You are on page 1of 42

4

Chapter 2

LITERATURE REVIEW

2.1 Introduction
This chapter presents an overview of previous work on related topics that provide the
necessary background for the purpose of this research. The literature review concentrates
on a range of earthquake engineering topics and structural modelling aspects. For the
understanding of seismic capacity, a review of literature is required in experimental
testing, current design practice, theoretical strength evaluation and modelling techniques
such as finite element modelling. The literature review begins with a coverage of general
earthquake engineering topics, which serves to set the context of the research.
At present, there is no information available on seismic performance of arched rib slab
systems. However, research on similar types of systems have been conducted and the
available literature on those projects reviewed in following sections.
2.2 Earthquake design techniques
The objective of design codes is to have structures that will behave elastically under
earthquakes that can be expected to occur more than once in the life of the building. It is
also expected that the structure would survive major earthquakes without collapse that
might occur during the life of the building. To avoid collapse during a large earthquake,
members must be ductile enough to absorb and dissipate energy by post-elastic
deformations. Nevertheless, during a large earthquake the deflection of the structure
should not be such as to endanger life or cause a loss of structural integrity. Ideally, the
Created with novaPDF Printer (www.novaPDF.com)
5
damage should be repairable. The repair may require the replacement of crushed concrete
and/or the injection of epoxy resin into cracks in the concrete caused by yielding of
reinforcement. In some cases, the order of ductility involved during a severe earthquake
may be associated with large permanent deformations and in those cases, the resulting
damage could be beyond repair.
Even in the most seismically active areas of the world, the occurrence of a design
earthquake is a rare event. In areas of the world recognised as being prone to major
earthquakes, the design engineer is faced with the dilemma of being required to design for
an event, which has a small chance of occurring during the design life time of the building.
If the designer adopts conservative performance criteria for the design of the building, the
client will be faced with extra costs, which may be out of proportion to the risks involved.
On the other hand, to ignore the possibility of a major earthquake could be construed as
negligence in these circumstances. To overcome this problem, buildings designed to these
prescriptive provisions would (1) not collapse under very rare earthquakes; (2) provide life
safety for rare earthquakes; (3) suffer only limited repairable damage in moderate shaking;
and (4) be undamaged in more frequent, minor earthquakes.
The design seismic forces acting on a structure as a result of ground shaking are usually
determined by one of the following methods:

- Static analysis, using equivalent seismic forces obtained from response spectra for
horizontal earthquake motions.
- Dynamic analysis, either modal response spectrum analysis or time history analysis
with numerical integration using earthquake records.

Created with novaPDF Printer (www.novaPDF.com)
6
2.2.1 Static analysis
Although earthquake forces are of dynamic nature, for majority of buildings, equivalent
static analysis procedures can be used. These have been developed on the basis of
considerable amount of research conducted on the structural behaviour of structures
subjected to base movements. These methods generally determine the shear acting due to
an earthquake as equivalent static base shear. It depends on the weight of the structure,
the dynamic characteristics of the building as expressed in the form of natural period or
natural frequency, the seismic risk zone, the type of structure, the geology of the site and
importance of the building.
The natural frequency, which is the reciprocal of natural period, can be calculated using
the following formulae (Smith et al., 1991) as given in Table 2-1.
Table 2-1: Formulae to calculate the fundamental natural frequency of a building
(Smith & Coull, 1991).
Formula Notation Type of lateral load resisting
system
N
o
= D
1/2
/0.091H D = base dimension in the direction
of motion in meters.
H = height of the building in meters
Reinforced concrete shear wall
buildings and braced steel frames
N
o
= 10/N N = number of storeys Moment resisting frame
N
o
= 1/C
T
H
3/4
C
T
= 0.035 for steel structures, 0.025
for concrete structures,
H = height of the building in feet
Moment resisting frame is the
sole lateral load resisting system.
N
o
= 46/H H = height of the building in meters For any type of building

The static equivalent earthquake load mainly depends on the accuracy of natural period
calculation. The Australian code (AS1170.4, 1993) recommends N
o
= 46/H formula to
Created with novaPDF Printer (www.novaPDF.com)
7
calculate the natural frequency of the building. The calculation of equivalent earthquake
force in Australian code is similar to the method recommended by UBC (1997).
2.2.2 Dynamic analysis
The dynamic time-history analysis can be classified as either linear elastic or inelastic
(Chopra, 1995). The linear elastic modelling and analysis of Reinforced Concrete (RC)
structures is a well-established technique. Several commercial packages for the 3-D elastic
analysis of structures are available and are in widespread use (e.g. SAP2000,ETABS,
SPACE GASS, etc.). However, the results of the linear analysis are not useful in the
determination of the actual behaviour of the RC structures and the seismic safety analysis
which depends more on inelastic displacement and deformation up to collapse than on
forces. It is necessary to take advantage of the inelastic capacity of various components of
the structure. The response spectrum approach is based on the linear force response of an
equivalent single degree of freedom (SDOF) system. There have been several
developments in the response spectrum approach including modification to account for
some non-linear effects such as inelasticity, ductility and the response modification factor.
The use of the capacity-spectrum technique in the evaluation of RC buildings has been
suggested (ATC40, 1996). The recent development in the field of displacement-based
response spectra (Bommer et al., 1988; Priestley et al., 2000) represents a promising
approach that may be adapted to the simple seismic assessment of buildings. In general,
the response spectrum approach has its limitations. It does not account for the different
failure modes and sequence of component failure. It does not provide information on the
degree of damage or the ultimate collapse mechanism of a deficient RC structure. The
inelastic analysis of structures requires a non-linear dynamic time-history procedure past
the elastic response and up to collapse (Chopra, 1995). The two principal approaches to
Created with novaPDF Printer (www.novaPDF.com)
8
model RC component behaviour are microscopic finite element (FE) analysis and
macroscopic phenomenological models. Although accurate, it is not feasible to analyse an
entire structure using microscopic FE models. It is practical to study the behaviour of
isolated elements such as a beam, column, connection, structural wall, slab-column and
slab-wall so that their macroscopic analytical models defined in terms of global
parameters are developed for use in the analysis of a complete structure.
RUAUMOKO (Carr, 1998) is one of the popular programs available to carry out time
history analysis for two or three dimensional frame structures, which has a loading input, a
discretely defined acceleration record (The actual acceleration record is digitised in 0.005,
0.01, 0.02 or 0.025-second time intervals). This program has various types of hysteretic
elements to represents the member behaviour. The commonly used simple element in
RUAUMOKO for reinforced concrete members is the modified Takeda, stiffness-
degrading model (Takeda et al., 1970). There are more complex elements like Fukada
degrading Tri-linear hysteresis are also available for more refined analysis. Li Xinrong
(Carr, 1998) reinforced concrete column hysteresis rule is available in Ruaumoko to model
concrete columns, which allows for the changes in the stiffness of reinforced column as
the axial force in the column changes. The commonly used concrete beam-column
interaction surface is used to model the columns.
2.2.2.1 Member stiffness
When analysing concrete frame structures for gravity loads, it is generally considered
acceptable to base member stiffness on the uncracked section properties and to ignore the
stiffness contribution of longitudinal reinforcement. This is due to, under service-level
gravity loads, the extent of cracking will normally be comparatively minor and relative
Created with novaPDF Printer (www.novaPDF.com)
9
rather than absolute values of stiffness are all that are needed to obtain accurate member
forces (Paulay et al., 1992).
Under seismic actions, however, it is important that the distribution of member forces be
based on the realistic stiffness values applying close to member yield forces, as this will
ensure that the hierarchy of formation of member yield conforms to assumed distributions.
The structural deformations due to seismic loading will generally be associated with high
stresses. Consequently extensive cracking in the tension zone of reinforced concrete
beams, columns or walls must be expected. The estimation of deflections for the purposes
of determining period of vibration and inter-storey drifts, will be more realistic if an
allowance for the effect of cracking on the stiffness of the member is made. The New
Zealand concrete code (SANZ, 1995) recommends a value for beam stiffness of I
e
= 0.4 I
g

for rectangular sections, and I
e
= 0.35 I
g
for T-beam sections. A more detail
recommendations for stiffness modelling of beams and columns are available (Carr, 1994;
Paulay et al., 1992). In recent papers published by Priestley (1998a) and Priestley et al.
(1998b) have highlighted that beam stiffness is heavily dependent on reinforcement
content, and hence on strength. The use of member stiffness based on just the second
moment of area of member, may lead to significant errors in calculation of building period
and the expected drift.
The recommended procedure of calculating the member stiffness to be used in time-
history analysis is as follows:
- The first step is to obtain the moment curvature curve for the beam section using a
specialised computer program such as RESPONSE (Bentz et al., 2000) that considers
Created with novaPDF Printer (www.novaPDF.com)
10
strain hardening effects and confinement of concrete, where appropriate. Figure 2-1
shows a typical moment-curvature curve for a doubly reinforced flanged T-beam.
- The nominal flexural strength (M
n
) is determined at a curvature equal to 5 times the
nominal yield curvature (see Figure 2-1), which involves an iterative solution.
- The effective stiffness can be calculated from Equation 2-1.


Equation 2-1
- The above procedure is carried out for both negative and positive moment-curvatures.
The average stiffness value is recommended for the seismic analysis. The average is
appropriate as a consequence of moment reversal along the beam length under seismic
loading conditions.

Figure 2-1: Effective bi-linear yield curvature [After (Priestley, 1998b)]
g
g c y
n
e
I
I E
M
I
(
(

Created with novaPDF Printer (www.novaPDF.com)


11
2.2.2.2 Effective flange width
The flange contribution to stiffness in L and T-beams is typically less than the contribution
to flexural strength (Paulay et al., 1992), as a result of the moment reversal occurring
across beam-column joints and the low contribution of tension flange to flexural stiffness.
Therefore, an effective flange width has to be evaluated both flexural compressive strength
and stiffness. These values are given in Figure 2-2.

Figure 2-2: effective flange width calculation after (Paulay et al., 1992)
Identical guide lines to determine the effective flange width for strength evaluation are
given in USA. (ACI-318, 2002) and New Zealand codes (SANZ, 1995), while slightly
different recommendations are given in British (BS8110, 1997) and Australian codes
(AS3600, 2001).
) ( 1 . 0
) ( 2 . 0
beams L l b b
beams T l b b
z w eff
z w eff
+ =
+ =

Where
z
l is the distance between points of zero bending moment.
Equation 2-2: effective flange width calculation [after (AS3600, 2001; BS8110, 1997)]
These effective flange widths are used in analytical work described later in chapter 5.

Created with novaPDF Printer (www.novaPDF.com)
12
2.2.3 Displacement-based seismic design
In recent years there have been extensive examinations of the current seismic design
philosophy, which is based on provision of a required minimum strength, related to initial
stiffness, seismic intensity and a force reduction or ductility factor considered to be a
characteristic of a particular structural system and construction material. There are
inappropriate two fundamental assumptions of the force-based design: (1) that the initial
stiffness of a structure determines its displacement response and (2) that a ductility
capacity can be assigned to a structural system regardless of its geometry, member
strength, and foundation conditions (Priestley et al., 2000).
The damage sustained by structures during seismic events is closely related to their
displacements and deformation. For this reason, deformation-based design approaches
have been developed to create a structure with controlled and predictable performance.
This design process is consistent with the capacity design philosophy, as it requires control
over deformation demand and supply of the energy dissipation zones. The direct
displacement-based design have now matured to the stage where seismic assessment of
existing structures, or design of new structures can be carried out to ensure that particular
deformation-based criteria are met.
2.3 Behaviour of reinforced concrete structures in earthquakes
The overall behaviour of a structure when subjected to earthquake forces is affected by a
number of factors.
Created with novaPDF Printer (www.novaPDF.com)
13
2.3.1 Factors affecting the earthquake performance of reinforced concrete
structures
As reported by (Sanders, 1995), the poor performance of buildings was generally due to a
combination of inadequate strength and stiffness of the overall seismic resisting system
and a poor distribution of strength and stiffness over successive storeys, leading to soft
storey formation, a lack of provision of an adequate load path through the structure leading
to partial or complete failure of the structure, and poor detailing of joints and connection
leading to various types of non- ductile failures.
- Ductility Capacity :
As described by Park (1992), the term ductility in structural design is used to mean the
ability of a structure to undergo large inelastic deformations in the post-elastic range
without a substantial reduction in strength. Ductility is an essential design requirement for
a structure to behave satisfactorily under severe earthquake excitation. The ductility
demand of a structure under seismic loading is dependent on the construction material, the
design elastic strength and the structural system.
The required ductility of a structure, element or section can be expressed in terms of the
maximum imposed deformations. Often it is convenient to express the maximum
deformation in terms of ductility factors, where the ductility factor is defined as the
maximum deformation divided by the corresponding deformation present when yielding
first occurs. The use of ductility factors permits the maximum deformations to be
expressed in non-dimensional terms as indices of post-elastic deformation for design and
analysis. Ductility factors have been commonly expressed in terms of the various
parameters related to deformations, i.e. displacements, rotations, curvatures and strains.
Created with novaPDF Printer (www.novaPDF.com)
14
- Effects of drift:
In flexible buildings, there can be relatively large lateral movements between consecutive
storeys, which is called the inter-storey drift. This can damage the structure and can also
lead to unacceptable damage to the cladding and non-structural elements. This effect can
be controlled with careful design and detailing. The control of the estimated lateral drift is
another design aspect, which has a significant effect on the seismic performance of
structures. Australian code (AS1170.4, 1993) requires that the maximum inter-storey drift
be restricted to 1.5% of the storey height.
- P-Delta effect:
P-delta effects reduce seismic performance because the effective lateral loads are
increased as lateral displacements increase. This has the effect of further increasing the
lateral displacement, and placing higher demand on the structural system. Damage will
therefore occur sooner than in similar systems without significant P-delta effect. The
importance of P-delta effects on the seismic performance of structures depends upon both
the extent of vertical load being carried by the lateral resisting system and the stiffness of
that system. If vertical loads are carried by columns, which are not part of the lateral load
resisting system, then P-delta effects are not likely to be significant. Stiffer structural
systems, such as shear walls, are less prone to P-delta effects because the lower lateral
displacements control the additional over turning moments due to vertical loads.
P-delta effects are significant for flexible systems, e.g. Moment-resisting frames, which
carry both vertical and lateral loads to the foundation. They are most significant for fully
ductile systems, because the relative values of vertical to lateral load are increased and the
lateral load resisting system is more flexible than for structures with limited ductility.
Created with novaPDF Printer (www.novaPDF.com)
15
Therefore P-delta effects in ductile systems are generally reduced somewhat below the
limiting drift values allowed by the code. P-delta effects should be included in determining
the deflection at the ultimate limit state, with some exceptions, e.g. Short period (stiff
structures), low structures, and structures that are designed to respond elastically. Sway
effects produced by vertical loads acting on the structure in its displaced configuration also
should be taken in to account. The extent to which such effects are included by designers
of flexible ductile systems which carry both vertical and lateral loads can have a
significant effect on the seismic performance of such structures, particularly when ground
motions may be substantially greater than those for which the structure has been designed
(Heidebrecht, 1997).
- Effects of strong beams and weak columns:
Under earthquake and gravity loading, the critical bending moments develop in the
vicinity of the frame joints. If these moments exceed the limit state capacity of the
sections, plastic hinges will develop. These hinges may develop mainly in beams, columns
or in a combination of locations. Beam hinge mechanism is more suitable for achieving
ductility in concrete frames than column mechanism because:
- A greater number of plastic hinges need to form before a collapse mechanism
develops leading to smaller inelastic rotations in each hinge.
- Columns are more critical because they carry the total gravity load from the
structure above and damage to them could lead to catastrophic failures.
- Beam hinges are more ductile because they carry lower axial loads than column
hinges.
Created with novaPDF Printer (www.novaPDF.com)
16
2.3.2 Strength and ductility of materials
2.3.2.1 Reinforcement
Figure 2-3(a) taken from Park (1992) shows typical stress-strain curves measured for
reinforcing bars under monotonic loading. In practice, the actual yield strength of the steel
will normally exceed the lower characteristic yield strength f
y
. Also, in the plastic hinge
regions during a major earthquake, the longitudinal reinforcement may reach strains in the
order of 20 or more times the strain at the first yield, and a further increase in steel stress
due to strain hardening may occur. The resulting increase in the flexural strength in
plastic hinge regions due to these two factors is of concern, since it is accompanied by an
increase in the shear forces, which could result in brittle failure, and an increase in the
column bending moments, which could cause column plastic hinges. A capacity design
procedure should be used to ensure that flexural yielding occurs only at the chosen plastic
hinge locations during a severe earthquake. In the capacity design procedure, when
designing other regions of the structure, it is assumed that actions are those associated with
the development of the maximum probable flexural strength at the plastic hinges, referred
to as the flexural over-strength. It is evident that the properties of the reinforcing steel to
be used in seismic design should be based on rigorous statistical analysis of the stress-stain
properties, to determine the lower and upper bounds of the flexural strength of reinforced
concrete elements.
Figure 2-3(b) shows stress-stain curves measured for reinforcing steel under cyclic
loading. The rounding of the stress-stain curve during loading reversals in the post
elastic range is due to the Bauschinger effect. This reduction in the tangent modules of the
steel at relatively low compressive stress during reversed loading makes the buckling of
Created with novaPDF Printer (www.novaPDF.com)
17
compression steel more likely than would be expected during monotonic loading. It is
very important that statistical information on the stress-stain properties of the reinforcing
steel used in seismic regions be available. A proper capacity design cannot be undertaken
without knowledge of the likely variations of the steel properties to obtain strength factors,
and adequate ductility of plastic hinges of members cannot be ensured if the steel is brittle
(Park, 1992).
(a)
(b)

Figure 2-3: Typical stress-strain curves for reinforcing steel (a) with monotonic loading (b) with cyclic
loading mainly in the tensile range of strain.
2.3.2.2 Concrete Behaviour
Figure 2-4 taken from (Mander et al., 1988) illustrates a typical non-linear stress-strain
relationship for confined and unconfined concrete. The confinement is provided by the
Created with novaPDF Printer (www.novaPDF.com)
18
lateral reinforcement. Concrete is a strain-softening material, unlike structural steel, which
is a strain-hardening material. Strain softening is a decline of stress at advance strain, and
is reflected in the moment-curvature diagrams of flexural members.

Figure 2-4: Non-linear stress-strain relation for confined and unconfined concrete.
2.3.3 Dynamic behaviour of multi-storey frames
It is shown from non-linear dynamic analysis that unexpected distribution of bending
moments may occur in columns of multi-storey frames, compared with the distribution
obtained from static lateral loading (Paulay et al., 1992). Static lateral load analysis
indicated that points of contraflexure exist generally close to mid height of columns.
However, non linear dynamic analysis suggests that at certain times during the response of
the structure to earthquake ground motions, the point of contraflexure in a column
between floors may be close to the beam-column joint and the column may even be in
single curvature. The reasons for the unexpected distribution of column bending moments
at some instants of time is the strong influence of higher modes of vibration, particularly
second and third modes (Paulay et al., 1992).
Created with novaPDF Printer (www.novaPDF.com)
19
The shift of point of contraflexure in the columns to well away from mid height position in
some cases means that the column moments induced may be much higher than the
moments obtained from the static lateral load analysis and may lead to plastic hinges
forming in columns. Thus, columns will need extra lateral reinforcement to provide
sufficient confinement for concrete.
Frames subjected to severe earthquake motions will undergo several reversals of loading
well into the inelastic range during an earthquake. The factors that affect the load
deflection relationship of concrete members subjected to large cyclic inelastic
deformations are:
1. The inelastic behaviour of the steel reinforcement: when subjected to reversed loading,
the stress strain curve becomes non-linear at a much lower stress than the initial yield
strength.
2. The extent of cracking of concrete: The opening and closing of cracks will cause a
deterioration of concrete, hence will result in stiffness degradation. The larger the
portion of load carried by the concrete, the larger the stress degradation.
3. The effectiveness of bond and anchorage: A gradual deterioration of bond between
concrete and steel occurs under high intensity cyclic loading.
4. The presence of shear: High shear forces will cause further loss of stiffness because of
increase in shear deformation in plastic hinge zones under reversed loading.
2.3.4 Bar slip and bond deterioration
Bar bond slip plays a significant role in the performance of reinforced concrete structures
such as in the case of inadequate anchorage of the beam bottom reinforcement. After
yielding of the beam longitudinal reinforcement the bond slip propagates to the beam
column joint causing additional rotation at the beamcolumn interface. When the bottom
longitudinal reinforcement starts to slip, pullout of the bottom reinforcement occurs which
Created with novaPDF Printer (www.novaPDF.com)
20
reduces the positive moment capacity substantially. This in turn will reduce the shear in
the joint. The beam will experience rigid body rotation with pronounced pinching (Paulay
et al., 1992).
2.3.5 Joint shear deformation
Joint shear deformation is an important component of the local and overall deformations
of the structure. Experimental measurements on specimens representing existing beam
column joints showed that joint shear deformation contributes over 30% of the story drift.
Shear failure in the joint element can be defined by compressive failure of deteriorated
concrete due to cracking defined by maximum strain in concrete and the tensile failure
when the reinforcement bar reaches the limit state. In spite of the tremendous advances in
the development of sophisticated models for the non-linear analysis of RC structures, the
accuracy and reliability of the results remain to be established. The lack of reliability with
current analysis methods is partly because of limitations in modelling and the adopted
simplifying assumptions (Miranda, 1996).
2.4 Performance assessment
2.4.1 Displacement ductility and capacity
Most researchers relate adequate performance to a certain level of displacement ductility
factor. Displacement ductility factor is defined as:

y
A
A
=
max
Equation 2-3
Where,
max
A

= Maximum displacement

y
A

= Displacement at yield
Created with novaPDF Printer (www.novaPDF.com)
21
The displacement ductility factor required for a typical structure is usually between 3 to 6.
Most design codes refer to this, as the ductility required of a structure responding to a
major earthquake. One disadvantage of using ductility factors as a performance criterion is
that very often the load-deflection relation for a structural component does not have a
well-defined yield point. Because of the difficulties in the definition of yield displacement,
some researchers (Durrani et al., 1985; Park, 1988) have suggested that the deformation
history used in quasi-static testing should be based on the drift ratio rather than the
ductility factor. Also, for the case of interior connections, where significant pinching of
the hysteretic responses occurs as a result of slippage of beam reinforcement, the ductility
factor becomes a meaningless parameter. Paulay (1988) suggested that structures
withstanding storey drift of up to 3% are satisfactory. A maximum inter-storey drift ratio
of 2% has also been a commonly accepted limit. The Australian earthquake loading code
(AS1170.4, 1993) also states that the design storey drift should not exceed 1.5%. It should
also be noted that the New Zealand loading code (NZS4203, 1992) states that the design
storey drift should not exceed 2% for h
n
s15m, where h
n
is the height from base of
building to the level of uppermost principal seismic weight.
2.4.2 Energy dissipation capacity
Energy dissipation capacity has been proposed by many investigators as a measure of
member performance. Energy dissipation capacity can be easily obtained as the area
within the hysteretic loops. However, the energy dissipation capacity of a test specimen is
dependent on several parameters that include material properties, reinforcing details,
geometry of the unit and peak deformations. Hence the use of the total energy dissipation
capacity in order to assess the performance of test specimens of different characteristics
and tested under different conditions would be doubtful.
Created with novaPDF Printer (www.novaPDF.com)
22
One of the more common approaches adopted for the measure of energy dissipation is the
use of equivalent viscous damping ratio (h
eq
). This h
eq
value is defined by Kitayama et al.
(1991) as the ratio of the energy dissipated within half a cycle to 2 times the strain
energy at peak of an equivalent linear elastic system. This h
eq
value is used to determine
the energy dissipated in a particular loading cycle, and to measure the degree of pinching
of the hysteretic loops. The definition of h
eq
is illustrated in Figure 2-5.

Figure 2-5: Definition of equivalent viscous damping ratio h
eq
[After (Quintero-Febres et al., 1997)]

2.5 Finite element analysis
The application of the finite element modelling (FEM) to RC structures has been
underway for the last 20 years, during which time it has proven to be a very powerful tool
in engineering analysis. The wide dissemination of computers and the development of the
finite element method have provided means for analysis of much more complex systems in
a much more realistic way.
For any type of structure, the more complicated its structural geometric configuration is,
the more a computer-based numerical solution becomes necessary. It has also been shown
Created with novaPDF Printer (www.novaPDF.com)
23
that experimental investigations are time consuming, capital intensive and even often
impractical. The FEM is now firmly accepted as a very powerful general technique for the
numerical solution of a variety of problems encountered in engineering. For concrete
structures in particular, because of complexities of concrete behaviour in tension and
compression together with integrity of concrete and steel, extreme difficulties are
encountered in modelling and obtaining closed form solutions, even for very simple
problems (Abdollahi, 1996).
The civil engineering structures are today designed with respect to the limit state of
serviceability and limit states of the strength and stability. These complex problems of a
different nature are possible to solve by FEM methods. Nonlinear elastic concrete models
have been extensively used in finite element analysis of RC structures with varying
degrees of success.
The main obstacle to finite element analysis of reinforced concrete structures is the
difficulty in characterizing the material properties. Much effort has been spent in search of
a realistic model to predict the behaviour of reinforced concrete structures. Due mainly to
the complexity of the composite nature of the material, proper modelling of such
structures is a challenging task. Despite the great advances achieved in the fields of
plasticity, damage theory and fracture mechanics, among others, a unique and complete
constitutive model for reinforced concrete is still lacking.
Created with novaPDF Printer (www.novaPDF.com)
24
Finite element analysis has advantages over most other numerical analysis methods,
including versatility and physical appeal. The major advantages of finite element analysis
can be summarised as (Cook et al., 2002):
Finite element analysis is applicable to any field problem.
There is no geometric restriction. The body analysed may have any shape.
Boundary conditions and loading are not restricted.
Material properties are not restricted to isotropy and may change from one element
to another or even within an element.
Components that have different behaviours, and different mathematical descriptions,
can be combined.
A finite element analysis closely resembles the actual body or region.
The approximation is easily improved by grading the mesh.
Some disadvantages of finite element analysis are:
It is fairly complicated, making it time-consuming and expensive to use.
It is possible to use finite element analysis programs while having little knowledge of
the analysis method or the problem to which it is applied. Finite element analyses
carried out without sufficient knowledge may lead to results that are worthless and
some critics say that most finite element analysis results are worthless (Cook et al.,
2002).
Specifically developed computer programs are used in finite element analyses of
reinforced concrete structures. However, many commercially available general-purpose
codes provide some kind of simplified material models intended to be employed in the
analysis of concrete structures.
Created with novaPDF Printer (www.novaPDF.com)
25
2.5.1 The material models
The program ANSYS, Version 8 (2003) was used in this study to model the test specimen.
It has been used successfully in the past to model beam-column subassemblage. Its
reinforced concrete model consists of a material model to predict the failure of brittle
materials, applied to a three-dimensional solid element in which reinforcing bars may be
included. The material is capable of cracking in tension and crushing in compression. It
can also undergo plastic deformation and creep. Three different uniaxial materials, capable
of tension and compression only, may be used as smeared reinforcement, each one in any
direction. Details of element types used for concrete and reinforcement are given in
chapter 5.
2.5.1.1 Failure Criteria for Concrete
The model is capable of predicting failure for concrete materials. As mensioned in the
previous section both cracking and crushing failure modes are accounted for. The two
input strength parameters i.e., ultimate uniaxial tensile and compressive strengths are
needed to define a failure surface for the concrete. Consequently, a criterion for failure of
the concrete due to a multi-axial stress state can be calculated (William et al., 1975).
A three-dimensional failure surface for concrete is shown in Figure 2-6. The most
significant nonzero principal stresses are in the x and y directions, represented by xp and
yp, respectively. Three failure surfaces are shown as projections on the xp-yp plane. The
mode of failure is a function of the sign of zp (principal stress in the z direction). For
example, if xp and yp are both negative (compressive) and zp is slightly positive (tensile),
cracking would be predicted in a direction perpendicular to zp. However, if is zero or
slightly negative, the material is assumed to crush (ANSYS, 2003).
Created with novaPDF Printer (www.novaPDF.com)
26

Figure 2-6: 3-D failure surface for concrete (ANSYS, 2003)


2.5.2 Non-linear solution
In non-linear analysis, the total load applied to a finite element model is divided into a
series of load increments called load steps. At the completion of each incremental solution,
the stiffness matrix of the model is adjusted to reflect nonlinear changes in structural
stiffness before proceeding to the next load increment. The program ANSYS (ANSYS,
2003) uses Newton-Raphson equilibrium iterations for updating the model stiffness.
Program ANSYS is used in finite element analysis in chapter 5.
Newton-Raphson equilibrium iterations provide convergence at the end of each load
increment within tolerance limits. Figure 2-7 shows the use of the Newton-Raphson
approach in a single degree of freedom nonlinear analysis.
Created with novaPDF Printer (www.novaPDF.com)
27
Displacement
Load
Converged Solutions

Figure 2-7: Newton-Raphson iterative solution (3 load increments) (ANSYS, 2003)

Prior to each solution, the Newton-Raphson approach assesses the out-of-balance load
vector, which is the difference between the restoring forces (the loads corresponding to the
element stresses) and the applied loads. Subsequently, the program carries out a linear
solution, using the out-of-balance loads, and checks for convergence. If convergence
criteria are not satisfied, the out-of-balance load vector is re-evaluated, the stiffness matrix
is updated, and a new solution is attained. This iterative procedure continues until the
problem converges (ANSYS, 2003).
2.5.2.1 Load stepping and failure definition for FE models
For the non-linear analysis, automatic time stepping in the ANSYS program predicts and
controls load step sizes. Based on the previous solution history and the physics of the
models, if the convergence behaviour is smooth, automatic time stepping will increase the
load increment up to a selected maximum load step size. If the convergence behaviour is
abrupt, automatic time stepping will bisect the load increment until it is equal to a selected
Created with novaPDF Printer (www.novaPDF.com)
28
minimum load step size. The maximum and minimum load step sizes are required for the
automatic time stepping.
In the FE study conducted by Kachlakev et al. (2001), it was shown that the convergence
of the models depended heavily on behaviour of reinforced concrete. Full size RC bridge
beam model was used by Kachlakev et al. (2001) to demonstrate the load stepping. Figure
2-8 shows the load-deflection plot of the beam with four identified regions exhibiting
different reinforced concrete behaviour. The load step sizes have been adjusted, depending
upon the reinforced concrete behaviour occurring in the model as shown in Table 2-2.

Figure 2-8: Reinforced concrete behavior in RC beam (After Kachlakev et al., 2001)
Table 2-2 Summary of load step sizes for beam model (After Kachlakev et al., 2001)

Table 2-2 shows the load step sizes used by Kachlakev et al. (2001) to obtained the
converged solution for non-linear analysis. As shown in the table, the load step sizes do
not need to be small in the linear range (Region 1). At the beginning of Region 2, cracking
Created with novaPDF Printer (www.novaPDF.com)
29
of the concrete starts to occur, so the loads have been applied gradually with small load
increments. Load step size of 0.91 kg (2 lb) is defined for the automatic time stepping
within this region. As first cracking occurs, the solution becomes difficult to converge. If a
load applied on the model is not small enough, the automatic time stepping will bisect the
load until it is equal to the minimum load step size. After the first cracking load, the
solution becomes easier to converge. Therefore the automatic time stepping increases the
load increment up to the defined maximum load step size, which is 34.05 kg (75 lb) for
this region. If the load step size is too large, the solution either needs a large number of
iterations to converge, which increases the computational time considerably, or it diverges.
In Region 3, the solution becomes more difficult to converge due to yielding of the steel.
Therefore, the maximum load step size is reduced to 11.35 kg (25 lb). A minimum load
step size of 0.45 kg (1 lb) has been defined to ensure that the solution will converge, even
if a major crack occurs within this region. Finally, for Region 4, a large number of cracks
occur as the applied load increases. The maximum load step size has been defined to be
2.27 kg (5 lb), and a 0.45 kg (1 lb) load increment is specified for the minimum load step
size for this region. For this study, a load step size of 0.45 kg (1 lb) is generally small
enough to obtain converged solutions for the models. It should be noted that the above
procedure cannot be used without at least having a rough idea of the load deformation
curve, therefore, this becomes trial and error iterative procedure.
The failure of the models has been defined when the solution for a 0.45 kg (1 lb) load
increment still does not converge. The program then gives a message specifying that the
models have a significantly large deflection, exceeding the displacement limitation of the
ANSYS program.
Created with novaPDF Printer (www.novaPDF.com)
30
2.5.3 Evolution of crack patterns
In ANSYS, stress and strain outputs are calculated at integration points of the concrete
solid elements. Figure 2-9(a) shows the integration points in a concrete solid element. A
cracking sign represented by a circle appears when a principal tensile stress exceeds the
ultimate tensile strength of the concrete. The cracking sign appears perpendicular to the
direction of the principal stress as illustrated in Figure 2-9(b). The smeared cracked pattern
in ANSYS can be displayed at each integration point or at element centroid.
(a)
(b)

Figure 2-9: (a) Integration points in concrete solid element (b) Cracking sign [After(ANSYS, 2003)]
2.6 Ribbed slab construction
Rib beam slab systems have long been regarded as one of the most economically efficient
forms of reinforced concrete Gravity Load Resisting Systems (GLRS). Specially for long
span slabs or slabs with very high-imposed loading, rib slab construction is extremely
economical and viable. The first rib slab system invented by Francois Hennebique, has
been patented in early 1900s,which had fallen into disuse due to the high cost of timber
and labour. The innovative long span light weight formwork system, Corcon, has been
developed and patented throughout the world, by Decoin Pty Ltd in response to an
increasing shortage of good quality plywood and the increasing need for safe, economical
and durable structural slab system. The Corcon slab, in contrast to conventional rib slab
Created with novaPDF Printer (www.novaPDF.com)
31
system, is easier and less expensive to construct from both labour and material points of
view. Corcon system has better performance with respect to environmental benefits such
as greenhouse gas emission, less use of embodied energy and reduction of whole of the
cycle cost in terms of maintenance and energy used (Dragh, 2000).
This reusable lightweight sheet metal form system optimises the traditional rib slab
construction (see Figure 2-10) by using corrugated arch metal sheet spanning over series
of sheet metal beam moulds to form the suspended concrete slab. The corrugated arched
metal sheet enables the rib beam spacing to be increased to 1200 mm from the
conventional 600 mm. A typical cross section of the Corcon formwork system is shown in
Figure 2-11. This system could be designed to span up to 9.0 m with simple reinforcement
and to span 14.0 m with post-tensioning.
600 mm
Typical
Plastic / plywood
Rib moulds

Figure 2-10: Typical conventional ribbed slab construction
Created with novaPDF Printer (www.novaPDF.com)
32


Figure 2-11: Typical cross section of Corcon slab formwork system
The use of ribs to the soffit of the slab reduces the quantity of concrete and reinforcement
and also the weight of the floor. The saving of materials of the conventional ribbed slab
systems will be offset by the complication of formwork and placement of reinforcement.
However, formwork complication is minimised by use of standard, modular, reusable
formwork, usually made from polypropylene or fiberglass. The Corcon formwork system
is further refined to achieve additional savings over the conventional formwork system,
such as reduction of scaffolding frames by over 50 % resulting in 50 % labour saving. The
reduction of the formwork supporting points are possible due to the fact that the sheet
metal rib forms are capable of spanning greater distance and the use of a special
bracketing system, which will prevent the buckling of formwork. The main advantage of
this contiguously cast rib slab system gives enhanced structural performance by making
use of the high shear resistance of the slab and the high flexural resistance of the ribs. The
slab between ribs is capable of supporting considerable superimposed dead and live load,
due to the natural arching action. The first Corcon slab has been installed in a two-storey
house in Queanbeyan, NSW in March 1995. Since then, over 80 000 m
2
has been placed
including 10 000 m
2
installed in Kuala Lumpur. Figure 2-12 illustrates the slab soffit of a
Corcon slab system.
Created with novaPDF Printer (www.novaPDF.com)
33

Figure 2-12: Corcon rib beam and slab soffit

2.6.1 Code Recommendation for Rib Slab Design
For conventional rib slab systems various minimum member sizes, proportions and rib
spacing have been specified in British and U.S. codes. ACI code(ACI-318, 2002) limits
the maximum clear rib spacing to 800 mm, New Zealand code (SANZ, 1995) limits the
maximum clear rib spacing to 750 mm whereas British code(BS8110, 1997) allows rib
spacing specified up to 1500 mm. The commentary to the U.S. and New Zealand codes
indicates that the size and spacing limitations for rib slab construction are based on
successful performance in the past and with an allowance of 10% higher shear stress
carried by concrete.
ACI and New Zealand codes recommend that rib depth excluding topping should not be
greater than 3-1/2 times the minimum width of the rib. The minimum width of the rib
should not be less than 100 mm. The minimum thickness of slab should not be less than 50
mm or one-twelfth of clear distance between ribs whichever is greater. In contrast BS
Code requires that the depth excluding topping should not be greater than four times the
Created with novaPDF Printer (www.novaPDF.com)
34
width of the rib. The minimum width of the rib is determined by the consideration of
cover, bar spacing, fire and durability. Bs code also specifies that the minimum thickness
of slab should not be less than 50 mm or one tenth of clear distance between ribs
whichever is greater.
Clearly, experimental work is required to explore performance of this new rib system,
which is quite different to the conventional system. Australian concrete structures Code
(AS3600, 2001) does not provide any design guidelines for rib slab construction.
2.7 Previous Relevant Experimental Work on Ribbed Slab System
There have been no investigations reported on seismic behaviour of concrete beam-arch
slab systems, both locally and internationally. However, there has been a limited amount
of research into reinforced concrete T-beams and beams with flange effect. As no directly
related research work was found, some relevant research projects were reviewed and
presented in following sections.
2.7.1 Research work carried out by Shao-Yeh et al. (1976)
The work covered by Shao-Yeh et al. includes an experimental and analytical study
program to investigate the inelastic behaviour of critical regions that may develop in a
beam near its connection with the column of a reinforced concrete ductile moment-
resisting space frame when subjected to severe earthquake excitations. In the experimental
program, a series of nine cantilever beams, representing half-scale models of the lower
story girder of a 20-story ductile moment-resisting reinforced concrete office building was
designed according to ACI (318-71) Code. These cantilever beams were designed in order
to study the effects of (1) the slab by testing T-beams with a top slab width equal to the
Created with novaPDF Printer (www.novaPDF.com)
35
effective width specified by the ACI (318-71) Code; (2) relative amounts of top and
bottom reinforcement by varying the amounts of bottom reinforcement.
The amount of instrumentation (mostly electronic transducers) used in the experimental
set up and provided valuable data for obtaining the overall response of the test beams, as
well as for studying in detail most of their deformation and resistance mechanisms. Data
from the continuously recorded hysteretic force deformation diagrams provided excellent
information on the overall beam behaviour since the history of stiffness degradation,
strength degradation and energy dissipation were easily deduced using such data.
Photogrammetric techniques were used for studying the deformation pattern of critical
regions in order to detect the nature of shear distortion. Shao-Yeh et al. concluded that
more realistic models, such as beam-column subassemblages, should be tested to study the
effect of critical regions near beam-column connections and the contribution of different
types of floor systems in the overall behaviour of these assemblages.
It was found that the stiffness degradation occurring in R/C beams has been identified to
be very sensitive to the loading history. It has been observed that once the peak
deformation of a cycle increased in either direction during inelastic load reversals the
initial stiffness and energy dissipation per cycle were observed to degrade during
subsequent reversals. Stiffness degradation also occurs due to repeated applications of
loading reversals at constant large beam displacement ductilities. In the low shear stress
situations, the Bauschinger effects of steel and bond deterioration have been considered
the main sources of stiffness degradation.
The failure of unsymmetrically reinforced beams (commonly found in T-beams with
unequal top and bottom reinforcement) subjected to reversals after flexural yielding,
Created with novaPDF Printer (www.novaPDF.com)
36
precipitated or accelerated by local buckling of the bottom bars near the beam support
when these bars were compressed during downward loadings. For the symmetrically
reinforced beams, failure appear to have been caused by the gradual loss of shear transfer
capability along large cracks, which opened up across the entire beam section.
It was identified that the energy dissipation capacity of R/C beams can be increased by
delaying the degradation of stiffness and the early failure of the beam, which may result
from buckling of the compression bars. More specifically, this can be achieved in the
following ways: by providing supplementary cross-ties to support the compression bars
unrestrained by corners ties. Using supplementary ties, over beam without such ties,
attained a 74% increase in the energy dissipation capacity. Increasing the amount of
bottom steel by 89 %, there was an improvement in the energy dissipation capacity by
55%.
It was found that the bond stress behaviour of anchored main bars in compression and
tension is different. The length required to develop applied compression forces along
cyclically loaded anchored main bars was less than that required to develop tension, i.e., a
larger maximum bond stress was developed along compression bars than along tension
bars. There were two areas where bond stress could not develop effectively. One was near
the beam-column interface, where bond disruption occurred as a consequence of the shear
that developed in the bar due to dowel action at the interface crack The other area where
bond could not properly develop was along the length where yielding takes place at the
peaks of cyclic loading. Here, bond disruption was mainly due to considerable contraction
of the bar.
Created with novaPDF Printer (www.novaPDF.com)
37
It was concluded that the main influence of the slab on the inelastic behaviour of T-beams
was the contribution of slab reinforcement to the top tensile steel area. The increase in
downward moment capacity due to slab reinforcement caused more energy dissipation per
cycle. However, this increase imposed higher compression in the bottom compression
zone, and higher shear force acting in the downward direction. These increased
compression and shear forces could cause early buckling of bottom bars and increase the
amount of shear degradation. These factors should be considered in the analysis and
design of the critical regions near beam-column connections. Therefore confinement of
compression bars in T-form beams (such as rib beams) is required.
It was identified by comparing the hysteretic behaviour of Beams with different lateral tie
reinforcement indicated the advantages of providing lateral supports for main compression
bars by means of stirrup-tie corners or by supplementary cross-ties. It was recommended,
therefore, that current provisions for the arrangement of lateral ties for longitudinal bars in
the columns also apply to compression bars in beams. Therefore, it may be essential to
keep ligatures in rib beams, even the shear requirement is not critical.
It was concluded that when full deformational reversals are expected to occur in the beam
critical regions near the column connections to improve energy dissipation capacity, it was
recommended that the bottom (positive moment) steel be at least 75 percent of the top
(negative moment) steel. This may be useful to check the performance of the rib beam
with different bottom steel percentage. As experimental program of this nature is
expensive, this could be easily done using FE analysis technique.
Created with novaPDF Printer (www.novaPDF.com)
38
2.7.2 Research work carried out by Durrani et al. (1987)
The work covered by Durrani et al. includes an experimental and analytical study program
to investigate the behaviour of interior beam-to-column connections including a floor slab
was studied by testing three subassemblages. The length of the beam and the height of the
columns represented one half of the span and storey height, respectively. This testing
arrangement was based on the assumption that for moment-resisting frames subjected to
lateral loading, the inflection points will occur approximately at mid span of beams and at
mid-height of columns and will remain stationary during load reversals. This assumption
results in test specimens that was convenient for laboratory testing. Despite this
simplification, such tests have given considerable insight into the behaviour of joints.
The test specimens have been designed based on the assumption that when the slab beam
was in negative bending, the beam longitudinal reinforcement and the slab longitudinal
reinforcement over the entire width of the slab would yield simultaneously. The columns
were designed to be at least 20 percent stronger than the slab beam to ensure the formation
of flexural hinges in the beams.
It was identified from the test results of subassemblage, the hysteretic loops have become
increasingly pinched after the 2% drift. This was attributed mainly to the opening and
closing of wide flexural cracks in the bottom of the main beams. It also observed that the
major flexural crack has been observed was at the beam-to-column interface of the
specimen.
Created with novaPDF Printer (www.novaPDF.com)
39
Cracking of the joint core due to shear in the joint was identified as an important factor
that affects the bond of reinforcing bars passing through the joint. It also identified that
there is a better anchorage of top bars compared to the bottom bars. This can be partly
attributed to the confinement of the upper portion of the joint by the slab. However, the
major factor was the larger amount of top steel than the bottom steel.
It has been observed that main beam top reinforcements starts yielding at the drift of 1.5 %
while the slab reinforcement remained elastic and the stain in the slab reinforcement
decreases with the distance away from the beam. However, once the main beam top
reinforcement yielded, the strain in the slab reinforcement has increased rapidly, the
reinforcing bars farther from the main beam experienced higher stain than did close to the
main beam. At a drift level of 4%, the reinforcement in the entire width of the slab had
yielded. Thus, for server earthquake loading, the contribution of the slab in calculating the
beam flexural strength cannot be ignored.
It was concluded that beams with slab sections (T beams) with unequal amounts of top and
bottom steel (more top steel than bottom steel), the range of strain demand during reversed
cyclic loading will be more server for the bottom steel. Thus, bond deterioration and bar
slip problems will be more significant for the bottom steel.
2.7.3 Research work carried out by Pantazopoulou et al. (2001)
The work covered by Pantazopoulou et al. includes an experimental and analytical study
program to investigate the effect of slab participation in seismic design. Until recently, it
was an established design practice to neglect the presence of the slab in estimating beam
stiffness and strength, except when the slab was located in the compression zone of the
beam (known as T beam design). Experimental evidence from tests on complete frames
Created with novaPDF Printer (www.novaPDF.com)
40
and slab-beam-column assemblies has illustrated that this practice resulted in the gross
underestimation of beam flexural strength in the assumed plastic hinge regions (at the face
of beam column connections). This neglected source of beam flexural over strength has
significant consequences in the realization of the objectives of the established capacity
design frame work for reinforced concrete (RC) where beam shear design, joint
dimensioning, and column flexure/shear detailing are controlled by the requirement of
beam flexural yielding.
Experimental studies have shown that at large drifts, the entire width of the slab might be
engaged as additional tension reinforcement to the beams subjected to hogging moments.
Therefore, in the design, the effect of increased slab participation on structural stiffness,
bar cut-offs, beam shear demands to be considered.
2.7.4 New Zealand Code (SANZ, 1995) recommendations
According to the New Zealand code (SANZ, 1995) recommendations, only some of the
reinforcement in slabs parallel and integrally built with a beam can be taken into
consideration in resisting negative moments at the supports of continuous beams. When
earthquake induced moments are to be resisted the tensile and compression forces in the
beams must be transferred to the core of the column beam joints. The effectiveness of
force transfer to the joint core from slab bars, situated a large distance from the column, is
doubtful. On the other hand the moment input capacity of the beams to the columns during
large inelastic lateral displacements of the frame must not be grossly underestimated if
columns are to be protected against early yielding. Code intended to permit the inclusion
of the slab steel, within the prescribed width limits, into the evaluation of the negative
Created with novaPDF Printer (www.novaPDF.com)
41
moment of resistance of the section and to require it to be considered when the over
strength of the section is being assessed.
Where transverse beams of comparable size to that under consideration, frame into a
column, a larger slab width is considered in recognition of a more efficient force transfer
to the column beam joint core. The four cases normally encountered are illustrated in
Figure 2-13.

Figure 2-13: All longitudinal steel placed within shaded area to be included in flexural resistance of
beam [After-(SANZ, 1995)]
2.7.5 Research work carried out by Scribner et al. (1982)
Scribner et al. have studied the influence of different arrangement of ligatures on the
behaviour of doubly reinforced flanged concrete T- beams during repeated reversed
inelastic flexure loading. The types of ligatures used are shown in Figure 2-14.
Created with novaPDF Printer (www.novaPDF.com)
42
Type-1
Type-2
Type-3

Figure 2-14: Different ligatures configurations used
The results indicated that the use of type1 and type 2 ligatures has no significant effect on
the cyclic flexural behaviour. However, the use of type 3 ligatures indicated little loss of
cyclic flexural capacity.
It was found that the shear requirement in rib slab design was not critical for the specimen
used in this project. The experimental details are given in chapter 3. The design
calculations for the test specimen are presented in AppendixB. Therefore, a V-shape
ligature (similar to type-3) to suit the rib slab construction was used in the rib beam test
assembly, as this will increase ease of construction.












Created with novaPDF Printer (www.novaPDF.com)
43
Abdollahi, A. (1996). "Numerical Strategies in the Application of the FEM to RC
Structures-1." Computers & Structures, Vol. 58, No. 6,pp1171-1182.
ACI-318 (2002). Building Code Requirements for Structural Concrete (ACI-318M-99)
and Commentary (ACI-318RM-99), American Concrete Institute, Farmington Hills, Mich.
ANSYS (2003). ANSYS v 8.0, Swanson Analysis System, U.S.
AS1170.4 (1993). Minimum Design Loads on Structures, Part 4: Earthquake loads,
Standard Association of Australia, Sydney, Australia.
AS3600 (2001). Concrete Structures, Standard Association of Australia, Sydney,
Australia.
ATC40 (1996). Seismic evaluation and retrofit of concrete buildings.Applied Technology
Council Report no ATC-40,Redwood City, California: ATC. A practical account of the
seismic evaluation approaches including the pushover analysis.
Bentz, E. C. and M. P. Collins (2000). RESPONSE-2000 Reinforced Concrete Sectional
Analysis Program Manual.
Bommer, J. J., A. S. Elnashai, G. O. Chlimintzas and D. Lee (1988). Review and
development of response spectra for displacement-based seismic design. Research report
no ESEE-98/3, Engineering Seismology and Earthquake Engineering Section, Imperial
College, London.
BS8110 (1997). Structural use of concrete, Part 1: Code of practice for design and
construction, British Standard Institute, London.
Carr, A. J. (1994). "Dynamic Analysis of Structures." Bulletin, New Zealand National
Society for Earthquake Engineering, Vol. 27, No 2, June 1994, pp129-146.
Carr, A. J. (1998). RUAUMOKO User Manual, University of Canterbury.
Chopra, A. K. (1995). Dynamics of structures: theory and applications to earthquake
engineering. Englewood Cliffs, New Jersey: Prentice Hall.
Cook, R. D., D. S. Malkus, M. E. Plesha and R. J. Witt (2002). Concepts and applications
of finite element analysis, fourth ed. John Wiley & Sons, New
York, 2002.
Dragh, P., et al (2000). "Tradition with Cost Savings." Concrete Engineering
International(March/April).
Durrani, A. J. and J. K. Wight (1985). "Behavior of Interior Beam-to-Column Connections
Under Earthquake-Type Loading." ACI Structural Journal 82(3): 343-349.
Mander, J. B., M. J. N. Priestley and R. Park (1988). "Theoretical Stress-Strain Model for
Confined Concrete." Journal of Structural Engineering, Vol. 114, No. 8, Augest 1998.
Created with novaPDF Printer (www.novaPDF.com)
44
Miranda, E. (1996). "Assessment of the seismic vulnerability of existing buildings."
Proceedings 11th World Conference on Earthquake Engineering, Acapulco, Mexico, 1996.
Paper no 513. New York: Elsevier Science.
NZS4203 (1992). "Code of Practice for General Structural Design and Design Loading for
Buildings, Standards New Zealand,New Zealand."
Park, R. (1988). Ductility Evaluation from Laboratory and Analytical Testing-State of the
Art Report, 9th World Conference on Earthquake Engineering, Tokyo-Kyoto, Janpan,
pp605-616.
Park, R. (1992). "Capacity design of ductile RC building structures for earthquake
resistance." The Structural Engineer, Vol. 70, No.16, Aug. 1992., pp 279-288.
Paulay, T. and M. J. N. Priestley (1992). Seismic design of reinforced concrete and
masonary buildings. USA, John Wiley & Sons,Inc.
Priestley, M. J. N. (1998b). "Brief Comments on Elastic Flexibility of Reinforced
Concrete Frames and Significance to Seismic Design." Bulletin, New Zealand National
Society for Earthquake Engineering, Vol. 31, No 4, December 1998, pp246-259. Vol.
31(No.4): 246-259.
Priestley, M. J. N. and M. J. Kowalaky (2000). "Direct Displacement-Based seismic
Design of Concrete Buildings." Bulletin, New Zealand National Society for Earthquake
Engineering, Vol. 33, No 4, December 2000, pp421-444.
Quintero-Febres, C. G. and J. K. Wight (1997). "Investigation on the Seismic Behaviour
of RC Interior Wide Beam-Column Connections." The University of Michigan, USA,
UMCEE 97-15.
Sanders, P. T. (1995). "Design and Detailing for Seismic Resistance- An Australian
perspective in Earthquake- Earthquake resistant design for reinforced concrete structures."
Steel Reinforcement Institute of Australia, pp 33-41.
SANZ (1995). NZS3101:Part 1&2: 1982- Concrete Structures Standard, Standards New
Zealand, Wellington [2 Vols: Code & commentry].
Scribner, C. F. and D. R. Wilhelm (1982). "Behaviour of T-Beam Sections with Varied
Shear Reinforcement." ACI Structural Journal,March-April 1982, pp 139-146.
Shao-Yeh, M. M., V. B. Vitelmo and P. P. Egor (1976). Experimental Studies on the
Hysteretic Behaviour of Reinforced Concrete Rectangular and T-Beams, Earthquake
Engineering Research Center, University of California, Berkeley, California.
Smith, B. S. and A. Coull (1991). Tall building structures, John Wiley, New York, 537 p.
Takeda, T., M. A. Sozen and N. N. Nielsen (1970). "Reinforced concrete response to
simulated earthquakes." Journal of Structural Division, A.S.C.E 96(ST12) : 2557-2573.
Created with novaPDF Printer (www.novaPDF.com)
45
William, K. J. and E. P. Warnke (1975). Constitutive Model for the Triaxial Behavior of
Concrete. Proceedings, International Association for Bridge and Structural Engineering,
Vol. 19, ISMES, Bergamo, Italy.

Created with novaPDF Printer (www.novaPDF.com)

You might also like