thesis (2)
thesis (2)
STRUCTURES IN BANGLADESH
MOHAMED KAMRUZZAMAN
October, 2011
SEISMIC VULNERABILITY OF SOME HISTORICAL MASONRY
STRUCTURES IN BANGLADESH
by
MOHAMED KAMRUZZAMAN
October, 2011
The thesis titled “Seismic Vulnerability of Some Historical Masonry Structures in
Bangladesh” Submitted by Mohamed Kamruzzaman, Roll No.: 040504348(F),
Session: April 2005, has been accepted as satisfactory in partial fulfillment of
requirement for the degree of Master of Science in Civil Engineering (Structural) on 29th
October, 2011.
BOARD OF EXAMINERS
It is hereby declared that this thesis or any part of it has not been submitted elsewhere for
the award of any degree or diploma.
_____________________
Mohamed Kamruzzaman
DEDICATION
Dedicated to My Mother
ACKNOWLEDGEMENT
All the praises is for Almighty, Allah. This work has become true for His wishing and
helping. The author wishes to express his profound gratitude and sincere appreciation to
his supervisor Dr. Munaz Ahmed Noor, Professor of the Department of Civil
Engineering, Bangladesh University of Engineering and Technology (BUET), Dhaka, for
his dynamic supervision, continuous guidance, helpful criticism, invaluable suggestion,
enthusiastic encouragement, strong support and persistent simulation. His fervent
guidance in every aspect of this work was the most valuable experience of the life of
author and also would remain so forever. The author also wishes to express his gratitude
and sincere appreciation to Prof. Dr. Mehedi Ahmed Ansary for his precious suggestion.
The author gratefully acknowledges the help, encouragement, cooperation received from
his friends and associates. The author is ever grateful to his parents for their inspiration,
encouragement and blessings.
i
ABSTRACT
ii
dynamic properties of the buildings and conducted in the vicinity of the buildings to
determine the period of the site and asses the potential effects of soil-structure interaction,
which could have a significant effect on the seismic performance of the buildings during
a severe earthquake.
In terms of average values of six indices, almost all mosques are safe except eight
(28.58%) mosques of the samples. For the high seismicity, these eight mosques are
vulnerable in the short direction only and required careful attention and deeper
investigation at risk. The results of the Microtremor tests were successful at obtaining the
site periods for the two locations, which were 0.26 second for Musa khan mosque and
0.13 second for Aambour Shah Shahi mosque. The dynamic response of the older
masonry structures during a severe earthquake can be significantly affected by soil-
structure interaction effects, and should be considered for possible retrofit design.
iii
TABLE OF CONTENTS
TITLE PAGE
Acknowledgement i
Abstract ii
Table of Contents iv
List of Figures ix
List of Tables xix
List of Symbols xx
List of Abbreviations xxii
CHAPTER 1: INTRODUCTION
1.1 General 1
1.2 Background and Present State of the Problem of the Study 1
1.3 Objectives with Specific Aims 2
1.4 Outline of Methodology 3
1.5 Scope of the Study 4
1.6 Organization of the Thesis 4
CHAPTER 2: LITERATURE REVIEW
2.1 Introduction 6
2.2 Empirical Methods of Seismic Vulnerability Assessment 9
2.2.1 Damage Probability Matrices 10
2.2.2 Vulnerability Index Method 14
2.2.3 Continuous Vulnerability Curves 16
2.2.4 Screening Methods 18
2.3 Analytical/Mechanical Seismic Vulnerability Assessment Methods 20
2.3.1 Analytically-Derived Vulnerability Curves and DPMs 21
2.3.2 Hybrid Methods 24
2.3.3 Collapse Mechanism-Based Methods 26
2.3.4 Capacity Spectrum-Based Methods 29
2.3.5 Fully Displacement-Based Methods 36
2.3.6 General Evaluation of Analytical Methods 43
2.4 Ground Motion Parameters for Seismic Hazard Analysis (Kramer 2003) 44
iv
2.5 Development of Fragility Curves 47
2.5.1 Limit States (Bazzurro et al. 2004) 48
2.5.2 Performance Based Engineering and Probabilistic Safety
Assessment 48
2.5.3 Fragility Curve 49
2.5.4 Computation of the Fragility Curves 50
2.6 In-Plane Behaviour of Existing Masonry Buildings 52
2.6.1 Shear behaviour of unreinforced masonry 52
2.6.2 Structural model 58
2.6.3 Capacity curve of a masonry wall 60
2.6.4 Ductility of masonry structures 64
2.6.5 Cyclic loading 68
2.6.6 Identification of damage grades according to the EMS 68
2.6.7 Vulnerability function of a masonry building 70
2.7 Out-Of-Plane Behaviour of Existing Masonry Buildings 70
2.7.1 Structural Modeling 71
2.7.2 Out-of-plane cracking 73
2.7.3 Out-of-plane failure 75
2.7.4 Influence on the vulnerability function 76
2.7.5 Some further remarks on the out-of-plane behaviour 77
CHAPTER 3: SIMPLIFIED METHOD ANALYSIS
` 3.1 Introduction 79
3.2 Methodology 81
3.2.1 In-Plan Area Ratio 82
3.2.2 Area to Weight Ratio 83
3.2.3 Base Shear Ratio 83
3.2.4 Out-of-plane indices 88
3.2.5 Preliminary Comparative Analysis 89
3.3 Seismic Hazard Zonation of Bangladesh 90
CHAPTER 4: MICROTREMOR ANALYSIS
4.1 Introduction 91
v
4.2 Background 91
4.3 Direct Measurement Technique 92
4.4 Ambient Vibration Technique 93
4.5 Microtremor Measurements 93
CHAPTER 5: SELECTED MASONRY STRUCTURES
5.1 Introduction 96
5.2 Chhota Sona Mosque 98
5.3 Ghorar Mosque 101
5.4 Jore Bangla Mosque 103
5.5 Golakata Mosque 105
5.6 Shait Gambuj Mosque 107
5.7 Noy Gambuj Mosque 109
5.8 Chunakhola Mosque 111
5.9 Ronbijoypur Mosque 113
5.10 Baba Adam Mosque 115
5.11 Rajbibi Mosque 117
5.12 Dhaniachak Mosque 119
5.13 Goaldi Mosque 121
5.14 Bagha Mosque 123
5.15 Sura Mosque 125
5.16 Kherua Mosque 127
5.17 Atiya Mosque 129
5.18 Shah Muhammad Mosque 131
5.19 Shah Niamatullah Mosque 133
5.20 Khawaja Shahbaz Mosque 135
5.21 Sat Gambuj Mosque 137
5.22 Lalbag Fort Mosque 139
5.23 Khan Muhammad Mridha Mosque 141
5.24 Miah Bari Mosque 143
5.25 Bazra Mosque 145
5.26 Dewanbari Mosque 147
vi
5.27 Karzon Hall Musa Khan Mosque 149
5.28 Bibir Mosque 151
5.29 Aambour Shah Shahi Mosque 153
CHAPTER 6: RESULT AND DISCUSSION
6.1 Introduction 155
6.2 Simplified Method Analysis 155
6.1.1 Chhota Sona Mosque 155
6.1.2 Ghorar Mosque 156
6.1.3 Jore Bangla Mosque 158
6.1.4 Golakata Mosque 159
6.1.5 Shait Gambuj Mosque 159
6.1.6 Noy Gambuj Mosque 159
6.1.7 Chunakhola Mosque 163
6.1.8 Ronbijoypur Mosque 164
6.1.9 Baba Adam Mosque 164
6.1.10 Rajbibi Mosque 164
6.1.11 Dhaniachak Mosque 169
6.1.12 Goaldi Mosque 169
6.1.13 Bagha Mosque 169
6.1.14 Sura Mosque 173
6.1.15 Kherua Mosque 173
6.1.16 Atiya Mosque 173
6.1.17 Shah Muhammad Mosque 177
6.1.18 Shah Niamatullah Mosque 177
6.1.19 Khawaja Shahbaz Mosque 177
6.1.20 Sat Gambuj Mosque 181
6.1.21 Lalbag Fort Mosque 181
6.1.22 Khan Muhammad Mridha Mosque 181
6.1.23 Miah Bari Mosque 181
6.1.24 Bazra Mosque 186
6.1.25 Dewanbari Mosque 186
vii
6.1.26 Karzon Hall Musa Khan Mosque 186
6.1.27 Bibir Mosque 186
6.1.28 Aambour Shah Shahi Mosque 186
6.1.29 Comparative Analysis 192
6.2 Microtremor Analysis 197
CHAPTER 7: CONCLUSIONS AND REOMMENDATIONS
7.1 General 200
7.2 Simplified Method Analysis 200
7.3 Microtremor Analysis 201
7.4 Recommendations 201
REFERENCES 202
APPENDIX A 214
APPENDIX B 216
APPENDIX C 225
APPENDIX D 229
viii
LIST OF FIGURES
Figure 2.1: The components of seismic risk assessment and choices for the vulnerability
assessment procedure; the bold path shows a traditional assessment method ……8
Figure 2.2: Example of a lognormal distribution of the estimated damage factor for a
given intensity showing the mean low (ML), mean best (MB), and mean high
(MH) estimates of the damage factor (adapted from McCormack and Rad
(1997))……………………………………………………………………………12
Figure 2.3: Vulnerability functions to relate damage factor (d) and peak ground
acceleration (PGA) for different values of vulnerability index (Iv) adapted from
Guagenti and Petrini (1989))…………………………………………………….15
Figure 2.4: Vulnerability curves produced by Spence et al. (1992) for bare moment-
resisting frames using the parameterless scale of intensity (PSI); D1 to D5 relate
to damage states in the MSK scale………………………………………………17
Figure 2.5: Example of the difference in the vulnerability point distribution using the
observations of low and mid-rise building damages after the 1995 Aegion
(Greece) earthquake for different ground motion parameters: (a) PGA, and (b)
Spectral displacement at the elastic fundamental period (Rossetto and Elnashai,
2003)……………………………………………………………………………18
Figure 2.8: Simulated (thick line) and observed (thin line) vulnerability functions for
MSK intensity VII (Barbat et al., 1996)…………………………………………23
Figure 2.9: Out-of-plane failure mechanisms (after D’Ayala and Speranza, 2002)……..25
Figure 2.11: Illustration of the estimation of damage from ground shaking in HAZUS
(FEMA, 1999, 2003)……………………………………………………………..30
ix
Figure 2.12: Modules involved in the LNECloss tool (Sousa et al., 2004)……………32
Figure 2.13: 1999 Kocaeli earthquake, Gölcük Coastal Zone: Results of various analyses
and comparison with the damage data for the mid-rise reinforced concrete frames
(Spence et al., 2003)……………………………………………………………...37
Figure 2.16: Deformed shapes for different limit states and in-plane failure modes……38
Figure 2.17: An example of the intersection of capacity areas and demand spectrum
(Calvi, 1999)……………………………………………………………………..39
Figure 2.18: Pushover results for a four-storey building with the simplified and three-
dimensional methodologies, and for different referential shear strength values...41
Figure 2.19: An illustration of the deterministic comparison of the limit state (LS)
displacement capacity and displacement demand in DBELA…………………...42
Figure 2.20: (a) Illustrative probability density functions of yield displacement capacity,
conditioned to period; (b) Illustrative probability density function of yield period
for a building class due to variability in structural properties; (c) Example JPDF
of capacity for a columnsway RC building class………………………………...42
Figure 2.22: Fragility curves for onset of damage, green, yellow, and red tags, and for
collapse of the building (Bazzurro et al. 2004)…………………………………..51
Figure 2.23: Different types of failure that can occur in masonry walls………………...53
Figure 2.24: Failure conditions for unreinforced masonry according to [Ga 85]………..54
Figure 2.25: Compression strength as a function of the angle after [Ga 85]…………….55
Figure 2.26: Internal forces and corresponding stress field of a wall element…………..56
x
Figure 2.29: Geometry of a pier + applied forces………………………………………61
Figure 2.31: Bending moment and shear force distribution of the bottom pier of a wall
due to a) the real forces and b) a virtual unit force………………………………63
Figure 2.32: Relationship between ultimate drift and normal stress of the wall elements
tested at the ETH Zürich (1984)…………………………………………………65
Figure 3.1: Location of earthquake epicenter, at and near Bangladesh, occurrence period
1869 to 2000 (Yasin, 2008)…………………………………………………….80
Figure 3.2: Isoseismal map of 1897 Great Indian earthquake (Ansary and Noor, 2006)82
Figure 3.4: Assumed thresholds as functions of PGA/g: (a) index 1, (b) index 2 and (c)
index 3 (Laurenco and Oliviera, 2004)…………………………………………..86
Figure 3.5: Assumed thresholds as functions of PGA/g: (a) index 4, (b) index 5 and (c)
index 6 (Laurenco and Oliviera, 2004)…………………………………………..87
xi
Figure 4.2: Microtremor observation…………………………………………………….95
Figure 4.3: Placing of sensor at the top of the building with proper direction and
microtremor set up……………………………………………………………….95
xii
Figure 5.22: Front Elevation of Chunakhola mosque…………………………………..113
xiii
Figure 5.46: Front Elevation of Kherua mosque…………………………………….....129
xiv
Figure 5.70: Front Elevation of Miah Bari mosque…………………………………….145
Figure 6.2: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Chhoto Sona Mosque………………………….157
Figure 6.3: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Ghorar Mosque. ………………………………158
Figure 6.4: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Jore Bangla Mosque. ………………………….160
xv
Figure 6.5: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Golakata Mosque. …………………………….161
Figure 6.6: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Shait Gambuj Mosque. ……………………….162
Figure 6.7: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Noy Gamboj Mosque. ………………………...163
Figure 6.8: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Chunakhola Mosque. …………………………165
Figure 6.9: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight ratio,
(c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of perimeter
wall) and PGA/g for Ronbijoypur Mosque. ………………………...166
Figure 6.10: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Baba Adam Mosque. ………………………….167
Figure 6.11: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Rajbibi Mosque. ………………………………168
Figure 6.12: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Dhaniachak Mosque. …………………………170
Figure 6.13: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Goaldi Mosque. ……………………………….171
Figure 6.14: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
xvi
perimeter wall) and PGA/g for Bagha Mosque. ……………………………….172
Figure 6.15: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Sura Mosque. …………………………………174
Figure 6.16: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Kherua Mosque. ………………………………175
Figure 6.17: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Atiya Mosque. ………………………………...176
Figure 6.18: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Shah Muhammad Mosque. …………………...178
Figure 6.19: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Shah Niamatullah Mosque. …………………...179
Figure 6.20: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Khawaza Shahbaz Mosque. …………………..180
Figure 6.21: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Sat Gambuj Mosque. ………………………….182
Figure 6.22: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Lalbag Fort Mosque…………………………...183
Figure 6.23: Relationship between Indices ((a) In-plane area ratio, (b)Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Khan Muhammad Mridha Mosque. …………184
xvii
Figure 6.24: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Miah Bari Mosque. …………………………...185
Figure 6.25: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Bazra Mosque. ………………………………..187
Figure 6.26: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Dewanbabi Mosque. ………………………….188
Figure 6.27: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Musa Khan Mosque. ………………………….189
Figure 6.28: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Bibir Mosque. ………………………………...190
Figure 6.29: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Aambour Shah Shahi Mosque. ……………….191
Figure 6.30: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g along X-axes for all structures. …………………...194
Figure 6.31: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g along Y-axes for all structures. …………………...195
xviii
LIST OF TABLES
Table 2.1: Format of the Damage Probability Matrix Proposed by Whitman et al.
(1973)....………………………………………………………………………….11
Table 2.2: Example of a Damage Model for Vulnerability Class C as Presented in EMS-
98…………………………………………………………………………………12
Table 2.3: Damage probability matrix for sample low-rise buildings (Singhal and
Kiremidjian, 1996)……………………………………………………………….23
xix
LIST OF SYMBOLS
IV = Vulnerability index
G = Ground index
U is a usage index
Sa = Spectral acceleration
ŠaLS = Median spectral capacity values associated with the onset of post-earthquake
structural limit state
βR = Aleatory uncertainty
βU = Epistemic Uncertainty
β = Total Uncertainty
xx
GAeff = Effective shearing stiffness of the pier
γ4 = slenderness ratio,
xxi
LIST OF ABBREVIATIONS
RS = Richter Scale
HAZUS = Hazard US
CONV= convolution
SRSS = square-root-of-the-sum-of-the-squares
xxii
MeBaSe = Mechanical Based Procedure for the Seismic Risk Estimation of Unreinforced
Masonry Buildings
LS = limit states
FC = Fragility Curve
EC = Eurocode
xxiii
CHAPTER ONE
INTRODUCTION
1.1 General
Seismic building structural planning is important, and since the architects device and
control the planning it follows that he/she is, even if ignorant, a major participant in the
seismic planning of the structure. It is useful for the architects to acquire an understanding
the way of the seismic resistant building configurations, in response to the forces that the
earthquake generates; the detail calculation can be left to the structural design engineers.
In the recent years it has become clear that architectural configuration, the size and shape
of the building, makes a major contributor to the success or failure of the building's
seismic performance.
The seismic risk of different building categories, engineered and non-engineered, has
recently been estimated by combining the information on seismic hazard and structural
vulnerability (Ansary and Noor, 2004). Retrofitting of these existing structures is very
much needed to improve the performance of these structures during earthquakes.
1
Due to the ageing process, as well as to environmental factors, many cultural heritage
masonry buildings, as structures planned and constructed in the past, are vulnerable to
dynamic loads, which may unpredictably induce a collapse of a portion of the building or
drive the whole structure to a rapid failure. However, the high vulnerability of historical
masonry buildings (mosques, temples, churches etc.) to seismic actions is mostly due to
the absence of adequate connections between the various building parts. This
characteristic leads to overturning collapse of the perimeter walls under seismic
horizontal acceleration. These buildings mainly consist of low height (up to three stories),
moderate spans (maximum of four or five meters), and large thickness of the walls (less
than 1/7 of the height) (Giuffrè 1995; Lourenco and Oliviera, 2004). However, this thesis
is focused on Bangladeshi ancient historical mosques, given:
− Their intrinsic greater structural vulnerability due to open plan, greater height to
width ratio and, often, the presence of thrusting horizontal members, vaulted
ceilings and timber roofs.
2
1.4 Outline of Methodology
This research deals with an investigation regarding the possibility of using simplified
methods of analysis and simple indices as indicators for fast screening and decision to
prioritize deeper studies of historical masonry buildings and to assess vulnerability to
seismic loading. These indices, both in-plane and out-of-plane, are established mostly on
the in-plan dimensions and height of the buildings. In general, the longitudinal direction
of the buildings (x) exhibits lower vulnerability than the transversal direction (y).
According to Eurocode 8 (CEN-EC8 2003), walls should only be considered as
earthquake resistant if the thickness is larger than 0.35 m, and the ratio between height
and thickness is smaller than nine.
The proposed geometrical indices of monuments located in different seismic areas are
compared with the respective seismic hazard, expressed by the peak ground acceleration
(PGA), defined for a 10% probability of exceedance in 50 years for a rock-like soil,
corresponding to a return period of 475 years. For Dhaka, Rajshahi and Narayanganj
areas, the peak ground acceleration was estimated to be 0.20g for a 475 year return period
for alluvium (Noor et al., 2005). The recognition of the likely existence of a correlation
between structural characteristics and seismic hazard is, therefore, sought.
Lourenco and Oliveira developed threshold values for analyzing historical structure by
this method (Lourenco and Oliveira, 2004). From this threshold one may identify whether
the structures are vulnerable or not. Vulnerable structure requires more detailed analysis.
In this thesis these threshold values will be used to identify the vulnerable historical
structures.
3
of selected historical masonry structures. The most significant peak of the Fourier
Spectrum is taken to be the dominant frequency of the site.
The aim of this thesis is to assess the seismic vulnerability of historical masonry mosque
by using simplified method and Microtremor measurements. So there are mainly two
parts of the study: a) simplified method Analysis, b) Microtremor measurements,
The usage of simplified methods applies to the design and construction of buildings and
civil engineering works in seismic regions. Its purpose is to ensure, that in the event of
earthquakes, e.g. human lives are protected, damage is limited, structures important for
civil protection remain operational.
In general, the approaches to the identification the dynamic properties of buildings can be
categorized into three main approaches: (1) empirical, (2) numerical analysis, and (3)
direct measurement approaches. The empirical approach provides simplified formulas for
estimating the fundamental periods of buildings in terms of geometric dimensions of the
buildings. The second approach, the numerical analysis, is normally used during the
design process. A finite element model of the building, which consists of the mass and
stiffness matrices of the system, is first formulated. Dynamic properties such as natural
frequencies and vibration mode shapes are obtained by the Eigen analysis. The third
approach is the direct measurement approach, which first measures dynamic responses of
existing buildings, and then identifies their dynamic properties from the measured
responses.
The thesis contains seven chapters. Chapter one introduces the background, objective,
methodology and scope of the study. Literature review has been presented in Chapter
two. Chapter three describes the Simplified method analysis and seismic hazard zonation
4
of Bangladesh. Microtremor measurements analysis is described in chapter four.
Description of selected historical masonry structures are presented in chapter five.
Chapter six presents the result and discussion of the simplified methods and microtremor
measurements analysis of historical masonry structures which located in different district
of Bangladesh. Finally a conclusion has been drawn and recommendations for future
studies in Chapter seven.
5
CHAPTER TWO
LITERATURE REVIEW
2.1 Introduction
Bangladesh is one of the most densely populated country in the world. Much of the
country lies in active seismic zones making the occurrence of deadly earthquakes a
frightening possibility. Many of the earthquakes have occurred in the northeastern and
southeastern part of the country. The Madhupur fault runs in Dhaka division though its
exact position is unknown. The border with the Indian state of Meghalaya is also a fault.
Dhaka and Chittagong have both suffered severe earthquakes in the past. The minor
earthquakes of magnitude ranging from 3-4 on the Richter Scale (RS) are the common
phenomenon in Bangladesh.
Many buildings and physical infrastructures, those were constructed 10, 20, 50, 100 and
200 years ago without considering seismic safety provisions, are now essentially needed
to strengthen or retrofit in order to protect human lives and environmental hazards.
In the last few decades, a dramatic increase in the losses caused by natural catastrophes
has been observed worldwide. Reasons for the increased losses are manifold, though
these certainly include the increase in world population, the development of new “super-
cities” (with a population greater than 2 million), many of which are located in zones of
high seismic hazard, and the high vulnerability of modern societies and technologies (e.g.,
Smolka et al., 2004). The 1994 Northridge (California, US) earthquake produced the
highest ever insured earthquake loss at approximately US$14 billion, and the US$150
billion cost of the 1995 Kobe (Japan) earthquake was the highest ever absolute
earthquake loss. Although the dollar value of economic losses in other parts of the world
may be far lower than in Japan and the US, the impact on the national economy may be
much greater due to losses being a larger proportion of the gross national product (GNP)
in that year. Coburn and Spence (2002) report the economic losses due to earthquakes
from 1972 to 1990; the three largest losses as proportions of the GNP are in the Central
American countries of Nicaragua (1972, 40% GNP), Guatemala (1976, 18% GNP) and El
Salvador (1986, 31% GNP). When the economic burden falls entirely on the government
(such as occurred after the 1999 Kocaeli earthquake in Turkey), the impact on the
national economy can be crippling; one possible solution is to privatise the risk by
offering insurance to homeowners and then to export large parts of the risk to the world’s
6
reinsurance markets (e.g., Bommer et al., 2002). In order to design such insurance and
reinsurance schemes, a reliable earthquake loss model for the region under consideration
needs to be compiled such that the future losses due to earthquakes can be determined
with relative accuracy.
The formulation of an earthquake loss model for a given region is not only of interest for
predicting the economic impact of future earthquakes, but can also be of importance for
risk mitigation. A loss model that allows the damage to the built environment and
important lifelines to be predicted for a given scenario (perhaps the repetition of a
significant historical earthquake) can be particularly important for emergency response
and disaster planning by a national authority. Additionally, the model can be used to
mitigate risk through the calibration of seismic codes for the design of new buildings; the
additional cost in providing seismic resistance can be quantitatively compared with the
potential losses that are subsequently avoided. Furthermore, the loss model can be used to
design retrofitting schemes by carrying out cost/benefit studies for different types of
structural intervention schemes.
Earthquake loss models should, ideally, include all of the possible hazards from
earthquakes: amplified ground shaking, landslides, liquefaction, surface fault rupture, and
tsunamis. Nevertheless, strong ground shaking is often the only hazard considered in loss
assessment methods; this is commonly an acceptable approach because as the size of the
loss model increases, the relative influence of the secondary hazards such as liquefaction
and landslides decreases (Bird and Bommer, 2004). Constructing an earthquake loss
model for a city, region or country involves compiling databases of earthquake activity,
ground conditions, ground-motion prediction equations, building stock and infrastructure
exposure, and vulnerability characteristics of the exposed inventory. The main aim of a
loss model is to calculate the seismic hazard at all the sites of interest and to convolve this
hazard with the vulnerability of the exposed building stock such that the damage
distribution of the building stock can be predicted; damage ratios, which relate the cost of
repair and replacement to the cost of demolition and replacement of the structures, can
then be used to calculate the loss.
7
building type due to a scenario earthquake. The various methods for vulnerability
assessment that have been proposed in the past for use in loss estimation can be divided
into two main categories: empirical or analytical, both of which can be used in hybrid
methods (see Figure 2.1).
Scenario Earthquake
Damage Scale
Fully
Field Expert Displacement
Typological Survey Judgement Based
Figure 2.1: The components of seismic risk assessment and choices for the
vulnerability assessment procedure; the bold path shows a traditional
assessment method
8
acceleration (PGA) have been used, whilst more recent proposals have linked the seismic
vulnerability of the buildings to response spectra obtained from the ground motions.
Each vulnerability assessment method models the damage on a discrete damage scale;
frequently used examples include the MSK scale (Medvedev and Sponheuer, 1969), the
Modified Mercalli scale (Wood and Neumann, 1931) and the EMS98 scale (Grünthal,
1998). In empirical vulnerability procedures, the damage scale is used in reconnaissance
efforts to produce post-earthquake damage statistics, whilst in analytical procedures this
is related to limit-state mechanical properties of the buildings, such as interstorey drift
capacity.
The evolution of vulnerability assessment procedures for both individual buildings and
building classes is described in the following sections, wherein the most important
references, applications and developments pertaining to each methodology are reported.
A larger emphasis has been placed on the vulnerability assessment of the built
environment at an urban scale for use in risk and loss assessment methodologies.
There are two main types of empirical methods for the seismic vulnerability assessment
of buildings that are based on the damage observed after earthquakes, both of which can
be termed “damage-motion relationships”: 1) damage probability matrices (DPM), which
express in a discrete form the conditional probability of obtaining a damage level j , due
9
2.2.1 Damage Probability Matrices
Whitman et al. (1973) first proposed the use of damage probability matrices for the
probabilistic prediction of damage to buildings from earthquakes. The concept of a DPM
is that a given structural typology will have the same probability of being in a given
damage state for a given earthquake intensity. The format of the DPM suggested by
Whitman et al. (1973) is presented in Table 2.1, where example proportions of buildings
with a given level of structural and non-structural damage are provided as a function of
intensity (note that the damage ratio represents the ratio of cost of repair to cost of
replacement). Whitman et al. (1973) compiled DPMs for various structural typologies
according to the damaged sustained in over 1600 buildings after the 1971 San Fernando
earthquake.
One of the first European versions of a damage probability matrix was produced by Braga
et al. (1982), which was based on the damage data of Italian buildings after the 1980
Irpinia earthquake, and this introduced the binomial distribution to describe the damage
distributions of any class for different seismic intensities. The binomial distribution has
the advantage of needing one parameter only which ranges between 0 and 1. On the other
hand it has the disadvantage of having both mean and standard deviation depending on
this unique parameter. The buildings were separated into three vulnerability classes (A, B
and C) and a DPM based on the MSK scale was evaluated for each class. This type of
method has also been termed ‘direct’ (Corsanego and Petrini, 1990) because there is a
direct relationship between the building typology and observed damage. The use of
DPMs is still popular in Italy and proposals have recently been made to update the
original DPMs of Braga et al. (1982). Di Pasquale et al. (2005) have changed the DPMs
from the MSK scale to the MCS (Mercalli-Cancani-Sieberg) scale because the Italian
seismic catalogue is mainly based on this intensity, and the number of buildings has been
replaced by the number of dwellings so that the matrices could be used in conjunction
with the 1991 Italian National Statistical Office (ISTAT) data. Dolce et al. (2003) have
also adapted the original matrices as part of the ENSeRVES (European Network on
Seismic Risk, Vulnerability and Earthquake Scenarios) project for the town of Potenza,
Italy. An additional vulnerability class D has been included, using the EMS98 scale
(Grüntal, 1998), to account for the buildings that have been constructed since 1980. These
buildings should have a lower vulnerability as they have either been retrofitted or
designed to comply with recent seismic codes.
10
Table 2.1: Format of the Damage Probability Matrix Proposed by Whitman et al.
(1973)
Damage Probability Matrices based on expert judgement and opinion were first
introduced in ATC- 13 (ATC, 1985). More than 50 senior earthquake engineering experts
were asked to provide low, best and high estimates of the damage factor (the ratio of loss
to replacement cost, expressed as a percentage) for Modified Mercalli Intensities (MMI)
from VI to XII for 36 different building classes. The low and high damage factor
estimates provided by the experts were defined as the 90% probability bounds of a
lognormal distribution, whilst the best estimate was taken as the median damage factor
(Figure 2.2). Weighted means of the experts’ estimates, based on the experience and
confidence levels of the experts for each building class, were included in the averaging
process. A lognormal distribution was then used to calculate the probability of a central
damage factor by finding the area below the curve within a given damage factor range.
Thus, a DPM could be produced for each intensity level for each building class. Examples
of the use of DPMs based on the ATC-13 approach for the assessment of risk and loss
include the city of Basel (Fah et al., 2001), Bogotá (Cardona and Yamin, 1997) and New
Madrid (Veneziano et al., 2002).
11
Figure 2.2: Example of a lognormal distribution of the estimated damage factor for
a given intensity showing the mean low (ML), mean best (MB), and mean
high (MH) estimates of the damage factor (adapted from McCormack and
Rad (1997))
A macroseismic method has recently been proposed (Giovinazzi and Lagomarsino, 2001,
2004) that leads to the definition of damage probability functions based on the EMS-98
macroseismic scale (Grünthal, 1998). The EMS-98 scale defines qualitative descriptions
of “Few”, “Many” and “Most” for five damage grades for the levels of intensity ranging
from V to XII for six different classes of decreasing vulnerability (from A to F). Damage
matrices containing a qualitative description of the proportion of buildings that belong to
each damage grade for various levels of intensity are presented in Table 2.2 for
vulnerability class C.
The problems related to the “incompleteness” of the matrices (i.e., the lack of information
for all damage grades for a given level of intensity) and the “vagueness” of the matrices
(i.e., they are described qualitatively) have been tackled by Giovinazzi and Lagomarsino
12
(2004) by assuming a beta damage distribution and by applying Fuzzy Set Theory,
respectively. The damage probability matrices produced for each vulnerability class have
been related to the building stock through the use of an empirical vulnerability index
which depends on the building typology, the characteristics of the building stock (e.g.,
number of floors, irregularity, etc.) and the regional construction practices. This
macroseismic method has already been applied in the risk assessment of the cities, Faro
(Oliveira et al., 2004), Lisbon (Oliveira et al., 2005), and Barcelona (Lantada et al.,
2004).
Following the introduction of DPMs based on intensity, the assessment of seismic risk on
a large scale was made possible in both an efficient and cost-effective manner because in
the past seismic hazard maps were also defined in terms of macroseismic intensity. The
use of observed damage data to predict the future effects of earthquakes also has the
advantage that when the damage probability matrices are applied to regions with similar
characteristics, a realistic indication of the expected damage should result and many
uncertainties are inherently accounted for. However, there are various disadvantages
associated with the continued use of empirical methods such as DPM’s:
13
• Seismic hazard maps are now defined in terms of PGA (or spectral ordinates) and
thus PGA needs to be related to` intensity; however, the uncertainty in this
equation is frequently ignored. When the vulnerability is to be defined directly in
terms of PGA, where recordings of the level of the ground shaking at the site of
damage are not available, it might be necessary to predict the ground shaking at
the site using a ground motion prediction equation; however, again the uncertainty
in this equation needs to be accounted for in some way, especially the component
related to spatial variability.
The “Vulnerability Index Method” (Benedetti and Petrini, 1984; GNDT, 1993) has been
used extensively in Italy in the past few decades and is based on a large amount of
damage survey data; this method is ‘indirect’ because a relationship between the seismic
action and the response is established through a ‘vulnerability index’. The method uses a
field survey form to collect information on the important parameters of the building
which could influence its vulnerability: for example, plan and elevation configuration,
type of foundation, structural and non-structural elements, state of conservation and type
and quality of materials. There are eleven parameters in total, which are each identified as
having one of four qualification coefficients, Ki , in accordance with the quality
conditions – from A (optimal) to D (unfavourable) – and are weighted to account for their
relative importance. The global vulnerability index of each building is then evaluated
using the following formula:
11
Iv = ∑ K i W i (2.1)
i =1
The vulnerability index ranges from 0 to 382.5, but is generally normalised from 0 to 100,
where 0 represents the least vulnerable buildings and 100 the most vulnerable. The data
from past earthquakes is used to calibrate vulnerability functions to relate the
vulnerability index (IV) to a global damage factor (d) of buildings with the same typology,
for the same macroseismic intensity or PGA. The damage factor ranges between 0 and 1
and defines the ratio of repair cost to replacement cost. The damage factor is assumed
14
negligible for PGA values less than a given threshold and it increases linearly up until a
collapse PGA, from where it takes a value of 1 (Figure 2.3).
Figure 2.3: Vulnerability functions to relate damage factor (d) and peak ground
acceleration (PGA) for different values of vulnerability index (Iv) adapted
from Guagenti and Petrini (1989))
The “Catania Project”, discussed in Faccioli et al. (1999) and GNDT (2000), used an
adapted vulnerability index method for the risk assessment of both masonry and
reinforced concrete buildings. Some modifications to the original vulnerability index
procedure were applied in that a rapid screening approach was used to define the
vulnerability scores of the buildings, following the guidelines of ATC-21 (ATC, 1988).
As in the original procedure, the vulnerability score was obtained from the weighted sum
of eleven parameters; however, only some were directly obtained from a field assessment,
while the rest were based on a range of values according to historic or recent construction
practices within the region, thus leading to a lower and an upper bound to Iv for each
15
building. Vulnerability functions for old masonry buildings were calibrated to the damage
observed after the 1976 Friuli and the 1984 Abruzzo earthquakes, and damage and peak
ground acceleration were correlated through the relationship proposed by Guagenti and
Petrini (1989) (see Figure 2.3).
The main advantage of ‘indirect’ vulnerability index methods is that they allow the
vulnerability characteristics of the building stock under consideration to be determined,
rather than base the vulnerability definition on the typology alone. Nevertheless, the
methodology still requires expert judgement to be applied in assessing the buildings, and
the coefficients and weights applied in the calculation of the index have a degree of
uncertainty that is not generally accounted for. Furthermore, in order for the vulnerability
assessment of buildings on a large (e.g., national) scale to be carried out using
vulnerability indices, a large number of buildings which are assumed to represent the
national building stock need to be assessed and combined with the census data (e.g.,
Bernardini, 2000); in a country where such data is not already available, the calculation of
the vulnerability index for a large building stock would be very time consuming.
However, in any risk or loss assessment model a detailed collection of input data is
required for application at the national scale.
Continuous vulnerability functions based directly on the damage of buildings from past
earthquakes were introduced slightly later than DPMs; one obstacle to their derivation
being the fact that macroseismic intensity is not a continuous variable. This problem was
overcome by Spence et al. (1992) through the use of their Parameterless Scale of Intensity
(PSI) to derive vulnerability functions based on the observed damage of buildings using
the MSK damage scale (Figure 2.4). Orsini (1999) also used the PSI ground-motion
parameter to derive vulnerability curves for apartment units in Italy. Both studies
subsequently converted the PSI to PGA using empirical correlation functions, such that
the input and the response were not defined using the same parameter.
16
occurred and each structural class. Empirical fragility curves with a binomial distribution
were derived as a function of PGA, Arias Intensity and effective peak acceleration. Rota
et al. (2006) have also used data obtained from post-earthquake damage surveys carried
out in various municipalities over the past 30 years in Italy in order to derive typological
fragility curves for typical building classes (e.g., seismically designed reinforced concrete
buildings of 1-3 storeys). Observational damage probability matrices were first produced
and then processed to obtain lognormal fragility curves relating the probability of
reaching or exceeding a given damage state to the mean peak ground acceleration at the
coordinate of the municipality where the damaged buildings were located. The PGA has
been derived using the magnitude of the event and the distance to the site based on the
attenuation relation by Sabetta and Pugliese (1987), assuming rock site conditions.
Alternative empirical vulnerability functions have also been proposed, generally with
normal or lognormal distributions, which do not use macroseismic intensity or PGA to
characterise the ground motion but are related to the spectral acceleration or spectral
displacement at the fundamental elastic period of vibration (e.g., Rossetto and Elnashai,
2003; Scawthorn et al., 1981; Shinozuka et al., 1997). The latter has been an important
development as it has meant that the relationship between the frequency content of the
ground motion and the fundamental period of vibration of the building stock is taken into
consideration; in general this has been found to produce vulnerability curves which show
improved correlation between the ground motion input and damage (see Figure 2.5).
Figure 2.4: Vulnerability curves produced by Spence et al. (1992) for bare moment-
resisting frames using the parameterless scale of intensity (PSI); D1 to D5
relate to damage states in the MSK scale
17
Figure 2.5: Example of the difference in the vulnerability point distribution using
the observations of low and mid-rise building damages after the 1995 Aegion
(Greece) earthquake for different ground motion parameters: (a) PGA, and
(b) Spectral displacement at the elastic fundamental period (Rossetto and
Elnashai, 2003)
The introduction of vulnerability curves based on spectral ordinates, rather than PGA or
macroseismic intensities, has also certainly been facilitated by the emergence of more and
more attenuation equations in terms of spectral ordinates.
Is=EoSDT (2.2)
where Eo is for the basic structural performance, SD is the sub-index concerning the
structural design of the building, and T is the sub-index for the time-dependent
deterioration of the building. The method to calculate Eo involves the calculation and
multiplication of an ultimate strength index C and a ductility index F, considering the
failure mode, the total number of storeys and the position of the storey under
examination. The influence of irregularity, stiffness and/or mass concentration of a
structure on the seismic performance should be accounted for by the sub-index SD. The
18
influence of deterioration and cracking is taken into account by the sub-index T , which is
based on the data found through a field investigation. Once the seismic performance
index I s has been calculated it should be compared with the seismic judgement index for
the structure I so to determine whether the building can be called “safe” against an
assumed earthquake ground motion (i.e., if I s> I so ).There are three possibilities
depending on the difference between I s and I so:
• I s> I so; corresponds to a low vulnerability condition for all three screening levels,
• I s< I so; is an uncertain condition when I s> I so is only slightly smaller than I so ( I s
falls within an interval between I so and a lower limit that is based on the observed
damage of reinforced concrete buildings) and thus requires a more detailed
assessment using the next screening level or nonlinear dynamic analysis.
I so =EsZGU (2.3)
where Es was originally taken as 0.6 for the second and third level screening and 0.8 for
the first (hence providing an increase in the demand due to a lack of reliability of the
method); Z is a zone index used to modify the intensity of the ground motion assumed at
the site of the building; G is the ground index used to account for ground-building
interaction, amplification in the top layer of ground, or topographical effects; and U is a
usage index which may be seen as a kind of importance factor concerned with the
function of the building. This seismic judgement index suggests a storey shear force
which is independent of the ground motion, but the force obviously depends on the
response of the building and so in the 1998 revised version of the Japanese Building
Standard Law the index I so is taken as the spectral acceleration (in terms of g) at the
period of response of the structure in question. This index should be distributed up the
height of the structure, and a triangular distribution is suggested.
19
walls and infill panels divided by total floor area) and a column index (area of columns
divided by total floor area). Damage data can be used to calibrate the “Priority Index” and
define the level of vulnerability of the building. Yakut (2004) has recently proposed a
“Capacity Index”, which considers the orientation, size and material properties of the
lateral load-resisting structural system as well as the quality of workmanship and
materials and architectural features such as short columns and plan irregularities.
Buildings are classified as either at low risk (and thus should not suffer severe damage) or
at high risk (and thus would probably not meet the life-safety performance level of recent
seismic codes). These rapid vulnerability assessment methods are of particular use for
prioritizing buildings for further, more detailed assessments that would be required to
design a rehabilitation scheme.
The Seismic Safety Screening Method (SSSM) has recently been proposed by Ozdemir et
al. (2005), which is an adaptation of the Japanese Seismic Index Method (JBDPA, 1990).
This rapid seismic safety evaluation method can be applied to structures with 6 storeys or
less and with reinforced concrete frame, shear wall or dual frame-shear wall structural
systems. As with the Japanese method, the principle of the method is that the seismic
capacity of a building is represented by a seismic index value, I s, which is a function of
its strength and ductility. If the building has a seismic index value that is lower than a
seismic demand index value (I D, dependent on the seismicity of the region of interest)
then detailed structural analysis would be necessary for further assessment of the
vulnerability; otherwise the building can be judged as ‘safe’. This method has been
calibrated to Turkish buildings using nonlinear static analyses of 12 buildings from
Zeytinburnu, Istanbul (Ozdemir et al., 2005).
The use of rapid screening methods has an important role to play in the definition of
prioritisation of buildings for seismic retrofit, but the use of such methods in large-scale
seismic risk models is limited due to the need to consider buildings individually in a
deterministic fashion, and thus this would not be economically feasible.
The emergence of more attenuation equations in terms of spectral ordinates and the
corresponding derivation of seismic hazard maps in terms of spectral ordinates, as
opposed to macroseismic intensity or PGA, has not only catered to the aforementioned
improvement of empirical methods, but also given rise to the development of analytical
20
methods. These methods tend to feature slightly more detailed and transparent
vulnerability assessment algorithms with direct physical meaning, that not only allow
detailed sensitivity studies to be undertaken, but also cater to straightforward calibration
to various characteristics of building stock and hazard. The latter is a definite
disadvantage of empirical methods. Such characteristics place the analytical type of loss
assessment approaches, as described below, in an ideal position for employment in
parametric studies that aim at the definition/calibration of urban planning, retrofitting,
insurance and other similar policies or initiatives.
Definition of random
characterization of Definition of
structural parameters damage states
Selection of
representative set
of earthquake
Selection of methodology Definition of criteria
for nonlinear analysis for identification of
damage states
Nonlinear analysis
Definition of probabilistic
distribution of damage
Although vulnerability curves and damage probability matrices have traditionally been
derived using observed damage data, recent proposals have made use of computational
analyses to overcome some of the drawbacks of the methods highlighted in the previous
21
section. Figure 2.6 summarises the basic components that are required to analytically
derive vulnerability curves or damage probability matrices.
Singhal and Kiremidjian (1996) developed fragility (or vulnerability) curves and damage
probability matrices for three categories of reinforced concrete frame structures using
Monte Carlo simulation. The probabilities of structural damage were determined using
nonlinear dynamic analysis with an ensemble of ground motions. For the DPMs,
Modified Mercalli Intensity was used as the ground-motion parameter, whilst spectral
acceleration was used for the generation of fragility functions. The major components of
the methodology consist of:
22
tagging survey of 84 buildings damaged by the 1994 Northridge earthquake, while using
a weighting system (Bayesian updating technique) to take into account the reliability of
different data sources (Singhal and Kiremidjian, 1998).
Table 2.3: Damage probability matrix for sample low-rise buildings (Singhal and
Kiremidjian, 1996)
23
Dumova-Jovanoska (2004) produced vulnerability curves/damage probability matrices
for reinforced concrete buildings built in the Skopje region. These earthquake damage-
intensity relationships were derived by analytically modelling the representative RC
buildings and by running the dynamic nonlinear analysis with a set of 240 synthetic
earthquake records. The damage to the structures was measured using the damage index
by Park and Ang (1985) and corresponding individual discrete damage states. A normal
probabilistic distribution was assumed for the probability of occurrence of damage.
Kappos et al. (1995, 1998) have derived damage probability matrices using a hybrid
procedure, in which parts of the DPMs for each intensity level were constructed using the
available data from past earthquakes following the “Vulnerability Index Procedure” as
discussed above. The remaining parts of the DPMs were constructed using the results of
nonlinear dynamic analysis of models that simulated the behaviour of each building class.
The time-history records were scaled to PGA values estimated by the seismic hazard
analysis; intensity and PGA were correlated using empirical relationships. A global
damage index was derived to correlate the structural response from the dynamic analysis
(ductility factors, displacements, etc.) with loss, expressed in terms of the cost of repair.
24
A total of 120 analyses of typical Greek buildings designed for the 1959 code were run
(for 6 structures, 10 ground motions and 2 intensities), and the statistical damage results
were combined with the observed damage from the 1978 earthquake in Thessaloniki.
Figure 2.8: Simulated (thick line) and observed (thin line) vulnerability functions for
MSK intensity VII (Barbat et al., 1996)
Barbat et al. (1996) used the Italian “Vulnerability Index Methodology” for a hybrid
vulnerability assessment of Spanish urban areas. A post-earthquake study was initially
performed for two earthquakes with a maximum intensity of VII on the MSK scale. The
structural and non-structural damage to masonry structures was analysed and correlated to
the vulnerability and damage indices used in the Italian methodology. Statistical analyses
were then performed to obtain the vulnerability function for the MSK intensity level VII.
A computer simulation process was subsequently used to obtain the vulnerability
functions at other intensity levels. Sixty hypothetical buildings with characteristics
obtained from the building stock in the area were generated using Monte Carlo simulation
(considering a uniform probability density function for the capacity parameters) in order
to simulate the behaviour of a complete urban zone, and a simplified analytical procedure
proposed by Abrams (1992) was chosen to model the capacity/demand relationship of
masonry structures. The vulnerability index was calculated for each building, this was
plotted against the global damage index (based on the capacity/demand ratios of the
structural members) for the MSK intensity level VII, and a curve was obtained by
regression analysis; this curve was then re-calibrated so as to match the field observations
(Figure 2.8). The difference between the curves in Figure 2.8 was assumed to be due to
the use of the proposed weighting factors for Italian buildings and so the weighting
factors were modified such that the observed and calculated vulnerability functions
25
matched. Once the calibration with the 60 random buildings had been carried out, the
vulnerability functions for the other intensity levels were then produced using 2000
hypothetical buildings in conjunction with the calibrated weighting factors.
The main difficulty in the use of hybrid methods is probably related to the calibration of
the analytical results, considering that the two vulnerability curves include different
sources of uncertainty and are thus not directly comparable. In the analytical curves the
sources of uncertainty are clearly defined during the generation of the curves whilst the
specific sources and levels of variability in the empirical data are not quantifiable. The
method used to calibrate the analytical vulnerability curves using empirical data should
depend on whether the aim is to include the additional uncertainty present in the empirical
data which is not accounted for in the analytical data or whether the aim is in fact to
improve the analytical model used to define the capacity of the building stock. If the
principle goal is the latter, then perhaps it would be preferable to calibrate the
vulnerability curves considering the median (50 percentile) values, and an adaptation of
the analytical model could be made such that the two median values coincide. In this way,
the observational data is used to calibrate the analytical model, but analytically-derived
vulnerability curves with their known and specified sources of uncertainty are used in the
loss model. This can be of use when only the uncertainty related to the capacity of the
building stock is required and not that related to the ground motion; for example, such
curves are needed when a probabilistic seismic hazard assessment is used to define the
ground motion, as this will ensure that the latter variability is not doublecounted (e.g.,
Bommer and Crowley, 2006).
Many recent proposals for analytical vulnerability assessment methods use collapse
multipliers calculated from mechanical concepts to ascertain whether a mechanism will
form and thus damage will occur; these procedures have been particularly applied to
masonry buildings.
VULNUS is a method that has been proposed for the vulnerability assessment of
unreinforced masonry buildings (URMB) using the fuzzy-set theory and the definition of
collapse multipliers (Bernardini et al., 1990). A collapse multiplier for in-plane behaviour
(I1) with shear failure at the ground floor is defined as the ratio of the in-plane shear
strength of the system of walls to the total weight:
26
min(V xV y )
I1 = (2.4)
W
where W is the total weight of the building, and Vx and Vy represent the strength at
midstorey height of the ground floor in the x - and y -directions. The latter can be
calculated considering the tensile strength of the masonry and the total area of the walls.
The collapse multiplier for the out-of-plane behaviour (I2) is found from the ratio of the
out-of-plane flexural strength of the most critical external wall to the total weight,
evaluated by summing the resistance of the vertical ( I 2' ) and horizontal strips ( I 2" )
Following the calculation of the two collapse multipliers, a third index I 3 is calculated,
which is the weighted sum of the scores of seven partial vulnerability factors:
Wi S i
I 3i = ∑ (2.6)
i 45 x3.15
where the score Si ranges from 0 (good) to 45 (poor), and the weight, Wi , is assigned
depending on the relative importance of the factors on the vulnerability of the building.
The mean absolute acceleration response A of the building (i.e., maximum base shear
divided by total weight) is calculated as well as an uncertainty factor a, which is obtained
from the fuzzy-set theory using I 3 . The probability of exceeding the D4 limit state (based
on the EMS98 damage scale) for a group of buildings is a function of the aforementioned
parameters: Vg = ƒ(I1,I2,A,a)
27
.
Figure 2.9: Out-of-plane failure mechanisms (after D’Ayala and Speranza, 2002)
The ultimate roof displacement ∆ui . is determined as a function of the ultimate rotation θu
of the structural elements:
∆ u ,1 = θ u ( H n − H k ) (2.7)
∆ u,2 = θ u H k (2.8)
∆ u ,1 = θ u ( H k − H k −1 ) (2.9)
28
The global seismic behaviour is then represented by the base shear coefficient Cb,i (i.e.,
base shear Vb,i divided by seismic weight W ) and the corresponding lateral (driftu )I (i.e.,
roof displacement divided by building height):
Vb ,i ∆ u ,i
C b ,i = ; (drift u ) i = (2.10)
W Hn
Seismic response is evaluated for 3nz mechanisms, where nz represents the number of
storeys; the seismic capacity corresponds to the lowest Cb,i of all the considered
mechanisms and the associated drift. The seismic capacity of building stock will vary
from the generic building due to different morphologic/geometric/structural
configurations and material properties. Hence, a number of building models are generated
based on the probabilistic distribution of the structural parameters. A Monte Carlo
simulation technique is applied to calculate “probabilistic capacity curves” which express
the probability of having a capacity lower than the assigned value: this can be likened to
the percentage of buildings with a value of capacity lower than a given threshold value.
A clear framework for the treatment of uncertainty in the geometric and material
properties in an urban environment is accounted for in the method by Cosenza et al.
(2005), whilst this is not the case for the VULNUS and FaMIVE procedures. However,
the main disadvantage of the procedure by Cosenza et al. (2005), and also FaMIVE, is
that no clear indication is given on how the capacity is to be convolved with the demand
to calculate the probability of exceeding given limit states. The VULNUS procedure does
allow the estimation of the probability of damage, but only for one limit state, i.e., the
collapse limit state. Thus, the use of these collapse-mechanism methods within a loss
assessment model appears to be somewhat limited at present.
HAZUS (Hazard US) is the outcome of a project conducted for the National Institute of
Building Science (NIBS), under a cooperative agreement with the Federal Emergency
Management Agency (FEMA), to develop a nationally applicable methodology for
estimating the potential losses from earthquakes on a regional basis (Whitman et al.,
1997; FEMA, 1999, 2003). The main modules of the methodology are as follows:
• The estimation of the Potential Earth Science Hazards (PESH); these include
ground motion, ground failure (liquefaction, landslides and surface fault rupture)
and tsunami/seiche.
29
• The inventory is divided into buildings, facilities, transportation systems, utility
systems, and hazardous material facilities. The inventory is then further
partitioned into pre-defined building classes with similar damage/loss
characteristics.
• Direct physical damage is calculated for each building class to find the probability
of none, slight, moderate, extensive and complete structural and non-structural
damage due to each type of PESH.
• Induced physical damage from inundation, fire, hazardous materials and debris is
estimated.
• Indirect economic losses due to downtime, that causes a chain reaction through
the regional economy, are estimated.
The vulnerability assessment component of the procedure is contained within the direct
physical damage module and is based on the Capacity Spectrum Method of ATC-40
30
(ATC, 1996). In this method, the performance point of a building type under a particular
ground-shaking scenario (or PESH) is found from the intersection of an acceleration-
displacement spectrum, representing the ground motion, and a capacity spectrum
(pushover curve), representing the horizontal displacement of the structure under
increasing lateral load (Kircher et al., 1997), as shown in Figure 2.11. The demand
spectrum is reduced for both damping and duration effects. The reduction of the spectrum
is applied to account for the hysteretic damping that occurs during inelastic behaviour of
the structure, where the damping is based on the area enclosed by the hysteretic loop at
peak response displacement and acceleration. A reduction factor is applied to the
hysteretic damping as a function of shaking duration to simulate degradation (e.g.,
pinching) of the hysteresis loop during cyclic response.
The capacity spectrum has been developed for each building class using model buildings
designed for different levels of design practice in the US. The performance point obtained
from this average building provides the displacement input into the limit state
vulnerability curves to give the probability of being in a given damage band. The
vulnerability curves are lognormal curves with a total logarithmic standard deviation, βSds,
which combines the uncertainty in the damage state threshold, βM(Sds), the variability in
the capacity (response) properties of the model building type, βC , and the uncertainty in
the response due to the spatial variability of the ground-motion demand, βD. The latter
two components of the variability are inter-dependent (as the damped demand spectrum is
dependent on the building capacity due to the nonlinearity of the capacity response and
thus energy dissipation) and thus need to be combined using a complex process of
convolution (CONV) of the probability distributions, before being combined with βM(Sds)
using the square-root-of-the-sum-of-the-squares (SRSS):
2
⎛ ⎡ ⎤⎞
β Sds = ⎜ CONV ⎢ β c β D S d , Sds ⎥ ⎟ + (β M ( Sds ) ) (2.11)
⎜ ⎟
⎝ ⎣ ⎦⎠
where , S d , Sds is the median spectral displacement for the damage state ds . Median
spectral displacement (or acceleration) values and the total variability are developed for
each of the model building types and damage states of interest through a combination of
performance data (from tests of building elements), earthquake field data, expert opinion
and judgment.
31
OPTION 1, OPTION 2, OPTION4
Macroseismic
Soil Column Units Column Soil Intensities
response
Seismic Action
OPTION4
at Surface
Vulnerability and
Fragility Liquefaction
Housing
Characterization
Inventory
# of Building
for Different
Damages States
# of
Human Inventory Human Loses Casualities
and Injuries
Economic Building
Direct Economic
Parameters Floor Area
Loss
(m2)
Figure 2.12: Modules involved in the LNECloss tool (Sousa et al., 2004)
Once the proportion of buildings in each damage band (none, slight, moderate, extensive,
and complete) has been obtained, a composite measure of damage, which aims to provide
an idea of the expected overall rebuilding cost, is calculated by summing the proportions
32
which are weighted by 0% for no damage, 2% for slight damage, 10% for moderate
damage, 50% for extensive damage, and 100% for complete damage.
The HAZUS methodology was originally derived as a tool for estimating the impact of
individual earthquake scenarios; it has been subsequently adapted (FEMA, 2001) to a full
loss assessment methodology wherein the hazard from all possible earthquakes, derived
from probabilistic seismic hazard assessment (PSHA), is considered. In this case, the
variability in the demand is already modelled in the PSHA and so needs to be removed
from the vulnerability curves to avoid double counting; the curves thus become lognormal
functions with a logarithmic standard deviation given by
A potential weakness of the method is that the capacity curves and vulnerability functions
published in the HAZUS manual have been derived for buildings in the US having a
limited range of storey heights; thus the application of this method to other parts of the
world requires additional research to be carried out. Hence, capacity curves and
vulnerability functions would need to be derived for the building stock under
consideration; however, there is difficulty involved in obtaining a physically realistic
representation of the inelastic response of the structure using pushover analysis. Although
this aspect can be somewhat improved using the displacement-based adaptive pushover
techniques (Antoniou and Pinho, 2004), a faithful representation of the real structural
behaviour requires a great deal of information about the structure, including
reinforcement details, which are unlikely to be well known for a large building stock.
HAZUS has been adopted all over the world for the loss assessment of urban areas. The
methodology in itself has not been adapted in any way, but the capacity curves and
fragility functions have been calibrated to the building stock under consideration.
The LNECloss tool is an automatic seismic scenario loss estimation methodology that
makes use of the capacity spectrum method and has been implemented in a GIS
environment (Sousa et al., 2004). However, an equivalent approach to the capacity
33
spectrum method to evaluate the performance point has been implemented in the
LNECloss code. This alternative equivalent scheme to compute the performance point
starts with the definition of an input motion in terms of a power spectral density function
(PSDF) and an equivalent stationary duration, T. The seismic input needs to be defined in
terms of response spectra, associated to some damping coefficient, and thus a linear
stochastic approach has been implemented in LNECloss to accomplish this
transformation. This methodology illustrated in Figure 2.12 is being developed and
implemented within the LESSLOSS Integrated Project (Calvi and Pinho, 2004) under the
earthquake disaster scenario predictions and loss modelling for urban areas sub-project
and is being applied to the city of Lisbon (see http://www.lessloss.org/main/).
Giovinazzi (2005) presented a mechanical procedure for the risk assessment of both
masonry and reinforced concrete frames. This uses simplified bilinear capacity spectra
(capacity or ‘pushover’ curves converted to plots of spectral acceleration versus spectral
displacement), which are derived using the equations and parameters available in seismic
design codes. The base shear coefficient, which can be related to the yield spectral
acceleration, is generally a function of seismic zone, soil conditions, building dynamic
response, structural type, and building importance, all of which can be obtained from the
codes. The yield spectral displacement is a function of the yield spectral acceleration and
the yield period of vibration; the latter being calculated using the simple formulae
available in seismic design codes which usually give the period as a function of the
building height. The ultimate spectral displacement is a function of the yield displacement
and the building ductility capacity, which has been assumed to be 2.5 for the reinforced
concrete buildings that have not been seismically designed, or which can be derived as a
function of the behaviour factor specified in the codes. The use of seismic design codes to
calculate simplified pushover curves of the building stock is a cost-effective method for
calibrating the vulnerability of buildings designed for different design codes. However,
code-based definitions of the yield period of vibration and base shear capacity of RC
structures could differ greatly from the actual properties of the building stock (e.g.,
Crowley, 2003) and in many countries the existing building stock has been constructed
without even conforming to any design code.
In the approach of Giovinazzi (2005), once capacity spectra for the building classes have
been derived, the next step is to use the Capacity Spectrum Method, which has been
developed and employed in HAZUS (FEMA, 1999, 2003). As has been described in the
34
review of HAZUS, the performance point of the building class is obtained by comparing
the capacity spectrum with the inelastic Acceleration- Displacement Response Spectrum
(ADRS), which is produced by using codified spectral shapes anchored to the PGA
obtained from the hazard analysis. Once the performance point is obtained it is inserted
into the vulnerability curves in order to obtain the probability of exceeding pre-defined
damage states. The method proposed by Giovinazzi for this stage of the risk assessment
differs from that proposed in HAZUS, as discussed in what follows.
The mean values for the displacement thresholds ( S d ,i ) employed by Giovinazzi (2005)
S d ,1 = 0.7d y (2.13)
S d , 2 = 1.5d y (2.14)
S d ,3 = 0.5(d y + d u ) (2.15)
S d ,1 = d u (2.16)
These thresholds have been based on expert judgement and have been verified on the
basis of the results of pushover analyses performed on prototype buildings, though details
of these analyses have not been reported. In order to model the variability in the
estimation of the damage suffered by a building (which is due to the uncertainty in the
capacity curve, the damage state definition and the seismic demand spectra), rather than
combining the logarithmic standard deviations assigned to each of these uncertainties (as
carried out in HAZUS), the proposed procedure is “to evaluate the overall uncertainty in
the damage estimation in such a way that it can represent the same dispersion of observed
damage data that….are well fitted by a binomial distribution”. Giovinazzi (2005) argued
that the binomial distribution had been shown to fit damage data by Braga et al. (1982)
and Spence et al. (2003); however, Giovinazzi admitted that there were problems with the
use of this distribution as it does not allow a variation of the scatter around the mean
damage band (i.e., it is fixed) and Spence et al. (2003) have shown that frame structures
may follow a very different, more complex distribution (see Figure 2.13 where observed
damage data from the Kocaeli earthquake of 1999 is presented and compared with the
results from various risk assessment procedures for reinforced concrete frames).
Indeed, the authors believe that forcing the damage distribution to follow a binomial
distribution could lead to erroneous damage predictions in many cases where non-ductile
35
reinforced concrete frames are considered in a loss model. The observed damage
distribution, as presented in Figure 2.13, can be easily understood by considering the
behaviour of such frames: once the first/yield limit state has been exceeded, the
inadequate confinement in these buildings will ensure that they will very rapidly fail the
higher limit states also due to the low levels of attainable limit state strains (and
consequently low levels of ductility). For this reason, the vulnerability curves for the
different damage states for these frames should be steep and close together. It is therefore
expected that the damage distribution will show larger proportions at the extreme damage
bands (i.e., slight and complete damage) with lower proportions being predicted for the
intermediate damage states for moderate ground-motion inputs. For well-designed
buildings, however, there should be a greater separation between the curves, representing
the greater postyield capacity due to the higher limit state strains of adequately confined
members, and thus a binomial/normal/lognormal type of damage distribution could be
expected.
36
Figure 2.13: 1999 Kocaeli earthquake, Gölcük Coastal Zone: Results of various
analyses and comparison with the damage data for the mid-rise reinforced
concrete frames (Spence et al., 2003)
Calvi (1999) considered the inherent variability in the structural properties within an
urban environment by assigning maxima and minima to the variables and assuming a
uniform probability distribution function. The period of vibration was calculated using the
empirical formula in EC8 (CEN, 2003), which directly relates the height of a building to
its period; again maxima and minima were applied to the parameters in this equation. The
range of the limit state displacement capacity for the building class can be plotted against
the range of the period of vibration, as presented in Figure 2.17. This capacity area can
then be directly compared with the displacement response spectrum; the area below the
spectrum represents the proportion of buildings failing, or exceeding, the limit state. The
displacement response spectra were adjusted to include the nonlinear response, wherein a
reduction of the spectral ordinates was applied to account for the energy dissipation
37
capacity of the structure as a function of the target displacement and the structural
response.
Figure 2.16: Deformed shapes for different limit states and in-plane failure modes
The methodology proposed by Calvi (1999) has subsequently been developed for
reinforced concrete buildings by Pinho et al. (2002) and Crowley et al. (2004, 2006),
leading to the Displacement-Based Earthquake Loss Assessment (DBELA) procedure.
For the masonry component of the methodology, a parallel extensive development was
carried out by Restrepo-Vélez and Magenes (2004), Restrepo-Vélez (2005), Modena et al.
(2005) leading to the procedure MeBaSe (Mechanical Based Procedure for the Seismic
Risk Estimation of Unreinforced Masonry Buildings). A brief description of the main
components of both MeBaSe and DBELA is provided in what follows.
Three limit states are considered in MeBaSe for the case of in-plane failure mechanisms,
namely LS1- LS2 for which just slight structural and non-structural damages occur, LS3
for which moderate structural damage and extensive non-structural damage occur, and
38
LS4 for which the collapse of the building is considered (Figure 2.16). Equations (2.17)
and (2.18) are used to compute the limit state functions for the case of in-plane failure
mechanisms, where the vector of random variables can be formed by the total height of
the building class, hT , the height of the openings at the failing storey, hsp , the drift limit
at yield, δy , the drift limit at the specified limit state, δLS , the resistant area of the walls in
the direction of minimum strength, Am , the ratio of areas of walls in both directions, γm ,
the referential shear strength of the masonry, km τkm , and the correction factor, Øc. This
correction factor allows to express the tri-dimensional response of a building by means of
a simplified bi-dimensional model, based solely on the shear strength of the walls, and it
is computed with Equation (2.19), where LW is the length of the piers and LT is the total
length of the perimeter walls.
Figure 2.17: An example of the intersection of capacity areas and demand spectrum
(Calvi, 1999)
k1 hT δ y + k 2 (δ LS − δ y )hsp
TLS = 2π (2.17)
1 ⎛ K2 ⎞
Am K 1τ km ⎜⎜1 + ⎟⎟
φc ⎝ Am (1 + γ m )τ km ⎠
1
T LS2 1 ⎛ K2 ⎞ 2
∆ LS = Am K 1τ km ⎜⎜1 + ⎟⎟ g (2.18)
4π φ c ⎝ Am (1 + γ m )τ km
2
⎠
39
τ km
φ c = 5.53 + 0.46 (2.19)
⎛ LW ⎞
⎜ L ⎟
⎝ T ⎠
Out-of-plane mechanisms, which were not considered by Calvi (1999), have been
included in MeBaSe by considering one-way and two-way bending mechanisms, whose
dynamic response has been described by Doherty et al. (2002) and Griffith et al. (2003)
by means of a tri-linear model.
The main improvements that have been applied to the method by Calvi (1999) for
reinforced concrete include theoretically improved structural and non-structural
displacement capacity equations for ground shaking, the derivation of an equation
between yield period and height for European buildings both with and without infill
panels (Crowley and Pinho, 2004, 2006), and the consideration of the vulnerability to
liquefaction-induced ground deformations (Bird et al., 2005, 2006). DBELA considers
three limit states based on the sectional strains in the steel and concrete, for the case of
structural limit states, and on interstorey drifts, in the case of non-structural limit states
(Crowley et al., 2004). The structural displacement capacity of reinforced concrete
members/structures is derived from the structural mechanics principles; beginning with
the yield strain of the reinforcing steel and the geometry of the beam and column sections
in the building class, the yield curvature is defined using the relationships suggested by
Priestley (2003). These beam and column yield curvatures are then multiplied by
empirical coefficients to account for shear and joint deformations to obtain a formula for
the chord rotation. This chord rotation is equated to base rotation and multiplied by the
height of the equivalent SDOF system to produce the yield displacement capacity. Post-
yield displacement capacity formulae are derived by adding a plastic displacement
40
component to the yield displacement, which is calculated by multiplying the limit state
plastic section curvature, the plastic hinge length, and the height or length of the yielding
member.
Figure 2.18: Pushover results for a four-storey building with the simplified and
three-dimensional methodologies, and for different referential shear strength
values
lb
∆ sy = 5e fh T y ε y (2.20)
hb
41
∆ NS1 = 5ϑ (10T y ) (2.21)
where e fh is the effective height coefficient, Ty is the yield period of vibration, ε y is the
yield strain of the reinforcing steel, hb is the length of the beam, b h is the depth of the
beam, S is the shape coefficient relating the roof displacement to the displacement at the
centre of the seismic force, and ϑ is the interstorey drift capacity of the non-structural
elements.
Figure 2.19: An illustration of the deterministic comparison of the limit state (LS)
displacement capacity and displacement demand in DBELA
One of the principal modifications to the methodology by Calvi (1999) that is common to
both DBELA and MeBaSe is the consideration of a joint probability density function
42
(JPDF) of displacement capacity and period (Figure 2.20(c)). This is calculated from the
variability of a vector of random variables present in the equations for displacement
capacity, which is conditioned on period (Figure 2.20(a)) and period of vibration (Figure
2.20(b)). Furthermore, the variability in the demand response spectrum may also be taken
into consideration in the calculation of the probability of exceeding a given limit state.
Notwithstanding its work-in-progress status, the DBELA and MeBaSe methods have
already been applied in the assessment of the seismic vulnerability of entire regions (e.g.,
Restrepo-Vélez and Magenes, 2004; Iaccino, 2004), in the carrying out of important
sensitivity studies for definition of priorities in terms of data collection needs (e.g.,
Crowley et al., 2005), and for the evaluation of different approaches to the definition of
the seismic hazard (e.g., Bommer and Crowley, 2006; Crowley and Bommer, 2006).
Finally, DBELA has also been proposed as a possible tool for the calibration of future
seismic design codes (e.g., Bommer et al., 2005). Future developments which are to be
considered include the following: the effect of duration of the demand on the response of
structures; the influence of infill panels on the displacement capacity of reinforced
concrete frames; the inclusion of model uncertainty due to the simplifying assumptions
that have been made in the modeling of RC buildings response; the prediction of
acceleration-sensitive non-structural damage; and improvements in the displacement
spectra predictions through the consideration of long-period ordinates, near-fault effects,
etc. Following the implementation of the aforementioned developments, it will then be
necessary to validate and perhaps even calibrate the methodologies using observed
damage data from recent earthquakes.
Perhaps one of the main reasons for which the analytical/mechanical methodologies have
only recently been proposed for use in risk and loss assessment is due to developments
which have occurred in the field of seismic hazard assessment, such as the derivation of
seismic hazard maps in terms of spectral ordinates as opposed to macroseismic intensity
or PGA. However, the benefits which the use of more advanced seismic hazard and
vulnerability models offer need to be weighed against the increased amount of
information, and thus time and money, which is required to construct an earthquake loss
model based on such analyses. For example, apart from the compilation of databases of
building stock inventory (which are required for any loss model), information on the
structural characteristics of the building stock is also required in order to calibrate the
43
analytical model, and databases of the earthquake activity, attenuation equations, and
ground conditions are needed to carry out a seismic hazard assessment in terms of
spectral ordinates. Nevertheless, the use of such detailed and transparent loss models
allows detailed sensitivity studies to be undertaken which provide valuable insight into
how much the loss results depend on the models, data, uncertainties and assumptions
employed.
There are still some outstanding issues with some analytical/mechanical methods which
need to be addressed, such as the capability of numerical models to accurately predict the
response of real structures; the accuracy in transforming numerical indices of damage into
actual damage of real structures; the capability of accounting for human errors in the
design and construction of buildings, which are often the main causes of catastrophic
collapses; and the need to extend the results obtained for few reference models to a large
class of structures, when accurate analyses are carried out.
2.4 Ground Motion Parameters for Seismic Hazard Analysis (Kramer 2003)
Ground motion parameters are essential for describing the important characteristics of
strong ground motion in compact, quantitative form. Many parameters have been
proposed to characterize the amplitude, frequency content, and duration of strong ground
motions; some describe only one of these characteristics, while others may reflect two or
three.
Amplitude Parameters
The most common way of describing a ground motion is with a time history. The motion
parameter may be acceleration, velocity, or displacement, or all three may be displayed.
Typically, only one of these quantities is measured directly with the others computed
from it by integration and/or differentiation.
Peak Acceleration
The most commonly used measure of the amplitude of a particular ground motion is the
peak horizontal acceleration (PHA). The PHA for a given component of motion is simply
the largest (absolute) value of horizontal acceleration obtained from the accelerometer of
that component. By taking the vector sum of two orthogonal components, the maximum
resultant PHA (the direction of which will usually not coincide with either of the
measured components) can be obtained. Horizontal accelerations have commonly been
used to describe ground motions because of their natural relationship to inertial forces;
44
indeed, the largest dynamic forces induced in certain types of structures (i.e., very stiff
structures) are closely related to the PHA. The PHA can also be correlated to earthquake
intensity (e.g., Trifunac and Brady, 1975a; Murphy and O'Brien, 1977; Krinitzsky and
Chang, 1987). Although this correlation is far from precise, it can be very useful for
estimation of PHA when only intensity information is available, as in the cases of
earthquakes that occurred before strong motion instruments were available
(preinstrumental earthquakes). A number of intensity-acceleration relationships have been
proposed. The use of intensity-attenuation relationships also allows estimation of the
spatial variability of peak acceleration from the isoseismic maps of historical earthquakes.
Vertical accelerations have received less attention in earthquake engineering than hor-
izontal accelerations, primarily because the margins of safety against gravity-induced
static vertical forces in constructed works usually provide adequate resistance to dynamic
forces induced by vertical accelerations during earthquakes. For engineering purposes, the
peak vertical acceleration (PVA) is often assumed to be two-thirds of the PHA (Newmark
and Hall, 1982). The ratio of PVA to PHA, however, has more recently been observed to
be quite variable but generally to be greater than two-thirds near the source of moderate
to large earthquakes and less than two-thirds at large distances (Campbell, 1985;
Abrahamson and Litehiser, 1989). Peak vertical accelerations can be quite large; a PVA
of 1.74g was measured between the Imperial and Brawley faults in the 1979 Imperial
Valley earthquake.
Ground motions with high peak accelerations are usually, but not always, more destruc-
tive than motions with lower peak accelerations. Very high peak accelerations that last for
only a very short period of time may cause little damage to many types of structures. A
number of earthquakes have produced peak accelerations in excess of 0.5g but caused no
significant damage to structures because the peak accelerations occurred at very high
frequencies and the duration of the earthquake was not long. Although peak acceleration
is a very useful parameter, it provides no information on the frequency content or
duration of the motion; consequently, it must be supplemented by additional information
to characterize a ground motion accurately.
McGuire, 1978; Duggal, 1989 and Boore et al. 1993 acceleration attenuation expression
is used to predict PGA. Boore expression is developed for rock and McGuire expression
is developed for both rock and alluvium and Duggal expression is developed for
alluvium. Public Works Research Institute (PWRI) in Japan proposes McGuire, 1978 and
45
Duggal, 1989 acceleration attenuation expression for alluvial soil. As bed soil of
Bangladesh is alluvium McGuire and Duggal expression can be applicable. Boore
expression can be used with amplification factor for alluvium. The controlling earthquake
that is expected to produce the strongest level of shaking is described in terms of its size
(usually expressed as magnitude) and distance from the site. The probability of
occurrence of the controlling earthquake is assumed to be one at the points in each source
zone closest to the site and zero elsewhere. Peak ground acceleration (PGA) is used to
characterize the seismic hazard.
b1 b2 b3 b4 b5 b6 b7 Gb Gc σ log PGA
The-notion of effective design acceleration, with different definitions, has been proposed
by at least two researchers. Since pulses of high acceleration at high frequencies induce
little response in most structures, Benjamin and Associates (1988) proposed that an
effective design acceleration be taken as the peak acceleration that remains after filtering
out accelerations above 8 to 9 Hz. Kennedy (1980) proposed that the effective design
acceleration be 25% greater than the third highest (absolute) peak acceleration obtained
from a filtered time history.
46
Response Spectra
Elastic response spectra assume linear structural force-displacement behavior. For many
real structures, however, inelastic behavior may be induced by earthquake ground
motions. An inelastic response spectrum (i.e., one that corresponds to a nonlinear force-
displacement relationship, can be used to account for the effects of inelastic behavior.
Spectral accelerations decrease with increasing ductility, but total displacements increase.
Response spectra reflect strong ground motion characteristics indirectly, since they are
"filtered" by the response of a SDOF structure. The amplitude, frequency content, and to
a lesser extent, duration of the input motion all influence spectral values. It is important to
remember that response spectra represent only the maximum responses of a number of
different structures. However, the response of structures is of great importance in
earthquake engineering, and the response spectrum has proven to be an important and
useful tool for characterization of strong ground motion. Application of response spectra
for dynamic analysis of structure in Bangladesh is described in Art. 2.5.7 of Part 6 of
BNBC 1993.
A limit-state fragility curve provides the conditional probability that the specified limit
state will be reached or exceeded as a function of the severity of the future ground motion
(Bazzurro et al. 2004). The limit states in fragilities are related to structure functionality
status and may be defined as global drift ratio, interstory drift ratio, maximum roof
displacement, story shear force or loss of dynamic capacity. The ground motion
intensities in the fragility functions can be spectral quantities, peak ground motion values,
47
earthquake intensities etc. Fragility curves involve uncertainties associated with dynamic
displacements produced by different ground motion records (even though they have same
intensity value), material properties, structure geometry, structure modeling and analysis
procedure.
The five limit states identified in increasing order of damage severity are:
Onset of Damage, OD: FEMA 356 and HAZUS manual define the onset of significant
nonlinear behavior (Immediate Occupancy in FEMA 356 and Slight Damage in HAZUS)
for different types of structures. No limitations on post-earthquake access are implied by
this limit state.
Yellow tag, Y: the structure is deemed fit for restricted occupancy. The access is limited
to specialized personnel only, until detailed engineering evaluation is completed.
Red tag, R: the structure is deemed unsafe. Access is not allowed until completion of
detailed engineering evaluation, retrofit or rebuilding.
Collapse, C: the structure has collapsed or is on the verge of global instability or local
collapse.
Hence, the onset of damage limit state lies within the green tag state boundaries while the
collapse state is, of course, the most severe stage of the red tag condition.
48
framework for evaluating uncertainty, performance and reliability of an engineered
system, and accordingly must play a central role in PBE. It is distinguished from
traditional deterministic approaches to safety assurance by its focus on why and how the
system might fail and by its explicit treatment of uncertainties, both in the phenomena
and in the analytical tools used to model them. One begins the PSA process by identifying
limit states (LS), or conditions in which the system ceases to perform its intended
functions in some way. In a (narrow) structural engineering sense, such limit states for
specific structural components and systems may be either strength or deformation-related.
With the limit states identified, the limit state probability can be expressed as,
In which D is a random variable (or random vector) describing the intensity of the
demand on the system, and P[LS/D = d] is the conditional limit state probability, given
that D = d, and the summation is taken over all possible values of D. The probability
P[D=d] defines the hazard and can be obtained from probabilistic hazard analysis. The
variable d is denoted the "control" or "interface" variable. The conditional probability,
P[LS/D = d] = FR(x), is the fragility. The fragility of a component or system defines the
conditional probability of its attaining or exceeding a performance limit state, which may
range from loss of function to incipient collapse, given the occurrence of a particular
operational or environmental demand. In a system of different environment limit states
can be generalized keeping equal safety or equal probability of the demand for the same
limit state.
As noted above, fragility (or vulnerability) can be described in terms of the conditional
probability of a system reaching a prescribed limit state (LS) for a given system demand
D = d, or probability of demand P(LS/D = d). Limit states related to structural behavior
range from unserviceability to various degrees of damage including incipient collapse.
Demands can be in the form of maximum force, displacement caused by earthquake
ground motions, or more generally a prescribed intensity measure of the ground motion,
over a given period of time. Expressed in this general manner, the fragility (or
vulnerability) is a function of the system capacity against each limit state as well as the
uncertainty in the capacity. The capacity controls the central location of the Fragility
Curve (FC) and uncertainty in the capacity and uncertainty in response of the structure to
49
the demand controls the shape (or dispersion) of the FC (Figure 2.21). For a deterministic
system with no capacity uncertainty, the FC is a step function. Strictly speaking, FC is
primarily a property of the system dependent on the limit state.
1.00
Smaller uncertainty fragility curve
0.50
0.25
The uncertainty inherent in building response and capacity for different ground motion
levels due to variability in construction and to uncertainty in structural evaluation process
is used to obtain the desired fragility curves for the different structural limit states. The
procedure has six steps (Bazzurro et al. 2004):
The fragility curve (Figure 2.22) for a given structural limit state LS (LS equal to onset of
damage, green, yellow, red-, or collapse state) provides the annual probability that the
intact building will end up in the specified limit state (or worse) given the occurrence at
the site of an earthquake ground motion of intensity Sa. The fragility curve, for the
yellow-tag state, for example, is denoted as FY(Sa). Based on the common lognormal
50
assumption, the curve’s estimation for the generic structural limit state LS requires two
parameter values, a median ŠaLS value and a measure of dispersion, β. The former is the
central value of the curve that correspond to an exceedance probability of 50%, the latter
controls its slope (the larger the β value, the flatter the curve). The former parameter is
referred to as the median spectral capacity value of that limit state. The fragility curve,
FLS(Sa), for the generic structural limit state LS is determined by plotting the values of
probability p = {0.05, 0.25, 0.5, 0.75 and 0.95} versus the corresponding values, Sa:
(
S a = S aLS e xβ (2.23)
where,
β =uncertainty
Sa
= percentile spectral acceleration
(
S aLS = median spectral acceleration to attain concerning limit state
x = −0.67,0.0,0.67
for 25 percentile, 50 percentile and 75 percentile value respectively, other values of x can
be obtained from table of the Gaussian distribution function (Benjamin and Cornell,
1970)..
Figure 2.22: Fragility curves for onset of damage, green, yellow, and red tags, and
for collapse of the building (Bazzurro et al. 2004)
It should be noted that in Figure 2.22 the fragility curve for the green-tag state is equal to
unity for all values of ground motions. This is because, by definition, the fragility curve
provides the likelihood that the building will be in the given limit state or worse if a
ground motion with specified Sa were to occur at the site. Because there are no structural
51
limit states less severe than green, it follows that the building will have a green tag or
worse with certainty for any level of ground motion. Note that fragility curve for onset of
damage (OD) is the steepest of the four because the value of β is the smallest. The
opposite is true for the fragility curve corresponding to the collapse state.
The behaviour of in-plane loaded masonry walls subjected to horizontal and vertical
forces has been investigated in various test programs. Figure 2.23 shows the three
possible modes of local failure that can occur depending on the condition of biaxial stress
(Page A. W., 1996).
• Flexural failure at the toe (region B) ensues from the development of tensile
cracks at the base of the wall, the increasing shear is carried by the compressed
masonry. Final failure occurs by overturning of the wall and/or crushing of the
compressed corner. This type of failure usually occurs for walls with a large
aspect ratio, i.e. a large ratio Vulnerability of masonry buildings of height to
length (small angle of inclination of resultant). It can display large displacements
(rigid body rocking motion) which reduce for increasing vertical loads.
For small vertical loads and especially when the mortar is rather weak, the diagonal
cracks usually occur in the mortar bed and butt joints (typical stepped cracks). In this case
the separated parts of the masonry wall can slide onto each other resulting in large
deformations. For large vertical loads and especially when the strength of the bricks is
rather low the diagonal cracks can also go through the bricks. In this case the separated
parts of the wall tend to slide with little ductility downwards along the more regular
diagonal crack.
52
C
A B
Figure 2.23: Different types of failure that can occur in masonry walls
The stress distribution within a masonry wall is rather complex and, depending on the
type of masonry, the geometry of the wall, the applied forces and the support conditions,
local failure can occur at the base or at the centre of the wall. However, unless major
discontinuities are present none of these local failures will lead to a complete collapse of
the wall. After local failure the wall will progressively degrade, final failure is usually a
combination of two or three different modes of failure.
Different failure criterions can be found in the literature using half empirical formulations
of the failure mechanisms. Flexural failure is usually defined by an ultimate stress
distribution in the wall section or by simple stability considerations. Shear failure is often
related to the diagonal tension capacity by principal stresses relationship using Mohr’s
Circle. A detailed discussion of those, however, is referred to the work by Bruneau which
gives an overview of different approaches that exist, the latter being also incorporated
into the “NEHRP (Natural Earthquake Hazard Reduction Program) guidelines for the
seismic rehabilitation of buildings” (FEMA 273).
The failure criterions used in the following were derived by Ganz H. R. (1985) based on
tests carried out at the ETH Zurich (Ganz H. R., 1984) and were the basis for the new
version of the Swiss Standard for Masonry (SIA177, 1980).
According to Ganz, the failure conditions for unreinforced masonry can be described by
five mechanisms, neglecting the tensile strength of masonry:
τ xy2 − σ xσ y ≤ 0 (2.24)
53
II)Compression failure of the stones
τ xy2 + σ y (σ y + f my ) ≤ 0 (2.26)
⎛ ⎛ π Φ ⎞⎞
τ xy2 + σ x ⎜⎜ σ x + 2c tan⎜ + ⎟ ⎟⎟ ≤ 0 (2.28)
⎝ ⎝ 4 2 ⎠⎠
Figure 2.24: Failure conditions for unreinforced masonry according to (Ganz H. R.,
1984)
The four independent material parameters are: Strength of the material in the x- and y-
direction, f mx and f my (x being the direction orthogonal to the mortar bed and y the
direction parallel to the mortar bed), the angle of internal friction and cohesion in the
mortar beds c.
In the three-dimensional stress field ( σ x , σ y and τ xy ) the failure conditions are represented
by two elliptical cones, two circular cylinders and a flat plane (Figure 2.24).
For application purposes it is convenient to transform the failure conditions into the
directions of principal stresses:
54
σ x = σ 2 .(cos α )2 + σ 1 .(sin α )2 (2.29)
⎛ σ1 ⎞
τ xy = σ 2 .⎜⎜ − 1⎟⎟. sin α . cos α (2.31)
⎝σ2 ⎠
Figure 2.25: Compression strength as a function of the angle after (Ganz H. R.,
1985)
Using as parameter the ratio of the two principal stresses σ 1 σ 2 the compression strength
can be presented as a function of the angle of inclination α of σ 2 to the orthogonal to the
mortar bed (Figure 2.25). Also indicated are the regions for which the different failure
mechanisms are valid. Following points are worth noting:
• For an angle of inclination greater than the angle of internal friction, α > Φ, the
uniaxial compression strength reduces dramatically and stays at a minimum up to
α = π/2
55
Also shown in Figure 2.25 is a simple approximation of the uniaxial compression strength
neglecting the cohesion in the mortar beds c. Thus for an angle of inclination bigger than
the angle of internal friction Φ the compression strength reduces to zero.
The value of Φ, the angle of internal friction, varies little for different types of masonry
and can be assumed to lie between 0.7 < tanΦ < 0.8 (Ganz H. R., 1985)
The ratio of f my f mx depends on the type of the stone units, on the workmanship of the
butt joints and on the type of assemblage, ranging typically from 0.3 to 0.5 for brick and
limestone masonry.
Figure 2.26: Internal forces and corresponding stress field of a wall element
Figure 2.26 shows the internal forces acting on a wall element of height h and length lW.
N, V, M1 and M2 are the normal force, the shear force, both considered constant over the
element height, and the bending moments at the top and bottom of the wall element
respectively. The thickness of the wall is denoted by t. It is assumed that the normal force
N is given by the vertical loads and therefore approximately known. Note that strictly
speaking, in the case of coupled walls, part of the total overturning moment is carried by
normal forces in the outer piers and hence the normal force N does not only depend on
the vertical loads but also on the total overturning moment and the coupling effect. In
fact, due to the cyclic nature of the earthquake, the normal force will vary from one cycle
to the other about its mean value given by the vertical loads. For reason of simplicity, this
56
variation due to the overturning moment is not considered in the following and the normal
force is simply calculated as a function of the vertical loads.
The moments M1 and M2 depend on the earthquake action represented by the shear force
V, the unknown in this problem.h0 is the height of zero moment measured from the
bottom of the wall element and depends on the structural model. In the case shown in
Figure 2.26 h0 > h, indicating a rather week coupling effect, however, the equations
developed in the following hold also for h0 < h.
In order to calculate the shear capacity of a wall element, i.e. the maximum shear force
the wall element can sustain, the lower bound theorem of plasticity is used. For this a
statically admissible stress field has to be found which satisfies equilibrium and the
material conditions. A simple stress field is shown in the right picture of Figure 2.26 with
two stress struts, one vertical and one inclined at an angle to the vertical. Assuming that
the reinforced concrete floors and/or the stiff joint regions can accommodate the stresses
at the nodes the normal force N can be divided into two components NV and Nn which
satisfy the equilibrium conditions:
Nv + Nn (2.32)
V = Nv tan α (2.35)
The internal forces have to satisfy the material conditions in the stress field, given by
Equations (2.24) - (2.28). For the simple stress field chosen, using the simplified
description of the compression strength in Figure 2.25, this reduces to the following three
conditions:
Nv
f inclined = ≤ f my (2.36)
l 2v .t.(cos α )
2
≤ ( f mx − f my )
Nn
f vertical = (2.37)
l 2 n .t
with
57
l 2v = l w − 2e2v (2.39)
l 2 n = l w − 2e2 n (2.40)
The shear capacity of the wall element is obtained by gradually increasing the shear force
V until one of the three conditions (2.36) - (2.38) is violated.
Equations (2.32) - (2.38) represent a system of seven equations with 8 unknowns: NV, Nn
e1n , e2 n , e1V , e2V , V and α. Thus the system is statically indeterminate and one parameter
can be chosen freely. However, using the lower bound theorem of plasticity, the shear
capacity calculated will always be a lower bound to the real shear capacity. As we are not
interested in a conservative value but in a description as close as possible to reality, the
one parameter should be chosen so as to maximise the admissible shear force V in the
wall section. This is a trial and error process. However good results can be obtained by
choosing e 2V = e 2 n .
Note that here the term admissible is used to stress the fact that the lower bound theorem
of plasticity only gives a lower bound approximation of the shear capacity. The true shear
capacity is independent of the stress struts chosen, only the approximation of the shear
capacity using the lower bound theorem of plasticity depends on the choice of the stress
struts and hence, can be maximised.
Observations have shown, that the damage in masonry buildings due to seismic action is
usually localised in some regions (piers and spandrels, shaded) while some other regions
remain almost undamaged. Considering the flow of the forces through the structure it
seems obvious that although the spandrels transfer the horizontal forces horizontally, all
forces, vertical and horizontal, have to be transmitted to the ground through the piers.
Thus they will be the most critical part of the building determining the capacity.
58
transverse wall
wall
Site investigations of damaged buildings after an earthquake have also shown that a
considerable part of the structural damage can lie inside the building. Thus it is important
for the evaluation of masonry buildings to consider not only the external but also the
internal walls. In order to identify the structural walls in masonry buildings, a very
common practice is to consider all walls with a thickness t ≥ 12cm to be structural walls.
It is also common to consider transverse walls as individual walls with no flange action
(Figure 2.27). This can be justified by the following considerations:
• The interlocking of a wall with the transverse wall is usually very weak, especially
inside the buildings, usually only three to four stones form the connection. Thus
the transfer of shear is not guaranteed.
• To create a flange action, the shear force induced must be distributed over the
whole transverse wall. This is not possible for longer walls. The introduction of an
effective width taking into account part of the transverse wall would be possible,
however, because the increase in resistance is very small, the influence of the
transverse wall is usually neglected.
In fact, in most cases the coupling effect varies throughout the building. The coupling
effect in the wall plane constituting a façade will be different from the coupling effect in a
wall plane in the interior of the building. Figure 2.28 shows three cases of coupled
masonry walls. In Figure 2.28 a) the opening is relatively small and the spandrels rather
deep producing a considerable coupling effect. This is a typical case of a façade wall
plane. In Figure 2.28 b) the opening is larger and hence the coupling effect will be
reduced. This is usually the case for door openings. In Figure 2.28 c) the opening extends
over the whole storey height. This is a typical example of modern interior wall planes.
59
a) b) c)
The analysis of a masonry building as a system of coupled walls using a frame model
assumes that the spandrels are able to carry the bending moments that balance the
bending moments in the piers. As the normal forces in the spandrels are relatively low,
the bending moment capacity of the spandrels is rather low. Thus the spandrels are
usually the first to crack reducing the coupling effect.
The bilinear capacity curve of a wall is determined by three parameters, the maximum
shear strength of the wall Vm, the nominal yield displacement at the top of the wall ∆y and
the nominal ultimate displacement at the top of the wall ∆u. In the next three sections it is
demonstrated how these three parameters can be determined for a masonry wall. In the
following, even though masonry structures do not yield, the point of transition between
the linear elastic and perfectly plastic region of the bilinear capacity curve will be referred
to as the “yield point” in analogy to reinforced concrete structures.
According to definition, a wall consists of several piers, one per storey, separated by
relatively stiff joint regions. It follows that the capacity of a wall is determined by the
capacity of the pier that fails first. The joint regions as well as the spandrels are not
considered, assuming that they can accommodate the internal forces that are required for
equilibrium conditions. As already mentioned, this is not really the case for the spandrels
which usually won’t be able to accommodate the forces required for equilibrium at the
point of failure of the first pier without damage, but will have cracked. This can be
considered in a very simplified way by introducing a reduced stiffness of the spandrels,
60
thus taking into account a reduced coupling effect due to the formation of early cracks in
the spandrels. Therefore, in order to calculate the capacity curve of a wall, the capacity of
a pier of height hP and length lW is considered first. Figure 2.29 is in essence a repetition
of Figure 2.26, the wall element being now a pier.
In Figure 2.29 is shown h0 < hP indicating a rather strong coupling effect. To take into
account the reduction of the coupling effect due to the formation of early cracks in the
spandrels a reduction of the stiffness of the spandrels by 50% is proposed.
M1=V. h0 (2.42)
For h0 < hp, the bending moment in the pier changes its sign as in the case of Figure 2.29.
For h0 < hp, the bending moment does not change its sign. This is the case in Figure 2.26.
Having obtained M1 and M2 from Equations (2.41) and (2.42), the lower bound theorem
of plasticity can be used to calculate the shear capacity of the pier, with h = hp. That pier
of a wall that yields first under the given force distribution determines the shear capacity
of the whole wall. In most cases it is the pier at ground level, but in cases where there is a
61
high variation of stiffness and mass over the height of the building, a pier at mid height
can yield first.
The effective stiffness of the linear elastic part can be determined using a secant stiffness
at Vm, with 0.6 < α < 0.75, and choosing Vm such that the bilinear curve is equivalent to
the experimental curve in an energetic sense. Comparisons with experimental results
show that in this case a stiffness reduction factor of 0.5 to 0.7 is appropriate:
Figure 2.30 shows the elastic displacement shape of a six-storey masonry building due to
a triangular distributed horizontal force using frame analysis.
It can be seen that the displacement at the bottom is less than one would expect for a
frame structure. This is due to the fact that the walls in a masonry structure are rather
massive and hence the built in condition at the base of the building becomes more
significant than in the case of reinforced concrete frame structures where the columns are
rather slender. Also given is the displacement using a constant drift over the building
height equal to the drift of the first storey and it can be seen that this simple
approximation gives good results for the top displacement.
Using the assumption of constant drift δ over the building height Htot the elastic top
displacement ∆ of a wall can be written as follows:
∆ = δ . Htot. (2.44)
62
The drift can be calculated using the principle of virtual work. Considering a pier of
height hp, with the point of zero moment at a height h0 , the bending moment and shear
force distributions due to a) the real forces and b) a virtual unit force are shown in Figure
2.31. This assumes that the drift of the first storey is in fact equal to the drift of the pier.
In reality the displacement over one storey is not uniform. The piers as the most slender
part of a wall will deform the most whereas the joint regions are rather stiff and will
deform less. Hence, the assumption of a constant drift equal to the drift of the pier
overestimates the linear elastic deflection. Nevertheless, for the purpose of this evaluation
method, the results are good enough.
The horizontal deformation at the top of the pier in the direction of the unit force can then
be calculated:
hP hP
M ( x).m( x) k .V ( x).V ( x)
d=∫ dx + ∫ dx. (2.45)
0
EI eff 0
GAeff
with
EIeff and GAeff are the effective flexural stiffness and the effective shearing stiffness of the
pier, respectively.
Figure 2.31: Bending moment and shear force distribution of the bottom pier of a
wall due to a) the real forces and b) a virtual unit force
Substituting the expressions for M(x), V(x), m(x) and v(x) in Equation (2.45) and
integrating from zero to hp the displacement at the top of the pier d is obtained:
3 2
V .hP M .h V .hP
d= + 1 P + k. . (2.46)
3.EI eff 2.EI eff GAeff
63
k is a form factor depending on the particular shape of the cross section. For a rectangular
cross section k = 6/5.
Substituting Equation (2.41) for M1 and dividing by the height of the pier hp an
expression for the drift δ is obtained:
⎛ hP .(3h0 − hP ) k ⎞
δ = V .⎜⎜ + ⎟
⎟
(2.47)
⎝ 6.EI eff GAeff ⎠
The yield displacement at the top of the wall ∆y can then be determined as the
displacement for V = Vm using Equations (2.47) and (2.45):
⎛ h .(3h0 − hP ) k ⎞
∆ y = Vm .H tot ⎜ P + ⎟ (2.48)
⎜ 6.EI GAeff ⎟
⎝ eff ⎠
Even though Figure 2.31 shows the bending moment and shear force distribution for h0 <
hP , Equation (2.47) is also true for the case h0 > hP, since the boundary conditions are
fully determined by h0 .
The point at which the shear strength of the masonry wall is reached does not necessarily
imply failure of the wall. Unreinforced masonry walls need not to be considered brittle; in
fact, measured behaviour of wall elements showed that unreinforced masonry can possess
considerable capacity for plastic deformations (ETZ Zurich, 1984)Plastic deformations as
large as eight times the yield deformations were observed! Based on a linear elastic-
perfectly plastic behaviour, the ultimate displacement ∆ u can therefore be expressed:
∆ u = µW .∆ y (2.49)
µW is the ductility of the wall. In the next section, the ductility of masonry is discussed.
64
i) Displacement ductility of a wall element µWE
properties expressed by the compression strength f mx and the boundary conditions bc:
⎛ h ⎞
µ = µ ⎜⎜ σ n , , f mx , bc ⎟⎟ . (2.50)
⎝ lW ⎠
Figure 2.32: Relationship between ultimate drift and normal stress of the wall
elements tested at the ETH Zürich (1984)
The dependence of the ductility on the normal stress is quantified using the results of the
masonry wall elements tested at the ETH Zürich by Ganz (1984]. Using a linear
interpolation of the test results (Figure 2.32), the ultimate drift of a wall element δn (in
[%]) is determined as a function of the normal stress σn (in [MPa]) acting on the wall
element:
65
σ u = 0.8 − 0.25σ n (2.51)
δu
µWE = . (2.52)
δy
Determining first the ultimate drift is more reliable since the value of µWE depends on the
yield point which, however, involves high uncertainties.
Equation (2.51) illustrates the fact that for low axial forces the cracks usually pass by the
bed joints in a diagonal pattern; the separated parts of the wall can slide onto each other
resulting in large relative deformations without significant loss of strength. For high axial
loads, however, the cracks pass through the brick units and as a result the separated parts
of the wall tend to slide with little deformation downwards along the more regular crack
surfaces.
As the tests were all carried out on test specimens with the same geometry (walls of the
same height, length and thickness) and with the same compression strength orthogonal to
the mortar bed f mx = 8.25 MPa, neither the influence of the geometry nor the influence of
the compression strength on the ductility could be studied. The influence of the geometry
was revealed by cyclic static tests on masonry test specimens of different aspect ratios
carried out at the Joint Research Centre of the European Commission, Ispra, Italy
(Anthoine A., 1994). The results show that in wall elements with high aspect ratios a
rocking mode develops leading to high drift capacities without apparent strength
degradation. Wall elements with low aspect ratios, however, tend to fail in shear leading
to reduced drift capacities. To take into account the different behaviour of masonry wall
elements with different aspect ratios it is suggested to enhance or reduce the ultimate drift
capacity of a pier in the following simplified manner:
⎧ hP
⎪ 0.8(0.8 − 0.25σ n ), < 0 .5
⎪ lW
⎪ h
δ u = ⎨0.8 − 0.25σ n ,0.5 < P < 1.5 (2.53)
⎪ lW
⎪ hP
⎪ 1.2(0.8 − 0.25σ n ), l > 1.5
⎩ W
This, however, needs further investigations, especially considering the energy dissipation
of a rocking mode and its implication on the strength reduction factor R.
66
An upper limit of the ductility of a wall element is set to µWE max = 12 .
In most cases, where the walls are quite slender, it is usually the bottom pier that first
reaches the limiting conditions and enters the plastic state. Assuming the upper part of the
walls as well as the spandrels to remain elastic, a pier sway mechanism is formed. This is
illustrated in Figure 2.33.
[
∆u = ∆ y + du − d y ] (2.54)
d y is the deformation of the bottom pier at yield, ∆ y is the top displacement of the wall
at the point of yield of the bottom pier (corresponding to the point of yield of the whole
wall) and d u the ultimate deformation of the bottom pier:
d u = µWE .d y . (2.55)
Thus the displacement ductility of the wall can be deduced assuming a linear elastic
displacement shape:
hP
µW = 1 + (µWE − 1) (2.56)
H tot
In the case of squat walls combined with low normal forces it is the upper part that first
reaches the limiting condition determined by sliding along the mortar beds (Equation
(2.38)). Some thoughts to this case can be found in (Lindemuth A. 2000).
67
2.6.5 Cyclic loading
So far, the loading considered has been applied monotonically. The effect of the cyclic
nature of the earthquake action has now to be taken into account. In reality the cyclic
behaviour of masonry structures is very complex and not very easily quantified. However,
for the purpose of this evaluation a very simple approach is used by introducing two
reduction factors, a force reduction factor RFF and a deformation reduction factor RFD. A
similar approach has already been used for reinforced concrete structures in the form of
an equivalent ductility factor (Fajfar P., 1998). The idea is to reduce the monotonic force
and deformation capacity of the structure to take into account the increased detoriation
due to cyclic loading. The study of the experimental results of Ganz (1984) and
Schwegler (1993) suggests the following reduction factors:
(Vm )cyclic and (Vm )monoton are the shear capacity, (µW )cyclic and (µW )monoton and are he
ductility of the wall element under cyclic loading and under monotonic loading
respectively.
Cyclic static tests on two masonry wall elements at the University of Illinois, however,
revealed virtually no detoriation due to cyclic loading. Reversing the lateral force closed
the previous opened cracks and resulted in an identical crack pattern. No reduction of
stiffness was observed. This would rather suggests reduction factors close to one.
In order to obtain the vulnerability function of a masonry building, i.e. the damage as a
function of the spectral displacement, the displacement at the top of the building ∆ is
associated with the damage grades according to the European Macroseismic Scale (EMS
98). For each damage grade, indicators are defined that allow the identification of the
points at which the building enters the next damage grade on the capacity curve of the
building.
A summary of the indicators of the damage grades for masonry buildings according to the
European Macroseismic Scale (EMS 98) is given in Table 2.4.
Even though in terms of financial loss damage grade 4 and damage grade 5, both indicate
100% loss (or even more, considering that the building first has to be pulled down
68
completely before a new building can be constructed, thus increasing the cost) it is
important to distinguish between these two, as in terms of casualties, the chance of
survival in a house which is very heavily damaged but has not collapsed yet is much
higher. However, in some cases the failure of the first wall will lead immediately to the
collapse of the building and hence DG4 = DG5.
It is not always possible to identify clearly all damage grades. A very common case is that
the first wall reaches its ultimate condition and fails, indicating damage grade 4, before
the last wall has yielded, indicating damage grade 3.
Damage
EMS 98 Identification
grade
Negligible to slight damage (no structural
damage, slight non-structural damage) point of onset of cracking,
Hairline cracks in very few walls. => stress distribution becomes
DG1
Fall of small pieces of plaster only. zero at the extreme fibre of the
Fall of loose stones from upper parts of build- wall section
ings in very few cases.
DG2 Moderate damage (slight structural
behaviour of the building
damage, moderate non-structural damage)
Cracks in many walls. becomes nonlinear,
Fall of fairly large pieces of plaster. Partial the stiffness of the building starts
col- to reduce,
lapse of chimneys. => yield of the first wall
Substantial to heavy damage (moderate
structural damage, heavy non-structural increased nonlinear behaviour of
damage) the building,
Large and extensive cracks in most walls. the stiffness of the building tends
DG3 Roof tiles detach. Chimneys fracture at the
to zero,
roof
-> yield of the last wall
line; failure of individual non-structural ele-
ments (partitions, gable walls).
Very heavy damage (heavy structural dam-
age, very heavy non-structural damage)
DG4 Serious failure of walls; partial structural fail- => failure of first wall
ure of roofs and floors.
Destruction (very heavy structural => drop of the base shear of the
DG5 damage) building Vb
Total or near total collapse. below 2/3 • Vbm
69
2.6.7 Vulnerability function of a masonry building
The top displacement ∆ can be plotted as a function of the spectral displacement Sd(f1).
Having identified the points on the capacity curve of the building at which the building
enters the next damage grades, these can be presented on the plot of top displacement ∆
versus spectral displacement Sd(f1) resulting in the vulnerability function of the building.
So far the in-plane behaviour of masonry walls aligned in the direction of the earthquake
was considered. However, masonry walls aligned orthogonal to the earthquake direction
can also fail in an out-of-plane mode and this may endanger the gravity load carrying
capability of a building (Figure 2.34).
The out-of-plane behaviour of masonry walls depends very much on the floor-wall
connections. For masonry walls which are properly anchored to the floors, the out-of-
plane behaviour is usually not critical and the vulnerability function of the building is
determined by its in-plane behaviour. In the case where the connection between
orthogonal walls and between walls and floors is rather poor, the walls might fail in an
out-of-plane mechanism before an in-plane mechanism can be triggered leading to a
correction of the vulnerability function established for in-plane behaviour. In the absence
of any floor-wall connections, the masonry walls behave like tall unrestrained cantilevers
which are most vulnerable to flexural out-of-plane failure determining the vulnerability
function of the building.
70
Furthermore, in reality the earthquake action does not correspond to one of the principal
directions of the building. Thus the walls are subjected to both, in-plane and out-of-plane
actions. However, this is not considered in the following.
Two different stages in the out-of-plane behaviour are distinguished, the occurrence of
cracking and failure.
As for in-plane behaviour, the floors are assumed to be completely rigid (no amplification
of the out-of-plane loaded wall accelerations by floor response) and hence the
accelerations at all points along a floor will be equal to the acceleration of the in-plane
loaded walls at the floor height. An estimate of the storey accelerations from the spectral
acceleration of the equivalent SDOF Sd(f1) is shown in Figure 2.35 (Paulay T., 1992)
At heights above the height hE of the equivalent SDOF system the storey acceleration at
the i-th storey ai is given by the mode shape (assumed to be linear) from the spectral
acceleration Sa(f1):
s a ( f1 )
ai = hi (2.58)
hE
At heights below hE the influence of the ground acceleration is taken into account by a
linear interpolation between ag and Sa(f1):
⎛ S a ( f1 ) − a g ⎞
a i = hi ⎜⎜ a g + ⎟⎟ (2.59)
⎝ hE ⎠
71
The capacity of unreinforced masonry subjected to out-of-plane action depends upon the
dimensions of the wall, the boundary conditions, the compressive stress and the tensile
strength of the masonry. Let’s consider a wall panel between two floor levels of height hst
and length l, the applied normal forces due to the gravitational loads indicated by N.
Assuming the acceleration at the i-th ai storey to be constant over the storey height, the
out-of-plane loading of a wall due to inertia is:
q = m . ai (2.60)
The maximum moment in the wall panel due to the out-of-plane loading will be
determined by the boundary conditions. Each wall panel has four boundaries. For a wall
panel of height hst the top and bottom boundary conditions are given by the floor-wall
connection. In the case of free boundary conditions at both sides, the wall panel can be
regarded as a one-way slab and the problem is reduced to a 2-D problem. This is shown in
Figure 2.36. In the case where the sides are somehow supported, the wall panel has to be
regarded as a two-way slab. The effect is to reduce the maximum moment due to the out-
of-plane loading. However, the influence of the supported sides on the maximum moment
reduces very rapidly with increasing length to height ratio and hence most out-of-plane
problems can be reduced to a 2-D problem.
Note, for design purpose, considering a 2-D problem always gives a conservative result.
Depending on the floor-wall connection, the maximum moment in the wall element can
be expressed as a function of the vertical force distribution q.
An example of this case would be a concrete floor system, where the normal forces are
usually high, preventing the top and the bottom of the wall element to rotate, or a timber
floor system with ties connecting the wall element to the floors.
For a simply supported beam (Figure 2.36 ii) the maximum moment is:
2
qh
M max = st (2.62)
8
72
An example of this case is a floor system with joints at the top and bottom of the wall
element.
In the case of a cantilever (Figure 2.36 iii) the maximum moment is:
2
qh
M max = st . (2.63)
2
On the analogy of in-plane behavior the onset of cracking is determined by the stress
distribution becoming zero at the extreme fibre neglecting the tensile strength of masonry
(Figure 2.37) (Paulay T., 1992).
73
Nt
M qcr = (2.64)
6
Knowing the normal force acting on a wall panel, the out-of-plane moment that causes
cracking can be determined using Equation (2.64). From this, using Equations (2.61) to
(2.63), depending on the floor-wall connection, the vertical force distribution that causes
cracking qcr and thus the acceleration at the i-the storey that causes cracking acri can be
calculated using Equation (2.60).
Hence, knowing acri , the spectral acceleration at which cracking occurs in an out-of-plane
mode Sa(f1)OPcr can be determined using Equations (2.58) and (2.59).
As can be seen from Figure 2.35, the storey acceleration increases with height, the
maximum storey acceleration being at the roof level. Since this is combined with the
lowest normal force N, and hence the lowest admissible Mqcr, cracking in an out-of-plane
mode occurs first in a wall panel at the upper floor level.
So far it was assumed that the floors are horizontal and the normal force is applied
centrically on the wall panel (cf. point of application of N in Figure 2.36). However, the
surcharge being applied on the floors, the floors will bend creating an additional out-of-
plane moment which can be additive or subtractive to the moment due to the storey
accelerations Mq.
Assuming the floors to be “built in” into the walls, the bending moment distribution is
given in Figure 2.38.
74
The moment at the fixed ends due to the horizontally distributed load si at the i-th storey
is:
s i L2
dm = (2.65)
12
On the other hand, if acts into the opposite direction it will counteract Mq , increasing the
“net” moment due to the storey acceleration that causes out-of-plane cracking. In reality,
the direction of the floor acceleration will change during the earthquake and hence there
will always be a point at which dm will be additive. It seems therefore appropriate to take
this effect into account using Equation (2.66).
The formation of cracks does not imply out-of-plane failure of the masonry wall panel.
After the onset of cracking, the crack will propagate through the thickness of the wall, the
maximum compressive stress will increase and the compression zone decrease until at
ultimate condition the compressive strength of masonry orthogonal to the mortar bed f mx
is reached. The stress distribution at ultimate can be approximated by a rectangular
distribution as shown in Figure 2.39 a) and hence the moment at ultimate is:
t−a
Mqu = N (2.67)
2
N
a= (2.68)
f xm l
1
m= n(1 − n ) (2.69)
2
75
with
M qu N
m= 2
and n = 2
(2.70)
lt f mx lt f mx
Thus, from the applied normal force N, the moment Mqu that causes out-of-plane failure
can be determined and hence the vertically distributed load qu that causes out-of-plane
failure using Equations (2.62) to (2.64) depending on the floor-wall connection. Using
Equation (2.60) the acceleration at the i-th aui storey that causes out-of-plane failure of the
wall panel can be calculated from which the spectral acceleration Sa(f1)OPu at which the
wall panel fails in an out-of-plane mode can be determined using Equations (2.58) and
(2.59).
Again, the influence of the additional out-of-plane moment due to bending of the floors
has to be taken into account, the “net” moment due to floor acceleration that causes out-
of-plane failure being:
So far out-of-plane cracking and out-of-plane failure have been considered and the
corresponding spectral accelerations at the fundamental frequency of the building,
Sa(f1)OPcr and Sa(f1)OPu, determined. To compare these values with the vulnerability
function, for in-plane behaviour, the spectral accelerations have to be converted into
76
spectral displacements. The spectral displacement at which out-of-plane cracking occurs
Sd(f1)OPcr can then be compared with the spectral displacement at which cracking occurs
in-plane, corresponding to damage grade 1, Sd(f1)DG1 . If Sd(f1)OPcr < Sd(f1)DG1, cracking in
an out-of-plane mode occurs first and thus the building enters earlier damage grade 1.
Hence the vulnerability function must be corrected by displacing the point at which the
buildings enters damage grade 1 down to Sd(f1)OPcr. If Sd(f1)OPcr > Sd(f1)DG1, cracking in an
out-of-plane mode occurs after cracking in an in-plane mode has occurred and thus no
correction of the vulnerability function is needed.
Considering out-of-plane failure, two different cases have to be distinguished. In the case
of a structural wall panel, the out-of-plane failure corresponds to damage grade 4 (serious
failure of walls) and the resulting spectral displacement Sd(f1)OPu has to be compared with
the spectral displacement corresponding to damage grade 4, Sd(f1)DG4. If Sd(f1)OPu <
Sd(f1)DG4, out-of-plane failure occurs before the in-plane failure. Hence the vulnerability
function must be corrected by displacing the point at which the building enters damage
grade 4 down to Sd(f1)OPu . If Sd(f1)OPu > Sd(f1)DG4, in-plane failure occurs before out-of-
plane failure and thus no correction of the vulnerability function is needed. In the case of
a gable wall, the resulting spectral displacement Sd(f1)OPu has to be compared with the
spectral displacement corresponding to damage grade 3, Sd(f1)DG3, as the failure of a gable
wall corresponds only to a moderate structural damage.
The calculation of the out-of-plane loading, expressed as the vertical force distribution q,
using Equations (2.61) to (2.63) depends strongly on h, the height of the wall panel. So
far it was assumed that the wall panel considered extends over a storey height i.e. h = hst.
In the absence of any floor-wall connection the overall out-of-plane behaviour of a wall
plane has to be assessed. In that case the wall plane will behave as a tall unrestrained
cantilever with a high vulnerability to out-of-plane loading. The same principles as
outlined above can be applied.
Even for good floor-wall connections the predicted out-of-plane resistance is rather low.
Although for masonry buildings with poor floor-wall connections the out-of-plane
behaviour is predominant, the results tend to be rather too conservative. One reason is the
neglect of the tensile strength of masonry which becomes particularly important in the
case of low compressive stresses such as gable walls. Furthermore, although the moment
77
that causes out-of-plane failure Mqu is calculated considering ultimate conditions, the
corresponding spectral acceleration Sa(f1)OPu is still assumed to be an elastic spectral
value without considering the inelastic behavior.
78
CHAPTER THREE
3.1 Introduction
During the last 240 years, a number of earthquakes including seven major earthquakes
(M>7) have affected Bangladesh (see Appendix A). In Figure 3.1 is shown Location of
earthquake epicenter, at and near Bangladesh, occurrence period 1869 to 2000 (Yasin,
2008). In 1897, The Great Indian Earthquake of magnitude 8.7 caused serious damages to
buildings in the northeastern part of India (including Bangladesh) and 1542 people were
killed. During this earthquake almost all existing masonry structures of Bangladesh were
either partially or fully destroyed. The epicenter was at 230 kms from Dhaka City. Figure
3.2 presents an isoseismal map of the Great Indian Earthquake, showing the area under
which the masonry structures were affected.
The extensive loss of life and property caused by earthquakes may be reduced to a
considerable degree by the adaptation and implementation of improved design and
construction procedures practicable within the context of the cultural and socio-economic
constraints prevailing in this country. When housing, built by traditional methods and
using conventional building materials, does not exhibit the necessary characteristics of
earthquake resistant housing, new designs and non-traditional building materials and
construction techniques need to be developed. It must be emphasized that carefully
conceived structural seismic planning and design with good materials, workmanship and
construction practices have generally been able to withstand the fury of earthquakes to a
large extent.
79
The approach followed (Laurenco and Oliviera, 2004) proposed in this paper aims at a
much more simple, fast, and low cost procedure based on a simplified geometric
approach for immediate screening of the large number of buildings at risk. The objective
is to evaluate the possibility of adopting simple indexes related to geometrical data as a
first (very fast) screening technique to define priority for further studies with respect to
seismic vulnerability. These fast techniques are to be used without actually visiting the
buildings, being therefore not accurate. It is expected that the geometrical indexes could
detect cases in serious risk and, thus, define priority for studies in countries/locations
without recent earthquakes. The historical buildings considered at possible risk may
deserve more detailed studies using advanced computer simulations, together with
adequate material and structural characterization.
In the case of urban areas, and in spite of the diversity, a common matrix can usually be
established for the seismic areas, more structural than technological. This consists of low
80
building height (up to three stories), moderate spans (maximum of four or five meters),
and large thickness of the walls (Giuffrè 1995; Lourenco and Oliviera, 2004).
3.2 Methodology
The usage of simplified methods of analysis usually requires that the structure is regular
and symmetric, that the floors act as rigid diaphragms, and that the dominant collapse
mode is in-plane shear failure of the walls (Meli 1998). In general, these last two
conditions are not verified by ancient masonry structures, meaning that simplified
methods should not be understood as quantitative safety assessment but merely as a
simple indicator of possible seismic performance of a building. The following simplified
methods of analysis and corresponding Indices are considered.
In-plane Indices:
Out-of-plane Indices:
81
- Index 6: Thickness to height ratio of perimeter walls
V Bhutan
Main Boundary Thrust
VII India
Brahmaputra Fault st
ru
Th
st
IX
ru
ga
Th
Na
ng
X VI
flo
Ha
Dauki Fault
VII
Haluaghat Fault
lt
au
lt
au
F
Bangladesh
tF
gra
lhe
Bo
Sy
Dhaka
Tripu
India
ra
Chitta
g
ong F
Myanmar
old B
elt
0 100 200
kilometres
Epicentre
Concealed Faults
Bay of Bengal
Thrust Faults
Fold Axis
Normal Faults
These methods can be considered as an operator that manipulates the geometric values of
the structural walls and columns and produces a scalar. As the methods measure different
quantities, their application to a large sample of buildings contributes to further
enlightening on their application. As aforementioned, a more rigorous assessment of the
actual safety conditions of a building is necessary to have quantitative values and to
define remedial measures, if necessary.
The simplest index to assess the safety of ancient constructions is the ratio between the
area of the earthquake resistant walls in each main direction (transversal y and
longitudinal x, with respect to the central axis of the mosque) and the total in-plan area of
the building. According to Eurocode 8 (CEN-EC8 2003), walls should only be considered
82
as earthquake resistant if the thickness is larger than 0.35 m, and the ratio between height
and thickness is smaller than nine. The first index, γ1,i reads:
where Awi is the in-plan area of earthquake resistant walls in direction “i” and S is the
total in-plan area of the building.
The nondimensional index, γ1,i is the simplest one, being associated with the base shear
strength. Special attention is required when using this index as it ignores the slenderness
ratio of the walls and the mass of the construction. Eurocode 8 recommends values up to
5–6% for regular structures with rigid floor diaphragms. In cases of high seismicity, a
minimum value of 10% seems to be recommended for historical masonry buildings (Meli
1998). For simplicity sake, high seismicity cases can be assumed as those where the peak
ground acceleration for soft soils, established for a 475 y.r.p., is equal or larger than
0.20g.
This index provides the ratio between the in-plan area of earthquake resistant walls in
each main direction (again, transversal y and longitudinal x) and the total weight of the
construction, reading:
where Awi is the in-plan area of earthquake resistant walls in direction “i” and G is the
quasi-permanent vertical action. This index is associated with the horizontal cross-section
of the building, per unit of weight. Therefore, the height (i.e. the mass) of the building is
taken into account; a major disadvantage is that the index is not nondimensional, meaning
that it must be analyzed for fixed units. In cases of high seismicity, a minimum value of
1.2 m2/MN seems to be recommended for historical masonry buildings (Meli 1998), but
on the basis of a recent work (Lourenço and Roque 2004), a minimum value of 2.5
m2/MN is adopted here for high seismicity zones.
The total design base shear for rigid structurs in a given direction shall be determined
from the following relation (BNBC):
V = 0.5ZIW (3.3)
83
Where,
Z= Seismic zone for a building site shall be determined based on the location of the
site on the Seismic Zoning Map provided in Figure 3.3.
84
I= Structure importance coefficient
Finally, the base shear ratio provides a safety value with respect to the shear safety of the
construction. The total base shear for seismic loading (VSd, base = FE) can be estimated
from an analysis with horizontal static loading equivalent to the seismic action (FE = βG),
where β is an equivalent seismic static coefficient related to the peak ground acceleration.
The shear strength of the structure (VRd, base = FRd) can be estimated from the contribution
of all earthquake resistant walls FRd,i = Σ Awi fvk, where, according to Eurocode 6 (CEN-
EC6 2003), fvk = fvk0 + 0.4σd. Here, fvk0 is the cohesion, which can be assumed equal to a
low value or zero in the absence of more information, σd is the design value of the normal
stress, and 0.4 represents the tangent of a constant friction angle φ, equal to 22º. The
index, γ3, reads:
If a zero cohesion is assumed (fvk0 = 0), γ3,i is independent from the building height,
reading:
For a non-zero cohesion, which is most relevant for low height buildings, γ3,i reads:
where Awi is the in-plan area of earthquake resistant walls in direction “i,” Aw is the total
inplan area of earthquake resistant walls, h is the (average) height of the building, γ is the
volumetric masonry weight, φ is the friction angle of masonry walls, and β is an
equivalent static seismic coefficient. Here, it is assumed that the normal stress in the walls
is only due to their self-weight, i.e. σd = γ × h, which is on the safe side and is a very
reasonable approximation for historical masonry buildings, usually made of very thick
walls.
Equation (3.6) must be used rather carefully since the contribution of the cohesion can be
very large. Within the scope of this work, a cohesion value of 0.05 N/mm2 is assumed.
This nondimensional index considers the seismicity of the zone, taken into account in β.
The building will be safer with increasing ratio (earthquake resistant walls/weight), i.e.
larger relation (Awi / Aw) and lower heights. For this type of buildings and action, a
minimum value of γ3,i equal to one seems to be acceptable.
85
(a)
0.1
safe
unsafe
0
0 0.25
PGA/g
(b)
y = 8.5716x + 0.3571
Index-2 (Area to weight ratio)
2.5
safe
1.0 unsafe
0
0 0.3x0.25 0.25
PGA/g
(c)
safe
Index-3 (Base shear ratio)
y =1
1
unsafe
0
0 0.1 0.2 0.3
PGA/g
Figure 3.4: Assumed thresholds as functions of PGA/g: (a) index 1, (b) index 2 and
(c) index 3 (Laurenco and Oliviera, 2004)
86
80 (d)
(0.15, 70)
safe
40
20 (0.4, 20)
0
0 0.1 0.2 0.3 0.4
PGA/g
(e)
0.3
Index-5 (t/h of columns)
( 0.4, 0.2)
0.2
safe
0.1 unsafe
( 0.15, 0.05)
0
0 0.1 0.2 0.3 0.4
PGA/g
0.1 (f)
( 0.4, 0.08)
Index-6 (t/h of perimeter walls)
0.05
safe9.5
unsafe
(0.15, 0.02)
0
0 0.1 0.2 0.3 0.4
PGA/g
Figure 3.5: Assumed thresholds as functions of PGA/g: (a) index 4, (b) index 5 and
(c) index 6 (Laurenco and Oliviera, 2004)
87
The adopted Indices measure rather different quantities and can hardly be compared.
Index 2 is dimensional, which means that it should be used with particular care. Index 1
and index 2 are independent of the ground acceleration. Therefore, assuming that the
buildings must have identical safety, these indexes should be larger with increasing
seismicity. For indexes 1 and 2, the seismicity is taken into account by considering that
the threshold value, defined above, is valid for a PGA value of 0.25 g and assuming its
linear variation with PGA/g, as illustrated in Figure 3.4 (a) and (b), see also Eurocode 8
(CEN-EC8 2003). On the other hand, index 3 should be constant in different seismic
zones, as it intrinsically considers the effect of seismicity. This index format is close to
the traditional safety approach adopted for structural design, being the threshold value
equaled to 1.0; see Figure 3.4 (c). Even though, Proposed threshold values are plotted for
Portugal, Spain, and Italian churches, in this thesis assumed that these threshold values to
be equally applicable for identifing the vulnerable historical selected structures.
Besides the three Indices exposed above, other key Indices related to structural
performance were computed for the monuments under analysis. In this study, three
geometric ratios concerning the structural out-of-plane behaviour of columns and walls in
main space were adopted, when applicable; slenderness ratio (γ4), and thickness to height
ratio of the columns (γ5), as well as thickness to height ratio of the perimeter walls (γ6),
were analyzed, reading:
where hcol is the free height of the columns, I and A are the inertia and the cross section
area of the columns, respectively, dcol and tcol are the (equivalent) diameter and thickness
of the columns, respectively, and twall and hwall are the thickness and the (average) height
of the perimeter walls, respectively. All of the out-of-plane Indices are dimensionless and
do not consider the local seismicity. If identical safety factors for the monuments are
assumed, these Indices should vary with increasing seismicity; namely, index 4 should
decrease and indexes 5 and 6 should increase. . For indexes 4, 5 and 6, Assumed
88
thresholds as functions of PGA/g are shown in Figure 3.4 (a), 3.4 (b) and 3.4 (c)
respectively,
Eqs. 3.1, 3.2, 3.4 and 3.5 can be recast in a similar format as a function of the ratio (Awi /
Aw), which allows direct comparison between the different methods, as
k1 = Aw / S; k2 = 1 / (γ x h) ; k3 = tanφ / β
Here, it is stressed that the ratio (Awi / Aw) represents the percentage of earthquake
resistant walls in a given direction in relation to the total area of earthquake
resistant walls in the building.
The new expressions for the scalar Indices indicate that they are all linearly dependent on
the ratio (Awi / Aw). This ratio provides direct information about the in plan stiffness of
the structure along each main direction and it is usually accepted that the sum of the
relations (Awi / Aw) for the two orthogonal directions can be larger than the unit value,
due to superposition of the areas in the two directions.
The Indices depend linearly also in the following quantities: (a) ratio between total area of
earthquake resistant walls and total in-plan area of the building (Index 1); (b) height of
the building (Index 2); (c) ratio between friction and equivalent seismic static coefficient
(Index 3). This stresses the fact that the Indices measure rather different quantities and
can hardly be compared between them. Index 2 is dimensional, which means that it
should be used with particular care. Index 1 and Index 2 are independent of the design
ground acceleration. Therefore, assuming that the buildings must have identical safety,
these Indices should be larger with increasing seismicity. On the other hand, Index 3
should be constant in different seismic zones, as it considers the effect of seismicity.
Finally, Index 3 format is close to the traditional safety approach adopted for structural
design.
89
3.3 Seismic Hazard Zonation of Bangladesh
The seismic map of Bangladesh is provided in Figure 48 Based on the severity of the
probable intensity of seismic ground motion and damages; Bangladesh has been divided
into three zones, i.e. Zone 1, Zone 2 and Zone 3 as shown in Figure 48 with Zone 3 being
the most severe (BNBC).
In order to understand the seismicity of the country, Bangladesh is divided into 3 zones.
Roughly it may be stated that the North-Eastern part (Zone-3 with zone factor = 0.25) of
Bangladesh is more prone for earthquake shocks than South-Eastern part (Zone-2 with
zone factor = 0.15) of Bangladesh. This does not mean in any way that strong earthquakes
are ruled out in the South-Eastern part and South-Western part (Zone-1 with zone factor =
0.075). It is estimated that nearly total area of Bangladesh is in the active seismic zone.
The proposed geometrical indices of monuments located in different seismic areas are
compared with the respective seismic hazard, expressed by the peak ground acceleration
(PGA). Hazard curve of different latitude and longitude with grid interval of 0.3 degree
are developed (after Yasin 2008). PGA for 0.002, 0.001 and 0.0004 annual probability of
exceedence are picked and used to plot the Seismic hazard map (PGA contour) of
Bangladesh for ten percent, five percent and two percent probability of exceedence in 50-
years period respectively (see Appendix B). The map developed using Boore et al.
(1993), McGuire (1978) and Duggal (1989) acceleration attenuation expressions (Yasin,
2008).
90
CHAPTER FOUR
MICROTREMOR ANALYSIS
4.1 Introduction
Bangladesh has long been believed to be a country with very medium seismic hazard. But
the practical design and analysis of buildings have not paid attention to seismic aspects.
However, in recent years, it has been revealed by many reliable reports that Dhaka, the
capital city of Bangladesh, has a potential risk from distant earthquakes due to the ability
of underlying land fill to amplify the ground motion. For this reason, the seismic design
and analysis of buildings cannot be neglected any more. Nevertheless, the dynamic
properties of buildings, which are important to seismic design and analysis, have been
limited studies in Bangladesh, particularly in low and medium-rise buildings. The
identification of dynamic properties of buildings is therefore required and it will provide
useful information for the development of design criteria of buildings in Bangladesh.
4.2 Background
91
members and the rigid foundations, is not appropriate to identify the dynamic properties
of buildings because the dynamic properties mainly depend on the total stiffness of the
buildings, which is also influenced by another assumption such as the stiffness of non-
structural members, and the flexibility of the foundations. Following to incomplete
modeling assumptions, the dynamic properties of buildings, which are calculated by
numerical analysis, are not reliable. Then, these will lead design engineers to make
mistakes when the dynamic properties of buildings are considered in design and analysis.
In order to identify correct dynamic properties, the most accurate approach is the direct
measurement approach because the properties are derived from actual dynamic response
of existing buildings. In recent years, several direct measurement techniques for
determining the dynamic properties of structures have been developed. These techniques
can be categorized into three basic methods:
In the forced vibration method, a structure is excited into a steady state response by
mechanical shakers and its response is measured. Plotting the amplitudes of the responses
against frequencies provides a frequency-response curve from which dynamic properties
can be determined. In the free vibration method, a step or an impulse force is applied to
the structure. A decay response curve can be utilized for system identification. The last
method is the ambient vibration method. The ambient responses of a structure, which are
generated by microtremor excitations such as wind forces, and traffic excitations, are
measured. Dynamic properties of the structure are extracted from processing signals in
the time domain technique or in the frequency domain technique. Both forced and free
vibration approaches are expensive compared to the ambient vibration approach due to
the need of mechanical shakers or impulse generators. Furthermore, in some situations
building operation is disturbed by a controlled excitation from forced or free vibration
test. In such a case, the ambient vibration measurement becomes an attractive option.
92
4.4 Ambient Vibration Technique
As mentioned above, the ambient vibration measurement requires processing signals for
identifying dynamic properties of structures. The signal processing techniques can be
categorized into time domain and frequency domain techniques. In the time domain
technique, response-time history is employed directly in the identification of dynamic
properties of structures. While in the frequency domain technique, response-time history
has to be firstly converted into frequency domain by Fourier analysis. Dynamic properties
are then extracted from the frequency spectrum, which is the plotting of Fourier
magnitude response against frequencies. Although both techniques can be employed to
identify the dynamic properties of buildings, the frequency domain technique gives a
better physical interpretation than the time domain technique because it presents the
response of buildings in the form of the frequency spectrum. This frequency spectrum can
be used directly to identify natural frequencies from the frequencies corresponding to the
peak values of Fourier magnitude and calculate vibration mode shapes from the spectral
ratio method. In addition, an algorithm in the time domain technique is more complicated
than frequency domain technique, which provides the algorithm in the form of Fast
Fourier Transform (FFT). According to this algorithm, the computational work in the
frequency domain technique can be significantly reduced. In this regard, the frequency
domain technique is an attractive technique to identify dynamic properties from the
ambient vibration measurement.
93
spectral ratio of the horizontal to the vertical components at the surface gives an estimate
of the period at which it peaks, corresponding to the site period.
The microtremor equipment consists of six velocity transducers; two horizontal and one
vertical, an amplifier, an analog-to-digital (A/D) converter and a laptop computer used for
data acquisition. For the selection of the test location, care is taken to avoid heavy traffic,
manholes, foundations and other underground structures. The sensors are placed so that
the two horizontal sensors are orthogonal, preferably facing North and East. The analysis
is carried out using Nakamura’s method, plotting the fourier spectrum of the buildings.
The most significant peak of the fourier spectrum is taken to be the dominant frequency
of the site. Microtremor instrument with sensor and battery are shown in Figure 4.1.
Figure 4.2 and Figure 4.3 show the Microtremor observation and Microtremor set up
respectively.
94
Figure 4.2: Microtremor observation
Figure 4.3: Placing of sensor at the top of the building with proper direction and
microtremor set up
95
CHAPTER FIVE
The selection of the monuments in this Simplified method and Microtremor analysis are
limited in the boundary of Bangladesh. There are selected 28 masonry structures; mainly
mosques which located in different district, built between 13th to 18th centuries in the
thesis.
Mosque had been built in Bangladesh from the very beginning of the Muslim rule. No
building of that time has survived. Bangladesh came within the ambit of Islam in the early
13th century, and from the time onward people in this region built numerous mosques. In
general the ground plan of mosques, built in greater Bengal until the end of the Mughal
rule (983A.H. /1775A.D.), falls into two broad categories - the Courtyard type and the
Enclosed type. With a single exception, all the mosques in greater Bengal had only one
enclosed prayer hall without an internal courtyard. It appears to be that the destruction of
ancient monuments took place at a greatly accelerated rate during the present and last
centuries. The devastating effect of the hydrological system or the changing of the river
course, the wild vegetation, the vandalism of building materials and the earthquakes were
responsible for the obliteration of most of the mosques in Bangladesh.
Thousands of mosques are serving the Muslim community of Bangladesh. Mosques were
built under the patronage of rulers and distinguished rich Muslims. Since last centuries
the common people started to build mosques gradually in their own initiative. These
mosques develop gradually and reflect the desire and demand of the community. In this
chapter, the rich Islamic heritage of this deltaic country can only be partially expressed.
There are 28 historical mosques which located in different districts are selected in these
analyses. Figure 5.1 shows the map of Bangladesh gives the location of all the structural
sites covered in this thesis numerically.
96
Figure 5.1: Map of Bangladesh showing the location of selected structures
97
9. Baba Adam mosque at Rampal, Munshiganj.
In the following article a short description of all the mosques are provided.
98
Archaeology and Museums, Government of Bangladesh (GoB). The mosque was built by
one Wali Muhammad during the reign of Sultan Husayn Shah.
The Chhoto Sona mosque is an isolated building from its present neighbourhood and
stands on an open piece of land in the southern side of a big tank. The area of the mosque
was enclosed with a boundary wall which does no longer exist. The front courtyard can
be reached through the main entrance gate located at the central axis. At a little distance
from the gate stand two tombs. H. Creighton suggests that the tombs belong to the
founder of the mosque and one of his kindred (Creighton 1871: XIV). On the south--
eastern side of the courtyard is the new grave of Bir Shresto Mohiuddin Jahangir buried
in 1971 during the Liberation War.
The mosque is a small building of great architectural merit in a seemingly good state of
preservation. A part of the western wall along with three domes fell during the earthquake
in 1897 and subsequently it was restored in 1900-07 by the Archaeological Department of
India (Annual Report ASI 1908-09: 45). It has only one prayer hall with simple
rectangular floor plan measuring 22.30x12.54 m from the inside. The hall is divided into
three bays and five aisles. The middle aisle is broader and divides the prayer chamber in
two parts. Each part consists of six quadrilateral areas covered by six semicircular domes.
99
The central nave is divided into three rectangular areas covered with three special forms
of vaults called Chouchala. It is derived from the traditional Bengali rural thatched roof
form. The prayer room is surrounded by a 1.78 m thick wall with an entrance from the
east by five pointed multi-foil arched openings. Corresponding to each aisle is an opening
in the eastern wall leading directly to a niche in the qibla in the western wall with five
mihrab niches. The total number of openings in both northern and southern walls is three.
Plan and front elevation of the mosque are shown in Figure 5.3 and Figure 5.4
respectively.
25.86 m
1.83m 1.3 m
1.78m
1.83m 3.6m 3.6m 3.6m 3.6m
4.5m
0.8m
0.8m
1.83m 16.1 m
1.83m
1.78m
1.83m
1.83m
1.83m
1.83m
1.83m
7.3 m
3.3m 3.3m 3.3m 3.3m 3.3m
100
Each of the four outside corners has a ten-sided turret enhancing the beauty and
monumentality of this mosque without having any functional and structural utility. There
are also two octagonal corner turrets at the back of the projected wall of the central
mihrab niche. Only the foundation of these corner turrets is in situ. The remaining parts
are lost. In the north-west corner of the prayer hall is a raised stone gallery reserved only
for the rulers or dignitaries. A flight of steps leads to the entrance platform for this
gallery, which is attached to the north-west side of the mosque. Like the traditional rural
bamboo hut, the roof of the mosque is sloped in both north-south and east-west direction
to remove rainwater. The rainwater disposes through four spouts in each side and the total
number of spouts is eight. All the walls are faced with black basalt stone up to the bottom
of the arches inside and up to the edge of the cornice outside.
Ghorar mosque (Figure 5.5) is an excellent example of the best conserved monument in
Barobazar by the Directorate of Archaeology in the southern region associated with the
history of legendary Saint Khan Jahan Ali. It is located in the village Barobazar under
Kaliganj Upazilla of Jhenidah district and about one kilometer west from the Jessore-
Jhenidah highway.
101
Barobazar in Jhenidah is believed to have derived its name from a certain twelve (Barn)
obscure Muslim saints who settled here prior to Khan Jahan Ali (Ahmed, N.1989: 28).
The architecture of Barobazar resembles with the Illyas Shahi architecture in Gaur and
Pandua than the style of Khan Jahan Ali that developed in the nearby Bagerhat. Neither
structure has any inscription, but according to its stylistic resemblance with the
architecture of Gaur and Adina; such as square room with foreroom, octagonal turret,
vertical offset and projected shallow panels of entrance, it may be assumed that this
structure was built during the Illyas Shahi period prior to the Khan Jahan architecture i.e
early 15th or late 14th century A.D.
9m
0.92m
1.1m
1.3m 6m 1.45m
12.53 m
1.45m
1.45m
102
by brick pendentives, upon which the dome rests. The foreroom is covered by three equal
hemispherical domes using pendentives at each corner. Plan and front elevation of the
mosque are shown in Figure 5.6 and Figure 5.7 respectively.
4.5 m
2.6 m 2.9 m 2.6 m
The Jore-Bangla mosque (Figure 5.8) stands at the village Barobazar under Kaliganj
upazilla in the district Jhenidah, about six kilometers west from the upazilla headquarters.
Recently one terracotta Arabic inscription, dated 800A.H./1397A.D., has been discovered
from the nearby Jore Bangla Dighi. (Banglapedia 2003). If this inscription is taken to
have any connection with the Jore Bangla mosque, the year 1397 A.D. can be assumed as
the date of its construction. Although uninscribed, the Jore Bangla mosque is dated back
to the late fourteenth century or early fifteenth century in the light of its close links with
dated monuments of that time such as Eklakhi Mausoleum at Adina or Chika building at
Gaur.
According to the ground plan the Jore Bangla mosque belongs to the group of square
shaped mosque but functionally it seems to be a Friday mosque with larger frontal shan
or courtyard. It measures externally 9.37m and the thickness of the surrounding wall is
1.42m. The prayer hall has three entrances on the eastern side; among them the central
entrance is higher and wider than the rest. The kibla wall contains three mihrab niches of
brick enclosed within rectangular frames. Among the three slender mihrab niches the
103
central one is bigger than the flanking ones. Finally the small brick pendentives make the
circular supporting area, upon which rests the dome of the hall. Plan and front elevation
of the mosque are shown in Figure 5.9 and Figure 5.10 respectively.
9.37 m
0.91m
0.9m
1.4m
1.4m
104
3.9 m
2m 2.15m 2m
The Golakata mosque (Figure 5.11) is situated on the northern side of the main road in
the westernmost part of the old city Barobazar under Kaliganj upazilla of Jhenidah
District This mosque bears no inscription, but according to its stylistic resemblance with
the architecture of early Ilyas Shahi period in Chhoto Pandua and Hazrat Pandua; it may
be assumed that this structure was built in early 15th or late 14th century A.D. prior to the
Khan Jahan's architectural development in Bagerhat. This ruined mosque was
reconstructed by the Archaeological Department of Bangladesh and thus presently seems
to be in good state of preservation.
This mosque conforms to the typical oblong enclosed mosques in Bengal built during the
early Islamic period. The interior of the prayer hall, measuring 11.83m by 8.22m
externally, is divided into two bays and three aisles surrounded by 1.4m thick brick wall.
The mosque has access only from the eastern side by three pointed entrance archways, of
which the central one is higher and wider that the flanking ones. Corresponding to the
entrance openings in the frontal wall, the kibla wall contains three highly ornamented
mihrab niches; among them the central one is bigger than the ones on each side. Two
freestanding stone pillars and eight partly concealed pilasters support the roof of six equal
hemispherical domes. The brick pendentives at each corner of a square grid transfer the
square supporting area into a circular base, upon which six hemispherical domes rest.
Plan and front elevation of the mosque are shown in Figure 5.12 and Figure 5.13
respectively.
105
Figure 5.11: Golakata mosque.
11.83 m
1m
1.1m
1.37m 1.4m
0.7m 8.22 m
0.7m
1.4m
106
4.5 m
2.3m 2.6m 2.3m
The Shait Gambuj mosque (Figure 5.14) is the most magnificent as well as the largest
enclosed type mosque in Bengal. Although the date of its erection is not known, it is
considered to be the Great Congregational mosque of Khan Jahan Ali, built during his
lifetime in the early 15th century at Bagerhat (O'Malley 1908: 27). It is commonly known
by a misleading name as Shait Gambuj mosque, or the mosque of sixty domes. As a
matter of fact it has as many as 81 domes: 70 circular domes upon the prayer hall, seven
chauchala domes upon the central aisle and four domes upon the corner towers.
The Shait Gambuj mosque is an oblong shaped prayer hall measuring externally 48.95 m
by 32.25 m with a 2.43 m thick surrounding brick wall. The hall is internally divided into
seven bays and eleven aisles. The central aisle is wider and higher than the side aisles.
The eastern wall consists of a row of eleven pointed-arch openings while each of the side
walls has seven. The central opening in each side is larger than others, which was used as
the entrance way and the rest six openings on each side were probably closed with a
perforated brick. In the northern wall, an evidence of such perforated brick jali is still in
situ. The western or the kibla wall has ten mihrab. The freestanding 60 pillars support the
whole roof. It can be possible that the name of the mosque derived not from the number
of domes (77), but from the number of the supporting pillars (60). Hence it may be stated
that the original name of this mosque was Shait (60) Khamba (local dialect for pillar)
mosque. In course of time, this Khamba (pillar) might have been confused with Gambuj
(dome). Plan and front elevation of the mosque are shown in Figure 5.15 and Figure 5.16
107
respectively.
48.95 m
1.83m
0.9m 0.9m 0.9m 0.9m 0.9m 1.7m 0.9m 0.9m 0.9m 0.9m 0.9m
1.8m
1.8m 2.44m
1.8m
1.8m 32.25 m
1.8m
4.4m
1.8m 3.4m 3.4m 3.4m 3.4m 3.4m 3.4m 3.4m 3.4m 3.4m 3.4m
0.9m
0.9m
1.8m
2.44m
1.8m 1.8m 1.8m 1.8m 1.8m 3.35m 1.8m 1.8m 1.8m 1.8m 1.8m
108
14m
7.31 m
3.31m
2.3m
The legendary Islamic hero of Bengal Khan Jahan Ali built numerous square shaped
small neighborhood mosques in addition to his large congregational mosque at Bagerhat.
This locally known nine-domed mosque belongs to the credential of this great builder
Khan Jahan Ali and was built in the middle of the fifteenth century. Noy Gambuj mosque
(Figure 5.17) is the finest example of a multi-domed square shaped plan situated at the
western side of the big Thakur Dighi near Khan Jahan Ali tomb at Bagerhat.
109
Figure 4.17: Noy Gambuj mosque.
17 m
1m
1m
0.5m
0.5 m
2.6m 17 m
2.6m
2.6m
110
5.5 m
2m 2.3m 2m
The Chunakhola mosque (Figure 5.20) is situated in the village Mogra, about one
kilometer west of the Shait Gambuj mosque in an isolated low mound in the midst of vast
111
flat cultivated field and has no road connection with the adjacent mosques. It belongs to
the group of mosques built by Khan Jahan Ali in the early 15th century (Hasan 1979:
153). This building has been restored recently by the Directorate of Archaeology. A
photo before the restoration work shows the extent of erosion due to salinity in the soil of
this area. It has eroded more of the lower part than the upper part. Directorate of
Archaeology has successfully restored the building; the lower part of the building is
almost rebuilt.
12.24 m
1m
1.1m
2.3m
It is a square shaped mosque measuring internally 7.64 m and the thickness of the
surrounding wall is 2.30 m. Similar to the other single domed mosques in Bagerhat it has
three entrances on the eastern side and two on northern and southern side. The opening on
the side walls is wider (1.50 m) than other frontal openings and has a raised sill at the
floor level, which proves that these openings were never used as entrance but as a
window for lighting and ventilation. The central entrance is elevated and wider than the
rest and is bordered within two rectangular frames with parallel terracotta details within
raised brick mouldings on the top of the frames. Among the three mihrab niches in the
kibla wall, the central one is bigger and the lateral ones are very low in height (1.25 m).
The mihrab details in this mosque are different from the rest in Bagerhat. The central
mihrab is bordered with geometric motifs and blind merlons; separated by lateral bands
of brick mouldings. The hall is roofed over by a semicircular dome, which was also
restored. Plan and front elevation of the mosque are shown in Figure 5.21 and Figure 5.22
respectively.
112
4.42 m
2m 2.6m 2m
Ronbijoypur mosque (Figure 5.23) is situated about one and a half kilometer east of the
Shait Gambuj mosque and on the northern side of the original Khan-e-Jahan road;
Khalifatabad-Bagerhat. The mosque stands on the southern embankment of the old bed of
Bhairab river facing the node of Khan-e-Jahan road from where the secondary road which
approaches the Mausoleum complex offshoots at a right angle towards south. The name
of this mosque is derived from the village called Ronbijoypur. It belongs to the group of
buildings of Khan Jahan Ali built in the middle of 15th century (Ahmed 1984:309-13).
The structure has gone under extensive restoration work for which some of the original
features may have been lost.
It measures internally 10.80 m each way and 16.40 m square externally. The surrounding
brick walls are thus 2.80 m thick and constructed with 2" x 5" x 6" bricks. The prayer hall
has access on the eastern side by three openings and each of the two walls has three
pointed-arch openings. The central opening of all the three side walls is higher and wider
than the flanking ones measuring 1.80 m wide. Corresponding to the three frontal
openings, the kibla wall is niched with three mihrabs, The huge dome over the square
prayer hall is a hemispherical one. The highest point reaches up to 39'3" from the ground
and the thickness of the dome gradually decreases at the top. The hall is covered by a
very spacious and the largest dome in Bangladesh. But unfortunately this historic ornate
113
feature of the mosque has been destroyed by plastering the dome during restoration
defacing its original character"(Banglapedia 2003: 371). Plan and front elevation of the
mosque are shown in Figure 5.24 and Figure 5.25 respectively.
16.4 m
1m
0.9m
1.2 m 2.8 m
1.2 m
2.8m
114
3.9 m
2.3m 2.3m 2.3m
The mosque (Figure 5.26) is the only surviving example in the ancient Hindu settlement
of Rampal upazilla in Munshiganj district, far away from the old capital Gaur. The Baba
Adam mosque, though renovated, is in a seemingly good state of conservation. It is
named after a locally famous saint Baba Adam, who died in a holy war, against a local
Hindu Raja. According to the local legend he was buried near the mosque in a simple
unadorned graveyard. According to an inscription fixed above the central entrance, the
mosque was built in 888A.H./l 483-84A.D. by Malik Kafur during the reign of Sultan
Jalal al-din Fath Shah (Ahmad 1960:118).
This mosque represents the six-domed verity of early Islmic period and has the usual
oblong shape measuring 6.64 m by 10.23 m internally with a 1.85 m thick surrounding
brick wall. The interior space of the mosque is split into six square areas by two bays and
three aisles. There are three entrances in the eastern facade and correspondingly, the kibla
wall contains three arched mihrab niches, the central one being larger than the ones on
each side. It is the only multi-domed mosque in Bengal without openings in the side wall.
Instead of openings, two 1.38 m wide and 0.79 m deep arched niches are placed in each
side wall. Plan and front elevation of the mosque are shown in Figure 5.27 and Figure
5.28 respectively.
115
Figure 5.26: Baba Adam mosque
13.93 m
1m
1.2m
3m 3m 3m
1.85m
0.7 m 10.34 m
0.7m
1.5 m
1.85 m
116
4.5 m
2.4m 2.4m 2.4m
From the literal meaning of the word Rajbibi (royal lady) it can be presumed that this
mosque was built at Gaur, Chapainawabganj by an influential family member of the ruler.
The time of founding cannot be ascertained as there is no inscription. According to S. M.
117
Hasan, "it may be placed between the early Ilyas Shahi period and the Husayn Shahi
period, that is, during the restored Ilyas Shahi dynasty which ruled from
841A.H./1437A.D. to 892A.H./1487A.D. (Hasan 1979: 208).
The mosque (Figure 5.29) belongs to the group of square-shaped plan with a foreroom.
The old photograph of the mosque shows that the outer surface, the foreroom and the
corner turrets had fallen down, but it was reconstructed by the Directorate of Archaeology
of Bangladesh in 1990. The mosque measures externally 17.78 m by 13.05 m and
internally the prayer hall is 8.85 m square with a 2.67 m wide foreroom. There are three
arched openings in the east facade and one at each side of the foreroom. The main hall is
accessed by three openings from the foreroom and has one opening at each of the north
and south side. Corresponding to the eastern openings there are three mihrab niches in the
kibla wall. The main prayer hall is roofed over by a semicircular dome. The foreroom is
divided into three unequal bays by two lateral arches running from the east to west wall.
The roof of the ruined foreroom is rebuilt by the restorer with three small domes. Plan
and front elevation of the mosque are shown in Figure 5.30 and Figure 5.31 respectively.
13.05 m
1m
1.2m
17.82 m
2.1m
2.1m
118
4.5 m
2.6m 2.6m 2.6m
The Dhaniachak mosque (Figure 5.32) is situated in the southern suburb of the old city
Gaur and presently in the village Shahbajpur under Shibganj upazilla of the District
Chapainawabganj. This mosque bears no inscription, but after comparing similar
mosques, such as the Shahi mosque at Basirhat (871A.H./1466A.D.), the Hathazari
mosque at Chittagong (879A.H./1474A.D.) and the Baba Adam mosque at Rampal
(888A.H./1483-84A.D.), it can be presumed that it was probably constructed in the late
fifteenth, when three-aisled mosques were popular in Bengal.
119
Figure 5.32: Dhaniachak mosque
17.28 m
1m
1.1m
1.94m
0.6m 12.66m
0.6m
1.8m
1.94m
120
3.9 m
2.7m 2.7m 2.7m
Goaldi mosque (Figure 5.35) is a very nice edifice. It was built following the style of
square shaped mosque found in Bagerhat during the regime of Khan Jahan Ali. This
small neighborhood mosque is located in the village Goaldi at Sonargaon upazilla under
Narayanganj district. According to an inscription slab which was once fixed over the
entrance door, this mosque was constructed in 925A.H./1519A.D. during the reign of
Sultan Alauddin Husayn Shah by the Mulla Hidjbr Akbar Khan (Cunnighaml882:143).
Before the restoration in 1975A.D. by the Directorate of Archaeology of Bangladesh, the
dome and major portions of the south, east and north walls had fallen down, the only
exception being the kibla on the west (Hasan 1984:186).
The Goaldi mosque has a square shaped prayer hall measuring 4.80 m square internally
with a 1.65 m wide surrounding wall. The layout of the mosque shows -a-strong deviation
by fixing the Mecca direction. The kibla wall of the mosque is 21' inclined towards south
from west. It has three entrances in the eastern wall. The central one is higher and wider
than the flanking ones. The wall on the south and north contains one arched opening.In
the kibla wall there are three mihrab niches corresponding to the entrance on the eastern
wall. The transformation of the square supporting area to a circular base is achieved,
firstly by four squinches at each corner and after that a slightly sloped pendentives to
make it a perfect circle, upon which the hemispherical dome rests. Plan and front
elevation of the mosque are shown in Figure 5.36 and Figure 5.37 respectively.
121
Figure 5.35: Goaldi mosque
0.61m
1.1m
0.9m
1.5m
122
4m
2m 2m 2m
123
The mosque (Figure 5.38) lies near the upazilla headquarter of Bagha, about 35 km south-
east of Rajshahi city. According to an inscription earlier fixed over the central entrance,
the mosque was built in 930A.H./1523-24A.D. by Sultan Nusrat Shah, (S. Ahmed 1960
:212). It is an isolated mosque standing on a high mound of earth and stands within a
compound surrounded by a low wall. It is accessed through two arched gateways from
the north and south. This mosque was seriously damaged in the earthquake of 1897 when
all the domes and the eastern wall collapsed (S. Ahmed 1960:213). The praiseworthy
restoration was undertaken in 1978 by the Directorate of Archaeology Bangladesh.
26.35 m
1m 1m 1m
0.8m 12.86 m
0.8m
1.37m
2.23m
1.1m
124
6.5m
5.5 m
2.3m 2.3m 2.3m 2.3m 2.3m
125
The Sura mosque (Figure 5.41) stands in the village Chorgacha under Ghorahat upazilla
in Dinajpur district; about six kilometers from the upazilla headquarter. Although
unadorned, the Sura mosque is dated back to the early sixteenth century in the light of its
close links with dated monuments of that time. "Recently an inscription of the time of
Alauddin Hussain Shah, dated at 910 A.H. /1504 A.D. has been discovered in the village
Champatali, a few miles away from the place. It records the construction of a mosque. If
this inscription is taken to have any connection with the mosque at Sura, the year
1504A.D. can be assumed as the date of its construction" (Banglapedia 2003: 480).
The Sura mosque belongs to the group of a square shaped room with a foreroom. It has a
4.87 m square prayer chamber flanked on the east by a 1.82 m wide foreroom and
measures externally 8.53 m by 12.18 m. Both forerooms and the prayer hall have three
entrances each on the eastern, northern and southern side. The kibla wall contains three
mihrab niches. Finally, the small brick pendentives make the circular supporting area,
upon which rests the dome of the hall. The transformed fore-room contains the brick
pendentive. It is roofed over by three small domes. Plan and front elevation of the
mosque are shown in Figure 5.42 and Figure 5.43 respectively.
8.53 m
0.76m
1.2m
12.18 m
1.83m
1.07m
1.83m
126
3.9 m
2.3m 2.3m 2.3m
This mosque (Figure 5.44) is located in the village Kherua about 1.5 km south of the
Sherpur district headquarters. A stone Persian inscription placed on the facade of the
mosque records that it was built by Mirza Murad Khan, son of Jahwar Ali Khan in the
year 989A.H. /1582A.D. (Ahmed 1960:161).
127
Figure 5.44: Kherua mosque
17.5 m
1.37m
2.01m
128
4.6 m
2.3m 2.3m 2.3m
A few number of mosques built in the outlying areas of Mughal Capital Dhaka during the
early Mughal Period reveal a happy blending of the Sultanate features with the new
Imperial Mughal features, which characterise a transitional phase in the development of
the mosque architecture in Bangladesh. It is located, about six kilometers south of Atiya
Union under Sadar Upazilla in the District Tangail. There are three inscriptional tablets in
this mosque. Two are placed over the central doorway of the eastern facade and the third
one is placed on the kibla wall. According to the original date-plaque inscription, which is
now preserved in the Bangladesh National Museum, the mosque was erected in
1019A.H./1610-11A.D. by Sayid Khan Panni, son of Bayazid Khan Panni in honour of
the great saint Pir Ali Shahansha Baba Kashmiri, who propagated Islam in this part of
Bangladesh (Banglapedia 2003: 321). His grave lies nearby.
The mosque (Figure 5.47) measures externally 16.51m by 10.52m and internally the
prayer hall is 6.71m square with a 3.55m wide fore-room. There are three arched
openings in the east facade and one on each side of the fore-room. The main hall is
accessed by three openings from the fore-room and has one opening on each of the north
and south side. Corresponding to the eastern openings there are three mihrab niches in the
kibla wall.The large central dome on the square prayer hall is supported on sequences at
each corner and the three domes over the foreroom are carried on pendentives. Plan and
front elevation of the mosque are shown in Figure 5.48 and Figure 5.49 respectively.
129
Figure 5.47: Atiya mosque
1m
1.2m
16.51 m
1.905m
0.8m
1.905m
130
6m
4.6 m
2.3m 2.3m 2.3m
131
This Shah Muhammad mosque (Figure 5.50) is located at Egarasindhur under Pakundia
upazilla of Kishoregonj district and about half a kilometer away from another famous
mughal mosque named Sadi's mosque (1652). The origin of its name is till unknown, but
probably after the name of the builder. The original inscription slab which was once fixed
over the entrance door is now missing. So its exact time of erection is not known.
However after analyzing its architectural style, many historians presumed that it may be
dated to about late 17th century.
The mosque proper is square in shape and measures externally 9.14 m each way and the
thickness of the external wall is 1.44 m. Like the other typical preMughal single domed
mosques it has also three pointed-arched entrance openings on the eastern side and a
single opening on the northern and southern side as well, flanked by a blind doorway on
either side of the central side opening. Corresponding to the three frontal openings, the
kibla wall is niched with three mihrabs. The hall is covered by a single hemispherical
dome on an octagonal shoulder and the dome being crowned with lotus and kalasa finial.
Plan and front elevation of the mosque are shown in Figure 5.51 and Figure 5.52
respectively.
9.14 m
0.91m
1.2m
1.44 m
132
6.5m
4.3 m
2.2m 2.2m 2.2m
This Mughal mosque (Figure 5.53) is a part of the Tahkhana complex, which consists of a
large residential enclave, the tomb of a holy saint Shah Niamatullah Wali and some
ruined service quarter. This complex is located about half a kilometer to the north-west of
the Chhoto Sona mosque. The building is not inscribed, but it is traditionally affirmed
that the whole tahkhana complex was built by Subandar Shah Shuja (1639-1660A.D.)
(Husain1997:117), who had his court at Rajmahal, in honour of the saint Shah
Niamatullah Wali, who lies buried in a nearby earliest known Mughal tomb structure in
Bangladesh. Recently the whole area of the tahkhana complex along with the mosque
proper has been extensively renovated by the Directorate of Archaeology of Bangladesh.
The mosque building is placed on the western end of a paved shan or courtyard, which is
encompassed by a low arcaded boundary wall with a small simple gateway in the middle
of the east side. The prayer hall has the usual oblong-shaped plan measuring 7.62m by
19.81m externally with a 1.37m thick surrounding brick wall. The prayer hall is entered
from the eastern side by three four-centered archway and the other two side walls have
also one pointed-arch openings each. Corresponding to the three frontal openings, the
kibla wall is niched with three mihrabs, The rectangular-shaped prayer room is divided
into three equal square bays by two wide transverse arches. With the help of triangular
brick pendentives at each corner, each square area is transformed into a circular
supporting area, upon which the three closely bulbous dome rest. Plan and front elevation
133
of the mosque are shown in Figure 5.54 and Figure 5.55 respectively.
0.76m
1.2m
1.37m
134
7m
4m
2.2m 2.2m 2.2m
One of the most well preserved mosque complex of Dhaka city belonging to the 17th
century is the Khawaja Shahbaz's mosque (Figure 5.56) complex. It is located to the east
of the modern mausoleum of three national leaders and to the south-east corner of the
Suhrawardy Uddyan. It is named after its builder Haji Khawaja Shahbaz, a well known
marchent of Dhaka during the reign of viceroy Muhammad Azam, whose tomb lies on
the eastern part of this complex. According to the inscription still placed over the central
doorway of the mosque, it was built in 1089A.H./1679A.D. (Banglapedia 2003:108).
The mosque is belongs to oblong shaped plan measuring 20.73m by 7.92m externally
with a 1.32m thick surrounding brick wall. Black Basalt stone was used in the plinth level
to protect the brick wall from the moisture and the salinity of the soil. Architecturally it
contains all the typical features of three domed mosque. The prayer hall is entered from
the eastern side by three archways and the other two side walls have also one arch
opening each. Corresponding to the three frontal openings, the kibla wall is niched with
three mihrabs. The rectangular shaped prayer room is divided into three equal square bays
by two wide transverse corbelled cusped arches supported by twin brick pilasters
embedded in the east and west walls, which seems to be the earliest example in
Bangladesh. With the help of half domed squinch at each corner, each square area is
transformed into a circular supporting area, upon which the dome supports.All the three
135
domes, are crowned with lotus and kalasa finial. The central one is slightly higher than
the flanking ones but the diameter is the same. Plan and front elevation of the mosque are
shown in Figure 5.57 and Figure 5.58 respectively.
20.73 m
0.76m
1.32m
1.32m
136
5.5m
4.42m
2.13m 2.13m 2.13m
Sat Gambuj mosque (Figure 5.59) is one of the well preserved seven domed mosques in
the ancient Mughal Capital Dhaka, located now in the Muhammadpur area of the city. "It
may have been built by Nawab Shaista Khan, as the local tradition remembers.
Stylistically it can be dated to the later half of the 17th century A.D." (Dani 1960: 220)
Picturesquely it was situated on the edge of a low swampy pool, which once formed a
part of the river Buriganga presently known as river Turag that now sifted its course to
about one km to the west.
The mosque building is positioned on the western end of a raised shan or courtyard
measuring 26.82m long and 25.60m wide, which is encompassed by a low boundary wall
with a gateway in the middle of the eastern side. The mosque building is oblong in plan,
measuring 8.22m wide by 17.67m long externally. The mosque is entered from the east
by three cusped arched openings, which leads directly to a niche in the kibla or western
wall. As a result the kibla wall has three arched mihrab niches, the central one being
larger than the flanking ones. The whole length of the rectanguhr hall is divided into three
unequal bays by means of two 1.05 m wide arches springing from the east and west walls.
The side bays are rectangular in shape and smaller in width, but the central one is bigger
and square. With the help of brick pendentives the square central bay is transformed into
an octagonal area and again by sequences at each octagonal corner formed a circular
137
supporting area, upon which the slightly bulbous dome supports. Plan and front elevation
of the mosque are shown in Figure 5.60 and Figure 5.61 respectively.
4m 17.67m 3.8m
0.76m
0.76m 3.75m 4m
1.3m
1.4m
1.07m 3.75m 5.1m 3.75m 8.22m
1.52m
1.37m
1.67m
1.37m 3.8m
138
1.5m
5.5m
2.9m 3m 2.9m
1.5m
The most magnificent, but unfortunately incomplete Mughal fort at Dhaka is the Lalbag
fort or fort Aurangabad, situated in the south western part of the old city once
overlooking the Buriganga River to the south. The viceroy of Mughal Bengal (1678A.D.)
Prince Muhammad Azam, son of Emperor Aurangazeb, was the originator of this fort.
Governor Shaista Khan continued the project, and after the premature death of his
daughter Bibi Pari the construction of the fort was not continued and the fort was left
unfinished (Dani 1969: 221). Among the remaining structures, there are a long fortified
wall, three gates, large tank, darbar hall, hammam, tomb of Pari Bibi, service block with a
rooftop garden and a mosque. This elegant mosque is located in the western part of the
central axis of the fort complex.
This fort mosque (Figure 5.62) is an unpretentious typical oblong shaped structure
measuring 19.5 m by 8.22 m externally with a 1.14 m thick surrounding brick wall. The
prayer hall is entered from the eastern side by three archways and the other two side walls
have also one pointed-arch openings each. The central archway is larger than the flanking
ones. Corresponding to the three frontal openings, the kibla wall is niched with three
mihrabs. The whole length of the rectangular hall is divided into three unequal bays by
means of two 1.05 m wide arches springing from the eastern and western walls. The side
bays are rectangular in shape and smaller in width, but the central one is bigger and
139
square. With the help of brick pendentives the square central bay is transformed into an
octagonal area and again by sequences at each octagonal corner formed a circular
supporting area, upon which the slightly bulbous dome supports. Plan and front elevation
of the mosque are shown in Figure 5.63 and Figure 5.64 respectively.
1.3m
1.37m
1.83m 3.35m
1.22m
140
6.75m
5.5m
3.35m
2.74m 2.74m
This mosque (Figure 5.65) is located at Atishkhana nighbourh, near Lalbag fort in Dhaka.
The mosque of Khan Muhammad Mridha was built in 1116 A.H./1706A.D., according to
an inscription placed in the frontal facade of this mosque (Hasan 1980: 59-60). The
nomenclature of the mosque derives from that of the builder Khan Muhammad Mridha,
who constructed it following the order of Quazi Ebadullah, who was probably the chief
Quazi of the then Mughal Dhaka under the Deputy Governor Farrukshah (Dani
1960:203).
The main sanctuary or prayer hall is smaller than the platform, which has also a
huzrakhana on the northern side. This mosque has the usual oblong plan, measuring
14.63m long and 7.31 m wide externally. There are three arched entrance doorways in the
eastern wall and correspondingly, the western kibla wall contains three mihrabs niches,
the central one being larger than the flanking ones. The interior is divided into three bays
by two lateral arches; each bay is covered by an onion shaped dome. Plan and front
elevation of the mosque are shown in Figure 5.66 and Figure 5.67 respectively.
141
Figure 5.65: Khan Muhammad Mridha mosque
1.1m
1.5m
142
7m
5m
2.74m 2.74m 2.74m
Just about ten kilometers from the divisional headquarter Barisal is a village named Ruiya
under the union Karapur, where the remnants of a large residential complex is found. This
complex comprises of two big tanks with associated ghat, ruins of several boundary walls
and foundation of few residential buildings. These recall the glorious past of this
settlement. This impressive two storied mosque structure is located on the north eastern
part of the complex, presently known as Miah Bari mosque (Figure 5.68). While this
edifice bears no inscription, according to the historian M. Hasan it was probably
constructed in the early eighteenth century (Hasan 1987: 23).
The main sanctuary or prayer hall is placed on the western side of the platform and has
the usual oblong plan, measuring 13.49m long and 6.1 m wide externally with a 1.65m
thick surrounding plastered brick wall. The prayer hall is entered from the eastern side by
three archway and the other two side walls have also one pointed-arch openings each.
The whole length of the rectangular hall is divided into three unequal bays by means of
two 1.05 m wide arches springing from the east and west walls. The side bays are
rectangular in shape and smaller in width, but the central one is bigger and square. With
the help of brick pendentives the square central bay is transformed into an octagonal area
and again by squinces at each octagonal corner formed a circular supporting area, upon
which the dome supports. Plan and front elevation of the mosque are shown in Figure
143
5.69 and Figure 5.70 respectively.
13.49m
1.1m
1.5m
144
6.5m
4m
2.6m 2.6m 2.6m
Perhaps the most graceful mosque complex in Bangladesh is the Bajra Shahi mosque
(Figure 5.71), popularly known by the local people as the Tajmahal of Bengal. Due to its
monumental scale and complexity it can be observed from the Dhaka-Noakhali Highway.
145
It is named after the village of Bajra under the Begumganj upazilla of Noakhali district,
on the west bank of a big derelict tank now used as a paddy field. This mosque is
seemingly in a good state of preservation due to the series of repair works and recent
refurbishing of the fa4ade. The inscription tablet dating the building to 1741A.D. is fixed
over its eastern central entrance (Banglapedia 2003:206). It was erected by the zamindar
Aman Ullah during the reign of the mughal emperor Muhammad Shah. Other Persian
inscription record that it was thoroughly repaired in the Bengali year between 1318-35 by
one Ali Ahmed and the name of the chief mason is also inscribed in Bengali which is
fixed on the outside of the southern wall.
15.77m
1.09m
1.28m
The prayed hall has the typical mughal oblong shaped plan measuring 15.77 m by 7.54 m
externally with a 1.28 m thick surrounding brick wall. The prayer hall is entered from the
eastern side by three alcove archways and the other two side walls have one pointed-arch
openings each. To articulate the main mihrab niche from outside, the kibla wall is
projected in the centre towards the west. The whole length of the rectangular hall is
divided into three unequal bays by means of two 1.06 m wide arches springing from the
east and west walls. The side bays are rectangular in shape and smaller in width, but the
central one is bigger and square. With the help of brick pendentives the square central bay
is transformed into an octagonal area. By introducing a series of sequences the octagonal
area is transformed into a circular supporting area, upon which the dome supports. The
two smaller rectangular side bays are converted into square supporting areas for the dome
by using two half domed vault springing from the eastern and western walls. Plan and
front elevation of the mosque are shown in Figure 5.72 and Figure 5.73 respectively.
146
5.5m
4.42m
2.6m 2.74m 2.6m
After having crossed the Gabtali Bridge or two kilometers on the Dhaka-Savar highway,
a traveller cannot miss on the right side of this highway a small village named Dewanbari
from which a spectacular Chini-tikri mosque sparkles in the middle, that recalls the
glorious past of the nearby merchant's house. This mosque (Figure 5.74) was built by a
leather wholesaler Hazi Janab Ali in 1880A.D. The whole compound is surrounded by a
low boundary wall on three sides and the main house on the north. Presently this
compound is accessed through a narrow street from the highway.
This mosque is oblong in plan with a fore-room. It measures externally 7.42m by 8.01m
and internally the prayer hall is 5.64m by 3.23m with a 2.11 m wide fore-room. There are
three arched openings in the east facade and one at each side of the fore-room. The main
hall is accessed by three openings from the fore-room and has one opening at each of the
north and south side. Corresponding to the eastern openings there are three mihrab niches
in the kibla wall. The main prayer hall is divided into three bays; the central one is square
and the flanking ones are rectangular in shape. The square bay is roofed over by a
semicircular dome in the centre and the side bays are roofed over by two semi-circular
half vaults springing from the northern and southern wall. The fore-room is divided into
two equal bays by three lateral arches running from the east to west wall. Two smaller
domes covering the fore-room are overlapped or interlocked with each other. Plan and
147
front elevation of the mosque are shown in Figure 5.75 and Figure 5.76 respectively.
0.46m
1.37m
1.14m
0.9m 0.9m
8.01m
0.89m
2m 2m
0.89m
0.9m 1.14m 0.9m
148
7.3m
4.3m
2.6m 2.6m 2.6m
This mosque is located at karzon hall in Dhaka Univercity.The musa khan mosque
(Figure 5.77) belongs to oblong shaped plan measuring 15.17 m by 7.54 m externally
with a 1.52 m thick surrounding brick wall. The prayer hall is entered from the eastern
side by three archways and the other two side walls have one pointed-arch openings each.
To articulate the main mihrab niche from outside, the kibla wall is projected in the centre
towards the west. The whole length of the rectangular hall is divided into three unequal
bays by means of two 1.06 m wide arches springing from the east and west walls. The
side bays are rectangular in shape and smaller in width, but the central one is bigger and
square. With the help of brick pendentives the square central bay is transformed into an
octagonal area. By introducing a series of sequences the octagonal area is transformed
into a circular supporting area, upon which the dome supports. The two smaller
rectangular side bays are converted into square supporting areas for the dome by using
two half domed vault springing from the eastern and western walls. Plan and front
elevation of the mosque are shown in Figure 5.78 and Figure 5.79 respectively.
149
Figure 5.77: Musa Khan mosque.
1.3m
1.5m
150
6.25m
5.5m
2m 2m 2m
151
The Bibir mosque is located at lalmatia in Dhaka city. The mosque (Figure 5.80) belongs
to the group of square shaped plan with a foreroom. The mosque measures externally
18.70m by 11.70m and internally the prayer hall is 8.40m square with a 4.12m wide fore-
room. There are two arched openings on the eastern side and three opening on the
northern and southern side. The main hall is accessed by three openings from the fore-
room and has one opening on each of the north and south side. Corresponding to the
eastern openings there are two mihrab niches in the kibla wall. And there is a large
central dome over the square prayer hall. Plan and front elevation of the mosque are
shown in Figure 5.81 and Figure 5.82 respectively.
18.7m
1.3m
1.7m
1.67m 1.6m
1.22m
6m
4.5m
2.6m 2.9m
152
5.29 Aambour Shah Shahi Mosque
One of the most well preserved mosque complex of Dhaka city belonging to the 17th
century is the Aambour Shah Shahi Mosque (Figure 5.83). It is located at kawran bazaar
in Dhaka City.
The mosque proper has the usual oblong shaped plan measuring 13.41m by 7.30m
externally with a 1.2m thick along east-west direction and 1.6m thick along north-south
direction brick wall. The prayer hall is entered from the eastern side by three archways
and the other two side walls have also one arch opening each. Corresponding to the three
frontal openings, the kibla wall is niched with three mihrabs. The rectangular shaped
prayer room is divided into three square bays by two wide transverse corbelled cusped
arches supported by twin brick pilasters embedded in the east and west walls. With the
help of half domed squinch at each corner, each square area is transformed into a circular
supporting area, upon which the dome supports.All the three domes, with a very low
shouldered dome on a cylindrical drum, are crowned with lotus and kalasa finial. The
central one is slightly higher than the flanking ones. Plan and front elevation of the
mosque are shown in Figure 5.84 and Figure 5.85 respectively.
153
13.41m
0.84m
1.3m
1.6m
6.75m
2.9m 5.5m
2.3m 2.3m
154
CHAPTER SIX
For the application of the simplified analysis methods, it was assumed that all the
masonry materials were similar; the volumetric weight of masonry was 18 kN/m3. The
values computed for the three in-plane indices and three out-of-plane indices are
graphically represented in Figure 6.2 to Figure 6.31, for the entire sample and for each
direction, as a function of the local parameter PGA/g (see Obtained index parameters for
the historical masonry structures in Appendix C). According to the BNBC, the seismic
map of Bangladesh is divided into three zones, i.e. Zone 1, Zone 2 and Zone 3 are shown
in Figure 3.3. PGA/g for 0.002, 0.001 and 0.0004 annual probability of exceedence are
picked and used to plot the Seismic hazard map (PGA contour) of Bangladesh for ten
percent, five percent and two percent probability of exceedence in 50-years period
respectively. The map developed using Boore et al. (1993), McGuire (1978) and Duggal
(1989) acceleration attenuation expressions are shown in Figure 27 to Figure 35 of Yasin
thesis (2008). Figure 6.1 shows using legend in Figure 6.2 to Figure 6.31.
The Chhoto Sona mosque is situated on the eastern side of a road connecting Gaur in
India and Chapainawabanj in Bangladesh. In order to the seismicity of the country, this
mosque is on Zone 1 (BNBC). For Chapainawabanj areas, the peak ground acceleration
was estimated to be 0.1g-0.275g for 0.002, 0.001 and 0.0004 annual probability of
exceedence. Six indices for Chhoto Sona mosque as a function of the PGA/g are shown in
Figures 6.2 (a) to (f). All In-plan indices present lower values in the transversal direction
(Y) than longitudinal direction (X). Index-1 (γ1) and -2 (γ2) is exceeded proposed
threshold by the mosque when the design ground acceleration above 0.15g. Index-3 (γ3)
155
and all out-of-plane indices are not exceeded it in each direction. Slenderness ratio,
thickness to height of column and perimeter wall is same in X- and Y-direction.
PGA/g for 0.002 Annual Probability of Exceedence (Boore et al., 1993) - X Direction
PGA/g for 0.002 Annual Probability of Exceedence (Boore et al., 1993) - Y Direction
Index-2 (Area to weight ratio)
PGA/g for 0.001 Annual Probability of Exceedence (Boore et al., 1993) - X Direction
6
PGA/g for 0.001 Annual Probability of Exceedence (Boore et al., 1993) - Y Direction
PGA/g for 0.0004 Annual Probability of Exceedence (Boore et al., 1993) - X Direction
PGA/g for 0.0004 Annual Probability of Exceedence (Boore et al., 1993) - Y Direction
Ghorar mosque is located in the village Barobazar under Kaliganj Upazilla of Jhenidah
district and about one kilometer west from the Jessore- Jhenidah highway. In order to the
seismicity of the country, this mosque is on low seismicity area in Bangladesh; i.e Zone 1
(BNBC). For Jhenidah area, the peak ground acceleration was estimated to be 0.075g-
0.225g for 0.002, 0.001 and 0.0004 annual probability of exceedence. Six indices for
Ghorar mosque as a function of the PGA/g are shown in Figures 6.3 (a) to (f). All In-plan
156
indices present lower values in the longitudinal direction (X) than transversal direction
(Y). The three in-plane indices are not exceeded proposed threshold by the mosque in X-
and Y-axis. The results of all out-of-plane indices are sufficient and quite similar in each
direction.
(a) 6 (b)
0.3
Index-1 (in-plan area ratio)
ratio)
0.1 2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
3
safe
ratio)
40
2 safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
w alls)
safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.2: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Chhoto Sona Mosque.
157
(a) 6 (b)
0.3
Index-1 (in-plan area ratio)
ratio)
0.1 2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
3
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
2 safe
safe ratio)
40
1
unsafe 20
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
walls)
safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.3: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Ghorar Mosque.
The Jore-Bangla mosque stands at the village Barobazar under Kaliganj upazilla in the
district Jhenidah, about six kilometers west from the upazilla headquarters. In order to the
158
seismicity of the country, this mosque is on low seismic Zone in Bangladesh; i.e Zone 1
(BNBC). For Jhenidah area, the peak ground acceleration was estimated to be 0.075g-
0.225g for 0.002, 0.001 and 0.0004 annual probability of exceedence. Six indices for
Ghorar mosque as a function of the PGA/g are shown in Figures 6.4 (a) to (f). All In-
plane indexes present lower values in the X-direction than Y-direction. The three in-plane
indices are not violated proposed threshold by the mosque in X- and Y-axis. The result of
all out-of-plane indices are also not exceeded it and quite similar in each direction (see
Figures 6.4 (d), (e), (f)).
The Golakata mosque is situated on the northern side of the main road in the westernmost
part of the old city Barobazar under Kaliganj upazilla of Jhenidah District.The three in-
plane indices and three out-of-plane indices are graphically represented in Figure 6.5 (a)
to (f) for Golakata mosque and for each direction, as a function of PGA/g are almost
similar to Jore-Bangla mosque.
The Noy Gambuj mosque is situated at the western side of the big Thakur Dighi near
Khan Jahan Ali tomb at Bagerhat in Bangladesh and on Zone 1 (BNBC). Six indexes for
this mosque as a function of the PGA/g are shown in Figures 6.7 (a) to (f). All In-plan
indices present lower values in the Y-axis than X-axis and all out-of-plane indices are
same in each axis. No indices are violated proposed threshold by this mosque except
159
Index-2 (γ2). This Index-2 is not sufficient when the design peak ground acceleration
above 0.25g.
(a) 6 (b)
0.3
Index-1 (in-plan area ratio)
ratio)
0.1 2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
In dex-3 (Base shear ratio)
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
walls)
safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.4: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Jore Bangla Mosque.
160
(a) 8 (b)
0.3
Index-1 (in-plan area ratio)
ratio)
4
0.1
unsafe 2
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness 60
3
safe safe
ratio)
2
40
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
walls)
safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.5: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Golakata Mosque.
161
0.3 (a) 4 (b)
ratio)
2
0.1
unsafe unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
3
safe
2 ratio) 40
safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
walls)
safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.6: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Shait Gambuj Mosque.
162
0.3 (a) 4 (b)
ratio)
2 safe
0.1
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
3
safe
ratio)
40
2
safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
safe
walls)
safe
0.1 0.2
unsafe
0 0 unsafe
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.7: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Noy Gamboj Mosque.
163
i.e Zone 1 (BNBC). The estimated peak ground acceleration for this mosque is similar to
Shait Gambuj and Noy Gambuj mosque. The In-plan indices and out-of-plane indices for
this mosque as a function of the PGA/g are shown in Figures 6.8 (a) to (f). All In-plan
indices present lower values in the Y-axis than X-axis and all out-of-plane indices are
same in each axis. No indices are exceeded proposed threshold by this mosque in each
axis.
It is situated about one and a half kilometer east of the Shait Gambuj mosque and on the
northern side of the original Khan-e-Jahan road; Khalifatabad-Bagerhat. In order to the
seismicity of the country, this mosque is also on low seismic Zone in Bangladesh; i.e
Zone 3 (BNBC). The estimated peak ground acceleration for this mosque is similar to
Shait Gambuj mosque, Noy Gambuj mosque and Chunakhola mosque. The In-plan
indices and out-of-plane indices for this mosque as a function of the PGA/g are shown in
Figures 6.9 (a) to (f). All In-plan indices present lower values in the Y-axis than X-axis
and all out-of-plane indices are same in each axis. No indices are exceeeded proposed
threshold by this mosque in each axis.
The mosque is located at Rampal upazilla in Munshiganj district and on the seismic Zone
2 in Bangladesh (BNBC). For Munshiganj area, the peak ground acceleration was
estimated to be 0.075g-0.225g for 0.002, 0.001 and 0.0004 annual probability of
exceedence. The In-plan indices and out-of-plane indices for this mosque as a function of
the PGA/g are shown in Figures 6.10 (a) to (f). All In-plan indices present lower values in
the Y-axis than X-axis and all out-of-plane indices are same in each axis. No indices are
exceeded proposed threshold by this mosque in each axis.
The Rajbibi mosque is situated on the western side of the tank Khania dighi at Gaur,
Chapainawabanj in Bangladesh. In order to the seismicity of the country, this mosque is
on lower seismicity area; i.e Zone 1 (BNBC). For Chapainawabanj areas, the peak ground
acceleration was estimated to be 0.1g-0.275g for 0.002, 0.001 and 0.0004 annual
probability of exceedence. Six indices for Rajbibi mosque as a function of the PGA/g are
shown in Figures 6.11 (a) to (f). All In-plan indices present lower values in the
164
longitudinal direction (X) than transversal direction (Y). Index-1 (γ1), -2 (γ2) and -3 (γ3)
are not exceeded proposed threshold by the mosque in X- and Y-axis. All out-of-plane
indices are not also violated it in each direction. Slenderness ratio, thickness to height of
column and perimeter wall is same in X- and Y-direction.
ratio )
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
3
In dex-3 (Base shear ratio)
unsafe
Index-4 (Sl enderness
60
2 safe
safe
ratio )
40
1
20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
safe
walls)
safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.8: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Chunakhola Mosque.
165
0.3 (a) 6 (b)
Index-1 (in-plan area ratio)
ratio)
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
3
ratio) safe
40
2 safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.6
0.2
0.4
w alls)
safe
safe
0.1
unsafe 0.2
0 0 unsafe
Figure 6.9: Relationship between Indices ((a) In-plane area ratio, (b) Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Ronbijoypur Mosque.
166
0.3 (a) 6 (b)
ratio)
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3
0 0.1 0.2 0.3
PGA/g PGA/g
(c) 80 (d)
3
unsafe
Index-3 (Base shear ratio)
Index-4 (Slenderness
60
2 safe
safe
ratio) 40
1
20
unsafe
0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
PGA/g PGA/g
0.2 0.4
safe
walls)
safe
0.1 0.2
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
PGA/g PGA/g
Figure 6.10: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Baba Adam Mosque.
167
0.3 (a) 6 (b)
Index-1 (in-plan area ratio)
In de x-2 (Are a to w e ig ht
0.2 4
safe
ra tio)
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
40
2
safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
safe
w alls)
safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.11: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Rajbibi Mosque.
168
6.2.11 Dhaniachak Mosque
The Dhunichak mosque is situated in the southern suburb of the old city Gaur and
presently in the village Shahbajpur under Shibganj upazilla of the District
Chapainawabganj in Bangladesh and on Zone 1 (BNBC). Six indices for Dhuniachak
mosque as a function of the PGA/g are shown in Figures 6.12 (a) to (f). All In-plan
indices present lower values in the transversal direction (Y) than longitudinal direction
(X). The three in-plan indices are not exceeded proposed threshold by the mosque in X-
and Y-axis. The results of all out-of-plane indices are sufficient and quite similar in each
direction.
The Goaldi mosque is located in the village Goaldi at Sonargaon upazilla under
Narayanganj district and on the seismic Zone 2 in Bangladesh (BNBC). For Sonargaon
area, the peak ground acceleration was estimated to be 0.075g-0.25g for 0.002, 0.001 and
0.0004 annual probability of exceedence. The In-plan indices and out-of-plane indices for
this mosque as a function of the PGA/g are shown in Figures 6.13 (a) to (f). The values of
In-plan indices are very nearest in the transversal direction (Y) and longitudinal direction
(X) and all out-of-plane indices are same in each direction. No indices are exceeded
proposed threshold by this mosque in each axis.
The Bagha mosque lies near the upazilla headquarter of Bagha, about 35 km south-east of
Rajshahi city. This mosque is on the seismic Zone 1 in Bangladesh (BNBC). For Rajshahi
area, the peak ground acceleration was estimated to be 0.115g-0.375g for 0.002, 0.001
and 0.0004 annual probability of exceedence. Six indices for this mosque as a function of
the PGA/g are shown in Figures 6.14 (a) to (f). The values of all In-plan indices are lower
in the Y-axis than X-axis and all out-of-plane indices are same in each axis.Index-1 (γ1)
and -2 (γ2) are exceeded proposed threshold by the mosque when the design ground
acceleration above 0.25g and 0.20g in Y-axis only respectively. Index-3 (γ3) and all out-
of-plane indices are not exceeded it in each axis.
169
0.3 (a) 6 (b)
ratio)
safe
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base she a r ratio)
unsafe
Index-4 (Slenderness
60
3
safe
ratio)
40
2
safe
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
safe
walls)
safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.12: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Dhaniachak Mosque.
170
0.3 (a) 6 (b)
In de x-2 (Are a to w e ig ht
0.2 4
safe
ra tio)
safe
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
3
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
2 safe safe
ratio)
40
1
20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w alls)
safe safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.13: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Goaldi Mosque.
171
0.3 (a) 6 (b)
ratio)
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
4 60 unsafe
safe
3
ratio)
40
2 safe
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
0.2 0.4
walls)
safe safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.14: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Bagha Mosque.
172
6.2.14 Sura Mosque
The Sura mosque stands in the village Chorgacha under Ghorahat upazilla in Dinajpur
district.This mosque is on the seismic Zone 2 in Bangladesh (BNBC). For Dinajpur area,
the peak ground acceleration was estimated to be 0.13g-0.425g for 0.002, 0.001 and
0.0004 annual probability of exceedence. All indices for this mosque as a function of the
PGA/g are shown in Figures 6.15 (a) to (f). Index-4 and -5 is exceeded proposed
threshold by the mosque when the peak ground acceleration above 0.4g in X- and Y-
direction. The three In-plan indices and thickness to height of perimeter walls of Sura
mosque are not exceeded it in each direction.
The Kherua mosque is located in the village Kherua at Sherpur in Bogra and on the high
seismic Zone (Zone 3) in Bangladesh (BNBC). Relationships between indexes and
PGA/g for this mosque are shown in Figures 6.16 (a) to (f). For Bogra area, the peak
ground acceleration was estimated to be 0.13g-0.45g for 0.002, 0.001 and 0.0004 annual
probability of exceedence.The values of all In-plan indices are lower in the Y-axis than
X-axis and all out-of-plane indices are quite similar in each axis. Index-1 (γ1) and -2 (γ2)
are exceeded proposed threshold by the mosque when the peak ground acceleration above
0.25g and 0.20g in Y-axis respectively. The value of base shear ratio (γ3) is greater than 1
of this Mosque (see Figure 6.16 (c)) in both directions. Slenderness ratio and thickness to
height of column of this mosque are not sufficient while PGA value above 0.40g; PGA/g
for 0.001 and 0.0004 annual probability of exceedence (Duggal, 1989) in each axis.
It is located, about six kilometers south of Atiya Union under Sadar Upazilla in the
District Tangail. In order to the seismicity, this mosque is on the seismic Zone 2 in
Bangladesh (BNBC) and estimated peak ground acceleration 0.115g-0.3g for 0.002,
0.001 and 0.0004 annual probability of exceedence. The In-plan indices and out-of-plane
indices for this mosque as a function of the PGA/g are shown in Figures 6.17 (a) to (f).
The values of In-plan indices are very nearest in the transversal direction (Y) to
longitudinal direction (X) and all out-of-plane indices are same in each direction. No
indices are exceeded proposed threshold by this mosque in each axis.
173
0.4 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
0.2 4
safe safe
0.1
unsafe 2
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
3
ratio) safe
40
2
safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
0.2 0.4
w alls)
safe safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.15: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Sura Mosque.
174
0.5 (a) 8 (b)
ratio)
safe 4
0.2 safe
0.1 unsafe 2
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
4 60 unsafe
safe
3
ratio) 40
2 safe
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
0.2 0.4
walls)
safe safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.16: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Kherua Mosque.
175
0.3 (a) 6 (b)
Index-1 (in-plan area ratio)
ratio)
safe safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
3
safe
ratio)
40
2
safe
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
walls)
safe safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.17: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Atiya Mosque.
176
6.2.17 Shah Muhammad Mosque
Shah Niamatullah Mosque is located about half a kilometer to the north-west of the
Chhoto Sona mosque at Chapainawabanj, on low seismic Zone in Bangladesh (BNBC).
In this area, the peak ground acceleration was estimated to be 0.1g-0.275g for 0.002,
0.001 and 0.0004 annual probability of exceedence. Six indices for this Mughal mosque
as a function of the PGA/g are shown in Figure 6.19 (a) to (f). All In-plane indices
present lower values in the Y-axis than X-axis. Index-1 (γ1) and -2 (γ2) is exceeded
proposed threshold by the mosque when the design ground acceleration above 0.225g in
Y-direction only. The value of base shear ratio (γ3) of this Shah Niamatullah Mosque is
sufficient (see Figure 6.19 (c)). The three out-of-plane indices are not exceeded it and
quite similar in each direction.
The Khawaja Shahbaz mosque is located to the east of the modern mausoleum of three
national leaders and to the south-east corner of the Suhrawardy Uddyan in Dhaka and on
Zone 2 in Bangladesh seismic zone (BNBC). Relationships between indices and PGA/g
for this mosque are shown in Figures 6.20 (a) to (f). For Dhaka area, the PGA value was
estimated to be 0.09g-0.225g for 0.002, 0.001 and 0.0004 annual probability of
exceedence.The values of all In-plan indices are lower in the Y-axis than X-axis, very
closed to threshold line in Y-axis and all out-of-plane indexes are quite similar in each
axis. Index-1 (γ1) and -2 (γ2) are exceeded proposed threshold by the mosque when the
peak ground acceleration above 0.20g in Y-axis only. The value of base shear ratio (γ3)
and the out-of-plane indices are not exceeded it by this mosque in each axis.
177
0.3 (a) 6 (b)
In dex-1 (in-plan area ratio)
ratio)
safe safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
(c) 80 (d)
3
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
2 safe
ratio) 40
safe
1
20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
0.2
0.2
walls)
safe safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.18: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Shah Muhammad Mosque.
178
0.3 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
safe 4 safe
0.1
unsafe 2
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
unsafe
4 Index-4 (Slenderness 60
safe
3
ratio)
safe 40
2
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w alls)
safe safe
0.1
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.19: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Shah Niamatullah Mosque.
179
0.3 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
4
safe safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
4 60 unsafe
safe
3 safe ratio) 40
2
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
walls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.20: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Khawaza Shahbaz Mosque.
180
6.2.20 Sat Gambuj Mosque
The Sat Gambuj mosque is located in the Muhammadpur area of the Dhaka city and on
Zone 2 in Bangladesh seismic zone (BNBC), the estimated PGA value 0.09g-0.225g for
0.002, 0.001 and 0.0004 annual probability of exceedence. All indices for this mosque as
a function of the PGA/g are shown in Figures 6.21 (a) to (f). In all In-plan and out-of-
plane indices, Index-2 (γ2) is violated proposed threshold by this mosque when the peak
ground acceleration above 0.20g in Y-axis only.
The Lalbag fort mosque is located in the western part of the central axis of the Lalbag fort
complex in Dhaka City.In order to the seismicity of the country, this mosque is on
seismic Zone 2 in Bangladesh (BNBC). ). The estimated peak ground acceleration for this
mosque is similar to Shait Gambuj and Sat Gambuj mosque. Six indices for this mosque
as a function of the PGA/g are shown in Figures 6.22 (a) to (f). The values of all In-plan
indices are lower in the Y-axis than X-axis and closed to threshold line in Y-axis. Index-1
(γ1) and -2 (γ2) are exceeded proposed threshold by this mosque when the design PGA
value above 0.175g in the Y-axis only. All out-of-plane indices are not exceeded it in
each direction and Slenderness ratio and thickness to height of column are quite similar in
X- and Y-direction (see Figures 6.20 (d) to (f).
This mosque is located at Atishkhana nighbourh, near Lalbag fort in Dhaka.The In-plan
indices and out-of-plane indices for this mosque as a function of the PGA/g are shown in
Figures 6.23 (a) to (f). The values of index-4 and -5 are quite similar in each direction. No
indices are exceeded proposed threshold by this mosque in each axis.
The Miah Bari mosque is situated about ten kilometers from the divisional headquarter
Barisal is a village named Ruiya under the union Karapur, on the low seismic Zone in
Bangladesh (BNBC). In this area, the peak ground acceleration was estimated to be
0.09g-0.25g for 0.002, 0.001 and 0.0004 annual probability of exceedence. Six indices for
this mosque as a function of the PGA/g are shown in Figure 6.24 (a) to (f). The results of
all indices of this are not exceeded proposed threshold.
181
0.4 (a) 8 (b)
In dex-1 (in-plan area ratio)
ratio)
0.2 4
safe safe
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
4 60 unsafe
safe
3 ratio)
safe 40
2
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w alls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.21: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Sat Gambuj Mosque.
182
0.3 (a) 6 (b)
ratio)
safe safe
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
4
safe
3 safe ratio) 40
2
20
1 unsafe
0 0
0 0.1 0.2 0.3 0 .4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
walls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.22: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Lalbag Fort Mosque.
183
0.4 (a) 6 (b)
Index-1 (in-plan area ratio)
In d e x-2 (A re a to w e ig h t
0.3
4
safe
ra tio )
0.2 safe
2
0.1
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
In d e x-4 (Sle n d e rn e ss
60 unsafe
3
ra tio ) safe
safe 40
2
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w a lls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.23: Relationship between Indices ((a) In-plane area ratio, (b)Area to weight
ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f) t/h of
perimeter wall) and PGA/g for Khan Muhammad Mridha Mosque.
184
0.4 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
0.2 4 safe
safe
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
5
Index-3 (Base shear ratio)
Index-4 (Slenderness
4 60 unsafe
safe
3 ratio) 40
safe
2
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
walls)
safe safe
0.1 0.2
unsafe
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.24: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Miah Bari Mosque.
185
6.2.24 Bazra Mosque
The Bajra mosque is named after the village of Bajra under the Begumganj upazilla of
Noakhali district. According to seismicity of Bangladesh, it is on the low seismic Zone
(BNBC). In this area, the peak ground acceleration was estimated to be 0.09g-0.175g for
0.002, 0.001 and 0.0004 annual probability of exceedence. The In-plan indices and out-
of-plane indices for this mosque as a function of the PGA/g are shown in Figures 6.25 (a)
to (f). The values of all out-of-plane indices are quite similar in X and Y direction. No
indices are exceeded proposed threshold by this mosque in each axis
The Dewanbari mosque is situated after having crossed the Gabtali Bridge or two
kilometers on the Dhaka-Savar highway, at aminbazar in Dhaka and on Zone 2 in
Bangladesh seismic zone (BNBC), the estimated PGA value 0.09g-0.225g for 0.002,
0.001 and 0.0004 annual probability of exceedence. All indices for this mosque as a
function of the PGA/g are shown in Figures 6.26 (a) to (f). The values of three In-plan
and three out-of-plane indices of the Dewanbari mosque are almost similar in X and Y
direction, not exceeded proposed threshold by this mosque.
This mosque is located at karzon hall in Dhaka University. The In-plan indices and out-
of-plane indices for this mosque as a function of the PGA/g are shown in Figures 6.27 (a)
to (f). The values of index-4 and -5 of the Musa khan mosque are quite similar in X and Y
direction. No indices are exceeded proposed threshold by this mosque in each direction.
The Bibir mosque is located at lalmatia in Dhaka city. All indices for this mosque as a
function of the PGA/g are shown in Figures 6.28 (a) to (f). All indices are satisfied
proposed threshold by this mosque in X and Y direction.
The Aambour Shah Shahi mosque is located at kawran bazar in Dhaka city. All indices
for this mosque as a function of the PGA/g are shown in Figures 6.29 (a) to (f). And data
processing of six indices for Aambour Shah Shahi mosque is shown in Appendix D. No
indices are exceeded proposed threshold by this mosque in X and Y direction.
186
0.3 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
4 safe
safe
0.1
2
unsafe
unsafe
0
0
0 0.1 0.2 0.3
0 0.1 0.2 0.3
PGA/g PGA/g
(c) 80 (d)
4
unsafe
Index-3 (Base shear ratio)
Index-4 (Slenderness
60
3
ratio) safe
safe 40
2
1 20
unsafe
0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
PGA/g PGA/g
0.2
0.2
wal ls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
PGA/g PGA/g
Figure 6.25: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Bazra Mosque.
187
0.3 (a) 6 (b)
ratio)
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
3
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
2 safe
safe ratio) 40
1
20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
safe 0.2
walls)
safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.26: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Dewanbari Mosque.
188
0.4 (a) 6 (b)
Index-1 (in-plan area ratio)
ratio)
0.2
2
0.1
unsafe unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
3
ratio) safe
safe 40
2
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w alls)
safe safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.27: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Musa Khan Mosque.
189
0.3 (a) 8 (b)
Index-1 (in-plan area ratio)
ratio)
4 safe
safe
0.1
unsafe 2
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
unsafe
Index-4 (Slenderness
60
3
safe
ratio) 40
2 safe
20
1
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2 0.4
walls)
safe safe
0.1 0.2
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.28: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Bibir Mosque.
190
0.4 (a) 8 (b)
Index-1 (in-plan area ratio)
safe
ratio)
0.2 4
safe
0.1 2
unsafe
unsafe
0
0
0 0.1 0.2 0.3 0.4
0 0.1 0.2 0.3 0.4
PGA/g PGA/g
(c) 80 (d)
4
Index-3 (Base shear ratio)
Index-4 (Slenderness
60 unsafe
3
ratio) safe
safe 40
2
1 20
unsafe
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
0.2
0.2
w alls)
safe
safe
0.1
unsafe
0 unsafe
0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
PGA/g PGA/g
Figure 6.29: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g for Aambour Shah Shahi Mosque.
191
6.2.29 Comparative Analysis
BNBC
Six indices for all masonry structures individually as a function of the PGA/g (BNBC) are
shown in Figure 6.2 to Figure 6.29. Figure 6.30 and Figure 6.31 present six indices all
masonry structures along the X-axis (NS) and Y-axis (EW) respectively. In terms of
average values, index-1 (γ1) presents lower values in the transversal direction (Y) than
longitudinal direction (X). But when a comparison is made using the proposed threshold,
No mosques exceed it in the X- and Y-direction (see Figure 6.30 (a) and Figure 5.31 (a)).
Index-2 (γ2), although being inversely proportional to the height of the buildings, presents
a situation similar to index-1. On average terms, index-2 (γ2) also presents lower values
in the y-direction, which can be justified by mosques geometry. As a result, this index is
exceeded by 0% and 3.6% of the monuments in X- and Y- directions, respectively (see
Figure 6.30 (b) and Figure 6.31 (b)). This index is mainly exceeded kherua mosque in y
direction only (see Figure 6.16 (b)). As it happened with both previous indices, index-3
(γ3) also presents lower values in the Y-direction. Individually, No mosques in X- and Y-
directions exceed it (see Figure 6.30 (c) and Figure 6.31 (c)). In out-of- plane indices can
be observed that maxima index-4 values (column’s slenderness) tend to decrease with
increasing seismicity and that both minima index-5 and index-6 values seem to increase
continuously with seismicity. These trends are depicted in Figure 6.2 (d), (e), (f) to Figure
6.31 (d), (e), (f) by threshold lines. Their general evolution with increasing seismicity was
expected since, for the same safety level, index-4 should decrease and both index-5 and
index-6 should increase. The values of out-of- plane indices are sufficient in the X- and
Y-directions (see Figure 6.30 (d), (e), (f) and Figure 6.31 (d), (e), (f)) for all mosques.
And all out-of- plane indices are almost same in X- and Y-directions for thier similar
dimension.
0.002, 0.001 and 0.0004 Annual Probability of Exceedence (Boore et al., 1993)
PGA for 0.002, 0.001 and 0.0004 annual probability of exceedence are picked and used to
plot the Seismic hazard map (PGA contour) of Bangladesh for ten, five and two percent
probability of exceedence in 50-years period respectively (Yasin, 2008). The PGA
contour developed using Boore et al. (1993) acceleration attenuation expression. Six
indices for all masonry structures individually as a function of the PGA/g for 0.002
annual probability of exceedence (Boore et al., 1993) are shown in Figure 6.2 to Figure
6.29. Figure 6.30 and Figure 6.31 present six indices all masonry structures as also
192
function of the PGA/g for 0.002 annual probability of exceedence (Boore et al., 1993)
along the X-axis (NS) and Y-axis (EW) respectively. In terms of average value of in-
plane indices, No mosques exceed proposed threshold in both X- and Y-directions (see
Figure 6.30 (a), (b), (c) and Figure 6.31 (a), (b), (c)). All out-of- plane indices show no
mosques violate it in the X- and Y-directions (see Figure 6.30 (d), (e), (f) and Figure 6.31
(d), (e), (f)).
In terms of average values, the three in-plane indices and three out-of-plane indices are
graphically represented in Figure 6.2 to Figure 6.31, for all mosques and for each
direction, as a function of PGA/g for 0.001 and 0.0004 annual probability of exceedence
(Boore et al., 1993) are almost similar to PGA/g for 0.002 annual probability of
exceedence (Boore et al., 1993) except Chhoto sona mosque. It is exceeded Index-2 (γ2)
threshold in only Y-direction for 0.0004 annual probability of exceedence (Boore et al.,
1993) (see Figure 6.2 (a)).
PGA for 0.002, 0.001 and 0.0004 annual probability of exceedence, Seismic hazard map
(PGA contour) of Bangladesh for ten, five and two percent probability of exceedence in
50-years period are developed using McGuire (1978) acceleration attenuation expression
(Yasin, 2008). Figure 6.2 to Figure 6.29 present six indices all masonry structures
individually as a function of the PGA/g for 0.002 annual probability of exceedence
(McGuire, 1978). Figure 6.30 and Figure 6.31 present six indices all masonry structures
as also function of the PGA/g for 0.002 annual probability of exceedence (McGuire,
1978) along the X-axis and Y-axis respectively. In terms of average values, the six
indices for all mosques and each direction, as a function of PGA/g for 0.002 annual
probability of exceedence (McGuire, 1978) are almost similar to PGA/g for 0.002 annual
probability of exceedence (Boore et al., 1993).
Six indices all masonry structures individually as a function of the PGA/g for 0.001 and
0.0004 annual probability of exceedence (McGuire 1978) are shown in Figure 6.2 to
Figure 6.29. Figure 6.30 and Figure 6.31 present six indices all masonry structures as also
function of the PGA/g for 0.001 and 0.0004 annual probability of exceedence (McGuire,
1978) along the X-axis and Y-axis respectively. In terms of average values, Index-1 (γ1)
is not violated any monuments in both X- and Y-directions (see Figure 6.30 (a) and
Figure 6.31 (a)). Index-2 (γ2) is exceeded by 0% and 7.15% of the monuments in X- and
Y- directions, respectively (see Figure 6.30 (b) and Figure 6.31 (b)). This index is mainly
193
exceeded by Chhoto sona mosque (see Figure 6.2 (b)) and Kherua mosque (see Figure
6.16 (b)) in Y direction only. The base shear ratio (γ3 of all mosques in X- and Y-
directions are sufficient (see Figure 6.30 (c) and Figure 6.31 (c)). In three out-of- plane
indices, no mosques are exceeded threshold in each direction (see Figure 6.30 (d), (e), (f)
and Figure 6.31 (d), (e), (f)). Zone 1
Zone 2
Zone 3
Zon e 1
Zon e 2
Zon e 3
0.5 (a) (b)
8
0.4
6
0.3
0.2 4
Zone 2
Zone 3
Z one 1
Z one 2
Z one 3
(c) 80 (d)
5
Index-4 (Slenderness
60
Index-3 (Base shear ratio)
unsafe
4
safe
ratio)
3 safe 40
2
20
1
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Zone 1
Zone 2
Zone 3
Zone 1
Zone 2
Zone 3
0.6
0.2
0.4
walls)
0 unsafe
0
0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.30: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g along X-axes for all structures.
194
Zone 1
Zone 2
Zone 3
Z one 1
Z one 2
Z one 3
0.4 (a) 8 (b)
0.2 4
0.1 2 safe
safe
unsafe unsafe
0
0
0.0 0.1 0.2 0.3 0.4 0.5
0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Zone 1
Zone 2
Zone 3
(c) 80 (d)
Z one 1
Z one 2
Z one 3
4
In d e x-4 (Sle n d e rn e ss
Index-3 (Base shear ratio)
60 unsafe
3
safe
ra tio )
safe 40
2
20
1
unsafe
0
0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Zone 1
Zone 2
Zone 3
Zone 1
Zone 2
Zone 3
0.6
0.2
walls)
0.4
0 unsafe
0
0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
PGA/g PGA/g
Figure 6.31: Relationship between Indices ((a) In-plane area ratio, (b) Area to
weight ratio, (c) Base shear ratio, (d) Slenderness ratio, (e) t/h of columns, (f)
t/h of perimeter wall) and PGA/g along Y-axes for all structures.
PGA for 0.002, 0.001 and 0.0004 annual probability of exceedence, Seismic hazard map
(PGA contour) of Bangladesh for ten, five and two percent probability of exceedence in
50-years period are developed using Duggal (1989) acceleration attenuation expression
195
(Yasin, 2008). All in-plane and out-of-plane indices as a function of the PGA/g for 0.002
annual probability of exceedence (Duggal, 1989) are presented Figure in 6.2 to Figure
6.31. In terms of average values, index-1 (γ1) presents lower values in the Y-axes than X
axes. In plan area ratio (γ1) is exceeded by 0% and 10.71% of the mosques proposed
threshold in X- and Y-directions respectively (see Figure 6.30 (a) and Figure 6.31 (a)).
This index is mainly exceeded by Shait Gambuj mosque (see Figure 6.6 (a)), Bagha
mosque (see Figure 6.14 (a)) and Kherua mosque (see Figure 6.16 (a)) in Y direction
only. Index-2 (γ2) is exceeded by 0% and 17.86% of the mosques in X- and Y-
directions, respectively (see Figure 6.30 (b) and Figure 6.31 (b)). This index is mainly
exceeded Chhoto sona mosque (see Figure 6.2 (b)), Shait Gambuj mosque (see Figure 6.6
(b)), Bagha mosque (see Figure 6.14 (b)), Kherua mosque (see Figure 6.16 (b)) and
Lalbag fort mosque (see Figure 6.22 (b)) in Y direction only. No mosques are exceeded
Index-3 threshold line. The values of all out-of- plane indices are sufficient in the X- and
Y-directions (see Figure 6.30 (d), (e), (f) and Figure 6.31 (d), (e), (f)).
In terms of average values, the three in-plane indices and three out-of-plane indices are
graphically presented in Figure 6.2 to Figure 6.31 for all masonry structures as a function
of the PGA/g for 0.001 annual probability of exceedence (Duggal, 1989). In plan area
ratio (γ1) is exceeded by 0% and 21.4% of the mosques proposed threshold in X- and Y-
directions respectively (see Figure 6.30 (a) and Figure 6.31 (a)). This index is mainly
exceeded by Chhoto sona mosque (see Figure 6.2 (a)), Shait Gambuj mosque (see Figure
6.6 (a)), Bagha mosque (see Figure 6.14 (a)), Kherua mosque (see Figure 6.16 (a)), Shah
Niamatulla mosque (see Figure 6.19 (a)) and Lalbag fort mosque (see Figure 6.22 (a)) in
Y direction only. Index-2 (γ2) is also exceeded it by 0% and 21.4 % of the mosques in X-
and Y- directions, respectively (see Figure 6.30 (b) and Figure 6.31 (b)). This index is
mainly exceeded by Chhoto sona mosque (see Figure 6.2 (b)), Shait Gambuj mosque (see
Figure 6.6 (b)), Noy Gambuj mosque (see Figure 6.7 (b)), Bagha mosque (see Figure 6.14
(b)), Kherua mosque (see Figure 6.16 (b)) and Lalbag fort mosque (see Figure 6.22 (b)) in
Y direction only. Index-3 (γ3) and All out-of- plane indices are not exceeded proposed
threshold in the X- and Y-directions (see Figure 6.30 (c) (d), (e), (f) and Figure 6.31 (c)
(d), (e), (f)) except Kherua mosque and Sura mosque. Slenderness ratio of Kherua
mosque and thickness to height of column of Sura mosque are not sufficient.
The in-plane indices and out-of-plane indexes are presented in Figure 6.2 to Figure 6.31
for all masonry structures as a function of the PGA/g for 0.0004 annual probability of
196
exceedence (Duggal, 1989). In plan area ratio (γ1) is Similar to 0.001 annual probability
of exceedence here. Index-2 (γ2) is exceeded threshold by 7.14% and 28.58% of the
mosques in X- and Y- directions, respectively (see Figure 6.30 (b) and Figure 6.31 (b)).
This index is mainly exceeded by Chhoto sona mosque (see Figure 6.2 (b)) and Shait
Gambuj mosque (see Figure 6.6 (b)) in both direction, and Noy Gambuj mosque (see
Figure 6.7 (b)), Bagha mosque (see Figure 6.14 (b)), Kherua mosque (see Figure 6.16 (b))
Shah Niamatulla mosque (see Figure 6.19 (b)), Khawaja shahbaz mosque (see Figure
6.20 (b)), and Lalbag fort mosque (see Figure 6.22 (b)) in Y direction only. Index-3 (γ3)
is almost similar to all PGA value. All out-of- plane indices are not exceeded proposed
threshold in the X- and Y-directions (see Figure 6.30 (d), (e), (f) and Figure 6.31 (d), (e),
(f)) except Kherua mosque and Sura mosque. Slenderness ratio and thickness to height of
column of Kherua mosque and Sura mosque are not sufficient.
The structural vibrations of the historical mosques are measured using Microtremor
equipment on the roof. For the two locations described earlier in this paper, both vertical
and horizontal measurements were recorded, with a sampling rate of 100Hz at Musa
Khan Mosque, Karzan Hall, Dhaka university and Aambour Shah Shahi Mosque, Kawran
bazar, Dhaka, and a duration of 10 minutes for each. The data processing was done with
Origin software. Fourier Spectrum are shown in Figure 6.32 and 6.33 (X, Y and Z
represent North-South direction, East-West direction and Up-Down direction
respectively), the pertinent information obtained is presented in Table 6.1.
197
2.0
X
1.8
Y
1.6 Z
1.4
Amplitude
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0 10 20 30
Frequency
The Musa khan mosque is located at karzon hall in Dhaka University. In this mosque, we
found resonance frequency (Structure and soil frequencies are 3.76 Hz and 3.2 Hz
respectively), Fourier spectrum in North-South and East-West direction are similar
(Figure 6.32).
2.0
X
Y
Z
1.5
Amplitude
1.0
0.5
0.0
0 10 20 30
Frequency
The Aambour Shah Shahi mosque is located at kawran bazar in Dhaka city. In this
mosque, Building and soil frequencies are not nearest, 7.66 Hz and 3.06 Hz respectively.
So it is not resonance frequency.
198
It was evident from the results of the microtremor tests that the site frequency for the
Dhaka district is in the range of 7.66 Hz. This raises the possibility of soil-structure
interaction and should be addressed for retrofit design considerations. The dynamic
response of the buildings during a severe earthquake can be significantly affected by soil-
structure interaction effects. This can also be of extra significance for older unreinforced
buildings such as those in this study.
199
CHAPTER SEVEN
7.1 General
The investigation presented in this thesis includes the application of the simplified
methods previously described to a sample of twenty eight monuments; mainly historical
mosques (4 from Chapainawabnganj, 3 from Jhenidah, 8 from Dhaka, 4 from Bagerhat
and 1 from Munshiganj, Sonargaon, Rajshahi, Dinajpur, Bogra, Tangail, Kishoreganj,
Barisal and Noakhali each district) selected according to the seismic level and to the
availability of information. Moreover, Musa khan mosque and Aambour Shah Shahi
mosque in Dhaka city were performed by microtremor measurement. This section
summarizes the major findings and conclusions obtained from the research. Furthermore,
some recommendations based on the observations of the research are discussed.
This thesis presents an investigation about the possibility of using simplified methods of
analysis and simple indices as indicators for fast screening and decision to prioritize
deeper studies in historical masonry buildings and assess vulnerability to seismic actions.
Eventhough, Lourenco and Oliveira used thresholds as functions of PGA/g to analyze
historical masonry structures (Portugal, Spain, and Italian churches) by this method for
evaluating seismic vulnerability of churches (Lourenco and Oliveira, 2004), In this thesis,
it is assumed that these threshold values to be equally applicable for identifing the
vulnerable historical masonry structures of Bangladesh. These indices are based mostly
on the in plan dimensions and height of the buildings. The simplified methods indicate
that, in Bangladesh, the average in-plan area of earthquake resistant walls and average
height are independent of the seismicity. This puzzling feature can be related to the short
memory of the ancient builders and the fact that major earthquakes in Bangladesh have
rather long return periods (over 200 years).
The usage of simplified methods was made, taking into consideration the in plan area of
the building, its height and seismicity, with the simultaneous verification of two indices,
one related to ratio of in plan area (γ1) and weight (γ2), and another related to the
maximum base shear force (γ3). The analysis of the out-of-plane indices shows that a
logical common trend can be established.
200
In general, the longitudinal direction of the buildings (x) exhibits much lower
vulnerability than the transversal direction (y). Indices (γ1) and (γ2) do not present a clear
trend with respect to seismicity; however, a slight trend observed associates the increase
of (γ1) and (γ2) with PGA growth. In terms of average values, Kherua mosque, Shah
niamatulla mosque Khawaja shahbaz mosque, Chhoto sona mosque, Shait Gambuj
mosque, Bagha mosque and Lalbag fort mosque are exceeded Index-1 and -2 proposed
thresholds in Y-direction when the design peak ground acceleration above 0.15g. For the
higher seismicity (design ground acceleration above 0.25 g), Noy Gambuj mosque is
exceeded Index-2 threshold line in also Y-direction only. About 71.42% of sample, No
in-plan indices exceed it in both direction. The base shear ratio (γ3) and the value of three
out-of- plane indices are sufficient almost all mosques in both direction. The results show
that almost all mosques are safe except eight (28.58%) mosques of the samples. For the
high seismicity, these eight mosques are vulnerable in the short or Y-direction only and
required careful attention and deeper investigation at risk.
7.4 Recommendations
201
REFERENCES
Ahmed, A. S. M. The Choto Sona mosque at Gaur: an example of the early Islamic
Architecture of Bengal, Karlsruhe-Germany, 1997
Ahmed, N, 1989. The Buildings of Khan Jahan in and around Bagerhat, Dhaka.
Aki, K. (1957). Space and time spectra of stationary stochastic waves, with special
reference to microtremors, Bull. of earthquake research institute, 35, 415-456.
Annual report of the Archaeological Survey of India, Eastern circle, Calcutta 1908-09
Ansary, M.A., Noor, M.A., 2004. Vulnerability Assessment of Existing Engineered And
Nonengineered Structures Of Dhaka City Using RVS And NDT Techniques,
October 2004, Agra, India.
Ansary, M.A. and Noor, M.A., 2006. Seismic Vulnerability of Historical Mosques of
Bangladesh.
Anthoine A., Magonette G., Magenes G.1994. “Shear-compression testing and analysis of
brick masonry walls”. Proceedings of the Tenth European Conference on
Earthquake Engineering, Vienna, Austria.
ATC (1985). “Earthquake Damage Evaluation Data for California,” Report ATC-13,
Applied Technology Council, Redwood City, California, U.S.A.
ATC (1988). “Rapid Visual Screening of Buildings for Potential Seismic Hazards: A
Handbook”, Report ATC-21, Applied Technology Council, Redwood City,
California, U.S.A.
ATC (1996). “Seismic Evaluation and Retrofit of Concrete Buildings”, Report ATC-40,
Applied Technology Council, Redwood City, California, U.S.A.
202
Banglapedia National Encyclopedia of Bangladesh, Asiatic Society of Bangladesh,
2003, Dhaka
Barbat, A.H., Yépez Moya, F. and Canas, J.A. (1996). “Damage Scenarios Simulation for
Seismic Risk Assessment in Urban Zones”, Earthquake Spectra, Vol. 12, No. 3,
pp. 371-394.
Bazzurro, P., Cornell, C.A., Menun, C., Luco, N., Motahari, M., 2004. Advanced Seismic
Assessment Guidlines. Stanford University, Stanford CA, Sep. 30.
Benjamin, J.R., Cornell, C.A., 1970. Probability, Statistics, and Decision for Civil
Engineers. Mc-Graw Hill, Inc., New York.
Benjamin, J.R. and Associates (1988). A Criteria for Determining Exceedance of The
Operating Basis Earthquake, EPRI Report NP-5930, Electric Power Research
Institute, Palo Alto, California.
Bird, J.F. and Bommer, J.J. (2004). “Earthquake Losses due to Ground Failure”,
Engineering Geology, Vol. 75, No. 2, pp. 147-179.
203
Bommer, J.J., Spence, R., Erdik, M., Tabuchi, S., Aydinoglu, N., Booth, E., del Re, D.
and Peterken, O. (2002). “Development of an Earthquake Loss Model for Turkish
Catastrophe Insurance”, Journal of Seismology, Vol. 6, No. 3, pp. 431-446.
Bommer, J.J., Spence, R. and Pinho, R. (2006). “Earthquake Loss Estimation Models:
Time to Open the Black Boxes?”, Proceedings of the First European Conference
on Earthquake Engineering and Seismology, Geneva, Switzerland, Paper No. 834.
Boore, D.M., Joyner, W.B., Fumal, T.E., 1993. Estimation of response spectra and peak
accelerations from western North America earthquakes: An interim report, Open-
File-Report 93-509, U.S. Geological Survey, Reston, Virginia, 72 pp.
Calvi, G.M. and Pinho, R. (2004). “LESSLOSS. A European Integrated Project on Risk
Mitigation for Earthquakes and Landslides”, Report 2004/02, European School for
Advanced Studies in Reduction of Seismic Risk (ROSE School), Pavia, Italy.
Cardona, O.D. and Yamin, L.E. (1997). “Seismic Microzonation and Estimation of
Earthquake Loss Scenarios: Integrated Risk Mitigation Project of Bogotà,
Colombia”, Earthquake Spectra, Vol. 13, No. 4, pp. 795-814.
204
CEN, 2003. Eurocode 6: Design of masonry structures, prEN 1996-1, Brussels.
Cosenza, E., Manfredi, G., Polese, M. and Verderame, G.M. (2005). “A Multi-Level
Approach to the Capacity Assessment of Existing RC Buildings”, Journal of
Earthquake Engineering, Vol. 9, No. 1, pp. 1-22.
Crowley, H. and Bommer, J.J. (2006). “Modelling Seismic Hazard in Earthquake Loss
Assessment with Spatially Distributed Exposure”, Bulletin of Earthquake
Engineering, Vol. 4, No. 3, pp. 249-273.
Crowley, H. and Pinho, R. (2006). “Simplified Equations for Estimating the Period of
Vibration of Existing Buildings”, Proceedings of the First European Conference
on Earthquake Engineering and Seismology, Geneva, Switzerland, Paper No.
1122.
Crowley, H., Pinho, R., Bommer, J.J. and Bird, J.F. (2006). “Development of a
Displacement-Based Method for Earthquake Loss Assessment”, Report 2006/01,
European School for Advanced Studies in Reduction of Seismic Risk (ROSE
School), Pavia, Italy.
205
D’Ayala, D. and Speranza, E. (2002). “An Integrated Procedure for the Assessment of
Seismic Vulnerability of Historic Buildings”, Proceedings of the 12th European
Conference on Earthquake Engineering, London, U.K., Paper No. 561.
Di Pasquale, G., Orsini, G. and Romeo, R.W. (2005). “New Developments in Seismic
Risk Assessment in Italy”, Bulletin of Earthquake Engineering, Vol. 3, No. 1, pp.
101-128.
Dolce, M., Masi, A., Marino, M. and Vona, M. (2003). “Earthquake Damage Scenarios of
the Building Stock of Potenza (Southern Italy) Including Site Effects”, Bulletin of
Earthquake Engineering, Vol. 1, No. 1, pp. 115-140.
Duggal, R., 1989. Estimation of Seismic Risk and Damage, Their Utilization as Design
Criteria. M Engg. Thesis, University of Tokyo, Japan.
Faccioli, E., Pessina, V., Calvi, G.M. and Borzi, B. (1999). “A Study on Damage
Scenarios for Residential Buildings in Catania City”, Journal of Seismology, Vol.
3, No. 3, pp. 327-343.
Fajfar P., 1998. “Capacity spectrum method based on inelastic demand spectra”. IKPIR
Report EE - 3/98, Institute of Structural Engineering, Earthquake Engineering and
construction IT, University of Ljubljana.
Fah, D., Kind, F., Lang, K. and Giardini, D. (2001). “Earthquake Scenarios for the City of
Basel”, Soil Dynamics and Earthquake Engineering, Vol. 21, No. 5, pp. 405-413.
FEMA (2001). “HAZUS99 Estimated Annualized Earthquake Loss for the United
States”, Report FEMA 366, Federal Emergency Management Agency,
Washington, DC, U.S.A.
206
FEMA (2003). “HAZUS-MH Technical Manual”, Federal Emergency Management
Agency, Washington, DC, U.S.A.
Giuffrè, A. 1995. Seismic damage in historic town centers and attenuation criteria, Annali
di Geofisica, 38(5–6), 837–843.
GNDT (2000). “The Catania Project: Earthquake Damage Scenarios for a High Risk Area
in the Mediterranean”, Proceedings of CNR-Gruppo Nazionale per la Difesa dai
Terremoti, Roma, Italy.
Griffith, M.C., Magenes, G., Melis, G. and Picchi, L. (2003). “Evaluation of Out-of-Plane
Stability of Unreinforced Masonry Walls Subjected to Seismic Excitation”,
Journal of Earthquake Engineering, Vol. 7, No. 1, pp. 141-169.
207
Guagenti, E. and Petrini, V. (1989). “The Case of Old Buildings: Towards a Damage-
Intensity Relationship”, Proceedings of the Fourth Italian National Conference on
Earthquake Engineering, Milan, Italy, pp. 145-153 (in Italian).
Hassan, A.F. and Sozen, M.A. (1997). “Seismic Vulnerability Assessment of Low-Rise
Buildings in Regions with Infrequent Earthquakes”, ACI Structural Journal, Vol.
94, No. 1, pp. 31-39.
Hasan, P, 1984. Eight Sultanate Mosques in Dhaka District, in the Islamic Heritage of
Bengal, ed. by G. Michell (1984), pp. 179-92.
Kappos, A.J., Pitilakis, K. and Stylianidis, K.C. (1995). “Cost-Benefit Analysis for the
Seismic Rehabilitation of Buildings in Thessaloniki, Based on a Hybrid Method of
Vulnerability Assessment”, Proceedings of the Fifth International Conference on
Seismic Zonation, Nice, France, Vol. 1, pp. 406-413.
Kennedy, R.P. (1980). Ground Motion Parameters Useful in Structural Design, presented
at the Conference on Evaluation of Regional Seismic Hazards and Risk, Santa Fe,
New Mexico.
208
Lantada, N., Pujades, L.G. and Barbat, A.H. (2004). “Risk Scenarios for Barcelona,
Spain”, Proceedings of the 13th World Conference on Earthquake Engineering,
Vancouver, Canada, Paper No. 423 (on CD).
McCormack, T.C. and Rad, F.N. (1997). “An Earthquake Loss Estimation Methodology
for Buildings Based on ATC-13 and ATC-21”, Earthquake Spectra, Vol. 13, No.
4, pp. 605-621.
Modena, C., Lourenço, P.B. and Roca, P. (editors) (2005). “Structural Analysis of
Historical Constructions – Possibilities of Numerical and Experimental
Techniques”, Taylor and Francis, London, U.K.
Molina, S. and Lindholm, C. (2005). “A Logic Tree Extension of the Capacity Spectrum
Method Developed to Estimate Seismic Risk in Oslo, Norway”, Journal of
Earthquake Engineering, Vol. 9, No. 6, pp. 877-897.
Murphy, J.R., O'brien, L.J., 1977. The correlation of peak ground acceleration amplitude
with seismic intensity and other physical parameters. Bulletin of the
Seismological Society of America, Vol. 67, pp. 877-915.
209
Nakamura, Y. (1989) “A method for Dynamic Characteristics Estimation of Subsurface
using Microtremor on the Ground Surface”, QR of RTRI, Vol. 30, No. 1, pp. 25-
31
Oliveira, C.S., Ferreira, M.A. and Mota de Sá, F. (2004). “Seismic Vulnerability and
Impact Analysis: Elements for Mitigation Policies”, Proceedings of the XI
Congresso Nazionale on L’ingegneria Sismica in Italia, Genova, Italy, Keynote
Lecture (on CD).
Oliveira, C.S., Mota de Sá, F. and Ferreira, M.A. (2005). “Application of Two Different
Vulnerability Methodologies to Assess Seismic Scenarios in Lisbon”, Proceedings
of the International Conference: 250th Anniversary of the 1755 Lisbon
Earthquake, Lisbon, Portugal, Paper No. 37.
Ozdemir, P., Boduroglu, M.H. and Ilki, A. (2005). “Seismic Safety Screening Method”,
Proceedings of the International Workshop on Seismic Performance Assessment
and Rehabilitation of Existing Buildings (SPEAR), Ispra, Italy, Paper No. 23.
Paulay T., Priestly M.J.N. 1992. “Seismic design of reinforced concrete and masonry
buildings”. John Wiley & Sons, New York.
Park, Y.J. and Ang, A.H.S. (1985). “Mechanistic Seismic Damage Model for Reinforced
Concrete”, Journal of Structural Engineering, Vol. 111, No. 4, pp. 722-739.
Priestley, M.J.N. (2003). “Myths and Fallacies in Earthquake Engineering, Revisited: The
Mallet Milne Lecture, 2003”, IUSS Press, Pavia, Italy.
210
Restrepo-Vélez, L.F. (2005). “A Simplified Mechanics-Based Procedure for the Seismic
Risk Assessment of Unreinforced Masonry Buildings”, PhD Thesis, European
School for Advanced Studies in Reduction of Seismic Risk (ROSE School),
Pavia, Italy.
Restrepo-Vélez, L.F. and Magenes, G. (2004). “Simplified Procedure for the Seismic
Risk Assessment of Unreinforced Masonry Buildings”, Proceedings of the 13th
World Conference on Earthquake Engineering, Vancouver, Canada, Paper No.
2561.
Rota, M., Penna, A. and Strobbia, C. (2006). “Typological Fragility Curves from Italian
Earthquake Damage Data”, Proceedings of the First European Conference on
Earthquake Engineering and Seismology, Geneva, Switzerland, Paper No. 386 (on
CD).
Sabetta, F., Goretti, A. and Lucantoni, A. (1998). “Empirical Fragility Curves from
Damage Surveys and Estimated Strong Ground Motion”, Proceedings of the 11th
European Conference on Earthquake Engineering, Paris, France, pp. 1-11.
Scawthorn, C., Iemura, H. and Yamada, Y. (1981). “Seismic Damage Estimation for
Low- and Mid-Rise Buildings in Japan”, Earthquake Engineering & Structural
Dynamics, Vol. 9, No. 2, pp. 93-115.
211
Shinozuka, M., Chang, S.E., Eguchi, R.T., Abrams, D.P., Hwang, H.H.M. and Rose, A.
(1997). “Advances in Earthquake Loss Estimation and Application to Memphis,
Tennessee”, Earthquake Spectra, Vol. 13, No. 4, pp. 739-758.
Smolka, A., Allmann, A., Hollnack, D. and Thrainsson, H. (2004). “The Principle of Risk
Partnership and the Role of Insurance in Risk Mitigation”, Proceedings of the 13th
World Conference on Earthquake Engineering, Vancouver, Canada, Paper No.
2020 (on CD).
Sousa, M.L., Campos Costa, A., Carvalho, A. and Coelho, E. (2004). “An Automatic
Seismic Scenario Loss Methodology Integrated on a Geographic Information
System”, Proceedings of the 13th World Conference on Earthquake Engineering,
Vancouver, Canada, Paper No. 2526 (on CD).
Spence, R., Coburn, A.W. and Pomonis, A. (1992). “Correlation of Ground Motion with
Building Damage: The Definition of a New Damage-Based Seismic Intensity
Scale”, Proceedings of the Tenth World Conference on Earthquake Engineering,
Madrid, Spain, Vol. 1, pp. 551-556.
Spence, R., Bommer, J., del Re, D., Bird, J., Aydinoglu, N. and Tabuchi, S. (2003).
“Comparing Loss Estimation with Observed Damage: A Study of the 1999
Kocaeli Earthquake in Turkey”, Bulletin of Earthquake Engineering, Vol. 1, No.
1, pp. 83-113.
Trifunac, M.D., Brady, A,G., 1975. On the Correlation of Seismic Intensity with Peaks of
Recorded Strong Ground Motion. Bulletin of the Seismological Society of
America, Vol. 65, pp. 139-162.
212
Veneziano, D., Sussman, J.M., Gupta, U. and Kunnumkal, S.M. (2002). “Earthquake
Loss under Limited Transportation Capacity: Assessment, Sensitivity and
Remediation”, Proceedings of the Seventh US National Conference on Earthquake
Engineering, Boston, U.S.A. (on CD).
Whitman, R.V., Reed, J.W. and Hong, S.T. (1973). “Earthquake Damage Probability
Matrices”, Proceedings of the Fifth World Conference on Earthquake
Engineering, Rome, Italy, Vol. 2, pp. 2531-2540.
Wood, H.O. and Neumann, F. (1931). “Modified Mercalli Intensity Scale of 1931”,
Bulletin of the Seismological Society of America, Vol. 21, No. 4, pp. 277-283.
Yeh, C.H., Jean, W.Y. and Loh, C.H. (2000). “Damage Building Assessment for
Earthquake Loss Estimation in Taiwan”, Proceedings of the 12th World
Conference on Earthquake Engineering, Auckland, New Zealand, Paper No. 1500.
Zimmerli B., Schwartz J., Schwegler G., 1999 “Mauerwerk - Bemessung und
Konstruktion”. Birkhäuser Verlag, Basel, Boston, Berlin.
213
APPENDIX A
During the last 240 years, a number of earthquakes have affected Bangladesh are
presented below:
Longitude Latitude
214
Time Place Position
Longitude Latitude
215
APPENDIX B
Graph B.1 Seismic hazard map of Bangladesh based on ten percent probability of
exceedence in 50-years period (using Boore et al. (1993) acceleration attenuation
expression), contour values are PGA in g
216
Graph B.2 Seismic hazard map of Bangladesh based on five percent probability of
exceedence in 50-years period (using Boore et al. (1993) acceleration attenuation
expression), contour values are PGA in g
217
Graph B.3 Seismic hazard map of Bangladesh based on two percent probability of
exceedence in 50-years period (using Boore et al. (1993) acceleration attenuation
expression), contour values are PGA in g
218
Graph B.4 Seismic hazard map of Bangladesh based on ten percent probability of
exceedence in 50-years period (using McGuire (1978) acceleration attenuation
expression), contour values are PGA in g
219
Graph B.5 Seismic hazard map of Bangladesh based on five percent probability of
exceedence in 50-years period (using McGuire (1978) acceleration attenuation
expression), contour values are PGA in g
220
Graph B.6 Seismic hazard map of Bangladesh based on two percent probability of
exceedence in 50-years period (using McGuire (1978) acceleration attenuation
expression), contour values are PGA in g
221
Graph B.7 Seismic hazard map of Bangladesh based on ten percent probability of
exceedence in 50-years period (using Duggal (1989) acceleration attenuation
expression), contour values are PGA in g
222
Graph B.8 Seismic hazard map of Bangladesh based on five percent probability of
exceedence in 50-years period (using Duggal (1989) acceleration attenuation
expression), contour values are PGA in g
223
Graph B.9 Seismic hazard map of Bangladesh based on two percent probability of
exceedence in 50-years period (using Duggal (1989) acceleration attenuation
expression), contour values are PGA in g
224
APPENDEX C
m2/MN
225
Sl. Index Index Index Index Index Index Remarks
No. Name of Mosques
1 2 3 4 5 6
m2/MN
226
Sl. Index Index Index Index Index Index Remarks
No. Name of Mosques
1 2 3 4 5 6
m2/MN
227
Sl. Index Index Index Index Index Index Remarks
No. Name of Mosques
1 2 3 4 5 6
m2/MN
228
APPENDEX D
229