering, Vol.7, special
issue 1, pp.107-140.
Goel, R. K., and Chopra, A. K. (1997), Period formulas for moment resisting frame
buildings, Journal of Structural Engineering, ASCE, Vol.123, No.11, pp. 1454 1461.
Goel, R.K. (2005), Performance of buildings during the January 26, 2001 Bhuj
Earthquake, hhttp://ceenve. ceng.calpoly.edu/goel/indian_ eqk/index.htmi, December
2003.

316

Goel, R.K. and Chopra, A.K. (1991). Effects of Plan Asymmetry in Inelastic Seismic
Response of One-story Systems, Journal of Structural Engineering, 117, 1492-1513.
Goel, R.K., and Chopra, A.K. (2004), Evaluation of modal and FEMA pushover
analyses: SAC buildings, Earthquake Spectra, Vol.20, No. 1, pp. 225 254.
Goel, R.K., and Chopra, A.K. (2005), Extension of modal pushover analysis to compute
member forces. Earthquake Spectra, Vol.21, No. 1, pp. 125 139.
Gosain, N.K., Brown, R.H., and Jirsa, J.O. (1977), Shear requirements for load
reversals on RC members, Journal of Structural Engineering, ASCE, Vol.103, No.7, pp.
1461 1476.
Grecea, D., and Dubina, D. (2003) Partial q-factor values for performance based design
of MR frames, Behaviour of Steel Structures in Seismic Areas, STESSA 2003,
Mazzolani, F. Ed. Naples, Italy, pp. 2329.
Griffith, M.C., Kawano, A., and Warner, R.F. (2002), Towards a direct collapse-load
method of design for concrete frames subjected to severe ground motions, Earthquake
Engineering and Structural dynamics, Vol. 31, No. 10, pp. 1879 1888.
Guler, K., Yuksel, E., and Kocak, A. (2008), Estimation of the fundamental vibration
period of existing RC buildings in Turkey utilizing ambient vibration records, Journal of
earthquake engineering, Vol.12, pp. 140 150.
Gunay, M.S., and Sucuolu, H. (2009), Predicting the Seismic Response of Capacity
Designed Structures by Equivalent Linearization, Journal of earthquake engineering,
Vol.13, No.5, pp.623 449.
Gupta, B., and Kunnath, S. K. (2000). Adaptive spectra-based pushover procedure for
seismic evaluation of structures,Earthquake spectra, Vol. 12, No. 2, pp. 367 391.
Hall, J.F. (1994), Northridge Earthquake January 17, 1994: Preliminary Reconnaissance
Report. Report No. 94 01, Earthquake Engineering Research Institute, Oakland,
California, U.S.A
Hall, J.F., Heaton, T.H., Halling, M. W., and Wald, D. J. (1995) Near source ground
motion and its effects on flexible buildings, Earthquake spectra, Vol.11, No.4, pp.569
605.

317

Hamburger, R.O. (1997), A framework for performance-based earthquakeresistive


design. The EERC-CUREE Symp. in Honor of Vitelmo V. Bertero, UCB/EERC Rep. No.
97/05, Earthquake Engineering Research Center, Univ. of California, Berkeley, California
, U.S.A., 101108.
Han, S.W., Moon, K.H., and Chopra , A.K. (2010), Application of MPA to estimate
probability of collapse of structures, Earthquake engineering and structural dynamics,
DOI: 10.1002/eqe.992.
Hanganu, A.D., Onate, E., and Barbat, A.H. (2002), A finite element methodology for
local/global evaluation in civil engineering structures, Computers & Structures, Vol.80,
pp. 1667-1680.
Haselton, C.B., and Deierlein, G.G. (2007), Assessing seismic collapse safety of
modern reinforced concrete frame buildings, PEER Rep. 2007/08. PEER Center, Univ.
of California, Berkeley, CA.
Haselton, C.B., Liel, A., Deierlein, G.G., Dean, B.S., and Chou, J.S. (2011), Seismic
Collapse Safety of Reinforced Concrete Buildings. I: Assessment of Ductile Moment
Frames, Journal of Structural Engineering, ASCE, Vol. 137, No. 4, pp.481-491.
Haselton, C.B., Liel, A., Deierlein, G.G., Dean, B.S.,and Chou, J.S. (2011), Seismic
Collapse Safety of Reinforced Concrete Buildings. II: Assessment of Ductile Moment
Frames. Journal of Structural Engineering, ASCE, Vol.137, No.4, pp.492-502.
Hatzigeorgiou, G.D., and Beskos, D.E. (2009), Inelastic displacement ratios for SDOF
structures subjected to repeated earthquakes Engineering structures, Vol.31, No.11,
pp.2744 2755.
Hatzigeorgiou, G.D., Beskos, D.E. (2007), Direct damage controlled design of concrete
structures, Journal of Structural Engineering, ASCE, Vol.133, pp.205215.
Hatzigeorgiou, G.D.. and Liolios, A.A. (2010), Nonlinear behaviour of RC frames under
repeated strong ground motions, Soil dynamics and earthquake engineering, Vol. 30,
No.10, pp.1010 1025.
Hisada, T., Nakagawa, K., and Izumi, M., (1962), Earthquake response of structures
having various restoring force characteristics, In proceedings of Japan National
conference on Earthquake Engineering, 1962, pp. 63 68.

318

Hong, L. and Hwang (2000) Empirical formula for fundamental periods of vibration for
Reinforced concrete buildings in Taiwan Earthquake Engineering and Structural
Dynamics 2000, 29:327-337.
http://Mid-America Earthquake Engineering Center, University of Illinois at UrbanaChampaign)
http://www.ngde.noaa.gov/seg/hazard/slideset/earthquake
http:/en.wikipedia.org
Huang, S. C., and Skokan, M.J. (2002) Collapse of the Tungshing building during the
1999 Chi-Chi earthquake in Taiwan, In proceedings of 7th U.S. National conference on
earthquake engineering, Boston.
Hueste, M.D., Bai, J.W. (2006). Seismic Retrofit of a Reinforced Concrete Flat-Slab
Structure: Part II - Seismic Fragility Analysis. Engineering Structures, Vol. 29, No.6,
pp.1178-1188.
Humar, J.L., and Wright, E.W. (1977). Earthquake Response of Steel-Framed Multistory
Buildings with Set-Backs, Earthquake Engineering & Structural Dynamics, Vol. 5, No. 1,
pp. 15-39.
Humar, J.M. (1990), Dynamics of structures, Prentice hall, NY
Humar, J.M., Lau, D., Pierre, J.R. (2001), Performance of buildings during the 2001
Bhuj Earthquake , Canadian journal of civil engineering, Vol. 28, pp. 979991.
Ibarra, L.F. (2003), Global collapse of frame structures under seismic excitation,
Doctoral dissertation (TR 152), Stanford university , U.S.A.
Ibarra, L.F., and Krawinkler, H. (2004), Global collapse of deteriorating MDOF systems,
In proceedings of 13th World Conf. on Earthquake Engineering, Vancouver, B.C.,
Canada, Paper No. 116.
Ibarra, L.F., Medina, R.A., and Krawinkler, H. (2005), Hysteretic models that incorporate
strength and stiffness deterioration. Earthquake Engineering and Structural Dynamics,
Vol.34, No.12, pp. 1489 1511.
IBC (2003), International building code 2003, Illiniosis, International code council (ICC),
2002 Inc.

319

IBC (2006), International building code 2006, Illiniosis, International code council (ICC),
Inc.
IS (1987), IS 875 (Part 2), Code of practice for design loads (other than earthquake) for
buildings and structures, Bureau of Indian Standards, New Delhi.
IS (2002), IS 1893 (Part 1)-2002: Indian Standard Criteria for Earthquake Resistant
Design of Structures, Part 1 General Provisions and Buildings (Fifth Revision), Bureau
of Indian Standards, New Delhi.
Iwan, W.D. (1966), A distributed-element model for hysteresis and its steady-state
dynamic response ,Journal of Applied Mechanics, Vol.33, No.42, pp.893 900.
Iwan, W.D. (1973), A model for the dynamic analysis of deteriorating structures , In
proceedings of fifth World Conference in Earthquake Engineering, Rome, Italy, pp. 17821791.
Izzuddin, B.A., Vlassis, A.G., A.Y. Elghazouli, A.Y., and Nethercot, D.A. (2008),
Progressive collapse of multi-storey buildings due to sudden column loss Part I:
Simplified assessment framework, Engineering structures, Vol.30, No.5, pp.1308
1318.
Jain, S. K., Lettis, W. R., Murty, C. V. R., and Bardet, J. P. (2002) Bhuj, India,
Earthquake of January 26, 2001, Reconnaissance Report, Earthquake spectra,
Supplement A to Vol.18.
Jangid, R.S., and Kelly, J.M. (2000), Torsional displacements in base isolated
buildings, Earthquake spectra, Vol. 16, No. 2, pp. 443-454.
Jarernprasert, S., Bazan, E., and Bielak, J. (2008), Inelastic torsional single story
systems, In proceedings of Fourteenth world conference on Earthquake Engineering, ,
Beijing, China.
Jennings, P. C., and Husid, R. (1968), Collapse of yielding structures during
earthquakes. Journal of Engineering Mechanics, ASCE, Vol.94, No.5, pp. 1045 1065.
Jeong, G.D., and Iwan, W.D. (1988), Effect of earthquake duration on damage of
structures, Eartquake engineering and structural dynamics, Vol.16, No.8, pp.1201
1211.

320

Jeong, S.H., and Elnashai, A.H. (2005), Analytical assessment of an irregular RC frame
for fullscale 3d pseudo-dynamic testing part i: analytical model verification, Journal of
earthquake engineering, Vol.9, No.1, pp.95 128.
Jones, P. (2012), Assessment of performance based procedures for Tall buildings A
Doctoral Thesis dissertation submitted to University of California: Irvine.
Jones, P., and Farzin, Z. (2009), Relative safety of high-rise and low-rise steel momentresisting frames in Los Angeles, The Structural Design of Tall and Special Buildings,
Vol. 19, pp.183-196.
Jones, P., and Farzin, Z. (2010), Seismic response of a 40-storey buckling-restrained
braced frame designed for the Los Angeles region, The Structural Design of Tall and
Special Buildings, DOI:10.1002/tal.687.
Kalkan, E., and Chopra , A. (2011), Modal-Pushover-Based Ground-Motion Scaling
Procedure, Journal of Structural Engineering, ASCE, Vol.137, No.3, pp. 298 310.
Kam, W.Y., Pampanin, S., Dhakal, R., Gavin, H.P., and Roeder, C. (2010), Seismic
performance of reinforced concrete buildings in the september 2010 darfield (canterbury)
earthquake, Bulletin of the New - Zealand society for earthquake engineering, Vol.33,
No.4, pp. 340 350.
Kamaris G.S., Hatzigeorgiou, G.D., Beskos D.E. (2009), Direct damage controlled
design of plane steel-moment resisting frames using static inelastic analysis, Journal of
mechanics and Materials, Vol. 4, pp.13751393.
Kamaris, G.S., Hatzigeorgiou, G.D., Beskos, D.E. (2013), A new damage index for
plane steel frames exhibiting strength and stiffness degradation under seismic motion,
Engineering Structures, Vol.46, pp.727736
Kanvinde, A.M. (2003), Methods to evaluate the dynamic stability of structuresShake
table tests and nonlinear dynamic analyses. Winner EERI Annual Student Paper
Competition, Earthquake Engineering Research Institute, Oakland, California, U.S.A.
Kappos, A. J. (1999), Evaluation of behaviour factors on the basis of ductility and
overstrength studies, Engineering structures, Vol.21, pp.823 835.

321

Kappos, A.J., and Panagopoulos, G. (2004). Performance-based seismic design of 3d


r/c buildings using inelastic static and dynamic analysis procedures, ISET Journal of
Earthquake Technology, Paper No.441,Vol.41, No.1, pp.141-158.
Kappos, A.J., and Scott, S.G. (1998), Seismic assessment of an R/C building with
setbacks using nonlinear static and dynamic analysis procedures, In Booth ED (ed)
Seismic design practice into the next century. Balkema, Rotterdam.
Kappos, A.J., and Stefanidou, S. (2010), A deformation based seismic design method
for 3D R/C irregular buildings using inelastic dynamic analysis, Bulletin of Earthquake
Engineering, DOI 10.1007/S10518-009-9170-1.
Karakostasa, C.Z., Athanassiadou, C.J., and Kappos., A.J., Lekidisa, V.A., (2007), Sitedependent design spectra and strength modification factors based on records from
Greece, Soil Dynamics and Earthquake Engineering, Vol. 27, pp. 10121027.plan
Karavasilis, T.L., and Bazeos, N. and Beskos, D.E. (2006) Maximum displacement
profiles for the performance based seismic design of plane steel moment resisting
frames ,Engineering Structures, Vol. 28, No. 1,pp. 9-22.
Karavasilis, T.L., and Bazeos, N. and Beskos, D.E.(2008a). Seismic response of plane
steel MRF with setbacks: Estimation

of inelastic deformation demands, Journal of

Construction and Steel research, Vol.64, pp.644-654.


Karavasilis, T.L., and Bazeos, N. and Beskos, D.E.(2008b). Estimation of seismic
inelastic deformation demands in plane steel MRF with vertical mass irregularities.,
Engineering Structures, Vol.30, pp.3265 3275
Karavasilis, T.L., and

Seo, C.Y. (2011). Seismic structural and non-structural

performance evaluation of highly damped self-centering and conventional systems


Engineering Structures, Vol.33, No.8, pp.2248 2258.
Karavasilis, T.L., Bazeous, N., and Beskos, D.E. (2007), Behavior Factor for
Performance-Based Seismic Design of Plane Steel Moment Resisting Frames, Journal
of Earthquake Engineering, Vol.11, pp.531 559.
Karavasilis, T.L., Kerawala., S., and Hale, E. (2012). Hysteretic model for steel energy
dissipation devices and evaluation of a minimal-damage seismic design approach for
steel buildings Journal of Constructional Steel Research, Vol. 70, pp. 358-367.

322

Karl V. Steinbrugge collection. Earthquake Engineering Research Center, Richmond.


(http://nisee.berkeley.edu/

visual

resources/steinbrugge

collection.html

and

<http://nisee.berkeley.edu/eqiis.html.
Kato, B., Akiyama, H., Suzuki, H., and Fukuzawa, Y. (1973), Dynamic collapse tests of
steel structural models, In proceedings of 5th world

conference on earthquake

engineeeirng, Rome Italy.


Kato, H., Tajiri, S., and Mukai , T. (2010), Preliminary Reconnaissance Report of the
Chile Earthquake 2010, Building research Institute, Japan.
Khandelwal, K., and El Tawil, S. (2007), Collapse Behavior of steel special moment
resisting frame connections, Journal of Structural Engineering, ASCE, Vol.133, No.5,
pp.646 655.
Khashaee, P. (2005) Damage-based seismic design of structures, Earthquake spectra,
Vol.21, No.2, pp. 371 387.
Khoury, W., Rutenberg, A., and Levy, R. (2005), On the seismic response of
asymmetric setback perimeter-frame structures, In Proceedings of the 4th European
workshop on the seismic behavior of irregular and complex structures, CD ROM.
Thessaloniki, August 2005.
Killar, V., and Fajfar, P. (1997), Simple push-over analysis of asymmetric buildings.
Earthquake Engineering and Structural Dynamics, Vol. 26, pp.233249.
Killar, V., and Fajfar, P. (2002), Seismic analysis of eccentric R/C buildings by the N2
methods, In Proceedings of the third European workshop on the seismic behaviour of
irregular and complex structures, CD ROM. Florence, September 2002.
Kim, J., and Choi, H. (2011), Use of rotational friction dampers to enhance seismic and
progressive collapse resisting capacity of structures, The Structural Design of Tall and
Special Buildings, Vol. 20, pp. 515537, Doi 10.1002/tal.563.
Kim, J., and Hong, S., (2011), Progressive collapse performance of irregular buildings,
The Structural Design of Tall Special Buildings, Vol. 20, pp. 721-734.
Kim, J., and Lee, Y.H. (2010), Progressive collapse resisting capacity of tube type of
structures, The Structural Design of Tall and Special Buildings, Vol. 19, pp.761-777. Doi
10.1002/tal.512.

323

Kim, S., and D Amore, E. (1999). Push-over analysis procedure in earthquake


engineering ,Earthquake Spectra, Vol. 115, No. 3, pp. 417 434.
King, S. A., Kiremidjian, A. S., Basoz, N., Law, K., Vucetic, M., Douroudian, M., Olson, R.
A., Eidnger, J. M., Goettel, K. A. and Horner, G. (I997) Methodologies for evaluating the
socio-economic consequences of large earthquakes", Earthquake spectra, Vol.13, pp.
565 584.
Kirac, N., Dogan, M., and Ozbasaran, H. (2009) Failure of weak-storey during
earthquakes, Engineering Failure Analysis, Vol. 18, 572581.
Kirac, N., Dogan, M., Ozbasaran, H. (2011), Failure of weak-storey during earthquakes,
Engineering Failure Analysis, Vol. ,pp. 572581
Kircher, C. A., Nassar, A. A., Kustu, 0. and Holmes, W. T. (1997). "Development of
building damage functions for earthquake loss estimation", Earthquake Spectra, Vol. 13,
No.4, pp. 663-682.
Kobayashi, H, Vidal, F, Feriche, D, Samano, T., and Alguacil, G. (1996). Evaluation of
dynamic behaviour of building structures with microtremors for seismic micro - zonation
mapping, In proceedings of the 11th World conference on earthquake engineering, ,
Acapulco, Mxico
Kobayashi, H., Midorikawa, S., Tanzawa, H., Matsubara, M. (1987). Development of
portable measurement system for ambient vibration test of building, Journal of Structural
Construction Engineering (Transactions of Architectural Institute of Japan), No. 378, pp
4856
Korean Building code (KBCS) (1988), The Ministry of Construction.
Kori, J.G., and Jangid, R.S. (2008), Semi-active friction dampers for seismic control of
structures, Smart Structures and Systems, Technopress, Vol. 4, No.4, 493-515
Kose, M.M. (2008), Parameters affecting the fundamental time period of RC buildings
with infill walls Engineering Structures, Vol. 31, pp.93-102.
Kowalsky, M. J., Priestley, M. J. N. and MacRae, G.A. (I995). "Displacement-based
design of RC bridge columns in seismic regions", Earthquake Engineering and Structural
Dynamics, Vol.24, No. 12, pp. 1623 1643.

324

Krlik, J., and Krlik Jr, J. (2009), Seismic analysis of reinforced concrete frame-wall
systems considering ductility effects in accordance to Eurocode, Engineering
Structures, Vol.31, No.12, pp. 2865 2872.
Kratzig, W.B., Meyer, I.F., and Meskouris, K. (1989), Damage evolution in reinforced
concrete members under cyclic loading, In proceedings of 5th international conference
on safety and reliability (ICOSSAR 89) , San Francisco CA, Vol II, pp. 795 802.
Krawinkler, H., and Seneviratna, G.D.P.K. (1998), Pros and cons of a pushover analysis
of seismic performance evaluation , Engineering Structures, Vol.20, pp. 452 464.
Krawinkler, H., Medina, R., and Alavi, B. (2003). Seismic drift and ductility demands and
their dependence on ground motions, Engineering Structures, Vol.25, No. 5, pp. 637
653.
Krawinkler, H., Zohrei M (1983), Cumulative damage in steel structures subjected to
earthquake ground motions, Computers and Structures, Vol. 16, pp.531541
Kreslin, M. and Fajfar, P. (2010) Seismic evaluation of existing complex RC building
Bulletin of Earthquake engineering, Vol.8, pp.363 385.
Kreslin, M. and Fajfar, P. (2012) The extended N2 mehod considering higher mode
effects both in plan and elevation Bulitten of Earthquake engineering, Vol.10, pp.695
715.
Kreslin, M. and Fajfar, P.(2011) Seismic evaluation of existing complex RC building
Earthquake Engineering and Structural dynamics, Volume 40, No.14,pages 1571
1589, November 2011.
Kukreti, A.R., and Abolmaali, A.S. (1999). Moment-rotation hysteresis behavior of top
and seat angle steel frame connections, Journal of Structural Engineering, ASCE,
Vol.125, No.8, pp. 810-820.
Kumar, M., Stafford, P.J., and Elghazouli, A.Y. (2013a) Influence of ground motion
characteristics on drift demands in steel moment frames designed to Eurocode
8,engineering structures, Engineering Structures, Vol. 52, pp. 502517
Kumar,M., Stafford, P.J., and Eighazouli, A.Y. (2013b), Seismic shear demands in multistorey steel frames designed to Eurocode 8, Engineering Structures, Vol.52, pp.69 87.

325

Kunnath, S.K., Hoffman .G., Reinhorn, A.M., and Mander, J.B. (1995), Gravity load
designed reinforced concrete buildings II, Evaluation and detailing enhancements, ACI
Structural Journal, Vol.92, No.3, pp.355 366.
Kunnath, S. K., Reinhorn, A.M., and Park, Y.J. (1990), Analytical modeling of inelastic
response of R/C structures, Journal of Structural Engineering, ASCE, Vol. 116, No.4 pp.
996-1027.
Kunnath, S.K., and Kalkan, E. (2004), Evaluation of seismic deformation demands using
nonlinear procedures in multistory steel and concrete moment frames, ISET Journal of
Earthquake Technology, Vol. 41, No. 1, pp.159-181.
Kunnath, S.K., Reinhorn, A.M., and Abel, J.F. (1991), Computational tool for evaluation
of seismic performance of reinforced concete buildings, Computers and Structures,
Vol.41, No.1, pp.157 173.
Kunnath, S.K., Reinhorn, A.M., and Lobo, R.F. (1992), IDARC version 3.0: A program
for the inelastic damage analysis of reinforced concrete structures, Report No. NCEER,
92-0022, National Center for Earthquake Engineering Research, State University of New
York at Buffalo, Buffalo, N.Y.
Kurama, Y.C.S.P., Pessiki, S.P., and Sauce , R., and Wu, S. (1994), Seismic behavior
of non ductile reinforced concrete structures, ASCE Structural Congress.
Ladinovic, D. (2008), Non-linear analysis of Asymmetric in plan buildings, Architecture
and Civil Engineering, Vol.6, No.1, pp.25-35.
Lagomarsino S (1993) Forecast models for damping and vibration periods of buildings. J
Wind Eng Ind Aerodyn 48:221239
Lee L, Chang K, Chun Y. (2000). Experimental formula for the fundamental period of RC
buildings with shear wall dominated systems. Structural Design of Tall and special
Buildings, Vol.9, No.4, pp. 295-307.
Lee, K., and Foutch, D.A. (2002). Performance evaluation of new steel frame buildings
for seismic loads, Earthquake Engineering and Structural Dynamics, Vol.31, No.3,
pp.653 670.

326

Lee, K., and Stojadinovic, B. (2004), Low-cycle fatigue limit on seismic rotation capacity
for US Steel moment connections, In Proceedings, 13th World Conference on
Earthquake Engineering, August 2004, Paper No. 90.
Leiva, G., and Wiegand, W. (1996). Analysis of a collapsed building using the
displacement seismic design approach, In proceedings of 11 th world conference on
earthquake engineering, Acapulco, Mexico, Paper No. 1512.
Lemaitre, J. (1992), A course on damage mechanics, Springer, Berlin
Lestuzzi, P., and Badoux, M. (2003) An experimental confirmation of the equal
displacement rule for RC structural walls, In proceedings of the first symposium on
concrete structures in seismic regions. 2003.
Li, W., and

Li, Q. (2010), Performance-based seismic design of complicated tall

building structures beyond the code specification, The Structural Design of Tall and
Special Buildings, Doi. 10.1002/tal.637.
Li, Y., Lu , X.,Guan, H., Ye, L. (2011), An improved tie force method for progressive
collapse resistant deisgn of reinforced concrete structures Engineering Structures,
Vol.33, pp.2931 2942.
Li, Y.R., and Jirsa, J.O. (1998), Nonlinear analyses of an instrumented structure
damaged in the 1994 Northridge earthquake, Earthquake Spectra, Vol.14, No.2, pp. 265
283.
Liang Su & Jitao Shi (2012) Displacement-Based Earthquake Loss Assessment
Methodology Adopting Two-Dimensional Equivalent Linearization Approach, Journal of
Earthquake Engineering, Vol. 16, No.3.
Liel, A.B. (2008), Assessing the collapse risk of Californias existing reinforced concrete
frame structures: metrics for seismic safety decisions, Doctoral Dissertation (2008),
Stanford University, U.S.A
Lieping, Y., Xinzheng, L., and Li Yi, l. (2010), Design objectives and collapse prevention
for building structures in mega-earthquake, Earthquake engineering and Engineering
Vibration, Vol.9, 99.189 199.
Lignos, D.G., and Krawinkler, H. (2009), Sidesway collapse of deteriorating structural
systems under seismic excitations, Report No. TB 172, John A. Blume Earthquake

327

Engineering Research Center, Department of Civil and Environmental Engineering.


Stanford University: Stanford, CA
Lignos, D.G., and Gantes, C.J. (2005), Seismic demands for steel braced frames with
stiffness irregularities based on modal pushover analysis, In Proceedings of the 4th
European workshop on the seismic behavior of irregular and complex structures, CD
ROM. Thessaloniki, August 2005.
Lobo, R.F. (1994), Inelastic dynamic analysis of reinforced concrete structures of three
dimensions, Ph.D. Dissertation, Department of Civil Engineering, State University of
New York, Buffalo, N.Y.

Lu, L., and Shi, J. (2012), Displacement-Based Earthquake Loss Assessment


Methodology Adopting Two-Dimensional Equivalent Linearization Approach, Journal of
earthquake engineering, Vol.16, No.3, pp. 425-442
Lu, X., Su, N., and Zhou, Y. (2011), Nonlinear time history analysis of a super-tall
building with setbacks in elevation, The structural design of tall and special buildings ,
DOI: 10.1002/tal.717.
Lu, X., Chen, Y., and Mao, Y. (2012), Shaking table model test and numerical analysis
of a supertall building with high-level transfer storey Structural design of tall and special
buildings, Vol.21, No.10, pp. pp. 699723.
Luchinni, A., Monti, G., and Kunnath, S. (2011), Nonlinear response of two way
asymmetric single storey bulding under biaxial excitation, Journal of structural
Engineering, ASCE, Vol.137, pp.34-40.
Lumantarna, E., Lam, N., and Wilson, J. (2011), Simple Seismic Design and
Assessment of Buildings Incorporating The Displacement Controlled Phenomenon, In
proceedings of Australian Earthquake Engineering Society 2011 Conference, 18-20
November, Barossa Valley, South Australia.
MacRae, G.A. (1994). P -Delta effects on single-degree-of-freedom structures in
earthquakes, Earthquake spectra, Vol. 10, No. 3, pp. 539 568.
Magliulo, G., Capozzi, V., and Ramasco, R. (2012), Seismic performance of R/C
frames with overstrength discontinuities in elevation, Bullitten of Earthquake
Engineering, Vol.10, pp.679 694.

328

Magliulo, G., Ramasco, R., and Realfonzo, R. (2002). A critical review of seismic code
provisions for vertically irregular frames, In Proceedings of the third European workshop
on the seismic behavior of irregular and complex structures, CD ROM. Florence,
September 2002.
Mahin, S.A., and Bertero, V.V. (1972), Rate of loading effect on uncracked and repaired
reinforced concrete members, Report EERC No. 73-6, Earthquake Engineering
Research Center, University of California at Berkeley, Berkeley, CA.
Maison, B.F., and Bonowitz, D. (2004), Discussion of Seismic performance evaluation
of pre-Northridge steel frame buildings with brittle connections, Journal of structural
engineering ASCE, Vol.130, No. 4, pp. 690 691.
Maison, B.F., and Hale, T.H. (2004). Case study of a Northridge welded steel momentframe building having severed columns, Earthquake spectra, Vol. 20, No. 3, pp. 951
973.
Malhotra, P. (2002) Cyclic-demand spectrum, Earthquake Engineering and Structural
Dynamics, Vol. 31, No.7, pp. 1441 1457.
Mander, J.B., Priestley, M.J.N., and Park, R. (1988), Theoretical stress-strain model for
confind concrete, Journal of Structural Engineering, Vol.114,No.8, pp. 1804-1826.
Manfredi, G. (2001), Evaluation of seismic energy demand, Earthquake Engineering
and Structural Dynamics, Vol. 30, No. 4, pp.485499.
Martin, S.C., and Villaverde, R. (1996), Seismic collapse of steel frame structures, In
proceedings of 11th World conference on earthquake engineering, Acapulco, Mexico,
paper No. 475.
Maruic, D., and Fajfar, P. (2005), On the inelastic seismic response of asymmetric
buildings under bi-axial excitation, Earthquake Engineering and Structural Dynamics,
Vol. 34, pp. 943963.
Masi, A., and Vona, M. (2010) Experimental and numerical evaluation of the
fundamental period of undamaged and damaged RC framed buildings Bullitten of
earthquake engineering, Vol.8, pp. 643656
MATLAB. (1997), The language of technical computing, Version 5. Natick, Mass, The
Mathworks Inc, 1997.

329

Mazzolini, F.M., and Piluso, V. (1996), Theory and design of seismic resistant steel
frames, London, Newyork: FN & SPON an imprint of Chapman and Hall 1996.
McCabe S.L., and Hall, W.J. (1989), Assessment of seismic structural damage, Journal
of structural engineering, ASCE, Vol.115, pp.2166 2183.
McGuire, R. K. (2001), "Deterministic vs. probabilistic earthquake hazard and risks", Soil
dynamics and earthquake engineering, Vol. 21, pp. 377 384.
McKenna, F. (1997), Object oriented finite element programming frameworks for
analysis, algorithms and parallel computing, Ph.D. Dissertation, University of California
at Berkeley, Berkeley, CA.
Medhakar. M.S., and Kennady, D.J.L. (2000), Displacement based design of buildings
Theory, Engineering Structures, Vol. 22,No.3, pp.201-209.
Medhakar. M.S., and Kennady, D.J.L. (2000), Displacement based design of buildings
Application, Engineering Structures, 22 , No.3, pp.210-221.
Medina, R.A., and Krawinkler, H. (2005). Strength design issues relevant for the seismic
design of moment-resisting frames, Earthquake spectra, Vol.21, No. 2, pp. 415 439.
Medina, R.A., and Krawinkler, H., (2003), Seismic demands for non-deteriorating frame
structures and their dependence on ground motions Report no. 144. Stanford (CA):
John A. Blume Earthquake Engineering Center. Department of Civil,Engineering,
Stanford University; 2003.
Mehanny, S.S.F., and Deierlein, G.G. (2001), Seismic damage and collapse
assessment of composite moment frames, Journal of structural engineering, ASCE,
Vol.127, No.9, pp. 1045 -1153.
Mehanny, S.S.F., and Howary, H.A.E. (2010), Assessment of RC moment frame
buildings in moderate seismic zones: Evaluation of Egyptian seismic code implications
and system configuration effects, Engineering structures, Vol.32, No.8, pp.2394 2406.
Menegotto, M., and Pinto, P.E. (1973), Method of analysis for cyclically loaded
reinforced concrete plate frames including changes in geometry and non-elastic behavior
of elements under combined normal force and bending, International Association of
Bridge and Structural Engineering (IABSE), Lisbon, Portugal, Vol.13, pp. 15-22.

330

Messele, H., and Tadese, K. (2002), The study of seismic behaviour buildings located
on different site in Addis Ababa (Ethiopia) by using microtremors and analytical
procedure, Joint Study on micro tremors and seismic microzonation in earthquake
countries, Workshop to Exchange Research Information, Hakone-Gora, Kanagawa,
Japan.
Meziane, Y.A., Djakab, E., Benouar, D., Esat, I.I. (2012) Vulnerability of existing
buildings: empirical evaluation and experimental measurements, Natural hazards,
Vol.62, pp.189206
Midorikawa, S. (1990). Ambient vibration tests of buildings in Santiago and Via del
Mar, A Report on the Chile-Japan Joint Study Project on Seismic Design of Structures,
The Japan International Co-operation Agency
Midorikawa, S. (2004). Dense strong-motion array in Yokohama, Japan, and its use for
disaster management. International Workshop on Future Directions in Instrumentation
for Strong Motion and Engineering SeismologyNATO Meeting, Kusadasi, Turkey, 17
21 May 2004
Ministry

of

National

education.

Bingol

after

earthquake.

(http://

bingol.meb.gov.tr/index.htm), May 4, 2003 (in Turkey).


Miranda, E. (1993), Site-dependent strength reduction factors, Journal of Structural
Engineering, ASCE, Vol.119, No.12, pp. 3503 3519.
Miranda, E., and Akkar, D. (2003), Dynamic instability of simple structural systems,
Journal of Structural Engineering, ASCE, Vol.129, No.12, pp.1722-1726.
Miranda, E., and Akkar, D. (2003). Dynamic instability of simple structural systems,
Journal of structural engineering, ASCE, Vol. 129, No. 12, pp. 1722 1726.
Miranda, E., and Bertero, V. V. (1994), Evaluation of strength reduction factors for
earthquake resistant design, Earthquake Spectra, Vol.10, No.2, pp. 357 379.
Mitchell, D., Tinawi, R., and Redwood, R.G. (1990), Damage to buildings due to the
1989 Loma Prieta earthquake a Canadian code perspective, Canadian journal of civil
engineering, Vol.17, No.5, pp.813 834.
Miyakoshi, J., Hayashi, Y., Tamura, K. and Fukuwa, N. (1997). Damage ratio functions
for buildings using damage data of the Hygo ken Nanbu earthquake, In proceedings of

331

the seventh international conference on structural safety and reliability (ICOSSAR,),


Kyoto.
Moehle, J.P. (I992), "Displacement-based design of RC structures subjected to
earthquakes", Earthquake spectra, Vol. 8, No. 3, pp. 403 428.
Moehle, J.P., and Alarcon, L.F. (1986), Seismic Analysis Methods for Irregular
Buildings, Journal of Structural Engineering, ASCE, Vol. 112, No. 1, pp. 35-52.
Moelhe, J.P. (1984), Seismic Response of Vertically irregular structures, Journal of
Structural Engineering, ASCE, Vol.110, pp.2002-2014.
Moghadam, A.S. (1998), Seismic Torsional response of Asymmetrical

Multi-Storey

Frame buildings, A Doctoral Dissertation submitted to McMaster University, Canada.


Moghadam, A.S., and Aziminejad, A. (2005), Interaction of torsion and P-Delta effects in
tall buildings, In proceedings of Thirteenth World Conference on Earthquake
Engineering Vancouver, B.C., Canada August 1-6, 2004, Paper No. 799.
Moghadam, A.S., and Tso, W.K. (1996b), Seismic response of regular asymmetrical
RC ductile frame buildings, In proceedings of European Workshop on the seismic
behavior of asymmetric and setback structures, Capri naples italy, pp. 37 57.
Mohammed, H.H., Ghobarah, A., and Aziz, T.S. (2008), Seismic Response of
Secondary Systems Supported by Torsionally Yielding Structures, Journal of
Earthquake Engineering, Vol. 12, pp. 932 952.
Mohhamadi, R. K., and Naggar, M. H. (2004), Modifications on equivalent lateral force
method, In proceedings of 13th World Conference on Earthquake Engineering, Paper
No. 928, CD-ROM Vancouver, B.C., Canada.
Mller, M., Johansson, B., and Collin, P. (1997), A new analytical model of inelastic
local flange buckling, Journal of Constructional Steel Research, Vol. 23 (1-3), pp. 43-63.
Mork, K.J. (1992), Stochastic analysis of reinforced concrete frames under seismic
excitation, Soil dynamics and earthquake engineering, Vol.11, No.3, pp.145 161.
Mostaghel, N. (1999), Analytical description of pinching, degrading hysteretic systems.
Journal of Engineering Mechanics, ASCE, Vol.125, No.2, pp. 216 224.

332

Mucciarelli, M., Masi, A., Gallipoli, M.R., Harabaglia, P., Vona, M., Ponzo, F., and Dolce,
M (2004), Analysis of RC building dynamic response and soil-building resonance based
on data recorded during a damaging earthquake (Molise, Italy, 2002), Bullitten of
Seismological Society of America, Vol. 94, No.5, pp. 1943 1953.
Musson, R. M. W. (2000) "Intensity-based seismic risk assessment", Soil Dynamics and
Earthquake engineering, Vol. 20, pp. 353 360.
Mwafy, A. M., and Elnashai, A. S. (2002) Calibration of force reduction factors of rc
buildings, Journal of Earthquake Engineering, Vol.6, No.2, pp.239273.
Mwafy, A.M., and Elnashai, A.S. (2002), Calibration of force reduction factors of RC
buildings,Journal of Earthquake Engineering,Vol.6, No.2, pp. 239273.
Nakashima, M., Inoue, K., and Tada, M. (1998), Classification of damage to steel
buildings observed in the 1995 Hyogoken - Nanbu earthquake, Engineering structures,
Vol. 20, pp. 271 281.
Nakashima, M., Matsumiya, T., Suita, K., and Liu, D. (2006), Test on full-scale threestory steel moment frame and assessment of ability of numerical simulation to trace
cyclic inelastic behavior, Earthquake Engineering and Structural dynamics, Vol.35,No.1,
pp. 3 19.
Nassar, A.A., and Krawinkler, H. (1991), Seismic Demands for SDOF and MDOF
Systems, Report No. 95, The John A. Blume Earthquake Engineering Center,
Department of Civil and Environmental Engineering, Stanford University, Stanford,
U.S.A.
National information centre for earthquake engineering, UCSD Laboratory, Tohoku
university.
Navarro, M., and Oliveira, C.S. (2004), Evaluation of dynamic characteristics of
reinforced concrete buildings in the City of Lisbon, In proceedings of 4th Assembly of
the Portuguese-Spanish of Geodesy and Geophysics, Figueira da Foz, Portugal.
Navarro, M., and Oliveira, C.S. (2006). Experimental techniques for assessment of
dynamic behaviour of buildings, In proceedings of Oliveira CS et al. (eds) Assessing
and managing earthquake risk, Chapter 8. GGEE 2, Springer, pp 159182.

333

NBCC (1990), National Building Code of Canada 1990, National Research Council of
Canada, Ottawa, Ontario, 1990.
NBCC (1995), National Building Code of Canada 1995, National Research Council of
Canada, Ottawa, Ontario 1995.
NBCC (2005), National Building Code of Canada 1995, National Research Council of
Canada, Ottawa, Ontario 1995.
Negro,, P., Mola, F., Molina, J., Magonette, G.E. (2004), Full-scale psd testing of a
torsionally unbalanced three-storey non-seismic rc frame, In proceedings of 13th World
Conference on Earthquake Engineering, Ppaer No. 928, Vancouver, Canada.
Nielsen, N.N., and Imbeault, F.A., (1971), Validity of various hysteretic systems , In
proceedings of third Japan National Conference on Earthquake Engineering, pp. 707714.
NISEE, National information center for earthquake engineering, UCSD Structural
Engineering and Earthquake Disaster-Research Laboratory.
Niu, Di-tao. and Ren, Li-jie. (1996), A modified seismic damage model with double
variables for reinforcedconcrete structures, Journal of Earthquake Engineering and
Engineering Vibration, Vol.16, pp.44 55.
Oliveira, C.S (2004), Actualizao das bases-de-dados sobre frequncias prprias de
estruturas de edifcios,pontes, viadutos e passagens de pees a partir de medies
expeditas in-situ, In proceedings of 5th Portuguese Conference on Earthquake
Engineering, University of Minho, Guimares (in Portuguese)
Oliveira, C.S., and Navarro, M. (2010) Fundamental periods of vibration of RC
buildings in Portugal from in-situ experimental and numerical techniques Bullitten of
Earthquake Engineering, Vol. 8, pp.609642.
OPCM 3274 - OPCM 3431 (2003), Preliminary Elements of Technical Code for
Construction in Seismic Zones (in Italian), italy.
Open SEES. (2005), Open system for earthquake engineering simulation, Pacific
Earthquake Engineering Research Center, http://opensees. berkeley.edu.

334

Oropeza, M., Favez, P., Lestuzzi, P. (2010), Seismic response of nonstructural


components in case of nonlinear structures based on floor response spectra method, Bulletin of Earthquake Engineering,Vol. 8, pp.387 400.
Otani, S., and M.A. Sozen, (1972), Behavior of multistory reinforced concrete frames
during earthquakes, Structural Research Series No. 392, Civil Engineering Studies,
University of Illinois, Urbana.
Otani. S. (2002), Non-Linear earthquake response analysis

of reinforced concrete

buildings, Lecture Notes presented at Pavia , Italy.


Paramasivam, P., Tan, K. H., & Murugappan, K. (1995). Finite element analysis of
partially prestressed steel fiber concrete beams in shear. Advanced Cement Based
Materials, Vol.2, No.6, pp. 231-239.
Panagiotakos, T.B., and Fardis, M.N. (2001b), Deformations of reinforced concrete
members at yielding and ultimate, ACI Structural Journal, Vol.98 (S-13), pp. 135-148.
Panagiotakos, T.B., and Fardis, M.N., (2001a). A displacement based design procedure
for R/C Buildings and comparison with EC 8. Earthquake Engineering and Structural
Dynamics, Vol.30, pp.1439-1462.
Panda, S.K. and Ramachandra, L.S. (2010), Buckling of rectangular plates with various
boundary conditions loaded by non-uniform inplane loads, International Journal of
Mechanical Sciences, Vol. 52 , pp. 819828.
Park, Y.J., and Ang, A.H.S. (1985 a), Mechanistic seismic damage model for reinforced
concrete, Journal of Structural Engineering, ASCE,Vol.III, PP. 740-757.
Park, Y.J., Ang, A.H.S., Wen, Y.H. (1985b), Seismic damage analysis of reinforced
concrete buildings. Journal of Structural Engineering ASCE, vol.III (1985 b), pp. 722739.
Park, Y.J., Ang, A.H.S., Wen, Y.H. (1987), Damage limiting aseismic design of
buildings. Earthquake Spectra, Vol.3, No.1, pp.1-26.
Patil, V.B., and Jangid, R.S. (2011a), Optimum Multiple Tuned Mass Dampers for the
Wind Excited Benchmark Building, Journal of Civil Engineering and Management,
Vol17, No.4, pp.: 540557.

335

Patil, V.B., and Jangid, R.S. (2011b), Response of wind-excited benchmark building
installed with dampers, The Structural Design of Tall and Special buildings, Vol.20,
No.4, pp.497-514.
Paul, D.K. PhD Professor & Head Buildings Vulnerability, Building Types and Common
Problems, Typical Earthquake Damage Pattern Department of Earthquake Engineering,
IIT Roorkee, Roorkee.
Paulay, T., and Priestley, M.J.N. (1992), Seismic design of reinforced concrete and
masonry buildings, John Wiley and Sons , Newyork.
Pauw, A. (1960), Static modulus of elasticity of concrete as affected by density, Journal
American Concrete Institute, Vol. 57, No.6, pp.679-687.
PEER (Pacific Earthquake engineering Research centre). 2006. Stroang ground motion
database 2006:http://peer.berkley.edu/.
Pekau, O.A., and Green, R. (1974), Inelastic structures with setbacks, In proceedings
of fifth world conference on earthquake engineering, Rome, Italy (1974).
Pekau, O.A., and Guimond, R. (1990), Accidental torsion in yielding symmetric
structures, Engineering structures, Vol.12, 1990, pp 98106.
Penelis, G.G., and Kappos A.J. (2005), Inelastic torsion effects in 3D pushover analysis
of buildings, In Proceedings of the fourth European workshop on the seismic behaviour
of irregular and complex structures, CD ROM, Thessaloniki, August 2005.
Penelis, G.G., and Kappos, A.J. (1997). Earthquake-resistant Concrete Structures,
London: E & FN SPON (Chapman &

Hall) 1997.

Penelis, G.G., Sarigiannis, D., Stavrakakis, E., and Stylianidis,K.C. (1978), A statistical
evaluation of damage to buildings in the Thessaloniki, Greece earthquake of June 20,
1978, In Proceedings of the 9th world conference on earthquake engineering,
Thessolanki Greece VII (1989), pp. 18792.
Peru, I., and Fajfar, P. (2005). On the inelastic torsional response of single-storey
structures under bi-axial excitation, Earthquake Engineering and Structural Dynamics,
Vol. 34, pp.931941.

336

Peru1, P. Fajfar, P., and Dolek, M. (2008), Simplified nonlinear seismic assessment of
structures using approximate SDOF-IDA curves, In proceedings of 14th world
conference on earthquake engineering, October 12-17, 2008, Beijing, China
Pincheira, J.A., Dotiwala, F.S. and DSouza, J.T. (1999), Seismic analysis of older
reinforced concrete columns, Earthquake Spectra, Vol.15, No.2, pp. 245-272.
Pinto, P.E. (2005), The eurocode 8part 3: the new european code for the seismic
assessment of existing structures, Asian journal of civil engineering, Vol.6, No.5, pp.
447 456.
Polese, M., Verdane, G.M., Marinello, C., Iervolino, I., and Manfredi, G. (2008)
Vulnerability Analysis for Gravity Load Designed RC Buildings in Naples Italy, Journal
of earthquake engineering, Vol. 12(S2), pp.234245.
Powell, G. H., and Allahabadi, R. (1988) Seismic damage prediction by deterministic
methods: concepts and procedures, Earthquake Engineering and Structural Dynamics,
Vol. 16, pp.719734.
Prakash, V., Powell, G. H., and Campbell, S. (1993), DRAIN-2DX: Basic program
description and user guide, Report No. UCB/SEMM- 1993/17, Univ. of California,
Berkeley, California , U.S.A..
Priestley, M.J.N. (2003) Myths and Fallacies in Earthquake engineering- Revisted
,Mallet miline Lecture series, IUSS Press, Pavia , Italy.
Priestley, M.J.N., Calvi, G.M., and Kowalsky, M.J. (2007), Displacement based seismic
design of structures, IUSS Press, Pavia , Italy.
Proceedings of the Eleventh European Conference on Earthquake Engineering, A. A.
Balkema, Rotterdam, 1998.
Rama raju, R., Meher Prasad, A., Muthumani, K., Gopalakrishnan, N., Iyer, N.R., and
Lakhmanan, N.(2013) Experimental studies on use of toggle brace mechanism fitted
with magnetorheological dampers for seismic performance enhancement of three-storey
steel moment-resisting frame model, Structural Control and health monitoring, Vol.20,
No.3, pp.373-386.
Ramberg, W., and Osgood, W.R. (1943), Description of stress-strain curves by three
parameters, Monograph No. 4, Publicazione Italsider, Nuova Italsider, Genova.

337

Reinhorn, A.M., Madan, A., Valles, R.E., Reichmann, Y., and Mander, J.B. (1995),
Modeling of masonry infill panels for structural analysis. Report No. NCEER-95-0018,
State University of New York at Buffalo, Buffalo, N.Y.
Ricci, P., Luca, F.D.., and Verdane, G.M. (2011 b). 6th April 2009 LAquila earthquake,
Italy: reinforced concrete building performance, Bullitten of Earthquake Engineering,
Vol.9, pp.285 - 305.
Ricci, P., Verderame, G.M., and Manfredi, G. (2011 a). Analytical investigation of elastic
period of infilled RC MRF buildings, Engineering structures, Vol.33, pp.308-319.
Riddell, R., and Newmark, N.M. (1979), Statistical analysis of the response of nonlinear
systems subjected to earthquakes, Structural research series no. 468. Urbana:
Department of Civil Engineering: University of Illinois; 1979.
Roeder, C.W., Schneider, S.P., and Carpenter, J.E. (1993), Seismic behavior of
moment-resisting steel frames: Analytical study. Journal of structural engineering,
ASCE, Vol.119, No.6, pp. 1866 1884.
Romo A, A.,Costa, A.A., Pauprio, E., Rodrigues, H., Vicente, R., Varum H., and Costa,
A. (2013), Field observations and interpretation of the structural performance of
constructions after the 11 May 2011 Lorca earthquake, Engineering Failure Analysis, All
rights reserved. http://dx.doi.org/10.1016/j.engfailanal.2013.01.040.
Romeo, X.,Costa, A., and Delgado,R. (2004). Seismic behavior of reinforced concrete
frames with setbacks, In proceedings of 13th world conference on earthquake
engineering, (2004), paper no. 2027.
Roufaiel, M.S.L., and Meyer, C. (1987a), Analytical modeling of hysteric behavior of
R/C frames , Journal of structural engineering, ASCE, Vol.113, No.3, pp.429 -444.
Roufaiel, M.S.L., and Meyer, C. (1987b), Reliability of concrete frames damaged by
earthquakes, Journal of structural engineering, ASCE, Vol.113, No.3, pp.445 -457.
Ruiz, S.E., and Diederich, R. (1989). The Mexico Earthquake of September 19, 1985
The Seismic Performance of Buildings with Weak First Storey, Earthquake
Spectra.Vol.5, No. 1, pp.:89-102.

338

Sabelli, R., Mahin, S., and Chang, C. (2003), Seismic demands on steel braced frame
buildings with buckling-restrained braces, Engineering structures, Vol.25, No.5, pp.655666
Sadasiva, V.K., Deam, B.L., and Fenwick, R. (2008), Determination of Acceptable
Structural Irregularity Limits for the Use of Simplified Seismic Design Methods, In
proceedings of

eighth pacific conference on earthquake engineering, Singapore,

December 2007.
Saffari, H., and Mansouri, L. (2012), An efficient nonlinear analysis of 2D frames using a
Newton-like technique Archives of

civil and mechanical engineering,

(2012)

http://dx.doi.org/10.1016/j.acme.2012.07.003.
Snchez, F.J., Navarro, M., Garca, J.M., Enomoto, T., and Vidal, F. (2002), Evaluation
of seismic effects on buildings structures using microtremor measurements and
simulation response, Structural Dynamics, Eurodyn 2002, Vol. 2,pp. 10031008,
Balkema
Santa-Ana, P.R.(2004), Estimation of strength reduction factors for elastoplastic
structures: Modification factors, Paper No. 126, CD-ROM Proceedings of 13th World
Conference on Earthquake Engineering, Vancouver, B.C., Canada.
Sasani, K. K., Freeman, S. A., and Paret, T.F. (1998), Multimode pushover procedure
(MMP) - A method to identify the effects of higher modes in a pushover analysis , I
proceedings of 6th U.S. National conference on earthquake engineering,Seattle, Wash.
Sasani, M., and Kropelnicki, J. (2007), Progressive collapse analysis of an RC
structure, The Structural Design of Tall and Special Buildings, Vol. 17, No. 4, pp.757
771.Doi 10.1002/tal.375
Satake, N., Suda, K., Arakawa, T., Sasaki, A., and Tamura, Y. (2003), Damping
evaluation using full-scale data of buildings in Japan, Journal of Structural Engineering
ASCE, Vol. 129, No. 4, pp. 470 477.
Satish Kumar, K., Muthumani, K., Gopalakrishnan, N., Sivarama Sarma, B., Reddy,
G.R., and Parulekar, Y.R. (2002), Seismic Response Reduction of structures using
Elasto-Plastic passive energy dissipation, ISET Journal of earthquake technology,
Paper No.421, Vol.39, No.3,pp.121-138.

339

Satish kumar, K., Jeyasekhar, A., and Muthumani, K. (2012), A Design methodology for
supplemental damping for seismic performance enhancement of frame structures, Asian
journal of civil engineering ,Vol. 13, No. 5, pp.659-678.
Scawthorn, C., and Johnson, G.S. (2000), Preliminary report on Kocaeli (Izmit)
earthquake of 17 August 1999, Engineering Structures, Vol. 22, pp.727745
Scawthorn, C., Iemura, H. and Yamada, Y; (1981), Seismic damage estimation for
lowand mid-rise buildings in Japan", Earthquake Engineering and Structural dynamics,
Vol. 9, pp. 93 115,
Schiff, A. J., ed. (1991), Philippines Earthquake Reconnaissance Report, Earthquake
spectra, Supplement A to Vol.7.
Schneider, S.P., Roeder, C.W., and Carpenter, J.E. (1993), Seismic behavior of
moment-resisting steel frames: Experimental study Journal of Structural Engineering,
ASCE, Vol. 119, No.6, pp.1885 1902.
SEAOC (1999), Recommended Lateral Force Requirements and Commentary,
Sacramento (CA); 1999.
SEAOC Vision

Committee (2000), Performance-based seismic engineering, Report

prepared by Structural Engineers Association of California, Sacramento, (CA) 1995.


Seneviratna, G. D. P. K., and Krawinkler, H. (1997), Evaluation of seismic MDOF effects
for seismic design, John A. Blume Earthquake Engineering Centre Report No. 120,
Department of Civil and Environmental Engineering, Stanford University.
Sextos, A.G., Katsanos, N.E., and Manolis, G.D. (2011), EC8-based earthquake record
selection procedureevaluation:Validation study based on observed damage of an
irregular R/c Building Anastasios, Soil Dynamics and Earthquake Engineering, Vol.31,
pp.583 597.
Sezen, H., Whittaker, A. S., Elwood, K. J., and Mosalam, K. M. (2003). Performance of
reinforced concrete buildings during the August 17, 1999 Kocaeli, Turkey earthquake,
and seismic design and construction practise in Turkey, Engineering Structures, Vol.25,
No.1, pp. 103 114.
Sezen, H., Whittaker, A.S., Elwood, K.J., and Mosalam, K.M. (2003), Performance of
reinforced concrete buildings during the August 17 1999 Kocaeli, Turkey earthquake,

340

and seismic design and construction practice in Turkey, Engineering Structures 2,


Vol.25, pp.10314.
Shafei, B,, Zarein, F., and Lignos, D.(2011). A simplified method for collapse capacity
assessment of moment resisting frame and shear wall structural systems, Engineering
Structures, Vol. 33, pp.11071116.
Shahrooz, B.M., and Moehle, J.P. (1990), Seismic Response and Design of Setback
Buildings, Journal of Structural Engineering, ASCE, Vol. 116, No. 5, pp. 1423-1439.
Shakib, H., and Ghasemi, A. (2007), Considering different criteria for minimizing
torsional response of asymmetric structures under near-fault and far-fault excitations,
International journal of Civil Engineering, Vol.5, No.4, pp.247-265.
Shinozuka, M., Chang, S. E., Eguchi, R. T., Abrams, D. P., Hwang, H. H. M. and Rose,
A. (I997), Advances in earthquake loss estimation and application to Memphis,
Tennessee", Earthquake spectra, Vol.13, No.4, pp. 739 758.
SNZ (1992), Code of practice for general structural design and design loadings for
buildings (NZS 4203), Standards association of New Zealand, Wellington, New Zealand,
1992.
Sobai, M., Abd El-Rahman, K., and Mady, M. (2008), Estimation of dynamic
characteristics of existing common reinforced concrete buildings in Egypt using ambient
vibration tests., In proceedings of 14th World Conference on Earthquake Engineering,
Bejing, pp 1116
Song, J., and Pincheira, J. (2000), Spectral displacement demands of stiffness and
strength degrading systems, Earthquake Spectra, Vol.16, No.4, pp. 817 851.
Sorace, S. (1998), Seismic damage assessment of steel frames, Journal of Structural
Engineering, ASCE, Vol.124, No.5, pp. 531 540.
SP-127, Earthquake-Resistant Concrete Structures - Inelastic Response and Design,
(S.K.Ghosh, editor)in 1991 Chapter 1
Spence, R., Peterken, O., Booth, E., Aydinoglu, N. del Re, D., Bommer, J. and Tabuchi,
S. (2002), Seismic loss estimation for Turkish catastrophe insurance", In Proceedings
Seventh US National Conference on Earthquake Engineering, CD-ROM, Paper No. 341.

341

Spyropoulos, P.J. (1982), Report on the Greek earthquakes of February 2425, 1981,
American Concrete International ACI, Vol. 2, pp.1115.
Sreekala, R., Lakshmanan, L., Muthumani, K.M., Gopalakrishnan, N., and Satish Kumar,
K. (2007), Modeling post peak behavior of concrete for seismic applications,
International workshop on earthquake hazards and mitigation, Guwahati, India.
Sreekala, R., Lakshmanan, L., Muthumani, K.M., Gopalakrishnan, N., and Mehar
Prasad, A. (2011), Development of a Simplified Damage Model for Beams Aiding
Performance Based Seismic Design, Advances in Structural Engineering, Vol. 7, No.4,
pp.307-318.
Stathopoulos, K.G. and Anagnostopoulos, S.A. (2005), Inelastic torsion of multi-storey
buildings under earthquake excitations, Earthquake Engineering and Structural
Dynamics, Vol. 34, pp. 14491465.
Stathopoulos, K.G., and Anagnostopoulos, S.A. (2003), Inelastic earthquake response
of single-story asymmetric buildings: an assessment of simplified shear-beam models,
Earthquake Engineering and Structural Dynamics, Vol.32, pp.18131831.
Stathopoulos, K.G., and Anagnostopoulos, S.A. (2010). Accidental design eccentricity:
Is it really important for inelastic response of

buildings to stroang earthquakes? ,

Engineering Structures, Vol.30, pp.782-797.


Stefano, M.D., Marini, E.M., and Rossi, P.P. (2006) Effect of Overstrength on the
Seismic Behaviour of Multi-Storey Regularly Asymmetric Buildings, Bulletin of
Earthquake Engineering, Vol. 4, pp.2342
Stephens, J.E., and Yao, J.T.P. (1987), Damage assessment using response
measurement, Journal of Structural Engineering, ASCE, Vol.113, No.4, pp. 787 801.
Stone, W.C., and Taylor, A.W. (1993), Seismic performance of circular bridge columns
designed in accordance with AASHTO/CALTRANS standards NIST Building science
series, National institute of standards and Technology, Gaithersburg, M.D.
Sivaselvan, M.A., and Reinhorn, A.(2000) Hysteretic models for Deteriorating inelastic
systems, Journal of Engineering Mechanics, ASCE, Vol.126, No. 6, pp. 633 640.

342

Su, N. F., Lu X. L., and Zhou Y. (2009), Shaking table model test of a super high-rise
building with setbacks in elevation, In Proceedings of the 3rd International Conference
on Advances in Experimental Structural Engineering. San Francisco, USA.
Sucuoglu, H., and Erberik, A., (2004), Energy-based hysteresis and damage models for
deteriorating systems, Earthquake engineering and structural dynamics, Vol.33, No.1,
pp. 69 88.
Sullivan, T.J., and Lago, A. (2012), Towards a simplified Direct DBD procedure for the
seismic design of moment resisting frames with viscous dampers, Engineering
Structures, Vol. 35, pp.140148.
Sullivan, T.J. (2011a), An Energy-Factor Method for the Displacement-Based Seismic
Design of RC Wall Structures, Journal of earthquake engineering, Vol.15, No.7, pp.1083
-1116
Sullivan, T.J. (2011b), An Energy-Factor Method for the Displacement-Based Seismic
Design of RC Wall Structures, Journal of Earthquake Engineering, Vol.15, No.7, pp.
Takeda, T., Sozen, M., and Nielsen, N. (1970), Reinforced concrete response to
simulated earthquakes, Journal of the Structural Division, Vol. 96 (ST12), pp. 25572573.
Takizawa, H. (1975), Non-linear models for simulating the dynamic damaging process
of low-rise reinforced concrete buildings during severe earthquakes, Earthquake
Engineering and Structural dynamics, Vol.4, No.1, pp. 73 -84.
Takizawa, H., and Jennings, P.C. (1980), Collapse of a model for ductile reinforced
concrete frames under extreme earthquake motions, Earthquake Engineering and
Structural Dynamics, Vol.8, pp.117-144.
Takizawa, H., and Jennings, P.C. (1980), Collapse of a model for ductile reinforced
concrete frames under extreme earthquake motions Earthquake engineering and
structural dynamics, Vol.8, pp. 117 144.
Tan, K.H., and Balendra, T. (2007), Retrofit of existing buildings for earthquake
resistance,

Journal

of

Earthquake

10.1142/S1793431107000110.

343

and

Tsunami,

Vol.

01,

DOI:

Tan, K.H., and Saha , M. (2008), Cracking Characteristics of RC Beams Strengthened


with FRP System, Journal of Composites for Construction, ASCE, Vol. 12, No. 5,
pp.513-521.
Tapan, M., Comert, M., Demir, C., Sayan, Y., Orakcal, K., and Ilki, A. (2013), Failures of
structures during the October 23, 2011 Tabanl (Van) and November 9, 2011 Edremit
(Van)

earthquakes

in

Turkey,

Engineering

Failure

Analysis,

http://dx.doi.org/10.1016/j.engfailanal.2013.02.013
Taskin, B., Sezen, A., Tugsal, U.L., and Erken, A. (2013), The aftermath of 2011 Van
earthquakes: evaluation of strong motion, geotechnical and structural issues, Bulliten of
earthquake engineering, Vol.11, pp.285 312.
TEC (2007), Ministry of Public Works and Settlement, Specification Structures To Be
Built In Disaster Areas, Part III Earthquake Disaster Prevention, Government of
Republic of Turkey, Turkey
Tena-Colunga, A. (2004), Evaluation of the seismic response of slender, setback RC
moment-resisting frame buildings designed according to the seismic guidelines of a
modern seismic code, In 13th world conference on earthquake engineering, Paper no.
2027, Canada (2004).
Thyagarajan, R.S., and Iwan, W.D. (1990), Performance characteristics of a widely used
hysteretic model in structural dynamics, In proceedings of fourth U.S. National
conference on Earthquake engineering, Earthquake engineering research institute (
EERI), Vol.2, pp. 177 186.
Tomaevic,M., and PolonaWeiss, W. (2010), Displacement capacity of masonry
buildings as a basis for the assessment of behavior factor: an experimental study,
Bullitten of earthquake engineering, Vol.8, pp.1267 1294.
Tremblay, R., and Poncet, L. (2005), Seismic performance of concentrically braced steel
frames in multistory buildings with mass irregularity, Journal of Structural Engineering,
Vol.131, pp.13631375.
Tsai, K.C., Hsiao, C.P., and Brunaeu, M. (2000), Overview of building damages in Chi
Chi Earthquake, Earthqauke engineering and Enginering Seismiology, Vol.2, No.1, pp.
93 108.

344

Tsai, M.H. (2012), A performance-based design approach for retrofitting regular building
frames with steel braces against sudden column loss, Journal of construction and steel
research, Vol. 77, pp.111.
Tsai, M.H., and Huang, T.C. (2012), Progressive Collapse Analysis of an RC Building
with Exterior Non-Structural Walls, In proceedings of twelfth East Asia-Pacific
Conference on Structural Engineering and Construction, Procedia Engineering, Vol. 14,
pp.377384.
Tso, W.K. (1994), Static eccentricity concept for torsiona moment estimation, Journal
of Structural Engineering, ASCE, Vol.116, No.5, pp. 1199 1212.
Tso, W. K., & Dempsey, K. M. (1980), Seismic torsional provisions for dynamic
eccentricity, Earthquake Engineering & Structural Dynamics, Vol.8, No.3, pp.275 289.
Tso, W.K. and Myslimaj, B. (2002). Effect of strength distribution on the inelastic
torsional response of asymmetric structural systems In: Proceedings of the 12th
European conference on earthquake engineering, CD ROM, London, September 2002.
Tso, W.K. and Sadek, A.W.(1985). Inelastic seismic response of simple eccentric
structures, Earthquake Engineering and Structural Dynamics, Vol. 13, No.2, pp.255269.
Tso, W.K., and Bozorgnia,Y. (1986),Effective eccentricity for inelastic seismic response
of buildings, Earthquake engineering and Structural Dynamics, Vol.14, No.3,pp.413427.
Tso, W.K., and Myslimaj, B. (2003). A yield displacement distribution-based approach
for strength assignment to lateral force-resisting elements having strength dependent
stiffness, Earthquake Engineering and Structural Dynamics, Vol. 32, pp.23192351.
Tso, W.K., and Sadek, A.W. (1989), Strength eccentricity concept for inelastic analysis
of asymmetrical structures, Engineering Structures, Vol.11, No.3, pp.189-194.
Uang, C.M. (1991), Establishing R (or Rw) and Cd factors for building seismic
provisions, Journal of Structural Engineering, ASCE, Vol. 117, No.1,pp. 1928.
Uang, C.M. (1993) An evaluation of two-level seismic design procedure, Earthquake
Spectra, Vol.9, No.1, pp. 121-135.

345

UBC (1988), International conference of building officials (ICBO), Uniform building


code, Whittier, California,1988.
UBC (1991), International conference of building officials (ICBO), Uniform building
code, Whittier, California, 1991.
UBC (1994), Uniform building code (UBC 94), International conference of building
officials (ICBO), Whittier, California, 1994.
UBC (1997), International conference of building officials (ICBO), Uniform Building
code, Whittier, California, 1997.
Valles, R.E., Reinhorn, A.M., Kunnath, S.K., Li, C., and Madan, A. (1996), IDARC2D: A
computer program for the inelastic analysis of buildings, Report, No. NCEER-96-0010,
State University of New York at Buffalo, Buffalo, NY.
Valmundsson, E.V. and Nau, J.M. (1997). Seismic Response of Building Frames with
Vertical Structural Irregularities, Journal of Structural Engineering, ASCE, Vol. 123, No.
1, pp. 30-41.
Vamvatsikos, D., and Cornell, C.A. (2002), Incremental dynamic analysis, Earthquake
Engineering and Structural Dynamics, Vol.31, No. 3, pp.491514.
Vamvatsikos, D., and Cornell, C.A. (2004), Applied incremental dynamic analysis
Earthquake Spectra, Vol.20, No. 2, pp.523553.
Vamvatsikos, D., and Cornell, C.A. (2005), Direct estimation of seismic demand and
capacity of multidegree-of-freedom systems through incremental dynamic analysis of
single degree of freedom approximation Journal of Structural Engineering, ASCE, Vol.
131, No.4, pp. 589 599.
Verdane, G.M., Luca, F.D., Ricci, P., and Manfredi, G. (2010), Preliminary analysis of a
soft-storey mechanism after the 2009 LAquila earthquake, Earthquake Engineering and
Structural dynamics, Vol.40, No.8, pp.925 944.
Vian, D., and Bruneau, M. (2003), Tests to structural collapse of single degree of
freedom frames subjected to earthquake excitations, Journal of structural engineering,
ASCE, Vol.129, No.12, pp. 1676 1685.

346

Vidic, T., Fajfar, P., and Fischinger, M. (1994), Consistent inelastic design spectra:
strength and displacement, Earthquake Engineering and Structural Dynamics, Vl. 23,
No.5, pp. 507 521.
Villaverde, R. (1991), Explanation for the numerous upper floor collapses during the
1985 Mexico City earthquake, Earthquake Engineering and Structural Dynamics,
Vol.20, No.3, pp. 223 241
Wallace, J.W., and Moehle, J. P. (1992), Ductility and detailing requirements of bearing
wall buildings, Journal of Structural engineering, ASCE, Vol. 118, No.6, pp.16251644.
Wang, M.L., and Shah, S.P. (1987), Reinforced concrete hysteresis based in the
damage concept, Earthquake Engineering and Structural Dynamics, Vol.15, No.8,
pp.993 1003.
Wang,

Q.,

and

Wang,

L.YY.

(2005)

Estimating periods of

vibration

of buildings with coupled shear walls, Journal of structural engineering ASCE, Vol.131,
No.12, pp.1931 1935.
Wen, Y.K. (1980), Equivalent linearization for hysteretic systems under random
excitation, Journal of Applied Mechanics, Vol. 47, pp. 150-154.
Westenenk, B., De la Llera, J.C., Jnemann, R., Hube, M.A., Besa, J.J., Carl Lders
Inaudi, J.A., Riddell, R., and Jordn, R. (2013) Analysis and interpretation of the seismic
response of RC buildings in Concepcin during the February 27, 2010, Chile
earthquake, Bulletin of

Earthquake Engineering, Vol.11, pp.69 91. DOI

10.1007/s10518-012-9404-5
Whitman, R. V., Anagnos, T., Kircher, C., Lagorio, H. C., Lawson, R. S. and Schneider,
P. (I997) "Development of a national earthquake loss methodology," Earthquake.
Spectra, Vol.13, No.4, pp.643 661.
Williamson, E.B. (2003), Evaluation of damage and P - Delta effects for systems under
earthquake excitation, Journal of Structural Engineering, ASCE, Vol.129, No.8, pp.1036
1046.
Wong, C.M., and Tso, W.K. (1994). Seismic Loading for Buildings with Setbacks,
Canadian Journal of Civil Engineering, Vol. 21, No. 5, pp. 863-871.

347

Wood, S.L. (1992), Seismic Response of R/C Frames with Irregular Profiles, Journal of
Structural Engineering, ASCE, Vol. 118, No. 2, pp. 545-566.
Wyllie, L.A., Abrahamson N., Bolt, B., Castro, G., Durkin, M.E., Escalante, L., Gates,
H.J., Luft, R., McCormick, D., Olson, R.S., Smith, P.D. and Vallenas, J. (1986). The
Chile Earthquake of March 3, 1985, Earthquake Spectra, Vol.2, No.2, pp. 293-371.
Yamazaki, F. and Murm, O. (2000) "Vulnerability functions for Japanese buildings based
on damage data from the 1995 Kobe earthquake", In Implications of Recent Earthquakes
on Seismic Risk (eds. A. S. Elnashai & S. Antoniou), Imperial College Press, pp. 91-102.
Yoshimura K, Kuroki M. Damage to Building Structures Causedby the 1999 Chi-chi
Earthquake in Taiwan, hhttp:// www.arch.oita-u.ac.jp/a-kou/taiwan1.pdfi, December
2003.
Youd T.L., Bardet J.P., and Bray J.D. (2000), Kocaeli, Turkey, earthquake of August 17,
1999 Reconnaissance Report, Earthquake spectra, Supplement A to Vol.16.
Zareian, F., and Krawinkler, H. (2009), Simplified performance-based earthquake
engineering, Report no. TB 169. Stanford (CA): John A. Blume Earthquake Engineering
Research Center. Department of Civil and Environmental Engineering, Stanford
University; 2009.
Zarein, F., Krawinkler, H., and Ibarra, L. (2009), Basic concepts and performance
measures in prediction of collapse of buildings under earthquake ground motions,The
Structural

Design

of

Tall

and

Special

Buildings,

Vol.19,

pp.167181.

DOI: 10.1002/tal.546.
Zhao, B., Taucer, F., & Rossetto, T. (2009), Field investigation on the performance of
building structures during the 12 May 2008 Wenchuan earthquake in China,
Engineering Structures, Vol.31, No.8,pp. 1707-1723.

348

APPENDIX A: DETAILS OF ADDITIONAL GROUND MOTIONS ADOPTED


In this section, an effort has been made to generalize the equations derived, so that they
have a wider applicability. To achieve this purpose, a set of ground motions have been
selected from ATC 3 - 06 (1978). On basis of these ground motions, NEHRP response
spectrum was derived and the NEHRP spectrum is being widely used in seismic design
codes like IBC 2003, UBC 97, IS 1893 2002, EC 8:2004 (Nassar and Krawinkler 1991;
Krawinkler and Seneviratana 1997; Al- Ali and Krawinkler 1998). This is evident from Figure
A.1 which shows similarity in the response spectrums adopted by different codes of practice.
Moreover, the code spectrums are used widely in large number of reported research works
[Kircher et al. 1997; Aschheim and Black 2000; Elnashai and Mwafy 2002; Sezan et al.
2003; Stathopoulos and Anagnostopoulos 2003; Stathopoulos and Anagnostopoulos 2005;
Dogangun and Livaoglu 2006; Karavasilis and Seo 2011]. Therefore, using these ground
motions (adopted by most of the seismic design codes to derive response spectrum) will
increase the applicability of the proposed relations. The system of ground motions adopted
have been presented in Table A.1.

Table A.1 Ground motion data pertaining to ATC 3-06 (1978) [NEHRP 1994, UBC 97,
IS1893:2002, EC8:2004; ,IBC 2003] for long period structures (PEER)
Name of Earthquake and Station

Date

Mu

1989

De
(Km)
7.1 09

PGA
Tc
SF
(m/s2)
4.94
0.610 3.38

Loma Prieta (Saratoga Aloha)


Loma Prieta Gilroy Array 6

1989

7.1 25

1.669

0.933 8.03

Loma Prieta (Stanford Linear)

1989

7.1 34

2.823

0.725 10.4

Loma Prieta (San Francisco, Diamond hill)

1989

7.1 75

1.10

0.550 13.2

Loma Prieta (San Francisco, Presidlo)

1989

7.1 79

1.941

0.655 11.9

Landers (Desert Hot springs)

1992

7.5 18

1.675

0.904 7.36

Landers (AMBOY 1992)

1992

7.5 44

1.432

0.252 7.49

Landers (Fort Irwin)

1992

7.5 62

1.190

0.305 11.2

Northridge (Arieta)

1994

6.7 9

3.02

0.700 4.17

Northridge (New Hall Fire station)

1994

6.7 19

5.71

0.612 39.2

Northridge (LA-Century City CC North)

1994

6.7 37

3.10

0.451 20.54

Northridge (LA-Baldwain Hills)

1994

6.7 31

1.65

0.401 11.9

Northridge (Castaic Old Ridge Route)

1994

6.7 23

5.58

0.410 2.46

Northridge (Sylmar country hospital)

1994

6.7 15

3.02

0.450 33.7

349

Sa/g

0.75
EC 8
NEHRP
UBC
IS 1893

0.5

0.25

0
0

2.5

7.5

10

T (Sec)
Figure A.1 Response spectrums adopted by different codes of practice

Figure A.2 Scaling of ground motions in Table A.1 for long period structures (PEER)

350

APPENDIX B :ADDITIONAL SEISMIC RESPONSE PARAMETERS

B.1 Equations to compute strength demands in members


The traditional methods of design are force based in which the design strength was obtained
by multiplying the forces with a predefined force reduction factor. The recent methods of
seismic design mainly focused on reducing the displacement or drift to a prescribed limit.
However, the force based methods are still preferred, as structures designed by this method
exhibit adequate seismic performance. This is mainly due to the fact that in this method the
factors like reserve strength, ductility and seismic energy dissipation capacity are accounted
for, which increases conservativeness of this method. Moreover, the seismic design codes
show differences regarding the aspect of force reduction factor; as some of the codes (SNZ
1992; NBCC 1995; UBC 97; NBCC 2005) consider the force reduction factor to account for
both reserve strength and ductility which results in higher value of forcer reduction factor.
But, some codes like EC 8 2004 consider force reduction factor to account for reserve
strength only. Furthermore, accurate estimation of over-strength factor is difficult due to its
dependence on large number of factors like actual material strength, reinforcement
confinement, contribution of non-structural components, fundamental time period, design
intensity level and ductility etc. The strength of the structural members is one of the most
important criteria which is considered in seismic design of buildings, and previous research
works (Haselton et al. 2011a; Elnashai and Mwafy 2002) suggest that the strength of the
structure can be effectively represented by the over-strength ratio (ratio between ultimate
strength and design base shear). In accordance with this definition, the values of over
strength factor for all the building models were determined. Based on the regression
analysis conducted on the seismic response databank the equation to compute overstrength factor has been proposed by the author as
4.13 0.31c 0.36q 0.0587

Vb
W

(B.i)

The detailed range of variation of over strength factor for the building models considered in
analytical study has been presented in the Appendix section. The proposed equation yielded
comparable results with dynamic analysis as evident from Figure B.1.

B.2 Determination of stress demands


The excessive stress concentration in a building may result in its premature failure. Hence,
the primary aim of the designer should be to keep the stress within prescribed limits. The
presence of irregularity in a building changes the stress demands. However, this aspect has

351

been ignored by the seismic design codes. Furthermore, the stress demands exhibited a
strong correlation with parameters like base shear, irregularity index and strength reduction
factor as indicated by seismic response database. Based on regression analysis conducted
on the seismic response data bank the equation to compute the stress demand () for the
irregular buildings has been proposed by the author as

38712.32 1682.34c 1825.4q 63.65Vb


H2

(B.ii)

where stress is computed in kN/m2. The detailed range of variation of overstrength factor for
the building models considered in analytical study has been presented in Appendix E. As
evident from Figure 4.45, the proposed equation yields comparable results with the dynamic
analysis with a correlation coefficient of 0.9884.

(Proposed Method)

1
R2 = 0.9945

0.9

0.8
0.8

0.9

(Dynamic Analysis)

Figure B.1 Comparison of propose equation with the dynamic analysis for a) strength

max (Proposed Method)

demand in terms of over-strength factor ()


1

0.79

R2 = 0.9884

0.58

0.37

0.16
0.16

0.37

0.58

0.79

max (Dynamic Analysis)

Figure B.2 Stress demand for irregular building models considered in the analytical
study

352

ADDITIONAL RESULTS FOR LONG PERIOD STRUCTURES


EC 8 DESIGN BASED SECTIONS

(a) Rectangular section (G 2 ground motion)

rd 0.000751 c 0.0000626H 0.00919q

I r 0.0821 c 0.0163 rd 0.0073q 0.617


H
i 0.006123 c 0.004312q 0.01347

(B1.1)
(B1.2)
(B1.3)
(B1.4)

V
3.83 0.228 c 0.435q 0.0738 b
W
(41713.76 1497.23 s 1632.23q 57.45Vb )

H2

cp 1.518 0.238T 0.2849 C 0.7312 3.68

pc
0.9341m' 0.1613Cs 0.077Clvcc
pl

Pc 0.2286 0.1582T 0.0131 C 0.0289 0.376

pc
0.0599m' 0.1511C s 0.073Clvcc
pl

(B1.5)
(B1.6)

(B1.7)

DM 0.0289c 0.0498q

(B1.8)

11 2.213 105 c5 1.321104 c4 7.987 103 c3 0.876 102 c2 0.5324c 1.592

(B1.9)

(b) Square section (G1 set of ground motions)

rd 0.000598 c 0.00000832H 0.00687q

I r 0.061 c 0.01102 rd 0.00775q 0.531

(B1.10)
(B1.11))

i 0.005023 c 0.003523q 0.01214

(B1.12)

Vb
W
(36521 1782.23 s 1913.27q 72.32Vb )

H 1.976

(B1.13)

4.23 0.3213 c 0.3783q 0.0613

(B1.14)

pc
0.9012m' 0.1513C s 0.091Clvcc
pl
pc
Pc 0.1983 0.1278T 0.0131 C 0.0353 0.387
0.0668m' 0.1231C s 0.068Clvcc
pl

(B1.15)

DI 0.0913q 0.0307 C

(B1.17)

11 1.874 105 c5 1.412 104 c4 6.235 103 c3 0.872 102 c2 0.591c 1.392

(B1.18)

cp 1.712 0.303T 0.3113 C 0.595 4.323

353

(B1.16)

(c) Square section (G2 set of ground motions)

rd 0.000810 c 0.00001023H 0.00934q

I r 0.0876 c 0.01129 rd 0.00689q 0.6119

(B1.19)

i 0.006165 c 0.003836q 0.01408

(B1.21)

Vb
W
(37834.36 1452.46 s 1763.23q 75.36Vb )

H 1.842

(B1.22)

(B1.20)

3.69 0.3109 c 0.3543q 0.05676

(B1.23)

pc
0.8123m' 0.1123C s 0.069Clvcc
pl
pc
Pc 0.2267 0.1510T 0.0209 C 0.0298 0.436
0.0632m' 0.1213C s 0.067Clvcc
pl

(B1.24)

DI 0.0432q 0.0562 C

(B1.26)

11 2.113 105 c5 1.356 104 c4 6.672 103 c3 0.983 102 c2 0.574c 1.479

(B1.27)

cp 1.569 0.249T 0.3761 C 0.611 4.724

(B1.25)

IS DESIGN BASED SECTIONS

(a) Rectangular section ( G1 set of ground motions)

rd 0.000654 c 0.00000903H 0.00701q

I r 0.0732 c 0.01450 rd 0.00780q 0.589

(B1.28)

i 0.005325 c 0.004034q 0.01319

(B1.30)

Vb
W
(40239.45 1598.12 s 1712.34q 72.98Vb )

H 1.913

(B1.31)

(B1.29)

3.952 0.2874 c 0.452q 0.0752

cp 1.579 0.241T 0.2912 C 0.602 5.018

pc
0.8271m' 0.1231Cs 0.074Clvcc
pl

Pc 0.2496 0.1785T 0.0156 C 0.0254 0.431

pc
0.0605m' 0.1309C s 0.075Clvcc
pl

(B1.32)
(B1.33)

(B1.34)

DI 0.0265 c 0.0516q

(B1.35)

11 2.539 105 c5 1.487 104 c4 7..981103 c3 0.821102 c2 0.5329c 1.613

(B1.36)

354

(b) Rectangular section ( G2 set of ground motions)


rd 0.000774 c 0.0000659H 0.00937q
I r 0.08320 c 0.01750

rd

0.00820q 0.621

H
i 0.006451c 0.004710q 0.01461

(B1.38)
(B1.39)

Vb
W
(45612.98 1361.47 s 1582.14q 62.13Vb )

H 1.783
3.684 0.2012 c 0.4254q 0.0710

cp 1.489 0.213T 0.2729 C 0.739 53.75

(B1.37)

(B1.40)
(B1.41)

pc
0.932m' 0.1312Cs 0.071Clvcc
pl

(B1.42)

pc
0.0627m' 0.1375C s 0.084Clvcc
pl

(B1.43)

Pc 0.2342 0.1618T 0.0139 C 0.0278 0.411

DI 0.0276 c 0.0508q

(B1.44)

11 2.342 105 c5 1.231104 c4 78.132 103 c3 0.754 102 c2 0.5023c 1.624

(B1.45)

(b) Square section (G1 set of ground motion)

rd 0.000631 c 0.0000875H 0.00673q

I r 0.0702 c 0.01276 rd 0.00704q 0.542

(B1.46)

i 0.005021 c 0.003872q 0.01217

(B1.48)

Vb
W
(46287.39 1296.34 s 1618.23q 78.38Vb )

H 1.812

cp 1.397 0.304T 0.3159 C 0.618 4.351 pc 0.8021m' 0.1123Cs 0.062Clvcc


pl

(B1.49)

(B1.47)

3.812 0.2945 c 0.4612q 0.05239

Pc 0.2549 0.1387T 0.0195 C 0.0386 0.497

pc
0.0886m' 0.1236C s 0.092Clvcc
pl

(B1.50)
(B1.51)
(B1.52)

DI 0.0274 c 0.0569q

(B1.53)

11 1.762 105 c5 1.231104 c4 65.872 103 c3 0.856 102 c2 0.521c 1.293

(B1.54)

355

(c) Square Section (G2 set of ground motions)

rd 0.000848 c 0.000123H 0.000931q

I r 0.0893 c 0.01011 rd 0.00713q 0.6115

H
i 0.006213 c 0.003854q 0.01631

(B1.55)
(B1.56)
(B1.57)

Vb
W
(39876.24 1386.12 s 1753.23q 81.26Vb )

H 1.729

cp 1.386 0.265T 0.3093 C 0.610 4.381 pc 0.8230m' 0.1239Cs 0.073Clvcc


pl
pc
Pc 0.2349 0.1623T 0.0234 C 0.0285 0.559
0.0829m' 0.1341C s 0.085Clvcc
pl

(B1.58)

DI 0.0413 c 0.0641q

(B1.62)

3.57 0.2896 c 0.3385q 0.05823

356

(B1.59)
(B1.60)
(B1.61)

APPENDIX C: SHORT PERIOD STRUCTURES


C.1 Introduction
The research work in the Chapters 1- 6 deal with the aspect of long period structures. The
long period structures are classified as per relation
Tb Tc (Long period structures)

(C1.1)

where Tb = Time period of the building, Tc = Critical time period of the ground motion
considered
However, in reality the fundamental time period of the structures may fall below the critical
time period of the ground motion as shown in equation C.1.2.In such a case, the structures
are referred as Short period structures
Tb Tc (Short period structures)

(C1.2)

As per EC 8:2004, the displacements for short period structures (Tb < Tc) can be computed
as

T
1 ( q 1) c

Tb

(C1.3)

where, = displacement ductility and q = Behavior factor


Nevertheless, the rule stated in equation C1.3 is derived on basis of SDOF system and
elastic analysis which is unrealistic. Therefore, there is necessity of a new rule which
overcomes these limitations and predict realistic estimate of the seismic demand. To
generate a condition such that critical time period of ground motion falls below the building
time period, a separate set of ground motions as described in Tables C1 and C2 are used.
The ensemble of the ground motions is presented in Figures C2 and C3. To determine the
seismic behavior of short period structures, the building models with height ranging from 4
storey to 12 storey have been adopted with input data as described in Chapter 2 with
modeling scheme as described in Table C.3.The setback geometries for adopted to study
the seismic behavior has been described in Figure C1.

C.2 Irregularity index for short period structures


In the present approach, a parameter called as irregularity index is proposed based on
dynamic characteristics of the frame to quantify the setback irregularity. The natural fre-

357

quency of vibration and mass participation factor are the most important dynamic
characteristics of the building under seismic excitation. Therefore, irregularity index can be
proposed based on any of these parameters, but the sensitivity analysis results on the
irregular building frames as shown in Table C.4 show the dominance of natural frequency of
vibration over the mass participation factor. The sensitivity index for these parameters is
calculated by the method suggested by Kose (2008). Therefore, the irregularity index to
quantify the setback irregularity is proposed as
1

s
k

irr
r

(C1.4)

are the modal combinations of frequency of vibration of the irregular and regular building
frames from mode 1to mode k.

17

34

52

71

18

35

53

72

19

20

21

22

23

40

41

36

54

73

37

55

74

38

39

56

75

57

76

58

77

78

24

25

42

59

79

10

26

43

60

80

27

44

61

81

11

45

62

63

82

83

12

13

28

46

64

84

14

29

47

65

85

30

48

66

86

15

31

49

16

32

50

33

51

67

68

69

87

88

89

Figure C.1 Different setback geometries considered in the analytical study of short
period structures

358

Table C.1 System of ground motions used for short period structures
Name of Earthquake and Station

Date

Mu

De

PGA

Tc

Imperial valley 1979 (EI Centro Array # 4)


Imperial Valley(Delta)
San Fernado (Pacoima Dam)
San Fernado (Castaic)
Northridge, LA - 116th St School
Northridge, LA - Cypress Ave
Northridge, LA Fletcher dr
Northridge (Hollywood Wiloughby Ave)
Northridge (LA N Faring Rd)
Northridge (LA-S Grand Ave)
Northridge (LA-Obregon Park)
Northridge (LA-Wonderland Ave.)
Imperial Valley-06, EC Country center FF

1979/10/15
15/10/1979
1971/02/09
09/02/1971
17/01/1994
17/01/1994
17/01/1994
17/01/1994
17/01/1994
17/01/1994
17/01/1994
17/01/1994
1979/10/15

6.5
6.5
6.6
6.6
6.7
6.7
6.7
6.7
6.7
6.7
6.7
6.7
6.5

15.5
(Km)
44.0
2.80
29.0
41.9
32.8
29.5
26.0
24.0
37.0
37.0
31.0
7.60

1.31
2
(m/s
3.44 )
12.01
2.63
2.04
2.06
2.35
2.41
2.68
2.84
3.10
1.65
2.41

0.60
0.61
0.60
0.60
0.60
0.60
0.60
0.91
0.61
0.65
0.50
0.50
0.60

Table C.2 Ground motion data pertaining to ATC 3-06 (1978) [NEHRP 1994, UBC 97,
IS1893:2002, EC8:2004, IBC 2003] for short period structures (PEER)
Name of Earthquake and Station
Loma trieta (Saratoga Aloha)
Loma prieta Gilroy Array 6
Loma prieta (Stanford Linear)
Loma prieta (San Francisco, Diamond hill)
Loma prieta (San Francisco, Presidlo)
Landers (Desert Hot springs)
Northridge (Arieta)
Northridge (New Hall Fire station)

Date
1989
1989
1989
1989
1989
1992
1994
1994

Mu De

PGA

Tc

7.1
7.1
7.1
7.1
7.1
7.5
6.7
6.7

4.942
(m/s )
1.669
2.823
1.10
1.941
1.675
3.02
5.71

0.610
0.933
0.725
0.550
0.655
0.904
0.700
0.612

09
(Km)
25
34
75
79
18
9
19

Table C.3 Detailed locations of irregularities

No. of storeys in which irregularity is present

Total cases

2,1

2,1

2,1

12

10

2,1

15

12

2,1

18

Where N represents number of storeys in the building


359

SF
3.38
8.03
10.4
13.2
11.9
7.36
4.17
39.2

Table C.4 Sensitivity analysis results

S.No

Parameter

Mass
participation
factor of
Frequency
vibration

Sensitivity index with respect to the deformation demand


parameter
rd
IDR
i
0.365

0.413

0.354

0.413

0.468

0.421

Figure C.2 Mean scaled spectrum for ground motions in Table C.2 for short period
structures

Figure C.3 Mean scaled spectrum for ground motions in Table C.3 for short period
structures

360

C.3 Behavior factor for short period structures


This section mainly aims at proposing the behavior factor for short period structures
incorporating the aspects of structural irregularity and cracking. The details of different
aspects and methodology to evaluate the behavior factor has been discussed in Chapter
4.Based on inelastic dynamic analysis conducted on short period structures, the relation
between the irregularity index (s) and the behavior factor for gross stiffness and cracked
stiffness is shown in Figure C.4. and C.5.This relation can be empirically expressed as
q 0.543s2 7.832s 10.55

(C.1.5)

qc 0.489s2 7.372s 11.35

(C.1.6)
The proposed behavior factors are observed to yield comparable results with dynamic
analysis as expressed in Figures C.6 and C.7.

4.8

R2 = 0.9831

4.05

3.3

2.55

1.8
0.4

0.7

Figure C.4 Relation between proposed irregularity index and behavior factor for short
period structures considering gross stiffness

361

5.2
R2 = 0.9896

qc

4.35

3.5

2.65

1.8
0.4

0.7

s (Cracked)
Figure C.5 Relation between proposed irregularity index and behavior factor for short
period structures considering cracked stiffness

q (Proposed equation)

R2 = 0.9885

0.7

0.4
0.4

0.7

q (Dynamic analysis)

Figure C.6 Comparison of behavior factor for short period structures using proposed
equation and dynamic analysis for gross stiffness

362

qc (Proposed equation)

R2 = 0.9863

0.7

0.4
0.4

0.7

qc (Dynamic analysis)
Figure C.7 Comparison of behavior factor for short period structures using proposed
equation and dynamic analysis
It is worthwhile to note that equations to estimate seismic response parameters and mean
values of seismic response parameters for short period structures (regression analysis
carried out using MATLAB software) have been presented in Appendix D and F. In case of
short period structures, the modified fundamental time period (incorporating irregularity and
cracking) can be directly obtained from the irregularity index.

363

This page is intentionally left blank

364

APPENDIX D: RESULTS FOR SHORT PERIOD STRUCTURES


EC 8 DESIGN BASED SECTIONS

(A) Rectangular beam column sections (G1 set of ground motions)

rd
Ir

q 0.879
1.13
1.381 s0.51 0.32

(D1.1)

rd0.89

(D1.2)

1.22 H

(1 0.122r 0.65 s2.12 0.33 )

(q 1)r 0.87
12.4i 0.54
19.45 s0.91

0.932 (2.46 0.28 s 0.081

(D1.3)

(D1.4)

Vb
0.265q 0.095r 0.079) 0.94
W

(D1.5)

1
(31238.32 1275.23 s 71.3Vb 52.8r 2231.32q)
H2

cp 1.823 0.1591 s0.8893 0.3296 0.6891 2.423

pc
0.7981m' 0.2671Cs 0.1378Clvcc 0.2387r
pl

Pc 0.2538 0.0246 s0.7792 0.01921 0.06651 0.191

pc
0.0387m' 0.0682Cs 0.00962Clvcc 0.1287r
pl

DI (0.0268 s0.8951 0.0641q 0.0721r ) 1.172

where r is the beam column strength ratio and is defined as r

(D1.6)

(D1.7)

(D1.8)

I bb
, and Ibb, Icc are moment of inertia of beams and columns
4 I cc

365

(B) Rectangular beam column sections (G2 set of ground motions)

rd

q 0.668
1.21
1.352 s0.46 0.37

I r0.763

rd0.856
1.187 H

(D1.9)
(D1.10)

(1 0.139r 0.598 s1.97 0.45 )

(q 0.983)r 0.94
11.7 i 0.569
15.26 s0.949

0.901 (2.236 0.249 s 0.0721

1
H

1.834

(D1.11)
(D1.12)

Vb
0.312q 0.109r 0.086) 0.963
W

(D1.13)

(34126.76 1033.12 s 68.23Vb 46.82r 2015.67q) 1.084

cp 1.651 0.1011 s0.8712 0.2893 0.6024 2.873

pc
0.7742m' 0.2894Cs 0.065Clvcc 0.210r
pl

Pc 0.2781 0.0313 s0.7893 0.02014 0.02014 0.06781

pc
0.0324m' 0.0768Cs 0.0112Clvcc 0.1104r
pl

DI (0.03124 s0.9563 0.0706q 0.0789r ) 1.791

(D.1.14)

(D1.15)

(D1.16)

366

(C) Square beam column sections (G1 set of ground motions)

rd
I

0.759
r

q 0.718
1.087
1.413 s0.531 0.374

rd0.803
1.196 H

(1 0.129r 0.605 s2.05 0.52 )

(q 1.123)r 0.893
21.76 s0.921

1
H

2.103

(D1.18)
(D1.19)

13.17i 0.584

0.986 (2.612 0.312 s 0.101

(D1.17)

(D1.20)

Vb
0.2013q 0.103r 0.096) 1.164
W

(D1.21)

(32321.76 1206.62 s 66.76Vb 45.12r 2341.65q) 0.9843

cp 1.9712 0.1023 s0.8784 0.3014 0.5913 2.105

pc
0.8112m' 0.1714Cs 0.1163Clvcc 0.2013r
pl

Pc 0.2813 0.03312 s0.7893 0.02043 0.06982 0.211

pc
0.04342m' 0.07165Cs 0.0123Clvcc 0.1132r
pl

DI (0.02781 s0.9682 0.06604q 0.07123r ) 1.165

(D1.22)

(D1.23)

(D1.23)

367

(D) Square beam column sections (G2 set of ground motions)

rd

q 0.698
1.14
1.383 s0.74 0.45

I r0.782

1.123
rd

1.021H

(D1.25)
(D1.26)

(1 0.121r 0.486 s2.35 0.439 )

(q 0.924)r 0.783
13.94i 0.563
0.948
17.56 s

0.838 (2.419 0.285 s 0.085

1
H

1.862

(D1.27)

(D1.28)

Vb
0.2106q 0.093r 0.0923) 1.113
W

(D1.29)

(32912.35 1144.56 s 59.73Vb 41.12r 1967.73q) 0.9931

pc
0.7981m' 0.2471Cs 0.1286Clvcc 0.2290r
pl
pc
Pc 0.2913 0.03496 s0.8013 0.02273 0.07214 0.2248
0.04621m' 0.07284C s 0.0148Clvcc 0.1217r
pl

(D1.30)

DI (0.02832 s0.9793 0.06482q 0.06934r ) 1.121

(D1.32)

cp 1.823 0.1231 s0.8893 0.3269 0.6891 2.243

368

(D1.31)

IS 456 DESIGN BASED SECTIONS


(A) Rectangular beam column sections (G1 set of ground motions)

rd

q 0.783
1.1213
1.265 s0.523 0.302

I r0.912

rd0.982
1.121H

(D.1.33)

(D1.34)

(1 0.11r 0656 s2.245 0.347 )

(q 0.923)r 0.912
11.11i 0.538
18.79 s0.886

0.775 (2.37 0.261 s 0.0743

1
H

1.878

(D1.35)

(D1.36)

Vb
0.289q 0.084r 0.0693) 0.916
W

(D1.37)

(33452.34 1123.45 s 76.45Vb 43.42r 2112.34q) 0.9913

cp 1.715 0.1472 s0.8376 0.3034 0.6612 2.254

pc
0.7705m' 0.2413Cs 0.1271Clvcc 0.2219r
pl

Pc 0.2642 0.0265 s0.7792 0.02158 0.06213 0.2013

pc
0.0321m' 0.07230C s 0.00989Clvcc 0.1129r
pl

DI (0.02721 s0.9680 0.0635q 0.07023r ) 1.129

(D1.38)

(D1.39)

(D1.40)

369

(B) Rectangular beam column sections (G2 set of ground motions)

rd

q 0.612
1.226
1.298 s0.773 0.4874

I r0.784

rd0.049
1.165H

(D1.41)
(D1.42)

(1 0.127r 0584 s2..13 0.552 )

(q 0.864)r 0.94
10.11i 0.532
0.914
14.27 s

0.849 (2.085 0.223 s 0.0673

1
H

1.834

(D1.43)

(D1.44)

Vb
0.339q 0.097r 0.079) 0.932
W

(D1.45)

(31126.76 1233.12 s 57.23Vb 38.65r 2815.67q) 1.084

pc
0.7821m' 0.3123Cs 0.079Clvcc 0.223r
pl
pc
Pc 0.2943 0.0385 s0.7993 0.01963 0.0573 0.1963
0.0425m' 0.0783C s 0.00118Clvcc 0.1254r
pl

(D1.46)

DI (0.03098 s0.9673 0.0713q 0.0813r ) 1.198

(D1.48)

cp 1.625 0.1132 s0.8513 0.2674 0.6175 3.042

370

(D1.47)

(C) Square beam column sections (G1 set of ground motions)


rd

q 0.823
1.065
1.346 s0.510 0.413

I r0.934

rd0.945
1.213H

(D1.49)
(D1.50)

(1 0.098r 0629 s2.110 0.329 )

(q 1.147)r 0.911
13.64i 0.605
0.943
19.13 s

0.784 (2.463 0.270 s 0.0823

1
H

2.0928

(D1.51)

(D1.52)

Vb
0.2762q 0.110r 0.0764) 1.129
W

(D1.53)

(30234.67 1312.34 s 60.785Vb 53.23r 2245.78q) 0.9671

pc
0.7873m' 0.2243Cs 0.096Clvcc 0.2071r
pl
pc
Pc 0.2856 0.03112 s0.8112 0.02218 0.06478 0.2113
0.04523m' 0.07318C s 0.00152Clvcc 0.1297r
pl

(D1.54)

DI (0.02423 s0.9434 0.0611q 0.0689r ) 1.034

(D1.56)

cp 1.8892 0.09876 s0.8562 0.2984 0.6412 2.125

371

(D1.55)

(D) Square beam column sections (G2 set of ground motions)

rd

q 0.602
1.162
1.289 s0.5342 0.4011
1.142
rd

I r0.805

(D1.57)
(D1.58)

(1 0.079r 0497 s2043 0.310 )

1.164 H
(q 0.87)r 0.765
11.39i 0.587
16.23 s0.923

0.869 (1.998 0.257 s 0.0634

1
H

1.763

(D1.59)
(D1.60)

Vb
0.2529q 0.101r 0.0743) 1.087
W

(D1.61)

(32312.65 1127.65 s 54.612Vb 49.83r 2011.54q) 0.9756

pc
0.7421m' 0.2134Cs 0.089Clvcc 0.1892r
pl
pc
Pc 0.3123 0.03245 s0.7918 0.02312 0.07109 0.2312
0.04683m' 0.07562C s 0.00168Clvcc 0.1319r
pl

(D1.62)

DI (0.02342 s0.9672 0.06894q 0.0721r ) 1.102

(D1.64)

cp 1.9132 0.09451 s0.8436 0.2782 0.5572 2.110

372

(D1.63)

APPENDIX E: MEAN VALUES OF SEISMIC RESPONSE PARAMETERS


FOR LONG PERIOD STRUCTURES
(A) MASS IRREGULAR BUILDINGS
1
M200

M400

0.9

M600
M800
M1000

0.8
1

10

11

Building Category

Figure E.1 Mean values of irregularity index for building models with mass irregularity
(Building categories 1 to 11 denote building storeys from 6 to 36)
6.5
M200

Mean q

5.2

M400

3.9

M600

2.6

M800
M1000

1.3
1

10

11

Building Category

Figure E.2 Mean values of behavior factor for building models with mass irregularity

rd (cm)

5
4

M200

M400
M600

M800

M1000

0
1

10

11

Building Category

Figure E.3 Mean values of roof displacement for building models with mass
irregularity

373

0.3

Ird (cm)

M200

0.2

M400
M600
M800

0.1

M1000

0
1

10

11

Building Category

Figure E.4 Mean values of interstorey drift for building models with mass irregularity

Mean i(rad)

0.033
M200
M400

0.022

M600
M800
M1000

0.011
1

10

11

Building Category

Figure E.5 Mean values of rotational demand for building models with mass
irregularity

Mean

3.2
M200

2.4

M400

1.6

M600
M800

0.8

M1000

0
1

10

11

Building Category

Figure E.6 Mean values of overstrength factor for building models with mass
irregularity

374

Mean max(kN/m2)

8000
M200

6000

M400

4000

M600
M800

2000

M1000

0
1

10

11

Building Category

Figure E.7 Mean values of stress demands for building models with mass Irregularity

Mean cp (g)

5.6
M200

4.95

M400

4.3

M600
M800

3.65

M1000

3
1

10

11

Building Category

Figure E.8 Mean values of collapse capacity for building models with mass
irregularity

Mean T (Sec)

3.5
2.9

M200

2.3

M400
M600

1.7

M800

1.1

M1000

0.5
1

10

11

Building Category

Figure E.9 Mean values of fundamental time period for building models with mass
irregularity

375

(B) STIFFNESS IRREGULAR BUILDINGS

Mean c

1
0.75

S25
S50

0.5

S75
S100

0.25
0
1

10

11

Building Category

Figure E.10 Mean values of irregularity index for building models with stiffness
irregularity
7.5

Mean q

S25

4.5

S50

S75
S100

1.5
0
1

10

11

Building Category

Figure E.11 Mean values of behavior factor for building models with stiffness
irregularity

Mean rd (cm)

6
4.5

S25
S50

S75
S100

1.5
0
1

10

11

Building Category

Figure E.12 Mean values of roof displacement for building models with stiffness
irregularity

376

Mean Ird (cm)

0.4
S25

0.3

S50
S75

0.2

S100

0.1
1

10

11

Building Category

Figure E.13 Mean values of interstorey drift for building models with stiffness
irregularity

Mean i (rad)

0.04
S25

0.03

S50
S75

0.02

S100

0.01
1

10

11

Building Category

Figure E.14. Mean values of rotational demand for building models with stiffness

Mean

irregularity
3
2.5
2
1.5
1
0.5
0

S25
S50
S75
S100

10

11

Building Category

Figure E.15 Mean values of overstrength factor for building models with stiffness
irregularity

377

Mean max (kN/m2)

9000
7500

S25
S50

6000

S75
S100

4500
3000
1

10

11

Building Category

Figure E.16 Mean values of stress demands for building models with stiffness
irregularity

Mean cp (g)

5
4.5

S25

S50

3.5

S75
S100

3
2.5
1

10

11

Building Category

Figure E.17 Mean values of collapse capacity for building models with stiffness
irregularity

Mean T (Sec)

3.5
2.9

S25

2.3

S50

1.7

S75
S100

1.1
0.5
1

10

11

Building Category

Figure E.18 Mean values of fundamental time period for building models with
stiffness irregularity

378

(C) STRENGTH IRREGULAR BUILDINGS

Mean c

1
0.75

ST25
ST50

0.5

ST75
ST100

0.25
0
1

10

11

Building Category

Figure E.19 Mean values of irregularity index for building models with strength
Irregularity

Mean q

7.5
6

ST25
ST50

4.5

ST75
ST100

3
1.5
1

10

11

Building Category

Figure E.20 Mean values of behavior factor for building models with strength
irregularity

Mean rd (cm)

6.3
5.1

ST25
ST50

3.9

ST75
ST100

2.7
1.5
1

10

11

Building Category

Figure E.21 Mean values of roof displacement for building models with strength
irregularity

379

Mean Ird (cm)

0.44
0.33

ST25
ST50

0.22

ST75
ST100

0.11
0
1

10

11

Building Category

Figure E.22 Mean values of interstorey drift for building models with strength
irregularity

Mean i(rad)

0.04
ST25

0.03

ST50
ST75

0.02

ST100

0.01
1

10

11

Building Category

Figure E.23 Mean values of rotational demand for building models with strength
irregularity

Mean

2.8
2.1

ST25
ST50

1.4

ST75
ST100

0.7
0
1

10

11

Building Category

Figure E.24 Mean values of overstrength factor for building models with strength
irregularity

380

Mean max (kN/m2)

9000
7300

ST25
ST50

5600

ST75
ST100

3900
2200
1

10

11

Building Category

Figure E.25 Mean values of stress for building models with strength irregularity

Mean cp(g)

4.6
4.1

ST25
ST50

3.6

ST75
ST100

3.1
2.6
1

10

11

Building Category

Figure E.26 Mean values of collapse capacity for building models with strength

Mean T(Sec)

irregularity
3.5
3
2.5
2
1.5
1
0.5
0

ST25
ST50
ST75
ST100

10

11

Building Category

Figure E.27 Mean values of fundamental time period for building models with strength
irregularity

381

(D) SETBACK IRREGULAR BUILDINGS

Mean c

1
0.75

SEA

0.5

SEB
SEC

0.25
0
1

10

11

Building Category

Figure E.28 Mean values of irregularity index for building models with setback
irregularity

7.5
Mean q

6
SEA

4.5

SEB

SEC

1.5
0
1

10

11

Building Category
Figure E.29 Mean values of behavior factor for building models with setback
irregularity

Mean rd (cm)

6.9
5.675
SEC

4.45

SEB
SEA

3.225
2
1

10

11

Building Category

Figure E.30 Mean values of roof displacement for building models with setback
irregularity

382

Mean Ird (cm)

0.5
0.4
SEC

0.3

SEB

0.2

SEA

0.1
0
1

10

11

Building Category

Figure E.31 Mean values of interstorey drift for building models with setback
irregularity

Mean I (rad)

0.05
0.04
SEC

0.03

SEB
SEA

0.02
0.01
1

10

11

Building Category

Figure E.32 Mean values of rotational demand for building models with setback
irregularity

Mean

2.5
2

SEC
SEB
SEA

1.5
1
1

10

11

Building Category

Figure E.33 Mean values of overstrength factor for building models with setback
Irregularity

383

Mean max (kN/m2)

10000
8250
SEC

6500

SEB
SEA

4750
3000
1

10

11

Building Category

Figure E.34 Mean values of stress for building models with setback irregularity

Mean cp (g)

4.3
3.85
SEC

3.4

SEB
SEA

2.95
2.5
1

10

11

Building Category

Figure E.35 Mean values of collapse capacity for building models with setback
Irregularity

Mean T(Sec)

3.7
2.9
SEA

2.1

SEB
SEC

1.3
0.5
1

10

11

Building Category

Figure E.36 Mean values of fundamental time period for building models with setback

384

(E) PLAN IRREGULAR BUILDINGS

Mean c

1
P1

0.9

P2
P3

0.8
1

10

11

Building Category

Figure E.37 Mean values of irregularity index for Building models with plan
irregularity
6.5
Mean q

5.2
3.9

P1

2.6

P3

P2

1.3
0
1

10

11

Building Category

Figure E.38 Mean values of behavior factor for building models with plan irregularity

Mean rd (cm)

5.2
3.9
P1

2.6

P2
P3

1.3
0
1

10 11

Building Category
Figure E.39 Mean values of roof displacement for building models with plan
irregularity

385

Mean Ird (cm)

0.36
0.24

P1
P2
P3

0.12
0
1

10 11

Building Category
Figure E.40 Mean values of interstorey drift for building models with plan irregularity

Mean i(rad)

0.035
P1
P2

0.023

P3

0.011
1

10

11

Building Category

Figure E.41 Mean values of rotational demand for building models with plan
irregularity

Mean

3
2.5

P1
P2
P3

2
1.5
1

10

11

Building Category

Figure E.42 Mean values of overstrength factor for building models with plan
irregularity

386

Mean (kN/m )

8500
7375
P1

6250

P2
P3

5125
4000
1

10

11

Building Category

Figure E.43 Mean values of stress for building models with plan irregularity

Mean cp (g)

5.3
P1

4.15

P2
P3

3
1

10

11

Building Category

Figure E.44 Mean values of collapse capacity for building models with plan
irregularity

Mean T (Sec)

3.2
2.4
P1
P2

1.6

P3

0.8
0
1

10

11

Building Category

Figure E.45 Mean values of fundamental time period for building models with plan
irregularity

387

APPENDIX F: MEAN VALUE OF SEISMIC RESPONSE PARAMETERS FOR


SHORT PERIOD STRUCTURES
(A) MASS IRREGULAR BUILDINGS
1

Mean s

M200
M400

0.9

M600
M800
M1000

0.8
1

Building Category

Figure F.1 Mean values of irregularity index for Building models with mass irregularity
(Building category 1 to 5 denote buildings with storeys ranging from 4 to 12)
3
M200

2.5

M400
M600
M800

M1000

1.5
1

Building Category

Mean rd (Cm)

Figure F.2 Mean values of behavior factor for building models with mass irregularity

2.15

M200
M400
M600

1.825

M800
M1000

1.5
1

Building Category

Figure F.3 Mean values of roof displacement for building models with mass
irregularity

388

Mean Ird(Cm)

0.24
M200

0.18

M400
M600
M800

0.12

M1000

0.06
1

Building Category

Figure F.4 Mean values of interstorey drift for building models with mass irregularity

Mean i(rad)

0.018
M200
M400

0.009

M600
M800
M1000

0
1

Building Category

Figure F.5 Mean values of rotational demand for building models with mass
irregularity
3.5

Mean

M200
M400

3.25

M600
M800
M1000

3
1

Building Category

Figure F.6 Mean values of overstrength factor for building models with mass
irregularity

389

Mean (kN/m2)

3600
M200
M400

2800

M600
M800
M1000

2000
1

Building Category

Figure F.7 Mean values of stress for building models with mass irregularity
3.3
M200

cp (g)

3.1

M400

2.9

M600
M800

2.7

M1000

2.5
1

Building Category

Figure F.8 Mean values of collapse capacity for building models with mass
irregularity

Mean T (Sec)

1.2
M200

0.9

M400
M600
M800

0.6

M1000

0.3
1

Building Category

Figure F.9. Mean values of fundamental time period for building models with mass
irregularity

390

(B) STIFFNESS IRREGULAR BUILDINGS

Mean s

1
0.75

S25
S50

0.5

S75
S100

0.25
0
1

Building Category

Figure F.10 Mean values of irregularity index for Building models with stiffness
4.8

Mean q

4.1

S25
S50

3.4

S75
S100

2.7
2
1

Building Category

Figure F.11 Mean values of behavior factor for building models with stiffness
irregularity

Mean rd (cm)

3
S25
S50

2.25

S75
S100

1.5
1

Building Category

Figure F.12 Mean values of roof displacement for building models with stiffness
irregularity

391

Mean Ird (cm)

0.32
S25
S50

0.23

S75
S100

0.14
1

Building Category

Figure F.13 Mean values of interstorey drift for building models with stiffness
irregularity

Mean i(rad)

0.022
S25
S50

0.015

S75
S100

0.008
1

Building Category

Figure F.14 Mean values of rotational demand for building models with stiffness
irregularity

Mean

3.4
S25
S50

3.075

S75
S100

2.75
1

Building Category

Figure F.15 Mean values of overstrength factor for building models with stiffness
irregularity

392

Mean (kN/m )

5000
S25

4000

S50
S75

3000

S100

2000
1

Building Category

Figure F.16 Mean values of stress for building models with stiffness irregularity

Mean cp(g)

3
2.75

S25
S50

2.5

S75
S100

2.25
2
1

Building Category

Figure F.17 Mean values of collapse capacity for building models with stiffness
irregularity

Mean T (Sec)

1.4
1.1

S25
S50

0.8

S75
S100

0.5
0.2
1

Building Category

Figure F.18 Mean values of fundamental time period


stiffness irregularity

393

for building models with

(C) STRENGTH IRREGULAR BUILDINGS

Mean s

1
ST25
ST50

0.8

ST75
ST100

0.6
1

Building Category

Figure F.19 Mean values of irregularity index for building models with strength
irregularity
5.2

Mean q

4.4

ST25
ST50

3.6

ST75
ST100

2.8
2
1

Building Category

Figure F.20 Mean values of behavior factor for building models with strength
irregularity

Mean rd (cm)

3.2
2.65

ST25
ST50

2.1

ST75
ST100

1.55
1
1

Building Category

Figure F.21 Mean values of roof displacement for building models with strength
irregularity

394

Mean I rd(cm)

0.35
ST25
ST50

0.26

ST75
ST100

0.17
1

Building Category
Figure F.22 Mean values of interstorey drift for building models with strength

Mean i(rad)

irregularity

0.022
ST25
ST50

0.015

ST75
ST100

0.008
1

Building Category

Figure F.23 Mean values of rotational demand for building models with strength
irregularity

Mean

3.4
ST25
ST50

3.05

ST75
ST100

2.7
1

Building Category

Figure F.24 Mean values of overstrength factor for building models with strength
irregularity

395

Mean (kN/m )

5500
4625

ST25
ST50

3750

ST75
ST100

2875
2000
1

Building Category

Figure F.25 Mean values of stress for building models with strength irregularity

Mean cp(g)

2.8
2.6

ST25
ST50

2.4

ST75
ST100

2.2
2
1

Building Category

Figure F.26 Mean values of collapse capacity for building models with strength
irregularity

Mean T (Sec)

1.6
1.3

ST25
ST50

ST75
ST100

0.7
0.4
1

Building Category

Figure F.27 Mean values of fundamental time period for building models with strength
irregularity

396

(D) SETBACK IRREGULARITY


1
0.75

SEC

0.5

SEB
SEA

0.25
0
1

Building Category

Figure F.28 Mean values of irregularity index for building models with setback
irregularity
4.5

Mean q

4
SEC

3.5

SEB
SEA

3
2.5
1

Building Category

Figure F.29 Mean values of behavior factor for building models with setback
irregularity
3.8

rd (cm)

3.35
SEC

2.9

SEB
SEA

2.45
2
1

Building Category

Figure F.30 Mean values of roof displacement for building models with setback
irregularity

397

Mean Ird(cm)

0.39
0.31
SEC

0.23

SEB
SEA

0.15
0.07
1

Building Category

Figure F.31 Mean values of interstorey drift for building models with setback
irregularity

Mean i(rad)

0.026

SEC

0.017

SEB
SEA

0.008
1

Building Category

Figure F.32 Mean values of rotational demand for building models with setback
irregularity

Mean

3.4

SEC

3.05

SEB
SEA

2.7
1

Building Category

Figure F.33 Mean values of overstrength factor for building models with setback
irregularity

398

Mean (kN/m )

6500
5375
SEC

4250

SEB
SEA

3125
2000
1

Building Category

Figure F.34 Mean values of stress for building models with setback irregularity

cp (g)

2.6

2.4

SEC
SEB
SEA

2.2

2
1

Building Category

Figure F.35 Mean values of collapse capacity for building models with setback
irregularity

Mean T (Sec)

1.8

1.3

SEC
SEB
SEA

0.8

0.3
1

Building Category

Figure F.36 Mean values of fundamental time period for building models with setback
irregularity

399

(E) PLAN IRREGULAR BUILDINGS

Mean s

P1

0.875

P2
P3

0.75
1

Building Category

Figure F.37 Mean values of irregularity index for Building models with plan
irregularity

Mean q

3.2

P1

2.6

P2
P3

2
1

Building Category

Figure F.38 Mean values of behavior factor for building models with plan irregularity

Mean rd(cm)

2.4

P1

2.05

P2
P3

1.7
1

Building Category

Figure F.39 Mean values of roof displacement for building models with plan
irregularity

400

Mean Ird(cm)

0.24

P1

0.205

P2
P3

0.17
1

Building Category

Figure F.40 Mean values of interstorey drift for building models with plan irregularity

Mean i(rad)

0.023

P1

0.015

P2
P3

0.007
1

Building Category

Figure F.41 Mean values of rotational demand for building models with plan
irregularity

3.5

P1

3.25

P2
P3

3
1

Building Category

Figure F.42 Mean values of overstrength factor for building models with plan
irregularity

401

Mean (kN/m )

4500

P1

3250

P2
P3

2000
1

Building Category
Figure F.43 Mean values of stress for building models with plan irregularity

Mean cp(g)

P1
P2
P3

0
1

Building Category

Figure F.44 Mean values of collapse capacity for building models with plan
irregularity

Mean T (Sec)

1.2

0.9

P3
P2
P1

0.6

0.3
1

Building Category

Figure F.45 Mean values of fundamental time period for building models with plan

402

APPENDIX G: ESTIMATION OF MEAN RELATION FACTORS


This section aims to estimate estimation of mean relation factors between seismic
responses evaluated by Ibarra et al. (2005) model and Dutta and Das (2002) model. The
Dutta and Das (2002) model has been described in Chapter 1 and Chapter 2. This model is
based on simple one storey idealized system and is efficient in capturing the effects of
stiffness and strength deterioration (Dutta and Das 2002). Moreover this model is simple
and requires less number of parameters as compared to Ibarras model but, Ibarras model
is more accurate as it is based on experimental test results. Therefore, this section aims to
estimate the mean relation factors between seismic response parameters estimated using
both models to achieve simplicity and accuracy in estimation of seismic response
parameters. To achieve this purpose, the seismic responses of long period and short period
structures are estimated for main cases (generating maximum seismic response) as
previously described. Based on the analytical study, the mean relation factor hm (expressed
as the ratio of seismic response evaluated using Ibarra et al.2005 hysteretic model to Dutta
and Das 2002 model) evaluated for irregular building models.
Table G.1. Mean relation factor hm for long period structures
Parameter

ST

SE

rd

1.008

1.0151

1.0372

1.061

1.021

Ird

1.012

1.0186

1.042

1.064

1.024

1.006

1.0137

1.0387

1.056

1.018

0.995

0.9891

0.9632

0.945

0.983

1.012

1.0325

1.055

1.076

1.022

cp

0.998

0.9865

0.9621

0.941

0.973

Pc

1.014

1.0156

1.039

1.054

1.024

1.003

1.0132

1.0276

1.0412

1.034

403

Table G.2 Mean relation factor hm for short period structures


Parameter

ST

SE

rd

1.003

1.0151

1.0372

1.061

1.021

Ird

1.007

1.0186

1.042

1.064

1.024

1.002

1.0137

1.0387

1.056

1.018

0.997

0.9891

0.9632

0.945

0.983

1.008

1.0325

1.055

1.076

1.022

cp

0.995

0.9865

0.9621

0.941

0.973

Pc

1.011

1.0156

1.039

1.054

1.024

1.002

1.0132

1.0276

1.0412

1.034

404

S-ar putea să vă placă și