You are on page 1of 160

NONLINEAR SEISMIC ANALYSIS OF MULTI-STORIED REINFORCED

CONCRETE BUILDINGS WITH MASONRY INFILL

MD. RUHUL AMIN

Department of Civil Engineering


BANGLADESH UNIVERSITY OF ENGINEERING AND TECHNOLOGY
Dhaka-1000.

February, 2011
NONLINEAR SEISMIC ANALYSIS OF MULTI-STORIED REINFORCED
CONCRETE BUILDINGS WITH MASONRY INFILL

A Thesis
by
MD. RUHUL AMIN

A thesis submitted in partial fulfillment of the requirements for the degree of


MASTER OF SCIENCE IN CIVIL ENGINEERING (STRUCTURAL)

Department of Civil Engineering


BANGLADESH UNIVERSITY OF ENGINEERING AND TECHNOLOGY
Dhaka-1000.

February, 2011

2
The Thesis Titled “Nonlinear Seismic Analysis of Multi-Storied Reinforced Concrete
Buildings with Masonry Infill”, Submitted by: Md. Ruhul Amin, Roll No.:
040404356(P), Session: April/2004; has been accepted as satisfactory in partial
fulfillment of the requirement for the degree of Master of Science in Civil
Engineering (Structural) on February 19, 2011.

BOARD OF EXAMINERS

____________________________
Dr. Raquib Ahsan
Professor, Chairman
Department of Civil Engineering, (Supervisor)
BUET, Dhaka-1000.

____________________________
Dr. Md. Zoynul Abedin
Professor and Head, Member (Ex-Officio)
Department of Civil Engineering,
BUET, Dhaka-1000.

____________________________
Dr. Ishtiaque Ahmed
Professor, Member
Department of Civil Engineering,
BUET, Dhaka-1000.

____________________________
Dr. Md. Mahmudur Rahman
Professor, Member (External)
Department of Civil Engineering,
Ahsanullah University of Science & Technology,
Tejgaon, Dhaka.

3
DEDICATION

To

My Parents

4
DECLARATION

It is hereby declared that except for the contents where specific reference have been
made to the work of others, the studies contained in this thesis is the result of
investigation carried out by the author. No part of this thesis has been submitted to
any other University or educational establishment for a Degree, Diploma or other
qualification (except for publication).

Signature of the Candidate

______________________
(Md. Ruhul Amin)

5
ACKNOWLEDGEMENT

Author wishes to express his deepest gratitude to his esteemed supervisor Dr. Raquib
Ahsan, Professor, Department of Civil Engineering, BUET for his encouraging
supervision al through the study. His systematic and invaluable guidance with
affectionate persuasion have helped me greatly during the study.

I would like to thank Shameem Ahmed, Md. Shohiduzaman Sarker for their immense
help in ABAQUS finite element modeling

My special gratitude to my parents, brothers and friends, the source of inspiration for
all my efforts and achievements.

6
ABSTRACT

Masonry infill wall in frame structures have long been known to affect strength,
stiffness and post peak behavior of the infilled framed structures. The interaction
between the wall and the frame under lateral load dramatically changes the overall
characteristics of the composite structure. In seismic areas, ignoring the composite
action of infill panel and frame is not always on the safe side. So, response to seismic
loads creates a major source of hazard during seismic events.

The present study consists of two parts. At first, lateral load – displacement curve for
column and masonry wall is obtained using non-linear FE analysis from which lateral
stiffness of column and infills has been found. Secondly, this stiffness which is
equivalent to the stiffness of a nonlinear spring is then used in a multi-degree of
freedom (MDOF) spring-mass-dashpot system. The MDOF system is modeled in
MATLAB where dynamic equilibrium equations have been solved by state-space
method. Detailed parametric study with different PGA values, earthquake frequency
and varying number of stories has been performed to obtain a guideline of collapse of
different types of buildings under seismic loading.

After performing nonlinear dynamic analysis, behavior of these structures is


compared. It is evident that natural frequency is completely different from code-
provision of natural frequency due to the presence of infill. Soft-storey frame
collapses earlier than bare and infilled frame for all seismic excitations. As, number of
storey increases, soft storey collapse initiates earlier than buildings with smaller
number of stories.

7
CONTENTS

Page
DECLARATION v
ACKNOWLEDGEMENT vi
ABSTRACT vii
CONTENTS viii

Chapter 1 INTRODUCTION
1.1 General 1
1.2 Objective and Scope of the Study 2
1.3 Methodology of Thesis Work 2
1.4 Organization of the Thesis 3

Chapter 2 LITERATURE REVIEW


2.1 General 4
2.2 Structural configurations and Irregularities 4
2.3 Soft Storey 7
2.3.1 Definition of Soft Storey in Different Codes 7
2.3.2 Seismic Behavior with Soft Storey 16
2.4 Failure Modes of Masonry Infilled RC Frames 19
2.5 Interaction of Frames with Infill Panels 25
2.6 Effect of Infill Panels on Member Forces 26
2.7 Effect of Infill Panels on Overall Seismic Response 28
2.8 Experimental Research on the Seismic Behavior of
30
Infilled Frames
2.9 Analysis of Infilled Frames 34
2.10 Discussion on Modeling of Infill Wall 36
2.10.1 Micro-model (Finite Element Method) 37
2.10.2 Macro-model (Equivalent Diagonal Strut) 37
2.11 Summary 38

8
Chapter 3 MODEL GENERATION AND SOLUTION
3.1 General 39
3.2 Finite Element Model - ABAQUS 39
3.2.1 General Description of ABAQUS Software 40
3.2.2 Overview of ABAQUS Analysis 40
3.2.3 ABAQUS Analysis Options 41
3.2.4 ABAQUS/Explicit Solution Scheme
42
Procedure
3.3 Finite Element Mesh 44
3.3.1 Solid (Continuum) Element 44
3.3.2 Truss Element 46
3.4 Material Properties 48
3.5 Sensitivity Analysis of FE Model 53
3.6 Seismic Methods of Analysis 57
3.7 Introduction to MATLAB 59
3.8 Programming in MATLAB and Solution Technique 60
3.9 Building Configurations and Parameters Used
65
for Nonlinear Dynamic Analysis

Chapter 4 ANALYSIS AND RESULTS


4.1 Introduction 74
4.2 Stiffness Determination 74
4.2.1 Lateral stiffness of RC Columns 74
4.2.2 Lateral Stiffness of masonry walls 80
4.3 Determination of natural frequency 87
4.4 Response of Different of Types of Frames 91
4.5 Comparison of Inter-storey Drift Ratio 98
4.5.1 Three (3) storied building 98
4.5.2 Six (6) storied building 101
4.5.3 Nine(9) storied building 104
4.5.4 Twelve (12) storied building 107
4.6 Comparison of Collapse Time 110

9
4.6.1 Three (3) storied building 110
4.6.2 Six (6) storied building 113
4.6.3 Nine(9) storied building 116
4.6.4 Twelve (12) storied building 119

Chapter 5 CONCLUSIONS AND RECOMMENDSATIONS


5.1 Conclusions 127
5.2 Suggestions for future research 128

REFERENCES 130

10
CHAPTER ONE
INTRODUCTION
1.1 GENERAL
Structural frames are often filled with masonry walls serving as partitions or as
cladding. In the structural design process, such filler walls are considered to be inert
“nonstructural” elements. The structure is assumed to carry the transverse loads by the
frame elements resisting primarily in flexure.

It is apparent from geometrical considerations that a reasonably tight fitted wall


having finite stiffness will impede deformations compatible with frame action. The
frame with filler wall is considerably stronger and stiffer than the frame alone.
Ignoring the interaction between the frame and the filler wall is tantamount to
neglecting a very important structural contribution. Also, the critical regions in the
frame-wall composite may not be the same as those in the frame alone and the
designer may have a risk on brittle links of the frame-wall composite. There is a
general agreement among researchers that infilled frames have greater strength as
compared to frames without infills. The presence of the infill will also increase the
lateral stiffness considerably. Due to the change in stiffness and the mass, the
dynamic characteristics will also change. Understanding the behavior of infilled
frames and having a satisfactory method of analysis will help us to have more realistic
and economical solutions. Earthquakes in past showed that, infills had an important
effect on the resistance and stiffness of buildings.

The behavior of the infilled frame under seismic loading is very complex and
complicated. Since the behavior is nonlinear and closely related to the link between
the frame and the infill, it is very difficult to predict it by analytical methods unless
the analytical models are supported and revised by using the experimental data. Due
to the complex behavior of such composite structures, experimental as well as
analytical research is of great importance to determine the strength, stiffness and
dynamic characteristics at each stage of loading. It is widely recognized that nonlinear
time-history analysis constitutes the most accurate way for simulating response of
structures subjected to strong levels of seismic excitation. This analytical method is
based on sound underlying principles and features the capability of reproducing the
intrinsic inelastic dynamic behavior of structures.
1
1.2 OBJECTIVES AND SCOPE OF THE STUDY
This study will examine building structures with reinforced concrete frames having
masonry infill under dynamic base excitation as time-history. Finite element analysis
technique shall be carried out with the following specific objectives:
 To investigate load-deflection response of a single column upto failure under
lateral load using FEM software.
 To investigate load-deflection response of masonry infill upto failure under
lateral load using FEM software.
 To develop a program in State-Space method using MATLAB by which
response of multi-degree of freedom system to ground motion will be
determined.
 To determine various structural properties for which typical frame-structure
buildings (with different number of storey) is most vulnerable to earthquake
excitation with particular frequency.

1.3 METHODOLOGY OF THESIS WORK


The proposed methodology consists of following steps:
 Column with different dimensions will be modeled with FEM software
ABAQUS. Then, lateral load - top deflection behavior under monotonic
loading will be observed.
 Masonry infill with 5 inch wall thickness will be modeled with FEM software
ABAQUS. Then, lateral load - top deflection behavior under both monotonic
loading will be observed. Both, column lateral stiffness and infill lateral
stiffness will be expected to decrease after elastic limit and a certain fraction
of it is available in the plastic range.
 Three types of buildings will be modeled for frames with full infill, frames
without any infill (bare frame) and frames with infill in upper stories and
having a soft storey at the ground floor considering the nonlinear effect of
column and infill. Modeling will be done in MATLAB where column and wall
will be replaced by nonlinear springs, floor mass will be concentrated at each
storey level and dashpots will be used for damping mechanism.
 Load-displacement response as obtained by FEM will be given as input.
Space-state function will be used for this purpose so that properties of both

2
column and infill can be simultaneously controlled.
 As elasto-plastic behavior of material is used in the model, energy will be
dissipated by nonlinear hysteresis of material and damping mechanism.
Analysis will be performed for 5% damping ratio.
 Analysis will be done for differenent earthquake frequencies, peak ground
accelerations and varying number of stories.

1.4 ORGANIZATION OF THE THESIS


Chapter 1 is a general introduction to the themes that will be dealt: the main topic is
described, the purposes of the work are set and a brief summary of the present job is
presented.

Chapter 2 presents the review of previously published literature in the field of infilled
reinforced concrete frame structures. It also reviews the general response of
reinforced concrete structures and performance of infill wall.

In Chapter 3, the features of the models used for non-linear analysis; after a short
description of the finite element code ABAQUS, the elements properties adopted in
the modeling phase, the materials properties assumed, the strut models used in the
analysis, the analysis procedure followed throughout this study is explained.
Information about dynamic analysis using MATLAB and description of case study
buildings and the parameters considered are also given.

In Chapter 4, the results of analysis are given in detail that includes numerical
evaluation of dynamic response of nonlinear MDOF system when subjected to
harmonic ground motions of different amplitudes and different frequencies. Behavior
of buildings, before and after cracking of infill walls has been observed and a
guideline for predominant frequency for different types of buildings was also given.

Chapter 5 presents the conclusions of the thesis and the final considerations achieved,
giving some suggestion for further works on this topic.

3
CHAPTER TWO
LITERATURE REVIEW

2.1 GENERAL
Walls are often created in buildings by infilling parts of the frame with stiff
construction such as bricks or concrete blocks. Unless adequately separated from the
frame, the structural interaction of the frame and infill panels must be allowed for in
the design. This interaction has a considerable effect on the overall seismic response
of the structure and on the response of the individual members. Many instances of
earthquake damage to both the frame members and infill panels have been recorded
(Stratta and Feldman, 1971). In this chapter, previous works on infilled framed will
be highlighted.

2.2 STRUCTURAL CONFIGURATIONS AND IRREGULARITIES


Configuration, or the general vertical and/or horizontal shape of buildings, is an
important factor in earthquake performance and damage. Buildings that have simple,
regular, symmetric configurations generally display the best performance in
earthquakes. The reasons for this are (a) nonsymmetric buildings tend to have twist
(i.e., have significant torsional modes) in addition to shaking laterally and (b) the
various ‘‘wings’’ of a building tend to act independently, resulting in differential
movements, cracking, and other damage. Rotational motion introduces additional
damage, especially at “re-entrant” or ‘‘internal’’ corners of the building. The term
‘‘configuration’’ also refers to the geometry of lateral load resisting systems as well
as the geometry of the building. Asymmetry can exist in the placement of bracing
systems, shear walls, or moment resisting frames (MRF) that are used to provide
earthquake resistance in a building. This type of asymmetry, of the (Lateral Force
Resisting Systems) LFRS, can result in twisting or differential motion, with the same
consequences as asymmetry in the building plan. An important aspect of
configuration is soft story, which is a story of a building significantly less stiff than
adjacent stories (i.e., a story in which the lateral stiffness is 70% or less than that in
the story above or less than 80% of the average stiffness of the three stories above;
BSSC 2001). Soft stories often (but not always) occur on the ground floor, where
commercial or other reasons require a greater story height, and large windows or
openings for ingress or commercial display (e.g., the building might have masonry
4
curtain walls for the full height, except at the ground floor, where these are replaced
with large windows for a store’s display). Due to inadequate stiffness, a
disproportionate amount of the entire building’s drift is concentrated at the soft story,
resulting in nonstructural and potential structural damage. Many older buildings with
soft stories but built prior to recognition of this aspect are likely to collapse due to
excessive ductility demands at the soft story.
The National Earthquake Hazard Reduction Program (NEHRP) has provisions for the
design of new buildings (BSSC 2001). When a building’s configuration is
‘‘irregular,’’ and it requires an increase in the required strength or adopts other
approaches, to deal with these irregularities. The NEHRP definitions for plan and
vertical irregularities are illustrated in Table 2.1 and Table 2.2.

Table 2.1: NEHRP 2000 Plan Structural Irregularities


1

Torsional irregularity — when diaphragms are


not flexible

Torisonal irregularity exists (1a) when the


maximum story drift, computed including
accidental torsion, at one end of the structure
transverse to an axis is more than 1.2 times the
average of the story drift at the two ends of the
structure. (1b) Extreme torisonal irregularity
exists when ratio>1.4.
2

Re-entrant corners

Plan configurations of a structure and its lateral-


force resisting system contain re-entrant
corners where both projections of the structure

5
beyond a re-entrant corner are greater than
15% of the plan dimension of the structure in
the given direction.
3

Diaphragm discontinuity

Diaphragms with abrupt discontinuities or


variations in stiffness including those having
cutout or open areas greater than 50% of the
gross enclosed diaphragm area or changes in
effective diaphragm stiffness of more than 50%
from one story to the next.
4

Out-of-plane offsets

Discontinuities in a lateral-force-resistance path


such as out-of-plane offsets of the vertical
elements.
5

Nonparallel systems

The vertical lateral-force-resisting elements are


not parallel to or symmetric about the major
orthogonal axes of the lateral-force-resisting
system.

Table 2.2: NEHRP 2000 Vertical Structural Irregularities

6
1

Stiffness irregularity — soft story

(1a) A soft story is one in which the lateral stiffness


is less than 70% of that in the story above or less
than 80% of the average stiffness of the three stories
above. (1b) An extreme soft story is for ratios of less
than 60% or less than 70%, respectively.
2

Weight (mass) irregularity

Mass irregularity exists where the effective mass of


any story is more than 150% of the effective mass of
an adjacent story. A roof that is lighter than the floor
below need not be considered.
3

Vertical geometric irregularity

Vertical geometric irregularity exists where the


horizontal dimension of the lateral-force-resisting
system in any story is more than 130% of that in an
adjacent story.

7
4

In-plane discontinuity in vertical lateral-force-


resisting elements

An in-plane offset of the lateral-force-resisting


elements greater than the length of those elements or
a reduction in stiffness of the resisting elements in
the story below.
5

Discontinuity in capacity — weak story

A weak story is one in which the story lateral


strength is less than 80% of that in the story above.
The story strength is the total strength of all seismic
resisting elements sharing the story shear for the
direction under consideration.

2.3 SOFT STOREY


A soft story, also known as a weak story, is defined as a story in a building
that has substantially less resistance, or stiffness, than the stories above or
below it. In essence, a soft story has inadequate shear resistance or
inadequate ductility (energy absorption capacity) to resist the earthquake-
induced building stresses. Although not always the case, the usual location of
the soft story is at the ground floor of the building. This is because many
buildings are designed to have an open first-floor area that is easily accessible
to the public. Thus the first floor may contain large open areas between
columns, without adequate shear resistance. The earthquake-induced building
movement also causes the first floor to be subjected to the greatest stress,
which compounds the problem of a soft story on the ground floor.

8
Concerning soft stories, the National Information Service for Earthquake
Engineering
(2000) states:
In shaking a building, an earthquake ground motion will search for
every structural weakness. These weaknesses are usually created by sharp
changes in stiffness, strength and/or ductility, and the effects of these
weaknesses are accentuated by poor distribution of reactive masses. Severe
structural damage suffered by several modern buildings during recent
earthquakes illustrates the importance of avoiding sudden changes in lateral
stiffness and strength. A typical example of the detrimental effects that these
discontinuities can induce is seen in the case of buildings with a “soft story.”
Inspections of earthquake damage as well as the results of analytical studies
have shown that structural systems with a soft story can lead to serious
problems during severe earthquake ground shaking. [Numerous examples]
illustrate such damage and therefore emphasize the need for avoiding the soft
story by using an even distribution of flexibility, strength, and mass.

2.3.1 Definition of Soft Storey in Different Codes


Different building codes define soft storey with respect to strength/stiffness
degradation compare with other stories. The followings are the some of them –
(i) ASCE 07 – Minimum Deisgn Loads for Buildings and Other Structures, USA
(2002): Soft storey is one in which the lateral stiffness is less than 70% of that in the
storey above or less than 80% of the average stiffness of the three storey above. An
extreme soft storey is one in which the lateral stiffness is less than 60% of that in the
storey above or less than 70% of the average stiffness of the three storey above. For
example, building on stilts will fall under this category.
(ii) Specification for Structures to be Built in Disaster Area, Turkey (1998): The case
where stiffness irregularity factor in each of the two orthogonal earthquake directions
is greater than 1.5 is considered as soft storey.
Stiffness irregularity factor is defined as ratio of the average storey drift at any storey
to average storey drift at the storey immediately above.
l ki D (D i ) /(D i D1 )

Storey drift D i , of any column or structural wall shall be determined by –

D i D d i D d i D1

Where, d i and d i D1 represents the displacement obtained from the analysis at the end

of any column or structural wall at stories I and i–1.

9
(iii) Indonesian Earthquake Code, Indonesia (1983): The ratio of floor mass to the
stiffness of a particular storey shall not differ by more than 50% of the average for the
structure.
(iv) Criteria for Earthquake Design of Structures, India (IS 1893: 2002):

kn
Soft storey when
k i D 0.7 k i D1
or
k n D1
k D 0.8{( 1/ 3 )( k Dk Dk )}
i i D1 i D 2 i D 3

k2

k1

Figure 2.1: Stiffness irregularities – soft storey.

(v) General Structural Design and Design Loading for Buildings, New Zealand (NZS
4203: 1992): A storey where the ratio of the inter-storey deflection divided by the
product of the storey shear and storey height exceeds 1.4 times the corresponding
ratio for the storey immediately above this level.

(vi) Bangladesh National Building Code, Bangladesh (BNBC 1993): Soft storey is
one in which the lateral stiffness is less than 70% of that in the storey above or less
than 80% of the average stiffness of the three storey above.

The following are five examples of buildings having a soft story on the ground
floor:

10
(a) Chi-chi earthquake in Taiwan on September 21, 1999: In Taiwan, it is
common practice to have an open first-floor area by using columns to support

the upper floors.

Figure 2.2: Damage due to a soft story at the ground floor. The damage
occurred during the Chi-chi earthquake in Taiwan on September
21, 1999 (Day, 2002). [Photograph from the USGS Earthquake
Hazards Program, NEIC, Denver]

In some cases, the spaces between the columns are filled in with plate-glass
windows in order to create ground-floor shops. Figure 2.2 shows an example
of this type of construction and the resulting damage caused by the Chi-chi
earthquake.

11
(b) Northridge earthquake in California on January 17, 1994: Many apartment
buildings in southern California contain a parking garage on the ground floor.
To provide an open area for the ground-floor parking area, isolated columns
are used to

Figure 2.3: Building collapse caused by a soft story due to the parking garage
on the first floor. The building collapse occurred during the
Northridge earthquake in California on January 17, 1994 (Day,
2002).

12
Figure 2.4: View inside the collapsed first-floor parking garage (the arrows
point to the columns). The building collapse occurred during the
Northridge earthquake in California on January 17, 1994 (Day,
2002).

support the upper floors. These isolated columns often do not have adequate
shear resistance and are susceptible to collapse during an earthquake. For
example, Figs. 2.3 and 2.4 show the collapse of an apartment building during
the Northridge earthquake caused by the weak shear resistance of the first-
floor garage area.

(c) Loma Prieta earthquake in California on October 19, 1989: Another


example of a soft story due to a first-floor garage area is shown in Fig. 2.5.
The four-story apartment building was located on Beach Street, in the Marina
District,

Figure 2.5: Damage caused by a soft story due to a parking garage on the first
floor. The damage occurred during the Loma Prieta earthquake in
California on October 17, 1989 (Day, 2002). [Photograph from the
Loma Prieta Collection, EERC, University of California, Berkeley]

13
San Francisco. The first-floor garage area, with its large open areas, had
inadequate shear resistance and was unable to resist the earthquake-induced
building movements.

(d) Izmit earthquake in Turkey on August 17, 1999: In terms of building


conditions, it has been stated (Bruneau, 1999): A typical reinforced concrete
frame building in Turkey consists of a regular, symmetric floor plan, with
square or rectangular columns and connecting beams. The exterior
enclosures as well as interior partitioning are of non-bearing unreinforced
brick masonry infill walls. These walls contributed significantly to the lateral
stiffness of buildings during the earthquake and, in many instances, controlled
the lateral drift and resisted seismic forces elastically. This was especially true
in low-rise buildings, older buildings where the ratio of wall to floor area was
very high, and buildings located on firm soil. Once the brick infills failed, the
lateral strength and stiffness had to be provided by the frames alone, which
then experienced significant inelasticity in the critical regions. At this stage,
the ability of reinforced concrete columns, beams, and beam-column joints to
sustain deformation demands depended on how well the seismic design and
detailing requirements were followed both in design and in construction. A
large number of residential and commercial buildings were built with soft
stories at the first-floor level. First stories are often used as stores and
mmercial areas, especially in the central part of cities. These areas are
losed with glass windows, and sometimes with a single masonry infill at
back. Heavy masonry infills start immediately above the commercial floor.
g the earthquake, the presence of a soft story increased deformation
ds very significantly, and put the burden of energy dissipation on the
y columns. Many failures and collapses can be attributed to the
deformation demands caused by soft stories, coupled with lack of
y of poorly designed columns. This was particularly evident on a
street where nearly all buildings collapsed towards the street.
his soft story condition are shown in Figs. 2.6 and 2.7.

14
Figure 2.6: Damage caused by a soft story at the first-floor level. The damage
occurred during the Izmit earthquake in Turkey on August 17, 1999
(Day, 2002). [Photograph by Mehmet Celebi, USGS]

15
Figure 2.7: Building collapse caused by a soft story at the first-floor level. The
damage occurred during the Izmit earthquake in Turkey on August
17, 1999 (Day, 2002). [Photograph by Mehmet Celebi, USGS]

(e) El Asnam earthquake in Algeria on October 10, 1980: An interesting


example of damage due to a soft story is shown in Fig 2.8 and described
below (National Information Service for Earthquake Engineering 2000):
Although most of the buildings in this new housing development remained
standing after the

Figure 2.8: Building tilting and damage caused by a soft story due to a ground-
floor crawl space. The damage occurred during the El Asnam
earthquake in Algeria on October 10, 1980 (Day, 2002).
[Photograph from the Godden Collection, EERC, University of
California, Berkeley]

earthquake, some of them were inclined as much as 20 degrees and dropped


up to 1 meter, producing significant damage in the structural and non-
structural elements of the first story. The reason for this type of failure was the
use of the “Vide Sanitaire,” a crawl space about 1 meter above the ground
level. This provides space for plumbing and ventilation under the first floor

16
slab and serves as a barrier against transmission of humidity from the ground
to the first floor. Unfortunately, the way that the vide sanitaires were
constructed created a soft story with inadequate shear resistance. Hence the
stubby columns in this crawl space were sheared off by the inertia forces
induced by the earthquake ground motion.

Although the above five examples show damage due to a soft story located on
the first floor or lowest level of the building, collapse at other stories can also
occur depending on the structural design (Figure 2.9). For example, after the
Kobe earthquake in Japan on January 17, 1995, it was observed that there
were a large number of 20-year and older high-rise buildings that collapsed at
the fifth floor, often due to designing the upper floors for a reduced seismic load.. The
cause was apparently an older version of the building code that allowed a weaker
superstructure beginning at the fifth floor. While damage and collapse due to a
soft story are most often observed in

17
Figure 2.9: Ten-story SRC building with third floor collapse during the Kobe
Earthquake (Day, 2002). [Photo courtesy of NIST; NIST, The January
17, 1995 Hyogoken (Kobe) Earthquake, U.S. NIST, Gaithersburg, MD,
1969]

Figure 2.10: Overview of a collapsed gas storage tank, located at a gas


storage facility near Sabanci Industrial Park, Turkey. The elevated
gas storage tank collapsed during the Izmit earthquake in Turkey
on August 17, 1999 (Day, 2002). [Photograph from the Izmit
Collection, EERC, University of California, Berkeley]

18
Figure 2.11: Close-up view of the columns that had supported the elevated
gas storage tank shown in Fig. 4.12. The columns did not have
adequate shear resistance and were unable to support the gas
storage tank during the Izmit earthquake in Turkey on August 17,
1999 (Day, 2002). [Photograph from the Izmit Collection, EERC,
University of California, Berkeley]

buildings, they can also be developed in other types of structures. For


example, Figs. 2.10 and 2.11 show an elevated gas tank that was supported
by reinforced concrete columns. The lower level containing the concrete
columns behaved as a soft story in that the columns were unable to provide
adequate shear resistance during the earthquake.

2.3.2 Seismic Behavior of Infill Frame with Soft Storey


The soft storey concept has technical and functional advantages over the conventional
construction. First, is the reduction in spectral acceleration and base shear due to
increase of natural period of vibration of structure as in a base isolated structure.
However, the price of this force reduction is paid in the form of an increase in spectral

19
displacement and inter-storey drift, thus entailing a significant P – Δ effect, which is a
threat to the stability of the structure (Figure 2.12).

Sa Sd
Sa1 Sd1

Sa2 Sd2

T1 T2 Tn T1 T2 Tn

Figure 2.12: (a) Design earthquake spectral acceleration (Sa) versus time period (Tn);
(b) Design earthquake spectral displacement (Sd) versus time period (Tn)
(Hart and Wong, 2000)

Secondly, a taller first storey is sometimes necessitated for parking of vehicles and/or
retail shopping, large space for meeting room or a banking hall, Figure 2.13. Due to
this functional requirement, the first storey has lesser stiffness of columns as
compared to stiff upper floor frames, which are generally constructed with masonry
infill walls.
Soft storey failure results from the combination of several unfavorable reasons, such
as torsion, excessive mass in upper floor, P – Δ effects and lack of ductility in the
bottom storey. It is not always necessary that all the first tall stories of the buildings
are soft stories, if the columns of first storey have designed on the basis of capacity or
ductility.

Roof

3rd floor

2nd floor

20 1st floor
Figure 2.13: Soft storey type construction (Vukazich, 1998)

The presence of walls in upper storeys makes them much stiffer than the open
ground storey. Thus, the upper storeys move almost together as a single
block, and most of the horizontal displacement of the building occurs in the
soft ground storey itself. In common language, this type of buildings can be
explained as a building on chopsticks. Thus, such buildings swing back-and-
forth like inverted pendulums during earthquake shaking (Figure 2.14a), and
the columns in the open ground storey are severely stressed (Figure 2.14b). If
the columns are weak (do not have the required strength to resist these high
stresses) or if they do not have adequate ductility, they may be severely
damaged (Figure 2.18) which may even lead to collapse of the building

Figure 2.14: Upper stories of open ground storey buildings move together as a single
block – such buildings are like inverted pendulums.

21
Several researchers have investigated the behaviour of soft-storey reinforced concrete
frames under seismic loading. Vasseva (1994) carried out a seismic analysis of frames
taking into account the geometrical nonlinearities. The analysis of frame with a
flexible first storey showed that the maximum displacements and restoring forces
occurred at different time steps and the influence of large axial forces was very strong
on the behaviour of structures. Arlekar, et al. (1997) highlighted the importance of
explicitly recognizing the presence of the open first storey in the seismic analysis of
buildings. The error involved in modeling such buildings as complete bare frames,
neglecting the presence of infills in the upper storeys, was brought out with different
analytical models. Elnashai (2001) analyzed the dynamic response of structures using
static pushover analysis. It was suggested that the nonlinear static analysis can be used
only as an alternative to predict the dynamic response of structure. The static
nonlinear analysis results obtained were closer to the inelastic time-history analysis.

Dolsek and Fajfar (2002) presented a mathematical model of a three-storey reinforced


concrete framed building with infill in the bottom two storeys using DRAIN-2DX. It
was concluded that it is possible to obtain a fair simulation of displacement time-
history with a wide range of mathematical models. However, a model which simulates
both the displacements and the shear forces requires appropriate modeling of both the
force-displacement relationships and the hysteretic rules of the diagonals representing
the infill. The lateral stiffness of brick masonry infilled plane frames was analyzed by
Asteris (2003). The influence of the masonry infill panel opening in the reduction of
the infilled frame stiffness was investigated by means of a new finite element
technique. Kanitkar R. and Kanitkar V. (2004) studied a five-storey building with soft
storey with the intent of reviewing the new provisions for the earthquake resistant
design of structures addressed in the code IS 1893:2002. It was observed that although
the new soft storey provisions of IS 1893 are a step in the right direction, more
investigations are required to completely define the resulting stilt buildings in terms of
their ductility capacities and stable inelastic action under the expected seismic loads.

2.4 FAILURE MODES OF MASONRY INFILLED RC FRAMES


Depending on relative properties of frame and infill, failure modes of masonry infilled
frame show variations. In other words, failure can occur in the frame elements or in
the infill. In estimating the lateral strength and lateral stiffness of masonry infilled
22
frame, it is necessary to find the most critical of the various modes of failure of the
frame and infill.

The usual modes for frame failure are tension failure of surrounding column elements
or shear failure of the columns or beams. These modes are given in Figure 2.15
(Smith and Coull, 1991). Tension failure of the column results from applied
overturning moments. Such mode may be critical one in infilled frames with high
aspect ratio and with very rigid frame elements. The tension steel acts as a flange of
the composite wall. However, in case of weak frame element, dominant modes of
failures are flexural or shear failure of column or beams at plastic hinge locations.

However, if the frame strength is enough to withstand, increasing lateral load results
in failure of infill. In addition to that, the failure may be a sequential combination of
the failure modes of frame and infill. For example, flexural or shear failure of the
columns will generally follow a failure of infill. In both case, failure modes of infill
show variety depending on geometric and material properties. Failure of the infill
occurs by one of the following modes; (a) Shear cracking along the interface between
the bricks and mortar (b) Tension cracking through the mortar joints and masonry (c)
Local crushing of the masonry or mortar in compression corner of the infill. Failure
modes of infill are presented in Figure 2.16 (Smith and Coull, 1991).

23
Figure 2.15: Failure Modes of Reinforced Concrete Frame (Smith and Coull, 1991).

Figure 2.16 : Failure Modes of Infill (Smith and Coull, 1991).

Shear failure of infill is directly related with the horizontal shear induced in the infill
panel by applied load. In addition to applied load, shear resistance of masonry plays an
important role. The resistance of masonry to shearing stress is usually considered to be
provided by the combined action of the bond shear strength and the friction between the
masonry and mortar. Also, vertical compressive stress level induced in infill panel by
applied load is important. When a vertical compressive stress is applied to masonry the
shear resistance is increased with the increase of friction between the masonry and the
mortar. However, friction effect is less effective for the case of perforated brick. Test
results (Polyakow, 1956) showed that for perforated brick the coefficient of internal
friction,µ, is about 0.15, while it varies between 0.6 to 1.7 for solid brick.

Diagonal tension cracking is the result of the diagonal force which produces a principle
tensile stress in the infill equal to tensile strength of the infill material. Smith and Carter
derived the lateral force cause diagonal crack on infill in terms of contact length between
frame and infill under the light of their experimental results. This relation showed that

24
greater value of the length to height ratio of infill or smaller value of λh (stiffer column
relative to the infill) result in greater diagonal strength of infill.

Compressive failure of infill is accompanied by a rapidly increasing rate of deflection.


Therefore, it can be said that compressive failure is a plastic type of infill failure. As done
for diagonal tension failure, compressive failure load is related with the contact length
between frame and infill by Smith & Carter (1969) according to experimental results. The
result of this relation can be concluded as follows; smaller value of λh results in greater
compressive strength of infill. This can be explained with that stiffening of column leads
to the reduction in lateral deflection. And stiffer column means smaller value of λh.
However, because of the weakness of the shearing and tensile modes relative to the
compressive failure mode, it is thought that a compressive failure would be unlikely to
occur in brickwork.

Apart from these three modes of failure, a forth mode of failure may take place. This is
sliding shear failure. If sliding shear failure of the masonry infill occurs, the equivalent
structural mechanism changes from the diagonally braced pin-jointed frame to the knee-
braced frame. Therefore, this type mechanism results in shear failure of surrounding
columns. This mechanism is shown in Figure 2.17 (Paulay and Priestley, 1992).

Apart from these three modes of failure, a forth mode of failure may take place. This is
sliding shear failure. If sliding shear failure of the masonry infill occurs, the equivalent
structural mechanism changes from the diagonally braced pin-jointed frame to the knee-
braced frame. Therefore, this type mechanism results in shear failure of surrounding
columns. This mechanism is shown in Figure 2.6 (Paulay and Priestley, 1992).

25
Figure 2.17: Sliding Shear Failure of Infill (Paulay and Priestley, 1992).

RC columns can experience two failure and shear failure. The column resistance due
to axial- by making the columns stronger than the beams. As a result, the beams,
rather than columns, absorb the earthquake energy and sustain damage in the process.
This resistance is determined, amongst other factors, by the total cross-sectional area
of vertical steel rebars. Shear failure is brittle and must be avoided in columns by
providing closely

26
(a) (b) (c)
Figure 2.18: Examples of column failure: (a) buckling of vertical column rebars due
to inadequately spaced horizontal ties (b) severe damage of a ground floor
column due to improper confinement of concrete and lapping of large
number of longitudinal bars (c) typical infrequent horizontal ties with 90
degree hooks, which were unable to confine concrete core

spaced transverse ties that enclose all the vertical bars. Tall and slender columns often
tend to be weaker than the framing beams, particularly when the column width in the
direction of framing is small. To prevent the undesirable “weak column-strong beam”
effect seismic design codes require the columns to be stronger than the beams. Since
columns are often wider than the beams framing into them and have a larger amount
of steel reinforcement than beams, the column width in the direction of frame action
should be generally equal to or greater than the width of beams framing into them.
Also, circular columns with spiral reinforcement tend to show superior earthquake
performance over rectangular columns of the same cross-sectional area. However,
spiral reinforcement is not common in design practice, particularly in columns of
rectangular or square shape. Further, the entire length of spiral must be made from a
single bar.
Also, the ends of the spiral need to be securely anchored into the beam-column joints
or beam-slab system.

Failure of masonry walls can occur:


• Due to In-Plane Action,
• Due to Out-of-Plane Action.
In masonry structures the roofs are supported by the walls and in case of their collapse,
they can not continue to stabilize the overall structure. Horizontal loads are mainly being
carried by in-plane action of walls (walls acting as shear walls). For this reason,
reinforcement of masonry walls in most cases aims to enhance in-plane capacity of walls.

The walls with in-plane action may collapse in three main failure modes: sliding, flexural
and shear; Figure 2-19 shows details. Sliding failure is defined as the horizontal
movement of entire parts of the wall on the single brick layer, vapor barrier or mortar bed.
Flexural failure, where the wall behaves as a vertical cantilever under lateral bending and,
either cracking in the masonry tension zone (opening of bed joints) or crushing at the wall
toe will limit the bearing capacity. Shear failure is characterized by critical combination

27
of principal tensile and compressive stresses as a result of applying combined shear and
compression, and leads to typical diagonal cracks. In practice mainly two types of shear
cracking can be observed, joint cracking by local sliding along the bed joint and diagonal
cracking associated with cracks running through the bricks as well as the joints. Both kind
of shear failure must be considered in design. Shear failure mainly occurs if the ratio of
the height to length of the wall is relatively low, but this is a common situation in practice
[Marzhan, 1998].

Figure 2.19: Different types of failure modes in masonry walls with in-plane action: (a)
Sliding Failure, (b) Flexural Failure, (c) Shear Failure [Marzhan, 1998]

In reality, masonry walls alternatively turn on their toe and in case to the deformation
required by the earthquake on the resistance of the wall, displacement passes the elastic
limit and starts to collapse at the edges. The nature of earthquake movements does not
allow to the structure to stay in this position and in the period of time tries to pull the
structure in the opposite direction. This reciprocation causes of time to separate the wall
from the foundation with every back and forth movement, this action is very dangerous
for the wall because of the following reasons:
(i) The impulses on the walls decrease the length of the wall on the foundation,
and this makes walls weaker and weaker in front of the further hits,
(ii) This staggering movement includes strong impulses and generates huge
values of horizontal forces on top of the wall (it is better to call in the slab),
Figure 2-4. This force can be big enough to generate shear failures. Experimental
tests show that short time after the start of the staggering effect, the toes of the
wall started to crush and after that shear failure appeared [Moghaddam, 1994],
(iii) Staggering movement decreases the stiffness of the wall on the one hand and
increases earthquake loads on the other hand, this effect generates bigger

28
responses on the structure. Studying the response spectrum shows by decreasing
the fundamental period of the structures lead the structures to have more response
acceleration in the slab (consequently more forces) and finally to the collapse of
the whole structure, Figure 2-20.

Figure 2.20: Staggering movement generates big forces in top of the wall

The walls with out-of-plane action are the perpendicular walls to the dominating
earthquake direction. They behave like a flat slab on line-supports (ground, roof and two
orthogonal shear walls). Inertia forces are generated by the mass of the walls due to the
earthquake acceleration. Bending failure can occur in continuously supported slabs (see
Figure 2-21). In lack of sufficient resistance in orthogonal wall connections, the
connections separate during earthquake loading.

29
Figure 2.21: Different types of failure in walls with out-of-plane action

2.5 INTERACTION OF FRAMES AND INFILL PANELS:


The effect of infill panels on the response of R/C frames subjected to seismic action is widely
recognised and has been subject of numerous experimental investigations, while several
attempts to model it analytically have been reported. It is well known that the assessment of
R/C frames with infilled walls is dependent on the material constituting the infills and the
geometry of both frame and infill. The possible effects of infills on frames are the following:
(i) The presence of infills does not affect the structural response. This can be the case
if the infills are very light and flexible, or completely isolated from the R/C frame, or so
brittle that a total failure is expected even for a moderate ground acceleration.
(ii) The infills are assessed to have a significant contribution on the response, and
they are expected to remain in the elastic range. In this case a linear elastic analysis can be
performed. The ductility capacity should be set to 1, unless inelastic structural wall behaviour
can be expected, with columns acting as tension or compression boundary members, and the
infill acting as a connecting shear element.

30
(iii) The infills are assessed to have a significant contribution to the response, and
they are expected to suffer significant damage during the seismic event. In this case the high
probability of the formation of a soft storey has to be recognised and taken into account.

In order to decide whether the first case is applicable to a given situation, the following
parameters should be examined: details of connections between infill and frame; ratio of the
stiffness of the infilled wall and the stiffness of the bare frame; ratio of the shear strength of
the infilled wall and the bare frame.

The decision as to whether cases (ii) or (iii) apply requires consideration of the likely infill
failure mechanisms.

2.6 EFFECT OF INFILL PANELS ON MEMBER FORCES


As mentioned above, considerable uncertainties are involved in estimation of the
seismic interaction between infill panels and structural frames. In a study of response
of low-rise buildings to moderate-strength ground shaking, modelling non-linear
response with DRAIN-2DX, Sritharan and Dowrick (1994) found that:
(i) infill-excluded frames had highly non-linear response;
(ii) infill-included frames responded elastically, despite the base shears being
up to four times larger than without infill;
(iii) the above results varied quantitatively but not qualitatively when element
stiffness were varied by factors of up to 2.
As an easier alternative to non-linear, time-history analyses, from a straightforward
finite-element response spectrum analysis some basic design information may be
derived. While to take such data as definitive seismic design criteria would be
misleading, sensible use of the computed forces in design would nevertheless be
much better than ignoring the presence of the infill, as has often been the case in the
past. By carrying out comparative analyses with and without the infill panels, at least
a qualitative idea of the effect of the infill can be obtained.
(a) Infill panels. The shear stresses computed in the infill panels should give a
reasonable indication whether or not the infill will survive the design earthquake.
Despite being very approximately determined, the shear stress level will also help in
determining what reinforcement to use in the panel and whether to tie the panel to the
frame.

31
(b) Frame members. The design of the beams and columns abutting the infill is
generally the least satisfactory aspect of this form of seismic construction. Because of
the approximations in the analytical model, the stresses in the frame members are ill
defined. Failures tend to occur at the tops and bottoms of columns, due to shears
arising from interaction with the compression diagonal which exists in the infill panel
during the earthquake (Stratta and Feldman, 1971). Unfortunately, no comprehensive
design criteria for this problem have yet been established, and further research
examining the frame rather than the panel stresses is required. In simple analyses, if
the analysis indicates the failure of the infill panels, the frame should be analysed with
any failed panels deleted, so that appropriate frame stresses may be taken into
consideration.

Because such structures often have highly non-linear behaviour arising from the
interaction between infill and frame, analytical modelling is a complex problem.
Crisafulli et al. (2000) have reviewed the different analytical procedures, discussing
how best to implement them.

32
(a) Non-plastered infill frame (b) Plastered infill frame
Figure 2.22: Failure Pattern of Infilled Frames.

2.7 EFFECT OF INFILL PANELS ON OVERALL SEISMIC RESPONSE


The principal effects of infill panels on the overall seismic response of structural
frames are:
(i) to increase the stiffness and hence increase the base shear response in most
earthquakes;
(ii) to increase the overall energy dissipation capacity of the building; and
(iii) to alter the shear distribution throughout the structure.
The more flexible the basic structural frame, the greater will be the above-mentioned
effects. As infill is often made of brittle and relatively weak materials, in strong
earthquakes the response of such a structure will be strongly influenced by the
damage sustained by the infill and its stiffness-degradation characteristics.

33
Figure 2.23: Horizontal seismic shear diagram for lift core of 20-storey hotel building
showing effect of brick partitions above the fourth floor (Dowrick, 2003).

To fully simulate the earthquake response of an infilled frame, a complex non-linear


time-dependent finite-element dynamic analysis would be necessary, as provided by
computer codes such as DRAIN-2DX. Such are the problems and effort involved in
using such software, e.g. in modelling the structural behaviour of normal masonry
infill, that more rudimentary dynamic analyses are more commonly used.

For many structures a response spectrum analysis in which the infill panels are
simulated by simple finite elements will be very revealing. Figure 2.23 shows the
results of such an analysis of a multi-storey hotel building, in which all of the
bedroom floors (fourth to twentieth) have alternate partitions in brickwork. Curve A
shows the horizontal earthquake shear distribution up the shear core ignoring the
brickwork, while curve B shows the shears when an approximate allowance for the
brick walls is made.

Allowing for the infill reduces the fundamental period of the structure from 1.96 s to
1.2 s, and correspondingly increases the base shear on the shear core from 21.0 MN to
31.0 MN. The effect on the distribution of shear is particularly dramatic; it can be
seen how the brick walls carry a large portion of the shear until they terminate at the
fourth-floor level; below this level the shear walls of the core must, of course, take the
total load.

In carrying out this simple type of dynamic analysis difficulty may be experienced in
selecting a suitable value of shear modulus G for the infill material. Not only is the G
value notoriously variable for bricks, but the infill material may not even have been
chosen at the time of the analysis. Either a single representative value may have to be
assumed, or it may be desirable to take a lower and a higher likely value of G in two
separate analyses for purposes of comparison.

Further examples of the effect of infill on mode shapes and periods of vibration of
structural frames are reported by Lamar and Fortoul (1969) and Sritharan and
Dowrick (1994). In their examples the first mode period was reduced by factors of

34
about 3–6 when comparing the infill-included with the infill-excluded cases. Mode
shapes of some structures also depend upon the distribution of infill.

2.8 EXPERIMENTAL RESEARCH ON THE SEISMIC BEHAVIOR OF


INFILLED FRAMES
A number of past studies (Klinger and Bertero 1978; Bertero and Brokken 1983;
Mander and Nair 1994) focused on evaluating the experimental behavior of masonry
infilled frames to obtain formulations of limit strength and equivalent stiffness. A
more rigorous analysis of structures with masonry infilled frames requires an
analytical model of the force – deformation response of masonry infill. While a
number of finite element models have been developed to predict the response of
infilled frames (Dhanasekar and Page 1986. Mosalam 1996, Malnotra 2002), such
micro modeling is time-consuming for analysis of large structures. Alternatively, a
macro model allows treatment of the entire infill panel as a single unit.

Historically, five different types of infill have been considered in the study of
the behavior of infilled frames: brick, clay tile, concrete block, plain concrete, and
reinforced concrete. The surrounding framing in these studies comprised either
reinforced concrete or steel members. The following literature review focuses on the
steel frame-RC infill wall system due to its distinctive behavior, with a brief summary
of other infilled frames. According to Liauw et al. (1983a, 1983b), infilled frames
may be divided into two categories: (1) those with connectors along the interfaces
between the frames and the infill walls are called integral infilled frames; and (2)
those without are called non-integral infilled frames. These definitions are used in the
following literature review.

Research on infilled frames started with the investigation of their static behavior
under monotonic lateral loading (e.g., Benjamin and Williams, 1957; Holmes, 1961)
with the intention of developing an effective method to predict their ultimate lateral
strength. With the recognition that the lateral loading imposed on infilled frames is
induced by dynamic phenomena producing reversing load histories, such as
earthquakes, wind, or explosions, researchers began to focus their efforts on the cyclic
load behavior of infilled frames. During the last three decades, experiments with three

35
different types of loading have been carried out to simulate dynamic phenomena,
especially seismic forces: cyclic static and dynamic load tests, pseudo-dynamic tests,
and shake-table tests.
Mallick and Severn (1968) performed half-cyclic dynamic load tests on small
scale, two-story infilled steel frames, where the cyclic load was applied to the infilled
frame in one direction only. The steel frames comprised 0.75 inch  0.75 inch square
steel bars and the story dimension was 24 inch  24 inch. The dynamic characteristics,
such as the damping ratio and the energy dissipation capacity, were compared
between infilled steel frames with and without interface shear connectors. It was
found that using a small number of shear connectors in loaded corners could prevent
the rotation of the infill walls inside the steel frames and increase the stiffness of the
system. However, they failed to see any strength increase with the use of shear
connectors. The frequencies and mode shapes of multi-story infilled steel frames were
obtained analytically using two models: a shear model in which the axial deformation
of the steel components was ignored, and a cantilever model in which the bending
deformation of the steel frame members was ignored. Test results showed that the
cantilever model was better than the shear model for analysis of multi-story infilled
steel frames, particularly for those with a height/span ratio greater than 2.
Liauw (1979) conducted both static and dynamic cyclic load tests on both
integral and non-integral steel frames with RC infill walls. The four-story steel frame
models comprised 22 mm  22 mm square steel bars, with the size of infill being
either 305 mm  610 mm  22 mm or 305 mm  610 mm  22 mm (height  width 
thickness). The reinforcement ratio for the infill wall was 0.56%. In contrast to the
conclusion drawn by Mallick and Severn (1968), the presence of the interface
connectors was shown to increase both frame stiffness and strength significantly.
Furthermore, failure of the integral infilled steel frames was induced largely by shear
between the steel frames and the infill walls, instead of by the diagonal compression
failure of the infill walls in the non-integral infilled steel frames. The dynamic
characteristics of these systems were studied further in another series of cyclic tests
(Liauw and Kwan, 1985) on similar infilled steel frames having three different
interface configurations: (i) no connectors; (ii) connectors welded only along the infill
wall/steel girder interface; and (iii) connectors welded along the entire infill wall/steel
frame interface. The tests showed that the infilled steel frames with the Type (iii)

36
interface configurations were the most reliable type of construction because they
possessed the highest energy dissipation capacity, the greatest damping ratio in the
nonlinear range of deformation and the slowest stiffness degradation.

In Japan, two sets of tests were conducted to investigate the behavior of portal
steel frames infilled with reinforced concrete, subjected to combined action of
constant gravity loading and static cyclic lateral loading (Makino et al., 1980). The
portal steel frames were approximately one-third scale and comprised wide-flange
sections. There were two specimens in each set, one of which had the strong axis of
the steel columns oriented perpendicular to the plane of the infill wall, and the other
having the weak axis of the steel columns oriented perpendicular to the plane of the
infill wall. The portal steel frames in the first set of tests were cyclically loaded well
into the inelastic range of behavior before casting of the infill walls. The two virgin
portal steel frames used in the second set of tests had the same sizes as those in the
first set, respectively. A few headed studs were employed with the objective of
preventing out-of-plane failure of the infill walls. The composite action of system
resulted in a uniformly distributed crack pattern in the RC infill walls. The infilled
frames having columns bent about their strong axis were shown to have ductile
behavior as good as those of typical bare steel frames. It was also concluded that an
infilled steel frame having the steel frame pre-loaded into the inelastic range of
behavior can have almost the same strength as a virgin steel frame with RC infill
walls. A tentative design recommendation was proposed by Makino (1984) with the
backing of more experimental research, in which the primary variable was the cross
section area ratio between the concrete and the steel column. Also in Japan, another
research group (Hayashi and Yoshinaga, 1985, 1986, 1987, 1994) performed a series
of similar tests on portal infilled steel frames, in which the width-to-height ratio was
chosen as the main variable, and interface connectors were either headed studs or
deformed bars.

Shake-table tests are believed to be the best approach for simulating the
behavior of a structure during an earthquake. However, shake-table tests are
expensive, especially for large scale models. Kwan and Xia (1995) reported a shake-
table test on steel frames infilled with light reinforced concrete walls (non-integral).
In their tests, the specimen comprised a pair of one-third scale infilled steel frames,
37
connected through reinforced concrete floor slabs at each story, representing a one-
bay, four-story structure. The steel frame comprised 40 mm  40 mm  2.5 mm high-
yield tubes and story dimension is 1125 mm  1500 mm (height  width).
Accelerograms from the EI Centro earthquake were used to excite the shake-table,
with progressively increasing magnitudes of acceleration being applied until failure of
the specimen occurred. It was observed that the infill walls separated from the steel
frames during early loading stages and acted as a diagonal compressive strut. The
specimen reached its maximum strength when the corners of the infill walls in the
first story were crushed. The specimen did not fail even when the applied peak
acceleration was increased to 1.50g, although the frames were badly damaged.
However, it was concluded that the RC infill walls would have collapsed out-of-plane,
had there not been steel plates attached to the outside of the RC walls to keep them in
place. It was also found that the natural frequency of the specimen decreased rapidly
as the applied peak acceleration of the simulated earthquake was increased, while the
damping increased from 1.7% to 11.0%.

Masonry is a traditional infill material for both steel and RC frame structures.
A number of experimental studies have been carried out to evaluate the effect of
masonry infill on the seismic behavior of the surrounding frames. Although it was
found that masonry infill can also significantly increase the stiffness and strength of
frame structures (Klinger and Bertero, 1978; Mehrabi et al., 1996; Negro and
Verzelett, 1996), designers in the United States are reluctant to treat unreinforced
masonry infill as a structural element in seismic regions due to several unfavorable
characteristics of the masonry infill. For example, Mander and Nair (1994) found that
the shear strength of brick infill walls was greatly affected by cyclic loading;
Mosalam et al. (1997) reported that there were significant pinching zones in the
hysteretic load-displacement curves from his cyclic load tests on two-bay masonry
infilled steel frames. Furthermore, shake-table tests indicated that it is often
unavoidable for the masonry infill to collapse out-of-plane as a result of increased
acceleration of ground motion (Dawe et al., 1989; Kwan and Xia, 1995).

2.9 ANALYSIS OF INFILLED FRAMES

38
During the last three decades, different approaches have been proposed for the
prediction of the ultimate strength of infilled steel frames subjected to monotonic
lateral load. Holmes (1961) proposed that the infill wall be replaced by an equivalent
diagonal strut having a width equal to one-third of the diagonal length of the infill
wall. In his elastic model, no axial deformation was included in the steel members and
the infilled steel frame achieved its maximum lateral strength when the equivalent
strut reached a limiting value of compressive strain. The corresponding deflection was
calculated based on the shortening of the equivalent strut. Stafford Smith (1966)
proposed an expression relating the width of the equivalent strut to the properties of
the frame and infill wall. The width of the equivalent strut varied with value of the
following non-dimensional factor:
1/ 4
 E th3 sin 2è 
ë h D  m 
 4 E I
c c 

1/ 4
 E t 3 sin 2è 
l D  m  (2.1)
 4 E I
c c h 

a D 0.175 L(lh) D0.4

where, l = ratio of infill-to-frame stiffness


Em = elastic modulus of the infill material, ksi
Ec = elastic modulus of the frame material, ksi
t = thickness of the infill wall, inches
h = height of a single story, inches
Ic = moment of inertia of the frame columns, inch4
 = slope of the infill diagonal relative to horizontal

Stafford Smith and Carter (1969) further related the width of the equivalent strut not
only to factor h, but also to the variation of the elastic modulus of the infill material
at different stress levels. Makino (1984) proposed a simplified formula to calculate
the width of the equivalent strut based on Stafford Smith and Carter’s work. In his
formula, the width of the equivalent strut was only related to the diagonal length of

39
the infill wall or the thickness of the infill wall. Liauw and Kwan (1983a) expressed
the equivalent strut width as a fraction of hcos:
0.86
bD (h cos è )  0.45(h cos è ) (2.2)
ëh

Where, the non-dimensional factor parameter h is defined in Eq. (1.3.1). This


relation was obtained by parametric study using the finite element method. Despite
the controversy on the determination of a reasonable width of the equivalent strut, the
equivalent strut concept was applied in the elastic design of multi-story infilled frames
by different researchers (Stafford Smith and Carter, 1969; Makino, 1984; Saneinejad
and Hobbs, 1995). It should be noted that all the previous equivalent strut methods
were established for the analysis of non-integral infilled frames.

A number of researchers have proposed analytical methods based on plastic


analysis theory, because the equivalent strut method neglects the contribution of the
steel frame and cannot fully represent the behavior of composite infilled steel frames
up to their ultimate strength levels. Wood (1978) applied plastic analysis to non-
integral steel frames, in which the collapse modes and loads depended on the bending
strength of the steel frames and the crushing stress of the infill wall. However, an
unrealistic assumption was made in his models that the whole infill wall would
behave in a perfectly plastic manner. Therefore, an empirical penalty factor, p, was
needed to reduce the effective crushing stress of the infill wall and to account for the
discrepancy between the theoretical predictions and experimental results. Four
different failure modes were identified in Wood’s work: a composite shear mode, a
shear rotation mode, a diagonal compression mode, and a corner crushing mode.
Nadjai and Kirby (1998) extended the shear mode to non-integral infilled steel frames
having different types of partially-restrained connections. The experimental and
analytical results showed that the shear mode overestimated the shear capacity of the
infill wall. Therefore, another panel reduction factor, , was used to further reduce the
value of the penalty factor, p. Based on the results of a nonlinear finite element
analysis and experimental observations, Liauw and Kwan (1983a and 1983b)
proposed plastic analysis models for both integral and non-integral infilled steel
frames, in which the infill wall always failed through crushing in the corner regions.

40
Furthermore, the formation of plastic hinges in the steel frames was determined based
on the distribution of the compressive crushing stress in the corner regions of the infill
walls. The contribution of interface friction was neglected in these models. Based on
nonlinear finite element analysis and model tests, Saneinejad and Hobbs (1995)
developed an inelastic method to calculate the ultimate lateral load and cracking load
of non-integral infilled steel frames. This method is similar to that proposed by Liauw
and Kwan (1983a, 1983b), except that the effect of the interface friction force was
included and plastic hinges were assumed to form only in the loaded corners.

2.10 DISCUSSION ON MODELING OF INFILL WALL

Recently a number of papers and documents have been published reporting on studies
of the behaviour, design, and analysis of infilled frames. A recent NZSEE Bulletin
article (Bell and Davidson, 2001) provides a review of macro (equivalent strut) and
micro models, and reports on some comparative studies between equivalent strut
models and FEM (‘solid’ element) models. The micro-model method is a finite
element method where the frame elements, masonry work, contact surface, slipping
and separation are modeled to achieve the results. This method has seems to be
generating the better results but it has not gained popularity due to its cumbersome
nature of analysis and computation cost. The macro-model which is also called
simplified model of equivalent diagonal strut method to study the global response of
the infilled frames. This method uses one or more struts to represent the infill wall.
The drawback of it is to the lack of its capability to consider the opening precisely as
found in the infill wall.

2.10.1 Micro-Model (Finite Element Method)


Achyutha, Jagadish et. al. (1985) investigated the elastic behaviour of single storey
infilled frame which had opening. The interface conditions such as slip, separation
and friction loss at the contact surface were considered using link element. They were
achieved by adjusting the axial, shear and tension force in the link element. The
opening was modlled by assigning very low value of infill thickness and Young’s
modulus of elasticity of infill but high value of Poisson’s ratio. It was reported that the
lateral stiffness of the structure decreases with the increase in opening size. The
principal stresses were maximum at the corners of the opening and the compression

41
ends when full contact was the condition which further increased by allowing
separation at interface. However, the author stated that the equivalent diagonal strut
mechanism may not be applicable for structures which have openings.

2.10.2 Macro-Model (Equivalent Diagonal Strut)


Many years of researches and experimental tests in the field of infilled frames consent
to asses that the influence of the infills on response of reinforced concrete structures
subjected to lateral loads is not negligible, on the contrary of what in the common
structural design is assumed up to now.

However, there are some problems to understand the interaction between the infills
and the boundary frame, and this is one of the main reason that led the researchers to
propose several models to try to fit the experimental results; is possible to divide these
models into two big classes: the local models (where the infills are modelled adopting
discrete or continuum models for masonry) and global models (where the infills are
replaced by single or multiple compression strut).

Figure 2.24: Global models (a) and local models (b) for infilled structures

How previously stated, the main aspect that affects the characterization of the infilled
frames under seismic loads is the interaction between the infills and the boundary
frames; experimental evidences have shown that the phenomena is influenced
essentially by the strength of the two materials, concrete and masonry, and by the
level of horizontal load applied to the structure: so it’s possible to analyze the pre-
peak phase by dividing it into three stages. At the beginning, when low forces (and

42
thus low deformations) are applied, there is no separation between the boundary
frames and the wall (if there are no gaps between the two component), and its
contribute in terms of stiffness is very high: this stage lasts just for very low values of
load, and so it’s supposed to be no such essential. Successively, when forces start to
increase to consistent values, a separation occurs between the wall and the frames
(both columns and beams), and so the resistant mechanism of the infills becomes very
similar to a compression strut, with compressive stresses concentrated at the
compressed corner and rapidly decaying in the central zone: in this stage there is a
quite small energy dissipation because cracking is still not reached. Finally, once the
crack strength has been reached, two cases are possible: shear collapse in the concrete
element if the infill is very resistant and the frame is very poor detailed, or diffusion
of the cracks in the infill panel with consequent growth of energy dissipation in
hysteretic cycles; three types of crack pattern have been in the most of the
experimental tests: horizontal slip crack (when the mortar is very weak), diagonal
cracks (stair-step configuration when the bricks are very strong or diagonal
configuration when also the mortar is of good quality), corner crushing (when both
masonry and frame are strong, and the strut failure mechanism is so fully developed).

2.11 SUMMARY
This chapter summarizes the experimental and theoretical research work conducted in
the area of infilled frame. Detailed model of building frames and their solution
techniques will be given in the next chapter.

43
CHAPTER THREE
MODEL GENERATION AND SOLUTION

3.1 GENERAL
Earthquake response analysis is an art to simulate the behavior of a structure
subjected to an earthquake ground motion based on dynamics and a
mathematical model of the structure. The correct analysis will depend upon
the proper modeling of the behavior of materials, elements, connections and
structure. Models may be classified mainly by essential difference in the
degree-of-freedom. The model, or the number of degree of freedom, should
be selected carefully considering the objective of the analysis. Sometimes
sophistication or complicated models are not only useless but also create
misunderstanding to interpret the results in practical problems. Therefore, it is
important to select an appropriate and simple model to match the purpose of
the analysis. Analytical models should also be based on physical observations
and its behavior under dynamic load.

The two most important factors in the analysis and design of building
structures are choosing an appropriate structural modeling method which
reflects the actual behavior of the system and deciding on the analyzing
technique to be performed on the structure.
In the first part of this chapter, the basic assumptions used in the modeling
studies are presented. these assumptions are divided into three categories: (i)
material behavior, (ii) element behavior and (iii) structural behavior.
In the second part, a lumped mass model of building is considered and
dynamic equilibrium equations will be solved using state-space method.

3.2 FINITE ELEMENT MODEL - ABAQUS


The FE method is a useful tool for analyzing problems with complex geometry,
material properties, and boundary conditions. This technique can produce
comprehensive and reliable results if the method is used properly. It is much
cheaper than the full-scale experimental testing. An accurate model to
simulate the actual structures is necessary to accomplish a successful
analysis.

39
In this study, ABAQUS version 6.8.1 was used. ABAQUS is a general-purpose
nonlinear finite element analysis program, which is used for stress, heat
transfer and other types of analysis in mechanical and structural engineering.
The pre- and post-processing work was performed by ABAQUS/CAE version
6.8.1 which is an additional graphical user interface module that allows a user
to execute a FE analysis process from start to finish. The FE model can be
viewed and checked interactively and the results (stresses, displacements,
etc.) can be visualized graphically.

3.2.1 General Description of ABAQUS Software


The computer code ABAQUS was used to perform the numerical analysis for the tested
masonry vault. ABAQUS is a suite of general-purpose, advanced nonlinear finite element
analysis (FEA) programs. ABAQUS is used throughout the world for stress, heat transfer,
and other types of analysis in mechanical, structural, civil, biomedical, and related
engineering applications. The ABAQUS suite consists of three core products:
ABAQUS/Standard, ABAQUS/Explicit and ABAQUS/CAE. Each of these packages
offers additional optional modules that address specialized capabilities.
The following items are some of the more significant features of the code:

(i) A wide range of element types, including continuum elements (1D, 2D, and
3D), beams, membranes and shells.

(ii) Element formulations suitable for large displacements, rotations and strains.

(iii) Material models for metals, sand, clay, concrete, jointed rock, plastics and
rubber.

(iv) User-defined subroutines permit inclusion of additional material models and


element types.

(v) Automatic time incrementation within an implicit time integration algorithm


(Hilber-Hughes-Taylor) for nonlinear dynamic analysis.

(vi) Fracture mechanics capability.

3.2.2 Overview of ABAQUS Analysis


The finite element analysis was carried in three stages, namely, pre-
processing, simulation and post-processing. The software Java (JCreatorPro)
was used as a pre-processor to create the finite element model as it is robust

40
in its interactive graphics, visualization and automatic mesh generation
capabilities. An input file which describes the model was then generated using
the preprocessor and appropriate parameters, constraints and prescribed
conditions were added to the input file via a text editor. Next, analysis was
carried out using the finite element code, ABAQUS/Explicit.

Post-processing was carried out using ABAQUS/CAE which provides a variety


of options for displaying the results such as color contour plots, animations
and x-y axis plots.

3.2.3 ABAQUS Analysis Options


ABAQUS/Standard enables a wide range of linear and nonlinear engineering
simulations to be carried out efficiently, accurately, and reliably. The direct-
integration dynamic procedure provided in ABAQUS/Standard uses the implicit
Hilber-Hughes-Taylor operator for integration of the equations of motion. In an
implicit dynamic analysis the integration operator matrix must be inverted, and a set
of simultaneous nonlinear equilibrium equations must be solved at each time
increment. This solution is done iteratively using Newton’s method. The Hilber-
Hughes-Taylor operator is unconditionally stable and, thus, there is no limit on the
size of the time increment that can be used for most analyses in ABAQUS/Standard
(accuracy governs the time increment in ABAQUS/Standard). The flexibility
provided by this integration allows ABAQUS/Standard to be applied to those portions
of the analysis that are well-suited to an implicit solution technique, such as static,
low-speed dynamic, or steady-state transport analyses.

ABAQUS/Explicit provides finite element solution techniques to simulate a wide


variety of dynamic and quasi-static events in an accurate, robust, and efficient manner
conducting the best analyses possible in the shortest amount of time. The explicit
dynamics procedure performs a large number of small time increments efficiently. An
explicit central-difference time integration rule is used together with the use of
diagonal or “lumped” element mass matrices. Each increment is relatively
inexpensive because there is no solution for a set of simultaneous equations.
However, the size of the time increment in an explicit dynamic analysis is limited,
because the central-difference operator is conditionally stable. The use of small
41
increments (dictated by the stability limit) is advantageous because it allows the
solution to proceed without iterations and without requiring tangent stiffness matrices
to be form. Contact conditions and other extremely discontinuous events are easily
handled by ABAQUS/Explicit.

3.2.4 ABAQUS/Explicit Solution Procedure


The explicit procedure by ABAQUS is based on the implementation of an explicit
integration rule along with the use of diagonal or “lumped” element mass matrices.
The equations of motion for the body are integrated using the explicit central
difference integration rule:
(i } 1 )
u (i }1) } u (i ) } }t (i }1)u / 2
(3.1)

(i } 1 ) (i } 1 ) 1 / (i )
u/ 2
} u/ 2
} (}t (i }1) } }t (i ) )u // (3.2)
2

where u/ is velocity and u// is acceleration. The superscript (i) refers to the increment
number and i − ½ and i − ½ refer to mid-increment values. The key to the
computational efficiency of the explicit dynamics procedure is the use of diagonal
lumped element mass matrices as the accelerations at the beginning of the increment
are computed by:
u //( i ) } M }1.( F (i ) } I (i ) ) (3.3)

where M is the diagonal lumped mass matrix, F the applied load vector and I the
internal force vector. The acceleration of any node is determined completely by its
mass and the net force acting on it, making the nodal calculations very inexpensive as
there are no simultaneous equations to solve.

Satisfying dynamic equilibrium at the beginning of the increment provides the


acceleration. Subsequently, the velocities and displacements are determined based on
Equation (3.1) and (3.2). The term “explicit” refers to the fact that the state at the end
of the increment is based solely on the displacements, velocities, and accelerations at
the beginning of the increment. This method integrates constant accelerations exactly.
For the method to produce accurate results, the time increments must be quite small

42
so that the accelerations are nearly constant during an increment. Since the time
increments must be small, analyses typically require many thousands of increments.
Fortunately, each increment is inexpensive as there are no simultaneous equations to
solve. Most of the computational expense lies in the element calculations to determine
the internal forces of the elements acting on the nodes. The element calculations
include determining element strains and applying material constitutive relationships
(the element stiffness) to determine element stresses and, consequently, internal
forces.

By default, the stable time increment is calculated automatically in


ABAQUS/Explicit. If the time increment is larger than this maximum amount of time,
the increment is said to have exceeded the stability limit. A possible effect of
exceeding the stability limit is numerical instability, which may lead to an unbounded
solution. However, it is generally not possible to determine the stability limit exactly.
Hence conservative estimates are used instead. The stability limit is defined in terms
of the highest frequency in the system.

The stability limit is defined in terms of the highest frequency in the system (ω max)

defined by the expression


2 T
}t stable } } min (3.4)
wmax r

The actual highest frequency in the system is based on a complex set of interacting
factors, and it is not computationally feasible to calculate its exact value.
Alternatively, a conservative estimate of the stable time increment can be redefined
based on element-by-element estimate as:
Le
}tstable } (3.5)
cd

where Le is the characteristic element dimension and cd the current effective,


dilatational wave speed of the material. As an approximation, the shortest element
distance can be used, but the resulting estimate is not always conservative. The wave
speed is a property of the material which can be expressed as

43
^ ^
r} 2 r
cd } (3.6)
r

^ ^
where r and r are the effective Lame’s constants, and ρ is the mass density.

3.3 FINITE ELEMENT MESH


Figure 3.1 shows the element families that are used most commonly in a stress
analysis. One of the major distinctions between different element families is the
geometry type that each family assumes. This section will describe the properties
of the elements that were employed in the current analysis, namely,
hexahedral elements and truss elements. A summary of the types of element
used for the various components of the model is presented in Table 3.1.

Figure 3.1: Commonly used element families

Table3.1: Properties of components in the finite element model


Structural Element Element designation
Material type
components type in ABAQUS/Explicit
RCC
Plain concrete Hexahedral C3D8R
Column
Reinforcement Truss T3D2
Brick
Clay brick Hexahedral C3D8R
masonry
Portland cement mortar Hexahedral C3D8R

44
3.3.1 Solid (Continuum) Element
The concrete, brick and mortar were modeled using hexahedral elements with
three translational degrees of freedom at each node. They were known to
perform better as compared to triangular or tetrahedral elements when their
shape is approximately rectangular which is typical of the rubber piece,
aluminum cap and the test specimen of the model. A good mesh of
hexahedral elements usually provides a solution of equivalent accuracy at
less cost and they offer better convergence rate than triangular and
tetrahedral elements, and sensitivity to mesh orientation in regular meshes is
not an issue.

Figure 3.2: 8-noded brick element Figure 3.3: 2-noded truss element

Solid (Continuum) elements in ABAQUS are named as follows:

C 3D 8 R

reduced integration
with hourglass control
number of nodes

two-dimensional (2D) or three-dimensional (3D)

solid / continuum / brick

The hexahedral elements available in ABAQUS/Explicit are linear reduced


integration elements (C3D8R) which have the advantage of alleviating
problems of shear-locking associated with fully integrated linear elements
when subjected to bending loads. However, they tend to suffer from hour-

45
glassing or zero-energy modes under the effects of bending. These elements
have just a single integration point located at the centroid. As such, it is
possible for them to distort in such a way that the strains calculated at the
integration point are zero, which in turn lead to uncontrolled distortion of the
mesh. To alleviate this problem, ABAQUS introduces a small amount of
artificial “hourglass stiffness” (Flanagan and Belytschko, 1981) in these
elements to limit the propagation of hourglass modes and is more effective
when a fine mesh is used.

3.3.2 Truss Element

Truss elements are used in two and three dimensions to model slender, line-like
structures that support loading only along the axis or the centerline of the element. No
moments or forces perpendicular to the centerline are supported. Truss elements in
ABAQUS are named as follows:
T 3D 2 H

optional: hybrid (H),


coupled temperature-displacement (T),
or piezoelectric(E)

number of nodes

two-dimensional (2D) or three-dimensional (3D)

truss

Reinforcement was modeled using truss elements, T3D2, which transmit only
axial force and have no resistance to bending. They are useful for modeling of
pin-jointed frames and thus can be used to model the reinforcement, which
serve to transmit the force from the top of column to the base without any
rotational resistance at the joints. The truss element used is a 2-node linear
element with three translational degrees of freedom at each node.

46
(a) (b) (c)
Figure 3.4: (a) FEM model of RCC Column, (b) close view main bar and tie
bar and (c) boundary condition: fixed support at bottom and
concentrated load at top.
Masonry is composite material which contains units and mortars in its body.
Due to different material properties of units and mortars and also due to
complex geometry of lay out, masonry is an anisotropic material as a whole
structure.

(a) (b)

47
(c) (d)
Figure 3.5: (a) FEM model of Masonry, (b) boundary condition: fixed at
bottom; close view of (c) brick and (d) mortar.

The majority of the proposed constitutive models for masonry, therefore, can be
classified in two categories: (i) the one-phase material models, treating masonry as an
ideal homogeneous material with constitutive equations that differ from those of the
components; and (ii) the two-phase material models where the components are
considered separately to account for the interaction between them. The second one is
used here.

3.4 MATERIAL PROPERTIES

After FE modeling of elements it is necessary to use materials with appropriate


material, properties. The flowing section will describe the materials and properties
used for the subsequent study.

(a) Concrete Characteristics


Stress-strain relationship for concrete is linear for small stress values. However once
stresses reach approximately 40% of the compressive strength of the concrete the
relationship becomes increasingly nonlinear (Figure 3.6).

If the load applied increases, the stress value will reach the peak of the nonlinear
curve. At this point the concrete is at its maximum compressive strength and any
additional loads will cause failure of the bond between the aggregate and the cement
paste. ABAQUS models the compressive stress-strain relationship as elastic and
plastic. Elastic properties are given for the initial linear behavior. Then specific

48
concrete properties are used to state the initial yield point and the stress and strain
values at which the concrete fails.
30

25

20

Stress (MPa)
15

10

0
0 0.001 0.002 0.003 0.004 0.005
Strain

Figure 3.6: Concrete stress – strain curve under uniaxial compression (Mander et al.,
1988).

When the concrete has been loaded into the inelastic range and is then unloaded some
elasticity will be lost. While this loss of elasticity is accounted for in the model the
unload/reload response is idealised as a straight line.

2.5

2
Stress (MPa)

1.5

0.5

0
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 0.0014
Strain
49
Figure 3.7: Concrete stress – strain curve under uniaxial tension (Mander et al., 1988).

The stress-strain relationship in tension is considered linear to its failure point. This is
considered to be 8-10 % of the total compression stress. Once cracks form the
concrete undergoes softening where it is still able to withstand small tensile stresses
as can be seen in Figure 3.7. The effect of this in relation to reinforced concrete will
be explained in the section on tension stiffening.

An elasto-plastic damage model is used to describe the nonlinear material properties


of concrete. Damage is caused by cracking or crushing under or compressive loading
conditions. Thus, tensile d t as well as compressive d c parts constitute the total

damage d.
1 } d } (1 } sc d t )(1 } st d c )

Where the two functions st , sc add in stiffness effects arising from closing and

reopening of cracks.

fc } rc Ec for 0 } f c  0.40 f c/ (3.7)


2
rc r 
Ec / }  c/ 
f c  rc 
f c/ 
 r  r  for 0 } rc  rc/
1 }  Ec c/ } 2    c/ 
 fc   rc 

r 
f c/  c/ r
 rc  ;
  rc  
r

r }1 }  /  
  rc  
for rco } rc  2rc/

Ec
where, r }
f c/
Ec }
rc/

50
rcu } rc
f c/  for 2rc/  rc  rcu
rcu } 2rc 0

(b) Reinforcing Steel Characteristics


Steel reinforcement is used in RC structures to provide the tensile strength
that concrete lacks. The properties of reinforcing steel, unlike concrete, are
generally well known and do not depend on environmental conditions or time
(creep). Reinforcement is usually classified on the basis of geometrical
properties, such as size and surface characteristics, and on the mechanical
properties, such as characteristic yield stress and ductility.

Typical stress-strain curves for reinforcing steel bars used in RC construction


are obtained from coupon tests of bars loaded monotonically in tension. For all
practical purposes, reinforcing steel exhibits the same stress-strain curve in
compression as in tension. Thus, the specification of a single stress-strain
relation is sufficient to define the material properties needed in the analysis of
RC structures. Properties of steel were modeled by means of a elastic-perfectly
plastic stress-strain response in compression and tension as:
fs } rs Es for rs  rsy (3.8)

f sy for rsy } rs  rsh

0 for rsh  rs
500
400
300
200
Stress (MPa)

100
0
-100-0.090 -0.068 -0.045 -0.023 0.000 0.023 0.045 0.068 0.090
-200
-300
-400
-500
Strain

51
Figure 3.8: Steel uniaxial stress-strain curve under tension and compression

(c) Brick Characteristics


In the market, there are several types of bricks that can be found and being used in
construction industry. The commonly used brick types are clay bricks, sand-lime
bricks and concrete bricks. These types of bricks have been standardized for its usage.
In this research, the type of bricks that to be used is clay bricks.
Coordinating size Work size
Length Width Height Length Width Height
10// 5 // 3// 9.5// 4.5// 2.75//
(250 mm) (125 mm) (75 mm) (240 mm) (115 mm) (65 mm)
[Note: The work sizes are derived from the corresponding coordinating sizes by the subtraction of a nominal
thickness of 3/8 inch (10 mm) for the mortar joint]
18
16
14
12
Stress (MPa)

10
8
6
4
2
0
0.E+00 2.E-04 4.E-04 6.E-04 8.E-04 1.E-03 1.E-03 1.E-03
Strain

Figure 3.9: Brick stress – strain curve under uniaxial compression (Hossain, 1997)

52
0.9
0.8
0.7
0.6
Stress (MPa) 0.5
0.4
0.3
0.2
0.1
0
0 1E-05 2E-05 3E-05 4E-05 5E-05 6E-05 7E-05
Strain
Figure 3.10: Brick stress – strain curve under uniaxial tension (Hossain, 1997)

It is assume that behavior of bricks is linear elastic until the damage (crack) is
detected. Usually, the crack goes through the full height (or width) of brick so it
is obvious that the failure condition must respect this.

(d) Mortar Characteristics


The model of mortar is based on an equivalent one-dimensional stress-strain
relation that depends also on 2D state of stress (through the Kupfer's failure
criteria that is used to obtain the limit stresses for the one-dimensional stress
strain relation) and on some special material properties such is fracture
energy of mortar.
8
7
6
5
Stress (MPa)

4
3
2
1
0
0.0E+00 5.0E-04 1.0E-03 1.5E-03 2.0E-03 2.5E-03
Strain

Figure 3.11: Mortar stress – strain curve under uniaxial compression (Hossain, 1997)

53
0.6

0.5

0.4

Stress (MPa)
0.3

0.2

0.1

0
0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05 5.0E-05 6.0E-05
Strain

Figure 3.12: Mortar stress – strain curve under uniaxial tension (Hossain, 1997)

The constitutive law for mortar is controlled by an equivalent uniaxial stress-


strain relation. The simplification of the problem from 2D to 1D is not ideal but
it is relatively easy to develop and understand. To make the uniaxial relation to
be more corresponding with the real behaviour the limits of the relation
(strenghts in tension and in compression) are computed from a 2D failure
criteria (the Kupfer's criteria is used here) and depends on the actual 2D
stresses.

3.5 SENSITIVITY ANALYSIS OF FE MODEL

Mesh sensitivity analysis using different mesh size and time period sensitivity
analysis using different loading rate was performed. For this, lateral
displacement is applied at the top of the column and brick masonry. For
column, it is obvious from figure 3.13 and 3.15 that mesh size less than 25
mm and rate of displacement application less than 12.5 mm/sec will be

54
adequate for FE analysis. For masonry, (figure 3.17 and 3.19) mesh size of 25
mm and displacement rate of 0.25 mm/sec was used.

200
195
190
Peak lateral force (kN)

185
180
175
170
165
160
155
150
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70
Mesh size (mm)

Figure 3.13: Peak lateral load vs. mesh size (results of a 375 mm × 375 mm ×
3 m column).

55
15.625 mm mesh
200
20.83 mm mesh
180 31.25 mm mesh
160 62.5 mm mesh
140

Lateral force (kN)


120
100
80
60
40
20
0
0 2 4 6 8 10
Lateral displacement (cm)

Figure 3.14: Lateral load vs. lateral displacement for different mesh size ((results of
a 375 mm × 375 mm × 3 m column).

190
170
Peak lateral force (kN)

150
130
110
90
70
50
0 5 10 15
Rate of displacement (cm/sec.)

Figure 3.15: Peak lateral force vs. rate of displacement application to a 375 mm ×
375 mm × 3 m column.

56
180
12.5 cm/second
160
2.5 cm/second
140

Lateral force (kN)


1.25 cm/second
120 0.833 cm/second
100
80
60
40
20
0
0 2 4 6 8 10 12
Lateral displacement (cm)

Figure 3.16: Lateral force vs. lateral displacement for different duration load
application to a 375 mm × 375 mm × 3 m column.

50
48
46
Peak lateral force (kN)

44
42
40
38
36
34
32
30
0.01 0.1 1 10
Rate of displacement (mm/sec.)

Figure 3.17: Peak lateral load vs. rate of displacement application to a 6 m


long × 1.575 m height × 125 mm thick masonry panel.

57
50
45
40
35

Lateral force (kN)


30
25 0.0625 mm/sec
20
0.0833 mm/sec
15
0.005 mm/sec
10
0.25 mm/sec
5
1.25 mm/sec
0
0 0.1 0.2 0.3 0.4 0.5
Lateral displacement (mm)

Figure 3.18: Lateral load vs. rate of displacement application to a 6 m long ×


1.575 m height × 125 mm thick masonry panel.

45
44
43
Peak lateral force (kN)

42
41
40
39
38
37
36
35
0 10 20 30 40 50
Mesh size (mm)

Figure 3.19: Peak lateral load vs. mesh size (results of a 6 m long × 1.575 m
height × 125 mm thick masonry panel).

58
45
40
35

Lateral force (kN)


30
25
20
18.75 mm mesh
15
22.5 mm mesh
10
28.125 mm mesh
5
37.5 mm mesh
0
0 0.1 0.2 0.3 0.4 0.5
Lateral displacement (mm)

Figure 3.20: Lateral load vs. lateral displacement for different mesh size
(results of a 6 m long × 1.575 m height × 125 mm thick masonry
panel).

3.6 SEISMIC METHODS OF ANALYSIS

Civil engineering structures are always designed to carry their own dead weight,
superimposed loads and environmental loads such as wind or waves. These loads are
usually treated as maximum loads not varying with time and hence as static loads. In
some cases, the applied load involves not only static components but also contains a
component varying with time which is a dynamic load. In the past, the effects of
dynamic loading have often been evaluated by use of an equivalent static load, or by
an impact factor, or by a modification of the factor of safety.

Many developments have been carried out in order to try to quantify the effects
produced by dynamic loading. Examples of structures where it is particularly
important to consider dynamic loading effects are the construction of tall buildings,
long bridges under wind-loading conditions and buildings in earthquake zones, etc.

Typical situations where it is necessary to consider more precisely the response


produced by dynamic loading are vibrations due to equipment or machinery, impact
load produced by traffic, snatch loading of cranes, impulsive load produced by blasts,
earthquakes or explosions. So it is very important to study the dynamic nature of
structures.

59
Dynamic characteristics of a damaged and undamaged body are, as a rule, different.
This difference is caused by a change in stiffness and can be used for the detection of
damage and for the determination of its parameters (crack magnitude and location).

Once the structural model has been selected, it is possible to determine the
seismically induced forces in the structures. There are different methods of
analysis which provide different degree of accuracy. The analysis process can
be categorized on the basis of three factors: the type of the externally applied
loads, the behavior of structure/ or structural materials and the type of
structural model selected (Figure 3.21).

Analysis process

External action Behavior of Type of model


structures/structural materials

Static analysis Elastic/linear analysis 1D

2D
Dynamic analysis Elastic-plastic/non-
linear analysis 3D

Figure 3.21: Method of analysis process (Syrmakezis, 1996)


Based on the type of external action and behavior of structure, the analysis
can be further classified as (Beskos and Anagnostoulos, 1977) –
(i) linear static analysis
(ii) linear dynamic analysis
(iii) non-linear static analysis
(iii) non-linear dynamic analysis
Structural behaviour is inherently nonlinear, particularly in the presence of large
displacements or material nonlinearities, the structural response can be accurately
caught only by means of nonlinear dynamic analyses. The non-linear dynamic
analysis (inelastic time history analysis) involves a time-step by step
integration of dynamic equilibrium equation. The general equation for a
dynamic response of a multi-degree-of-freedom system subjected to ground
motion is given by

60
[ M ]{x // (t )} } [C ]{x / (t )} } [ K ]}x(t )}} }[ M ]{x // (t )} (3.9)
g

where, [M] = mass matrix


[C] = damping matrix
[K] = stiffness matrix
{x // (t )} = acceleration vector

{x / (t )} = velocity vector

{x(t )} = displacement vector

[ M ]{x // (t ) = inertia force due to earthquake


g

3.7 INTRODUCTION TO MATLAB


The basic element in MATLAB is the matrix, and unlike other computer
languages it does not have to be dimensioned or declared. MATLAB’s original
objective was to be the tool to solve mathematical problems in linear algebra,
numerical analysis, and optimization; but it quickly evolved as the preferred
tool for data analysis, statistics, signal processing, control systems,
economics, weather forecast, and many other applications. Over the years,
MATLAB evolved creating an extended library of specialized built-in functions
that are used to generate among other things two-dimensional (2-D) and 3-D
graphics and animation and offers numerous supplemental packages called
toolboxes that provide additional software power in special areas of interest
such as
(i) Curve fitting
(ii) Optimization
(iii) Signal processing
(iv) Image processing
(v) Filter design
(vi) Neural network design
(vii) Control systems
(viii) Statistics

3.8 PROGRAMMING IN MATLAB AND SOLUTION TECHNIQUE

61
The lumped parameter or mass-spring-dashpot model (Figure 3.21) is at best a cruse
approximation of the actual geometry of the structure, typically in the form of a beam
or frame model with discrete lumped masses at a small number of degrees-of-
freedom. The simplicity of the geometric model allows increased complexity on the
loading side and in the nonlinear dynamic response. The LPM is used to determine
overall dynamic response to actual earthquake ground motion input. The nonlinear
beam element or spring characteristics describing the hysteretic behavior of individual
stories or subassemblages of the building can be obtained from Structural component
model of FEM models subjected to cyclic loading.

m3
m3

k3
c
m2
m2
k2
c
m1
m1

k1
c

Figure 3.21: Lumped mass model

Parameter studies in support of design considerations performed with LPMs


investigate the dynamic response to different ground motion inputs, soil conditions in
the form of boundary springs, and the dynamic interaction between various parts of
the structure, including higher mode effects. LPMs are invaluable tools to obtain a
basic understanding of the dynamic response characteristics of a building, but rely
heavily on the correct definition of component hysteresis which had to include not
only the nonlinear material behavior but also effects resulting from the true geometry
of the structure such as changes in axial load level due to overturning, uplift, and
coupling. LPMs are not suitable to predict local or global failure mechanisms or
distress levels in individual structural components.

62
The multi-story structure as shown in Figure 3.23, 3.24 and 3.25 may be idealized as a
multi-story shear building by assuming that the mass is lumped at the floor and roof
diaphragms. The diaphragms are infinitely rigid, and the columns are axially
inextensible but laterally flexible. The dynamic response of the system is represented
by the lateral displacement of the lumped mass with the number of degrees of
dynamic freedom, or modes of vibration ‘n’ being equal to the number of masses. The
resultant vibration of the system is given by the superposition of the vibrations of each
lumped mass. Each individual mode of vibration has its own period and mat be
represented by a single-degree-of-freedom system of the same period, ad each mode
shape, or eigen vector, remains of constant relative shape regardless of the amplitude
of the displacement. A reference amplitude of a given mode shape may be assigned
unit value to give the normal mode shape. The actual amplitudes must be obtained
from initial conditions. The mode of vibration with the longest period (lowest
frequency) is termed as the fundamental modes with shorter periods (higher
frequencies) are termed higher modes or harmonics.

Analysis of multi-degrees of freedom will be done where number of degrees of


freedom equals the number of story. Floor mass was considered as the lumped mass.
Let first consider a three degrees of freedom system. By taking free body for each part
(Figure 3.21) we get the following equations -
Equilibrium equation for mass ( m ) at first floor
1

m x // } k x } k (x } x ) } cx / } c(x / } x / ) } }m x //
1 1 1 1 2 2 1 1 2 1 1 g

m x // } }k x } k (x } x ) } cx / } c(x / } x / ) } m x //
1 1 1 1 2 2 1 1 2 1 1 g

(k } k ) k
x // } } 1 2 x } 2c x / } 2 x } c x / } x //
1 m 1 m 1 m 2 m 2 g
1 1 1 1

Equilibrium equation for mass ( m ) at second floor


2

m x // } k (x } x ) } k (x } x ) } c(x / } x / ) } c(x / } x / ) } }m x //
2 2 2 2 1 3 3 2 2 1 3 2 2 g

63
m x // } }k (x } x ) } k (x } x ) } c(x / } x / ) } c(x / } x / ) } m x //
2 2 2 2 1 3 3 2 2 1 3 2 2 g

k c / k2 } k3 2c / k 3 c /
x // } 2 x } x } x } x } x } x } x //
2 m 1 m 1 m 2 m 2 m 3 m 3 g
2 2 2 2 2 2

Equilibrium equation for mass ( m ) at third floor


3

m x // } k (x } x ) } c(x / } x / ) } }m x //
3 3 3 3 2 3 2 3 g

m x // } }k (x } x ) } c(x / } x / ) } m x //
3 3 3 3 2 3 2 3 g

k c / k3 c /
x // } 3 x } x } x } x } x //
3 m 2 m 2 m 3 m 3 g
3 3 3 3

 0 1 0 0 0 0 
 x1   k 1 } k 2 x   0 
/
c k2 c
 //  } }2 0 0   1/   
m1 m1 m1 m1 }1
 x1     x1   
x  
/ 0 0 0 1 0 0  x   0 
 2//  }  k 2 c
}
k2 } k3
}2
c k3 }}
c   2/  }   x g//
x 2   m 2 m2 m2 m2 m2 m 2  x 2  } 1
x /   0 0 0 0 0 1  x3   0 
 3//   k3 c k c   /  } 1
x 3   0 0 } 3 }   x 3   
 m3 m3 m3 m3 

64
 0 1 0 0 0 0 
 k1 } k 2 c k2 c 
} m }2
m1 m1 m1
0 0 
 1

 0 0 0 1 0 0 
}A}}  k 2 c
}
k2 } k3
}2
c k3 c 
 m2 m2 m2 m2 m2 m2 
 0 0 0 0 0 1 
 k3 c k c 
 0 0 } 3 } 
 m3 m3 m3 m3 

0 1 0 0 0 0 0 0 
} 1 0 1 0 0 0 0 0 
    
0 0 0 1 0 0 0 0 
}B}}   }C}}   }D}}  
} 1 0 0 0 1 0 0 0 
0 0 0 0 0 1 0 0 
     
} 1 0 0 0 0 0 1 0

For multi-degrees of freedom system it needs to create generalized expressions for


matrix coefficients. For n-storied frame
Size of matrix A = 2n  2n
(2mth-1) row of matrix A= … 0 0 0 0 0 1 0 0 0
0…
km c km } km }1 c k m }1 c
2mth row of matrix A = …. 0 0 } }2 0
mm mm mm mm mm mm

0….
Size of matrix B = 2n  1
Size of matrix C = 2n  2n
Diagonal elements of C are 1 and other elements are zero.
Matrix D is always zero.

The state-space method is based on transforming the N second-order coupled


equations into a set of 2N first-order coupled equations by augmenting the
displacement response vectors with the velocities of the corresponding coordinates
(Newland, 1989).

65
The State-Space block implements state-space system a system whose behavior is
defined by
x / } Ax } Bu
(3.10)
y } Cx } Du

Where x is the state vector, u is the input vector, and y is the output vector. The matrix
coefficients must have these characteristics, as illustrated in the following diagram:
(i) A must be an n-by-n matrix, where n is the number of states.
(ii) B must be an n-by-m matrix, where m is the number of inputs.
(iii) C must be an r-by-n matrix, where r is the number of outputs.
(iv) D must be an r-by-m matrix.

n m

n A B

r C D

The block accepts one input and generates one output. The number of columns in the
B and D matrices determines the input vector width. The number of rows in the C and
D matrices determines the output vector width.
ss function: ss Create state-space model or convert LTI model to state-space.
sys = ss(A,B,C,D) creates a continuous-time state-space (‘ss’) model ‘sys’ with
matrices A,B,C,D. The output ‘sys’ is a ‘ss’ object.
D = 0 mean the zero matrix of appropriate dimensions.
sys = ss creates an empty ‘ss’ object.
sys = ss(A,B,C,D);
This declares a state-space model.
The first two dimensions of A,B,C,D determine the number of states, inputs, and
outputs, while the remaining dimensions specify the array sizes.

lsim operatior: LSIM Simulate time response of LTI models to arbitrary inputs.
lsim(A,B,C,D,U,T) plots the time response of the linear system:

66
x / } Ax } Bu
(3.11)
y } Cx } Du

the input time history U. Matrix U must have as many columns as there are inputs, U.
Each row of U corresponds to a new time point, and U must have length (T) rows.
The time vector T must be regularly spaced. lsim (A,B,C,D,U,T,X0) can be used if
initial conditions exist. X0 is the initial state vector, used for setting condition.

3.9 BUILDING CONFIGURATIONS AND PARAMETERS USED FOR


NONLINEAR DYNAMIC ANALYSIS:

6 m = 20 ft

6 m = 20 ft

6 m = 20 ft

6m 6m 6m
= 20 ft = 20 ft = 20 ft
Figure 3.22: Plan view of building

Three types of building frames (Figure 3.23, 3.24 and 3.25) will be studied
with following configuration
(a) Model I: Frame modeled bare frame. However, masses of the walls as in
model II are included in the model.
(b) Model II: Frame has one half brick infill masonry walls (5 inch thick) in all
storeys, including the first storey.

67
(c) Model III: Frame has no walls in the first storey and one half brick infill
masonry walls (5 inch thick) in the upper storeys.

6 m = 20 ft

3 m = 10 ft

Figure 3.23: Model I – No infill/bare frame

6 m = 20 ft

2.4 m = 8 ft 3 m = 10 ft

Figure 3.24: Model II – Fully infilled frame

68
6 m = 20 ft

2.4 m = 8 ft 3 m = 10 ft

Figure 3.25: Model III – Soft storey frame

Each frame having a bay distance (clear) of 6 m (= 20 ft) in both directions with floor-
to-floor-height of 3 m (=10 ft). Each floor of the building with a slab thickness of
137.5 mm (= 5.5 inch) supports a live load of 40 psf, partition wall load of 80 psf and
floor finish of 30 psf.
5.5
Self-weight of slab =  150 psf = 68.75 psf
12

Partition wall = 80 psf


Floor finish = 30 psf
Seismic mass = 178.75 psf (excluding live load)
Seismic mass lumped at floor level (for an interior frame) = 178.75 psf × (20 ft × 60
ft) = 214500 lb. = 97323.049 kg. [Q 1 kg. = 2.204 lb.]
5% damping (typical for building) is used.
Number of storey considered: 3 storey, 6 storey, 9 storey and 12 storey (for each
frame configuration).

69
Earthquake ground motions frequencies considered: 0.25 Hz., 0.50 Hz., 0.75 Hz., 1
Hz., 1.25 Hz., 1.50 Hz., 1.75 Hz., 2 Hz., 2.25 Hz., 2.5 Hz., 2.75 Hz., 3 Hz.
Earthquake peak ground accelerations (PGA) considered: 1 m/s2, 2 m/s2, 3 m/s2.
Earthquake strong motion duration: 5 seconds
Details of earthquake input signal shown in Figure 3.26 to Figure 3.37.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)

Figure 3.26: Earthquake ground motion with frequency = 0.25 Hz. with strong motion
duration of 5 seconds.

70
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.27: Earthquake ground motion with frequency = 0.50 Hz. with strong motion
duration of 5 seconds.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.28: Earthquake ground motion with frequency = 0.75 Hz. with strong motion
duration of 5 seconds.

71
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.29: Earthquake ground motion with frequency = 1 Hz. with strong motion
duration of 5 seconds.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.30: Earthquake ground motion with frequency = 1.25 Hz. with strong motion
duration of 5 seconds.

72
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.31: Earthquake ground motion with frequency = 1.50 Hz. with strong motion
duration of 5 seconds.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.32: Earthquake ground motion with frequency = 1.75 Hz. with strong motion
duration of 5 seconds.

73
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.33: Earthquake ground motion with frequency = 2.00 Hz. with strong motion
duration of 5 seconds.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.34: Earthquake ground motion with frequency = 2.25 Hz. with strong motion
duration of 5 seconds.

74
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.35: Earthquake ground motion with frequency = 2.50 Hz. with strong motion
duration of 5 seconds.

3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2
Ground Acceleration (meter/sec2)

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.36: Earthquake ground motion with frequency = 2.75 Hz. with strong motion
duration of 5 seconds.

75
3
PGA=1 m/s2
PGA=2 m/s2
2 PGA=3 m/s2

Ground Acceleration (meter/sec2) 1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
Figure 3.37: Earthquake ground motion with frequency = 3.00 Hz. with strong motion
duration of 5 seconds.

76
CHAPTER FOUR
ANALYSIS AND RESULTS
4.1 INTRODUCTION:
After detailed finite element modeling in ABAQUS, lateral stiffness of column and
infill was obtained which will be used in MATLAB program to obtain response of
different types of building frames. Providing data for no, of stories, ground motion
frequency, ground acceleration amplitude, relative stiffness of frame-to-infill,
response in terms of interstorey drift and time to collapse are studied and presented in
this chapter.

4.2 STIFFNESS DETERMINATION


Siffness of frame which comes from column and beam (beams are considered to
infinitely stiff along its longitudinal direction) and stiffness of wall need to be
determined first. These stiffnesses, which is equivalent to nonlinear shear springs is
lumped mass model, will be given as input in MATLAB program

4.2.1 Lateral Stiffness of RCC Columns


In order to obtain lateral stiffness, lateral displacement applied at column top which
continues to increases until collapse occurs. Different sizes square columns with a
height of 3 m (=10 feet) are selected which is fixed at bottom. All columns are reinfo-

Table 4.1: Lateral deformation of single column (3 m = 10 ft height)


Displacement at
Number of
Column Size Elastic limit peak value of
Storey
lateral force (inch)
3 300 mm × 300 mm 14.89598 mm 43.96575
(three) (12// × 12//) (0.595839 inch) (1.75863 inch)
6 375 mm × 375 mm 12.11885 mm 41.68625 mm
(six) (15// × 15//) (0.484754 inch) (1.66745 inch)
9 450 mm × 450 mm 8.930025 mm 35.77125 mm
(nine) (18// × 18//) (0.3572 inch) (1.43085 inch)
12 525 mm × 525 mm 9.063225 mm 33.376 mm
(twelve) (21// × 21//) (0.362529 inch) (1.33504 inch)

74
-ced with 3.5% steel with a tie spacing 10 mm @ 300 mm c/c (= #3 @ 12 inch c/c).
Since every column, in real situation, has to carry loads from upper floors – so, a
concentrated force (1007.5239 kN for 300 mm square column, 2015.0403 kN for 375
mm square column, 3102.6302 kN for 450 mm square column and 4030.0838 kN for
525 mm square column) in vertical downward direction is also applied. For all these,
stiffness increases linearly upto certain limit (elastic range) and then it continues to
degrades due progressive cracking and damage (Figure 4.1 – 4.10). Ratio of inelastic
to elastic stiffness varies in between 0.24 to 0.30 (Table 4.2).

Table 4.2: Lateral stiffness of single column (3 m = 10 ft height)


Ratio of Inelastic
Column Size Elastic Stiffness Inelastic Stiffness Stiffness to Elastic
Stiffness
300 mm × 300 mm 1.626 kN/mm 0.4371 kN/mm 0.265495
(12// × 12//) (9138.24 lb./inch) (2624.16 lb./inch)
375 mm × 375 mm 4.8707 kN/mm 1.2833 kN/mm 0.2644493
(15// × 15//) (27357 lb./inch) (7250.4 lb./inch)
450 mm × 450 mm 12.161 kN/mm 3.6556 kN/mm 0.300600
(18// × 18//) (67831 lb./inch) (20390 lb./inch)
525 mm × 525 mm 21.859 kN/mm 5.247 kN/mm 0.240030
(21// × 21//) (122851 lb./inch) (29489 lb./inch)

40
35
30
lateral force (kN)

25
20
15
10
5
300 mm x 300 mm column
0
0 20 40 60 80 100
lateral displacement (mm)

75
Figure 4.1: Lateral force vs. lateral displacement at column top for 300 mm square
column up to collapse.
40
elastic
inelastic
30 y = 0.4317x + 18.752
lateral force (kN)
R2 = 0.9542

20
y = 1.626x
R2 = 0.9957
10

0
0 10 20 30 40 50
lateral displacement (mm)

Figure 4.2: Lateral force vs. lateral displacement at column top for 300 mm square
column up to ultimate value.

100

80
lateral force (kN)

60

40

20
375 mm x 375 mm column
0
0 20 40 60 80 100
lateral displacement (mm)

76
Figure 4.3: Lateral force vs. lateral displacement at column top for 375 mm square
column up to collapse.

120
110 elastic
100 inelastic
90 y = 1.2883x + 48.722
lateral force (kN)
80
R2 = 0.9456
70
60
50
40 y = 4.8707x
30 R2 = 0.9913
20
10
0
0 10 20 30 40 50 60
lateral displacement (mm)

Figure 4.4: Lateral force vs. lateral displacement at column top for 375 mm square
column up to ultimate value.

220
200
180
160
lateral force (kN)

140
120
100
80
60
40
20 450 mm x 450 mm column
0
0 10 20 30 40 50 60 70
lateral displacement (mm)
77
Figure 4.5: Lateral force vs. lateral displacement at column top for 450 mm square
column up to collapse.

220 elastic
200
inelastic
180
y = 3.6556x + 93.901
160
lateral force (kN)

140 R2 = 0.9254
120
100
80 y = 12.161x
60 R2 = 0.9987
40
20
0
0 10 20 30 40
lateral displacement (mm)

Figure 4.6: Lateral force vs. lateral displacement at column top for 450 mm square
column up to ultimate value.

350
300
lateral force (kN.)

250
200
150
100
50
525 mm x 525 mm column
0
0 10 20 30 40 50 60
lateral displacement (mm)

78
Figure 4.7: Lateral force vs. lateral displacement at column top for 525 mm square
column up to collapse.

330 elastic
300
inelastic
270 y = 5.247x + 173.76
240 R2 = 0.8899
lateral force (kN)

210
180
150
120 y = 21.859x
90 R2 = 0.9985
60
30
0
0 10 20 30 40
lateral displacement (mm)

Figure 4.8: Lateral force vs. lateral displacement at column top for 525 mm square
column up to ultimate value.

300 mm x 300 mm column


330 375 mm x 375 mm column
300 450 mm x 450 mm column
270 525 mm x 525 mm column
240
lateral force (kN)

210
180
150
120
90
60
30
0
0 10 20 30 40 50 60 70 80 90 100
lateral displacement (mm)
79
Figure 4.9: Lateral force vs. lateral displacement at column top for different column
sizes (upto collapse)

300 mm x 300 mm column-elastic portion


300 mm x 300 mm column-inelastic portion
330 375 mm x 375 mm column-elastic portion
300 375 mm x 375 mm column-inelastic portion
450 mm x 450 mm column-elastic portion
270 450 mm x 450 mm column-inelastic portion
lateral force (kN)

240 525 mm x 525 mm column-elastic portion


525 mm x 525 mm column-inelastic portion
210
180
150
120
90
60
30
0
0 10 20 30 40 50
lateral displacement (mm)

Figure 4.10: Lateral force vs. lateral displacement at column top for different column
sizes (upto ultimate value)

4.2.2 Lateral Stiffness of Masonry Wall


In order to obtain lateral stiffness, lateral displacement applied at wall top (bottom
fixed) which continues to increases. As masonry is a brittle one, failure will occur
suddenly. Different wall panels (125 mm thick wall is considered common for
partition walls in residential buildings) of height 2.4 meter was considered. For all
panels, stiffness value remains almost same (Figure 4.11 – 4.12). So, only 6 meter
long wall panel is considered.

80
3 m x 2.4 m x 125 mm panel
110 4.5 m x 2.4 m x 125 mm panel
100 6 m x 3 m x 125 mm panel
7.5 m x 2.4 m x 125 mm panel
90 9 m x 2.4 m x 125 mm panel
80
lateral force (kN)
70
60
50
40
30
20
10
0
0 0.5 1 1.5 2 2.5 3 3.5 4
lateral displacement (mm)

Figure 4.11: Lateral force vs. lateral displacement at wall top for different wall sizes
(upto collapse)
45
40
35
lateral force (kN)

30
25
20
15 3 m x 2.4 m x 125 mm panel
10 4.5 m x 2.4 m x 125 mm panel
6 m x 2.4 m x 125 mm panel
5 7.5 m x 2.4 m x 125 mm panel
9 m x 2.4 m x 125 mm panel
0
0 0.05 0.1 0.15 0.2 0.25
lateral displacement (mm)

81
Figure 4.12: Lateral force vs. lateral displacement at wall top for different wall sizes
(value upto 1st masonry cracking).

Table 4.3: Lateral stiffness and deformation of masonry wall [125 mm (= 5 inch
thick) and 2.4 m (= 8 ft) height]
Elastic limit
(Deformation at
Panel dimension Aspect ratio Elastic Stiffness
Masonry
Cracking)
3 m×2.4 m×125 mm 3 m/2.4 m 184.086601 kN/mm 0.139892 mm
(10/ × 8 / × 5//) = 1.25 (1034608.25 lb./inch) (0.005596 inch)
4.5 m×2.4 m×25 mm 4.5 m/2.4 m 204.285308 kN/mm 0.2177165 mm
(15/ × 8 / × 5//) = 1.875 (1148129.52 lb./inch) (0.0087087 inch)
6 m×2.4 m×25 mm 6 m/2.4 m 211.325392 kN/mm 0.2089253 mm
(20/ × 8 / × 5//) = 2.5 (1187696.38 lb./inch) (0.008357 inch)
7.5 m×2.4 m×125 mm 7.5 m/2.4 m 213.586433 kN/mm 0.197752 mm
(25/ × 8 / × 5//) = 3.125 (1200403.94 lb./inch) (0.007910 inch)
9 m×2.4 m×125 mm 9 m/2.4 m 213.838022 kN/mm 0.20317 mm
(30/ × 8 / × 5//) = 3.75 (1201817.93 lb./inch) (0.008127 inch)

80
70
60
lateral force (kN)

50
40
30
20
10 6 m x 2.4 m x 125 mm panel
0
0 0.5 1 1.5 2 2.5
lateral displacement (mm)

82
Figure 4.13: Lateral force vs. lateral displacement at wall top for 6 m × 2.4 m × 125
mm masonry panel (upto collapse).

lateral force (kN) 40

30 y = 211.33x
R2 = 0.9965

20

10

0
0 0.05 0.1 0.15 0.2 0.25
lateral displacement (mm)

Figure 4.14: Lateral force vs. lateral displacement at wall top for 6 m × 2.4 m × 125
mm masonry panel (value upto 1st masonry cracking).

Table 4.4: Ratio of frame stiffness to column stiffness at different levels for 3 (three)
storied building.
Number of Frame stiffnes
Storey level Frame Type
storey Column stiffness
3 1st storey Bare frame 1*
Infilled frame 98.47481181**
Soft storey frame 1*
2nd storey Bare frame 1

83
Infilled frame 98.47481181
Soft storey frame 98.47481181
3rd storey Bare frame 1
Infilled frame 98.47481181
Soft storey frame 98.47481181

Frame stiffnes  Column stiffnes =  Wall stiffness


* =
Column stiffness  Column stiffness

4  1.626 = 3  0
= =1
4  1.626

Frame stiffnes  Column stiffnes =  Wall stiffness


** =
Column stiffness  Column stiffness

4  1.626 = 3  211.325
= = 98.4748
4  1.626

Storey stiffnes without infill 4  1.626


[ =  100% = 1.015488%; which is
Storey stiffnes with infill 4  1.626 = 3  211.325

less than 60% (extreme soft-storey case)]

Table 4.5: Ratio of frame stiffness to column stiffness at different levels for 6 (six)
storied building.
Number of Frame stiffnes
Storey level Frame Type
storey Column stiffness
6 1st storey Bare frame 1*
Infilled frame 33.54030098**

84
Soft storey frame 1*
2nd storey Bare frame 1
Infilled frame 33.54030098
Soft storey frame 33.54030098
3rd storey Bare frame 1
Infilled frame 33.54030098
Soft storey frame 33.54030098
4th storey Bare frame 1
Infilled frame 33.54030098
Soft storey frame 33.54030098
5th storey Bare frame 1
Infilled frame 33.54030098
Soft storey frame 33.54030098
6th storey Bare frame 1
Infilled frame 33.54030098
Soft storey frame 33.54030098

Frame stiffnes  Column stiffnes =  Wall stiffness


* =
Column stiffness  Column stiffness

4  4.8707 = 3  0
= =1
4  4.8707

Frame stiffnes  Column stiffnes =  Wall stiffness


** =
Column stiffness  Column stiffness

4  4.8707 = 3  211.325
= = 33.5403
4  4.8707

Storey stiffnes without infill 4  4.8707


[ =  100% = 2.98148%; which
Storey stiffnes with infill 4  4.8707 = 3  211.325

is less than 60% (extreme soft-storey case)]

Table 4.6: Ratio of frame stiffness to column stiffness at different levels for 9 (nine)
storied building.

85
Number of Frame stiffnes
Storey level Frame Type
storey Column stiffness
9 1st storey Bare frame 1*
Infilled frame 14.03297788**
Soft storey frame 1*
2nd storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
3rd storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
4th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
5th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
6th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
7th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
8th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788
9th storey Bare frame 1
Infilled frame 14.03297788
Soft storey frame 14.03297788

Frame stiffnes  Column stiffnes =  Wall stiffness


* =
Column stiffness  Column stiffness

4  12.161 = 3  0
= =1
4  12.161

Frame stiffnes  Column stiffnes =  Wall stiffness


** =
Column stiffness  Column stiffness

4  12.161 = 3  211.325
= = 14.0329
4  12.161

Storey stiffnes without infill 4  12.161


[ =  100% = 7.12611%; which is
Storey stiffnes with infill 4  12.161 = 3  211.325

less than 60% (extreme soft-storey case)]

86
Table 4.7: Ratio of frame stiffness to column stiffness at different levels for 12
(twelve) storied building.
Number of Frame stiffnes
Storey level Frame Type
storey Column stiffness
12 1st storey Bare frame 1*
Infilled frame 8.250745414**
Soft storey frame 1*
2nd storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
3rd storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
4th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
5th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
6th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
7th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
8th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
9th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
10th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
11th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414
12th storey Bare frame 1
Infilled frame 8.250745414
Soft storey frame 8.250745414

Frame stiffnes  Column stiffnes =  Wall stiffness


* =
Column stiffness  Column stiffness

4  21.859 = 3  0
= =1
4  21.859

87
Frame stiffnes  Column stiffnes =  Wall stiffness
** =
Column stiffness  Column stiffness

4  21.859 = 3  211.325
= = 8.2507
4  21.859

Storey stiffnes without infill 4  21.859


[ =  100% = 12.12018%; which
Storey stiffnes with infill 4  21.859 = 3  211.325

is less than 60% (extreme soft-storey case)]

4.3 NATURAL FREQUENCY DETERMINATION


Response of building frame to ground motion depends upon the natural frequency of
the frame itself. The natural frequency of building frame depends on various
parameters. Frame height, width primarily controls natural frequency. Empirically,
the natural frequency is 10 / n , where n is the number of story of the building. Natural
frequency further depends upon infill condition of the frame. Frames, of same height,
with soft story, with infill and with no infill certainly possess different natural
frequencies, as infill contributes stiffness of frame.

0.05
1st storey
0.04 2nd storey
3rd storey
0.03

0.02
Inter Storey Drift(meter)

0.01

-0.01

-0.02

-0.03

-0.04

-0.05
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.15: Free vibration – 3 storey – bare frame

88
-4
x 10
1
1st storey
0.8 2nd storey
3rd storey
0.6

0.4

Inter Storey Drift(meter) 0.2

-0.2

-0.4

-0.6

-0.8

-1
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.16: Free vibration – 3 storey – infilled frame

0.04
1st storey
2nd storey
0.03 3rd storey

0.02
Inter Storey Drift(meter)

0.01

-0.01

-0.02

-0.03

-0.04
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.17: Free vibration – 3 storey – soft storey frame

89
0.015
1st storey
2nd storey
3rd storey
0.01
4th storey
5th storey
6th storey
Inter Storey Drift(meter)
0.005

-0.005

-0.01

-0.015
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.18: Free vibration – 6 storey – bare frame

-4
x 10
8
1st storey
2nd storey
6 3rd storey
4th storey
4 5th storey
6th storey
Inter Storey Drift(meter)

-2

-4

-6

-8
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.19: Free vibration – 6 storey – infilled frame

90
0.1
1st storey
0.08 2nd storey
3rd storey
0.06 4th storey
5th storey
0.04 6th storey

Inter Storey Drift(meter) 0.02

-0.02

-0.04

-0.06

-0.08

-0.1
0 1 2 3 4 5 6 7 8 9 10
Time(sec)
Figure 4.20: Free vibration – 6 storey – soft storey frame

To determine natural frequency of frame free vibration is required. For this, first some
initial ground movement is applied. After a certain time external excitation (ground
movement) stops, then frame continued to oscillate in a constant manner. Here
building frame has no additional damping. The results obtain from MATLAB is
presented in table 4.8.

Table 4.8: Natural frequency for different types of frames


Natural Time Natural Frequency
No. of Storey Type of Frame
Period (sec) (Hz.)
3 Bare Frame 0.52 1.92308
Infilled Frame 0.06 16.6667
Soft Storey 0.40 2.5
6 Bare Frame 0.96 1.04167
Infilled Frame 0.10 10
Soft Storey 0.56 1.78571
9 Bare Frame 1.4 0.71429
Infilled Frame 0.38 2.63158
Soft Storey 0.74 1.35135
12 Bare Frame 1.8 0.55556
Infilled Frame 0.64 1.5625
Soft Storey 0.88 1.13636
15 Bare Frame 2.26 0.44248
Infilled Frame 0.92 1.08696
Soft Storey 1.12 0.89286

91
Determination of natural frequency is important, because, response of structure to
ground motion depends mainly on frequency ratio (ratio of ground frequency and
natural frequency). Natural frequencies have been decreased with the increases of
storey no.. Natural frequency of frame with soft storey (ground storey has no infill)
falls in between bare and infilled frame (Figure 4.21). Empirical 10/n for
determination of natural frequency is valid for bare frame only.
18
Bare Frame
16 Infilled Frame
Natural Frequency (Hz.)

14 Soft-Storey Frame
10/n
12
10
8
6
4
2
0
0 3 6 9 12 15
Number of Storey

Figure 4.21: Natural frequency vs. Number of storey for various types of frames.

4.4 RESPONSE OF DIFFERENT TYPES OF FRAMES


Inter storey drift vs. time has been plotted, using the proposed program, has been
plotted for building frames of different infill conditions for several ground motion
parameters. As inelastic analysis with 5% damping is considered, the response pattern
does not follow any regular patters (as in elastic analysis), because structure’s
stiffness is continuous changing with time. This is true for all storey and buildings
types considered.

92
1
1st storey
2nd storey
0.8
3rd storey

0.6

Inter Storey Drift(meter)


0.4

0.2

-0.2

-0.4

-0.6

-0.8
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.22: Inter storey drift vs. time for 3 storied bare frame
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 3
Type of Frame: Bare
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.923
-4
x 10
1
1st storey
0.8 2nd storey
3rd storey
0.6

0.4
Inter Storey Drift(meter)

0.2

-0.2

-0.4

-0.6

-0.8

-1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.23: Inter storey drift vs. time for 3 storied infilled frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 3
Type of Frame: Infilled
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 16.667

93
0.15
1st storey
2nd storey
0.1 3rd storey

0.05

Inter Storey Drift(meter)


0

-0.05

-0.1

-0.15

-0.2
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.24: Inter storey drift vs. time for 3 storied soft storey frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 3
Type of Frame: Soft Storey
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 2.50
0.1
1st storey
2nd storey
0.08
3rd storey
4th storey
0.06 5th storey
6th storey
Inter Storey Drift(meter)

0.04

0.02

-0.02

-0.04

-0.06

-0.08
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.25: Inter storey drift vs. time for 6 storied bare frame.

Analysis Type: Inelastic


Damping: 5%
No. of Storey: 6
Type of Frame: Bare
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.041

94
-3
x 10
4
1st storey
2nd storey
3 3rd storey
4th storey
5th storey
2 6th storey

Inter Storey Drift(meter)


1

-1

-2

-3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.26: Inter storey drift vs. time for 6 storied infilled frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 6
Type of Frame: Infilled
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 10

0.5
1st storey
2nd storey
0.4 3rd storey
4th storey
0.3 5th storey
6th storey
Inter Storey Drift(meter)

0.2

0.1

-0.1

-0.2

-0.3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.27: Inter storey drift vs. time for 6 storied soft storey frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 6
Type of Frame: Soft
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.786

95
0.12
1st storey
0.1 2nd storey
3rd storey
0.08 4th storey
5th storey
0.06 6th storey

Inter Storey Drift(meter)


7th storey
0.04 8th storey
9th storey
0.02

-0.02

-0.04

-0.06

-0.08
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.28: Inter storey drift vs. time for 9 storied bare frame.

Analysis Type: Inelastic


Damping: 5%
No. of Storey: 9
Type of Frame: Bare
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 0.714
0.03

0.02

0.01
Inter Storey Drift(meter)

-0.01 1st storey


2nd storey
3rd storey
-0.02 4th storey
5th storey
6th storey
-0.03 7th storey
8th storey
9th storey
-0.04
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)

Figure 4.29: Inter storey drift vs. time for 9 storied infilled frame.

96
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 9
Type of Frame: Infilled
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.351
0.15

0.1

0.05
Inter Storey Drift(meter)

-0.05 1st storey


2nd storey
3rd storey
-0.1 4th storey
5th storey
6th storey
-0.15 7th storey
8th storey
9th storey
-0.2
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.30: Inter storey drift vs. time for 9 storied soft storey frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 9
Type of Frame: Soft
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.351
0.15

0.1

0.05
Inter Storey Drift(meter)

0 1st storey
2nd storey
3rd storey
-0.05 4th storey
5th storey
6th storey
-0.1 7th storey
8th storey
9th storey
-0.15 10th storey
11th storey
12th storey
-0.2
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)

97
Figure 4.31: Inter storey drift vs. time for 12 storied bare frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 12
Type of Frame: Bare
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 0.556
0.06

0.04
Inter Storey Drift(meter)

0.02

1st storey
2nd storey
0 3rd storey
4th storey
5th storey
-0.02 6th storey
7th storey
8th storey
-0.04 9th storey
10th storey
11th storey
12th storey
-0.06
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)

Figure 4.32: Inter storey drift vs. time for 12 storied bare frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 12
Type of Frame: Infilled
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 0.556

98
0.1
1st storey
2nd storey
0.08 3rd storey
4th storey
0.06 5th storey
6th storey

Inter Storey Drift(meter)


7th storey
0.04 8th storey
9th storey
0.02 10th storey
11th storey
12th storey
0

-0.02

-0.04

-0.06
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time(sec)
Figure 4.33: Inter storey drift vs. time for 12 storied soft storey frame.
Analysis Type: Inelastic
Damping: 5%
No. of Storey: 12
Type of Frame: Soft
Peak Ground Acceleration (PGA in m/s2): 2
Input Frequency (Hz.): 1.5
Natural Frequency (Hz.): 1.136

4.5 COMPARISON OF INTER-STOREY DRIFT RATIO OF DIFFERENT


TYPES OF FRAMES
From the previous plots, maximum inter story drift has been picked and inter story
drift ratio (inter story drift divided by storey height) plotted against frequency and
frequency ratio for different ground accelerations.

4.5.1 Three (3) Storied Frames


Bare frame is vulnerable to frequency range of 1.5 to 2 Hz. (frequency ratio of 0.78 to
1.04) and soft storey frame of 1.75 to 2.25 Hz. (frequency ratio of 0.91 to 1.17) for all
PGA values. Infilled frame is not vulnerable within the frequency range considered.
Trend is similar for all PGA values (Figure 4.34 – 4.39); higher PGA value gives
higher value of drift and vice versa.

99
0.3
Bare Frame
0.25 Infilled Frame

Interstorey Drift Ratio


Soft-Storey Frame
0.2

0.15

0.1

0.05

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)
Figure 4.34: Interstorey Drift Ratio vs. Frequency for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

0.3

0.25
Interstorey Drift Ratio

0.2

0.15 Bare Frame


Infilled Frame
0.1 Soft-Storey Frame

0.05

0
0 0.5 1 1.5 2
Frequency Ratio

Figure 4.35: Interstorey Drift Ratio vs. Frequency Ratio for 3 (three) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

100
0.35
Bare Frame
0.3 Infilled Frame

Interstorey Drift Ratio


Soft-Storey Frame
0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.36: Interstorey Drift Ratio vs. Frequency for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.
0.35
0.3
Bare Frame
Infilled Frame
Interstorey Drift Ratio

Soft-Storey Frame
0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2
Frequency Ratio

Figure 4.37: Interstorey Drift Ratio vs. Frequency Ratio for 3 (three) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

101
0.45 Bare Frame
0.4 Infilled Frame

Interstorey Drift Ratio


0.35 Soft-Storey Frame
0.3
0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.38: Interstorey Drift Ratio vs. Frequency for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.
0.45
0.4
Bare Frame
Infilled Frame
Interstorey Drift Ratio

0.35 Soft-Storey Frame


0.3
0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2
Frequency Ratio

102
Figure 4.39: Interstorey Drift Ratio vs. Frequency Ratio for 3 (three) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

4.5.2 Six (6) Storied Frames


Bare frame is vulnerable to frequency range of 1 to 1.5 Hz. (frequency ratio of 0.96 to
1.44) and soft storey frame of 0.5 to 0.75 Hz. (frequency ratio of 0.27 to 0.42) for all
PGA values. Infilled frame is not vulnerable within the frequency range considered
(Figure 4.40 – 4.45).
0.6

0.5
Interstorey Drift Ratio

0.4 Bare Frame


Infilled Frame
0.3
Soft-Storey Frame
0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.40: Interstorey Drift Ratio vs. Frequency for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

103
0.6
Bare Frame
0.5 Infilled Frame
Soft-Storey Frame

Interstorey Drift Ratio


0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3
Frquency Ratio (Hz.)
Figure 4.41: Interstorey Drift Ratio vs. Frequency Ratio for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

2
Bare Frame
1.8
Infilled Frame
1.6
Soft-Storey Frame
Interstorey Drift Ratio

1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.42: Interstorey Drift Ratio vs. Frequency for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

104
6
Bare Frame
5 Infilled Frame

Interstorey Drift (meter)


Soft-Storey Frame
4

0
0 0.5 1 1.5 2 2.5 3
Frequency Ratio (Hz.)

Figure 4.43: Interstorey Drift Ratio vs. Frequency Ratio for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

4.5
Bare Frame
4
Infilled Frame
3.5 Soft-Storey Frame
Interstorey Drift Ratio

3
2.5
2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

105
Figure 4.44: Interstorey Drift Ratio vs. Frequency for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.
4.5
4
3.5
Bare Frame
Infilled Frame

Interstorey Drift Ratio


Soft-Storey Frame
3
2.5
2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3
Frquency Ratio (Hz.)

Figure 4.45: Interstorey Drift Ratio vs. Frequency Ratio for 6 (six) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

4.5.3 Nine (9) Storied Frames


Bare frame is vulnerable to frequency range of 0.25 to 0.75 Hz. (frequency ratio of
0.35 to 1.05) and soft storey frame of 0.75 to 1.25 Hz. (frequency ratio of 0.55 to
0.93) for all PGA values. Infilled frame is not vulnerable within the frequency range
considered (Figure 4.46 – 4.51).

106
0.35
Bare Frame
0.3 Infilled Frame
Soft-Storey Frame

Interstorey Drift Ratio


0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.46: Interstorey Drift Ratio vs. Frequency for 9 (nine) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

0.35
Bare Frame
0.3 Infilled Frame
Soft-Storey Frame
Interstorey Drift Ratio

0.25
0.2
0.15
0.1
0.05
0
0 1 2 3 4 5
Frequency Ratio

107
Figure 4.47: Interstorey Drift Ratio vs. Frequency Ratio for 9 (nine) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

0.7
Bare Frame
0.6 Infilled Frame
Soft-Storey Frame
Interstorey Drift Ratio

0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.48: Interstorey Drift Ratio vs. Frequency for 9 (nine) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.
0.7
Bare Frame
0.6 Infilled Frame
Soft-Storey Frame
Interstorey Drift Ratio

0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5
Frequency Ratio

108
Figure 4.49: Interstorey Drift Ratio vs. Frequency Ratio for 9 (nine) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

0.7
Bare Frame
0.6 Infilled Frame
Soft-Storey Frame
Interstorey Drift Ratio

0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.50: Interstorey Drift Ratio vs. Frequency for 9 (nine) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.
0.7
Bare Frame
0.6 Infilled Frame
Soft-Storey Frame
Interstorey Drift Ratio

0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5
Frequency Ratio

109
Figure 4.51: Interstorey Drift Ratio vs. Frequency Ratio for 9 (nine) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

4.5.4 Twelve (12) Storied Frames


Bare frame is vulnerable to frequency range of 0.25 to 0.75 Hz. (frequency ratio of
0.45 to 1.35) and soft storey frame of 0.5 to 1 Hz. (frequency ratio of 0.44 to 0.88) for
all PGA values. Infilled frame is not vulnerable within the frequency range considered
(Figure 4.52 – 4.57).

0.45
0.4
0.35
Interstorey Drift Ratio

0.3
0.25
Bare Frame
0.2
Infilled Frame
0.15 Soft-Storey Frame
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.52: Interstorey Drift Ratio vs. Frequency for 12 (twelve) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

110
0.45
0.4
0.35
Interstorey Drift Ratio 0.3
0.25
Bare Frame
0.2
Infilled Frame
0.15 Soft-Storey Frame
0.1
0.05
0
0 1 2 3 4 5 6
Frequency Ratio

Figure 4.53: Interstorey Drift Ratio vs. Frequency Ratio for 12 (twelve) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

0.9
0.8
0.7
Interstorey Drift Ratio

0.6
0.5
0.4 Bare Frame
Infilled Frame
0.3
Soft-Storey Frame
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

111
Figure 4.54: Interstorey Drift Ratio vs. Frequency for 12 (twelve) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.
0.9
0.8
Interstorey Drift Ratio 0.7
0.6
0.5
0.4 Bare Frame
Infilled Frame
0.3
Soft-Storey Frame
0.2
0.1
0
0 1 2 3 4 5 6
Frequency Ratio

Figure 4.55: Interstorey Drift Ratio vs. Frequency Ratio for 12 (twelve) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

2.5
Bare Frame
Infilled Frame
2
Soft-Storey Frame
Interstorey Drift Ratio

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

112
Figure 4.56: Interstorey Drift Ratio vs. Frequency for 9 (nine) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.
2.5
Bare Frame
Infilled Frame
2
Soft-Storey Frame
Interstorey Drift Ratio

1.5

0.5

0
0 1 2 3 4 5
Frequency Ratio

Figure 4.57: Interstorey Drift Ratio vs. Frequency Ratio for 9 (nine) – Storied
Building Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

4.6 COMPARISON OF COLLAPSE TIME OF DIFFERENT TYPES OF


FRAMES
A frame is considered to be failed if it’s interstorey drift ratio equal to 0.02 (life safety
level, ATC-40). From the time response plots, time needed to reach 0.02 has been
picked and plotted against frequency and frequency ratio for different ground
accelerations.

4.6.1 Three (3) Storied Frames


Minimum collapse time for bare frame 0.65 sec after shaking (when frequency 0.75
Hz. and PGA 1 m/s2), 0.24 sec after shaking (when frequency 1 Hz. and PGA 2 m/s2)
and 0.32 sec after shaking (when frequency 1.25 Hz. and PGA 3 m/s2). Minimum
collapse time for soft storey frame 0.58 sec after shaking (when frequency 0.75 Hz.
and PGA 1 m/s2), 0.36 sec after shaking (when frequency 2.75 Hz. and PGA 2 m/s2)

113
and 0.22 sec after shaking (when frequency 1.75 Hz. and PGA 3 m/s2). Infilled frame
does not collapse within the frequency range considered (Figure 4.58 – 4.63).

4
Bare Frame
3.5 Soft-Storey Frame
3
Time to Fail (Sec)

2.5
2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3
Input Frequency (Hz.)

Figure 4.58: Collapse Time vs. Frequency for 3 (three) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 1 m/s2.

4
3.5 Bare Frame
Soft-Storey Frame
3
Time to Fail (Sec)

2.5
2
1.5
1
0.5
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Frequency Ratio

114
Figure 4.59: Collapse Time vs. Frequency Ratio for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

1
Bare Frame
0.9
Soft-Storey Frame
0.8
0.7

Time to Fail (Sec)


0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.60: Collapse Time vs. Frequency for 3 (three) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 2 m/s2.

1
Bare Frame
0.9
Soft-Storey Frame
0.8
0.7
Time to Fail (Sec)

0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2
Frequency Ratio

115
Figure 4.61: Collapse Time vs. Frequency Ratio for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

0.6
Bare Frame
0.5 Soft-Storey Frame

0.4
Time to Fail (Sec)

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.62: Collapse Time vs. Frequency for 3 (three) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 3 m/s2.

0.6
Bare Frame
0.5 Soft-Storey Frame

0.4
Time to Fail (Sec)

0.3

0.2

0.1

0
0 0.5 1 1.5 2
Frequency Ratio

Figure 4.63: Collapse Time vs. Frequency Ratio for 3 (three) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

116
4.6.2 Six (6) Storied Frames
Minimum collapse time for bare frame 0.5 sec after shaking (when frequency 0.75
Hz. and PGA 1 m/s2), 0.37 sec after shaking (when frequency 0.75 Hz. and PGA 2
m/s2) and 0.28 sec after shaking (when frequency 1.25 Hz. and PGA 3 m/s2).
Minimum collapse time for soft storey frame 0.42 sec after shaking (when frequency
1 Hz. and PGA 1 m/s2), 0.23 sec after shaking (when frequency 2 Hz. and PGA 2
m/s2) and 0.17 sec after shaking (when frequency 2.5 Hz. and PGA 3 m/s2). Infilled
frame does not collapse within the frequency range considered (Figure 4.64 – 4.69).

6
Bare Frame
5 Soft-Storey Frame

4
Time to Fail (Sec)

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.64: Collapse Time vs. Frequency for 6 (six) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 1 m/s2.

117
6
Bare Frame
5 Soft-Storey Frame

Time to Fail (Sec) 4

0
0 0.5 1 1.5 2 2.5 3
Frequency Ratio (Hz.)
Figure 4.65: Collapse Time vs. Frequency Ratio for 6 (six) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 1 m/s2.

0.8
Bare Frame
0.7 Soft-Storey Frame
0.6
Time to Fail (Sec)

0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.66: Collapse Time vs. Frequency for 6 (six) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 2 m/s2.

118
0.8
Bare Frame
0.7 Soft-Storey Frame
0.6

Time to Fail (Sec)


0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3
Frequency Ratio (Hz.)

Figure 4.67: Collapse Time vs. Frequency Ratio for 6 (six) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 2 m/s2.

0.7
Bare Frame
0.6 Soft-Storey Frame

0.5
Time to Fail (Sec)

0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

Figure 4.68: Collapse Time vs. Frequency for 6 (six) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 3 m/s2.

119
0.7
Bare Frame
0.6 Soft-Storey Frame

0.5

Time to Fail (Sec)


0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3
Frequency Ratio (Hz.)

Figure 4.69: Collapse Time vs. Frequency Ratio for 6 (six) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 3 m/s2.

4.6.3 Nine (9) Storied Frames


Minimum collapse time for bare frame 0.5 sec after shaking (when frequency 0.75
Hz. and PGA 1 m/s2), 0.36 sec after shaking (when frequency 1 Hz. and PGA 2 m/s2)
and 0.27 sec after shaking (when frequency 1.25 Hz. and PGA 3 m/s2). Minimum
collapse time for soft storey frame 0.36 sec after shaking (when frequency 1.25 Hz.
and PGA 1 m/s2), 0.22 sec after shaking (when frequency 2 Hz. and PGA 2 m/s2) and
0.18 sec after shaking (when frequency 2.25 Hz. and PGA 3 m/s2). Infilled frame does
not collapse within the frequency range considered (Figure 4.70 – 4.75).

3.5
Bare Frame
3
Soft-Storey Frame
2.5
Time to Fail (Sec)

2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

120
Figure 4.70: Collapse Time vs. Frequency for 9 (nine) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 1 m/s2.

3.5
Bare Frame
3
Soft-Storey Frame
2.5
Time to Fail (Sec)

2
1.5
1
0.5
0
0 1 2 3 4 5
Frequency Ratio
Figure 4.71: Failure Time vs. Frequency Ratio for 9 (nine) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 1 m/s2.

2
1.8 Bare Frame
1.6 Soft-Storey Frame
1.4
Time to Fail (Sec)

1.2
1
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)
Figure 4.72: Collapse Time vs. Frequency for 9 (nine) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 2 m/s2.

121
2
1.8 Bare Frame
1.6 Soft-Storey Frame
1.4
Time to Fail (Sec) 1.2
1
0.8
0.6
0.4
0.2
0
0 1 2 3 4 5
Frequency Ratio
Figure 4.73: Collapse Time vs. Frequency Ratio for 9 (nine) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 2 m/s2.

0.5
0.45
0.4
0.35
Time to Fail (Sec)

0.3 Bare Frame


0.25 Soft-Storey Frame
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)
Figure 4.74: Collapse Time vs. Frequency for 9 (nine) – Storied Building Frame with
Peak Ground Acceleration (PGA) of 3 m/s2.

122
0.5
0.45
0.4
0.35
Time to Fail (Sec) 0.3
0.25
0.2
0.15
0.1 Bare Frame
0.05 Soft-Storey Frame
0
0 1 2 3 4 5
Frequency Ratio
Figure 4.75: Collapse Time vs. Frequency Ratio for 9 (nine) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 3 m/s2.

4.6.4 Twelve (12) Storied Frames


Minimum collapse time for bare frame 0.49 sec after shaking (when frequency 0.75
Hz. and PGA 1 m/s2), 0.36 sec after shaking (when frequency 1 Hz. and PGA 2 m/s2)
and 0.28 sec after shaking (when frequency 1.25 Hz. and PGA 3 m/s2). Minimum
collapse time for soft storey frame 0.37 sec after shaking (when frequency 1 Hz. and
PGA 1 m/s2), 0.23 sec after shaking (when frequency 1.75 Hz. and PGA 2 m/s2) and
0.18 sec after shaking (when frequency 2 Hz. and PGA 3 m/s2). Infilled frame does
not collapse within the frequency range considered (Figure 4.76 – 4.81).

4.5
Bare Frame
4 Soft-Storey Frame
3.5
3
Time to Fail (Sec)

2.5
2
1.5
1
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)

123
Figure 4.76: Collapse Time vs. Frequency for 12 (twelve) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 1 m/s2.

4.5
Bare Frame
4 Soft-Storey Frame
3.5
3
Time to Fail (Sec)

2.5
2
1.5
1
0.5
0
0 1 2 3 4 5 6
Frequency Ratio
Figure 4.77: Collapse Time vs. Frequency Ratio for 12 (twelve) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 1 m/s2.

3
Bare Frame
2.5 Soft-Storey Frame

2
Time to Fail (Sec)

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)
Figure 4.78: Collapse Time vs. Frequency for 12 (twelve) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 2 m/s2.

124
3
Bare Frame
2.5 Soft-Storey Frame

2
Time to Fail (Sec)
1.5

0.5

0
0 1 2 3 4 5 6
Frequency Ratio
Figure 4.79: Collapse Time vs. Frequency Ratio for 12 (twelve) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 2 m/s2.

0.45
0.4
0.35
0.3
Time to Fail (Sec)

0.25
0.2
0.15
0.1
Bare Frame
0.05
Soft-Storey Frame
0
0 0.5 1 1.5 2 2.5 3 3.5
Input Frequency (Hz.)
Figure 4.80: Collapse Time vs. Frequency for 12 (twelve) – Storied Building Frame
with Peak Ground Acceleration (PGA) of 3 m/s2.

125
0.45
0.4
0.35
0.3
Time to Fail (Sec)
0.25
0.2
0.15
0.1
Bare Frame
0.05
Soft-Storey Frame
0
0 1 2 3 4 5
Frequency Ratio
Figure 4.81: Collapse Time vs. Frequency Ratio for 12 (twelve) – Storied Building
Frame with Peak Ground Acceleration (PGA) of 3 m/s2.

1.2

1
Time to Fail (Sec)

0.8

0.6

0.4
Bare Frame

0.2 Soft-Storey Frame

0
0 0.5 1 1.5 2 2.5 3 3.5
PGA (m/s^2)

Figure 4.82: Collapse Time vs. Peak Ground Acceleration (PGA) (for 3 Storied
Building with Earthquake Frequency of 1.5 Hz.)

126
0.7
0.6
0.5

Time to Fail (Sec)


0.4
0.3
0.2 Bare Frame

0.1 Soft-Storey Frame

0
0 0.5 1 1.5 2 2.5 3 3.5
PGA (m/s^2)

Figure 4.83: Collapse Time vs. Peak Ground Acceleration (PGA) (for 6 Storied
Building with Earthquake Frequency of 1.5 Hz.)

1
Bare Frame
0.8
Soft-Storey Frame
Time to Fail (Sec)

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3 3.5
PGA (m/s^2)

127
Figure 4.84: Collapse Time vs. Peak Ground Acceleration (PGA) (for 9 Storied
Building with Earthquake Frequency of 1.5 Hz.)

0.6

0.5

0.4
Time to Fail (Sec)

0.3

0.2
Bare Frame

0.1
Soft-Storey Frame

0
0 0.5 1 1.5 2 2.5 3 3.5
PGA (m/s^2)

Figure 4.85: Collapse Time vs. Peak Ground Acceleration (PGA) (for 12 Storied
Building with Earthquake Frequency of 1.5 Hz.)

1.2
Bare Frame
1 Soft-Storey Frame
Time to Fail (Sec)

0.8

0.6

0.4

0.2

0
0 2 4 6 8 10 12 14
Number of Storey

128
Figure 4.86: Collapse Time vs. Number of Storey (for Earthquake Frequency 1.5 Hz.
and Peak Ground Acceleration (PGA) of 1 m/s2).

1
Bare Frame
Soft-Storey Frame
0.8
Time to Fail (Sec)

0.6

0.4

0.2

0
0 2 4 6 8 10 12 14
Number of Storey

Figure 4.87: Collapse Time vs. Number of Storey (for Earthquake Frequency 1.5 Hz.
and Peak Ground Acceleration (PGA) of 2 m/s2).

0.6
Bare Frame

0.5 Soft-Storey Frame


Time to Fail (Sec)

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10 12 14
Number of Storey

129
Figure 4.88: Collapse Time vs. Number of Storey (for Earthquake Frequency 1.5 Hz.
and Peak Ground Acceleration (PGA) of 3 m/s2).

130
CHAPTER FIVE
CONCLUSIONS AND RECOMMENDATIONS

5.1 CONCLUSIONS
An extensive computational investigation has been performed on reinforced concrete
framed buildings having masonry infilled panels to identify the relative importance of
different types of structures. Focus is primarily given on reasonable estimation of the
frequency (or frequency ratio) of ground motion, which is an important parameter for
earthquake resistant design. Three, six, nine and twelve storied building frames of
different column sizes were analyzed for different infill conditions by the program
considering nonlinear material behavior with 5% damping. The findings of the study
presented in chapter 4 are summarized below:
(i) Natural frequency obtained after applying free vibration is different from that
obtained from code specified empirical formula. Natural frequency of soft
storey buildings falls within the range between full infill and bare frame. The
code empirical formula for natural time period holds good for bare frame as
well as for the soft-storey frame. The difference is obvious due to the presence
of masonry infill, which changes the mass and stiffness of the structures
(Figure 4.21).

(ii) From all the time-response analysis of different types of buildings, it was
observed that interstorey drift ratio is almost zero for infilled frame irrespective
of ground motion amplitude and frequency. Interstorey drift at 1st storey level
of soft-storey frame have a higher value than the bare frame for all cases and
also interstorey drift / ground floor displacement is the highest in soft storey
construction than other stories.

(iii) For 3 (three) storied building with excitation (f) = 1.5 Hz., when PGA = 0.1g to
0.25g, the bare frame fails before the soft-storey frame; when PGA = 0.25g to
0.3g, the soft-storey frame fails before the bare frame (Figure 4.82).

(iv) For 6 (six) storied building with excitation (f) = 1.5 Hz., when PGA = 0.1g to
0.125g, the bare frame fails before the soft-storey frame; when PGA = 0.125g
to 0.3g, the soft-storey frame fails before the bare frame (Figure 4.83).

127
(v) For 9 (nine) and 12 (twelve) storied building with excitation (f) = 1.5 Hz., the
soft-storey frame fails before the bare frame irrespective of any PGA values
(Figure 4.84 – 4.85).

(vi) For storey number less than or equal to 6 (six), with PGA = 0.1g and excitation
(f) = 1.5 Hz., the bare frame fails before the soft-storey frame. With the same
excitation and storey number greater than 6 (six), soft-storey frame fails first
(Figure 4.86).

(vii) For storey number less than or equal to 5 (five), with PGA = 0.2g and
excitation (f) = 1.5 Hz., the bare frame fails before the soft-storey frame. With
the same excitation and storey number greater than 5 (five), soft-storey frame
fails first (Figure 4.87).

(viii) For any number of storey, with PGA > 0.3g and excitation (f) = 1.5 Hz., the
soft-storey frame fails first (Figure 4.88).

Vulnerability of structure not only depends on the number of storey but also on
ground motion frequency and it’s amplitude. For all the cases, frame with full infill
performs the best. In high PGA values, soft-storey becomes more vulnerable as
collapse time is shortened compared to the bare frame, irrespective of frequency.
Increasing number of storey makes the soft storey building more vulnerable than the
bare frame.

5.2 RECOMMENDATION FOR FUTURE RESEARCH


i. Only plane frame is considered in this study. A more rigorous three-
dimensional study considering the out-of-plane strength and stiffness of
masonry can be carried out.
ii. The study was carried out for harmonic loads only. Arrangements should be
made to input more general and realistic time series or amplitude spectra of
seismic vibrations.
iii. Bay width is kept same in the adjacent span. Future work can be done bay
varying bay sizes.
iv. Only three types of frames are studied. As soft storey can occur any floor,
other frames having different infill orientation in the upper storey can be
investigated.

128
v. Only plain masonry infill behavior was observed. Detailed analysis could be
done using reinforced masonry infill considering effect of plastering.
vi. Column base is assumed to be fixed at bottom. In reality, it is surrounded by
soil – so, more work can be done considering soil-foundation-structure
interaction.
vii. Presence of openings within the infill due to doors and widows can be
incorporated.
viii. Axial stiffness and damping of beam may be considered.

129
REFERENCES

Akkar, S., Yazgan, U., and Gulkan, P. (2005), “Drift Estimates in Frame
Buildings Subjected to Near Fault Ground Motions”, Journal of Structural
Engineering, ASCE, Vol. 131, No. 7, pp. 1014–1024.

Alarcon, L. F. and Moehle, J. P. (1986), “Seismic Analysis Methods for Irregular


Buildings”, Journal of Structural Engineering, ASCE, Vol. 112(1), pp. 35–52.

Ali, S. and Page, A. W. (1985), "An Elastic Analysis of Concentrated Loads on


Brickwork", Masonry Int., Vol. 6, 1985, pp. 9-21.

Ali, S. and Page, A. W. (1988), "Concentrated Loads on solid Masonry Walls - A


Parametric Study and Design Recommendations", Proc. Inst. of Civil Engineers,
Voi. 85, 1988, pp. 271-289.

Ali, S. and Page, A. W. (1988), "Finite Element Model for Masonry Subjected to
Concentrated Loads", Journal Str. Div., Proc. ASCE, Vol. 114, No ST8, 1988, pp.
1761-84.

Ali, S. and Page, A. W. (1987),"Non-linear Finite Element Analysis of Masonry


subjected to Concentrated Loads", Proc. Inst. of Civil Engineers, Vol. 83, 1987,
pp. 815-832.

Arlekar, J. N., Jain S. K. and Murty, C. V. R. (1997), “Seismic Response of RC


Frame Buildings with Soft First Storeys”, Proceedings of the CBRI Golden
Jubilee Conference on Natural Hazards in Urban Habitat, 1997, New Delhi, India.

Arnold, C. & Elsesser, E. (1980), “Building Configurations: Problems &


Solutions”, Proceedings of the Seventh World Conference on Earthquake
Engineering, Vol. 4, pp. 153–160.

Arnold, C. & Reitherman, R. (1982), “Building Configuration & Seismic Design”,


John-Wiley & Sons, New York.

Asteris, P. G. (2003), “Lateral Stiffness of Brick Masonry Infilled Plane Frames”,


Journal of Structural Engineering, ASCE, Vol. 129, No. 8, pp. 1071–1079.

ATC – 40 (1996), “Seismic Evaluation and Retrofit of Concrete Buildings”,


Volumn I, Applied Technology Council, Redwwood City, California 94065, USA.

Bakhteri, J. and S. Sambasivam. (2003), “Mechanical Behavior of Structural


Brick Masonry: An Experimental Evaluation”, Proceeding of 5th Asia-Pacific
structural Engineering and construction conference. Malaysia. 305- 317.

130
Barzegar, F. & Maddipudi, S. (1997), “Three-Dimensional Modelling of Concrete
Structures, I: Plain Concrete”, ASCE Journal of Structural Engineering, Vol. 123,
No. 10, pp. 1339–1346.

Bell, D.K. and Davidson, B.J., (2001),“Evaluation of Earthquake Risk Buildings


with Masonry Infill Panel”, No.4.02.01 NZSEE 2001 Conference.

Bertero, V. & Brokken, S. (1983), “Infills in Seismic Resistant Buildings”, Journal


of Structural Engineering, ASCE, Vol. 109, No. 6, pp. 1337–1361.

Bertero, V. & Klingner, R. E. (1978), “Earthquake Resistance of Infilled Frames”,


Journal of Structural Engineering, ASCE, Vol. 104, No. ST6, pp. 973–989.

Beskos, D. E. and Anagnostoulos, S. A. (1977), “Advances in earthquake


Engineering – Computer Analysis and Design of Earthquake Resistant
Structures: A Handbook”, Computational Mechanics Publications,
Southampton, UK.

BNBC (1993), “Bangladesh National Building Code”, Housing and Building


Research Research Institute, Mirpur, Dhaka.

Bruneau, M. (1999). “Structural Damage: Kocaeli, Turkey Earthquake,


August 17, 1999”, MCEER Deputy Director and Professor, Department of
Civil, Structural and Environmental Engineering, University at Buffalo.

Building Seismic Safety Council, BSSC (1997), “NEHRP Recommended


Provisions for Seismic Regulations for Buildings and Other Structures”, Report
No. FEMA 302/303. Federal Emergency Management Agency, Washington, DC.

Building Seismic Safety Council, BSSC (2001), “NEHRP Recommended


Provisions for Seismic Regulations for Buildings and Other Structures”, Report
no. FEMA 368. Federal Emergency Management Agency, Washington, DC.

Buonopane, S. G. and White, R. N. (1999), “Pseudodynamic Testing of Masonry


Infilled Reinforced Concrete Frame”, Journal of Structural Engineering, ASCE,
Vol. 125, No. 6, pp. 578–589.

Chopra, A. K. (1987), “Simplified Earthquake Analysis of Buildings”, Proceedings


of the Sessions at Structures Congress’ 87 related to Dynamics of Structures, page
139-155.

Chopra, A. K., D. P. Clough, & Clough, R. W. (1973), “Earthquake Resistance of


Buildings with a Soft First Storey”, Journal of Earthquake Engineering &
Structural Dynamics, Vol. 1, No. 4, pp. 355.

131
Crisafulli, F. J., Carr, A. J. and Park, R. (2000), “Analytical Modeling of Infilled
Frame Structures – A General View”, Bulletin of the New Zealand Society for
Earthquake Engineering. Vol. 33, No. 1, pp. 30-47.

Crisafulli, F. J., and Carr, A. J. (2007), “Proposed Macro-Model for the Analysis
of Infilled Frame Structures”, Bulletin of the New Zealand Society for Earthquake
Engineering, Vol. 40, No. 2, pp. 69-77.

Clough, R. W. and Penzien, J. (1975), “Dynamics of Structures”, Mc-Graw Hill,


London.

Craig, R. R. (1981), “Structural Dymanics – An Introduction to Computer


Methods”, John Wiley & Sons.

Day, R. W. (2002), “Geotechnical Earthquake Engineering Handbook”, Mc-Graw


Hill, London.

Dhanasekar, M., Page, A. W., Kieeman P. W. (1985), "The Failure of Brick


Masonry under Biaxial Stresses", Proc. Inst. of Civil Engineers, Vol. 79, Part 2,
1985, pp. 295-313.

Dhanasekar, M., Page, A. W., Kieeman P. W. (1982), "The Elastic Properties of


Brick Masonry", Int. Joumal of Masonry Construction, Vol. 2, No 4, 1982, pp.
155- 160.

Dhanasekar, M., Page, A. W., Kieeman P. W. (1984), "A Finite Element Model
for the In-Plane Behavior of Brick Masonry", Proc. 9th Australasian Conference
on Mechanisms of Structures, 1984, pp. 262-267.

Dhanasekar, M., Page, A. W., Kieeman P. W. (1985), "Biaxial Stress-Strain


Relations for Brick Masonry", Journal of Structural Engineering, ASCE, Vol. 111,
No 5, 1985, pp. 1085-1100.

Dolsek M and Fajfar P (2001), “Soft storey effects in uniformly infilled


reinforced concrete frames”, Journal of Earthquake Engineering, Volume
5, No. 1, pp. 1–12.

Dowrick, J. D. (2003), “Earthquake Risk Reduction”, John Wiley and Sons, USA.

Ewing, R. D., Kariotis J. C., Englekirk R. E. and Hart, G. C. (1987), “Analytical


Modeling for Reinforced Masonry Building and Componets – TCCMAR
Category 2 Program”, Proceedings of the Third Meeting of the Joint Technical
Coordinating Committee on Masonry Research, Tomamau, Japan.

Fintel, M. & Khan, F. R. (1969), “Shock-Absorbing Soft Storey Concept for


Multistory Earthquake Structures”, ACI Journal, Vol. 66, No. 5, pp. 381–390.

132
Ganju, T. N. (1977), "Nonlinear Finite Element Analysis of Clay Brick Masonry",
Proc. 6th Australasian Conf. on Mech. of Structures and Materials, 1977, pp. 59-
65.

Guinea, G. V., Hussein, G., Elices, M. and Planas, J. (2000), “Micromechanical


Modeling of Brick-Masonry Fracture”, Cem. Concr. Res., 30, 731-737.

Güler, K. (1996), “Dynamic Behavior of a Building having Vertically Irregular


Structural System”, Proceedings of the European Workshop on the Seismic
Behavior of Asymmetric and Setback Structures, Capri, IT, pp. 267–278.

Gulkan, P., Ascheim, M., and Spence, R. (2002), “Reinforced Concrete Frame
Building with Masonry Infills”, WHE Report 64 (Turkey), World Housing
Encyclopedia, Earthquake Engineering Research Institute and International
Association for Earthquake Engineering.

Hart, G. C., Low, Y-K, Jaw J-W and Englekirk, R. E. (1989), “Structural
Componet Model for University of Colorado Flexural Walls”, U. S. – Japan
Coordinated Program for Masonry Building Research, Report No. 2.1-6,
December.

Hart, C. G. and Wong, K. (2000), “Structural Dynamics for Structural Engineers”,


John Wiley and Sons, New York, USA.

He, L. and Priestley, M. J. N. (1990), “Inelastic Structural Component Model for


Flanged Masonry Structural Walls”, Sixth Meeting of the U. S. – Japan Joint
Technical Coordinating Committee on Masonry Research, Seattle, Washington,
USA.

Healey, T. J. & Sozen, M. A. (1978), “Experimental Study of the Dynamic


Response of a Ten-Storey Reinforced Concrete Frame with a Tall First Storey”,
University of Illinois at Urbana-Champaign, (Report No. UILU-ENG-78-2012).

Hegemier, G. A., (1978)., "On the Behavior of Joints in Concrete Masonry", Proc.
North American Conference, Masonry Society, USA, 1978, pp. 4.1-4.21.

Hegemier, G. A., Nunn, R. O., Arya, S. K. (1978), "Behavior of Concrete


Masonry Under Biaxiai Stresses", Proc. North American Masonry Conference,
1978, pp. 1.1- 1.28.

Holmes, M. (1963), “Combined Loading on Infilled Frames”, Proceedings of the


Institution of the Institution of Engineers, Vol. 25, pp. 31–38.

133
Homes, M. (1961), “Steel Frames with Brick Work and Concrete Infilling”,
Procedings of Institute of Civil Engineers, London, England, Vol. 73, Part 2, pp.
473–478.

Hossain, M. M. (1997), “Non-Linear Finite Element Analysis of Wall-Beam


Structures”, Ph.D Thesis, BUET, BUET, Dhaka-1000.

Hurty, W. C. & Rubinstein M. F. (1967), “Dynamics of Structures”, Prentice Hall


of India Private Limited, New Delhi.

IBC (2000), “International Building Code”, International Code Council”, Falls


Church, VA, USA.

IS 1893–2002, “Indian Standard Code – Criteria for Earthquake Resistant


Design of Structures”, India.

Jain S. K. and Murthy, C. V. R. (1997), “Beneficial Influence of masonry Infills on


Sismic Performance of RC Frame Buildings”, Proceedings of 12th World
Conference on Earthquake Engineering, New Zealand, Paper no. 1790.

Kanitkar, R. and Kanitkar, V. (2004), “Seismic Performance of Conventional


Multi-Storey Buildings with Open Ground Floors for Vehicular Parking”, The
Indian Construction Journal, Vol. 78, No. 2, pp. 99–104.

Klingner, R. E (1980), “Mathematical Modelling of Infilled Frames”, Reinforced


Concrete Structures Subjected to Wind & Earthquake Forces, Publication SP-63,
American Concrete Institute, Detroit, pp. 1–25.

Lourenço, P. B. (1996), “Computational Strategies for Masonry Structures”, Ph.


D. thesis, Delft University of Technology, Delft, The Netherlands.

Lourenço, P. B., Rots, J. G. and Van Der Pluijm, R. (1999), “Understanding the
Tensile Behavior of Masonry Parallel to the Bed Joints: a Numerical Approach”,
Masonry International, 12(3), 96-103.

Lourenco, P. B., Rots, J. G., Blaauwendraad J. (1995), "Two Approaches for the
Analysis of Masonry Structures: Micro-and Macro-Modeling", Heron, Vol. 40,
No. 4, 1995, pp. 313-340.

Lourenco, P. B., Rots, J. S. , (1993), "On the use of micro-modelling for the
analysis of masonry shear-walls", Proceedings of the Second International
Symposium on Computer Methods in Structural Masonry, Wales, U. K., 6-8 April
1993, pp. 14-26.

134
Madan, A., Reinhorn, A. M., Mander, J. B. & Valles, R. E. (1997), “Modeling of
Masonry Infill Panels for Structural Analysis”, Journal of Structural Engineering,
ASCE, Vol. 123, No. 10, pp. 1295–1297.

Mander, J. B., Priestly, M. J. N and Park, R. (1988), “Theoretical Stress-Strain


Behaviour of Confined Concrete”, Journal of Structural Engineering, ASCE, Vol.
114, No. 8, pp. 1827–1849.

Martel, R. R. (1929), “Effects of Earthquakes on Buildings with a Flexible First


Storey”, Bulletin of the Seismological Society of America, Vol. 19.

Marzhan, G. (1998), “ The Shear Strength of Dry-Stacked Masonry Walls. Institüte


für Massivbau und Baustofftechnologie”, Universität Leipzig. Lacer No.3.

Mehrabi, A. B., Shing, P. B., Schuller, M. P. & Noland, J. N. (1996),


“Experimental Evaluation of Masonry Infilled RC Frames”, Journal of Structural
Engineering, ASCE, Vol. 122, No. 3, pp. 228–237.

Meirovitch, L. (1980), “Computational Methods in Structural Dynamics”, Sijthoff


& Noordohoff, Netherlands.

Meirovitch, L. (1997), “Principles and Techniques of Vibrations”, Prentice-Hall


International, Inc., New Jersey.

Mezzi, M. (2004), “Architectural and Structural Configuration of Buildings with


Innovative Seismic Systems”, 13th World Conference on Earthquake Engineering,
August (Paper No. 1318), Vancouver, B. C., Canada.

Minimum Design Loads for Buildings and Other Structures (2002) – American
Society of Civil Engineers (ASCE), Reston, Virginia, USA.

Moehle, J. P. & Sozen, M. A. (1978), “Earthquake Simulation Tests of a Ten-


Storey Reinforced Concrete Frame with a Discontinued First-Level Beam”,
University of Illinois at Urbana-Champaign, (Report No. UILU-ENG-78-2014).

Mogaddam, H. A. & Dowling, P. J. (1987), “The State of the Art in Infilled


Frames”. Imperial College of Science & technology, Civil Engineering
Department, London, UK, ESEE Report No. 87-2.

Moghaddam, H. (1994), “Seismic Design of Masonry Structures”, ISBN 964- 6379-


38-9.

135
Munshi, J. A. and Ghosh, S. K. (1995),“Analysis of Seismic Performance of a
Code Designed Reinforced Concrete Building”, Journal of Engineering
Structures, Elsevier Science Ltd., Vol. 20, No. 7, pp. 608–616.

Murthy, C. V. R. (2005), “Earthquake Tips – Learning Earthquake Design and


Construction”, National Information Centre of Earthquake Engineering, IIT,
Kanpur, India.

Newland, D. E. (1987), “On the Modal Analysis of Nonconservative Linear


Systems”, Journal of Sound and Vibration, 112 (1), pp. 69–96.

Newland, D. E. (1989), “Mechanical Vibration Analysis and Computation”,


Longman, Harlow and John Wiley, New York.

National Earthquake Hazard Reduction Program (1997), “NEHRP Provisions for


the Development of Seismic Regulations for Buildings – Working Draft”, NEHRP-
97, Washington, DC.

NISSE (2000), “National Information Service for Earthquake Engineering


Software Library”, University of California, Berkeley
(http://nisee.berkeley.edu/data/strong_motion/sacsteel/ground_motions.ht
ml).

NZS 4203 (1992), “General Structural Design and Design Loading for
Buildings”, Wellington, New Zealand.

Page, A. W. (1978), "Finite Element Model for Masonry", Journal Str. Div., Proc.
ASCE, Vol. 104, No ST8, 1978, pp. 1267-1285.

Paulay, T. and Priestley, M. J. N (1992), “Seismic Design of Reinforced Concrete


and Masonry Buildings”, John Wiley & Sons, Inc.,

Perera, R. (2005), “Performance Evaluation of Masonry-Infilled RC Frames


under Cyclic Loading based on Damage Mechanics”, Engineering Structures, No.
27, pp. 1278–1288.

Polyakow (1956), “Masonry in framed buildings”, Gosudarstvennoe isdatel’stvo


Literatury po stroitel’stvu i arckhitecture, Moscow.

Prakash. V., Powell, G. H and Campbell S. (1995), “Static & Dynamic Analysis of
Plane Structures”, NISEE, Earthquake Engineering Research Center, University
of California, Berkeley.

Qi, X. and Pantazopoulou (1991), “Response of RC Frames under Lateral Loads”,


Journal of Structural Engineering, ASCE, Vol. 117, No. 4, pp. 1167–1188.
136
Samarasinghe, W. and Hendry A. W. (1980), "The Strength of Brickwork under
Biaxial Tension-Compression", Proc. 7th Int. Symp. on Load Bearing Brickwork,
London, 1980.

Saneinjad, A. and Hobbs B. (1995), “Inelastic Design of Infilled Frames”, Journal


of Structural Engineering, ASCE, Vol. 121, No. 4, pp. 634–643.

Seible, F., LaRovere, H. L. and Kingsley, G. R. (1990), “Nonlinear Analysis of


Reinforced Concrete Masonry Shear Wall Structures – Cyclic Loading”, The
Masonry Society Journal, Vol. 9, No. 1, August.

Smith, B. S. (1962), “Lateral Stiffness of Infilled Frames”, Journal of Structural


Division, Proceedings of the ASCE, Vol. 88, No. ST 6, pp. 183–199.

Smith, B. S. (1966), “Behaviour of Square Infilled Frames”, Journal of Structural


Division, Proceedings of the ASCE, Vol. 92, No. ST 1, pp. 381–403.

Smith, B. S. and Carter, C. (1969), “A Method of Analysis for Infilled Frames”,


Proceedings of the Institution of Civil Engineers, Vol. 44, pp. 31–48.

Smith, B. S. and Coull, A. (1991), “An Infilled-Framed Structures”, Tall Building


Structures Analysis and Design, John Wiley & Sons, inc., pp. 168–174.

Smith, B. S. and Coull, A. (1991), “Tall Building Structures: Analysis and


Design”, John Wiley & Sons, Inc.

Smith, B. S., Carter, C. (1970), "Distribution of Stresses in Masonry Walls


subjected to Vertical Loading", Proc. 2nd Int. Brick Masonry Conference, 1970,
pp. 119-124.

Smith, B. S., Carter, C., Chowdhury, J. R. (1970), "The Diagonal Tensile Strength
of Brickwork", The Structural Engineer, Vol. 48, No 4, 1970, pp. 219-225.

Smith, B. S., Rahman K. M. K. (1972), "The Variation of Stress in Vertically


Loaded Brickwork Walls", Proc. Inst. of Civil Engineers, Vol. 43, 1972, pp. 6890.

Smith, B.S. and Carter, C. (1969), “A Method of Analysis for Infilled Frames”,
Proc. ICE, Vol. 44, pp.31-48, September.

Stratta, J. L. and Feldman, J. (1971), “Interaction of infill walls and


concrete frames during earthquakes”, Bulletin of Seismological Society
America, Vol. 61, No. 3, pp: 609–612.

137
Syrmakezis, C. A. (1996), “Tentative Guidelines for Protection and
Rehabilitations”, CISM Course on Protection of the Architectural Heritage
against Earthquake, Springer Wien, New York.

Tso, W. K. and Zelman, I. M. (1970), “Concrete Strength Variations in Actual


Structures”, ACI Journal, Vol. 87, pp. 981–988.

Van Zijl, G. P. A. G. (1996), “A Discrete Crach Modeling Strategy for Masonry


Structures”, Structural Engineering, Mechanics and Computation, 745-752.

Van Zijl, G. P. A. G. (1996), “Shear Transfer across Bed Joints in msonry: a


Numerical Study”, TU-DELFT Rep. No. 03.21.0.22.28, Delft University of
Technology, Delft, The Netherlands.

Veletsos, A. S. and Ventura, C. E. (1986), “Modal analysis of non-classically


damped linear systems”, Earthquake Engineering and Structural Dynamics, 14,
pp. 217–243.

Vukazich, S. E. (1998), “The Apartment Owner’s Guide to Earthquake Safety”,


San Jose State University, 1998.

138

You might also like