You are on page 1of 424

Strength and ductility of high-strength concrete shear walls

under reversed cyclic loading

Author:
Dabbagh, Hooshang
Publication Date:
2005
DOI:
https://doi.org/10.26190/unsworks/17186
License:
https://creativecommons.org/licenses/by-nc-nd/3.0/au/
Link to license to see what you are allowed to do with this resource.

Downloaded from http://hdl.handle.net/1959.4/27467 in https://


unsworks.unsw.edu.au on 2023-08-03
STRENGTH AND DUCTILITY OF
HIGH-STRENGTH CONCRETE SHEAR WALLS
UNDER REVERSED CYCLIC LOADING

by
Hooshang Dabbagh

A thesis submitted as partial fulfilment of


the requirements for the degree of
Doctor of Philosophy

School of Civil and Environmental Engineering


The University of New South Wales
Sydney, Australia

December 2005
CERTIFICATE OF ORIGINALITY

I hereby declare that this submission is my own work and to the best of my knowledge
it contains no material previously published or written by another person, nor material
which to a substantial extent has been accepted for the award of any other degree or
diploma at the University of New South Wales or any other educational institution,
except where due acknowledgment is made in the thesis. Any contribution made to the
research by others, with whom I have worked at the University of New South Wales or
elsewhere, is explicitly acknowledged in the thesis.

I also declare that the intellectual content of this thesis is the product of my own work,
except to the extent that assistance from others in the project’s design and conception or
in style, presentation and linguistic expression is acknowledged.

_______________________
Hooshang Dabbagh
Dedicated to my wife, Farnaz, and my children,
Koshiar and Katayoun
ABSTRACT

This study concerns the strength and behaviour of low-rise shear walls made from
high-strength concrete under reversed cyclic loading. The response of such walls is often
strongly governed by the shear effects leading to the shear induced or brittle failure. The
brittle nature of high-strength concrete poses further difficulties in obtaining ductile
response from shear walls.

An experimental program consisting of six high-strength concrete shear walls was carried
out. Specimens were tested under inplane axial load and reversed cyclic displacements with
the test parameters investigated being longitudinal reinforcement ratio, transverse
reinforcement ratio and axial load. Lateral loads, lateral displacements and the strains of
reinforcement in edge elements and web wall were measured. The test results showed the
presence of axial load has a significant effect on the strength and ductility of the shear
walls. The axially loaded wall specimens exhibited a brittle behaviour regardless of
reinforcement ratio whereas the specimen with no axial load had a lower strength but higher
ductility. It was also found that an increase in the longitudinal reinforcement ratio gave an
increase in the failure load while an increase in the transverse reinforcement ratio had no
significant effect on the strength but influenced the failure mode.

A non-linear finite element program based on the crack membrane model and using
smeared-fixed crack approach was developed with a new aggregate interlock model
incorporated into the finite element procedure. The finite element model was corroborated
by experimental results of shear panels and walls. The finite element analysis of shear wall
specimens indicated that while strengths can be predicted reasonably, the stiffness of edge
elements has a significant effect on the deformational results for two-dimensional analyses.
Therefore, to capture the deformation of walls accurately, three-dimensional finite element
analyses are required.

The shear wall design provisions given in the current Australian Standard and the Building
Code of American Concrete Institute were compared with the experimental results. The
comparison showed that the calculated strengths based on the codes are considerably
conservative, specially when there exists the axial load.

i
ACKNOWLEDGEMENTS

I would like to express my deepest appreciation and thanks to my supervisor Associate


Professor Stephen J. Foster for his continual support, guidance and encouragement
throughout the course of this research project and the writing stages of the thesis.

I would like to thank Professor B. V. Rangan at the Curtin University of Technology for
his interest in this research and serving as co-supervisor.

Special thanks to Mark D’urso for his great cooperation and assistance in undertaking
the experimental tests.

I am thankful to the technical and laboratory assistance provided by the staff of the
School of Civil and Environmental Engineering, Heavy Structure Laboratory in
Randwick (Ronald Moncay, Frank Scharfe and Saffwan Ramadan) and the Concrete
Materials Laboratory in UNSW Kensington campus (Bill Terrey). Special thanks to
Chris Gianopoulos in Randwick Heavy Structure Laboratory for his great assistance and
managements during the experimental works.

In the undertaking this research, I have been supported by a scholarship from the Iranian
Ministry of Science, Research and Technology. This support is thankfully
acknowledged. The laboratory testing was funded through an Australian Research
Discovery Grant (Foster and Rangan 2001-2003) and this support is also gratefully
acknowledged.

My gratitude is extended to Shamseddin Nejadi, Kak Tien Chong, Adnan Malik,


Gregory Lee and Mohammad Bazyar for their friendship and discussion which have
made my study at the University of New South Wales such an enjoyable experience.

Special thanks to my parents for their love and prayers that made this possible.

Finally, very special thanks to my wife, Farnaz, and my children, Koshiar and
Katayoun, for their many sacrifices during the development of this research. Without
their love, encouragement and support, this thesis would not have been possible to
complete.

ii
TABLE OF CONTENTS

ABSTRACT i

ACKNOWLEDGEMENT ii

TABLE OF CONTENT iii

NOMENCLATURE ix

CHAPTER 1
INTRODUCTION
1.1 General 1-1
1.2 Structural Characteristics of Shear Walls 1-2
1.3 Research Significance 1-6
1.4 Objectives 1-7
1.5 Scope 1-8
1.6 Organisation of Thesis 1-10

CHAPTER 2
LITERATURE REVIEW
2.1 Introduction 2-1
2.2 Constitutive Models 2-1
2.2.1 General 2-1
2.2.2 Concrete 2-2
2.2.3 Reinforcing Steel 2-14
2.2.4 Shear Transfer 2-17
2.3 Finite Element Analysis 2-35
2.3.1 Overview 2-36
2.3.2 Finite Element Analysis of RC Panels and Shear Walls 2-44

iii
2.4 Experimental Studies 2-52
2.4.1 Barda et al (1977) 2-53
2.4.2 Oesterle et al. (1975, , 1978, , 1984) 2-55
2.4.3 Maier and Thurlimann (1985) 2-57
2.4.4 Lefas et al. (1990) and Lefas and Kotsovos (1990) 2-59
2.4.5 Pilakoutas and Elnashai (1995) 2-61
2.4.6 Gupta and Rangan (1996) 2-63
2.4.7 Kabeyasawa and Hiraishi (1998) 2-66
2.4.8 Palermo and Vecchio (2002) 2-68
2.4.9 Farvashany (2004) 2-71
2.5 Summary 2-72

CHAPTER 3
FINITE ELEMENT MODELLING
3.1 Introduction 3-1
3.2 Crack Membrane Model 3-2
3.2.1 Background 3-2
3.2.2 Concepts and Relationships 3-3
3.3 Constitutive Relationships 3-10
3.3.1 Constitutive Models for Concrete 3-10
3.3.2 Constitutive Model for Reinforcing Steel 3-16
3.3.3 Constitutive Models for Interaction between Concrete 3-16
and Steel
3.4 Finite Element Procedure 3-29
3.4.1 Finite Element Formulation 3-30
3.4.2 Implementation 3-36

CHAPTER 4
NUMERICAL EXAMPLES
4.1 Introduction 4-1
4.2 Monotonic Loading 4-2
4.2.1 Shear Panels 4-2

iv
4.2.2 Shear Walls 4-10
4.2.3 Shear-Critical Reinforced Concrete Beams 4-13
4.3 Cyclic Loading 4-18
4.3.1 Shear Panels 4-18
4.3.2 Shear Walls 4-22
4.4 Conclusion 4-25

CHAPTER 5
EXPERIMENTAL PROGRAM
5.1 Introduction 5-1
5.2 Test Parameters 5-1
5.3 Test Specimens 5-2
5.3.1 Dimensions 5-2
5.3.2 Reinforcement Layout 5-4
5.3.3 Material Properties 5-7
5.4 Construction of Test Specimens 5-14
5.4.1 Formwork 5-15
5.4.2 Assembling the Reinforcement 5-18
5.4.3 Casting 5-18
5.5 Instrumentation 5-23
5.5.1 Strain Gauges 5-23
5.5.2 Linear Variable Displacement Transducers (LVTDs) 5-27
5.6 Test Set-up 5-28
5.7 Testing Procedure 5-29
5.7.1 Specimen Installation and Loading 5-34
5.7.2 Data Recording 5-35

CHAPTER 6
EXPERIMENTAL RESULTS
6.1 Introduction 6-1
6.2 Observed Response of Specimens 6-2
6.2.1 Specimen SW1 6-2

v
6.2.2 Specimen SW2 6-12
6.2.3 Specimen SW3 6-23
6.2.4 Specimen SW4 6-32
6.2.5 Specimen SW5 6-41
6.2.6 Specimen SW6 6-47
6.2 Out-of-plane Displacement 6-57
6.4 Summary of Test Results 6-58

CHAPTER 7
ANALYSIS OF EXPERIMENTAL RESULTS
7.1 Introduction 7-1
7.2 Effects of Experimental Parameters 7-1
7.2.1 Transverse and Longitudinal Reinforcement 7-1
7.2.2 Axial Load 7-6
7.3 Test Results vs. Those of Gupta and Rangan (1996) 7-7
7.4 Comparison of the Wall Strengths with Design Code Predictions 7-11
7.5 Simplified Strut-and-Tide Model by Rangan (1997) 7-14
7.6 Finite Element Analysis of Shear Wall Specimens
7.6.1 Two Dimensional Finite Element Analysis 7-15
7.6.2 Three Dimensional Finite Element Analysis 7-22

CHAPTER 8
CONCLUSIONS
8.1 Summary 8-1
8.2 Concluding Remarks 8-3
8.3 Recommendations for Future Study 8-5

REFERENCES R-1

APPENDIX A
FINITE ELEMENT IMPLEMENTATION
A.1 Type of Element A-1

vi
A.2 Nonlinear Solution Procedure A-3
A.3 Convergence Criteria A-6
A.4 Program RECAP A-7

APPENDIX B
RESPONSE OF LVDTS
B.1 Specimen SW1 B-2
B.2 Specimen SW2 B-6
B.3 Specimen SW3 B-11
B.4 Specimen SW4 B-16
B.5 Specimen SW5 B-21
B.6 Specimen SW6 B-26

APPENDIX C
LOADS OF MACALLOY BARS
C.1 Specimen SW1 C-2
C.2 Specimen SW2 C-5
C.3 Specimen SW3 C-8
C.4 Specimen SW5 C-11
C.5 Specimen SW6 C-14

APPENDIX D
STRAINS OF EDGE ELEMENT REINFORCEMENT
D.1 Specimen SW1 D-2
D.2 Specimen SW2 D-5
D.3 Specimen SW3 D-7
D.4 Specimen SW4 D-10
D.5 Specimen SW5 D-13
D.6 Specimen SW6 D-16

vii
APPENDIX E
STRAINS OF WALL REINFORCEMENT
E.1 Specimen SW1 E-2
E.2 Specimen SW2 E-5
E.3 Specimen SW3 E-8
E.4 Specimen SW4 E-11
E.5 Specimen SW5 E-14
E.6 Specimen SW6 E-17

APPENDIX F
DESIGN CODE CLAUSES FOR SHEAR WALLS
F.1 Australian Standard (AS 3600, 2001) F-1
F.1.1 Strength in Flexural F-1
F.1.2 Strength in Shear F-3
F.2 American Concrete Institute Code (ACI 318, 2002) F-5
F.2.1 Strength in Flexural F-5
F.2.2 Strength in Shear F-5
F.3 Sample Calculation F-7

APPENDIX G
STRUT-AND-TIE MODEL

viii
NOMENCLATURE

A area of finite element

Ag gross cross-sectional area of shear wall

Al cross-sectional area of longitudinal reinforcement in shear wall

An average contact areas for a unit crack surface in the normal direction

As cross-sectional area of reinforcement

At average contact areas for a unit crack surface in the tangential


direction

Avs cross-sectional area of shear reinforcement in shear wall

a effective aggregate size

a1 , a2 Menegotto and Pinto parameters representing the Bauschinger effect


on the cyclic response of steel reinforcement

B strain displacement matrix

bf width of flange in shear wall

sec
D12 secant constitutive matrix in the 1_2 coordinate system

Dc12 constitutive matrix of the cracked concrete in the 1-2 coordinate


system

D cr12 constitutive matrix of crack in the orthotropic 1-2 coordinate system

D cts constitutive matrix of concrete tension stiffening

Dmax maximum aggregate size

Dsec
nt secant constitutive matrix in the n - t coordinate system

Ds constitutive matrix of steel reinforcement

D sc12 constitutive matrix of intact concrete between the cracks in the


orthotropic 1-2 coordinate system

ix
D xy constitutive matrix of reinforced concrete element in the global
X-Y coordinate system

dN distance from extreme compressive fibre to the neutral axis of wall


section

dw effective transverse length of shear wall

E modulus of elasticity

Ec initial elastic modulus of concrete

Ec1 , Ec 2 uniaxial secant moduli of concrete in the orthotropic 1-2 coordinate


system

Ects secant modulus corresponding to concrete tension stiffening

Ectsx , Ectsy secant moduli due to concrete tension stiffening in X and Y directions

Es elastic modulus of steel reinforcement

Es sec secant elastic modulus of steel reinforcement

E sx , E sy secant elastic moduli of steel reinforcement in X and Y directions,


respectively

Ew strain hardening modulus of elasticity for steel reinforcement

Fc internal compressive force due to the contribution of concrete in shear


wall

Fcr strength of shear wall corresponding to the onset of cracking

Fs internal force resulted from all steel bars in shear wall

Fy yield strength of shear wall

f c* biaxial compressive strength of concrete

f c' cylinder compressive strength of concrete

f cc cube compressive strength of concrete

f cp concrete peak stress

x
f ct tensile strength of concrete

f si stress in each layer of steel reinforcing bars

f su ultimate stress of reinforcement

f sy yield stress of reinforcement

G shear modulus of elasticity

Gc shear modulus of concrete

Gc12 secant shear moduli of cracked concrete in the 1-2 coordinate system

Gcr shear modulus of cracked concrete

Gcr12 secant shear modulus of crack in the orthotropic 1, 2-coordinate


system

Gf fracture energy

Gsc12 secant shear modulus of intact concrete between cracks in the


orthotropic 1-2 coordinate system

Hw height of shear wall

K stiffness of shear wall

K stiffness matrix

Kc stiffness matrix for concrete

Ks stiffness matrix for reinforcement

K ( wcr ) effective ratio of contact area in crack

k element stiffness matrix in the X-Y coordinate system

k decay factor in the concrete stress-strain relationship

k3 factor that accounts for the difference in compressive strengths of


in-situ concrete with the concrete test cylinder

Lw length of shear wall

l length of reinforcement throughout element

xi
lch characteristic length over which the fracture energy is dissipated

M* bending moment at the section of shear wall

N* axial load at the section of shear wall

n factor in the concrete stress-strain relationship equal to Ec ( Ecp − Ec )

pk ratio of total volume of the aggregates over the total volume of the
concrete

pl ratio of longitudinal reinforcement in shear wall

pt ratio of transverse reinforcement in shear wall

pw ratio of shear reinforcement in shear wall

R Bauschinger parameter used in Menegotto and Pinto model for steel


reinforcement

R0 Menegotto and Pinto parameter representing the Bauschinger effect on


the cyclic response of steel reinforcement

r displacement ratio ( δ cr / wcr )

s spacing of reinforcement

s rm average crack spacing measured normal to the cracks

srmx 0 , srmy 0 average crack spacings of uniaxial tension chords in the X and Y
directions, respectively

T transformation matrix

Tε strain transformation matrix

tf thickness of flange in shear wall

tw thickness of web in shear wall

V* shear force at the section of shear wall

Vu shear strength of shear wall

Vuc shear strength contributed by concrete in shear wall

xii
Vus shear strength contributed by shear reinforcement in shear wall

wcr crack width

α parameter to determine the rectangular stress block of compressive


concrete

α1 , α 2 , α 3 softening parameters of the concrete response in tension

αd dowel effect parameter

β scaling factor applied to the uniaxial compressive stress-strain curve


to determine the biaxial compressive stress-strain curve

βs shear retention factor

Δy displacement corresponding to the yield strength of shear wall

Δu displacement corresponding to the ultimate strength of shear wall

Δ εc total concrete strain increment

Δ ε cr crack strain increment

Δ ε ic intact concrete strain increment

Δσcx, Δσcy X- and Y-component concrete stresses due to tension stiffening

δ cr tangential displacement of crack surfaces

δs crack slip

{ε c } strain vector of the cracked concrete

{ε cr } strain vector of crack

{ε sc } strain vector of intact concrete between cracks

ε strain

ε1u , ε 2u equivalent uniaxial strains in the orthotropic 1-2 coordinate system

εa concrete strain just before beginning the unloading

εc concrete strain at the crushing taken as 0.003

xiii
ε c1 , ε c 2 average strains of concrete in the 1-2 coordinate system

ε cp
*
biaxial concrete strain corresponding to the biaxial concrete strength

ε cp strain corresponding to the concrete peak stress

ε cr1 , ε cr 2 average normal strains due to cracks in the 1-2 coordinate system

εe elastic component of concrete strain

εm average strain over the length of the element

ε nn , ε tt average normal strains in the n ( t ) direction

εo strain at the intersecting point of asymptotes in Menegotto and


Pinto model for steel reinforcement

εp plastic component of concrete strain

εr strain at the last strain reversal in Menegotto and Pinto model for steel
reinforcement

ε sc1 , ε sc 2 average strains of intact concrete between cracks in the orthotropic


1-2 coordinate system

ε sh strain corresponding to the onset of strain hardening in reinforcement

ε to strain corresponding to the concrete tensile strength

εy yield strain of steel reinforcement

∅ diameter of the reinforcing bar

∅x, ∅y diameter of the reinforcing bar in X (Y) directions

γ parameter to determine the rectangular stress block of compressive


concrete

γ c12 average shear strain of concrete in the 1-2 coordinate system

γ cr12 average shear strain due to cracks in the 1-2 coordinate system

γ nt average shear strain in the n - t coordinate system

γ sc12 average shear strain of intact concrete between cracks in the


1-2 coordinate system

xiv
η ratio of the value of concrete strain over the concrete strain
corresponding to the concrete peak stress ( ε c ε cp )

λ ratio of tensile stresses in the concrete due to tension stiffening over


the concrete tensile stress

λx , λy ratios of tensile stresses in the concrete due to tension stiffening in the


X and Y directions, respectively, over the concrete tensile stress

μ roughness coefficient of crack interface

μd measure of displacement ductility

μf coefficient of friction in crack interface

μn normal reduction factor

v12 , v21 Poisson’s ratios

θ inclination of contact areas in concrete crack

θc angle between major principal stress and the global X axis

θr angle between a vector normal to the cracks and the global X axis

ρ steel reinforcement ratio

ρx, ρy steel reinforcement ratios in the global X- and Y-directions,


respectively

ρv ratio of reinforcement crossing the crack interface

{σ c } stress vector of cracked concrete

{σ cr } stress vector of crack

{σ sc } stress vector of intact concrete between cracks

σ stress

σa concrete stress just before beginning the unloading

σ c1 , σ c 2 average concrete stresses in the orthotropic 1-2 coordinate system

σ cm average stress in concrete between cracks

xv
σcn,σct concrete stresses in the n - t coordinate system

σ con normal contact stress in crack

σ cr normal stress over crack surface

σ cr1 , σ cr1 average stresses over the crack surfaces in the orthotropic
1-2 coordinate system

σ ctsm average concrete tensile stiffening stress

σ nn , σ tt average normal stresses in the n - t coordinate system

σo stress at the intersecting point of asymptotes in Menegotto and


Pinto model for steel reinforcement

σr stress at the last stress reversal in Menegotto and Pinto model for steel
reinforcement

σ sc1 , σ sc 2 average stress of intact concrete between cracks in the orthotropic


1-2 coordinate system

σ sm average stress in steel bar between cracks

σ s min minimum stress in steel bar between cracks

σ sr stress in the steel bar at the cracks

σsx, σsy steel reinforcement stresses in the global X-Y coordinate system

σx, σy normal stress in the global X-Y coordinate system

τb bond shear stress

τ b0 plastic bond strength before yielding of the reinforcing steel

τ b1 plastic bond strength after yielding of the reinforcing steel

τ c12 average concrete shear stress in the orthotropic 1-2 coordinate system

τ cnt concrete shear stress in the n - t coordinate system

τ co focal point in the aggregate interlock relationship

τ cr shear stress over crack surface

xvi
τ cr12 average shear stress over the crack surfaces in the orthotropic
1-2 coordinate system

τ cr max maximum shear capacity of crack

τ nt average shear stresse in the n - t coordinate system

τu ultimate shear capacity due to aggregate interlock

τxy shear stress in the global X-Y coordinate system

Ω(θ ) distribution function of the crack surface orientations

ξ normalised strain history parameter in Menegotto and Pinto model for


steel reinforcement

ψ displacement ratio ( δ cr / wcr )

xvii
CHAPTER 1

INTRODUCTION

1.1 General

Reinforced concrete (RC) shear walls are frequently used in multistorey buildings to
resist the lateral loads due to wind forces and seismic effects on buildings and the
vertical loads due to dead and live loads transmitted by floors. So shear walls are
subjected to axial forces, bending moments and shear forces. In practice, there are two
different types of shear walls: cantilever and squat shear wall (Figure 1.1). The
cantilever shear walls act as a cantilever beam and its design is usually governed by
flexural behaviour. Squat shear walls are walls with a ratio of height to length less than
2 and are usually found in low-rise buildings or in the lower storeys of medium to
high-rise buildings (Paulay and Priestley, 1992). The response of such walls is often
strongly governed by the shear effects leading to shear induced brittle failure and tensile
cracks. In Australia and other western countries using cantilever shear walls is a popular
practice whilst the squat shear wall is common in Japan, China and Southeast Asia
(Mau and Hsu, 1986).

The use of shear walls as a lateral load resisting system emerged during the 1950s and
1960s, when the modern approach to earthquake engineering of building commenced.
Over the last two decades, the improvements in material technology and in the
production of high strength concrete (HSC) have resulted in a number of projects in the
(a) (b)

Figure 1.1 - (a) Squat shear wall (b) Cantilever shear wall.

Australia, United States, Canada, Europe and Japan that have been completed using
such concrete (Nawy, 2001, among many others). The HSC exhibits superior
performance allowing shear walls (eg. lift core walls) to be thinner, thereby increasing
the amount of rentable floor area. Nevertheless, the brittle nature of HSC being
considerably greater than that of normal strength concrete (NSC), poses many
difficulties for designers, particularly in obtaining ductile response from shear walls
constructed using HSC and subjected to the reversed cyclic loading.

1.2 Structural Characteristics of Shear Walls

To design shear walls against seismic actions, various levels of protection including the
preservation of functionality, the different degrees of damage and prevention of loss of
life are generally described. In conjunction with these three levels of seismic protection
the specific structural properties termed as stiffness, strength and ductility need to be
considered. A typical response of a reinforced concrete shear wall subjected to
monotonic loading is illustrated in Figure 1.2.

1-2
Δ
F

(a) Loading and the ensuing deformation

Load

Ductile failure
Fy

0.75 Fy Idealised response

Brittle failure
Fcr K
1

Displacement
Δy Δu

(b) Load-displacement relationship

Figure 1.2 – Typical response of RC shear wall (Paulay and Priestley, 1992).

The stiffness of a shear wall relates the lateral applied load to the resulting lateral
displacement. The stiffness is defined as the slope of idealised linear elastic response,
K, where

K = Fy / Δ y (1.1)

as shown in Figure 1.2 where Fy is the load at yield and Δ y is the corresponding

displacement at the point of yielding. For the real load-displacement curve, K can be
determined based on the effective secant stiffness at a load of 0.75 Fy (Paulay and

Priestley, 1992). The force Fcr , in Figure 1.2, is the load corresponding to the onset of

cracking and Δ u is the displacement corresponding to ultimate failure of the wall.

1-3
Strength is used as a criterion to evaluate the level of protection against the damage
due to seismic loading. Since the inelastic response during seismic loading is the main
source of concrete damage, the shear wall must have adequate strength to resist internal
actions by an elastic seismic response. Therefore, the desired strength can be expressed
in terms of the force Fy derived from an elastic analysis and based on the stiffness

described previously.

The ability of a shear wall to resist collapse associated with sustaining large
deformations is known as its ductility. Displacement ductility can be quantified by the
ratio of the total imposed displacement Δ u to the displacement at yielding. That is:

μ = Δu / Δ y (1.2)

where μ is the measure of displacement ductility.

Since concrete is an inherently brittle material, the reinforcing steel bars must be
distributed throughout the concrete shear wall in such a way that the composite can
exhibit a ductile response in order that it function under large inelastic deformations
caused by severe seismic loading. Therefore, detailing of shear walls becomes an
important issue for designers.

The failure of a well-detailed cantilever shear wall is usually in a flexural mode similar
to that of beams. However, depending on the different parameters such as geometrical
dimensions, boundary conditions, the way lateral loads are imposed and the
reinforcement detailing, squat shear walls may fail in any of three modes: diagonal
tension, diagonal compression or sliding shear (Paulay and Priestley, 1992). The
diagonal tension failure mode will occur whenever transverse reinforcement is
insufficient to carry shear forces or is insufficiently detailed (Figure 1.3a). When
adequate the transverse reinforcement is provided but the wall is subjected to a high
shear stress, concrete may crush under diagonal compression (Figure 1.3b). This is
common in the squat walls with edge elements. Finally, for walls with sufficiently
detailed transverse reinforcement but low quantities of longitudinal reinforcement in the

1-4
(a) Diagonal Tension (b) Diagonal Compression (c) Sliding Shaer

Figure 1.3 - Shear failure modes in squat shear walls (Paulay and Priestley, 1992).

web, failure can be due to yielding of longitudinal reinforcement leading to a sliding


displacement along base of the wall (Figure 1.3c). This last failure mode is particularly
important for walls subjected to cyclic reversals in displacement.

The cyclic behaviour of concrete shear walls shows pinched hysteretic loops with
significant strength and stiffness degradation as number of cycles increases (Figure 1.4).
The reduction in stiffness of the system is as a result of both the concrete cracking and
reinforcement yielding. The loss of stiffness results in the hysteresis curve becoming
shallower.
Load

Envelope Curve

Displacement

Pinching

Figure 1.4 – Cyclic Behaviour of Shear Walls.

1-5
The hysteresis loops also show a change of slope during the secondary loops resulting
in a significant increase of stiffness during reloading, known as pinching effect
(Paulay and Priestley, 1992) (Figure 1.4). This is caused by the closing of previously
formed cracks during the reloading phase. This effect can significantly reduce the
energy absorption at each cycle, represented by the area enclosed within each loop and
thus decrease the system damping.

1.3 Research Significance

Fintel (1991) documented the superiority of shear walls, over other load resisting
systems, to resist lateral forces resulting from earthquake events. This investigation as
well as an extensive number of other publications (as listed by Paulay and
Priestley, 1992) show that the ductility of shear walls is of paramount importance. To
design a RC shear wall to behave in a ductile manner two issues are critical
(Fintel, 1991). First, the wall as well as joints and other members in the building must
be appropriately detailed. Second, the strength of wall is required to be governed by the
flexural behaviour rather than in shear. In other words, because a shear failure is
significantly less ductile compared with flexural failure it should not be permitted to
occur. To achieve this, the shear capacity of a wall must be known and be larger than
the shear corresponding to its moment capacity. It is also important to be understood
what occurs between the beginning of shear cracking and the shear failure. However,
the lack of a similar level of confidence for the shear design of walls as is presently
available for the flexural design leads one to recognise an immediate need for
developing an analytical and experimental understanding of shear response of shear
walls (Fintel, 1991). A rational understanding of shear behaviour of concrete shear
walls in general, and HSC shear walls in particular, is therefore vital.

Extensive information is available with regard to design of normal strength concrete


(NSC) walls. In design methods, the flexural strength and the ductility capacity may be
calculated using the conventional analysis of reinforced concrete sections subjected to
axial force and bending moment (Paulay and Priestley, 1992). Since HSC is more brittle

1-6
than the NSC, it is important that designers can calculate the difference in the safety
between walls designed in the NSC and the HSC alternatives. However, there is a lack
of the experimental data and proper models necessary for this to be accomplished. More
importantly, the current models and design codes developed for the design of NSC
shear walls are empirical and cannot be directly transferred to the design of HSC shear
walls.

To ensure that analytical models predict the behaviour of shear walls accurately,
experimental data is required. Experimental data can also provide useful information for
better understanding the cyclic response of shear walls. While a large number of NSC
shear walls tested under monotonic and cyclic loading have been reported in the
literature (Barda et al., 1977, Cardenas et al., 1980, Maier and Thurlimann, 1985, Lefas
and Kotsovos, 1990, Lefas et al., 1990, Pliakoutas and Elnashai, 1995a, Palermo, 1998
among many others), only a limited number of research programs on HSC shear walls
subjected to monotonic loading (Kabeyasawa et al., 1993, Gupta and Rangan, 1996,
Farvashany, 2004) and cyclic loading (Kabeyasawa et al., 1993) have been carried out
and many important parameters affecting the cyclic behaviour of HSC walls remain
unclear. Hence, further experiments regarding the behaviour of HSC shear walls,
especially under cyclic condition, are required.

In summary, although shear walls provide good resistance against wind and earthquake
loads, the current information about the cyclic response of such walls cast using HSC is
not sufficient and further experimental and analytical information is needed so that they
can be designed safely and detailed appropriately.

1.4 Objectives

The main aim of this study is the investigation of strength and ductility of HSC shear
walls under reversed cyclic loading, so as to provide the analytical models and baseline
experimental data needed to safely design shear walls with high strength concrete under
seismic loads. The objectives and expected outcomes to reach this are as follows:

1-7
• To design and undertake an experimental program to investigate the general
response, ultimate strength and ductility characteristics of high strength concrete
squat shear walls loaded in axial compression and in cyclic shear stresses.

• To develop a finite element procedure and appropriate constitutive models for


predicting the strength and ductility of reinforced high strength concrete shear
walls.

• To verify the applicability of the finite element model by comparing the results
of analyses with the experimental observations available in the literature.

• Analysis of shear walls tested as a part of this research project to determine their
strength and ductility and comparing the analytical results with the experimental
data.

• To compare the strength of shear walls tested in the experimental program with
the calculated ultimate strengths based on the design codes.

1.5 Scope

This study is concerned with the behaviour and strength of framed HSC squat shear
walls subjected to in-plane constant vertical load and reversed cyclic loading
(Figure 1.5). The research involves an analytical and an experimental program.

The experimental program involves tests on wall specimens that represent


approximately one-third scale models of a prototype shear wall in a multistorey
building. The compressive strength of the concrete used in the walls is around 80 MPa.
To investigate the shear response and strength of HSC shear walls, the specimens were
designed to fail in a shear mode. The main parameters included in the study are the
vertical load, the longitudinal reinforcement ratio and the transverse reinforcement ratio.
Other factors such as overall dimensions of the test specimens and the reinforcement
ratio of top, bottom and edge elements were kept constant.

1-8
P

V Top Element

Wall
Edge Element

Bottom Element

Figure 1.5 - Typical shear wall specimen under loading.

The analytical program includes the development of the crack membrane model as a
two-dimensional smeared-fixed crack finite element model. Then, the model is
implemented into a computer program, verified against data found through the literature
and used to analyse the shear wall specimens with main emphasis on the investigation
of the strength and ductility of shear walls. The analysis of other cyclic aspects such as
hysteretic characteristics is not included in the analytical program.

The investigation of fatigue effects is outside the scope of this work. Also excluded
from the scope of this thesis, while not underestimating its importance, is the bond
deterioration between concrete and reinforcement.

1-9
1.6 Organisation of Thesis

In Chapter 2, a review of the literature is undertaken on the constitutive models for


reinforced concrete, the aggregate interlock effect, the finite element modelling of shear
walls, and the experimental works carried out on shear walls.

In Chapter 3, constitutive relationships for concrete and reinforcing steel are derived.
Then a finite element model for the analysis of reinforced concrete elements in shear
walls is developed.

In Chapter 4, the finite element model outlined in Chapter 3 is verified by comparing


the results obtained in numerical analyses with existing experimental test data over a
range of reinforced concrete panels, beams and shear walls.

In Chapter 5 a description of the experimental program carried out on 6 HSC squat


shear wall specimens including the test parameters, material properties, specimen
details, and testing is presented.

Chapter 6 contains a summary of the qualitative and quantitative results of the


experimental program. The main concentration is on the cyclic response, crack patterns,
failure mechanism, and strength and ductility aspects.

In Chapter 7, the experimental results of the shear wall specimens given in Chapter 6
are discussed. The wall strengths are also compared to those calculated based on design
codes as well as strut-and tie model. Then the results obtained from the analysis of test
specimens using the two-dimensional finite element procedure proposed in Chapter 3 as
well as a three-dimensional finite element model are compared with the experimental
data and discussed.

Finally in Chapter 8 the results and outcomes of the thesis are summarised and
conclusions and recommendations for future research are given.

Other pertinent information is listed in the Appendices A to G at the end of the report.

1-10
CHAPTER 2

LITERATURE REVIEW

2.1 Introduction

In this chapter a review of the previous works carried out to investigate the structural
behaviour of reinforced concrete shear walls is presented. The literature review focuses
on three areas related to the main objectives of this research including constitutive
modelling, the finite element analysis and experimental works. A summary of major
outcomes obtained from the literature and the proposed research framework is
expressed at the end of this chapter.

2.2 Constitutive Modelling

2.2.1 General

The challenge in numerical modelling of reinforced concrete arises from its composite
nature. The important aspects such as cracking, crushing, tension stiffening,
compression softening, aggregate interlock and bond-slip cause non-linear behaviour of
reinforced concrete members. Under cyclic loading, there are further complexities
including stiffness degradation in concrete, the Bauschinger effect in reinforcing steel,
bond degradation, crack opening and closing and unloading and reloading process. The
accuracy and reliability of numerical modelling of reinforced concrete is governed by
the underlying constitutive relationships used in analysis.

Over the past decades many constitutive models have been proposed to analyse
reinforced concrete structures that can be classified into elasticity-based, plasticity-
based, progressive damage-based, micromechanics and endocrinic models (CEB, 1996).
An extensive review of these models is available in the literature (ASCE Committee
447, 1982, Chen, 1982, ASCE Committee 447, 1993, CEB, 1996). Amongst these,
elasticity-based models that use a Hookean formulation in incremental form are
frequently used (ASCE Committee 447, 1982, CEB, 1996). In the following, a review
of the constitutive models expressed for the concrete and the reinforcing steel are
presented. The emphasising is on those using in the elasticity-based approach.

2.2.2 Concrete

Concrete in Compression

One of the first experimental investigations into the behaviour of plain concrete under
cyclic loading was conducted by Sinha et al. (1964). The experiment was undertaken on
concrete cylinders with compressive strengths from 20 to 28 MPa and subjected to
repeated axial compressive loading in order to determine the main factors governing the
cyclic response of concrete. To investigate the effects of load history, the load cycles
were applied in two different manners including complete and partial unloading.
Figure 2.1 shows the response of concrete under cyclic loading. Based on the test
results, the following qualitative conclusions were drawn:

• The stress-strain paths caused by cyclic loading do not go beyond an envelope


curve regardless of the previous load history. Such envelope curve can be
considered unique for the each strength of concrete and expressed by stress-strain
curve obtained under monotonic compressive loading to failure.

2-2
• The unloading and reloading paths present different curves such that a hysteresis
feature reflecting the energy dissipation is exhibited in each cycle. Two
distinctive pattens represented by two mathematical families of curves are
recognised for these curves.

• The locus of points where the unloading and reloading curves of each cycle
intersected may be defined as the shakedown limit at which the strains stabilise.
Stresses above this shakedown limit produce additional stains showing further
damage due to cycling whereas maximum stresses at or below this limit lead to
the stabilisation of strains and a closed hysteresis loop is formed in subsequent
cycles.

• The value of shakedown limit depends on the minimum stress in the cycle so that
for complete unloading this limit is near the envelope curve while for partial
unloading it takes place at a lower value of stress.

Envelope Curve

Figure 2.1 – Experimental stress-strain curves of concrete under complete unloading


(Sinha et al., 1964).

2-3
To represent the concrete response analytically, a polynomial relationship was
adopted for the envelope curve. Also, the unloading and reloading paths were modelled
using parabolic and linear equations independent of the previous load history,
respectively, although subsequent studies (Karsan and Jirsa, 1969, Bahn and Hsu, 1998)
indicated the dependency of unloading and reloading responses on the previous load
history. The analytical cyclic response could show the test results qualitatively. Sinha et
al. (1964) introduced some main characteristics of the cyclic behaviour of concrete and
established a sound basis for the future studies in this area.

Based on a fracture mechanics treatment, Shah and Winter (1966a, 1966b) carried out a
series of tests on prismatic specimens subjected to cyclic axial compressive loading.
Test results indicated that the shakedown limit is approximately equivalent to the
critical load at which the number of microcracks in mortar begins to increase sharply,
and a continuous pattern of microcracks begins to form. As a result, undamaged
portions that carry the load reduce and the stress-strain relationship becomes even more
nonlinear. The onset of major microcracking was reported at 70% to 90% of the
ultimate load.

To gain further insight into the response of plain concrete under different cyclic
compressive loading histories, Karsan and Jirsa (1969) performed an experimental
study on 46 short rectangular concrete columns with the cylinder compressive strengths
from 24 to 35 MPa. The specimens were tested under four different loading regimes
including monotonic increasing loading to failure, cycles to envelope curve, cycles to
envelope curve adding a specified strain increment during each cycle and cycles
between maximum and minimum stress levels.

Generally, the envelope curve for the cyclic response of test specimens coincided with
the stress-strain curve obtained under monotonic loading to failure regardless the load
history. Defining the concept of common point as the intersecting point of unloading
and reloading curves at each cycle, it was concluded that the points of intersection due
to load cycles to the envelope curve represent an upper limit (shakedown limit) on the
common points, whereas the cycles with lower stress levels lead to reducing and
stabilising the common points at a lower bound. The lower and upper limits, termed as

2-4
the stability limit and the common point limit, respectively, were reported to be
' '
corresponding to the stress levels of 0.63 f c and 0.76 f c , i.e. 74% and 90% of the
specimen ultimate strength (Figure 2.2), showing the concrete response is largely
dominated by the effects of microcracking.

Researchers found out that unloading and reloading curves are not unique but
depend on the previous load history. Introducing the concept of nonrecoverable
strain (or plastic strains) as the strain corresponding to a zero stress on the unloading
or reloading curves, the shape of these curves was found to be significantly
influenced by this factor. Based on the test results, a parabolic relationship was
proposed to determine the normalised plastic strains using the normalised strains on
the envelope curve at the commencement of unloading. As well, a parabolic
relationship was assumed for reloading curve to address the additional strains for
stress beyond the common point. Adopting the relationship of Smith and Young
(1955) for the envelope curve, the test results could be quite well simulated by
analytical model.

Figure 2.2 – Loading and unloading curves (Karsan and Jirsa, 1969).

2-5
While the previous works concerned the plain concrete, Brown and Jirsa (1971) tried to
simulate the behaviour a series of 12 reinforced concrete cantilever beams subjected to
reversed cyclic loading using the stress-strain curves for concrete and for the reinforcing
steel. The concrete model in compression was generally based on studies by Karsan
and Jirsa (1969). To address the reversal effects they considered that after cracking no
concrete compressive stress is developed until the cracks are completely closed. This
was experimentally found to occur whenever the tensile strain of concrete reaches a
value of 40% of the tensile strain corresponding to concrete cracking. The model was
corroborated against the load-deflection and load-rotation responses obtained from tests
and a reasonable agreement was observed. It should be noted that since the experimental
responses of RC beams were significantly dominated by the reinforcement response, the
concrete model was not as critical.

Park et al. (1972) studied experimentally and theoretically the moment-curvature and
load-deflection responses for reinforced concrete beams and columns subjected to cyclic
loading. The cyclic model used for concrete is shown in Figure 2.3. For unloading it
was assumed that 0.75 of previous stress is lost with no reduction in strain followed by a
linear path of slope 0.25 Ec to a zero stress level where Ec is the initial elastic modulus

of concrete. For reloading the strain had to re-reach the value corresponding to the zero
stress level of the last unloading and after that the path was similar to that of unloading,
as shown in Figure 2.3. The tensile response of concrete was considered brittle linear
elastic with no tensile strength after cracking. Although the concrete model could be
simply implemented in a numerical procedure, it suffered three major drawbacks. First,
the plastic strains are independent of the loading history. Second, the reloading path
terminates at the last unloading stress so that the damage attributed to the cycling
loading is ignored. Finally, the tension stiffening effect was not incorporated in the
model. Although predicted responses compared well with the test results, the
experimental responses seem to be dominated by the reinforcement response, similarly
to the specimens tested by Brown and Jirsa (1971), and therefore the concrete model
was not as critical.

2-6
Figure 2.3 – Stress-strain curve for concrete under cyclic loading (Park et al., 1972).

Darwin and Pecknold (1976) carried out one of the first finite element analysis of
reinforced concrete panels subjected to cycling loading. The model for concrete under
cyclic loading (Figure 2.4) was developed based on the experimental results of Karsan
and Jirsa (1969). The unloading and reloading curves were considered to terminate and
initiate at the plastic strain points, respectively. The reloading curve was represented by
a line initiated from the plastic strain, passed through the common point and ending at
the envelope curve. The unloading path was modelled using three straight lines with the
slopes of initial elastic modulus, the same as the slope of reloading line and zero,
respectively. The first and second lines intersected at the turning point determined
experimentally. The concrete in tension was assumed as a linear elastic brittle material
so that after cracking the tensile tangential stiffness along the principal tensile stress
direction was reduced to zero. The concrete model along with that of steel
reinforcement was incorporated into a finite element program and corroborated against
the response of shear panel W-4 tested by Cervenka and Gerstle (1972). While the
predicted results were in agreement with the first cycle, they deviated gradually during
the second cycle.

2-7
Figure 2.4 – Stress-strain curve for concrete under cyclic loading (Darwin and
Pecknold, 1976).

Concrete in Tension

Similar to the concrete in compression, the tensile behaviour of concrete subjected to


cyclic loading can be modelled using two groups of curves including a stress-strain
envelope curve and the unloading and reloading curves. Test results have shown that the
envelope curve can be approximated by the stress-strain curve of concrete under
monotonic loading (Yankelevsky and Reinhardt, 1989). The main characteristics of this
curve is as follows (Hillerborg, 1980, ASCE Committee 447, 1982, CEB, 1996):

• The stress-strain curve consists of an ascending branch before ultimate strength


and a descending branch after that (Figure 2.5(a)).

• The stress-strain relationship in ascending part is linear almost up to the peak


stress with an elastic modulus equal to the initial tangential modulus of elasticity
in compression.

2-8
• The descending curve qualitatively includes a very steep branch and then a more
flat branch. Since the exact shape of stress-strain descending curve does not
have significant influence on the most of practical applications, its selection is
primarily a matter of practical convenience.

While a linear relationship for ascending branch is commonly considered, for the
descending branch a variety of curves such as a linear model (Hillerborg et al., 1976),
bilinear model (Petersson, 1981), multilinear model (Gustafsson, 1985) and more
complex nonlinear models (Lin and Scordelis, 1975, Cornelissen et al., 1985,
Gopalaratnam and Shah, 1985) have been proposed. Figures 2.5 (b) and (c) show the
linear and bilinear models, where f ct and ε t 0 are the tensile strength and corresponded

strain, respectively; α1 , α 2 and α 3 are softening parameters dependant on fracture


energy. These parameters were proposed by Hillerborg et al. (1976) and
Petersson (1981), respectively, as follows:

2 Ec G f
α1 = 0 ; α 2 = α 3 = (2.1)
lch f ct2

and

1 2 18 Ec G f
α1 = ; α 2 = α 3 + α1 ; α 3 = (2.2)
3 9 5 lch f ct2

where Ec is the initial elastic modulus of concrete, G f is the fracture energy and lch is

a characteristic length over which the fracture energy is dissipated.

2-9
fct fct fct

α1 fct

εto εto α3εto εto α 2εto α3 εto

(a) (b) (c)

Figure 2.5 – Tensile envelope curve of concrete (a) typical response (b) linear
descending model (Hillerborg et al., 1976) (c) bilinear descending model
(Petersson, 1981).

When the cracking occurs in concrete, the stresses are redistributed throughout the
concrete element. The stress redistribution results in non-monotonic variation of
stresses, even for monotonic external loading. Therefore, unloading and reloading
behaviours are significant not only for cyclic loading but also for monotonic loading
and should be incorporated into the concrete model. Yankelevsky and Reinhardt
(1987b, 1989) conducted a series of tests to investigate the tensile cyclic response of
concrete subjected to different types of loading including monotonic displacement-
control tensile test, complete unloading, reversed loading up to a low level of
compressive stress and reversed loading until a greater compressive stress level (type 1
to 4, respectively) (Figure 2.6). It was found that the curvature of the unloading and
reloading branches was related to the strain at unloading and the strain at the reloading,
respectively.

2-10
Figure 2.6 – Uniaxial cyclic response of concrete in tension for different loading regimes (Yankelevsky and Reinhardt, 1987b, 1989).

2-11
The number of proposed models for unloading and reloading response is limited
(Gylltoft, 1984, Reinhardt, 1984, Rots et al., 1985, Yankelevsky and Reinhardt, 1987b,
1989). Prior to the peak stress, both unloading and reloading paths are usually assumed
to be the same(linear elastic) (ASCE Committee 447, 1982, CEB, 1996). After the peak
stress, however, different relationships have been suggested. Rots et al. (1985) assumed
that both unloading and reloading run on a same line starting from the origin of the
coordinate system (Figure 2.7(a)). This model approximates the real behaviour roughly
and is path-independent. Gylltoft (1984), considering a set of lines parallel to the initial
ascending branch for modelling the unloading and reloading behaviour, proposed a
better approximation to the real response of concrete (Figure 2.7(b)), although it is also
path-independent. Foster and Marti (2003) employed a simple model including the
history-dependant aspect to describe unloading and reloading (Figure 2.7(c)). The whole
strain was assumed to be composed of an elastic component ε e and a plastic

component ε p . The unloading was considered to cross through a sarin of ε p / 2 and

reloading path was assumed along the same path.

fct fct fct

Ec
1
Ec
ε ct ε ct ε p /2 εp ε ct

(a) (b) (c)

Figure 2.7 – Unloading and reloading models: (a) Rots et al. (1985), (b) Gylltoft (1984)
(c) Foster and Marti (2003).

2-12
Based on the test results, Yankelevsky and Reinhardt (1989) introduced a focal point
model to describe the response of concrete subjected to cyclic tension. Figure 2.8
illustrates the graphical procedures used to form a set of unloading and reloading paths.
This model presented a good approximation of cyclic response of concrete and can
capture the concrete damage on reloading as well as crack closing. The mathematics to
describe this is trivial.

Figure 2.8 – Focal point model of Yankelevsky and Reinhardt (1989).

2-13
2.2.3 Reinforcing Steel

The constitutive stress-strain relationship of reinforcing steel bars in the cyclic loading
is expressed using two groups of curves. The first group is the backbone envelope curve
and the second group includes unloading and reloading curves. One significant
consideration to select an appropriate constitutive model for reinforcing steel is its
numerical efficiency. Particularly, for a non-linear analysis of large RC structures with a
huge amount of numerical processing, the model should be as simple as possible (CEB,
1996).

Tests conducted by Belarbi and Hsu (1994) indicated that the monotonic stress-strain
curve of steel bars can be used to approximate the backbone envelope curve. This curve
generally consists of elastic, plastic, strain hardening and strain softening zones as
shown in Figure 2.9(a). In some grades of steel, however, the response may not include
a distinctive plastic zone (Figure 2.9(b)), thus the intersection of main stress-strain
curve and a line with a slope equal to initial elastic modulus and crossing through the
strain of 0.002 is considered as the effective yield point (refer section 5.3.3). In
Figure 2.9, f sy and ε y are stress and strain corresponding to yield point, respectively,

ε sh is the strain corresponding to the onset of strain hardening, and f su and ε u are the
ultimate stress and corresponded strain, respectively. While several analytical models
have been proposed for envelope curve of reinforcement response, more common
idealised stress-strain curves have been bilinear (Vecchio, 1989, Belarbi and Hsu, 1994,
Elmorsi et al., 1998, Kwak and Kim, 2004a, 2004b, Mansour and Hsu, 2005) and
trilinear (Stevens et al., 1987, Vecchio, 1999, Foster and Marti, 2003, Palermo and
Vecchio, 2003) as shown in Figure 2.10.

Depending on the analysis for monotonic or reversed cyclic loading, different types of
unloading and reloading curves may be considered. In the case of monotonic as well as
cyclic loading, the unloading and reloading are usually modelled using a similar linear
path with a slope equal to initial elastic modulus. This idealisation provides a reasonable
approximation for reinforcing steel (ASCE Committee 447, 1982, CEB, 1996). Under a
reversed cyclic condition, however, observations have shown that the beyond the initial

2-14
fs fs

strain softening zone strain softening zone


fsu
fsy fsy
strain hardening zone strain hardening zone
plastic zone
elastic zone
elastic zone

εy εsh εu εs 0.002 εy εs

(a) (b)

Figure 2.9 – Typical stress-strain curve for steel reinforcement: (a) with plastic zone
(b) without plastic zone.

fs fs
fsu fsu
fsy
fsy

εy εu εs εy εsh εu εs

(a) (b)

Figure 2.10 – Idealised stress-strain relationship for steel reinforcement: (a) bilinear
model (b) trilinear model.

yield stress, the unloading path exhibits the Bauschinger effect in the region of reversed
loading, which refers to the softening of the steel modulus (Bauschinger, 1887).

The models of the cyclic stress-strain response of steel reinforcement may be classified
in two major categories including macroscopic models and microscopic models (Popov
and Ortiz, 1979, CEB, 1996). The former is based on the measured stress-strain

2-15
relationship while the latter is based on the dislocation theory. Although the
microscopic models are obtained from sound theory, they are too complex to use in a
nonlinear analysis of RC structures (Balan et al., 1998). The most widely used models
to express the hysteretic response of reinforcing steel fall into the first category. Within
this group, two main approaches can be distinguished. In the first approach the
constitutive relationship is expressed in a form ε = f (σ ) (Ramberg and Osgood, 1943,
Ma et al., 1976) and in the second one it is in the form σ = f (ε ) (Menegotto and Pinto,
1973, Filippou et al., 1983, Chang and Mander, 1994). Since in the finite element
method the strains are usually derived first from the strain-displacement relationship,
the second approach seems to be more advantageous. The most popular model in this
context is the model originally proposed by Mentegotto and Pinto (1973) in which the
non-linear relationship is considered as shown in Figure 2.11(a) and expressed by

⎡ (1 − b) ⎤
σ = σ r + ⎢b ε + ⎥ (σ o + σ r ) (2.3)
⎢⎣ (1 + ε R )1 R ⎥⎦

ε − εr
ε= (2.4)
εo − εr

where σ and ε are the stress and strain of reinforcing steel respectively; σ o and ε o are
the coordinates of the point where the asymptotes of the branch under consideration
intersect; σ r and ε r are the stress and strain at the point where the last strain reversal
with stress of equal sign took place; Es and Ew are the initial and the strain hardening

modulus of elasticity respectively as well as the slopes of the asymptotes; b is the strain
hardening ratio equal to Ew Es and R is a transition parameter to account for the
Bauschinger effect that is given by

a1 ξ
R = R0 + (2.5)
a2 + ξ

where ξ is the normalised strain history parameter that is updated following a strain
reversal as shown in Figure 2.11(b); R0 , a1 and a2 are material parameters determined

2-16
Figure 2.11 –Hysteretic model of Menegotto and Pinto (1973): (a) parameters (b)
definition of parameter R (ξ ) .

from experiment. For the non-prestressed steel, Menegotto and Pinto (1973)
recommended R0 = 20 , a1 = 18.5 and a2 = 0.15 .

2.2.4 Shear Transfer

When a shear force is transmitted through an uncracked RC member, the concrete


response is considered to be linear elastic. As the principal tensile stress at some
location develops and reaches the cracking strength of concrete, a crack forms normal to
the direction of the principal tensile stress. After cracking, two major mechanisms
namely aggregate interlock and dowel action contribute to the shear transfer. Aggregate
interlock is essentially a material property depending on aggregate type and shape and
the strength of cement paste while the dowel action is a structural property relating to
detailing of the reinforcement, the geometry of section and the loads or constraints
(CEB, 1996). The contribution of dowel action is less pronounced in comparison to that

2-17
of the aggregate interlock (Walraven, 1980) and to analyse the RC structures with low
amount of reinforcement such as shear walls, it may be ignored (Elmorsi et al. 1998).

Aggregate Interlock

Cracked concrete transmits a significant amount of shear force attributed to sliding of


the rough crack surfaces along each other and interlocking of aggregate particles. This
phenomenon is termed aggregate interlock (Fenwick and Paulay, 1968) and is
accompanied with a restraining normal force along the crack due to the reinforcing steel
as well as the boundary restraints. Additionally, relative tangential and normal
displacements of the two crack surfaces named as the crack slip and the crack width,
respectively, occur over the crack interface. Figure 2.12 shows the shear stress τ cr , the

normal stress σ cr , the crack slip δ cr and the crack width wcr which occurred over a
rough crack. As the crack slip is developed at the crack interface, there is a tendency for
widening of the crack known as crack dilatancy resulting in increasing axial stress in the
reinforcing bars. The relationship of a planar crack may be written in the form (Bazant
and Gambarova, 1980):

⎧dσ cr ⎫ ⎡ D11 D12 ⎤ ⎧dwcr ⎫


⎨ ⎬=⎢ ⎨ ⎬ (2.6)
⎩ dτ cr ⎭ ⎣ D21 D22 ⎥⎦ ⎩ dδ cr ⎭

where D11 , D12 , D21 and D22 are crack stiffness coefficients which are dependant on
τ cr , σ cr , δ cr , wcr and possibly other parameters. The crack stiffness matrix is
asymmetric and non-positive definite, so that crack response tends to be unstable.
However, the restraint provided by the reinforcement usually stabilises the response.

2-18
σcr
τcr

wcr
δcr

Figure 2.12 – Concrete rough crack.

Early tests conducted to investigate the internal shear force distribution in RC beams did
not take into consideration the shear forces transmitted by the interface shear transfer
mechanism. Taylor (1959) was the first to undertake tests to study this mechanism and
its effect on diagonal cracking. The results of these tests, as well as those of other
experimental investigations such as Fenwick (1966), Fenwick and Paulay (1968) and
Sharma (1969), revealed the significant contribution of the interface shear transfer
mechanism to resist shear forces with between 33 to 50 percent of the total shear
applied at a section carried by interface shear. It was also found that the shear stiffness
of beams increased with decreasing initial crack width and increasing concrete strength.
However, due to limitations of tests undertaken on beams a systematic investigation of
the effects of major parameters such as initial crack width and aggregate size, the basic
load-displacement relationships could not be determined. During the 1970s a series of
direct shear tests were carried out to study the effect of initial crack width, concrete
strength, aggregate size, normal restraining stiffness, shear stress intensity and cyclic
shear stresses (Loeber, 1970, Taylor, 1970, Houde and Mirza, 1972, White and Holley,
1972, Paulay and Loeber, 1974, Laible et al., 1977). Figure 2.13 shows a typical
specimen used in direct shear tests in which a shear force, V, and a normal stress, σ n ,
are applied. The tests were conducted under different conditions including constant
crack width, constant and variable confinement stiffness, constant confinement stress
and constant crack dilatancy. The main conclusions drawn from the tests were:

2-19
V
preformed crack

σn σn
shear plane

Figure 2.13 – Typical direct shear test specimen.

• the aggregate shape and size had little influence on the shear stress-shear
displacement relationships.

• the shear displacement across the crack was largely affected by the crack width.

• the response of specimens observed prior to the termination of test consisted of


the three phases shown in Figure 2.14. In the first phase an initial free crack slip
occurred before the aggregates projecting across the crack came into contact with
each other or the cement matrix. During the second phase, the interlocking of
aggregates was established accompanied with a low shear stiffness. Stage 3,
complete interlocking with a linear elastic response and increased shear stiffness.

• the shear stiffness of specimens increased with increasing normal pressure and
decreasing the crack width.

• when the crack width as well as the shear stress over the crack surfaces increased
the shear stiffness of crack gradually decreased.

2-20
τcr
termination of test

phase 1 phase 3

phase 2

δ cr

Figure 2.14 – Typical shear stress-displacement response of concrete crack.

• the magnitude of crack width dominated the response of specimens tested under
cyclic loading such that the shear displacement accumulated after each load cycle
was proportionally related to the crack width.

• the cyclic shear forces resulted in the degradation of the crack surfaces so that the
initial free slip increased with load cycling.

• the shear stiffness of a crack observed during the first load cycle was about 1/3
to 1/2 of shear stiffness of following cycles. Moreover, the following cycles up to
about 17th and similarly between the 18th and 32nd, load repetition were
practically coincident (Figure 2.15) (Paulay and Loeber, 1974).

• upon unloading, residual displacements ranged from 70 to 90 percent of


maximum shear displacement corresponding to the onset point of unloading were
observed.

• the interface shear transfer was fully activated at a shear stress and shear
displacement of approximately 10% and 70% of the maximum shear stress and
shear displacement, respectively, during the cycle considered.

The studies discussed above focused on the shear stiffness rather than shear strength of
cracked concrete sections. In studies by Mattock and his co-workers (Hofbeck et al.,

2-21
1969, Mattock and Hawkins, 1972, Mattock et al., 1976) a number of experimental tests
were undertaken to investigate the shear strength of cracked reinforced concrete
expressed in terms of the ultimate shear stress. These works were used as a base for the
design relationships of shear friction described in the ACI 318 building code (ACI
Committee 318, 1971, 1977). The specimens were tested either as uncracked or
precracked along the shear plane and only the failure load and amount of transverse
reinforcement were reported. The test results showed that the strength of the uncracked
specimens were higher than those of the initially precracked specimens, although the
shear stiffness and upper limit on shear transfer for the both groups were similar
(Figure 2.16). Considering that the shear transfer can be described in terms of friction, a
shear-frictional equation was proposed by Mattock and Hawkins (1972):

6
Number of cycle 1 2 17

to continue
Shear Stress (MPa)

0
0 1 2 3 4 5 6 7 8 9
Shear Displacement (mm)

Figure 2.15 – Cyclic response of shear stress due to aggregate interlock ((Paulay and
Loeber, 1974).

2-22
12

10

Initially uncracked
8
τcr (MPa)
6

Initially cracked
4

0
0 2 4 6 8 10 12

ρvfy (MPa)

Figure 2.16 – Crack shear stress τ cr versus crossing reinforcement ratio ρ v f y


(Hofbeck et al., 1969, Mattock and Hawkins, 1972).

τ cr max = ρ v f y μ (2.7)

where ρ v is the crossing reinforcement ratio, f y is the yield stress of reinforcement and

μ is the coefficient of friction and is dependant on the construction procedure. This


study dealt with the ultimate shear friction capacity of cracks and did not investigate the
relationship between crack shear stress and crack displacements. Later on, Mattock
(2001) re-evaluated the design equations based on the large number of tests carried out
on the concrete with strengths ranging from 17 to 100 MPa and obtained a set of
modified design equations.

Although by the early 1970s some models had been proposed for certain aspects of the
phenomenon of aggregate interlock (Hofbeck et al., 1969, Houde and Mirza, 1972,
Paulay and Loeber, 1974), comprehensive analytical models of the problem were

2-23
lacking. Shear transfer was introduced into the finite element models of RC structures,
by Suidan and Schnobrich (1973) using the relationship:

Gcr = β s Gc (2.8)

where Gcr and Gc are the shear modulus of cracked and uncracked concrete,

respectively, and β s is a shear retention factor ( 0 ≤ β s ≤ 1 ). In the early studies the β s


was generally considered constant but later models often adopted a decreasing functions
for β s to consider the effect of increasing crack width with increasing applied load (Dei

Poli et al., 1990). Further discussion regarding β s is presented in the section 2.3 of this
chapter. The use of an empirical shear retention factor, however, does not adequately
and comprehensively represent the nature of aggregate interlock and the dependency of
shear stiffness on the crack properties.

Fardis and Buyukozturk (1979) considered the rough and irregular shape of crack as a
superposition of low frequency sinusoidal components identified by the location and
size of aggregates and matrix upon high frequency ones comprised of small asperities
and protruding particles (Figure 2.17). These components were termed general and local
roughness, respectively. The general roughness of a crack interface was assumed to
retain its shape during relative movement, whereas the local roughness was considered
to be ground and smoothed. Ignoring the effects of local roughness and introducing the
shape of the general roughness as functions of the coordinate x (in the tangential
direction of the crack interface), the forces developed due to material internal friction at
the contact points were determined. Since some of the independent variables such as the
detailed shape of the crack are a priory unknown, the model was not a predictive tool
but functional model. Applying a multiple regression analysis for the available
experimental data and determining these parameters, a model between the response and
main variables such as the crack width and the stress resultants at the crack location was
developed. The model could qualitively predict the response of concrete crack (Fardis
and Buyukozturk, 1979). The limitation of model lies in its attempt to describe a
stochastic crack surface profile with a deterministic function.

2-24
y

Figure 2.17 – General and local roughness of crack interface (Fardis and Buyukozturk,
1979).

During the 1980s, the experimental data collected in the 1970s and the early 1980s
(Walraven, 1980) made it possible to develop much more realistic and reliable models.
These models may be classified into two categories: empirical models and physical
models (Feenstra et al., 1991a). The models of first category include empirical
formulations derived from the experimental results, while the models in the latter are
based on rational assumptions of the shape of the crack surface. Among the empirical
models, the most reliable and comprehensive are those that originate from the rough
crack theory (Bazant and Gambarova, 1980, Divakar et al., 1987, Yoshikawa et al.,
1989) and the model expressed by Walraven and Reinhardt (1981), whereas those of the
physical models are the two-phase model (Walraven, 1980, Walraven, 1981) and the
contact density model (Li and Maekawa, 1987, Li et al., 1989). Although the
constitutive relationship is assumed to be incrementally linear, due to the complexity of
the problem the less general but simpler constitutive laws based on the total deformation
approach have been mostly proposed. The major drawback of these models is the
path-independency while the inelastic response of aggregate interlock is normally path-
dependent.

Bazant and Gambarova (1980) introduced the rough crack model considering only the
local roughness of crack interface. The number of contact points along the crack
interface was assumed to be infinite so that the stress-displacement relationships for a
crack could be expressed as continuous. The main assumptions were:

2-25
• the interface stresses attributed to the wedging effect are assumed to be primarily
dependant on the displacement ratio r = δ cr wcr .

• for large values of the displacement ratio, r , the shear stress τ cr must exhibit an
asymptote due to microcracking and crushing in the mortar close to the
aggregates.

• for large values of crack width wcr (e.g. wcr > Dmax / 2 where Dmax is the
maximum aggregate size), the stresses vanish because the contact between crack
surfaces is lost.

The shear stress τ cr and the normal stress σ cr are expressed as functions of main
parameters as follows:

τ cr = f t (r , wcr / Dmax , f c' ) (2.9)

σ cr = f n ( wcr ,τ cr ) (2.10)

By optimising the fits of Paulay and Loeber’s experimental data (1974) the following
relationships were identified:

3
a3 + a4 r
τ cr = τ u r (units of mm and MPa) (2.11)
1 + a4 r 4

σ cr = −
a1
(a2 τ cr )p (units of mm and MPa) (2.12)
wcr

where a1 = 0.000534 ; a2 = 145.0 ; a3 = 2.45 / τ 0 ; a4 = 2.44 (1 − 4 / τ 0 ) ; τ 0 = 0.245 f c' ;


2
τ u = 0.001τ 0 / [ 0.001 + ( wcr / Dmax ) 2 ] ; p = 1.30 [1 − 0.231 /(1 + 0.185 wcr + 5.63 wcr )] and

f c' is the compressive cylindrical strength of concrete in MPa. The response diagram of
model is presented in Figure 2.18.

2-26
10

8 wcr = 0.1 mm
τcr (MPa) 0.5 mm
6 1.0 mm
4

-2 1.0 mm
σcr (MPa)

-4 0.5 mm

-6 0.1 mm
-8

-10
0 0.5 1 1.5 2

δcr (mm)

Figure 2.18 – Typical response diagram for f c' = 30 MPa (Bazant and Gambarova,
1980).

Based on the concepts of the rough crack model, the functional relationships among the
four major parameters δ cr , wcr , τ cr and σ cr have been expressed in different ways.
Divakar et al. (1987) considered the functional relationships:

τ cr = g t (δ cr , σ cr ) (2.13)

wcr = g n (δ cr , σ cr ) (2. 14)

while Yoshikawa et al. (1989) focused on the crack slip and confinement stress as
follows:

δ cr = ht (τ cr , wcr ) (2.15)

2-27
σ cr = hn (τ cr , wcr ) (2.16)

Similarly to the original rough crack model by Bazant and Gambarova (1980), in these
studies the empirical relationships drawn on the experimental data were proposed to
determine the crack response.

The rough crack model may be extended to cyclic loads. The CEB (1996) reported a
extended cyclic model by Gambarova (1980) drawn on the rough crack model in which
the proposed loading and reloading equations were similar to Eqs. 2.14 and 2.15, while
different unloading formulations were expressed.

Based on extensive and comprehensive experimental tests on the aggregate interlock


mechanism in normal weight concrete (Walraven et al., 1979) and data-fit analysis,
Walraven and Reinhardt (1981) introduced the following empirical linear formulations
(in units of mm and MPa) that fit their experimental data with the greatest accuracy:

f cc −0.80 −0.707
τ cr = − + [1.80 wcr + (0.234 wcr − 0.20) f cc ]δ cr (2.17)
30

f cc −0.63 −0.552
σ cr = − + [1.35 wcr + (0.191 wcr − 0.15) f cc ]δ cr (2.18)
20

in which δ cr ≥ 0 ; τ cr ≥ 0 ; σ cr ≤ 0 and f cc is the compressive cube strength of


concrete in MPa. Figure 2.19 illustrates the model.

The original model lacked an upper bound for the shear stress. Later, using the same
experimental data Vecchio and Collins (1986) proposed the following equation as the
maximum shear stress that the crack could carry:

f c'
τ cr max = (units of mm and MPa) (2.19)
24 wcr
0.31 +
Dmax + 16

2-28
10
wcr = 0.1 mm
0.5 mm
8
τcr (MPa)
6

4 1.0 mm
2

-2
1.0 mm
-4
σcr (MPa)

-6

-8
0.1 mm 0.5 mm
-10
0 0.5 1 1.5 2

δcr (mm)

Figure 2.19 - Response diagram for f cc = 37.5 MPa (Walraven and Reinhardt, 1981).

Although this model can be simply employed in a finite element model, the initial free
slip included in the model leads to some instability into the computational algorithm
(Lai, 2001). Moreover, the model has been validated for the crack width up to 1 mm and
the model will lose its accuracy when the crack width is too large so that the ratio
τ cr / δ cr (i.e. the crack shear stiffness) as well as σ cr / δ cr degrades dramatically and
becomes zero after a crack width limit regardless of the maximum aggregate size. For
example for a concrete with the compressive cube strength of 37.5 MPa and maximum
aggregate size of 19 mm, the shear stiffness becomes zero at a crack width of about
1.3 mm.

2-29
The two-phase model is a rational and comprehensive analytical model introduced by
Walraven (1980) that was developed to model the aggregate interlock in the cracks
without reinforcing steel bars. The main assumptions of the model were as follows:

• concrete is assumed as a two-phase material consisting of a very stiff spherical


inclusions embedded in a soft matrix.

• Fuller’s curve is adopted for the grading of the aggregate along a crack plane.

• the inclusions and the matrix are assumed to be in partial contact attributed to
interface displacements. The active contact areas are related to interface
displacements using the geometrical relationships and the statistical distribution
of aggregates.

• while the compressive contact strength of matrix is considered as a function of


concrete strength, the shear contact strength is assumed as a linear function of the
compressive contact strength using a friction coefficient.

Applying the equilibrium conditions in the crack interface, the stress-displacement


relationships were given as:

τ cr = σ pu ( An + μ At ) (2.20)

σ cr = −σ pu ( At − μ An ) (2.21)

where An and At are the average contact areas for a unit crack surface in the normal

and tangential directions of the crack interface, respectively; σ pu is the maximum

compressive strength of matrix and μ is the friction coefficient between inclusions and
matrix. The relationships proposed to determine An and At were sophisticated

functions of δ cr , wcr , Dmax and relative aggregate volume pk defined as the ratio of
total volume of the aggregates over the total volume of the concrete. While a quite good
agreement between the predicted values by the model and the experimental data was

2-30
observed by Walraven (1980), its application in the finite element models has been
limited due to its complex relationships resulting in some difficulties.

The contact density model of Li and Maekawa (1987) and Li et al. (1989) is based on
two proposals and three assumptions. The proposals describe the geometry of a crack
surface and the direction of contact stress as follows:

• a crack surface contains a set of contact areas (i.e. contact units) with different
inclinations θ varied from − π / 2 to π / 2 . A probabilistic contact density
function Ω (θ ) can be used to describe the distribution of the crack surface
orientations.

• the direction of each contact stress is proposed to be fixed and normal to the
initial contact direction denoted by θ .

The main assumptions are:

• The contact density function Ω(θ ) is independent of the size and the grading of
aggregates and given by

Ω(θ ) = 0.5 cos θ (2.22)

• a simple elastic-perfectly plastic model is assumed for the normal contact stress
(σ con ) with regard to the matrix and interface deformations.

• the effective ratio of contact area K ( wcr ) representing the effect of wcr on the
loss of contact area, is given by

0.5 Dmax
K (wcr ) = 1 − exp (1 − ) (2.23)
wcr

2-31
Applying the equilibrium conditions, the shear stress τ cr and the normal stress σ cr
transferred across the crack are given by

π/ 2
τ cr = ∫ σ con . K ( wcr ) . At . Ω(θ ) . sin θ dθ (2.24)
− π/ 2

π/ 2
σ cr = ∫ σ con . K ( wcr ) . At . Ω(θ ) . cos θ dθ (2.25)
− π/ 2

where At is the total area corresponding to the nominal unit area of crack surface.

For reinforced concrete with an ordinary reinforcement ratio, where the assumption of
smeared crack is valid, a simplified path-independent stress transfer model was
formulated (Li et al., 1989, Maekawa et al., 2003):

ψ2
τ cr = 3.83 3 f c' (2.26)
1 +ψ 2

⎡π ψ2 ⎤
σ cr = 3.83 3 f c' ⎢ − cot −1ψ − ⎥ (2.27)
⎢⎣ 2 1 +ψ 2 ⎥⎦

where ψ = δ cr / wcr and f c' is in MPa. The response diagram of the simplified model is
demonstrated in Figure 2.20. The model is simple and rational and can be easily
incorporated into the finite element modelling to analyse the RC structures subjected to
either monotonic or cyclic loading (Okamura and Maekawa, 1991, Maekawa et al.,
2003). Nevertheless, the model will lose its accuracy if the crack width is too large
(typically greater than 1 mm) since the maximum shear transfer in this model is

independent of wcr and Dmax and only dependent on f c' (Maekawa et al., 2003).

2-32
10
wcr = 0.1 mm 0.5 mm
8 1.0 mm
τcr (MPa)
6

-2
σcr (MPa)

-4

-6
1.0 mm
-8
0.1 mm 0.5 mm
-10
0 0.5 1 1.5 2

δcr (mm)

Figure 2.20 - Response diagram for f c' = 30 MPa (Li et al., 1989).

Feenstra et al. (1991a, 1991b) carried out a comprehensive comparative study to


evaluate the performance of different aggregate interlock models including the rough
crack model of Bazant and Gambarova (1980), the two-phase model of Walraven
(1980), the empirical model of Walraven and Reinhardt (1981) and the simplified
contact density model of Li et al. (1989). In general, all models showed a reasonable
agreement with experimental data, though a large scatter was observed specially at
smaller crack widths. They concluded that the contact density model, compared with the
other models, seemed to have the edge from a numerical point of view.

The aggregate interlock capacity of cracks in high-strength concrete (HSC) is reduced


compared with normal strength concrete. For HSC, since the strength of the matrix is
more than that of the aggregate particles, the cracks tend to cross the individual pieces
of aggregates rather than going around them. Based on an experimental investigation
undertaken on RC beams with compressive strength from 21 to 92 MPa,

2-33
Mphonde (1988) concluded that contribution of aggregate interlock decreased
considerably as the compressive strength of concrete increased.

Walraven (1995) carried out tests on pre-cracked push-off specimens made of concrete
with a compressive cube strength of 115 MPa. The tests included both plane and
reinforced cracks. It was found that the fracture of the aggregate particles upon crack
formation in high-strength concrete resulted in smoother crack surfaces and as a result a
significant reduction in the shear capacity of cracks due to aggregate interlock was
observed. The reduction for plain cracks was about 65 percent of the value which would
be achieved without particle fracture, while that of reinforced cracks was about 25 to 45
percent. Walraven concluded that because of better bond between reinforcing bars and
the high strength concrete, the yield stress was reached at a smaller crack width
compared with normal strength concrete and therefore the number of contact areas was
smaller.

Based on the contact density model of Li and Maekawa (1987), Ali and White (1999)
developed a method for use in design. The model was used to determine the ultimate
shear capacity and validated for a wide range of concrete strengths (20 to 100 MPa).
Nevertheless, the model did not provide the full response of the crack interface
subjected to the crack displacements.

Angelakos et al. (2001) conducted tests on large RC beams with compressive strengths
ranging from 20 to 100 MPa to study their shear strengths. The maximum shear stress
of a crack was calculated using Eq. 2.22 in which the smoother interface of cracks in
HSC was addressed by reducing the effective aggregate size a from the nominal
diameter to zero as the concrete strength increases from 60 to 70 MPa. For concrete
strengths of 70 MPa or higher, the a was taken as zero.

2-34
2.3 Finite Element Analysis

Since the first use of finite element method for modelling of simply supported
reinforced concrete beams (Ngo and Scordelis, 1967), the method has been significantly
developed and become a powerful and general tool to simulate and study the behaviour
of reinforced concrete structures as reported comprehensively by ASCE (1982, 1993)
and CEB (1996). Although the progressing in speed and accuracy of analyses in recent
years has led to the wide use of the nonlinear finite element analysis softwares in many
types of practical applications, the use of them in reinforced concrete remains complex.
Besides the significance of the type of finite element and the size of mesh used to
idealise the structure, the main challenge in using this method to model RC structures
arises from the composite nature of the material. The accuracy and reliability of finite
element analysis is largely governed by the abilities of the underlying constitutive
relationships to capture different types of nonlinear behaviour such as the nonlinear
stress-strain relationships of concrete and reinforcing steel, cracking, crushing, tension
stiffening, compression softening, aggregate interlock, dowel action and bond slip.
Moreover cyclic loading introduces further complexities such as stiffness and strength
degradation in concrete, the Bauschinger effect in reinforcing steel and bond
deterioration between concrete and reinforcement. As well, incompatibility of models
and approaches is another important problem, so that using some models from one
analytical approach to another or combining them with other models may yields errors
in the finite element analysis (Vecchio and Palermo, 2001). As a result, to simulate the
response of reinforced concrete members and specially shear walls under general
loading, the selection of proper and compatible material models and numerical
procedures is essential. In the following, a review of the major approaches and the
previous works conducted on the nonlinear finite element analysis of reinforced
concrete with emphasis on the finite element analysis of RC shear walls and panels is
presented.

2-35
2.3.1 Overview

The dominate method for the finite element analysis of reinforced concrete consists of
developing separate models for the concrete and for the reinforcing steel and then
combining those models to make the constitutive matrices at the element level as well
as the structure level. The main approaches for modelling of concrete and reinforcing
steel in reinforced concrete structures are discussed herein.

Concrete Modelling

Reinforced concrete as a composite material exhibits a highly nonlinear behaviour. This


nonlinearity arises primarily from the nonlinear behaviour of the constituent materials
of steel and concrete, which dominate the pre- and post-cracking responses. A major
reason for nonlinearity in reinforced concrete is the tensile cracking which causes a
considerable redistribution of stress within the intact concrete as well as the stress
transfer from concrete to the reinforcement. The success of a nonlinear finite element
analysis for RC structures essentially depends on realistic crack modelling, along with
appropriate material representations of reinforcing steel and concrete. In general, two
major approaches have been used to model the cracking in concrete structures named as
the discrete crack model and the smeared crack model. The former approach models a
crack as a geometrical discontinuity in concrete, while the latter treats a cracked
concrete as a continuum.

In the early discrete crack model that was first introduced by Ngo and Scordelis (1967),
cracking was modelled as a separation of nodes along element edges. Post-cracking
effects such as tension stiffening and bond-slip can be taken into account using linkage
elements or interface elements between the crack surfaces. Although this approach was
used in the other early studies on the finite element modelling of RC structures (Nilson,
1968, Franklin, 1970, Mufti et al., 1970), two major drawbacks involved in using the
model in the finite element modelling greatly restrict the ease and speed of analysis.
First, the crack propagation is limited to follow a predefined path along the element
boundaries so that to improve the analytical results, the topology of the finite element
mesh is required to be changed continuously throughout the analysis. Second, because

2-36
of considering the additional degree of freedoms for the separated nodes along the new
crack faces, the time and cost of computation increase and the efficiency decreases
(although with today’s high speed computers this time and cost penalty is less restrictive
than it once was). To overcome these issues some modifications were proposed such as
remeshing techniques with respect to the potential direction of crack propagation at each
crack increment (Cervenka and Gerstle, 1971, Ingraffea and Saouma, 1984, Bocca et al.,
1991) and the techniques in which the crack is permitted to extend through the finite
elements (Ortiz et al., 1987, Belytschko et al., 1988). While the mentioned drawbacks
have made the discrete crack model less popular than other techniques (such as, for
example, smeared cracks), for those problems that involve a few dominate cracks or
where the local material behaviour is of interest, it is most useful and likely to be the
main choice. Nevertheless, the behaviour of many reinforced concrete structures
contained closely spaced reinforcement such as shear panels and walls are dominated by
distributed parallel cracks. In such cases, the smeared crack model is often considered to
be more effective and beneficial than the discrete crack model (ASCE Committee 447,
1982, 1993).

In the smeared crack model, introduced by Rashid (1968), once cracking occurs at a
integration point in an finite element, the cracks are assumed to be parallel and finely
distributed (or smeared) over the entire region that point represents. Thus the cracked
concrete can be considered as a continuum and the average stress-strain relationship can
be expressed in a continuous manner so that upon cracking the initial isotropic stress-
strain relationship can be simply replaced by an orthotropic stress-strain relationship.
Consequently, cracking is treated as a reduction of average material stiffness over the
area in directions of the major orthotropic axes and the constitutive matrix is
correspondingly modified to reflect this. In comparison to the discrete crack model, the
smeared crack model neither changes the topology of original finite element mesh nor
imposes any constraints on crack propagation directions. Despite this, classical smeared
crack approaches suffer from a deficiency when dealing with localised cracking
(Bazant, 1983), however, the advantages of the smeared crack approach have made it
more widely adopted than the discrete crack approach by researchers to model cracking
in the large scale reinforced concrete structures. Throughout the literature, three major

2-37
variants of smeared crack approach have been used; the fixed crack model, rotating
crack model and multi-directional fixed crack model.

In the fixed crack model that was firstly used by Rashid (1968), the crack takes place
normal to the direction of maximum principal stress once it reaches the concrete tensile
stress and thereafter the crack orientation is taken as fixed throughout the loading
process. The cracked concrete is assumed as an orthotropic material with the orthotropic
axes n and t , normal and parallel to the crack direction, respectively. Considering a
two dimensional cracked concrete membrane, the orthotropic constitutive relationship
may be given by (ASCE Committee 447, 1993)

⎧σ nn ⎫ ⎧ε nn ⎫
⎪ ⎪ sec ⎪ ⎪
⎨ σ tt ⎬ = D nt ⎨ ε tt ⎬ (2.28)
⎪τ ⎪ ⎪γ ⎪
⎩ nt ⎭ ⎩ nt ⎭

where Dsec
nt is the secant constitutive matrix; σ nn , σ tt and τ nt are the average stresses

and ε nn , ε tt and γ nt are the average strains in the n - t coordinate system, respectively.
In the early application of the fixed crack model (Rashid, 1968, Cervenka and Gerstle,

1971) Dsec
nt is expressed as

⎡0 0 0 ⎤
Dsec
nt = ⎢⎢0 E 0⎥⎥ (2.29)
⎢⎣0 0 0⎥⎦

In other words, after cracking the stiffness normal to the crack, the shear stiffness and
the Poisson’s ratio for cracked concrete are taken to be zero. Later, because of
numerical problems arising from the immediate drop of stiffness, a reduced shear
stiffness, β s G was incorporated into the model as follows (Suidan and Schnobrich,
1973) :

⎡μn E 0 0 ⎤
sec
Dnt = ⎢⎢ 0 E 0 ⎥⎥ (2.30)
⎢⎣ 0 0 β s G ⎥⎦

2-38
where G is the shear stiffness of intact concrete and β s is a shear retention factor and

ranged between 0 and 1. While foe many problems the exact amount of β s is not

crucial to the solution, a value equals 0.2 is a commonly adopted value. The use of β s
may be also thought as a way to account for aggregate interlock and dowel effects.
Additionally, to address the strain softening effect in tensile concrete a “normal
reduction factor” μ n has been inserted in the constitutive matrix:

⎡μn E 0 0 ⎤
sec
Dnt = ⎢⎢ 0 E 0 ⎥⎥ (2.31)
⎢⎣ 0 0 β s G ⎥⎦

where μ n is considered as a function of the strain normal to the crack.

The fixed crack model has been used to analyse the reinforced concrete structures under
both monotonic and cyclic loading. It usually predicts a stiffer response compared with
experimental results especially when a structure is anisotropically reinforced (Crisfield
and Wills, 1989) because, while the fixed crack model does not allow the crack
directions to rotate, in real structures the direction of principal stresses is continuously
changed as the loading is increased and consequently a shear stress on the crack
surfaces is developed. This matter led to developing the rotating crack model in which
the cracks can rotate during loading process and the multi-direction fixed crack model
in which new cracks are allowed to develop at the different orientations.

The rotating crack model was originally developed by Cope et al. (1980) to take into
account crack reorientation. While the basic relationships of rotating crack model are
essentially similar to the fixed crack model, the main distinction between them is that in
the rotating crack model the axes of orthotropy are always aligned with the direction of
major principal stresses as the load increases. Furthermore to make the rotating crack
model simpler, the assumption of co-axiality of the principal stresses and the principal
strains is usually incorporated into the model (Vecchio, 1989). Consequently,
considering the cracked concrete as an orthotropic material, the stress-strain constitutive
relationship may be written based on the principal 1-2 axes as

2-39
⎧σ 11 ⎫ ⎧ε 11 ⎫
⎪ ⎪ sec ⎪ ⎪
⎨σ 22 ⎬ = D12 ⎨ε 22 ⎬ (2.32)
⎪τ ⎪ ⎪γ ⎪
⎩ 12 ⎭ ⎩ 12 ⎭

sec
where D12 is the secant constitutive matrix; σ 11 and σ 22 are the average stresses and
ε 11 and ε 22 are the average strains in the directions of 1- and 2-principal axes
respectively. The co-axiality assumption requires that the shear stress τ 12 and the shear
strain γ 12 are always zero and hence the need to employ the shear reduction factor β
or a shear stress-strain relationship is removed. In general, the rotating crack model has
been criticised for two shortcomings. First, the rotation of crack direction implies that
the concrete damage is not permanent and it depends only on the current stress state and
therefore it is load history-independent (Bazant, 1983, Noguchi, 1985). Second, with
respect to either new crack direction or new axes of orthotropy at the beginning of each
load step, a different stress-strain relationship should be used and hence, a whole family
of stress-strain curves are required whereas in the fixed crack approach only few curves
are used. Nevertheless, the model can be simply implemented into a computer program
and many researchers have successfully used the model for analysing the behaviour of
RC structures subjected to monotonic loading (Gupta and Akbar, 1983, Milford, 1984,
Vecchio, 1989, Hu and Schnobrich, 1990 among others). Stevens et al. (1987) were the
first researchers that developed the model to include the reversed cyclic loading. While
the model was quite successful at the element level, the analytical procedure at the
structure level encountered some numerical problems and their attempt failed. Later
Vecchio (1999) and Palermo and Vecchio (2003) successfully employed the rotating
crack model to analyse the cyclic response of RC shear walls. Crisfield and Wills
(1989) investigated the application of the fixed and rotating crack models to the analysis
of a number of RC panels tested by Vecchio and Collins (1982). The main conclusions
derived from this investigation can be summarised as:

Applying the fixed crack model to the RC members that fail in a ductile manner with
response dominated by yielding in the reinforcement steel, the shear retention factor
does not affect on the final collapse load, although the load-deflection is dependant on
it.

2-40
When the straining pattern changes as the loading is increased, the fixed crack model
may result in overestimated failure loads, where the rotating crack model yields a
failure load that is less than or equal to that given by the fixed crack model.

If the concrete fails in a shear induced mode, the rotating crack model leads to an
overestimation of failure loads and often indicates a failure attributed to yielding of the
reinforcing steel.

The application of a simple, but effective orthogonal fixed crack model is more general
than rotating crack model. Nevertheless in absence of such fixed crack model, the
rotating crack model appears to offer many advantages.

The multi-directional fixed crack model was originally developed by De Borst and
Nauta (1985). The model is a refined version of the fixed crack model in which the
crack orientation is still assumed constant after cracking takes place. However, new
cracks are allowed to develop in different directions if the angle between two
consecutively formed cracks exceeds a threshold angle. Based on this concept, the
rotating crack model may be conceived as a multi-directional fixed crack model with a
zero threshold angle (ASCE Committee 447, 1993). The main concept in this model is
that the total concrete strain increment Δ ε c is decomposed into an intact concrete strain

increment Δ ε ic and a crack strain increment Δ ε cr which can be expressed as

Δ ε c = Δ ε ic + Δ ε cr (2.33)

where Δ ε cr is the summation of the strain contributions of all cracks at one integration
point and is expressed by

Δ ε cr = Δ ε cr 1 + Δ ε cr 2 + ... + Δ ε crn (2.34)

So the cracks can be treated separately and while the effect of new cracks on the total
response of reinforced concrete is considered, the influence of the old cracks is taken
into account as well. In other words the model is path-dependent and sensitive to load
history. The application of the multi-directional fixed crack model is limited compared

2-41
with alternative approaches. This arises from its complex nature as well as the need to a
large number of computations to build the total constitutive matrix particularly under
cyclic conditions (Crisfield and Wills, 1989). The model was successfully used by
Bolander and Wight (1991) for investigating the behaviour of RC shear walls under
monotonic loading and Xu (1991) employed the model to simulate the cyclic response
of RC members. While the analytical results at element level were generally
satisfactory, because of numerical problems the approach was not successful at the
structure level.

Reinforcing Steel Modelling

Three major approaches have been used to model the reinforcing steel in the finite
element analysis of reinforced concrete including the smeared steel model, the discrete
steel model and the embedded steel model (Figure2.21).

In the smeared steel approach, the reinforcing steel is assumed to be distributed


uniformly over a concrete element at a particular orientation and expressed using the
reinforcing steel ratio. Since the reinforcement is considered to be smeared, the
complete compatibility between reinforcing steel and concrete is imposed. The
constitutive matrix of reinforced concrete finite element is considered to be composed
of individual contributions of concrete and steel reinforcement. This model is usually
helpful for analysing reinforced concrete structures consisting of uniformly distributed
reinforcement (eg. shear walls) not having to define separate reinforcing bars (ASCE
Committee 447, 1982, Crisfield and Wills, 1989).

The discrete steel approach was the first reinforcing steel model used in the finite
element analysis of RC structures (Ngo and Scordelis, 1967). In this model the
reinforcing steel is modelled as at common separate one-dimensional truss elements
connected to the concrete element edges at common nodal points. Bond-slip effects
between steel and concrete can be included by inserting the linkage elements that is the
important advantage of this model. The main drawback of this model is that the

2-42
direction and location of steel reinforcement are restricted to the boundaries of the
concrete elements.

In the embedded steel approach, each steel reinforcing bar is considered to be an axial
member included into the concrete element in such a way that the displacements of
embedded member and those of surrounding concrete element are compatible. Similar
to the smeared steel model, the bond between reinforcing steel and concrete is often
assumed to be perfect. The one exception to this is the incorporation of a stepped plastic
bond slip made into the FE formulation by Foster and Marti (2002, , 2003). The main
advantage of embedded reinforcement model is that there is no limitation for
representing the direction and location of the steel reinforcement. Using the principle of
virtual work the element stiffness matrix, K , can be determined as (Chang et al., 1987)

K = Kc + Ks (2.35)

in which

K s = As Es ∫l B
T
TT T B dl (2.36)

where K c and K s are the stiffness matrices for concrete and reinforcement,

respectively; l , As and Es are the length, the cross-sectional area and the elastic
modulus of reinforcement, respectively; B is the strain displacement matrix and T is
the strain transformation matrix for reinforcement with respect to global coordinate
system.

2-43
(a) (b) (c)

Figure 2.21 – Reinforcing steel modelling: (a) smeared steel model (b) discrete steel
model (c) embedded steel model.

2.3.2 Finite Element Analysis of RC Panels and Shear Walls

The accuracy and reliability of the finite element analysis of reinforced concrete is
largely dependant on material constitutive models, the capability to capture the major
nonlinear aspects and the numerical solution procedures. Since the cyclic loading
introduces further complexities compared with monotonic loading, limited finite
element analysis to predict the cyclic response of RC shear walls has been performed.
The progress of finite element models describing the cyclic behaviour of RC shear walls
and panels may be reviewed in two periods. During the decades 1970s and 1980s,
although the numerical problems associated with employing inappropriate constitutive
models limited the application of finite element analyses especially at structure level,
the investigations led to establishing fundamental concepts, recognising and quantifying
the influence of different aspects of concrete and reinforcing steel on cyclic response
and developing reliable and stable numerical procedures. In the second period, from
1990 to present, some refined finite element models have been quite successfully
developed to predict the cyclic response of RC shear walls and panels.

2-44
The Period from 1970 to 1989

Cervenka and Gerstle (1971, 1972) carried out one of the earliest studies to analyse the
behaviour of RC shear panels under monotonic and cyclic loading using the finite
element method. They used the smeared-fixed crack approach with only one open crack
allowed at a point associated with smeared reinforcing steel. The cracked concrete was
modelled as an orthotropic material without stiffness perpendicular to crack and shear
stiffness tangential to crack which could carry the stresses only parallel to crack
directions. The reinforcing steel was assumed as an elastic-perfectly plastic material
while the response of concrete was modelled as being elastic-perfectly plastic and
elastic-brittle in compression and tension, respectively. To extend the model toward
including cyclic conditions, linear unloading and reloading relationships with the slope
equal to initial elastic modulus were applied. While the resulting model could
reasonably predict the response of two shear panels tested by Cervenka and
Gerstle (1972) under monotonic loading, it failed to simulate that of a shear panel under
reversed cyclic loading. This failure is attributed to not including important post-
cracking aspects such as tension stiffening and aggregate interlock effects into the
model and, more importantly, the exclusion of major effects exhibited under cyclic
conditions such as the energy dissipation in the concrete and steel during unloading and
reloading phases and the degradation of concrete stiffness and strength. With regard to
the shear panels tested under monotonic loading, the experimental results were
governed by yielding of the reinforcing steel and, therefore, the concrete response was
not as critical.

Darwin and Pecknold (1974, 1976) were the first to develop a nonlinear finite element
model for the analysis of reinforced concrete membranes that included major cyclic
aspects. Adopting the smeared-fixed crack approach, the cracked concrete was modelled
as an orthotropic material with two orthogonal open cracks allowed at a point. The
model included a shear retention factor to represent the aggregate interlock and dowel
action but tension stiffening and bond deterioration between concrete and steel
reinforcement were ignored. The reinforcing steel was taken as smeared with an
elastic-perfectly plastic strain hardening material model, including the Bauschinger
effect. To incorporate effects due to the biaxial stress states, Darwin and Pecknold

2-45
introduced the concept “equivalent uniaxial strain” which was later adopted by many
researchers. A more realistic stress-strain relationship for concrete proposed by Sanez
(1964) was used and some important aspects observed in cyclic response of concrete,
including the degradation of concrete stiffness and strength and the energy dissipation at
each hysteretic cycle, were included in the model. As the nonlinear solution procedure
an incremental-iterative tangent stiffness approach was adopted. The model was
corroborated against the shear panel W4 tested by Cervenka and Gerstle (1972) for two
and one half cycles. While the analytical results were in a reasonable agreement with
the experimental results for the first cycle, they began to deviate during following cycle.
Palermo (2001) attributed the deviation of the model from the test data to the numerical
procedure used in the finite element analysis rather than the constitutive models.

Developing an empirical bond-slip model, Shipman and Gerstle (1979) studied the
effect of bond deterioration on the cyclic response of the panel W4 (Cervenka and
Gerstle, 1972). Comparing the numerical results with those by Darwin and
Pecknold (1974), it was found that the effect due to the degradation of concrete strength
and stiffness is more significant than that of the bond deterioration in the cyclic
response of the shear panels.

To analyse a slender shear wall tested by Cardenas et al. (1980), Aktan and Hanson
(1980) proposed a model in which the total response of the wall was decomposed into
linear and nonlinear parts. The nonlinear joint element was modelled based on previous
extensive experimental results on shear walls. The shear stiffness was also assumed to
be a function of normal stress acting on the element after closing the crack.
Nevertheless, the experimental moment-curvature curve could not be reproduced well
by analytical results, even though the shear pinching characteristics were fairly
represented. Furthermore, since it is very difficult to develop the material model for the
nonlinear joint element for general structural members, the application of the model is
limited.

Noguchi (1985) reviewed a large number of finite element models of reinforced


concrete members subjected to reversed cyclic loading. He pointed out that the general
drawback of these early models was the inability to accurately represent the unloading

2-46
stiffness and the area of hysteresis loops. In other words the degradation of concrete
stiffness and the energy dissipation were the challenging issues to be simulated. It is
attributed to this fact that almost all studies previous to 1985 emphasised the modelling
rather the material nonlinearity. Studies through the 1980s, however, showed that the
modelling of the post-cracking effects such as tension stiffening and compression
softening are vital to the analyses of reinforced concrete structures (Vecchio and
Collins, 1982, 1986, Stevens et al., 1987).

To take into account the findings concerning the post-cracking response of reinforced
concrete in finite element analysis, Stevens et al. (1987) developed a more
comprehensive constitutive model for the concrete based on the modified compression
field theory of Vecchio and Collins (1986) and employing a smeared-rotating crack
approach for modelling the concrete. Besides considering the effects due to the material
nonlinearity and the biaxial stress state in the concrete, post-cracking effects such as
tension stiffening, compression softening and bond-slip were incorporated into the finite
element model. The model accounts for the cyclic aspects of concrete response
containing the degradation of concrete strength and stiffness as well as the energy
dissipation during each unloading and reloading cycle. Adopting a discrete beam
element for reinforcement, the stress-strain relationship of the reinforcing steel was
modelled to include the reduced yielding caused by cracking in addition to the strain
hardening and the Bauschinger effect. The proposed model was verified against the
experimental results of three panels tested by the researchers under cyclic loading.
When it was used at the structure level, however, numerical problems arose from the
complexity of constitutive relationships and most analyses could not be completed.

The Period from 1990 to Present

Xu (1991) developed a finite element model using the multi-directional fixed crack
approach as a solution to the drawbacks of the fixed crack and rotating crack models.
Considering multiple non-orthogonal cracks at a cracked point, the model could account
for the nonlinear response due to each crack and superimpose the effects from all

2-47
cracks. Reinforcing steel was taken as smeared with a constitutive stress-strain
relationship including yielding, stress hardening and the Bauschinger effect. Using
trilinear unloading and linear reloading relationships, the energy dissipation was
included in the cyclic model. In addition the model incorporated the degradation of
strength and stiffness of concrete due to cyclic damage. Although the model showed
good success at the element level, at the structure level like many previous attempts, the
complexity of model led to the numerical problems during the analyses.

The progress and improvement of techniques used in finite element modelling and
computational procedures as well as the better understanding of main aspects of the
cyclic response of reinforced concrete has led to a number of successful cyclic analyses
of RC shear walls (Okamura and Maekawa, 1991, Sittipunt and Wood, 1993, Elmorsi et
al., 1998, Vecchio, 1999, Palermo and Vecchio, 2003, Kwak and Kim, 2004a, Palermo
and Vecchio, 2004). The main features of these models are summarised in Tables 2.1,
2.2 and 2.3. The following conclusions are drawn:

The equivalent uniaxial strain approach has been successfully employed to simulate the
response of RC shear walls subjected to cyclic loading. The main reason for this success
is attributed to the use of empirical relationships expressing the complex behaviour of
concrete being directly incorporated into the constitutive relationships (Kwan and
Billington, 2001).

As extensive cracking in shear walls is usually expected, the smeared crack model has
been the dominate approach adopted by researchers.

Both fixed crack and rotating crack approaches have been used successfully for
modelling of RC structures under reversed cyclic loading.

Most researchers have modelled the reinforcing steel as smeared throughout the shear
walls.

The finite element type used in the studies ranges from 4-node isoparametric membrane
elements to 12-node quadrilateral isoparametric elements.

2-48
The tangent stiffness formulation is commonly used for analysis of cyclic response of
RC structures as shown in Table 2.1. Although the secant stiffness method has been
criticised by some that it cannot be effectively employed to analyse the response to
general loading conditions (ASCE Committee 447, 1982), the behaviour of shear walls
under reversed cyclic loading has been successfully modelled using this method
(Vecchio, 1999, Palermo and Vecchio, 2003, 2004).

Post-cracking effects including tension stiffening and compression softening have been
included in the finite element models. Assuming perfect bond between reinforcement
and concrete, however, the bond-slip effect has been ignored by most researchers except
by Okamura and Maekawa (1991). It has been argued that since shear walls usually
contain the low amounts of reinforcing steel the effect of bond-slip is not significant.

The rotating crack approach eliminates the need to express a separate shear transfer
model whereas the studies undertaken based on the fixed crack approach have employed
the cyclic shear transfer models. Because typically shear walls have low reinforcement
ratios, Okamura and Maekawa (1991) and Elmorsi et al. (1998) neglected the
contribution of dowel action and assumed that the cracked shear stiffness is due to only
aggregate interlock. In contrast, Sittipunt and Wood (1993) considered the shear
stiffness to be attributed to both effects.

The major cyclic characteristics of concrete modelling applied by researchers include


the unloading and reloading curves, the degradation of concrete strength and stiffness
and the crack-closing transition curve. As it is shown in Table 2.2, the cyclic model of
concrete employed by Vecchio (1999) compared with other models is simpler not
including some cyclic modelling aspects such as strength degradation, nonlinear
unloading curve for concrete indicating energy dissipation nor a crack-closing transition
curve. Although this model yielded the reasonable prediction for the cyclic behaviour of
shear walls dominated by reinforcing steel response, it failed for those that showed
pinched hysteretic loops indicating shear dominant behaviour, where the other models
were successful.

2-49
Table 2.1 – Comparing the main features of finite element modelling used by various
researchers to analyse the cyclic response of RC shear walls

Researchers Approach for Crack Steel Element Iterative


Constitutive Modelling Modelling Type Solution
Relations Technique

Okamura orthotropic smeared-fixed smeared 8-node tangent


and equivalent orthogonal isoparametric stiffness
Maekawa uniaxial strain crack membrane
(1991) element

orthotropic smeared-fixed discrete, 4-node tangent


Sittipunt and
equivalent orthogonal 2-node truss isoparametric stiffness
Wood
uniaxial strain crack element membrane
(1993)
element

Elmorsi et orthotropic smeared-fixed smeared 12-node tangent


al. equivalent orthogonal quadrilateral stiffness
(1998) uniaxial strain crack membrane
element

Vecchio orthotropic smeared- smeared 4-node secant


(1999) equivalent rotating crack isoparametric stiffness
uniaxial strain membrane
element

Palermo and orthotropic smeared- smeared 4-node secant


Vecchio equivalent rotating crack isoparametric stiffness
(2003, , uniaxial strain membrane
2004) element

Kwak and orthotropic smeared- smeared 4-node tangent


Kim equivalent rotating crack isoparametric stiffness
(2004) uniaxial strain membrane
element

2-50
Table 2.2 - Comparing the main features of concrete modelling used by various researchers to analyse the cyclic response of RC shear walls

Cyclic Characteristics Post-Cracking Effects

Researchers Unloading Reloading Strength Stiffness Crack - Tension Compression Shear Bond-
curve curve degradation degradation Closing stiffening softening transfer slip
Transition across cracks
Curve
Okamura and Nonlinear in both Linear in both Not included Included Linear Included Included Only aggregate Joint
compression and compression and interlock element
Maekawa tension tension between
(1991) concrete
elements
Sittipunt and Trilinear in Bilinear in Included only in Included Nonlinear Included Included Aggregate Not
compression, compression, compression interlock and included
Wood linear in tension linear in tension dowel action
(1993)

Elmorsi et al. Bilinear in Linear in both Included only in Included Nonlinear Included Included Only aggregate Not
compression, compression and compression interlock included
(1998) linear in tension tension

Vecchio Elastic linear in Linear in both Not included Included Not included Considered Included Not included Not
compression, compression and both cases with because of using included
(1999) secant linear in tension and without rotating crack
tension tension model
stiffening

Palermo and Nonlinear in both Linear in both Included in both Included Linear Included Included Not included Not
compression and compression and compression and because of using included
Vecchio tension tension tension rotating crack
(2003, , 2004) mode

Kwak and Nonlinear in Linear in both Not included Included Nonlinear Included Included Not included Not
compression, compression and because of using included
Kim linear in tension tension rotating crack
(2004) mode

2-51
Table 2.3 – Comparing the main features of reinforcing steel modelling used by
researchers to analyse the cyclic response of RC shear walls

Researchers Basic Constitutive Reduced Yielding Strain Bauschinger


Stress-Strain Stress due to Hardening Effect
Curve Embedding or Effect
Bond

Okamura and Bilinear Not included Included using a Included using a


Maekawa linear relation proposed model
(1991)

Sittipunt and Linear for elastic and Not included Included using a Included using a
Wood plastic regions and nonlinear relation proposed model
(1993) nonlinear for strain
hardening

Elmorsi et al. Bilinear Included using the model Included using a Included using the
(1998) of Stevens (1987) linear relation modelled of Menegotto
and Pinto (1973)

Vecchio Trilinear Not included Included using a Included using the


(1999) linear relation model of Seckin (1981)

Palermo and Trilinear Not included Included using a Included using the
Vecchio linear relation model of Seckin (1981)
(2003, , 2004)

Kwak and Bilinear Included using the model Included using a Included using the
Kim of Belarbi and Hsu linear relation modelled of Menegotto
(2004) (1994) and Pinto (1973)

2.4 Experimental Studies

Prior to the 1970s, although the shear walls had been employed in a large number of
buildings, the knowledge about inelastic response of reinforced concrete shear walls
subjected to cyclic loading was limited. Many experimental investigations by various
research groups were undertaken during the 1970s and later. Over the past 30 years,
many tests have been carried out on shear wall specimens, simulating single or

2-52
multi-storey shear walls. The majority of tests have been undertaken on normal strength
concrete under monotonic loading. Experimental works on HSC shear walls subjected
to cyclic condition are few. With respect to these limitations and to reach some
conclusions on the behaviour of shear walls, in general, results of tests conducted on
normal strength concrete shear walls are also discussed. However, in order to remain
within the scope of the present research, the review is concentrated on behaviour of
low-rise shear walls.

2.4.1 Barda et al (1977)

Barda et al. (1977) tested eight framed reinforced concrete shear wall specimens
(Figure 2.22) subjected to monotonic and reversed cyclic loading without axial load.
The compressive strength of the concrete ranged from 21 to 29 MPa. The main
parameters of experiment included the height-to-length ratio (0.25 to 1.0) and vertical
reinforcement and horizontal reinforcement ratios (0 to 0.5 percent). The main findings
from the experimental study were reported as follows:

• The longitudinal reinforcement in the edge elements did not have influence on
the shear strength of walls.

• Increasing the height-to-length ratio for a shear wall, where other properties were
maintained as invariable, led to decreasing the shear strengths.

• The failure load of shear walls under reversed cyclic loading was approximately
10% less than that of similarly detailed shear walls under monotonic loading.
The main reasons for this are the existence of cracks in orthogonal directions and
increasing degradation of bond between reinforcement and concrete under cyclic
loading.

• Specimens without edge elements failed suddenly compared with the gradual
failure of framed shear walls due to the frame action provided by the end
elements.

2-53
150 mm
2405 mm 1525 mm
A

Hv

B B
457 mm

3048 mm A 1220 mm

Section A-A
102 mm
1220 mm

610 mm

102 mm

572 mm 1904 mm 572 mm

Section B-B

Figure 2.22 – Typical shear wall specimen tested by Barda et al. (1977).

This research was one of the earliest investigations on shear walls and the design
equations included in the ACI 318 Code are based on its outcomes. Some important
issues such as the effects of axial load and strength of concrete, for example, were not
investigated in this study.

2-54
2.4.2 Oesterle et al. (1976, 1978, 1984)

As part of an extensive experimental program conducted by the Portland Cement


Association (PCA), Oesterle et al. (1976, 1978, 1984) carried out a large number of
tests to investigate the inelastic behaviour of reinforced concrete shear walls subjected
to monotonic and reversed cycling loading. The test specimens were similar to that
tested by Barda et al. (1977) with different wall cross sections including flanged, barbell
and rectangular shapes as shown in Figures 2.23 and 2.24. The compressive strength of
the concrete ranged from 21 MPa to 55 MPa and the yield stress of reinforcing steel was
about 420 MPa. The reinforcement details for each specimen were selected based on the
provisions of the 1977 ACI Building Code (1977). The ratio of longitudinal
reinforcement was approximately 0.3 percent with the amount of transverse
reinforcement varied from 0.3 to 1.38 percent. In addition to the wall cross section, the
primary experimental parameters included the axial compression load, the compressive
strength of concrete, confinement reinforcement in boundary elements, the amount of
transverse shear reinforcement, the level of shear stress, load history and lap splices
within plastic regions. Based on the experimental results the following conclusions
were drawn:

• The strength and ductility of specimens under reversed cyclic loading are less
than those of similar specimens but under monotonic loading.

• The response of shear wall specimens under reversed cyclic loading was ductile.
This aspect was attributed to the special reinforcement details used in this study.

• Transverse reinforcement greater than the amount recommended by ACI


Building Code (1977) did not improve the total shear strength.

'
• The axial load corresponding to an axial compressive stress of 0.1 f c increased
the ductility of the walls.

• The confinement reinforcement in the boundary elements had a considerable


effect on the inelastic response of the walls.

2-55
• Failure of shear walls under reversed cyclic loading is limited by web crushing.

• Web crushing depended on the amount of shear stress and the level of
deformation.

203 mm
2360 mm 1220 mm
A

4570 mm

1910 mm 102 mm
5383 mm

A
610 mm

3050 mm 1220 mm

Section A-A

Figure 2.23 – Typical shear wall specimen tested by Oesterle et al. (1984).

2-56
102 mm
(a)
1910 mm

305 mm
102 mm
(b)
305 mm 1300 mm 305 mm

102 mm

910 mm
(c)

102 mm 1706 mm 102 mm

Figure 2.24 – The cross-section of specimens tested by Oesterle et al. (1984):


(a) rectangular, (b) barbel, (c) flanged .

2.4.3 Maier and Thurlimann (1985)

Maier and Thurlimann (1985) undertook tests on 10 low-rise shear wall specimens, of
dimensions shown in Figure 2.25, to investigate strength and behaviour under vertical
and horizontal loads. Eight specimens were tested under monotonic loading and two
specimens (S5 and S7) were tested under reversed cyclic loading. The height-to-length
ratio of the specimens was about 1.0 and the ratios of reinforcement were varied as 0.0,
0.57 and 1.0 percent in the horizontal direction and 0.57, 1.0 and 2.5 percent in the
vertical direction. The compressive strength of concrete was 29 to 37 MPa and the yield
stress of reinforcement was about 500 MPa. The test variables included the shape of
wall cross section, openings in compression zone, arrangement and amount of

2-57
reinforcement and amount of axial load. The main conclusions drawn on the
experimental results are summarised as follows:

• The horizontal resistance is a function of the cross-sectional geometry, the


percentage of longitudinal reinforcement and the axial load. Increasing the last
two parameters and adding boundary elements improve the strength of shear
walls.

• The amount of transverse reinforcement has a negligible effect on the ultimate


load but improves the ductility of shear walls.

A 240 mm
1200 mm

100 mm
1820 mm

B B

1700 mm A 700 mm
380 mm
150 mm

Section A-A
100 mm
400 mm

100 mm

260 1180 mm 260


150 mm

Section B-B

Figure 2.25 – Typical shear wall specimen tested by Maier and Thurlimann (1985).

2-58
• A high axial load or a high percentage of longitudinal reinforcement decreases
the ductility of shear walls and results in brittle fracture.

• Under lateral cyclic loading and low axial load, no reduction in ultimate strength
and no significant change in ductility were observed as compared with similar
specimen but under monotonic loading. However at higher axial load, the failure
load is decreased due to the load reversal.

• In general, the observed failure mode was lateral splitting of the concrete and
spalling of concrete cover in the compression corner of the specimens.

• Under high axial load, a horizontal fracture zone over the length of specimens
was observed.

While the study provided useful information regarding the behaviour of squat shear
walls cast with the normal strength concrete, the effects due to the use of high strength
concrete remained to investigate.

2.4.4 Lefas et al. (1990) and Lefas and Kotsovos (1990)

Lefas et al. (1990) undertook an experimental program on the behaviour of shear walls
loaded under either monotonic or cyclic loading. It the first part of the experiment
programme, thirteen shear walls, shown in Figure 2.26, were tested under a combination
of the constant axial load and the horizontal monotonic loading. The experimental
parameters investigated were the height-to-width ratio, the axial load, the concrete
strength and the ratio of transverse reinforcement. Two groups of shear walls with the
height-to-width ratios equal to 1 and 2 were tested. The wall cross-sections were
rectangular with additional reinforcement confined by stirrups used at the edge
boundaries. The compressive strength of the concrete was varied from 30 MPa to
55 MPa and 500 MPa grade reinforcement was used. The ratio of longitudinal
reinforcement for all specimens was constant and equal to 2.4 percent. The horizontal

2-59
reinforcement ratio ranged from 0.4 percent to 1.1 percent. The main findings from the
monotonic test were:

• The transverse reinforcement did not significantly influence the shear strength of
the walls.

• The strength and deformation characteristics of specimens were independent of


the compressive strength of concrete within the range of 30 to 55 MPa.

150 mm
A

1300 mm

65 mm
1750 mm

B B
300 mm

1150 mm A 200 mm

Section A-A
200 mm

250 mm 650 mm 250 mm

Section B-B

Figure 2.26 – Typical shear wall specimen tested by Lefas et al. (1990).

2-60
• While the axial load decreased either vertical or horizontal deformation it caused
an increase in both the horizontal failure load and the lateral stiffness
characteristics. For a higher ratio of height-to-width this increase became further
pronounced.

In the second part of the investigation, Lefas and Kotsovos (1990) tested one shear wall
under monotonic loading and three specimens in the different regimes of reversed cyclic
loading. The test specimens were similar to those tested previously and shown in Figure
2.5. The longitudinal and transverse reinforcement ratios for all specimens were equal to
1.5 percent and 0.35 percent, respectively. The following conclusions were drawn on
the experimental results:

• The shear wall specimens showed a flexural response before the failure with
considerable energy dissipation.

• The strength and ductility of the specimens were independent of the cyclic
loading regime.

This study led to further information about the affects of some major parameters on the
response of shear walls subjected to monotonic loading. It also showed the
independency of cyclic response of shear walls to the reversed cyclic loading regime.
All specimens failed in a flexural mode and, therefore, the response of shear walls
failing in shear was not investigated.

2.4.5 Pilakoutas and Elnashai (1995)

An experimental research program was conducted by Pilakoutas and Elnashai (1995a,


1995b) to study the behaviour of six RC shear walls under severe cyclic loading to
destruction. The shear wall specimens had a rectangular cross section with a height-to-
width ratio equal to 2 as shown in Figure 2.27. The walls were built using concrete with
a compressive strength between 32 and 46 MPa. The yield stress of reinforcement was
approximately 500 MPa. The cyclic loading was imposed on the specimens at a slow

2-61
rate in displacement increments of 2 mm until failure occurred. At each displacement
level two full excursions performed. All specimens had no imposed axial load. The
main test parameters included the ratios of transverse and longitudinal reinforcement in
the web wall and the longitudinal and tide reinforcement ratios as well as the width of
the boundary elements. The following conclusions were drawn on the experimental
results:

• The failure mode was primarily dependant on the amount and distribution of the
wall transverse reinforcement.

• The shear force is transmitted through the concrete in compression and the link
reinforcement in the tensile zone. The walls failed after the yielding of link
reinforcement and when the shear capacity of compressive concrete was
exceeded.

• The wall transverse reinforcement beyond the amount needed to carry the
maximum lateral load did not have significant effects on the strength and
ductility aspects of specimens.

• Although the shear deformation was a considerable component of total


deformation, the overall energy dissipation was significantly attributed to the
flexural action.

• The walls tested without applied axial load exhibited a longitudinal extension
due to cyclic loading, particularly in the lower levels of the walls.

2-62
600 mm

Stiff Top Beam

1200 mm

60 mm

Stiff Bottom Foundation

Section A-A

Figure 2.27 – Typical shear wall specimen tested by Pliakoutas and Elnashai (1995a).

2.4.6 Gupta and Rangan (1996)

Gupta and Rangan (1996) conducted tests on eight HSC shear wall specimens subjected
to constant vertical and monotonically increasing horizontal loads. The test specimens
were similar to those tested by Maier and Thurlimann (1985), as shown in Figure 2.28,
but built by using high strength concrete. The compressive strength of concrete ranged
from 60 MPa to 80 MPa and the nominal yield stress of the reinforcement was
400 MPa. The major test parameters of the study were the ratio of horizontal
reinforcement, the ratio of vertical reinforcement and axial load.

2-63
200 mm
1300 mm
A

1000 mm
1600 mm

B B

400 mm
1800 mm A 575 mm
100 mm

Section A-A
75 mm
375 mm

100 mm

400 mm 1000 mm 400 mm


100 mm

Section B-B

Figure 2.28 - Typical test specimen tested by Gupta and Rangan (1996).

2-64
For each specimen the full axial load was initially applied and then the horizontal load
was monotically applied in increments of 50 kN before cracking and 10 kN after
cracking up to failure. A summary of test results is given in Table 2.4.

Based on the study the following conclusions were drawn:

• The lateral failure load was increased and the ductility was reduced with an
increase in the vertical load.

• An increase in longitudinal reinforcement ratio led to an increase in the


horizontal failure load. This increase was greater when the vertical load was zero.

• Increasing the transverse reinforcement ratio gave only a small increase in the
lateral failure load.

Table 2.4 - Summary of test results reported by Gupta and Rangan (1996)

Specimen Equivalent ratio Equivalent ratio Axial Failure Failure


of transverse of longitudinal load transverse load mode
reinforcement reinforcement
(percent) (percent) (kN) (kN)

S-1 0.5 1.0 0.0 428 shear

S-2 0.5 1.0 610 720 shear

S-3 0.5 1.0 1230 851 shear

S-4 0.5 1.5 0.0 600 shear

S-5 0.5 1.5 610 790 shear

S-6 0.5 1.5 1230 970 shear

S-7 1.0 1.0 610 800 shear

S-F 0.5 1.0 310 487 flexural

2-65
The conclusions from the study are similar to those for tests carried out on specimens
cast with normal strength concrete. Although this study did not involve tests with cyclic
loading, it is one of few available on HSC shear walls.

2.4.7 Kabeyasawa and Hiraishi (1998)

Kabeyasawa and Hiraishi (1998) reported the results of an extensive experimental


research program undertaken on 21 HSC shear walls as a part of a five-year national
research project in Japan. The specimens were approximately one-quarter scale with the
same cross sectional dimensions (Figure 2.29). Four series of tests were conducted as
follows:

• RC flexural tests: six specimens designed to yield first in flexure and then reach
different ultimate deformations.

• RC preliminary shear tests: two specimens subjected to non-symmetric lateral


loading designed to fail in shear mode prior to yielding the longitudinal
reinforcement.

• RC shear tests: eight specimens subjected cantilever loading designed to fail in


shear mode prior to yielding the longitudinal reinforcement.

• RC bi-directional tests: five specimens tested to investigate the deformation


capacity of HSC flexural walls subjected to bi-directional lateral earthquake
loads.

The concrete strength and the yield stress of reinforcing steel used in walls varied from
60 to 137 MPa and from 753 to 1420 MPa, respectively. The concrete strength, the
reinforcing steel strength, transverse and longitudinal reinforcement ratios and axial and
horizontal load conditions were systematically varied for the investigation of strength
and ductility characteristics in both flexural or shear failure modes. It was concluded
from the experimental results that:

2-66
• The high strength concrete can effectively be used to build ductile shear walls
failing in a flexural mode, though the pinching effect observed in the hysteresis
aspect should be considered in design and analysis.

• The flexural strength and deformations could be evaluated using the conventional
flexural theory.

• The strength of walls failing in shear with yielding of web reinforcement could
fairly be estimated using the arch and truss model by modifying the effective
factors for HSC. For walls failing in shear without yielding of web
reinforcement, however, the model overestimated the shear strength.

A
600 mm
4200 mm

3000 mm

200 mm

B B
600 mm

2000 mm A 500 mm

Section A-A
150 mm

80 mm

200 1300 mm 200 mm


150 mm

Section B-B

Figure 2.29 – Typical shear wall specimen tested by Kabeyasawa et al. (1993).

2-67
2.4.8 Palermo and Vecchio (2002)

Palermo and Vecchio (2002) tested two large scaled squat flanged shear walls, DP1 and
DP2, subjected to lateral displacement reversals. The experimental program was
designed to investigate the behaviour of shear walls under cyclic displacements and the
main test parameter included in the study was axial load. The geometrical properties of
the specimens were similar as shown in Figure 2.30.

640 mm
A

2020 mm
3280 mm

95 mm 75 mm

B B

4000 mm A 4415 mm
620 mm

Section A-A
685 mm

462.5 mm 462.5 mm
3045 mm

3075 mm
685 mm

Section B-B

Figure 2.30 - Typical test specimens by Palermo and Vecchio (2002).

2-68
The compressive strength of concrete used in wall and end flanges was 20 MPa and the
yielding and ultimate stresses of reinforcement were 605 MPa and 652 MPa,
respectively. The ratios of longitudinal and transverse reinforcement were 0.73 percent
and 0.79 percent, respectively. Specimen DP1 had a constant axial load of 960 kN,
while specimen DP2 had no axial load. After applying the axial load the shear walls
were displaced laterally in 1 mm increments.

The test results showed that the axial load had an important effect on the response of
shear walls under reversed cyclic loading. As it can be observed in Figure 2.31 and
Table 2.5, the maximum load resisted by DP1 was about 40 percent more than that of
DP2. Also DP1 was more ductile and dissipated more energy than DP2. Furthermore,
the hysteresis loops of DP2 showed more pinching. It is important to note that the squat
shear walls specimens presented greatly pinched hysteresis response with small energy
dissipation and their behaviour was extensively dominated by shear induce mechanisms.

The failure modes of the two specimens were different. Whereas DP1 failed because of
crushing of concrete over a widespread zone of web wall, the failure of DP2 was due to
a sliding shear plane extending the entire length of the web wall. The researchers
concluded that the stiffness of the flange wall had key role in the failure mode so that
stiff flanges resulted in forming vertical slip planes in DP1. In DP2, however, probably
weaker concrete in the top zones caused a shear sliding plane to form near the top slab.

Table 2.5 - Summarised Test Results of Palermo and Vecchio (2002)

Axial Positive Direction Negative Direction


load
Corresponding
Specimen Ultimate Corresponding Ultimate Failure mode
displacement
load Displacement load
(mm)
(kN) (kN) (mm) (kN)
Concrete
DP1 940 1298 11.14 -1255 -11.09
Crushing

DP2 0 904 9.15 -879 -9.08 Sliding Shear

2-69
(a)

(b)

Figure 2.31 - Load-Deformation response: (a) Specimen DP1; and (b) Specimen DP2
(Palermo and Vecchio, 2003, 2004).

2-70
2.4.9 Farvashany (2004)

To investigate the questions raised in the earlier research project conducted at the Curtin
University of Technology (Gupta and Rangan, 1996, 1998), Farvashany (2004) carried
out tests on seven HSC shear walls subjected to monotonic loading. The specimens
were similar to those of Gupta and Rangan (1996) but with an increased height-to-width
ratio equal to 1.25. The dimensions of the specimens are shown in Figure 2.32 with the
walls detailed to fail in shear. The compressive strength of concrete ranged from 83 to
100 MPa and the yield stress of reinforcement used in the walls was about 500 MPa.
The main test parameters of the study were axial load, compressive strength of concrete
and longitudinal and transverse reinforcement ratios. Fro the experimental results
Farvashany concluded that:

• An increase in the axial load gave an increase in the horizontal failure load but a
decrease in lateral deformation at the failure load.

• An increase in the ratio of longitudinal reinforcement in the wall resulted in an


increasing in the lateral load at failure.

• The increase in transverse reinforcement ratio did not have considerable effect
on the failure load but it increased the ductility.

• The failure load of specimens increased with increasing the concrete compressive
strength.

This study, accompanied with that of Gupta and Rangan (1996), provides a considerable
data base to validate the analytical models and resulted in valuable information
regarding to the behaviour of HSC low-rise shear walls under monotonic loading.

2-71
200 mm
1300 mm 375 mm
A

1100 mm
90 mm 75 mm
1600 mm

B B

300 mm
1800 mm A 575 mm
100 mm

Section A-A
75 mm
375 mm

90 mm

460 mm 880 mm 460 mm


100 mm

Section B-B

Figure 2.32 – Typical shear wall specimen tested by Farvashany (2004).

2.5 Summary

The analytical, numerical and experimental researches related to the behaviour of shear
walls under monotonic and cyclic loading have been reviewed. Based on the elasticity-
based equivalent uniaxial stains approach, a variety of constitutive models for concrete,
steel reinforcement and the interaction between them have been introduced. These
models can be successfully implemented into a finite element procedure to analysis of
reinforced concrete structures, if they capture major aspects of their behaviour such as
non-linear stress-strain relationships of concrete and steel, concrete cracking,

2-72
compression softening, tension stiffening and so on. Finite element methods based on
the smeared crack model have been commonly used to analyse shear walls. Two major
approaches in this area are fixed and rotating crack model. While the former is a more
general model and has been used by many researches for analysis of shear walls, the
latter is simpler and is more easily implemented. The experiments carried out on normal
strength concrete shear walls have shown the effects of major parameters such as
longitudinal and transverse reinforcement ratios, axial load, the ratio of height-to-length,
edge elements and loading regime on the behaviour of walls. Some of these parameters
have been also investigated by a few researchers that conducted tests on high-strength
concrete shear walls under monotonic loading. However, experimental data on the
behaviour of high-strength concrete shear walls under reversed cyclic loading is very
limited and the study on this area is essentially required.

2-73
CHAPTER 3

FINITE ELEMENT MODELLING

3.1 Introduction

The non-linear finite element analysis of RC structures involves the simulation of their
response under loading and ambient conditions. The significant advantage of this
method is that it can be applied to a wide range of problems. The reliability of a
simulated response is largely dependant on the nonlinear constitutive material models
used as well as the finite element procedure adopted for the analysis. The constitutive
models should not only permit consistent interpretations of past behaviour of RC
members but also predict accurately their future response. The separate constitutive
models describing the different phenomena should be combined and integrated into a
consistent theoretical framework for the treatment of the entire response.

Starting from a description of the crack membrane model (CMM) as the fundamental
theory of analysis, this chapter presents several constitutive material models used in this
study. Then, the finite element implementation of constitutive models and the numerical
algorithms used in non-linear analysis are described.
3.2 Crack Membrane Model

3.2.1 Background

The compression field theory (CFT) (Mitchell and Collins, 1974, Collins, 1978, Collins
and Mitchell, 1981) formulated equilibrium, compatibility and stress-strain relationships
in terms of average stresses and averages strains assuming the cracks as uniformly
distributed and smeared across an element. The CFT was based on a number of basic
concepts and described using Mohr’s circles for the transformations of strains and
stresses with the full response of a RC membrane determined by adding the stress-strain
relationships in principal directions of the concrete to that of the reinforcing steel.

The basic assumption behind the CFT is that the directions of the principal average total
strains are the same as the directions of principal average concrete stresses and, thus, the
cracks are taken as rotating and stress-free. The main drawback of the CFT was that the
tension stiffening effects were neglected so that the predicted response of a RC structure
is too soft. To address this weakness, as well as to determine a better constitutive
relationship for concrete, Vecchio and Collins (Vecchio and Collins, 1982, 1986)
developed the modified compression field theory (MCFT) by introducing empirical
relationships between average total strains and the average tensile and compressive
stresses in the concrete to account for tension stiffening and compression softening
effects, respectively.

The MCFT contributed significantly to the progress in the research areas of structural
concrete in the 1980s and the 1990s and was accepted by many analysts. Although the
MCFT led to more realistic results, the link to limit analysis was lost as equilibrium was
expressed in the form of average stresses across an element (Kaufmann and
Matri, 1998). Also, the empirical tension stiffening relationship for concrete has and
continues to undergo debate with various researchers proposing various models and
variations to that originally formulated (Collins and Mitchell, 1991, Okamura and
Maekawa, 1991, Belarbi and Hsu, 1994, Bentz, 1999).

3-2
Kaufmann and Marti (1998) proposed an alternative tension stiffening model by
developing a two-dimensional representation of the tension chord model (TCM) of
Marti et al. (1998) and combining it with the basic components of the MCFT. The new
model was named the crack membrane model (CMM) for the analysis of RC
membranes. The CMM maintains equilibrium at the crack faces and, thus, there is no
need to describe the equilibrium in terms of average stresses across an element. In this
way the link to limit analysis is maintained. Moreover the tension stiffening effect is
developed as a rational model based on a simple bond shear stress-slip constitutive
relationship. Assuming the crack faces to be stress free and the concrete average
principal stresses and the average principal strains being coincident, the model was
implemented as a rotating crack procedure. Although a generalised CMM involving
shear and normal stresses acting on the cracks was expressed (Kaufmann and Marti,
1998), no progress in this direction has yet been implemented (Marti, 2005). In this
study, the CMM is considered in its more general form and developed as a fixed crack
model.

3.2.2 Concepts and Relationships

Considering a plane concrete element subjected to in-plane stresses, uniformly


intersected by a system of parallel cracks and orthogonally reinforced by a regular grid
of steel bars, as shown in Figure 3.1, from equilibrium at the cracks, it can be shown
that

σ x = σ cn cos 2 θ r + σ ct sin 2 θ r + τ cnt sin (2θ r ) + ρ xσ sx (3.1a)

σ y = σ cn sin 2 θ r + σ ct cos 2 θ r − τ cnt sin(2θ r ) + ρ yσ sy (3.1b)

τ xy = 0.5(σ cn − σ ct )sin (2θ r ) − τ cnt cos(2θ r ) (3.1c)

where θr is the angle between a vector normal to the cracks and the global X-axis
(-π/2 ≤ θr ≤ π/2); σx, σy and τxy are the in-plane normal and shear stresses in the global
XY-coordinate system, respectively; σcn and σct are the concrete stresses normal and

3-3
parallel to the direction of cracking, respectively; τcnt is the corresponding shear stress;
ρx and ρy are the steel reinforcement ratios in the global X- and Y-directions and σsx and
σsy are the stresses in the X- and Y-reinforcement, respectively. It is assumed that the
element is sufficiently large compared to the crack and bar spacings and that the internal

forces are uniformly distributed over a distance of several cracks and bar spacings. The
average tensile stresses in the concrete between cracks in the X and Y directions, as per
the tension chord model shown in Figure 3.2, can be written as

λ x f ct
σ ctsx = ⋅ g (τ b ) (3.2a)
2

λ y f ct
σ ctsy = ⋅ g (τ b ) (3.2b)
2

σy ρ x σsx
(a)
τxy
(c) σcncos θ r
θr τcnt cos θ r
σx
+
1

σx σx τcnt sin θr
τxy
σct sin θ r
σy
(b) (d) ρ y σsy

σct cos θ r + σcn sin θr


t

θr τ cnt cos θr θ r τcnt sin θr


τxy
σy
1

Figure 3.1- Orthogonally reinforced membrane subject to plane stress: (a) applied
stresses; (b) axis notation; (c) and (d) stresses at a crack (Kaufmann and
Marti, 1998).

3-4
(a)

m
sr
m
cft

sr
m
sr

srmy
θr
srmx

(b) Y λ y fct

s rmy s rmy s rmy


X

Tension Stiffening
λ x f ct

Stresses
srmx srmx

Figure 3.2- Tension stiffening stresses: (a) in material axis directions and (b) in
orthogonal tension chord of cracked membrane (Kaufmann and Marti,
1998).

where f ct is the concrete tensile strength, λ x f ct and λ y f ct are the maximum tensile

stresses in the concrete due to tension stiffening in the X and Y directions, respectively,
and g (τ b ) is a function of bond shear stress τ b and is discussed further in the next

parts. The tension stiffening factors, λ x and λ y , are determined based on the crack

spacings as shown in Figure 3.2 and are given by

3-5
Δσ cx s rm
λx = = (3.3a)
f ct s rmx0 cos θ r

Δσ cy s rm
λy = = (3.3b)
f ct s rmy0 cos θ r

where Δσcx and Δσcy are the X- and Y-component stresses due to tension stiffening, s rm

is the average crack spacing measured normal to the cracks and s rmx0 and s rmy0 are

average crack spacings of uniaxial tension chords in the X and Y directions,


respectively. The crack spacings in a fully developed crack pattern in the X and Y
directions, s rmx and s rmy , are limited by (Marti et al., 1998)

s rmx0
≤ s rmx ≤ s rmx0 (3.4a)
2

s rmy0
≤ s rmy ≤ s rmy0 (3.4b)
2

leading to 0.5 ≤ λ x ≤ 1 and 0.5 ≤ λ y ≤ 1 .

Modelling steel-concrete bond shear stress as a stepped, rigid-perfectly plastic


constitutive relationship (Figure 3.3(a)), the TCM gives the crack spacings for uniaxial
tension chords in X and Y directions as (Marti et al., 1998)

f ∅ (1 − ρ x )
srmx 0 = ct x (3.5a)
2τ b 0 ρx

srmy 0 =
(
f ct ∅ y 1 − ρ y ) (3.5b)
2τ b0 ρy

3-6
(a) τb
τ b0
τ b1

δy δ

1
2 fct
1 θc

cos θc
Δσcx

θc

(b) Δσcy
sin θc

Figure 3.3- Tension chord model: (a) bond stress versus slip relation ship and (b)
tension stiffening stress components (Foster and Marti, 2003).

where ∅x and ∅y are the diameters of the reinforcing bars and ρ x and ρ y are the ratios

of reinforcement in X and Y directions, respectively, and τ b0 is the plastic bond

strength before yielding of the reinforcing steel and reduces to τ b1 after yielding.

Considering the fixed crack approach, the crack is geometrically fixed once first
cracking takes place and then the cracked concrete is treated as an orthotropic material
with the fixed 1- and 2- orthotropic axes that are, respectively, normal and tangential to
the direction of the crack (Figure 3.4).

Applying equilibrium of forces at the point of cracking (Figure 3.3(b)) gives

Δ σ cx cos 2 θ c + Δ σ cy sin 2 θ c = f ct (3.6)

3-7
Y

2 Reinforcement
1
θr
X

Figure 3.4– Orthotropic and global axes in a cracked concrete element.

where θ c is the angle between the global X axis and the major principal stress,

measured midway between the cracks. At the point of cracking θ c = θ r where θ r is the
angle of principal stresses at the cracks. Substituting Eq. (3) into Eq. (6) results in the
Vecchio and Collins (1986) crack spacing equation:

−1
⎡ cos θ r sin θ r ⎤
s rm = ⎢ + ⎥ (3.7)
⎢⎣ s rmx0 s rmy0 ⎥⎦

Foster and Marti (2002) showed that while the crack spacing can be rationally and fully
determined using Mohr’s failure criteria, Eq.(3.7) is a reasonable approximation to the
exact solution. To analyse the post-cracking behaviour, the crack orientation and crack
spacing are assumed fixed and equal to their values at the moment of cracking.

For fully developed cracks spaced at s rm across a continuum, and assuming no


transmission of normal stresses in the concrete across the cracks, the crack width is
given by (Foster and Marti, 2002) as

wcr = s rm (ε 1 + ν 12 ε 2 − λ f ct 2 E c ) (3.8)

3-8
where ε1 and ε2 are the average strains for element in 1- and 2- directions respectively,
ν12 is Poisson’s ratio for expansion in the 1-direction resulting from stress in the 2-
direction, λ is the uniaxial tension stiffening factor (0.5 ≤ λ ≤ 1.0) and Ec is the initial
modulus of elasticity for concrete.

If the dowel action across the cracks is ignored and cracks are assumed to be
continuous, the stresses in the solid concrete between the cracks could be considered the
same as stresses on the cracks (Bazant and Gambarova, 1980). By substitution of
Eq. (3.2) into Eq. (3.1) and writing in terms of the orthotropic 1-2 coordinate system,
the average stresses in an element due to any boundary tractions are given by

λ x f ct
σ x = σ c1 cos 2 θ r + σ c 2 sin 2 θ r + τ c12 sin (2θ r ) + ρ xσ sx + g (τ b ) (3.9a)
2

λ y f ct
σ y = σ c1 sin 2 θ r + σ c 2 cos 2 θ r − τ c12 sin (2θ r ) + ρ yσ sy + g (τ b ) (3.9b)
2

(σ c1 − σ c 2 )
τ xy = sin (2θ r ) − τ c12 cos(2θ r ) (3.9c)
2

where σ c1 , σ c 2 and τ c12 are average tensile, compressive and shear stresses in the
concrete in the orthotropic 1-2-axis system, respectively. It is to be noted that prior to
cracking the concrete is taken to carry tensile stresses in a linear elastic manner. After
cracking, concrete tensile stresses result from two independent mechanisms, namely
tension softening and tension stiffening. Whilst taking the former into account is
especially important in analysing the concrete members with little or no reinforcing
steel (Vecchio, 2000b), the latter has a significant effect on the general response of
reinforced concrete members. It is to the advantage of the CMM that both effects are
incorporated into Eq.(3.9) separately and, thus, the model can be applied for a wide
range of problems without modification including problems involving fracture.

3-9
3.3 Constitutive Relationships

The accuracy and reliability of nonlinear finite element analysis of reinforced concrete
structures depends on the underlying constitutive relationships applied in each finite
element domain. To simulate the nonlinear response of cracked reinforced concrete; a
combined model based on separate material models for the concrete state, reinforcing
steel state and the state of interaction between concrete and reinforcement need to be
considered (Foster et al., 1996).

3.3.1 Constitutive Models for Concrete

To establish the constitutive models for concrete, a diverse number of approaches have
been used and can be classified as the elasticity-based approach, plasticity-based
approach, damaged-based approach, micro modelling and endochronic theory
(CEB, 1996). Among these the elasticity-based approach is more popular because of its
simplicity, the capability of modelling concrete under various states of loading
(monotonic, cyclic, etc.) and giving reasonably good results (ASCE Committee 447,
1982, Bahlis and Mirza, 1987). The elasticity-based approach includes a variety of
models, the most common being the orthotropic, equivalent uniaxial, model. Many
researchers have employed the non-linear elasticity approach to simulate the behaviour
of RC shear walls under cyclic loading (Okamura and Maekawa, 1991, Sittipunt and
Wood, 1995, Elmorsi et al., 1998, Vecchio, 1999, Kwak and Kim, 2004, Palermo and
Vecchio, 2004). A common version of this model is based on the concept of “equivalent
uniaxial strains” introduced by Darwin and Pecknold (1976) , with the constitutive
relationships for concrete founded on the uniaxial stress-strain curves modified to take
into account the effects of stresses in other directions.

Adopting orthotropic, equivalent uniaxial strain approach in this research, the


constitutive relationships for concrete in tension and compression are expressed using
two sets of curves. The behaviour of concrete subjected to biaxial states of stress is
taken into account by modifying the uniaxial response of the concrete.

3-10
Concrete in Tension

Before cracking, concrete in tension is considered as a linear elastic material. Adopting


the bilinear softening model of Petersson (1981) for the post-cracking response, the
envelope curve of concrete in tension is expressed by the monotonic stress-strain curve
as shown in Figure 3.5(a). The main softening parameters of the model are given by

1 2 18 Ec G f
α1 = ; α 2 = α 3 + α1 ; α3 = (3.10)
3 9 5 l ch f ct2

where Ec is the initial elastic modulus of concrete, G f is the fracture energy and lch is

a characteristic length over which the fracture energy is dissipated. For RC elements
analysed using the smeared crack approach, this length is taken as the crack spacing.

The unloading stiffness modulus for concrete in tension is expressed by

σa
Ed = (3.11)
εe + ε p 2

σc1 σc1
f ct f ct

σa Ed
α1f ct Ec Ec
1
εc1 1 1 εc1
εtp α2ε tp α3ε tp εp/2 εp/2 εe εa

(a) (b)

Figure 3.5- Concrete in uniaxial tension: (a) envelope stress-strain curve (b) unloading
and reloading curves (Foster and Marti, 2003).

3-11
where σ a is the concrete stress just before beginning the unloading and ε e and ε p are

the elastic and plastic components of strain ε a corresponding to σ a and are given by

ε e = σ a Ec
(3.12)
ε p = εa − εe

Concrete in Compression

In this study, the uniaxial stress-strain curve proposed by Thorenfeldt et al. (1987) is
adopted (Figure 3.6). This model is a convenient expression, which accurately describes
the shape of the ascending and descending branches of high-strength concrete cylinder
stress-strain curve (Collins et al., 1993). The curve forms the envelope for concrete in
compression with the compressive stress expressed as


σ c = − f cp (3.13)
n − 1 + η nk

(
where η = ε c ε cp and n = Ec Ec − Ecp . )

In Eq. 3.13, ε c is the concrete strain, ε cp is the strain corresponding to the peak stress

f cp on the curve that is equivalent to the compressive strength of concrete, Ec is the

initial elastic modulus of the concrete, Ecp = f cp ε cp and k is a decay factor that

controls the post-peak response. Collins and Porasz (1989) calibrated the decay factor
for conventional and high strength concrete and proposed it as follows:

ε ≤ ε cp ............ k = 1 (3.14a)

ε > ε cp ............ k = 0.67 + f cp 62 ≥ 1.0 (3.14b)

3-12
where f cp is in MPa.

σc

f c'

σa

Ed
Ec
1 1 εc
ε cp εa

Figure 3.6 – Envelope stress-strain curve for concrete under uniaxial compression.

For unloading of concrete in compression, the stiffness modulus E d defined by


Filippou et al. (1983) is used; that is

σa
Ed =
ε a − ε a (0.1 + 0.15 ε a ε cp )
(3.15)

where σ a and ε a define the point of unloading (refer Figure 3.6).

Biaxial Behaviour of Concrete

The strength and stress-strain response of concrete subjected to biaxial states of stress
are different compared with uniaxial conditions and vary as the functions of the
combinations of stresses. Appropriate modelling of concrete in the two-dimension using

3-13
the orthotropic equivalent uniaxial approach needs modification of uniaxial concrete
response in tension and in compression so that the biaxial behaviour of concrete can be
captured. Figure 3.7 shows the biaxial strength envelope for concrete used by Foster and
Marti (2003) and is used in this study.

Commonly for a two-dimensional concrete model three biaxial states of stress are
considered: biaxial compression (C-C), biaxial compression and tension state (C-T) and
the biaxial tension (T-T). In biaxial compression the strength of concrete is greater than
the uniaxial compressive strength (Kupfer and Gerstle, 1969, Liu et al., 1972) due to the
confinement effect. Under combinations of compression and tension the effects of
tensile cracking causes the compressive strength of concrete to be reduced and this is
known as the compression softening effect (Robinson and Demorieux, 1977, Vecchio
and Collins, 1982, 1986, Miyakawa et al., 1987, Belarbi and Hsu, 1995). Finally, when
concrete is subjected to biaxial tension the strength is close to that in uniaxial tension
(Kupfer and Gerstle, 1969).

The biaxial compressive strength of concrete f c* may be written as the in-situ uniaxial

compressive strength f cp modified by a strength factor β as

−σ2c /fcp
(0.6, 1.25)
(1.15, 1.15)
1.0

0.6 (1.25, 0.6)

1.0 −σ1c /fcp


0.6

3-14
Figure 3.7- Strength envelope curve for concrete in biaxial stress (Foster and Marti,
2003).

f c* = β f cp (3.16)

In the Eq. 3.16, the coefficient β is considered as a scaling factor applied to the
*
compressive stress-strain curve depending on the state of stress. The strain ε cp

corresponding to f c* may be considered the same as the uniaxial peak strain ε c

(Vecchio and Collins, 1986) or modified using the same factor β as follows (Vecchio
and Collins, 1982, Belarbi, 1991):

*
ε cp = β ε cp (3.17)

The latter approach is used in this study (Figure 3.8).

For the C-C state, β is a confinement factor that can be determined from the biaxial
strength envelope (Figure 3.7). For the T-C state, β is a strength reduction factor
obtained from the modified compression field theory (Vecchio and Collins, 1986) and is
expressed as

1
β= ≤ 1.0 (3.18)
ε
0.8 + 0.34 c1
ε cp

where ε c1 is the tensile strain in the material 1-direction.

The biaxial tensile strength of concrete f ct* for the C-T and T-T states, based on the
biaxial strength envelope (Figure 3.7), is given by

σ c2
f ct* = f ct if ≤ 0.6 (3.19a)
f c'

3-15
σ σ
f ct* = 2.5 f ct (1 − c 2 ) if 0.6 < c 2 ≤1.0 (3.19b)
f c' f c'

−σ c
β fcp
fcp

β >1
β fcp
β =1

Ed β <1
1
βε cp ε cp βε cp −ε c

Figure 3.8– Scaling of biaxial compressive stress-strain curves (after Foster and
Marti, 2003)

3.3.2 Constitutive Model for Reinforcing Steel

The response of reinforcing steel subjected to uniaxial loading is expressed using a


tri-linear stress-strain curve with the properties as defined in Figure 3.9. The unloading
modulus for the steel is taken as equal to initial elastic modulus, that is E d = E s .

3.3.3 Constitutive Models for Interaction between Concrete and Steel

Concrete Tension Stiffening

Based on the CMM and the work of Marti et al. (1998) and Kaufmann and
Marti (1998), considering a stepped rigid-perfectly plastic bond slip relationship as
shown in Figure 3.10, Foster and Marti (2003) developed a concrete tension stiffening
model. By the tension chord model, the mean concrete tensile stiffening stress σ ctsm

3-16
between cracks is determined for various loading stages depending on the conditions of
the reinforcing steel stresses at and between cracks as follows:

σs
fu
fw 1
Eu

Ew
fsy 1
Ed
1

Es
1 εs
εy εw εu

Figure 3.9– Trilinear stress-strain model for reinforcing steel .

a) if σ cr ≤ f sy then

τ b λf ct
σ ctsm = (3.20)
τ b0 2
where
ε m Es ∅
τb = ≤ τ b0 (3.21)
2srm

b) if σ s min ≤ f sy < σ sr then

ρ
σ ctsm = (σ sr − σ sm ) (3.22)
(1 − ρ )
where (Kaufmann and Marti, 1998)

σ sm = f sy −
(σ sr − f sy )2 ∅ ⎛ τ b0 ⎞
⎜⎜ ( ) τ τ s
− 1⎟⎟ + σ sr − f sy b0 − b0 rm (3.23)
4τ b1srm ⎝ τ b1 ⎠ τ b1 ∅

and

3-17
2
τ b0 srm

− ( f sy − Esε m )τ b1∅srm ⎛⎜⎜ ττb0 − EEs ⎞⎟⎟ + EEs τ b0τ b1 srm2
⎝ b1 w⎠ w ∅ (3.24)
σ sr = f sy +
⎛ τ b0 Es ⎞
0.5 ⎜⎜ − ⎟⎟
⎝ τ b1 Ew ⎠

3-18
c) if f sy < σ s min then

τ b λf ct
σ ctsm = (3.25)
τ b1 2
where
ε m Essec ∅
τb = ≤ τ b1 (3.26)
2 srm

In Eqs. 3.20 to 3.26, ε m is the average strain over the length of the element; σ sm and

σ s min are the average stress and the minimum stress in steel between cracks,
respectively; σ sr is the stress in the steel at the cracks that is the maximum stress, σ cm

is average stress in concrete between cracks and E s and E s sec are the initial elastic

modulus and the secant stiffness modulus of reinforcing steel corresponding to the
strain ε m , respectively. Figure 3.10 illustrates the distributions of shear bond, concrete
and steel stresses between two cracks of a cracked reinforced concrete tension chord.

In applying the equations described above to the X and Y direction reinforcement λ ,


∅, ρ , s rm , ε m , σ ctsm , σ sm and σ sr must be replaced by appropriate values for

each direction such as λ x or λ y and so on.

Finally, the secant modulus corresponding to concrete tension stiffening is obtained by


relating the mean stress in the concrete, between cracks, to the mean strain described by

Ects = σ cm ε m (3.27)

3-19
s rm / 2 C.L s rm / 2 C.L s rm / 2 C.L
crack
τb
Fc / 2
Fs Fs − Fc
Fc / 2
τb τb τb

τ b0 τ b0
τ b1 τ b1

σc σc σc

σ ctsm
σ ctsm σ ctsm

σs σs σs

σ sr
σ sr σ sm
f sy f sy
f sy σ sm
σ sr σ sm

(a) (b) (c)

Figure 3.10 - Bond shear, concrete and steel stresses between cracks: (a) before yielding
and (b) yielding of a part of length (c) yielding of whole length.

Shear Transfer

After cracking, as loading continues and if the direction of the crack is fixed, shear
forces are transmitted across the cracks through aggregate interlock and dowel action.
Analytical studies show that the contribution of the aggregate interlock is more
pronounced than that of the dowel action (Walraven, 1980). Since the ratio of
reinforcement in RC shear walls is usually low, for a finite element analysis based on
the fixed-crack approach the stiffness due to dowel action may be ignored (Okamura
and Maekawa, 1991, Elmorsi et al., 1998). In this study, the shear transfer is assumed to
be attributed to the aggregate interlock effect. Among the variety of aggregate interlock
models discussed in Chapter 2, the model expressed by Walraven and Reinhardt (1981)
can be simply employed in a finite element model. However, three issues limit its

3-20
effective application in the analysis of RC members. First, free initial crack slip results
in some instability into the numerical solution. Second, it does not address the effect of
reinforcement crossing a crack interface on the shear response resulting from the
aggregate interlock. Finally, the model was verified against the experimental results
obtained from the tests carried out on normal strength concrete. The aggregate interlock
model adopted in this study is developed, primarily, from the experimental works of
Walraven who carried out one of the most extensive and comprehensive experimental
test studies on the aggregate interlock mechanism in conventional strength concrete
(Walraven et al., 1979, Walraven, 1980) and HSC (Walraven, 1995). The major
findings for conventional strength concrete based on these observations can be
summarised as:

• Dowel action does not contribute significantly to the shear resistance across
cracks for reinforcement ratios between 0.56% and 2.24%.

• The behaviour of the externally reinforced specimens without internal


reinforcement crossing the crack and with internal reinforcement crossing the
crack was fundamentally different. The contribution of aggregate interlock to
resist against shear forces is greater for the case where reinforcing bars cross
cracks than is the case for unreinforced specimens. It was concluded that local
compressive struts formed in the region around the reinforcing bars due to
secondary cracking and that this effect is the probable cause of the difference in
behaviour (Figure 3.11). It should be noted that this phenomenon is distinct from
shear forces carried by dowel action but, nevertheless, is a “dowel effect”.

• Specimens without reinforcement crossing the cracks exhibited an initial slip


before interlocking of the aggregate particles, whereas, for the case where
reinforcing steel passes through a crack no initial slippage before take up of load
was observed.

• Increasing the reinforcement ratio and the concrete strength led to an increase in
shear resistance. Variation of bar diameters, however, for a constant
reinforcement ratio had no significant effect on the behaviour.

3-21
co secondary cracking
m pre
ssi
on
str
ut

Figure 3.11 – Compression struts in the vicinity of the reinforcing steel due to
secondary cracking.

A problem with the Walraven (1981) empirical model, however, is that for crack
openings of greater than approximately 1.5 mm, the shear capacity across the crack
becomes negative and ,therefore, is taken as zero (see section 2.3). To avoid
discontinuity and subsequent numerical problems due to that, the Walraven data was
re-evaluated and a new model is proposed. Using a regression analysis, it was found for
pre-cracked, unreinforced, normal strength concrete that the shear stress can be given by

1.18 f cp0.75 δ s
τ cnt − τ co = 2
(3.28)
0.1 + 1.6 wcr + 3 wcr

where δ s and wcr are the crack slip and crack width, respectively (in mm), fcp is the

compressive strength of the in-place concrete (in MPa) and τco defines the focal point
and accounts for the initial slip for the case of pre-cracked concrete. The focal point is
taken as

τ co = − f cp / 24 ...... for reinforced concrete (3.29)

(after Walraven and Reinhardt, 1980) and

τ co = 0 ...... for reinforced concrete (3.29b)

Note, Eq. (3.29b) comes from the experimental observations of Walraven (1980) and
Milford (1984) which show that for the case where reinforcing steel passes through a
crack, no initial slip occurs before the take up of load.

3-22
If the cube compressive strength of concrete f cc is used, the coefficient of 1.18 in

Eq. 3.28 and the focal point must be taken as 1.0 and τ co = − f cc 30 , respectively. The
aggregate interlock model developed compares well with the results of the Walraven
tests, as shown in Figures 3.11 to 3.13 for concrete with the cube compressive strengths
of 13.4, 37.6 and 56.1 MPa (corresponding to cylinder compressive strength of 10.7,
30.1 and 44.9 MPa). The normalised shear stress versus the crack width also shows a

good correlation ( R 2 = 0.97 ) as illustrated in Figure 3.14.

The limiting shear strength for the case of cracking in unreinforced concrete is taken as
that adopted by (Vecchio and Collins, 1986), that is

f cp
τ cr max = (units of mm and MPa) (3.30)
0.31 + 24 wcr (Dmax + 16)

where Dmax is the maximum size of the aggregate. Eq. 3.30 provides a plastic cap to
the shear on the crack given by Eqs. 3.28 and 3.29.

Walraven (1980) also carried out tests on cracks in reinforced concrete with cube
compressive strengths of 19.9, 30.7, 38 and 56.1 MPa. In addition to the compressive
strength, other test parameters were the ratio of reinforcement crossing the crack
interface, the diameter of reinforcement, maximum aggregate size and the angle
between reinforcement and crack interface. It was found that the contribution of the
compressive struts that form around the reinforcement is significantly affected by the
concrete strength and the reinforcement ratio. To take into account the contribution of
the dowel effect to increase the shear capacity along the cracks, Eqs. (3.28) to (3.30) are
rewritten as

α d 1.18 f cp0.75δ s α d f cp0.5


τ cnt = ≤ (3.31)
2
0.1 + 1.6wcr + 3wcr 0.31 + 24 wcr (a + 16 )

where α d is a dowel effect parameter.

3-23
10 0.1 0.2 0.3 0.4

6
wcr=0.7
τcr (MPa)

2
f cc =13.4 MPa
0
0 0.5 1 1.5 2
-2
proposed model
Walraven's model
-4
δcr (mm)

Figure 3.12 – Verification of proposed model as well as Walraven model for


concrete with a cube compressive strength of 13.4 MPa.

10 0.1 0.2 0.3 0.4 0.7

6
wcr=1.0 mm
τcr (MPa)

2
f cc =37.6 MPa
0
0 0.5 1 1.5 2
-2
proposed model
Walraven's model
-4

δcr (mm)

Figure 3.13 - Verification of proposed model as well as Walraven model for


concrete with a cube compressive strength of 37.6 MPa.

3-24
0.1 0.2 0.3 0.4 0.7
10

8
wcr=1.0
6
τcr (MPa)

2
f cc =56.1 MPa
0
0 0.5 1 1.5 2
-2
proposed model
Walraven's model
-4
δcr (mm)

Figure 3.14 - Verification of proposed model as well as Walraven model for


concrete with a cube compressive strength of 56.1 MPa.

τ cnt + τ o 1
⋅ (1 mm)
6 δs 0.75
fcc Cube Strength
13.4 MPa
5 37.6 MPa
56.1 MPa
4

3
τ cnt + τ o f cc0 . 75
=
2 δs 0 .1 + 1 .6 w + 3 w 2

1 R2 = 0.97

0
0 0.5 1 1.5 2 2.5 3
wcr (mm)

Figure 3.15 – Verification of aggregate interlock model (Eq. (3.28) ) against the
test results of Walraven (1980).

3-25
The dowel effect parameter used in this study was calculated by undertaking a
regression analysis and is given by

13.5 ρ
αd = 1+ (3.32)
3 f cp

where ρ is the reinforcement ratio perpendicular to the crack interface. For the case of
unreinforced concrete ρ = 0.0 and α d = 1. The adopted coefficient α d is compared
with the Walraven’s test data in Figure 3.16. The data scatter is not unreasonable
considering the simplicity of the model and the limited number of test data found in
literature. Figure 3.17 shows the calculated shear stress versus crack slip using the
proposed model for cracked reinforced concrete with fcp = 40 MPa and ρ = 0.01.

4.0
α = 1 + 13 . 5 ρ 3 f cp
3.5

3.0

2.5
+20%
2.0

1.5 -20%
1.0

0.5

0.0
0.01 0.02 0.03 0.04 0.05 0.06 0.07

ρ 3 f cp

Figure 3.16 - Comparisons of coefficient α with the test data of Walraven (1980).

3-26
30
fcp = 40 MPa
ρ = 0.01
25

wcr = 0.2 mm
20

(MPa)
15 0.5 mm

τ cnt
10 1.0 mm

2.0 mm
5

0
0 1 2 3 4 5
δ s (mm)

Figure 3.17 - Shear stress versus slip given by Eq. 3.31 for fcp = 40 MPa and ρ = 0.01.

For the analysis of RC members made from high strength concrete (HSC), the shear
capacity along the cracks produced from interlocking of the cracks is significantly
reduced compared with that of normal strength concrete (Walraven, 1995, Ali and
White, 1999, Lubell et al., 2004). Since in HSC the matrix can be stronger than the
aggregate particles, cracks can cut the aggregate particles forming a relatively smooth
surface. By test observations, the cylinder compressive strength at which the aggregate
particles fracture depends on the type of aggregate used but can be as low as 70 MPa for
some aggregates (Walraven, 1995, Lubell et al., 2004). In tests undertaken by Walraven
(1995) on cracked unreinforced HSC with a cube compressive strength of 115 MPa and
under a constant crack width condition showed that the shear resistance decreased to
about 35 percent of the value which would be obtained without particle fracture. For
cracked reinforced concrete made from HSC, this reduction is 55 to 75 percent of the
values where cracking occurs around (rather than through) aggregate particles.

To account for this phenomenon, a roughness coefficient μ is incorporated into the


Eq. 3.31 as follows:

3-27
μ α d 1.18 f cp0.75δ s μ α d f cp0.5
τ cnt = ≤ (3.33)
2
0.1 + 1.6wcr + 3wcr 0.31 + 24 wcr (a + 16 )

where

μ = 1.0 for f cp ≤ 70 MPa


(3.34)
μ = 0.35 for f cp > 70 MPa

The predicted values of shear reduction ratio for the test specimens of Walraven (1995)
made from HSC with a cube compressive strength of 115 MPa (equivalent to a cylinder
compressive strength of 100 MPa) are shown in Figure 3.18. The reinforcement ratio
crossing the cracks ranged from 0.0028 to 0.0126. The predicted values of reduction
ratios are between 0.4 and 0.75 with a mean of 0.65 and are in reasonable agreement
with the test results.

0.3 0.6
8

6
wcr=1.0 mm
4
τcr (MPa)

2
f cc =115 MPa
0
0 0.5 1 1.5 2 2.5
-2

-4

δcr (mm)

Figure 3.18 – Verification of proposed model for concrete with a cube compressive
strength of 115 MPa..

3-28
The secant shear stiffness due to aggregate interlock is calculated as

τ s τ
Gcr = cnt = rm cnt (3.35)
γ cnt δs

Since the maximum crack width for the test data used in regression analysis was
1.0 mm, the proposed model may lose its accuracy if the crack width is too large.
However, the crack width of 1.0 mm may be regarded as a practical upper bound in
actual engineering structures (Maekawa et al., 2003). Nevertheless, to apply the model
for the analysis of RC members and structures this restriction should be considered.

3.4 Finite Element Procedure

The accuracy, stability and capacity for convergence of finite element model used to
analyse the reinforced concrete structures depends not only on the applied material
models but also on other important aspects including the finite element formulation and
numerical implementation. To achieve accurate and stable solutions with a good
convergence, an understanding of the aspects of the finite element procedures as well as
the nature of problems being studied is essential. The assumptions made in the
description of finite element model are summarised as follows:

• The response of concrete and reinforcing steel bars as well as their interaction is
modelled separately. These models are then superimposed to obtain the element
response.

• The smeared crack approach is adopted to describe the behaviour of cracked


concrete.

• Cracking in more than one direction is represented by a system of orthogonal


cracks.

3-29
• The crack direction is assumed to be fixed with load history (fixed crack model).

• The shear transfer across crack interfaces is explicitly taken into account using a
developed aggregate interlock model.

• Reinforcing steel bars are assumed to carry stress only along the axis of the bars.

• The reinforcing pars are taken to be perfectly anchored at their ends.

In this section the finite element formulation developed based on CMM is discussed.

3.4.1 Finite Element Formulation

Adopting the smeared-fixed crack approach, the cracked concrete is treated as an


orthotropic material with the fixed 1- and 2- orthotropic axes that are respectively
normal and parallel to the crack direction as shown in Figure 3.19. The crack angle θ r
indicates the fixed angle between the 1-2 coordinate system and the X-Y global
coordinate system.
rm
S

Y
rm

1
S

θr
1

Figure 3.19 – Cracked reinforced concrete element with orthogonal material and global
coordinate systems.

3-30
A modified form of equivalent uniaxial strain model (Darwin and Pecknold, 1974) is
used to obtain the biaxial stress-strain relationships as proposed by Foster and
Marti (2003). That is;

⎧ ε1u ⎫ 1 ⎡ 1 v12 ⎤ ⎧ ε c1 ⎫
⎨ ⎬= ⎢ ⎥⎨ ⎬ (3.36)
⎩ε 2u ⎭ 1 − v12 v21 ⎣v21 1 ⎦ ⎩ε c 2 ⎭

where v12 and v21 are the Poisson’s ratios, ε1 and ε 2 are the strains in the
orthotropic 1-2 coordinate system and ε1u and ε 2u are the equivalent uniaxial
strains in the orthotropic 1-2 coordinate system that would exist in each direction
when the stress is zero in the other.

The biaxial stresses in the orthotropic 1-2 coordinate system, σ 1 and σ 2 , is determined
as

⎧σ c1 ⎫ ⎡ Ec1 0 ⎤ ⎧ ε1u ⎫
⎨ ⎬=⎢ ⎨ ⎬ (3.37)
⎩σ c 2 ⎭ ⎣ 0 Ec 2 ⎥⎦ ⎩ε 2u ⎭

where Ec1 and Ec 2 are the secant moduli in the orthotropic stress 1- and 2- directions
obtained from the appropriate uniaxial stress-strain curve.

Stresses and strains for a cracked concrete element are related in the familiar manner

⎧ σ c1 ⎫ ⎧ ε c1 ⎫
⎪ ⎪ ⎪ ⎪
⎨σ c 2 ⎬ = Dc12 ⎨ ε c 2 ⎬ (3.38)
⎪τ ⎪ ⎪γ ⎪
⎩ c12 ⎭ ⎩ c12 ⎭

where σ c1 , σ c 2 and τ c12 are the stresses in the material 1-2 coordinates; ε c1 , ε c 2 and

γ c12 are strains in the material 1-2 coordinates and Dc12 is the constitutive matrix of the
cracked concrete with respect to the 1-2 coordinate system.

3-31
The cracked concrete is assumed as a material composed of two components, the solid
concrete between cracks and the crack, with the Dc12 obtained based on manipulation
of the material matrices of its components. Assuming the cracks as smeared at an
average spacing srm , the interface displacements over crack surfaces can be replaced
with average strains as follows:

δs
ε cr1 = (3.39a)
srm

ε cr 2 = 0 (3.39b)

wcr
γ cr12 = (3.39c)
srm

where ε cr1 , ε cr 2 and γ cr12 are the average strains due to cracks in the 1- and 2-

directions, δ s is the slip along the crack and wcr is crack width. Since the shear strain
of intact concrete between cracks is negligible in comparison to the shear strain of
crack, it is assumed that γ cr12 ≅ γ c12 .

Then the average stresses over the crack surfaces, σ cr1 , σ cr1 and τ cr12 are determined
as

⎧ σ cr1 ⎫ ⎧ ε cr1 ⎫
⎪ ⎪ ⎪ ⎪
⎨σ cr 2 ⎬ = Dcr12 ⎨ ε cr 2 ⎬ (3.40)
⎪τ ⎪ ⎪γ ⎪
⎩ cr12 ⎭ ⎩ cr12 ⎭

where Dcr12 is the constitutive matrix of crack and given by

⎡ d cr11 0 d cr12 ⎤
Dcr12 = ⎢⎢ 0 0 0 ⎥⎥ (3.41)
⎢⎣d cr 21 0 Gcr12 ⎥⎦

3-32
in which Gcr12 is the secant shear modulus at the crack derived from an appropriate

constitutive relationship which includes the effects of aggregate interlock; d cr11 is the

crack stiffness coefficient in the normal direction on crack and d cr12 and d cr 21 are the
crack stiffness coefficients due to the crack dilatancy. It should be noted that although
the crack response shows a tendency for instability it is usually stabilised by the
restraint provide by the reinforcement and boundary conditions (Bazant and
Gambarova, 1980).

For the solid concrete between cracks the relationship between the average stresses
( σ sc1 , σ sc 2 and τ sc12 ) and the average strains ( ε sc1 , ε sc 2 and γ sc12 ) in the 1-2
coordinate system is given by

⎧ σ sc1 ⎫ ⎧ ε sc1 ⎫
⎪ ⎪ ⎪ ⎪
⎨σ sc 2 ⎬ = D sc12 ⎨ ε sc 2 ⎬ (3.42)
⎪τ ⎪ ⎪γ ⎪
⎩ sc12 ⎭ ⎩ sc12 ⎭

where D sc12 is the constitutive matrix of solid concrete between the cracks in the
1-2 coordinate system and is expressed as

⎡ Ec1 ν 12 Ec1 0 ⎤
1 ⎢ ⎥
D sc12 = ν 21Ec 2 Ec 2 0 ⎥ (3.43)
1 −ν 12ν 21 ⎢
⎢⎣ 0 0 (1 −ν 12ν 21 )Gsc12 ⎥⎦

In Eq. (3.43) Ec1 and Ec 2 are the secant moduli in the orthotropic 1- and 2- directions

determining from the uniaxial constitutive relationships and Gsc12 is the secant shear

modulus of solid concrete in the 1-2 coordinate system. The Gsc12 adopted here is that
derived by Attard et al. (1996) and is given by

(1 −ν 12ν 21 )Gsc12 = 1 [Ec1 (1 −ν 12 ) + Ec 2 (1 −ν 21 )] (3.44)


4

3-33
After cracking it is assumed that there is no transmission of lateral tensile strains across
the cracks. As a result, once cracking occurs and the cracked concrete is treated as an
orthotropic material, ν 21 is considered equal zero. When cracking takes place in the
material 2- direction both ν 21 and ν 12 are taken equal zero.

Ignoring dowel action and considering the cracks as continuous with regular spacing,
the stresses in the solid concrete between the cracks are equal to the stresses on the
cracks (Bazant and Gambarova, 1980). These assumptions also yield that the average
strains of the cracked concrete are the sums of the strains of the solid concrete between
the cracks and of the strains due to the cracks. These conditions can be expressed as

{σ c } = {σ sc } = {σ cr } (3.45)

{ε c } = {ε sc }+ {ε cr } (3.46)

where {σ c } , {σ sc } and {σ cr } are the stress vectors and {ε c }, {ε sc } and {ε cr } are the
strain vectors of the cracked concrete element. Substituting Eqs. (3.38), (3.40) and
(3.42) into the Eq. (3.46) the constitutive matrix of cracked concrete is determined as

−1
Dc12 = Dcr[ −1
12 + D sc12 ] −1
(3.47)

Neglecting the small effects of normal stress and the crack dilatancy
( d cr11 = d cr12 = d cr 21 = d cr 22 = 0 ), Dc12 is determined from the following relationship

⎡ Ec1 ν 12 Ec1 0 ⎤
1 ⎢ν E ⎥
Dc12 = ⎢ 21 c 2 Ec 2 0 ⎥ (3.48)
1 −ν 12 ν 21
⎢⎣ 0 0 (1 −ν 12 ν 21 )Gc12 ⎥⎦

where Gc12 is the secant shear moduli of cracked concrete in the 1-2 coordinate system
given by

3-34
Gsc12 + Gcr12
Gc12 = (3.49)
Gsc12 Gcr12

The constitutive matrix for the cracked concrete, Dc12 , is transformed into the global X-
Y coordinate system by

Dcxy = TεT Dc12 Tε (3.50)

in which Tε is the strain transformation matrix given by

⎡ C2 S2 − CS ⎤
⎢ ⎥
Tε = ⎢ S 2 C2 CS ⎥ (3.51)
⎢2CS − 2CS C2 − S2⎥
⎣ ⎦

where C = cos θ r and S = sin θ r .

Adding the contributions of the reinforcement and the concrete tension stiffening to the
constitutive matrix of cracked concrete, the constitutive matrix for the reinforced
concrete element, D xy , is derived as

D xy = Dcxy + Dcts + D s (3.52)

where Dcts and D s are the concrete tension stiffening component and the reinforcing

steel component, respectively. Combining Dcts and D s together the following equation
can be expressed

⎡ Ectsx + ρ x Esx 0 0⎤
Dcts + D s = ⎢⎢ 0 Ectsy + ρ y Esy 0⎥⎥ (3.53)
⎢⎣ 0 0 0⎥⎦

where Ectsx and Ectsy are the secant moduli due to concrete tension stiffening in X and

Y directions, respectively; Esx and Esy are the secant moduli of reinforcing steel in X

3-35
and Y directions, respectively, and ρ x and ρ y are the reinforcement ratios in X and Y

directions, respectively. Secant moduli are determined from the constitutive


relationships defined for the concrete tension stiffening and reinforcing steel.

Finally, the element stiffness matrix, k , in the X-Y coordinate system is determined as
follows:

k = t ∫ BT D xy B dA (3.54)
A

where t is the element thickness, A is the element area, and B is the strain
displacement matrix.

3.4.2 Implementation

The finite element formulation outlined in the previous section is incorporated into the a
displacement-based finite element program named as RECAP, developed by Foster and
Gilbert (1990) for linear and nonlinear analysis of reinforced concrete structures. The
numerical implementation used in this study including the element type, solution
algorithms, convergence criteria and the finite element program RECAP are described ,
in brief, in Appendix A. A flow chart showing how the element is integrated into the
program is given in Figure 3.20.

3-36
Input data

Element Module
Increase time and external nodal forces Determine element stiffness matrix
t = t + Δ t ; P = k P0
k = ∫ BT D xy B dA
A

Determine element strains


ε = Bu Determine structure stiffness matrix
K = ∑ki

Determine element
secant stiffness moduli Determine element stresses
Ec1 ; Ec 2 ; Gc12 ; Gcr ; Es ; Ects σ = Dε

Determine component Determine structure internal forces


constitutive matrices
Q = ∫ BT σ dA
Dcxy ; D s ; Dcts
A

Determine element constitutive matrix Determine out-of-balance forces


D xy = Dcxy + D s + Dcst R = P−Q

Calculate structure nodal


displacements
ui +1 = ui + Δ ui where Δ ui = K −1R

No
Convergence tolerance
is reached

Yes

Yes Output stresses and


Next displacement or load
strains
step
No

Stop

Figure 3.20 – Solution Procedure Flowchart.

3-37
CHAPTER 4

NUMERICAL EXAMPLES

4.1 Introduction

To verify the finite element model outlined in Chapter 3 at the element level as well as
member level, the results of analyses are compared with experimental data obtained
from a number of tests undertaken on reinforced concrete shear panels, shear walls and
shear-critical beams, the results of which are presented in this chapter. Both loading
conditions, monotonic and cyclic, for shear panels as well as shear walls are considered.
In cyclic tests, the envelope curve, stiffness, strength and the displacement or strain
corresponding to the strength are investigated to test if the finite element model can be
used to calculate the strength and ductility of cyclically loaded RC members;
particularly RC shear walls. For all finite element analyses presented here, the concrete
was modelled using four node isoparametric plane stress elements with numerical
integration using a 2×2 gauss quadrature (refer to Appendix A). The steel reinforcement
was modelled as smeared throughout the concrete element either directly in the RC
formulation or by discrete bar elements placed at the element boundaries.
4.2 Monotonic Conditions

4.2.1 Shear Panels

Meyboom (1987)

Meyboom (1987) tested three panels; PP1, PP2 and PP3, with panels PP2 and PP3
being partially prestressed with unbonded tendons. As the panels were uniformly
stressed over their entirety, they were modelled using a single 4-node finite element of
unit in-plane dimensions. The prestressing steel was modelled using discrete bar
elements. Details of the material properties and geometry are given in Table 4.1 and in
Figure 4.1. For all specimens ν = 0.2 and τb0 / 2 = τb1 = fct . For specimens PP2 and PP3
the applied prestress was 2.08 MPa and 4.39 MPa, respectively.

In Figure 4.2, the results of the finite element modelling are compared with the
experimental data for shear stress versus shear strain, shear stress versus the principal
strain angle and shear stress versus crack width. The peak strengths obtained from the
fixed crack model were 4.78 MPa for panel PP1, 5.55 MPa for panel PP2 and 5.13 MPa
for panel PP3. Also plotted are the results for the rotating crack model as reported by
Foster and Marti (2003). The results show a slight improvement in the match against the
experimental data for the fixed crack model, compared to the rotating crack model, with
the most significant change being for panel PP2.

Vecchio and Collins (1982)

Vecchio and Collins undertook an extensive testing programme for panels stressed
under bi-axial conditions with the main parameter being varying quantities and
strengths of reinforcement in each of two orthogonal directions. Two panels PV19 and
PV20 tested under pure shear have been chosen as they failed by crushing of the
concrete after yielding of the reinforcement in one direction while, in the second
direction (the X-direction), the reinforcement remained elastic before the ultimate load

4-2
V/2 V/2

V/2 Panel f cp (MPa) f ct (MPa) Ec (GPa) ε cp

PP1 27.0 2.7 28.0 0.0021


reinforced concrete PP2 28.1 2.8 28.0 0.0024
t = 287 mm element
PP3 27.7 2.8 28.0 0.0019

V/2
unbonded prestressed V/2
steel element
Figure 4 1 - FE mesh for modelling of Meyboom (1987) panels PP1, PP2 and PP3 and concrete properties.

Table 4.1 - Reinforcing steel properties for panels PP1, PP2 and PP3. (Meyboom, 1987)

Panel Location ρ ∅ (mm) E s (GPa) E w (GPa) Eu (GPa) ε sy εw εu

PP1 X steel 0.0194 19.5 200 0.0 2.47 0.0024 0.014 0.091
PP1 Y steel 0.0065 11.3 192 0.0 3.90 0.0025 0.050 0.091
PP2 X steel 0.0130 16.0 202 0.0 2.88 0.0024 0.050 0.100
PP2 Y steel 0.0065 11.3 192 0.0 3.90 0.0025 0.050 0.091
PP2 Prestress 0.0029 15.2 202 0.0 15.7 0.0045 0.011 0.025
PP3 X steel 0.0065 11.3 192 0.0 3.90 0.0025 0.050 0.091
PP3 Y steel 0.0065 11.3 192 0.0 3.90 0.0025 0.050 0.091
PP3 Prestress 0.0059 15.2 202 0.0 15.7 0.0045 0.011 0.025

4-3
Panel Shear Stress v Shear Strain Shear Stress v tanθ r Shear Stress v Crack Width
6 6 6
Yield in Y steel
Onset of
5 Crushing 5 5

4 4 4
τxy (MPa)

τxy (MPa)

τxy (MPa)
3 3 3
Srm = 222 mm
PP1 l = 1.0
2 2 Experimental 2
Experimental CMM: RCM Experimental
CMM: RCM CMM: FCM CMM: RCM
1 1 1
CMM: FCM CMM: FCM

0 0
0 0.01 0.02 0.03 1 1.2 1.4 1.6 1.8 0 1 2 3
γxy tan (θ ε ) Maximum w cr (mm)

6 6 6
Onset of
Crushing
5 5 5

4 Yield in 4 4
Experimental
τxy (MPa)

τxy (MPa)

τxy (MPa)
Prestress. CMM: RCM
3 3 CMM: FCM 3 Srm = 260 mm
Yield in λ = 1.0
PP2 Y steel
2 2 2
Experimental Experimental
1 CMM: RCM 1 1 CMM: RCM
CMM: FCM CMM: FCM

0 0
0 0.01 0.02 0.03 1 1.5 2 2.5 3 0 1 2 3 4
γxy tan (θ ε ) wc (mm)

6 6 6
Experimental
5 5 CMM: RCM 5
CMM: FCM
4 Yield in 4 4
τxy (MPa)

τxy (MPa)
τxy (MPa)

Prestress.
3 3 3
Yield in Onset of Srm = 314 mm
PP3 Crushing
2
Y steel
2 2 λ = 1.0
Experimental Experimental
1 CMM: RCM 1 1 CMM: FCM
CMM: FCM CMM: RCM
0 0
0 0.005 0.01 0.015 0.02 0.025 1 1.5 2 2.5 3 0 1 2 3 4 5 6
γxy tan (θ ε ) wc (mm)

Figure 4.2 - Comparison of FE results against experimental data for shear stress versus
shear strain, strain angle at crack and crack widths for Meyboom (1987)
panels PP1, PP2 and PP3.

4-4
was reached. As such, significant rotations in the principal strain directions were
observed before failure and the panels are some of the more challenging to model and
one where significant differences between a rotating crack model and a fixed crack
model might be expected.

As the panels were uniformly stressed over their entirety, they were modelled using a
single 4 node finite element of unit in-plane dimensions. Details of the geometry of the
panels, the material properties and the results of the analyses are presented in
Figures 4.3 and 4.4.

For panels PV19 and PV20 the peak calculated shear stresses obtained by the rotating
crack model of Foster and Marti (2003) were 3.34 and 3.87 MPa, respectively,
compared the experimental values of 4.0 MPa and 4.3 MPa, respectively. For the fixed
crack model, the peak shear stresses calculated for PV19 and PV20 were 4.20 and 4.51
MPa, respectively (26 and 16 percent, respectively, higher those that obtained by the
rotating crack model). For both models, the elements failed by crushing of the concrete
after yielding of the Y-direction reinforcing steel but before yielding of the X-direction
steel, as observed in the experiment. As is expected, both the rotating and fixed crack
models provide identical results to the point of yielding of the Y-steel after which the
solutions diverge as the friction along the crack becomes significant.

4-5
x-direction steel: 5
890 ρ = 0.0179 ∅ = 6.35 mm
Es = 200 GPa Ew = 0 GPa
Eu = 3.5 GPa εsy = 0.0023 4

Shear Stress (MP


y εw = 0.055 εu = 0.135
y-direction steel: 3
ρ = 0.0071 ∅ = 4.01 mm
890

x Es = 200 GPa Ew = 0 GPa


t = 70 mm Eu = 2.0 GPa εsy = 0.0015 2
εw = 0.055 εu = 0.135 Panel PV19
concrete: Fixed Crack Model
fcp = 19.0 MPa fct = 1.5 MPa 1
Rotating Crack Model
Ec = 17.3 GPa εcp = 0.0022
Experiment
0
0 0.01 0.02 0.03
Shear Strain

Figure 4.3 - Panel PV19 (Vecchio and Collins, 1982): Dimensions, reinforcement details and shear stress versus shear strain.

4-6
x-direction steel: 5
890 ρ = 0.0179 ∅ = 6.35
mm
Es = 200 GPa Ew = 0 GPa 4

Shear Stress (MP


y Eu = 3.5 GPa εsy = 0.0023
εw = 0.055 εu = 0.135
y-direction steel:
3
890

x ρ = 0.0089 ∅ = 4.01
mm 2
t = 70 mm Es = 200 GPa Ew = 0 GPa
Panel PV20
Eu = 2.0 GPa εsy = 0.0015
εw = 0.055 εu = 0.135 1 Fixed Crack Model
concrete: Rotating Crack Model
Experiment
fcp = 19.6 MPa fct=1.5MPa
Ec = 21.8 GPa εcp = 0.0018 0
0 0.01 0.02 0.03
Shear Strain

Figure 4.4 - Panel PV20 (Vecchio and Collins, 1982): Dimensions, reinforcement details and shear stress versus shear strain.

4-7
Zhang and Hsu (1998)

To verify the finite element model against the high-strength concrete shear panels, Two
panels VB2 and VB4 of Zhang and Hsu (1998) tested under pure shear have been
selected as they have different reinforcement ratios in the x and y directions and are
more challenging to model.

As the panels were uniformly stressed over their entirety, they were modelled using a
single 4 node finite element of unit in-plane dimensions. Details of the geometry of the
panels, the material properties and the results of the analyses are presented in
Figures 4.5 and 4.6.

The peak calculated shear stresses of then panels VB2 and VB4 were 8.32 and
4.65 MPa, respectively, compared the experimental values of 9.14 and 4.86 MPa,
respectively, and show a good overall correlation between test data and calculated
results. For both analyses, failure was by crushing of the concrete after yielding of the
Y-direction reinforcing steel but before yielding of the X-direction steel, as observed in
the experiment.

4-8
x-direction steel concrete
1397 mm
ρ = 0.0359 ∅ = 19.5 mm fcp = 97.6 MPa
Es = 200 GPa Ew = 0 GPa fct = 3.3 MPa
Eu = 3.4 GPa εsy = 0.0023 εcp = 0.00245
εw = 0.010 εu = 0.050 Ec = 39.5 GPa
y-direction steel

1397 mm
y

x
ρ = 0.0120 ∅ = 11.3 mm
t = 178 mm Es = 200 GPa Ew = 0 GPa
Eu = 3.2 GPa εsy = 0.0022
εw = 0.010 εu = 0.050

12
Shear Stress (MP

10
8
6
VB2
4
FEA
2 Experiment
0
0 0.005 0.01 0.015 0.02
Shear Strain

Figure 4.5 - Panel VB2 (Zhang and Hsu, 1998): Dimensions, reinforcement details and
shear stress versus shear strain.

x-direction steel concrete


1397 mm
ρ = 0.0180 ∅ = 19.5 mm fcp = 96.9 MPa
Es = 200 GPa Ew = 0 GPa fct = 3.3 MPa
Eu = 3.4 GPa εsy = 0.0023 εcp = 0.00230
εw = 0.010 εu = 0.050 Ec = 39.5 GPa
y-direction steel
1397 mm
y

ρ = 0.0060 ∅ = 11.3 mm
t = 178 mm Es = 200 GPa Ew = 0 GPa
Eu = 3.2 GPa εsy = 0.0022
εw = 0.010 εu = 0.050

6
Shear Stress (MP

5
4
3
VB4
2
FEA
1 Experiment
0
0 0.005 0.01 0.015 0.02 0.025
Shear Strain

Figure 4.6 – Panel VB4 (Zhang and Hsu, 1998) : Dimensions, reinforcement details and
shear stress versus shear strain.

4-9
4.2.2 Shear Walls

Lefas et al. (1990)

To verify the proposed finite element model in the member level it is applied to the two
RC shear walls SW14 and SW24 of Lefas et al. (1990) with the height-to-length ratios
of 1 and 2, respectively. The geometry and reinforcement details of walls have been
briefly described into section 2.4.4. The cube concrete strength for the SW14 and SW24
were 42.1 and 48.3 MPa, respectively, and their reinforcement ratios were 1.1 and
0.8 percent in longitudinal direction, respectively, and 2.4 and 2.5 percent in transverse
direction, respectively. The steel bars used in longitudinal and transverse directions
were of 8 and 6.25 mm diameter, respectively, with the yield stress of 470 and 520
MPa, respectively. Additional horizontal reinforcement in the form of stirrups confined
the wall edges. The walls with no axial load were monotically loaded in the lateral
direction. Figure 4.7 shows the finite element mesh used for analysing the walls.

The results of finite element analysis versus test data for the wall SW14 and SW24 are
illustrated in Figures 4.8 and 4.9, respectively. The maximum load measured in test for
walls SW14 and SW24 were 266 kN and 120 kN, respectively; while the predicted
values are 242 kN and 107 kN (9 percent and 11 percent lower than the test data,
respectively). The computed displacements at the peak load were 9.1 mm for wall
SW14 and 19.5 mm for wall SW24 and compare reasonably with the measured data of
11.7 and 19.3 mm, respectively. For these walls, the model indicates a generally stiffer
response than the measured values under initial loading and service conditions. This is
possibly due to movements in the testing apparatus not considered in the perfect
boundary conditions implied in the numerical modelling. For both walls, the failure was
in a flexural mode with yielding of the longitudinal reinforcement, as observed in the
experiments.

Generally, before cracking the stiffness of simulated responses are higher than observed
ones, but after cracking it becomes closer to observed amount. It can be concluded that
the finite element model yields a stiffer response for lower height-to-length ratios.
Considering inconsistency and uncertainty in material properties and boundary
conditions, the predicted responses are reasonable.

4-10
Figure 4.7 – Finite Element mesh for the Lefas et al. (1990) walls: (a) SW14 and (b) SW24.
4-11
300

250

200
Load (kN

150

100
SW14
50 FEA
Experiment
0
0 2 4 6 8 10 12 14
Displacement (mm)

Figure 4.8 – Comparison of the results of finite element analysis and the test data of
Lefas et al. (1990) wall SW14.

140

120

100
Load (kN

80

60
SW24
40
FEA
20
Experiment
0
0 5 10 15 20 25
Displacement (mm)

Figure 4.9 - Comparison of the results of finite element analysis and the test data of
Lefas et al. (1990) wall SW24.

4-12
4.2.3 Shear-Critical Reinforced Concrete Beams

Bresler and Scordelis (1963)

Although the Bresler and Scordelis (1963) beams have been modelled by many with
good accuracy between numerical and experimental results (for example Vecchio
(1989), and Foster and Gilbert (1990)). Vecchio (2000a) highlighted a significant flaw
in previous methodologies where tension stiffening and tension softening were typically
incorporated using a single model. In the case of Vecchio (2000a) for concrete members
with low quantities of reinforcement, a check of stresses along a crack determined that
slip at the crack would occur at loads significantly below those of the experiment unless
a significant residual tensile stress was carried across the cracks. Vecchio later
developed the disturbed stress field model to account for the identified weakness
(Vecchio, 2000b).

As for Vecchio (2000a), in the CMM:FE rotating crack implementation of Foster and
Marti (2003) it was recognised that lacking from the implementation was the allowance
for shear transfer parallel to the crack faces via aggregate interlock and dowel action
leading to an under-estimation of the capacity of lightly reinforced members. Examples
of beams with low volumes of web reinforcement are beams OA1, OA2 and OA3 of
Bresler and Scordelis (1963). The beams were modelled using the rotating and fixed
crack implementations of the CMM:FE model using 4-node membrane elements. The
beam dimensions and details are presented in Figure 4.10 and the FE meshing used for
the analyses is shown in Figure 4.11. Also Figure 4.12 shows the finite element results
versus test data for beams OA1, OA2 and OA3. Both the fixed and rotating versions of
the model compare well with the load deflection response to the point of failure. The
rotating crack model, however, consistently failed at loads lower than those observed in
the experiment. Conversely, the failure loads calculated by the fixed crack
implementation compare well with the test results. The increase in failure load for the
fixed crack model over that of the rotating crack model was between 21 and 38 percent.
In all cases, for both the rotating and fixed models, failure was by beam shear, as
observed in the experiments. For the fixed crack model, the response beyond failure for
these tests was difficult to capture using conventional arc length or displacement control
due to bifurcations caused by snap-back leading to instabilities in the solution routine.

4-13
P
b = 305 mm

465
560

0A1 ... 4#9 0A1 ... L = 3655 mm


0A2 ... 5#9 0A2 ... L = 4570 mm
0A3 ... 6#9 L 0A3 ... L = 6400 mm

concrete:
OA1: OA2: OA3:
fcp = 22.6 MPa fct = 1.9 MPa fcp = 24.5 MPa fct = 2.0 MPa fcp = 37.6 MPa fct = 2.5 MPa
Ec = 20.0 GPa εcp = 0.0020 Ec = 17.3 GPa εcp = 0.0020 Ec = 32.0 GPa εcp = 0.0025

steel:
fy = 556 MPa
Es = 206 GPa

Figure 4.10 - Bresler and Scordelis (1963) beams OA1, OA2 and OA3.

4-14
ρ x = 0.0 P/2 L
C
ρ y = 0.0
All load and reaction
plates 25 mm thick
418 steel plate

143 ρ x = 0.089
ρ y = 0.0
158 125 1613 150

Beam OA1 P/2 L


C

symmetry about
418 centreline

143 ρ x = 0.111
ρ y = 0.0
158 125 2070 150

Beam OA2 ρ x = 0.134 P/2 L


C

ρ y = 0.0

418

143

158 125 2984 150

Beam OA3

Figure 4.11 - FE mesh for Bresler and Scordelis (1963) beams.

4-15
400 0A1 0A2 0A3

300

Load (kN)
200

CMM:FE - Fixed
100 CMM:FE - Rotating
Experiment

0
0 5 10 15 20 25 30 35
Deflection (mm)

Figure 4.12 – Load-deflection response for Bresler and Scordelis beams.

Leonhardt and Walther (1966)

Leonhardt and Walther (1966) tested nine deep beams to study non-flexural behaviour
of reinforced concrete members. In this study, the two span deep-beam DWT2 (shown
in Figure 4.13) is analysed using the finite element mesh shown with one half the beam
modelled with symmetry.

The results of the analysis of beam DWT2 are plotted in Figure 4.14 for the load versus
midspan displacement and it is seen that the rotating CMM:FE formulation (without
residual concrete tension strength after cracking) under-calculates both the stiffness and
the failure load of the member after cracking. The model failure load was 985 kN, 80
percent of the experimental failure load of 1230 kN and is consistent with the plastic
collapse load of 953 kN. An inspection of the displaced shape at failure revealed that
after yielding of the reinforcement a shear slip occurs in the elements adjacent to the
central support thickening. For the rotating crack model, however, no load can be
transmitted to the support via shear along the cracked surface. In the fixed CMM:FE
formulation the stiffness results are more consistent with the measured data with a
portion of the load being transferred through aggregate interlock and dowel action. The
failure load was 1090 kN (89 percent of the reported failure load).

4-16
concrete:
R = 0.094P 480 480 480 360
fcp = 27.1 MPa fct = 2.5 MPa
εcp = 0.0020 S = 0.250P
Ec = 27.1 GPa
T = 0.156P
R S T T S R L
C A
P/2 P/2
steel:
Es = 206 GPa Ew = 5.0 GPa
εsy = 0.002 εsw = 0.05 Midspan
3 support
region 1: thickening
ρx = 0.0195 ρy = 0.0020 1∅6
1490
∅x = 8.0 mm ∅y = 5.0 mm 1∅6
1∅6
region 2: 1∅6
ρx = 0.0195 ρy = 0.0020 2∅6
∅x = 6.0 mm ∅y = 5.0 mm
110
region 3: 2 4∅8 100
120 1 A
ρx = 0.0019 ρy = 0.0020 Section A-A
∅x = 5.0 mm ∅y = 5.0 mm
2∅5 e/w
160 1280 160

Figure 4.13 - Details of Leonhardt and Walther (1966) beam DWT2 and FE mesh.
4-17
1400
Exp. Failure Load = 1230 kN
1200

1000 953 kN

Load (kN) 800

600

400
Experimental
200 CMM:FE - Fixed
CMM:FE - Rotating
0
0 0.5 1 1.5 2
Midspan Displacement (mm)

Figure 4.14 - Load versus midspan displacement for beam DWT2.

4.3 Cyclic Loading

4.3.1 Shear Panels

Stevens et al. (1987) tested three large-scale panels SE8, SE9 and SE10 under reversed
cyclic condition. The geometry dimensions of the test specimens are presented in
Figure 4.15. The material properties of concrete and reinforcing steel used in the panels
are given in Table 4.2. Since the same stress state appears over the entire region in an
element, only one element is used to analyse the cyclic response of panels. While the
SE8 and SE9 are under cyclic pure shear, the SW10 is subjected to cyclic shear stress
together with proportional biaxial normal stresses. Panels SE8 and SE10 have a large
difference in the reinforcement ratios in the two directions, causing the shear failure of
concrete after the yielding of one reinforcement layer.

4-18
Figure 4.15 – Panels dimensions and application of stresses.

Table 4.2 – Properties of shear panels tested by Stevens et al. (1987)

Specimen SW8 SW9 SW10

Concrete
f c' (MPa) 37.0 44.2 34.0
ε co 0.0026 0.0027 0.0022
Ec (MPa) 30000 33000 29000
Reinforcement in X-Direction
Ø (mm) 20 20 20
f yx (MPa) 492 422 422
ρ sx 0.0293 0.0293 0.0293
Reinforcement in Y-Direction
Ø (mm) 10 20 10
f yy (MPa) 479 422 479
ρ sy 0.0098 0.0293 0.098
Loading
τ xy cycled cycled cycled
fx = f y 0 0 − τ xy 3

4-19
Results of the analyses are compared with the test data in Figures 4.16 to 4.18. The
cyclic envelope curve for Panel SE8 gave a shear stress at ultimate of 6.0 MPa,
compared with the numerical calculation of 7.0 MPa with, in each case, significant
yielding of the reinforcing steel (Figure 4.16). Since the analysis is based on the
monotonic loading, the reduction of resistance of the cyclically loaded test specimen is
attributed to micro damage resulted from the cyclic loading. Thus, in this case the
analysis leads to a 17 percent over-calculation of the peak load. Similarly, the calculated
ultimate shear stresses of the panels SE9 and SE10 were 11.3 and 9.5 MPa,
respectively, giving a 17 and 14 percent over calculation of the capacity, respectively.

The maximum shear strain just prior to complete failure of panels the SE8, SE9 and
SE10 were calculated as 0.0120, 0.0089 and 0.0092, respectively, and showed a very
good agreement with measured data of 0.0116, 0.0089 and 0.0092, respectively. The
panel stiffness compares well with the test data. The failure modes were also predicted
to be the same as test observation. It can be concluded failure mode, the maximum shear
strain and the stiffness were calculated well by the proposed finite element procedure
but the ultimate load of the monotonic analysis is about 17 percent greater than that for
the cyclic test.

12
Panel SE8
Negative Direction
10 Positive Direction
Analysis

8
Shear Stress (MP

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Shear Strain

Figure 4.16 - Comparison of FE results against experimental envelope curve for shear
stress versus shear strain for Stevens (1987) panel SE8.

4-20
12

10

8
Shear Stress (MP

4
Panel SE9
Negative Direction
2
Positive Direction
Analysis
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Shear Strain

Figure 4.17 - Comparison of FE results against experimental envelope curve for shear
stress versus shear strain for Stevens (1987) panel SE9.

12

10

8
Shear Stress P

4
Panel SE10

2 Negative Direction
Positive Direction
Analysis
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Shear Strain

Figure 4.18 - Comparison of FE results against experimental envelope curve for shear
stress versus shear strain for Stevens (1987) panel SE10.

4-21
4.3.2 Shear Walls

Oesterle et al. (1976)

The finite element model is also verified versus to the shear walls B1 and B2 tested
under reversed cyclic loading at the Portland Cement Association (PCA)
(Oesterle et al. ,1976). These walls have been used by other researchers to verify
proposed finite element models (Sittipunt and Wood, 1993, Elmorsi et al., 1998,
Vecchio, 1999, Kwak and Kim, 2004). The wall dimensions have been described in
Section 2.4.2. The material properties and the reinforcement ratio are given in Table 4.3
and the finite element mesh is illustrated in Figure 4.19.

Table 4.3 – Material properties for the walls B1 and B2 of PCA


(Oesterle et al., 1976)

Shear Wall B1 B2

Concrete
f c' (MPa) 53 53.7
Ec (MPa) 32000 32700
Reinforcement in X-Direction
f yx (MPa) 521 533
ρ sx 0.0031 0.0063
Reinforcement in Y-Direction
f yy (MPa) 521 533
ρ sy 0.0029 0.0029
Reinforcement in boundary element
f y (MPa) 450 410
ρ 0.0111 0.0367

4-22
Figure 4.19 – Finite element mesh for the PCA shear walls B1 and B2.

4-23
The nodes at the base of wall are considered completely restricted against horizontal
and vertical translations. Top slab is assumed to be very stiff and distributes the lateral
load to the entire cross section. Figures 4.20 and 4.21 show the results of finite element
analysis compared with the experimental data for the envelope curve of load-
displacement response for the walls B1 and B2, respectively.

The finite element analysis gives a reasonable overall prediction of the behaviour for
both walls with an excellent correlation observed for wall B1. The ultimate strength of
the B1 and B2 are calculated as 288 kN and 695 kN, respectively, compared to the
measured data of 285 kN and 724 kN, respectively. The failure displacement of the B1
is also predicted well; for B2 is the displacement at failure if underestimated by 19
percent (102 mm compared to 126 mm for the test) but, in general, the overall response
is not unreasonable. The experimental and numerical failure mode for the both walls is
flexural with yielding of the longitudinal reinforcement before failure. Also the
calculated stiffness after cracking is in good agreement with the experimental data.

300

200
Load (kN

100 PCA-B1
Negative Direction
Positive Direction
Analysis
0
0 20 40 60 80 100 120
Top Deflection (mm)

Figure 4.20 - Comparison of FE results against experimental envelope curve for load
versus deflection for Oesterle et al. (1976) wall B1.

4-24
800

600
Load (kN

400

Wall B2
200
Negative Dir.
Positive Dir.
Analysis
0
0 20 40 60 80 100 120 140
Top Deflection (mm)

Figure 4.21 - Comparison of FE results against experimental envelope curve for load
versus deflection for Oesterle (1976) wall B2.

4.4 Conclusions

In this chapter the finite element model developed in Chapter 3 for the analysis of 2D
plane-stress reinforced concrete members was tested against a number of experimental
tests at the element level as well as member level. The main propose was to test the
predicted strength and displacement against the experimental data so that the model can
be used to analyse the response of membrane structures in shear. A good correlation
was observed between the numerical and experimental data for the fixed crack finite
element presented. Furthermore, the response of shear-critical members that was
dominated by the behaviour of concrete component in member was calculated well. In
the case of shear walls with a stiffer edge ends, the model revealed a underestimated
displacement. The comparison of proposed fixed crack model with the rotating crack
finite element model developed by Foster and Marti (2003) shows that as expected, the
fixed crack model results in a stiffer response. For the numerical examples, analysed by
both models, up to a 26 percent increase in strength was observed for the fixed crack
approach over that of the more conservative rotating crack model. Also, it leads to a

4-25
little improvement in the calculated results specially for those RC members with the
different amounts of reinforcement ratios in orthogonal X- and Y- directions.

The fixed crack model presented in this paper represents a complete approach to the
analysis of reinforced concrete membranes subjected to monotonic loading with
compression softening, bond and tension stiffening, tension softening, aggregate
interlock and dowel effects all included into the formulation. A fully rational approach
is used in the development of the finite element model with empirical fits for the various
constitutive sub-models.

Comparisons of the monotonic model developed in this study with the results of cyclic
envelope curves for shear panels tested by Stevens (1987), indicate that the monotonic
model over-predicts the failure load by an order of 15 to 20 percent. This needs
consideration when using monotonic modelling techniques to predict the behaviour of
members that are subjected to cyclical damage.

4-26
CHAPTER 5

EXPERIMENTAL PROGRAM

5.1 Introduction

The experimental work conducted on six high strength concrete, squat, shear walls
under reversed cyclic loading is presented in this chapter. Squat shear walls refer to
those shear walls with a ratio of height-to-width less than 2 or 3 (Paulay and Priestley,
1992) that find wide application as seismic load resisting systems in low-rise buildings.
They may also make a major contribution to the lateral load resistance over the first few
stories above foundation level in high-rise buildings. In the following, the test
parameters, details of test specimens, the material properties, the manufacture process,
test setup, instrumentation and the test procedure used for testing are detailed. (Foster
and Rangan, 1999)

5.2 Test Parameters

The test parameters included in the experimental program were the ratio of longitudinal
reinforcement, the ratio of transverse reinforcement and the axial load. The previous
experiments conducted on the normal strength concrete shear walls (Barda et al., 1977,
Oesterle et al., 1978, 1984, Maier and Thurlimann, 1985, Lefas, 1988) showed that
while the increase in longitudinal reinforcement ratio as well as axial load increased the
ultimate strength of shear walls, the transverse reinforcement did not have a
considerable effect. On the other hand, the ductility of shear walls was increased with
either an increase in the transverse reinforcement ratio or a decrease in the axial load,
whereas, the effect of longitudinal reinforcement ratio on ductility was negligible. These
conclusions were also evidenced for HSC shear walls under monotonic loading by
Gupta and Rangan (1996, 1998) and Farvashany (2004). Nevertheless, it is prudent to
investigate all parameters on strength and ductility when testing walls subjected to
cyclic loading.

5.3 Test Specimens

To investigate the effects of test parameters on the cyclic behaviour of HSC shear walls,
six specimens were tested identified as SW1, SW2, SW3, SW4, SW5 and SW6. The test
specimens used in this study were nominally identical in geometry to those specimens
tested by Gupta and Rangan (1996) so that a comparison between cyclic testing and
monotonic testing can be investigated. Four specimens SW1, SW2, SW3 and SW6 were
constructed to study the effect of reinforcement ratio on ultimate transverse load. To
investigate the effects of axial load, specimen SW4 had no axial load. Finally, specimen
SW5 contained different amount of reinforcement in both the transverse and
longitudinal directions and is used to study the effects of transverse as well as
longitudinal reinforcement ratio on the overall response of specimens. All specimens
were designed to fail in shear.

5.3.1 Dimensions

The scale of test specimens was approximately one-third of shear walls used in a
multistorey building. The specimens were made up the web wall, edge elements and
stiff top and bottom slabs. The height-to-width ratio of the walls was equal to 1, as
shown in Figure 5.1. The geometrical dimensions of all specimens were identical. The

5-2
top slab ( 1300 × 575 × 200 mm ) was designed to be sufficiently stiff to distribute the
lateral and axial loads on the test walls. The bottom slab ( 1800 × 575 × 400 mm ) was
also stiff and clamped to the test setup representing a rigid foundation. To simulate
columns or cross-walls that may exist at the ends of a wall in a multistorey building, the
shear wall specimens were constructed with edge elements (1000 × 375 × 100 mm ). The
clear dimensions of the web wall were 1000 mm in height, 800 mm wide and 75 mm in
thickness.

200 mm
1300 mm
A

Top Slab

150 mm Edge Element


1000 mm

Wall 75 mm
1600 mm

B B
400 mm

Bottom Slab

1800 mm A 575 mm
100 mm

Section A-A
375 mm

100 mm

400 mm 1000 mm 400 mm


100 mm

Section B-B

Figure 5.1 – Dimensions of wall specimens.

5-3
5.3.2 Reinforcement Layout

The reinforcement arrangements for the specimens are shown in Figures 5.2 and 5.3.
For all specimens the reinforcement details for the top slab, bottom slab and edge
elements were the same. The web walls, however, contained varying ratios of
longitudinal and transverse reinforcement, as shown in Table 5.1. Except for the closed
ties in the edge elements that were built using the plain round bars (R bars), deformed
N-grade (normal ductility) bars or wire (W) with a nominal yield stress of 500 MPa
were used in the manufacture of all specimens. For stress development purposes, in
accordance with the Australian Standard (AS 3600, 2001), the main reinforcing bars
used in the specimens were provided with 90-degree hooks at the ends of the bars. For
the stirrups and closed ties 135-degree hooks were used. The longitudinal reinforcement
in the wall and edge elements was taken well into the top and bottom slabs. A clear
cover of 20 mm for the reinforcement in top and bottom slabs and 10 mm for the
reinforcement in wall and edge elements was provided.

The web walls were reinforced using two parallel layers of steel bars of 6 and 8 mm.
The bars in the wall were spaced such that an equivalent ratio of transverse
reinforcement as well as longitudinal reinforcement varied from, approximately, 0.5 to
1.3 percent (Table 5.1). The top slab consisted of two layers of steel bars of 16 mm
diameter and stirrups of 12 mm diameter at a spacing 40 mm at the edges and 100 mm
spacing toward the centre. The bottom slab was reinforced using the main reinforcement
of 16 mm diameter in three top layers and reinforcement of 24 mm diameter in two
bottom layers. The stirrups were 12 mm diameter bars at a spacing of 80 mm at edges
and 100 mm toward the centre. Since all specimens were designed to fail in a shear
failure mode, the edge elements contained an excess of longitudinal reinforcement to
carry flexurally generalled forces. The reinforcement was calculated based on the
monotonic analysis of specimens using the nonlinear finite element program RECAP
(refer to Appendix B). Applying a trial-and-error procedure, the reinforcement was
adjusted until the calculated transverse load corresponding to flexural failure mode was
approximately 30 percent larger than the load corresponding to shear failure mode. The
longitudinal reinforcement was enclosed by closed ties made of 4 mm plain bars placed
at a centre-to-centre spacing of 100 mm.

5-4
200 mm
1300 mm
A

2 x 3N12 @75 mm 2 x 6N16


2 x 6N16 2 x 3N12 @ 100 mm
N12
R4 @ 100 mm R4 @ 100 mm
Dv
2 x 6N16 2 x 6N16

1000 mm
1600 mm

B B
Dv Dh

Dh
3 x 4N16 N 12

3 x 4N16

400 mm
2 x 5N12 @ 80 mm
2 x 3N12 @ 100 mm
2 x 4N24
2 x 4N24
1800 mm A 575 mm

2 x 6N16 Dv Dh 2 x 6N16 Section A-A


575 mm

2 x R4 @ 100 mm

400 mm 1000 mm 400 mm

Section B-B

Figure 5. 2 – Reinforcement details of shear wall specimens.

5-5
Table 5.1 – Reinforcement details of web walls
Transverse Reinforcement Longitudinal Reinforcement
Equivalent Equivalent
Specimens Dh
Reinforcement
Dv
Reinforcement
ratio ratio
(refer Figure 5.2) (refer Figure 5.2)
(percent) (percent)

SW1 2 × 6 W6 @ 167 mm 0.5 2 × 5 W6 @ 160 mm 0.5

SW2 2 × 10 W8 @ 100 mm 1.3 2 × 8 W8 @ 100 mm 1.3

SW3 2 × 10 W6 @ 100 mm 0.8 2 × 5 W8 @ 160 mm 0.8

SW4 2 × 10 W6 @ 100 mm 0.8 2 × 5 W8 @ 160 mm 0.8

SW5 2 × 6 W6 @ 167 mm 0.5 2 × 8 W8 @ 100 mm 1.3

SW6 2 × 7 W8 @ 143 mm 1.0 2 × 6 W8 @ 133 mm 1.0

Figure 5.3 – Reinforcement cage for the specimen SW6.

5-6
5.3.3 Material Properties

Concrete

The concrete mixes used in the test specimens were cast by local commercial ready-mix
concrete producers. A concrete mix giving a nominal 28-day compressive strength of
80 MPa and with a maximum aggregate size of 10 mm was used for the mixes. The
concrete mix I was a self-compacting HSC concrete but as it did not satisfy the expected
physical aspects, the other two mixes were provided as the conventional HSC with a
slump of 100 mm.

For each concrete mix, eighteen 100 × 200 mm cylinders were cast to measure the
compressive strength at regular time intervals. As well, twelve 150 × 300 mm cylinders
were cast to determine the splitting tensile strength and elastic modulus and
compressive strength of the concrete. For each concrete mix, the compressive strength
was measured at 7 days and 28 days after casting and at the time of the test while the
splitting tensile strength and elastic modulus were measured on the 28th day and the day
of the wall test. The test results are given in Tables 5.2 to 5.5.

Figures 5.4 to 5.6 give typical concrete stress-strain responses for each concrete mix.
These stress-strain curves have been compared with the relationship of
Thorenfeldt et al. (1987) for the high strength concrete with reasonable agreement
between the test results and the model.

Table 5.2 – Density of concrete


I II III
Concrete Mix
(SW1 & SW2) (SW3 & SW4) (SW5 & SW6)
Average Density
(
kg/m 3 ) 2304 2469 2401

5-7
Table 5.3 – Compressive strength of concrete (100 mm dia. × 200 mm high cylinders)
Samples Average
Concrete Age
a b c Strength
Mix (day)
(MPa) (MPa) (MPa) (MPa)
7 54.2 55.6 - 55
I
(SW1 & SW2) 28 70.0 66.0 66.5 68
382 87.7 85.2 85.0 86
7 69.1 69.7 67.9 69
II
(SW3 & SW4) 28 80.6 81.9 80.8 81
355 96.0 97.2 95.3 96
7 65.4 63.3 - 64
III 28 76.8 76.2 76.5 77
(SW5 & SW6) 36 82.9 82.9 - 83
38 84.3 83.7 82.6 84

Table 5.4 –Splitting tensile strength of concrete (150 mm dia. × 300 mm high cylinders)
Samples Average
Concrete Age
a b c Strength
Mix (day)
(MPa) (MPa) (MPa) (MPa)
I 28 4.4 4.6 5.6 4.9
(SW1 & SW2) 382 5.4 5.7 6.8 6.0
II 28 4.9 4.7 4.2 4.6
(SW3 & SW4) 355 6.9 8.3 6.3 7.2
III 38 5.5 5.4 5.3 5.4
(SW5 & SW6)

Table 5.5 – Elastic modulus of concrete (150 mm dia. × 300 mm high cylinders)
Samples Average
Concrete Age Elastic
Mix (day) a b c
Modulus
(MPa) (MPa) (MPa)
(MPa)
I 28 32830 33210 32840 32960
(SW1 & SW2) 382 35910 38410 38890 37740
II 28 43280 42200 42810 42760
(SW3 & SW4) 355 42020 43100 45900 43670
III 28 41810 40280 39390 40490
(SW5 & SW6) 38 38300 40000 40510 39600

5-8
90

80

70

60
Stress ( MPa )

50
Thorenfeldt et al.
(1987)
40

30 Concrete Mix I
(SD-1&2)
20

10

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01 0.011 0.012 0.013
Strain

Figure 5.4 – 28-day compressive stress-strain curve for concrete mix I.

90

80

70

60
Stress ( MPa )

Thorenfeldt et al.
50 (1987)

40

30 Concrete Mix II
(SD-3&4)
20

10

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01 0.011 0.012 0.013
Strain

Figure 5.5 – 28-day compressive stress-strain curve for concrete mix II.

5-9
90

80

70

60
Stress ( MPa )

Thorenfeldt et al.
50 (1987)

40

30
Concrete Mix III
20 (SW-5&6)

10

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01 0.011 0.012 0.013
Strain

Figure 5.6 – 36-day compressive stress-strain curve for concrete mix III.

Reinforcing Steel

Except for the closed ties used in the edge elements that were built using the plain bars
of grade R (Round Bars) with a nominal yield strength about 250 MPa, deformed
reinforcing bars of grade N (Normal Ductility) with the nominal yield stress equal to
500 MPa were used in the manufacture of all specimens. With the exception of the 4
mm R-grade bars, three samples of each reinforcing bar used in the project were tested
under uniaxial tension and the results are presented in Figures 5.7 to 5.11. The yield and
ultimate strength of the reinforcement samples are summarised in Table 5.6. For the N6
and N8 bars, the yield strength was determined based on a 0.2 percent offset.

5-10
6 mm dia. bars
700

600

500
Yield Stress
Stress ( MPa )

400

300

N6a
200
N6b
N6c
100 0.002 Line

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Strain

Figure 5.7 - Stress-strain curve of 6 mm diameter reinforcement.

8 mm dia. bars
700

600

500
Stress ( MPa )

Yield Stress
400

300
N8a
200 N8b
N8c
0.002 Line
100

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Strain

Figure 5.8 - Stress-strain curve of 8 mm diameter reinforcement.

5-11
12 mm dia. bars
700

600

500
Stress ( MPa )

400

300

200 N12a
N12b
100 N12c

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Strain

Figure 5.9 - Stress-strain curve of 12 mm diameter reinforcement.

16 mm dia. bars
700

600

500
Stress ( MPa )

400

300

200 N16a
N16b
N16c
100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
Strain

Figure 5.10 - Stress-strain curve of 16 mm diameter reinforcement.

5-12
24 mm dia. bars
700

600

500
Stress ( MPa )

400

300

200 N24a
N24b
N24c
100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Strain

Figure 5.11 - Stress-strain curve of 24 mm diameter reinforcement.

Table 5.6 – Properties of Reinforcing Steel


Reinforcement Sample Yield Strain Yield Ultimate Elastic Modulus
Strength Strength
(MPa) (MPa) (GPa)
A 0.00339 534 593 199
B 0.00305 535 597 198
W6
C 0.00320 540 601 198
Average 0.00320 536 597 198
A 0.00322 498 536 178
B 0.00350 497 534 178
W8
C 0.00320 498 536 182
Average 0.00330 498 535 179
A 0.00339 569 643 196
B 0.00323 567 651 204
N12
C 0.00340 577 652 197
Average 0.00330 571 649 199
A 0.00280 535 641 207
B 0.00305 538 633 202
N16
C 0.00303 531 639 202
Average 0.00300 535 638 204
A 0.00351 520 619 195
B 0.00326 526 626 200
N24
C 0.00354 525 624 189
Average 0.00340 524 623 195

5-13
Macalloy Bars

Macalloy bars (or post-tension bars) of 29.5 mm diameter were used in the experiment
to impose the axial load upon the specimens, to restrain the steel base to the strong floor
and to fix the lateral loading assemble to the top slab of specimens (Figure 5.12). From
the uniaxial test the elastic modulus was determined as 190 GPa. Figure 5.13 shows the
typical elastic stress-strain curve for MacAlloy bars.

5.4 Construction of Test Specimens

The test specimens were constructed in three stages including the building up the
formwork, assembling the reinforcement and concrete casting. In the following section
these stages are described in detail.

Figure 5.12 – Macalloy bars used to impose axial load upon a test specimen.

5-14
500
Elastic response only
450

400

350
Stress (MPa)

300

250
Es = 190 GPa
200
1
150

100

50

0
0 0.0005 0.001 0.0015 0.002 0.0025

Strain

Figure 5.13 – Typical Elastic Stress-Strain curve of Macalloy bars.

5.4.1 Formwork

Using 18 mm thick plywood, two timber forms were designed and built up to achieve
the specimen dimensions within ± 0.5 % tolerance (Figures 5.14 and 5.15). The
plywood was screwed together and two levels of timber were bolted around the forms to
constraint the formwork while casting the concrete. To minimise the bleeding of fresh
concrete, the formwork was built such that the specimens were cast horizontally.
Considering a systematic procedure for casting the whole specimen to eliminate the
construction joint, each form included three components named as main form, central
piece and edge piece. The main form was first attached to the base. The central piece
and edge pieces were fixed with main form when the level of cast concrete reached to
the level 1 and level 2, respectively, as shown in Figure 5.14. To install the prestressing
bars that would be used to impose axial load, except for the specimen SW4, six holes of

5-15
40 mm diameter were drilled into the top and bottom parts of form and then PVC tubes
were placed in position to form blockouts (Figure 5.16).

Base

Main Form

Central Piece
Edge peice

Edge peice
A A

Top Veiw

Central Piece Level 3


Edge Piece Level 2
Level 1

Main Form
Base

Section A-A

Figure 5.14 – Form of test specimen.

5-16
Figure 5.15 – Photograph of forms.

Figure 5.16 – Photograph of holes in form to install prestressing bars.

5-17
5.4.2 Assembling the Reinforcement

For each specimen, the reinforcement was firstly cut and bent according to the details.
The reinforcement cage was then assembled as shown in Figure 5.17. After finishing the
reinforcing cage, strain gauges were placed on steel bars at desired locations and
covered by silicone for protection (Figure 5.18).

5.4.3 Casting

The day before casting the forms were cleaned and greased. Then the reinforcement
cage was inserted into the form and lifting anchors were attached to it in two locations
in order to facilitate lifting and handling of specimen after curing (Figure 5.19). The
wires of strain gauges, outside of the forms, were covered in plastic to protect them
against damage during casting and curing.

Figure 5.17 – Photograph of reinforcement cage for the specimen SW6.

5-18
Figure 5.18 – Strain gauges mounted on the steel bars.

Two specimens were cast together using one concrete mix. The concrete was poured
into the form avoiding segregation. Three steps were used for pouring concrete
(Figure 5.20). First, the concrete was poured up to the level 1 equivalent to the top of
wall surface as shown in Figure 5.14. After vibrating the concrete, the central piece was
fixed in its position and the concrete was filled up to level 2, equivalent to the top
surface of edge elements. Finally the edge pieces were screwed to the main form and the
concrete was filled up to level 3 and finished as shown in Figure 5.21. Simultaneously
with the specimens, the control cylinders were cast and vibrated in two layers
(Figure 5.22).

5-19
Figure 5.19 – Photograph of form before casting for the specimen SW6.

Figure 5.20 – Photograph of casting the specimen SW4.

5-20
Figure 5.21 – Photograph of a cast specimen.

Figure 5.22 – Cast control cylinders.

5-21
Figure 5.23 – Specimens during curing period.

Figure 5.24 – Completed test specimens.

5-22
To minimise moisture loss, both the specimens and the cylinder samples were covered
immediately after casting. The specimens and the concrete cylinders were stripped two
days after casting and immediately covered with Hessian sheets. Curing of concrete was
continued for seven days keeping the Hessian wet at all times (Figure 5.23). After seven
days, the specimens and concrete cylinders were stripped and left to air dry until the day
of testing (Figure 5.24).

5.5 Instrumentation

5.5.1 Strain Gauges

To measure the strains on the prestressing, MacAlloy bars used for applying the axial
loading and the reinforcement bars during the testing, 18 electrical strain gauges were
attached at selected locations on the reinforcing steel in the wall and edge elements and
on the MacAlloy bars, as shown in the Figures 5.25 to 5.31. Six strain gauges were
placed on the MacAlloy bars at the mid hight of the walls (Figure 5.25). In the edge
elements, six strain gauges, three each side of the wall, were glued to the external
reinforcement layer to record the strains near the base of the edge elements at the
locations shown in Figure 5.26. Six strain gauges were used in the web wall with three
of these located near the top and bottom slabs and at the mid-height on the transverse
reinforcement and three placed on the longitudinal reinforcement near the edge
elements and at the mid-length. In both directions the gauges were glued to the
reinforcement mesh of the East face. Depending on the reinforcement details for each
specimen, the location of strain gauges was varied slightly, as shown in the
Figures 5.27 to 5.31.

Due to the casting and vibrating operation, some of the strain gauges were accidentally
damaged or their wires were broken. To facilitate the track of strain measuring, the
strain gauges were numbered from 1 to 18 so that the loading rods, the edge elements
and the web wall included the strain gauges of No. 1 to 6, No. 7 to 12 and No. 13 to 18,
respectively. The gauge adopted numbering is shown in Figures 5.25 to 5.28.

5-23
MacAlloy Bar
67 mm
500 mm 500 mm

4 3

500 mm
5 2

500 mm
6 1
67 mm

(a) East Face (b) West Face

Figure 5.25 - Locations of strain gauges in the rods used for axial loading.

100 mm

154 mm
20 mm
12

288 mm
11
10

East Face

Top Flange, External Face

100 mm
89 mm

East Face
221 mm

7
355 mm

8
9

Wall Specimen, East Face Bottom Flange, External Face

Figure 5.26 – Locations of strain gauges in the specimen flanges.

5-24
417 mm

83 mm 83 mm

80 mm
13

18 17 14 16

80 mm
480 mm

400 mm
15

500 mm

Figure 5.27 – Locations of strain gauges in the wall SW-1, East face.

450 mm

50 mm 50 mm
50 mm

13

14
18 17 16
400 mm

400 mm

15
50 mm

500 mm

Figure 5.28 – Locations of strain gauges in the wall SW-2, East face.

5-25
450 mm

50 mm 50 mm

80 mm
13

18 17 16

80 mm
14
480 mm

400 mm
15

500 mm

Figure 5.29 – Locations of strain gauges in the walls SW-3 and SW-4, East face.

417 mm

82 mm 83 mm
50 mm

13

14
16 17 16
50 mm
400 mm

400 mm

15

500 mm

Figure 5.30 – Locations of strain gauges in the wall SW-5, East face.

5-26
500 mm

71 mm 71 mm

67 mm
13

14
18 17 16

67 mm

467 mm
400 mm

15

429 mm

Figure 5.31 - Locations of strain gages in the wall SW-6, East face.

5.5.2 Linear Variable Displacement Transducers (LVTDs)

To measure the displacements of the test walls, six linear variable displacement
transducers (LVDTs) were installed for each specimen as shown in Figure 5.30. The
LVDTs were labelled by an L and numbered as shown. Gauge L1 was used to record
the lateral displacement of the top slab that is equivalent to the vertical displacement in
this experiment with respect to the orientation of the specimen. Gauges L2 and L3
monitored the diagonal displacement of the wall corners and L4 and L5 monitored the
longitudinal displacement of the edge elements. Finally, L6 was installed to check the
out-of-plane displacement of the wall, if any. An additional LVDT was used to monitor
the lateral displacement during testing and control the loading system which has not
been shown here. The range of L1, L2 and L3 was ± 75, while that of L4, L5 and L6
was ± 15 mm. The accuracy of the all LVDTs was 0.01 mm.

5-27
L4

L2

L3

L5

L6 L6

L1 L1

Side View Front View

Figure 5.32 – Locations of LVDTs.

5.6 Test Set-up

The specimens were tested in the stiff testing frame of the structural engineering
research laboratory at UNSW (Figures 5.33 and 5.34). Due to the vertical orientation of
the actuator, the specimens were tested on their sides as shown in Figures 5.34. The test
set-up was composed of two major components named as the fixing and loading
systems.

The fixing system included specially built assemblies to restrain the specimen and
prevent it from lifting or sliding during testing. The bottom slab of the specimens was
fixed against sliding using a rigid steel base attached to the laboratory strong floor and a
pair of steel beams attached to the testing rig at the top of the specimen. The base was
built using two steel 310UC137 beams fixed to the strong floor using 4 pairs of
prestressed bars as shown in Figures 5.35 and 5.36. A special connection including two

5-28
symmetrical components was used to fix the bottom of the specimen to the strong floor
(Figures 5.36). The top of the specimen was held in place by two 310UC137 beams
which were connected together and to the rig as shown in Figure 5.37. To prevent the
specimen against lifting an axial compressive force set at 2000 kN for the unloaded
specimen was applied to the top of the specimen base to precompress the specimen
base using a hydraulic jack supported by a stiff beam attached into the testing frame
(Figures 5.33 and 5.38). The axial load also increased the friction force between steel
base and the concrete strong floor. The axial load applied by this jack varied throughout
the test depending on the reactive force required by the applied load. At all times,
however, the base of the specimen was kept in compression.

The axial load was approximately constant during testing. A displacement control
method was used to impose the cyclic lateral displacements to the top slab of specimens
through a 5000 kN capacity actuator mounted to the testing rig (Figure 5.39). The
connection between the load cell and the top slab was provided using special built
assemblies held down together by four prestressed MacAlloy bars. Prior to testing these
rods were tensioned to ensure that the assemblies tightly attached to the both ends of top
slab. The axial load was applied onto the specimens (except for the specimen SW4) by
six prestressed MacAlloy bars symmetrically inserted into the holes formed during
casting operation as previously discussed (Figure 5.40). After fixing the specimens in
the right position, each prestressed bars was tensioned to 200 kN using a hydraulic jack
and anchored so that a total applied axial load was 1200 kN. The stiff top slab was used
to uniformly distribute the axial load along the wall.

5.7 Testing Procedure

The test procedure includes the installation of specimen, loading process and data
recording, as described below. On the day of testing, the concrete cylinder samples were
tested to measure the compressive strength, tensile strength and elastic modulus.

5-29
East West North South

(a) (b)
Testing Rig

Actuator

Hyraulic Jack
Specimen

Specimen

Steel Base

Srong Floor Strong Floor

Figure 5.33 – Test set-up: (a) front view (b) side view.

5-30
Figure 5.34 – Photograph of test set-up, West view.

Figure 5.35 – Steel base fixed to strong floor.

5-31
Figure 5.36 – Connection between the specimen and the base.

Figure 5.37 – Restraining top beams, East view.

5-32
Figure 5.38 – Hydraulic jack and stiff beam.

Figure 5.39 - Lateral loading system , front view.

5-33
Figure 5.40 – Prestressing MacAlloy bars used to apply axial load.

5.7.1 Specimen Installation and Loading

The test specimen was placed and fixed in the test rig several days before the test. The
day before testing the axial load was applied to the specimen, the LVDTs were placed in
their positions and the data acquisition system connected. To check that all systems
worked correctly, an initial lateral load of 50 kN was applied to the each specimen and
then released.

On the day of testing, the wall specimens were subjected to a combination of constant
axial load and cyclic lateral loading as shown in Figure 5.41, except for specimen SW4,
in which the axial load was zero. To study the pre-cracking as well as post-cracking
behaviour of specimens, lateral loading was applied using a displacement control. The
specimens were tested under reversed cyclic conditions displacing them laterally, along
the axis of the web wall, in 4-mm increments in the negative (downward) and positive

5-34
(upward) directions (Figure 5.41). Since the wall specimens had high strength and
stiffness and their behaviours were approximately brittle and linear until failure, using
the displacement increments of 4 mm seems to be logical. A loading rate of 30 minutes
per cycle was maintained until the specimens experienced a significant loss of capacity.
For the cyclic tests, two repetitions at each displacement level were imposed for each
phase. The loading history applied to the specimens is shown in Figure 5.42.

5.7.2 Data Recording

Two computers were used to record the test data. A computer connected to the Instron
actuator, was used to control the displacement (and load) applied on the each specimen.

Reaction Pre-Load (2000 kN for zero cyclic lateral load)

_ Cyclic Lateral Load

Axial Load

Figure 5.41 – Loading System applied on the walls.

5-35
20
Phase 4
16 Phase 3
12 Phase 2
Lateral Displacement (mm)

8
Phase 1
4
0
-4
-8
-12
-16
-20
0 1 2 3 4 5 6 7 8 9
Cycle No.

Figure 5.42 – Loading history applied on the test specimens.

Data from the load cell, LVDTs and strain gauges were collected by a data logger which
was connected to an HBM amplifier. The data, converted into digital readings by the
data logger, was stored in the computers and was used to graph and analyse the test
results (presented in Chapter 6). Additionally, during and after testing photographs were
taken of the damaged specimens.

5-36
CHAPTER 6

EXPERIMENTAL RESULTS

6.1 Introduction

In this chapter, experimental results are presented for the six shear wall specimens
described in Chapter 5. For each specimen, the main aspects including the applied
loading, the lateral load versus displacement observations, the crack pattern, the failure
mode, the strain of the reinforcing steel bars and the response of wall as well as
edge elements are given. The location of strain gages and LVDTs are shown in
Figures 5.23 to 5.29 and Figure 5.30, respectively. The positive and negative
quantities for lateral load as well as displacement refer to pulling of the specimen (up)
and pushing of the specimen (down), respectively. Positive strains are tensile and
negative strains compressive. The axial load described in this chapter is always
compression. A summary of the test results is given at the end of the chapter. The
separate load-displacement response under each lateral loading phase for the wall
specimen as well as the load displacement diagrams for the LVDTs are illustrated in
the Appendix B. Load-strain curves for the MacAlloy bars used to apply the axial load
for all specimens are given in Appendix C. The strain versus lateral load curves for the
strain gauges installed in the edge elements and in the walls are shown in the
Appendices D and E, respectively.

The photographs showing cracking during the test were taken using a digital camera
and the cracks were marked later with a digital pen so that the paths could be clearly
identified (as the occupational health and safety requirement did not allow the cracks
to be marked during testing). In these figures the closed cracks are not marked. At the
completion of testing, the cracks were highlighted with thick Texta pen and the photos
taken after testing shows the Texta pen marked.

6.2 Observed Response of Specimens

6.2.1 Specimen SW1

Specimen SW1 had longitudinal and transverse reinforcement ratios of about 0.5
percent. The specimen was subjected to a combination of axial load and lateral
reversed cyclic loading. The age of the specimen at the time of testing was 382 days.
The compressive strength of concrete on the day of test was 86 MPa.

Loading History

The MacAlloy bars were initially loaded to a combined axial load of 1200 kN and then
the lateral displacements were cyclically imposed in increments of 4 mm until
complete failure took place (see Figure 5.40) with two complete cycles for each
displacement cycle increment. Totally, five displacement cycles were applied on the
specimen prior to failure. Using the strains measured in the MacAlloy bars during the
test (refer to Appendix C) the overall axial load applied to the specimen is calculated
and presented in Figures 6.1 and 6.2. As the figures show, there was some change in
the axial load due to the lateral displacements during the testing. The axial load varied
from 1200 kN at the zero position to a larger amount at the peak points of each cycle.
The maximum axial load occurred at the failure point of last cycle and was 1288 kN,
an increase of 7.3 percent compared with the initial axial load. In general, the axial
load of bottom MacAlloy bars did not change during the negative direction of lateral
loading, but increased in positive direction (see Appendix C). This was inversed for

6-2
1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure 6.1 – Axial load vs. lateral load for the specimen SW1.

1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.2 – Axial load vs. lateral Displacement for the specimen SW1.

6-3
the top MacAlloy bars. The axial load in the middle MacAlloy bars did not change
significantly for loading either in the positive or negative directions.

Load-Displacement Response

The load-displacement response of the specimen SW1, as recorded by LVDT L1, is


shown in Figure 6.3. Testing was terminated after the completion of first excursion
towards 12 mm. In the positive direction, the lateral load at which the failure occurred
was 992 kN and corresponded to a displacement of 11.93 mm. In the negative
direction the peak load was -1007 kN at a displacement of -11.99 mm. The wall did
not exhibit a ductile post-peak response or any significant energy dissipation. No
pinching effects were observed during unloading. There was little difference between
the path of the first and second cycles for the 4 mm displacement cycle. For the second
displacement cycle, the wall stiffness was less than that of the first due to damage that
occurred during the cyclic loading. After failure a significant degradation in stiffness is
observed.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.3 – Load-displacement response of the specimen SW1.

6-4
Crack Pattern

During the phase one, some minor cracks formed trough the central part of wall from
the bottom corner of a wall side toward the top corner on the opposite side at an angle
of about 50 degrees with respect to the bottom slab. As the displacement level was
increased in the second phase, new minor cracks formed both sides of the diagonal and
the width and length of previous cracks increased. Some cracks were also observed in
the edge elements during this phase. Figures 6.4 and 6.5 show the cracks in both
negative and positive directions of lateral load. In phase three, a major shear crack of
45 degrees occurred along the main diagonal zone and crossed into the edge elements
leading to the failure of the specimen (Figure 6.6).

Figure 6.4 – Cracks under the applied load in the negative direction for SW1.

6-5
Figure 6.5 – Cracks under the applied load in the positive direction for SW1 (closed
cracks not drawn).

Figure 6.6 – The specimen SW1 after testing.

6-6
Failure Mode

Failure occurred at the first excursion of phase 3 (12 mm of displacement) in the


positive direction when the lateral load reached to 977 kN. The failure was sudden
with that the lateral load dropping sharply after the peak load. The wall failed in a
shear mode with crushing of the concrete along the main diagonal cracks that had
formed in the web (Figure 6.6). The failure surface was at an angle of about 45
degrees and extended into the edge elements as shown in Figure 6.5. At failure, while
the longitudinal reinforcement was bent due to the movement of top piece of wall
above the major crack, most of the transverse reinforcing bars crossing the failure
surface had snapped (Figure 6.7). The edge elements experienced some flexural
cracking but no other significant damage was apparent.

Figure 6.7 – Reinforcement at the failure of specimen SW1.

6-7
Reinforcement Strains

All strain gauges attached to the reinforcement in the edge elements recorded strains
less than the yield strain (3000 μ ε for the 16 mm diameter bars). A maximum strain
of 1892 μ ε at a load of 990 kN was recorded for strain gauge eight in bottom edge
element. The maximum strain recorded for the top edge element was 1428 μ ε at
-1006 kN for strain gauge 10 (Figures 6.8 and 6.9). The response of reinforcing steel
in compression was approximately linear but showed nonlinearity in tension after a
strain of 500 μ ε due to the flexural cracking of the concrete in the edge elements and
the increasing contribution of the reinforcement to carry load. The results for all strain
gauges are given in Appendix D.

Except for the strain gauge 17 located to the transverse bar at the mid-height of the
reinforcement in both the transverse and longitudinal directions had not yielded prior
to failure of the wall. The maximum strain measured in strain gauge 17 was 5261 με
at the negative peak load of the cycle five equal to -1006 kN (Figure 6.11). It was
evident that the reinforcement strain experienced a large jump in the third
displacement with gauges 15 and 17 indicating significant energy dissipation in the
reinforcing steel compared with the others gauges (see Appendix E).

Displacements of Web Wall and Edge Elements

LVDTs L2 and L3 were used to measured the diagonal deformation of the web wall
and the results are plotted in Figures 6.12 and 6.13. They show a low ductility
response of the wall. The response of the wall is dominated by shear and the energy
dissipation during the cyclic loading was low. A maximum diagonal displacement of
4.0 mm was recorded during cycle five.

6-8
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.8 – Load-strain response of Strain gauge 8 for the SW1.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.9 – Load-strain response of Strain gauge 10 for the SW1.

6-9
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2500 -2000 -1500 -1000 -500 0 500 1000 1500 2000 2500

Strain(μm/m)

Figure 6.10 – Load-strain response of Strain gauge 15 for the SW1.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000 5000 6000

Strain(μm/m)

Figure 6.11 – Load-strain response of Strain gauge 17 for the SW1.

6-10
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.12 – Load-displacement response of LVDT L2 for the SW1.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.13 – Load-displacement response of LVDT L3 for the SW1.

6-11
The response of edge elements, as monitored by LVDT L5, showed the element to be
stiff (Figure 6.14). At the peak-load in cycle 5 a maximum displacement of 0.6 mm
was recorded for the bottom edge element. The bottom boundary element acted like a
rigid plate in the negative loading direction and a linear elastic element in the positive
loading direction. This was inverse for the top edge element (refer to Appendix B). No
residual displacements were observed in either of the edge elements.

6.2.2 Specimen SW2

Specimen SW2 had longitudinal and transverse reinforcement ratios equal to 1.3
percent. The specimen was subjected to a combination of axial load and lateral
reversed cyclic loading. The age of the specimen at the time of testing was 388 days.
The compressive strength of concrete on the day of test was 86 MPa.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.14 – Load-displacement response of LVDT L5 for the SW1.

6-12
Loading History

The loading applied on the specimen SW2 was similar to that of the specimen SW1.
After imposing four loading phases on the specimen, a sudden failure occurred in the
first excursion into phase 5 in the negative direction. The displacement was continued
to complete the cycle. Because of an error in entering the control data into the
displacement control system, the peak-displacement of cycles did not match the
pre-programmed values. The actual imposed displacements in the positive direction
cycles were 2.5, 6.1, 9.2 and 12.5 mm whereas those of negative directions were -6.2,
-8.2, -12.9 and -17.6 mm. The total axial load applied to the specimen from the
contributions of the MacAlloy bars is given in Figures 6.15 and 6.16. Like specimen
SW1 the applied load was symmetric with respect to the zero load position. The
maximum axial load recorded at the negative direction of cycle 8 was 1305 kN, an
increase of 8.7 percent relative to the initial axial load. The variation of axial load for
each MacAlloy bar is given in Appendix C.

1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure 6.15 – Axial load vs. lateral load for the specimen SW2.

6-13
1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure 6.16 – Axial load vs. lateral Displacement for the specimen SW2.

Load-Displacement Response

Figure 6.17 shows the load-displacement response of the specimen SW2. In the
positive direction, a maximum lateral load of 1190 kN corresponding to a
displacement of 12.52 mm was recorded. The values for the negative direction were
-1192 kN and -17.53 mm, respectively. Failure was initiated just at the completion of
the forth loading phase in the positive direction (second cycle). On the next excursion
into the negative direction, the load reached just -816 kN at a displacement of
-13.6 mm. Although the energy dissipated by SW2 was greater than SW1, the
post-peak response was not ductile and as for wall SW1, no pinching effect was
observed. For the both excursions of loading phases one to three, the loading and
unloading curves followed a similar path. During phase four, however, the first and
second curves were different with the progressing damage. The damage also caused a
decrease in the slope of the loading and unloading curves as the stiffness of the wall
was degraded.

6-14
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure 6.17 – Load-displacement response of the specimen SW2.

Crack Pattern

The crack pattern and propagation at the early loading phases is shown in Figure 6.18
and were similar to the SW1, although the cracks distributed throughout the wall more
uniformly compared with wall SW1. During phase three the regions of inclined cracks
near the bottom wall corners started to gradually crush (Figure 6.19). As the loading
progressed, a widespread shear sliding damage zone formed along the width of wall
near the base slab (Figure 6.20). In the middle regions of wall near the base slab, no
significant cracking was observed. A few shear cracks developed toward the middle of
top slab. In addition to the wall cracks, flexural tension cracks occurred in the edge
elements along the width of the elements.

6-15
Figure 6.18 – Cracks under the applied load in the positive direction for SW2 (closed
cracks not indicated).

Figure 6.19 – Cracks under the applied load in the negative direction for SW2 (closed
cracks not indicated).

6-16
Figure 6.20 – Shear sliding damage zone for SW2.

Figure 6.21 – Specimen SW2 after testing.

6-17
Failure Mode

During the first and second loading phases, the cracking occurred across the diagonals
of wall at approximately 45 degrees to the horizontal. During the third and forth
phases, however, a severe cracked zone developed along the width of wall near the
bottom slab (Figure 6.19). Crushing of the concrete was observed at this stage in the
end regions of this zone. At the peak-load of the second excursion of phase 4 in the
positive direction, a partial shear crack occurred. As the loading was progressed,
another major shear crack occurred with failure at an absolute load higher than the first
occurred during the first excursion of the phase 5 in negative direction. A sliding shear
plane resulted after phase three and the wall failed by shear sliding (Figure 6.20). The
edge elements did not fail but because of crushing the concrete under severe axial load
in the edge elements, longitudinal cracks formed in the stiffened edges accompanied
by concrete splitting (Figure 6.21). There was no evidence of either the longitudinal or
transverse reinforcement crossing the failure zone having fractured.

Reinforcement Strains

The maximum strain measured by the gauges installed in the edge elements was
2450 μ ε in gauge nine in the bottom edge element at P = 983 kN and is less than the
yield strain for the longitudinal bars (Figures 6.22) . The maximum strain recorded for
the top edge element was 1340 μ ε at P = -1193 kN for strain gauge 11 (Figures 6.23).
While the response of the reinforcing steel for the loading in the negative direction
was approximately linear, the reinforcement showed stiffness degradation associated
with residual strains for the positive direction of loading. This observation was more
significant for the last loading phase. The flexural and bond cracking of concrete in the
edge elements, specially near the edge reinforcement, resulted in increasing the
contribution of the reinforcement to carry load. Other strain gauges showed a linear
response in both compression and tension (see appendix D).

6-18
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.22 – Load-strain response of Strain gauge 9 for the SW2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000
Strain (μm/m)
Figure 6.23 – Load-strain response of Strain gauge 11 for the SW2.

6-19
Figures 6.24 and 6.25 show the response of the mid-longitudinal and mid-transverse
reinforcement, respectively, for the web wall. The response of strain gauge 14 is
symmetric with similar stiffness in each loading direction. The maximum recorded
strain was 2860 μ ε at the last loading phase, less than the yield strain. Prior to the last
excursions, the strains recorded for the gauge 17 were less than the yield strain. At the
last excursion, however, the maximum strain of 3550 μ ε exceeded the yield strain of
3330 μ ε . The graph shows an initial residual strain at P = 0 after cracking. As the
load-displacement progressed, the residual (plastic) strain continued to increase. The
other strain gauges showed a linear response without yielding of the reinforcing steel
(see Appendix E).

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.24 – Load-strain response of Strain gauge 14 for the SW2.

6-20
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000

Strain(μm/m)

Figure 6.25 – Load-strain response of Strain gauge 17 for the SW2.

Web Wall and Edge Element Displacements

The LVDTs installed in the diagonal direction of web wall showed, generally, a linear
response without significant energy dissipation (Figure 6.26). At the last loading
excursion a considerable degradation of stiffness, attributed to the final failure, was
observed. A peak displacement across the diagonal of -3.1 mm was recorded during
the last cycle before failure.

The response of edge elements of the SW2 was similarly very stiff without any
dissipation of energy (Figure 6.27). For the bottom edge element, a maximum
displacement of 1.1 mm at the peak-load in the positive loading direction was
recorded.

6-21
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.26 – Load-displacement response of LVDT L2 for the SW2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.27 – Load-displacement response of LVDT L5 for the SW2.

6-22
6.2.3 Specimen SW3

Specimen SD-3 had longitudinal and transverse reinforcement ratios of 0.8 percent
with the specimen subjected to a combination of axial load and lateral reversed cyclic
loading. The age of the specimen at the time of testing was 355 days. The compressive
strength of concrete on the day of test was 96 MPa.

Loading History

Specimen SW3 was tested under displacement cycles accompanied by an axial load of
1200 kN. The lateral loading was applied to the specimen through the complete phases
of 4, 8 and 12 mm and a half-cycle in the negative direction at the 16 mm phase. At
this stage the test was terminated, as the wall had failed in both directions and its
stiffness was significantly decreased. Figures 6.28 and 6.29 demonstrate the variation
of overall axial load versus the lateral loads and displacements. The axial loads at the
peak points of the negative and positive directions were 1302 kN and 1300 kN,
respectively. These values are 8.5 percent higher than the applied load at P = 0 of
1200 kN. After failure, the axial load dropped sharply to the 1172 kN and then
decreased further as failure progressed.

Load-Displacement Response

As for the previous specimens, the behaviour of SW3 was dominated by shear action
with little ductility (Figure 6.30). In the positive direction, the values for the maximum
lateral load and the corresponding displacement were 1107 kN and 12.31 mm,
respectively, occurring in the first excursion of loading phase three. In the negative
direction, the maximum load and corresponding displacement were -1090 kN and
-12.12 mm, respectively. The response included two jumps at failure in the positive
and negative directions. Similar to SW1 and SW2, the energy dissipation was low but,
in contrast to those walls, a slight pinching effect was observed. Prior to failure,

6-23
1400

1350

1300

1250
Axial (kN)

1200

1150

1100

1050

1000
-1500 -1000 -500 0 500 1000 1500
Lateral Load (kN)

Figure 6.28 – Axial load vs. lateral load for the specimen SW3.

1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.29 – Axial load vs. lateral displacement for the specimen SW3.

6-24
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure 6.30 – Load-displacement response of the specimen SW3.

the loading and unloading paths for the excursions of each loading phase were similar.
While the degradation in the wall stiffness for the first two loading phases was slight,
it was greater for phase three. Although after failure in the positive direction, the wall
stiffness sharply dropped, it increased gradually during following load stage in the
negative direction. After failure in the negative direction, however, the response of the
wall softened significantly.

Crack Pattern

The diagonal inclined cracking initially appeared during phase one and then developed
further throughout the wall during the second phase. In the next phases the length and
width of these cracks increased and new cracks started to surface between the initial

6-25
primary cracks and were spaced uniformly throughout the wall (Figures 6.31
and 6.32). At failure in the positive direction, a major crack at about 45 degrees was
observed extending into the edge elements (Figure 6.33). Failure occured in the
negative direction, although no similar major crack was visible. Similar to SW1, some
flexural cracks were visible near the bottom regions of the edge elements.

Failure Mode

The failure was accompanied with a major crack at P = 1052 kN in the positive
loading direction and corresponded to a displacement of 10.86 mm. Crushing of the
concrete between the diagonal cracks and spalling of concrete cover were visible. The
failure crack extended in the edge elements. Some of transverse reinforcement was
observed to be fractured at the completion of the test. The specimen after testing is
shown in Figure 6.33.

Figure 6.31 – Cracks under the applied load in the negative direction for SW3 (closed
cracks not indicated).

6-26
Figure 6.32 – Cracks under the applied load in the positive direction for SW3 (closed
cracks not indicated).

Figure 6.33 – The specimen SW3 after testing.

6-27
Reinforcement Strains

The strain gauges at the bottom edge element showed a linear response with a small
strain (see gauge 7 data plotted in Figure 6.34). The stiffness response was similar for
both loading directions. The strain gauges installed in the bottom edge element,
however, showed greater strains as well as stiffness degradation, particularly in the
negative loading direction (Figure 6.35). The measured strains were less than yield
with the maximum recorded strain of 2810 μ ε in gauge 11 at the last loading
excursion.

For the longitudinal web reinforcement, only the strain recorded in gauge 15 was a
little more than the yield strain with a maximum strain of 3970 μ ε recorded for the
last loading phase (Figure 6.36). While the initial cracking led to an increase in the
residual strain at P = 0, the residual strains did increase further during the test. The
stiffness for the both loading directions were approximately equal.

In contrast to longitudinal reinforcement, strain gauges 17 and 18 connected to the


transverse steel bars recorded strains greater than yield (Figures 6.37 and 6.38). The
maximum strain was 5290 μ ε in gauge 17. Both responses included increase in the
residual strain during each subsequent cycle.

Displacements of Web Wall and Edge Elements

Similar to the previous specimens, the response recorded for the LVDTs installed
across the diagonal of the web as well as those installed on the edge elements showed
a stiff and approximately linear response (Figures 6.39 and 6.40). Energy dissipation
in the diagonal directions was low, while in the edge element direction it was close to
zero. The maximum displacements in the diagonal and edge element directions were
3.5 mm and 0.8 mm, respectively.

6-28
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure 6.34 - Load-strain response of Strain gauge 7 for the SW3.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure 6.35 – Load-strain response of Strain gauge 11 for the SW3.

6-29
1400
1200
1000
800
600
Lareral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -4000 -2000 0 2000 4000 6000

Strain(μm/m)

Figure 6.36 – Load-strain response of strain gauge 15 for the SW3.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000

Strain(μm/m)

Figure 6.37 – Load-strain response of strain gauge 17 for the SW3.

6-30
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -4000 -2000 0 2000 4000 6000

Strain(μm/m)

Figure 6.38 – Load-strain response of strain gauge 18 for the SW3.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-16 -12 -8 -4 0 4 8 12 16

Displacement (mm)

Figure 6.39 – Load-displacement response of LVDT L2 for the SW3.

6-31
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.40 – Load-displacement response of LVDT L5 for the SW3.

6.2.4 Specimen SW4

Specimen SW4 had longitudinal and transverse reinforcement ratios of 0.8 percent.
The specimen was subjected to the lateral reversed cyclic loading but had no applied
axial load. The age of the specimen at the time of testing was 358 days. The
compressive strength of concrete on the day of test was 96 MPa.

Loading History

Wall SW4 was subjected to only lateral loading with four phases of lateral loading
completed with failure occurring during the first excursion into phase three. The
maximum loads recorded in the negative and positive directions were -684 kN and
752 kN, respectively, and corresponded to displacements of -7.9 and 12.1 mm,
respectively.

6-32
Load-Displacement Response

The lateral load-displacement response for the specimen SW4 is shown in Figure 6.41.
The maximum load in the positive direction was 753 kN at a corresponded
displacement of 12.3 mm. In the negative direction, the maximum load was -683 kN
and corresponded to a displacement of -7.9 mm. Although after the failure the wall
stiffness in the negative direction dropped sharply, the wall showed some ductility in
the positive direction with the wall stiffness degraded gradually and showing
significant energy dissipation with a notable pinching effect.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.41 – Load-displacement response of the specimen SW4.

6-33
Crack Pattern

Minor shear cracks appeared during the first cycle. The next cycles led to extending
these cracks and the appearance of the secondary cracks. Some minor flexural cracks
were also visible in the edge elements. Advancing the cyclic displacements, the shear
cracks propagated the full height of the web wall with an orientation of about 45
degrees to the bottom slab (Figures 6.42). The cracks were approximately equally
spaced throughout the middle region of the web. At failure, a major diagonal crack
formed from the corner of the top edge element to the opposite corner of bottom edge
element (Figure 6.43). Prior to failure, the cracks that had slide during loading were
realigned during unloading. The grinding of the crack surfaces caused localised
crushing along the cracks (Figure 6.44). Some minor flexural cracks formed along to
the width of edge elements. The state of the wall at the completion of the test is shown
in Figure 6.45.

Figure 6.42 – Cracks under the applied load in the positive direction for SW4 (fully
closed cracks not showed).

6-34
Figure 6. 43 – Cracks under the applied load in the positive direction for SW4 (closed
cracks not indicated).

Figure 6.44 – Cracks when the loading continued after the failure load for SW4 cracks
not digitally marked).

6-35
Figure 6.45 – The specimen SW4 after testing.

Failure Mode

The wall failed during cycle five at a load of -683 kN, corresponding to a displacement
of -8.0 mm. The failure was induced by shear action with a major diagonal crack
forming in the web wall (Figure 6.46). The crack extended into the edge elements at
both sides toward the intersection points with the top and bottom slabs at an angle of
about 45 degrees. Before failure the edge elements experienced flexural cracking, but
no significant damage was visible.

After the completion of the test, it was observed that the longitudinal reinforcing bars
located near the failure surface were bent, whereas, some transverse bars in this region
had fractured (Figure 6.46).

6-36
Figure 6.46 – Reinforcement at the failure surface of SW4.

Reinforcement Strains

The strain gauges attached to the reinforcement in the edge elements recorded strains
less than yield. The strains were small with exception of strain gauge 9 where a
maximum strain of 1840 μ ε was measured during the last excursion (Figure 6.47).

The strain measured on the longitudinal web reinforcement indicated that the bars did
not yield, or at least did not yield at the location of the gauge (strain gauge 15 recorded
the maximum strain of 2930 μ ε ). In the transverse reinforcing bars, however, strains
greater than the yield strain were recorded by the strain gauges 17 and 18 with the
maximum strain measured by gauge 18 of 7170 μ ε at the last loading excursion
(Figure 4.49). Significant residual plastic strains at P = 0 are seen in Figure 4.49,
particularly for the later loading phases.

6-37
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000
Strain (μm/m)

Figure 6.47 – Load-strain response of Strain gauge 9 for the SW4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.48 – Load-strain response of Strain gauge 15 for the SW4.

6-38
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-8000 -6000 -4000 -2000 0 2000 4000 6000 8000

Strain(μm/m)

Figure 6.49 – Load-strain response of Strain gauge 18 for the SW4.

Displacements of Web Wall and Edge Elements

The response of LVDT L2, installed in the diagonal direction, is presented in


Figure 6.50. The figure shows a degree of energy dissipation occurred during the test.
Prior to the peak load, the response was similar to that observed for the other
specimens. After failure, however, the wall was considerably more ductile with greater
energy dissipation than previous walls. The maximum displacement recorded for
LVDT L2 was 11.6 mm.

The behaviour of the edge elements for SW4 was similar to those of other specimens
(Figure 6.51). For the bottom edge element, a maximum displacement of 1.5 mm was
recorded before failure.

6-39
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.50 – Load-displacement response of LVDT L2 for the SW4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.51 – Load-displacement response of LVDT L5 for the SW4.

6-40
6.2.5 Specimen SW5

Specimen SW-5 had longitudinal and transverse reinforcement ratios equal to 1.3
percent and 0.5 percent, respectively. The specimen was subjected to a combination of
axial load and lateral reversed cyclic loading. The age of the specimen at the time of
testing was 38 days. The compressive strength of concrete on the day of test was
83 MPa.

Loading History

Wall SW5 was tested under the combination of lateral loading and an axial load of
1200 kN. The lateral loading was applied through the phases one to three and the first
excursion into phase four. Because of a problem with the performance of LVDT L1
used to control the applied displacement during the phase two of testing, the test was
paused and the gauge replaced. As a result, the programmed displacements for the
phases three and four were not completely matched. From the strains measured in the
Macalloy bars during the test (refer to Appendix C) the total axial load applied to the
specimen was calculated and is given in Figures 6.52 and 6.53. The axial load
increased with increasing lateral displacement. The maximum axial load occurred at
the first excursion of the phase three in the negative direction and was 1316 kN at a
displacement of 15.9 mm. This amount was about 9.6 percent greater than 1200 kN.

Load-Displacement Response

The load versus lateral displacement for SW5 is presented in Figure 6.54. The
dissipation of energy was low and no pinching effect was observed. The specimen had
a low level of ductility with low energy dissipation. The maximum loads and their
corresponding displacements were 1073 kN at 13.51 mm in the positive direction and
-1134 kN at -10.41 mm in the negative direction. At the failure points, in both the

6-41
1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure 6.52 – Axial load vs. lateral load for the specimen SW5.

1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.53 - Axial load vs. Displacement for the specimen SW5.

6-42
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.54 – Load-displacement response of the specimen SW5.

positive and negative directions, a jump in load is seen for little increase in
displacement. The stiffness of the wall decreased as the loading phases advanced.
After the second failure point, the wall significantly lost stiffness. A subtle difference
is observed between the loading and unloading paths in the two completed excursions
before the failure.

Crack Pattern

Diagonal cracks developed uniformly throughout the web of the wall


(Figures 6.55 and 6.56). These cracks were accompanied with some minor flexural
cracks in the edge elements near the bottom slab. A major diagonal crack developed at
the failure load of the negative direction (Figure 6.56) and extension of the diagonal
crack into the edge elements was evident. The state of the specimen at the completion
of testing is shown in Figure 6.57.

6-43
Figure 6.55 – Cracks under the applied load in the positive direction for SW5 (fully
closed cracks not indicated).

Figure 6.56 – Cracks under the applied load in the negative direction for SW5 (closed
cracks not indicated).

6-44
Figure 6.57 – The specimen SW5 after testing.

Failure Mode

Similar to that of specimens SW1, SW3 and SW4, the wall SW5 failed in a sudden
shear mode after crushing the concrete struts between the diagonal cracks and at the
corners of the edge elements. The failure occurred during phase three in the negative
direction at a load of -1047 kN corresponding to a displacement of -9.65 mm. After
failure the stiffness decreased but the wall still resisted the positively applied load until
a load of 819 kN, after which the wall failed in the positive direction. Continuing the
loading, the damaged zone around the failure crack widened (Figure 6.57). After
testing, it was observed that some of the transverse reinforcement crossing the failure
crack had fractured (Figure 6.58). During testing some flexural cracks occurred at the
edge elements, although they did not cause any major damage.

6-45
Figure 6.58 – Snapping the transverse reinforcement crossed the failure crack of SW5.

Reinforcement Strains

All strain gauges attached to the reinforcement in the edge elements recorded strains
less than the yield strain. The strains recorded in these were generally small and the
responses were approximately linear. The maximum strain was recorded in the strain
gauge 9 as 860 μ ε at a load of 1083 kN (Figure 6.59).

In the web all strain gauges attached to the transverse and longitudinal reinforcement
recorded the strains smaller than the yield strain. The maximum recorded strain in the
longitudinal direction was 1570 μ ε in gauge 15 and in the transverse direction was
2740 μ ε in gage 17 (Figures 6.60 and 6.61). Although strains less than yield were
recorded, fracture of the web reinforcement crossing the major failure crack, dictates
that strains greater than yield occurred in the web steel in these regions. Significant
residual, plastic, strains were measured after cracking in strain gauge 17.

6-46
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure 6.59 – Load-strain response of Strain gauge 9 for the SW5.

Displacements of Web Wall and Edge Elements

The displacement across the diagonal (LVDT L2) and along the edge element
(LVDT L5) are plotted in Figures 6.62 and 6.63, respectively. The response were
similar to those for walls SW1 to SW3.

6.2.6 Specimen SW6

Specimen SW6 had reinforcement ratios of 1.0 percent in each of the longitudinal and
transverse directions. The specimen was subjected to a combination of axial load and
lateral reversed cyclic loading. The age of the specimen at the time of testing was 35
days and the compressive strength of concrete on the day of test was 83 MPa.

6-47
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2000 -1500 -1000 -500 0 500 1000 1500 2000

Strain (μm/m)
Figure 6.60 – Load-strain response of Strain gauge 14 for the SW5.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.61 – Load-strain response of Strain gauge 17 for the SW5.

6-48
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.62 – Load-displacement response of LVDT L2 for the SW5.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.63 – Load-displacement response of LVDT L5 for the SW5.

6-49
Loading History

A lateral cyclic loading accompanied with an axial load of 1200 kN was applied to
SW6. The wall failed during the first excursion into phase four in the positive
direction. After failure, the wall exhibited an extreme degradation of stiffness and the
test was stopped before completing the second excursion in the positive direction.
Using the strains measured in the Macalloy bars (refer to Appendix F) the axial load
applied to the specimen was calculated and is plotted in Figures 6.64 and 6.65. The
maximum axial load measured during the test was 1316 kN (9.7 percent higher than
initial load).

Load-Displacement Response

The behaviour of SW6 was similar to the earlier tests on specimens SW2 and SW3
and no pinching effects were observed in the cyclic response (Figure 6.66). The wall
had little ductility with a sharp decrease in the load at failure and little energy
dissipation. In the positive direction, the maximum lateral load was 1141 kN
corresponding to a displacement of 14.0 mm, occurring during the first excursion into
phase four. In the negative direction the maximum load was -1101 kN at -16.8 mm.
The wall stiffness degraded considerably after failure in both loading directions.

Crack Pattern

Cracking for the SW6 was similar to that for SW5. Diagonal cracks started to develop
uniformly throughout the web wall during the early loading phases (Figures 6.67 and
6.68) with increasing length and width of cracks as loading progressed. Some minor
flexural cracks occurred in the edge elements near the bottom slab. A major diagonal
crack developed at the failure load in the negative direction extending into the edge
elements (Figure 6.69). After failure, slip across the cracks was observed.

6-50
1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure 6.64 – Axial load vs. lateral Load for the specimen SW6.

1400

1350

1300
Axial Load (kN)

1250

1200

1150

1100

1050

1000
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab(mm)

Figure 6.65 – Axial load vs. lateral Displacement for the specimen SW6.

6-51
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure 6.66 – Load-displacement response of the specimen SW6.

Failure Mode

The wall failed in shear with crushing of the concrete between the diagonal cracks and
at the region near the bottom slab (Figures 6.68). The concrete in the edge elements at
the ends of inclined cracks also showed significant crushing (Figure 6.69). By
continuing the loading after failure, the concrete between the struts crushed and a
severe damaged zone formed (Figure 6.69).

Reinforcement Strains

The maximum strain recorded in the edge elements was 1250 μ ε in strain gauge 9
(Figure 6.70) with the response being relatively linear.

6-52
Figure 6.67 – Cracks under the applied load in the positive direction for SW6 (closed
cracks not indicated).

Figure 6.68 – Cracks under the applied load in the negative direction for SW6 (closed
cracks not indicated).

6-53
Figure 6.69 – The specimen SW6 after testing.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure 6.70 – Load-strain response of strain gauge 9 for the SW6.

6-54
The maximum strain recorded in the edge elements was 1250 μ ε in strain gauge 9
(Figure 6.70) with the behaviour being close to linear.

Other than strain gauge 15, the strain gauges of the web wall did not record strains
greater than the yield. For gauge 15, attached to the longitudinal reinforcement,
maximum strain was 3830 μ ε at the last excursion just after the failure (Figure 6.71).
The maximum strain recorded in the transverse direction was 1800 μ ε in the strain
gauge 17 (Figure 6.72).

Displacements of Web Wall and Edge Elements

The response observed for the both diagonal and edge element LVDTs were similar to
those observed in the previous tests and are shown in (Figures 6.73 and 6.74). The
peak displacement across the diagonal was -2.5 mm (LVDT L2). In the edge element
the displacement at failure was 1.0 mm (LVDT L5).

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000 5000

Strain (μm/m)

Figure 6.71 – Load-strain response of strain gauge 15 for the SW6.

6-55
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure 6.72 – Load-strain response of strain gauge 17 for the SW6.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.73 – Load-displacement response of LVDT L2 for the SW6.

6-56
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-12 -8 -4 0 4 8 12

Displacement (mm)

Figure 6.74 – Load-displacement response of LVDT L5 for the SW6.

6.3 Out-of-Plane Displacement

The LVDT L6 was installed on the top slab to monitor the out-of-plane displacement
of the test specimens during testing. The plots of results are given in the Appendix B.
From the plots, it is observed that the out-of-plane effect for specimens was not
significant. The maximum twisting was 0.00386 radians measured for SW3 at a load
of 1107 kN and occurring during the last cycle in the positive direction. Although for
most of specimens, the response curves were approximately symmetrical around the
zero-load position, for the SW3 the larger displacements were recorded as the load
increased indicating a slight instability against the twisting. However, the magnitude
of twisting corresponding to the failure was not significant.

6-57
6.4 Summary of Test Results

Six wall specimens SW1 to SW6 were tested under lateral reversed cyclic loading.
Five of the specimens also had an axially applied load of 1200 kN. The lateral load
was imposed using displacement control in 4-mm increments with two repetitions
applied for each displacement level.

Wall SW4, with no axial load, exhibited some ductility in the positive loading
direction, although the response in the negative direction was more brittle. For the
other specimens, the responses in the both loading directions were non-ductile and
without significant energy dissipation. The difference in behaviour between specimen
SW4 and the others is directly attributed to the applied axial load. A summary of the
results for the load-displacement response for each specimen is given in Table 6.1.

In general, the crack patterns in the web of each wall were similar with diagonal shear
cracks uniformly spaced throughout the web. These cracks started from the bottom
corner of wall and reached diagonally to the top corner of the opposite side in both
directions. The orientation of diagonal cracks was generally about 45 to 50 degrees
with respect to the bottom slab. At failure, some diagonal cracks extended into the
edge elements resulting in severe damage. Accompanying the shear cracks, some
minor flexural cracks were observed in the edge elements near the bottom slab.

Except for wall SW4, failure was sudden and brittle with failure of all walls (including
SW4) in a shear mode. After failure, the wall stiffness decreased significantly and the
specimens showed little residual strength (except for SW4). Except for specimen
SW2, some of the transverse reinforcing bars had fractured before the completion of
the test.

Prior to the failure, the strains recorded by the strain gauges placed on the longitudinal
reinforcement of the edge elements as well as walls were less than the yield strain. In
contrast, the strains measured in the transverse reinforcement of web, particularly near
the mid-height of wall and near the bottom slab, yielded before failure.

6-58
Table 6.1 – Summary of experimental results
Reinforcement Ratio Peak Lateral Load Corresponding Displacement
Concrete
(%) Axial Load (kN) (mm)
Specimen Strength
(kN)
(MPa) Transverse Longitudinal Downward Upward Downward Upward

SW1 86 0.5 0.5 1200 -1007 992* -11.99 11.93

SW2 86 1.3 1.3 1200 -1192 1190* -17.53 12.52

SW3 96 0.8 0.8 1200 -1090 1107* -12.12 12.31

SW4 96 0.8 0.8 0.0 − 683* 753 -7.94 12.32

SW5 83 0.5 1.3 1200 − 1134* 1073 -10.41 13.51

SW6 83 1.0 1.0 1200 -1101 1141* -16.82 14.01

Note: * The direction of first failure.

6-59
The general response of wall and edge elements recorded by LVDTs was stiff and linear
without significant energy dissipation. Only for the LVDT installed diagonally on the
SW4, was some ductility observed.

6-60
CHAPTER 7

ANALYSIS OF EXPERIMENTAL RESULTS

7.1 Introduction

This chapter contains a discussion of experimental results of the test specimens. The
influence of test variables on the wall response is discussed. To study the effects of
cyclic and monotonic loading, the test results are compared to the those of selected
walls tested by Gupta and Rangan (1996, 1998). The failure load of walls is also
compared to that calculated by the provisions of design codes (AS 3600, 2001, ACI
Committee 318, 2002) as well as the strut-and-tide model proposed by Rangan (1997).

7.2 Effects of Experimental Parameters

7.2.1 Transverse and Longitudinal Reinforcement

Specimens SW1, SW3, SW6 and SW2 were subjected to 1200 kN of axial load and had
the reinforcement ratios of 0.5, 0.8, 1.0 and 1.3 percent equivalent to the ρ f sy of 2.7, 4,

5 and 6.5 MPa, respectively, in both transverse and longitudinal directions, respectively.
Other parameters including reinforcement details in the edge elements as well as the top
and bottom slabs were kept the same. The failure loads of these walls were respectively
measured as 992, 1107, 1141 and 1190 kN.

Figure 7.1 shows the failure load versus ρ f sy for these walls. It can be seen that an

increase in the reinforcement ratio increased the shear strength. The strength of the SW2
with ρ f sy = 6.5 MPa (i.e. reinforcement ratio of 1.3%) was 20% higher than that of wall

SW1 with ρ f sy = 2.7 MPa (i.e. the reinforcement ratio of 0.5%). The rate of increase in

the shear strength prior to walls with ρ f sy < 4 MPa (i.e. 0.8%) was greater than

for ρ f sy > 4 . A linear relationship between lateral failure load and reinforcement ratio in

the range of 4 to 6.5 MPa (i.e. 0.8% to 1.3%) was observed.

Although the walls SW3 and SW6 had the concrete strength of 96 MPa and 83 MPa,
respectively, the shear strength of SW6 was higher than SW3. This implied that the
influence of reinforcement ratio on the strength of these walls was more significant than

1300
Lateral Failure Load (kN)

1200 SW2

SW6
SW3
1100

SW1
1000

900
0 2 4 6 8

ρ fsy (MPa)

Figure 7.1 – Relationship between the lateral failure load and reinforcement ratio.

7-2
that of the concrete strength. This argument is also applied for the walls SW1 and SW2
(with a concrete strength of 86 MPa).

To investigate the deformation characteristics of the wall response, the envelope curves
for the cyclic response of the walls SW1 to SW3 and SW6 have been illustrated in
Figure 7.2. Envelope curves had a similar trend approximately, with the exception for
the displacement corresponding to the failure load. In the positive loading direction wall
SW2 showed slightly a higher stiffness compared to the other specimens, whereas in the
negative loading direction this occurred for the SW3. This seems to be due to the test
method used in this research. In general, the difference was not very considerable and it
can be concluded that the reinforcement ratio did not affect the stiffness of the test walls
significantly. A similar trend was observed in the specimens S-3 and S-6 tested
monotically by Gupta and Rangan (1996). Walls S-3 and S-6 were subjected to an axial
load of 1230 kN with the longitudinal reinforcement ratio of 1.06% and 1.61%,
respectively. Other test parameters for these specimens were the same. It was observed
that the load-displacement responses for the specimens were very similar.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600 SW1 (ps=0.5%)
-800 SW3 (ps=0.8%)
-1000 SW6 (ps=1.0%)
SW2 (ps=1.3%)
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure 7.2 – Envelope curves for the specimens SW1 to SW3 and SW6.

7-3
The displacement corresponding to the failure load in the negative loading direction for
the specimens SW2 and SW6 was quietly greater than that for the SW1 and SW3. In
comparison, this difference was smaller in the positive loading direction. This was
attributed to the higher strength of the SW2 and SW6.

While the SW2 failed due to a sliding shear plane that was evident in the web wall near
the base, for other specimens the diagonal compression failure occurred. Moreover,
except for the SW2, in the other specimens some transverse reinforcement crossing the
failure surface of web wall fractured. Considering that the SW2 had the highest amount
of transverse reinforcement among all specimens, the failure mode in the HSC squat
shear walls cyclically loaded can be considered to be significantly dependant on the
transverse reinforcement amount. This is in agreement with the findings on NSC shear
walls (Paulay and Priestley, 1992).

The hysteresis aspects for SW1, SW2 and SW6 demonstrated similarities (see
Chapter 6). Cyclic responses were without pinching effect and with very low energy
dissipation. The SW3 response had a slight pinching that seems to be due to testing.
However, similar to other previously mentioned specimens, the energy dissipation was
low. The load-displacement curves for the specimens showed that prior to failure, there
was a slight difference between the loading paths of the first and second cycle at each
displacement level caused by cyclic damages. Also, the general trend for unloading and
reloading characteristics for the specimens was similar.

The SW1 and SW5 had the longitudinal reinforcement ratios of 0.5% and 1.3%,
respectively, and other parameters were the same. The comparison of envelope curves
of their cyclic responses (Figure 7.3) shows the failure load of SW5 was about 14%
higher than that of SW1. As a result, the wall strength was influenced by the amount of
longitudinal reinforcement ratio so that it increased when this ratio increased. Besides,
the trends of curves in the positive loading direction were similar, while in the negative
loading direction a stiffer response for the SW5 in the last cycles was observed.

7-4
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000 SW1
SW5
-1200
-1400
-16 -12 -8 -4 0 4 8 12 16

Displacement (mm)

Figure 7.3 - Envelope curve of cyclic response for the specimens SW1 and SW5.

The hysteresis aspects for the SW5 demonstrated similarities with the SW1. The cyclic
response of SW5 was without pinching effect and with the low energy dissipation.
Similar to the SW1, SW5 failed in a diagonal compression mode. It was concluded that
the longitudinal reinforcement ratio did not significantly affect the cyclic response or
the failure mode.

The envelope curves of the response of the specimens SW2 and SW5 have been shown
in Figure 7.4. Walls SW2 and SW5 had transverse reinforcement ratios of 1.3% and
0.5%, respectively, with the other parameters kept the same. In comparison to the SW5,
wall SW2 showed slightly a stiffer response in the positive loading direction. In the
negative loading direction, however, it showed a softer response but with a similar
strength. In other words the response of SW2 was more ductile than that of SW5. This
behaviour was apparently due to the higher volume of transverse reinforcement in the
SW2.

7-5
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000 SW2
SW5
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20
Displacement (mm)

Figure 7.4 - Envelope curve of cyclic response for the specimens SW2 and SW5.

7.2.2 Axial Load

Specimens SW3 and SW4 had the same reinforcement details but were subjected to
axial loads of 1200 kN and 0 kN, respectively. The failure loads for these walls were
1107 and 683 kN, respectively, that is wall SW3 was 62 percent stronger than SW4.
Gupta and Rangan (1996) reported that the increase in shear capacity for HSC shear
wall loaded monotically was about 30% of the increase in the axial load on the wall.
The observations from these tests are consistent with those of Gupta and Rangan.

7-6
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
SW3
-1200 SW4
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure 7.5 – Envelope curve of cyclic response for the specimens SW3 and SW4.

The response of SW4 was softer than that of SW4 (Figure 7.5) and showed a more
ductile trend in the positive loading direction. In addition, while wall SW3 had a brittle
behaviour and failed in a shear mode in both loading directions, the failure of the SW4
was brittle in the negative loading direction while more ductile in the positive direction.
Since the other parameters were kept the same, the axial load applied on the SW3 can
be concluded to be the main reason for this more brittle behaviour as well as higher
strength.

7.3 Test Results vs. those of Gupta and Rangan (1996)

Since the specimens tested in this study are geometrically similar to those of Gupta and
Rangan (1996) and the strengths of concrete used in both tests are fairly close, a
comparative study between some selective walls can be undertaken to investigate the
effects of load reversal.
7-7
Table 7.1 shows three comparative groups. The first group consists of the wall S-1 of
Gupta and Rangan and the walls SW2 and SW5. The load-displacement envelope
curves of wall responses are illustrated in Figure 7.6. While during early loading stages
in the positive loading direction the curves of the S-6 and SW2 are similar, the SW5
showed a softer response. For early loading stages in the negative loading direction,
however, both SW2 and SW5 presented a softer response compared with the Gupta
specimen S-6. Since the concrete strength of SW2 and SW5 is higher than that of the
S-1, the softer response of these walls seems to be caused by the cyclic damage. Wall
S-6 failed in a lower load compared to the SW2 and SW5 showing the greater shear
capacity for the latter due to the higher concrete strength.

In second group, the specimen S-1 of Gupta and Rangan is compared with the SW4,
where both walls did not have axial load. The load-displacement envelope curve of the
walls shows a similar stiffness before the displacement of 4 mm which is the
displacement level of first loading phase of SW4. In other words no cyclic damage was
experienced by the SW4 at this stage.

Table 7.1 – Comparison of test results with those of Gupta and Rangan (1996)

Comparative Specimen f c'


Axial ρ h f syh ρ v f syv Failure Corresponding
Load Load Displacement
Group (MPa) (MPa) (MPa)
(kN) (kN) (mm)

S− 6* 71 1230 2.9 8.0 970 13.0


1 SW2 86 1200 6.5 6.5 1190 12.5
SW5 83 1200 2.7 6.5 1134 10.4
S− 1* 79 0 2.9 5.5 426 23
2
SW4 96 0 4.3 4.0 683 7.9
S− 3* 69 1230 2.9 5.5 851 6.5
3 SW1 86 1200 2.7 2.7 992 11.9
SW6 83 1200 5.0 5.0 1141 14.0

Note: * Test specimens of Gupta and Rangan (1996).

7-8
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800 S-6
-1000 SW5
-1200 SW2
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20
Displacement (mm)

Figure 7.6 – Load-displacement envelope curve of the walls S-6, SW2 and SW5.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000 S-1
-1200 SW4
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24
Displacement (mm)

Figure 7.7 – Load-displacement envelope curve of the walls S-1 and SW4.

7-9
After failure of S-1, its response was in a ductile manner similarly with the response of
SW4 in the positive loading direction. The strength of SW4 was, however, higher than
that of S-1 due to the effect of concrete strength.

In the third group the wall S-3 of Gupta and Rangan is compared with the walls SW1
and SW6. Figure 7.8 shows the response of these walls. Although the concrete strength
of SW1 and SW6 are higher than that of S-3, the response of these walls is softer than
the S-3. However, the responses of SW1 and SW6 were also softer for the
displacements less than 4 mm (before cycling) and, therefore, the softer response of
SW1 and SW6 seems to be due to not only cyclic damage but probably also due to
difference in ρ f sy for the longitudinal reinforcement. The strength and corresponding

displacement for the SW1 and SW6 are greater than S-3 due primarily to the higher
concrete strength.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800 S-3
-1000 SW1
-1200 SW6
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20
Displacement (mm)

Figure 7.8 – Load-displacement envelope curve of the walls S-3, SW1 and SW4.

7-10
7.4 Comparison of the Wall Strengths with Design Code Predictions

In this section, the test failure loads are compared with the calculated strengths based on
the Australian Standard for Concrete Structures (AS 3600, 2001) and American
Building Code for Concrete Structures (ACI Committee 318, 2002). The requirements
of both codes for flexural and shear design of shear walls have been described in the
Appendix F.

The calculated shear and flexural strengths for the specimens SW1 to SW6 are
presented in Table 7.2. The flexural strengths calculated based on both codes are the
same. The shear strengths calculated by the codes, however, are different. Since the
flexural strengths of the walls are higher than the corresponding shear strengths, both
codes predict that the walls fail in a shear mode which is in agreement with the test
results.

The ratios of the test failure loads over the calculated shear strengths by both codes for
the test specimens are given in Table 7.3. The mean value and standard deviation of
ratios for AS 3600 are 2.16 and 0.54, respectively, representing an under calculation of
the actual wall strengths. For ACI 318 considering the limit of maximum shear strength
of shear walls, the mean and standard deviation of ratios are 2.31 and 0.28, respectively,
whereas without considering this limit the values are 1.82 and 0.21, respectively. These
quantities show underestimated results especially when the maximum shear strength
limit is considered. The maximum shear strength limit included in the ACI 318 code is
based on the tests carried out by Cardenas et al. (1980) on the 13 shear walls with the
concrete strength of 39 to 52 MPa and with the height-to-width ratios of 1, 1.9 and 3.4.
This limit, however, seems to be a conservative estimation of maximum shear strength
for the HSC low-rise shear walls and needs to be enhanced. For the walls SW1 to SW3
and SW6 with a similar axial load and the same amount of reinforcement in both the
transverse and longitudinal directions, the increase in the reinforcement ratio resulted in
a decrease in the test/code ratio by AS 3600 and ACI 318 without considering the
maximum limit of shear strength and, inversely, an increase in this ratio by ACI 318
including the maximum limit (Figure 7.6). Gupta and Rangan (1996) reported a similar
trend for the test/code ratios of shear walls loaded monotically.

7-11
Table 7.2 – Comparison of test failure loads with predicted loads by codes

Calculated Shear Strength


(kN)
Failure Calculated
Load AS 3600 ACI 318 Flexural
Specimen (2001) (2002)
from Test Strength
(kN) Considering Not considering (kN)
maximum shear maximum shear
strength limit strength limit
SW1 992 395 462 524 1633
SW2 1190 651 462 780 1726
SW3 1107 506 488 629 1673
SW4 683 506 369 369 1200
SW5 1134 391 454 522 1725
SW6 1141 526 454 657 1687

Table 7.3 – Correlation of test loads and predicted shear strength by codes

Test Failure Load / Calculated Shear Strength

ACI 318
Specimen (2002)
AS 3600
(2001) Considering Not considering
maximum shear maximum shear
strength limit strength limit
SW1 2.51 2.15 1.89
SW2 1.83 2.58 1.53
SW3 2.19 2.27 1.76
SW4 1.35 1.85 1.85
SW5 2.90 2.50 2.17
SW6 2.17 2.51 1.74
Mean 2.16 2.31 1.82
Standard Deviation 0.54 0.28 0.21

While the relationships for calculation of the shear strength of shear walls given in
ACI 318 include the effect of axial load, those of AS 3600 do not. Comparison of the
test/code ratios by the codes (Table 7.3) shows that the shear strength of specimens with
the axial load, estimated by ACI 318 without considering the maximum limit of shear
strength, is closer to the test results. The best estimation is for the strength of the wall

7-12
SW2 with a test/code ratio of 1.53. The highest test/code ratio, however, is for the wall
SW5 with a value of 2.17. Since the ratio of transverse reinforcement for the SW2 and
SW5 (1.5% and 0.5%, respectively) is the main difference between the walls, it seems
the estimation of the strength for walls with different amounts of reinforcement in
transverse and longitudinal directions by ACI 318 (as well AS 3600) is more
conservative. On the other hand, since the requirements of AS 3600 for the calculation
of the wall strength do not include the effect of axial load, for wall SW4 with the zero
axial load the estimation of AS 3600 is better than ACI 318 (test/code ratios of 1.35 and
1.85, respectively). As a result, the enhancement of relationships on shear walls by
AS 3600 to include the effect of axial load and by the both codes to address the
difference between the amounts of reinforcement in transverse and longitudinal
directions is essential.

3
SW1 SW3 SW6 SW2
Test Failure Load /Calculated Load

2.5

AS 3600
1.5
ACI 318 (considering the maximum shear strength )

ACI 318 (not considering the maximum shear strength )

1
0 2 4 6 8

ρ fsy (MPa)

Figure 7.9 – The relationship between ρ f sy reinforcement ratio and test/code ratio.

7-13
7.5 Simplified Strut-and-Tide Model by Rangan (1997)

The test results are compared to the predicted strength of the walls using a simplified
strut-and-tide model developed by Rangan (1997) with the results presented in Table
7.3. The model, expressed in the Appendix G, was verified against a range of
experimental data conducted on the NSC as well as HSC shear walls subjected to
monotonic loading (Barda et al., 1977, Cardenas et al., 1980, Maier and Thurlimann,
1985, Oesterle, 1986, Lefas, 1988, Gupta and Rangan, 1996, Kabeyasawa and Hiraishi,
1998) with a reasonable agreement with the test results (Rangan, 1997). For the
specimens, tested in this study under reversed cyclic loading, the mean and standard
deviation values for the test/calculated ratios are 1.01 and 0.15, respectively. In general,
there is a good agreement between the results of test and those of the model.
Nevertheless, the strength of the SW4 (with no axial load) is unsafely calculated with a
test/calculated ratio of 0.76. Consequently, the model needs to be corroborated against
more experimental results for shear walls tested under reversed cyclic loading and , in
particular, with low axial loads.

Table 7.4 - Correlation of test loads and predicted shear strength by strut-and-tide model

Calculated Shear
Failure Load Test Failure Load
Strength by
from Test Strut-and-Tide Model
Specimen Calculated Shear Strength
(kN)
(kN)
SW1 992 1044 0.95
SW2 1190 1044 1.14
SW3 1107 1148 0.96
SW4 683 901 0.76
SW5 1134 1013 1.12
SW6 1141 1013 1.13
Mean 1.01
Standard Deviation 0.15

7-14
7.6 Finite Element Analysis of Shear Wall Specimens

7.6.1 Two Dimensional Finite Element Analysis

Shear wall specimens SW1 to SW6 were analysed using the two dimensional
finite element model described in Chapter 3. The material properties and
reinforcement details for the specimens are given in Table 7.5. The 2D finite
element mesh, shown in Figure 7.10, consisted of 182 four-node isoparametric
plane stress elements. The reinforcement steel in the web wall and edge elements
was modelled as smeared. The top and bottom slabs were modelled as stiff
elements and the bottom slab was assumed fully fixed at the base. A constant
axial load of 1200 kN distributed over the top slab was applied (except for the
wall SW4) and the lateral load was incrementally increased until the failure.

The observed and calculated envelope curves for the load-displacement responses of the
test specimens SW1 to SW6 are shown in Figures 7.11 to 7.16, respectively. Also, the
calculated results for the failure loads and the corresponding displacements for the shear
walls are given in Table 7.6. As wall SW4 exhibited two different behaviours in the
negative and positive loading directions, brittle and ductile, respectively, the data for
both cases is shown in Table 7.6.

A comparison between the experimental and calculated results shows that the 2D finite
element model gives a good prediction of the failure loads with a mean of the ratio of
the test data over the calculated results of 0.93 and standard deviation of 0.05. For the
wall SW4 with no axial load, the ratios in the negative and positive loading directions
are 0.71 and 0.64, respectively, showing the severe cyclic damage compared with the
monotonic loading. In other words, the presence of axial load appears to reduce the
damage attributed to cyclic loading.

7-15
Table 7.5 – Material properties of shear wall specimens

Shear Wall SW1 SW2 SW3 SW4 SW5 SW6

Concrete
f c' (MPa) 86 96 83
ε co 0.0025
Ec (MPa) 37700 43700 39600
web wall reinforcement in transverse direction
f yt (MPa) 536 498 536 536 536 498
ρ st 0.005 0.013 0.008 0.008 0.005 0.01
web wall reinforcement in longitudinal direction
f yl (MPa) 536 498 498 498 498 498
ρ sl 0.005 0.013 0.008 0.008 0.013 0.01
reinforcement in edge elements
f yl (MPa) 535
ρ sl 0.064

7-16
Figure 7.10 – 2D finite element mesh used to model the shear wall specimens.

The comparison of the test and calculated displacements, however, shows that the 2D
finite element model greatly under predicts the lateral drifts of the test shear walls. The
mean of the ratios of the test over calculated displacements corresponding to the failure
loads is 2.42 with a standard deviation of 0.08.

7-17
1400

1200

1000
Lateral Load (k

800 Negative Direction


Positive Direction
FEA
600

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.11 – 2D numerical results versus experimental envelope curve for SW1.

1400

1200

1000
Lateral Load (k

800 Negative Direction


Positive Direction
600 FEA

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7. 12 – 2D numerical results versus experimental envelope curve for SW2.

7-18
1400

1200

1000
Lateral Load (k

800 Negative Direction


Positive Direction
600 FEA

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.13 – 2D numerical results versus experimental envelope curve for SW3.

1400

1200

1000
Lateral Load (k

800 Negative Directiont


Positive Direction
FEA
600

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.14 – 2D numerical results versus experimental envelope curve for SW4.

7-19
1200

1000

800
Lateral Load (k

Negative Direction
600 Positive Direction
FEA

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.15 – 2D numerical results versus experimental envelope curve for SW5.

1400

1200

1000
Lateral Load (k

800 Negative Direction


Positive Direction
600 FEA

400

200

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.16 – 2D numerical results versus experimental envelope curve for SW6.

7-20
Table 7.6 – 2D numerical results versus test data

Displacement Corresponding to
Failure Load (kN)
Failure Load (mm)
Specimen
Test Analysis Test /Analysis Test Analysis Test /Analysis

SW1 992 1050 0.95 11.9 4.6 2.59


SW2 1190 1200 0.99 12.5 5.7 2.19
SW3 1107 1200 0.92 12.3 4.7 2.62
SW4* 683 1060 0.64 7.9 5.4 1.46
SW4** 753 1060 0.71 12.3 5.4 2.28
SW5 1134 1080 1.05 10.4 4.5 2.31
SW6 1141 1190 0.96 14 5.5 2.55
* 0.92 2.23
Mean
** 0.93 2.42
* 0.06 0.20
Standard Deviation
** 0.05 0.08

* Negative loading direction for SW 4


* * Positive loading direction for SW 4

Foster and Rangan (1999) reported a similar under prediction of displacements when
they analysed the squat HSC shear walls of Gupta and Rangan (1996) using a 2D finite
element model. They concluded that this problem is due to an over calculation of the
stiffness contribution of the edge elements to the lateral displacement of the walls.
When a 3D finite element model was used to analyse the walls, the calculated results
showed vastly improved correlations to the test data both for the failure loads and the
load-displacement response (Foster and Rangan, 1999). The same problem has been
reported for the 2D cyclic finite element analysis of large-scale squat shear walls under
monotonic loading (Vecchio, 1998) and cyclic loading (Palermo and Vecchio, 2004).
Generally, concentrating the full width of the edge elements into a single finite element
in the 2D has some significant implication as follows from Foster and Rangan (1999)
and Palermo and Vecchio (2004):

7-21
• The very stiff edge elements are assumed fully connected to the web elements.
Thus, the degree of lateral and vertical confinement they provide to the web may
be overestimated.

• The shear lag effect that occurs in the out-of-plane direction is not considered in
the model.

• The ability of the edge elements to carry a lateral shear can be overestimated
when full fixity to the web is assumed.

These factors can contribute to overestimating the stiffness of shear walls when
conducting a 2D finite element analysis.

In general, it can be concluded that the analysis of the shear walls tested in this study
using the developed 2D finite element model, could capture the failure load as well as
the failure mechanism, but overestimated the wall stiffness. To capture the wall stiffness
and the wall displacements under loading a 3D finite element model is required.

7.6.2 Three Dimensional Finite Element Analysis

To compare the calculated results of the 2D and 3D finite element analyses, the shear
wall specimens SW1 to SW6 were modelled using a 3D finite element model developed
by Foster and Rangan (1999). Details of the 3D FE formulation are given in Foster and
Rangan (1999). The 3D mesh (shown in Figure 7.17) consisted of 40 non-linear
20-node isoparametric brick elements to model the edge elements and web wall, 96
linear-elastic brick elements to model the top and bottom slabs and 119 by 3-node bar
elements to model the reinforcing steel. Considering the symmetry of walls, to reduce
computation time only one half the shear walls was modelled. The walls were analysed
under a constant 1200 kN axial load (600 kN for the ½ wall modelled) and
incrementally loaded laterally until failure.

7-22
Figure 7.17 – 3D finite element mesh used to model the shear walls.

Figures 7.18 to 7.23 show the calculated load-displacement response versus the
envelope curve of the experimental data. The distribution of stresses in the other walls
was similar to that of the wall SW4 shown in Figure 7.24. Extensive concrete crushing
was calculated in the compression region located in the bottom corner of the walls. This
was consistent with the observed failure mechanism except for the wall SW2 which
exhibited a shear sliding failure. The 3D analysis results are in reasonable agreement
with the failure loads and a better predict the displacements corresponding to the failure
loads than for the 2D analyses. The mean of the ratios of the test data over the
calculated results for failure loads is 0.99 with a standard deviation of 0.06. The mean
of the ratios of the test over calculated displacements corresponding to the failure loads
is 1.37 with a standard deviation of 0.13. Comparison of means for the 2D and 3D
analyses shows that the both models result in a good estimation of failure loads but only
the 3D model is capable of capturing the displacements. Also, similar to 2D analyses, in
the 3D modelling the failure mechanisms for the wall specimens were predicted
correctly compared with the test observations.

7-23
Before first cracking in the walls, the 3D finite element model provided a stiffer
response compared to the observed response. After cracking, however, the stiffness of
wall responses was calculated very well. In other words the contribution of edge
elements in the whole stiffness of the shear walls has been predicted well. The over-
calculations of the initial stiffness suggests some support rotation (or movement) as the
test specimen bedded into the testing frame during the initial stages of loading.

The results of both 2D and 3D analyses, similarly to the experimental results, show that
the strength of the HSC shear walls tested in this study does not change significantly for
the variation of the reinforcement ratios considered (i.e. 0.5 to 1.3 percent).

1200

1000

800
Load (kN

600

400
SW1
Negative Direction
200
Positive Direction
FEA
0
0 4 8 12 16 20
Displacement (mm)

Figure 7.18 – 3D numerical results versus experimental envelope curve for SW1.

7-24
1400

1200

1000
Load (kN

800

600

400
SW2
Negative Direction
200 Positive Direction
FEA
0
0 4 8 12 16 20

Displacement (mm)

Figure 7.19 – 3D numerical results versus experimental envelope curve for SW2.

1400

1200

1000
Load (kN

800

600

400 SW3
Negative Direction
200 Positive Direction
FEA
0
0 4 8 12 16 20
Displacement (mm)

Figure 7.20 – 3D numerical results versus experimental envelope curve for SW3.

7-25
1200

1000

800
Load (kN

600

400
SW4
Negatige Dirction
200 Positive Direction
FEA

0
0 4 8 12 16 20
Displacement (mm)

Figure 7.21 –3D numerical results versus experimental envelope curve for SW4.

1200

1000

800
Load (kN

600

400
SW5

200 Negative Direction


Positive Direction
FEA
0
0 4 8 12 16 20
Displacement (mm)

Figure 7.22 – 3D numerical results versus experimental envelope curve for SW5.

7-26
1200

1000

800
Load (kN

600

400
SW6
200 Negative Direction
Positive Direction
FEA
0
0 4 8 12 16 20
Displacement (mm)

Figure 7.23 – 3D numerical results versus experimental envelope curve for SW6.

Figure 7.24 – Distribution of stresses throughout the wall SW4.

7-27
Table 7.7 - 3D numerical results versus test data

Displacement Corresponding to
Failure Load (kN)
Failure Load (mm)
Specimen
Test Analysis Test /Analysis Test Analysis Test /Analysis

SW1 992 1009 0.98 11.9 7.7 1.55


SW2 1190 1236 0.96 12.5 9.6 1.30
SW3 1107 1174 0.94 12.3 9.2 1.33
SW4* 683 782 0.87 7.9 10.6 0.75
SW4** 753 782 0.96 12.3 10.6 1.27
SW5 1134 1021 1.11 10.4 8.2 1.27
SW6 1141 1134 1.01 14 9.2 1.52
* 0.98 1.29
Mean
** 0.99 1.37
* 0.08 0.29
Standard Deviation
** 0.06 0.13

* Negative loading direction for SW 4


* * Positive loading direction for SW 4

7-28
CHAPTER 8

CONCLUSIONS

8.1 Summary

The main aim of this study is the investigation of the strength and ductility of
high-strength concrete shear walls subjected to reversed cyclic loading with emphasis
on low rise shear walls.

The analytical, numerical and experimental research related to the behaviour of shear
walls under monotonic and cyclic loading were reviewed. Based on the elasticity-based
equivalent uniaxial stains approach, a variety of constitutive models for concrete, steel
reinforcement and the interaction between them have been reviewed. These models can
be successfully implemented into a finite element procedure to analysis of reinforced
concrete structures, if they capture major aspects of their behaviour such as non-linear
stress-strain relationships of concrete and steel, concrete cracking, compression
softening, tension stiffening and so on. Finite element methods based on the smeared
crack model have been commonly used to analyse shear walls. Two major approaches
in this area are fixed and rotating crack model. While the former is a more general
model and has been used by most researches for analysis of shear walls, the latter is
simpler and can be implemented easily. The experiments carried out on normal strength
concrete shear walls have shown the effects of major parameters such as longitudinal
and transverse reinforcement ratios, axial load, the ratio of height-to-length, edge
elements and loading regime on the behaviour of walls. Some of these parameters have
been also investigated by a few researchers that conducted tests on high-strength
concrete shear walls under monotonic loading. However, experimental data on the
behaviour of high-strength concrete shear walls under reversed cyclic loading is very
limited. A research project consisting of an experimental and an analytical program was
carried out to fulfil this requirement.

An experimental program consisting of six wall specimens was undertaken. The scale of
test specimens was one-third of shear walls used in a multistorey building. The walls,
cast with concrete strengths ranging from 80 to 90 MPa, had the same geometrical
properties and were made up the web wall, edge elements, and stiff top and bottom
slabs. The height, length and thickness of walls were 1000 mm, 1000 mm and 75 mm,
respectively. The test parameters were longitudinal reinforcement and transverse
reinforcement ratios (0.5, 0.8, 1.0 and 1.3 percent) and axial load (0 and 1200 kN). The
wall specimens were tested under inplane constant axial load and cyclic incremental
displacements of 4 mm until the failure. Axial load was imposed on each wall using six
prestressed MacAulay bars anchored in the top and bottom slabs. The results of tests
including the cyclic load-displacement response of the walls, the monitored axial loads,
the strains in the reinforcement in edge element as well as the wall, the longitudinal
dilation of edge elements and diagonal dilation of web walls were presented. These
results provide useful data to investigate the effects of test parameters, to investigate
cyclic aspects of wall responses and to corroborate future finite element modelling.

Based on the crack membrane model (Kaufmann and Marti, 1998), a two dimensional
smeared-fixed crack finite element model was developed. The 2D model, however, was
found to predict an overly stiff response. A significant contribution, nevertheless, of the
analytical work was in developing a new aggregate interlock model. The model is
simple and can be easily incorporated into a finite element procedure. It has been
successfully verified against shear panel tests. While aggregate interlock models
commonly concern with plain concrete cracks, the proposed model deals with cracks in
reinforced concrete so that the effect of reinforcement crossing the crack interface on
the aggregate interlock phenomenon has been quantified and incorporated into the
model. The model also covers both normal and high strength concrete. Incorporating the

8-2
proposed model into the crack membrane model leads to a more general tool to analysis
reinforced concrete structures.

8.2 Concluding Remarks

Based on the experimental work presented in this thesis, the following conclusions are
made:

• Imposed axial load significantly influenced the ultimate strength and ductility of
low-rise high-strength concrete shear walls. Presence of axial load increased the
wall strength but resulted in a brittle response with almost no ductility for walls.
Whilst this conclusion is consistent with the findings of Mayer and
Thurlimann (1985) and Gupta and Rangan (1996, 1998) that used squat wall
specimens as same as this research, it is inconsistent with the findings of tests
carried out on cantilever walls by Oesterle et al (1981). The behaviour of squat
shear walls is governed by shear action and concrete response. In contrary, the
behaviour of cantilever shear walls is dominated by flexural action and
reinforcement response. The increase of axial load in cantilever shear walls
causes a decrease of stresses in the tensile reinforcement and therefore the
increase of capacity of walls to carry further lateral load before reinforcement
yielding.

• An increase in both longitudinal and transverse reinforcement ratios caused an


increase in the wall strength but did not affect the brittle response observed for
them. For the walls with higher amounts of reinforcement, the deformation was
slightly greater than that of walls with lower reinforcement ratios.

• An increase in only longitudinal reinforcement ratio caused an increase in wall


strength. Test results indicated that an increase of 160% in the longitudinal
reinforcement ratio resulted in an increase of failure load of about 14%. The low
increase of strength despite a very large increase in reinforcement ratio may be
attributed to the amounts of reinforcement used in this study (0.5% to 1.3%) so
8-3
that their effects in shear walls with a high concrete strength is not as significant
as those in normal strength concrete shear walls.

• An increase in the transverse reinforcement ratio did not affect the wall strength
considerably, but caused an increase in the wall deformation at failure.

• The failure mode of the walls was influenced by the transverse reinforcement
ratio. While the wall SW2 with the maximum transverse reinforcement ratio of
1.3 percent failed in a sliding shear mode, other shear walls failed in a diagonal
compression mode.

• The stiffness of shear walls with different reinforcement ratios but same axial
load was generally close to each other. The stiffness of walls tested seems to be
strongly dominated by the stiffness of edge elements rather than the wall
reinforcement.

• Comparison of results for shear walls tested by Gupta and Rangan (1996) under
monotonic loading with the results of this study indicated a slightly softer
response for the latter case attributed to the damage that occurred in the walls
tested under cyclic loading.

• The enhancement of relationships on shear walls by AS 3600 to include the


effect of axial load and by AS 3600 and ACI 318 codes to address the difference
between the amounts of reinforcement in transverse and longitudinal directions is
required.

• Although there is a good agreement between the results of test and those of the
strut-and-tie model of Rangan (1997), the model needs to be corroborated against
more experimental results for shear walls tested under reversed cyclic loading
and , in particular, with low axial loads.

Based on the finite element analysis, the following considerations are derived:

8-4
• A 2D smeared-fixed crack finite element model based on the crack membrane
model of Kaufmann and Marti (1998) can be used to model panels in shear but
where 3D effects are important (such as out-of-plane stiffening elements) 3D
models are needed.

• Generally, the shear wall strength is calculated well using the 2D finite element.

• 2D finite element analysis assuming fully effective edge elements, over predicts
the wall stiffness. As a result, the displacement corresponding to the failure load
is calculated smaller than real one. To calculate the displacement as well as
ductility of squat shear wall a 3D finite element analysis is needed to be applied.

• In comparison to the 2D finite element model, the 3D finite element analysis


provides a better calculation of the wall displacements and stiffnesses.

8.3 Recommendations for Future Study

Further experimental testing should address the following research needs:

• Reinforcement ratios ranged from 0.5 to 1.3 percent did not significantly affect
the general cyclic behaviour of high-strength concrete shear walls. Experimental
data is required to investigate the effects of reinforcement ratios higher than 1.3
percent in the response of walls subjected to cyclic loading.

• Considering the brittle failure of walls with low amount of transverse


reinforcement ratios, experimental program to determine the minimum
requirements of transverse reinforcement in high-strength shear wall needs to be
carried out.

• Experimental tests consisting of high-strength concrete shear wall specimens


with diagonal wall reinforcement is required to investigate its effect on the cyclic
behaviour and particularly the ductility of walls.

8-5
• In the aggregate interlock model proposed in this study, the effect of
reinforcement crossing the crack interface has been quantified based on the two
parameters of reinforcement ratio and concrete strength due to limited data
available in the literature for cracks in reinforced concrete. To reach a more
detailed aggregate interlock model with other parameters such as the
reinforcement yield stress, more experimental tests on the cracks with crossing
reinforcement are needed.

• More data is required to refine the proposed aggregate interlock model for
high-strength concrete and investigate the degradation of the shear stress at
cracks.

The finite element model developed in this study may be extended in the following
areas:

• The presence of edge elements significantly affect the results obtained from
analysis of wall specimens using two dimensional finite element model. To
obtain more accurate numerical results, the extension of current work to a three
dimensional finite element model is required.

• To simulate the cyclic behaviour of reinforced concrete members, the two


dimensional finite element model presented can be extended to the case of cyclic
loading. For shear walls with stiff edge elements, in particular, a three
dimensional finite element formulation extended to cyclic condition is needed.

• The two dimensional finite element model developed in this study can be
implemented for other structural systems under biaxial states of stress such as
plates, slabs and shells.

8-6
REFERENCES

ACI Committee 318 (1971), "Building Code Requirements for Reinforced Concrete,"
American Concrete Institute, Detroit, Michigan, 78 pp.

ACI Committee 318 (1977), "Building Code Requirements for Reinforced Concrete,"
American Concrete Institute, Detroit, Michigan, 102 pp.

ACI Committee 318 (2002), "Building Code Requirements for Structural Concrete and
Commentary," American Concrete Institute, Farmington Hills, Michigan,
443 pp.

Aktan, H. M. and Hanson, R. D. (1980), "Nonlinear Cyclic Analysis of Reinforced


Concrete Plane Stress Members," in SP-63: Reinforced Concrete Structures
Subjected to Wind and Earthquake Forces, ACI, Detroit, Michigan,
pp. 135-152.

Ali, M. A. and White, R. N. (1999), "Enhanced Contact Model for Shear Friction of
Normal and High-Strength Concrete," ACI Structural Journal, Vol. 96, No. 3,
pp. 348-360.

Angelakos, D., Bentz, E. C. and Collins, M. P. (2001), "Effect of Concrete Strength and
Minimum Stirrups on Shear Strength of Large Members," ACI Structural
Journal, Vol. 98, No. 3, pp. 290-300.

AS 3600 (2001), "Australian Standard for Concrete Structures," Standard Australia


International, Sydney, Australia, 176 pp.
ASCE Committee 447 (1982), "State-of-the-art Report on Finite Element Analysis of
Reinforced Concrete," ASCE, New York.

ASCE Committee 447 (1993), "Finite Element Analysis of Reinforced Concrete


Structures II," ASCE, New York, 717 pp.

Attard, M. M., Nguyen, D. M. and Foster, S. J. (1996), "Finite Element Analysis of Out
of Plane Buckling of Reinforced Concrete Walls," Computers & Structures,
Vol. 61, No. 6, pp. 1037-1042.

Ayoub, A. and Filippou, F. J. (1998), "Nonlinear Finite Element Analysis of Reinforced


Concrete Shear Panels and Walls," ASCE Journal of Structural Engineering,
Vol. 124, No. 3, pp. 298-308.

Bahlis, J. B. and Mirza, M. S. (1987), "Nonlinear analysis of Planar Reinforced


Concrete Structures," Canadian Journal of Civil Engineering, No. 14,
pp. 771-779.

Bahn, B. Y. and Hsu, T. T. C. (1998), "Stress-Strain Behavior of Concrete under Cyclic


Loading," ACI Structural Journal, Vol. 95, No. 2, pp. 178-193.

Balan, T. A., Filippou, F. C. and Popov, E. P. (1998), "Hysteretic Model of Ordinary


and High-Strength Reinforcing Steel," ASCE Journal of Structural Engineering,
Vol. 124, No. 3, pp. 288-297.

Barda, F., M., H. J. and G., C. W. (1977), "Shear Strength of Low Rise Walls with
Boundary Elements," in SP-53: Reinforced Concrete Structures in Seismic
Zones, ACI, Detroit, USA, pp. 149-202.

Barzegar, F. (1995), "Constitutive Models for Nonlinear Finite Element Analysis of


Reinforced Concrete Structures," Report No. R-346, University of New South
Wales, Sydney, 74 pp.

R-2
Bathe, K. J. (1996), "Finite Element Procedures," Prentice Hall, New Jersey, 1037 pp.

Bauschinger, J. (1887), "Variations in the elastic limit of iron and steel [summarised
translation]," Journal of Iron and Steel Institute, Vol. 1, pp. 442-444.

Bazant, Z. P. (1983), "Comment on Orthotropic Models for Concrete and


Geomaterials," ASCE Journal of Engineering Mechanics, Vol. 109, No. EM3,
pp. 849-865.

Bazant, Z. P. and Gambarova, P. G. (1980), "Rough Cracks in Reinforced Concrete,"


ASCE Journal of Structural Division, Vol. 106, No. ST4, pp. 818-842.

Belarbi, A. (1991), "Stress-Strain Relationships of Reinforced Concrete in Biaxial


Tension-Compression," PhD Thesis, Department of Civil Engineering,
University of Huston, Huston, Texas, 471 pp.

Belarbi, A. and Hsu, C. T. T. (1994), "Constitutive Laws of Concrete in Tension and


Reinforcing Bars Stiffened by Concrete," ACI Structural Journal, Vol. 91,
No. 4, pp. 456-474.

Belarbi, A. and Hsu, C. T. T. (1995), "Constitutive Laws of Softened Concrete in


Biaxial Tension-Compression," ACI Structural Journal, Vol. 92, No. 5,
pp. 562-573.

Belytschko, S., Elwi, A. E. and Murray, D. W. (1988), "Effect of Modelling on NLFEA


of Concrete Structures," ASCE Journal of Structural Engineering, Vol. 114,
No. 7, pp. 2323-2342.

Bentz, E. (1999), "Sectional Analysis of Reinforced Concrete Structures," PhD, Civil


Engineering, University of Toronto, Toronto, Canada.

R-3
Bocca, P., Carpinteri, A. and Valente, S. (1991), "Mixed Model Fracture in Concrete,"
International Journal of Solids and Structures, Vol. 27, No. 9, pp. 1139-1153.

Bolander, J. E. and Wight, J. K. (1991), "Finite Element Modeling of


Shear-Wall-Dominate Buildings," ASCE Journal of Structural Division, Vol.
117, No. 6, pp. 1719-1739.

Bresler, B. and Scordelis, A. C. (1963), "Shear Strength of Reinforced Concrete


Beams," ACI Journal, Vol. 60, No. 1, pp. 51-72.

Brown, R. H. and Jirsa, J. O. (1971), "Reinforced Concrete Beams Under Load


Reversals," ACI Journal, Vol. 68, No. 3, pp. 380-390.

Cardenas, A. S., Russell, H. G. and Corley, W. G. (1980), "Strength of Low-Rise


Structural Walls," in SP-63: Reinforced Concrete Structures Subjected to Wind
and Earthquake Forces, ACI, Detroit, Michigan, pp. 221-241.

CEB (1996), "RC Elements under Cyclic Loading," Thomas Telford, London, 190 pp.

Cervenka, V. and Gerstle, K. H. (1971), "Inelastic Analysis of Reinforced Concrete


Panels: Theory," International Association of Bridge and Structural
Engineering, Vol. 31-II, pp. 31-45.

Cervenka, V. and Gerstle, K. H. (1972), "Inelastic Analysis of Reinforced Concrete


Panels: Experimental Verification and Application," International Association of
Bridge and Structural Engineering, Vol. 32-II, pp. 23-39.

Chang, G. A. and Mander, J. B. (1994), "Seismic Energy Based Fatigue Damage


Analysis of Bridge Columns. Part I: Evaluation of Seismic Capacity,"
Report No. NCEER-94-0006, State University of New York, Buffalo,
New York.

R-4
Chang, T. Y., Taniguchi, H. and Chen, W. F. (1987), "Nonlinear Finite Element
Analysis of Reinforced Concrete Panels," ASCE Journal of Structural
Engineering, Vol. 113, No. 1, pp. 122-140.

Chen, W. F. (1982), "Plasticity in Reinforced Concrete," McGraw-Hill, New York.

Collins, M. P. (1978), "Towards a Rational Theory for RC Members in Shear,"


ASCE Journal of Structural Division, Vol. 104, No. ST4, pp. 649-666.

Collins, M. P. and Mitchell, D. (1981), "Shear and Torsion Design of Prestressed and
Non-Prestressed Concrete Beams," PCI Journal, Vol. 25, No. 5, pp. 96-118.

Collins, M. P. and Mitchell, D. (1991), "Prestressed Concrete Structures," Prentice Hall,


New Jersey, 766 pp.

Collins, M. P., Mitchell, D. and MacGregor, J. G. (1993), "Structural Design


Considerations for High-Strength Concrete," ACI Concrete International,
Vol. 15, No. 5, pp. 27-34.

Collins, M. P. and Prorasz, A. (1989), "Shear Strength for High Strength Concrete,"
CEB Bulletin of Design Aspects of High Strength Concrete, No. 193, pp. 75-83.

Cook, R. D., Malkus, D. S., Plesha, M. E. and Witt, R. J. (2002), "Concepts and
Applications of Finite Element Analysis," John Wiley & Sons, US, 719 pp.

Cope, R. J., Rao, P. V., Clark, L. A. and Norris, P. (1980), "Modelling of Reinforced
Concrete Behaviour for Finite Element Analysis of Bridge Slabs," proceedings
Numerical method for Nonlinear Problems, Taylor, C., a and b, ed., Swansea,
pp. 457-470.

R-5
Cornelissen, H. A. W., Hordijk, D. A. and Reinhardt, H. W. (1985), "Experiments and
Theory for the Application of Fracture Mechanics to Normal and Lightweight
Concrete," proceedings International Conference on Fracture Mechanics of
Concrete, Wittman, F. H., ed., Amsterdam, Netherlands.

Crisfield, M. A. and Wills, J. (1989), "Analysis of R/C Panels Using Different Concrete
Models," ASCE Journal of Engineering Mechanics Division, Vol. 115, No. 3,
pp. 578-597.

Darwin, D. and Pecknold, D. A. (1974), "Inelastic Model for Cyclic Biaxial Loading of
Reinforced Concrete," Report No. 409, University of Illinois, Urbana, Illinois,
169 pp.

Darwin, D. and Pecknold, D. A. (1976), "Analysis of RC Shear Panels Under Cyclic


Loading," ASCE Journal of Structural Division, Vol. 102, No. ST2,
pp. 355-369.

De Borst, R. and Nauta, P. (1985), "Non-Orthogonal Cracks in a Smeared Finite


Element Method," Engineering Computations, Vol. 2, pp. 35-46.

Dei Poli, S., Di Prisco, M. and Gambarova, P. G. (1990), "Stress Field in Web of RC
Thin-Webbed Beams Failing in Shear," ASCE Journal of Structural
Engineering, Vol. 116, No. 9, pp. 2496-2515.

Divakar, M. P., Fafitis, A. and Shah, S. P. (1987), "Constitutive Model for Shear
Transfer in Cracked Concrete," ASCE Journal of Structural Engineering,
Vol. 113, No. 5, pp. 1046-1062.

Elmorsi, M., Kianoush, M. R. and Tso, W. K. (1998), "Nonlinear Analysis of Cyclically


Loaded Reinforced Concrete Structures," ACI Structural Journal, Vol. 95,
No. 6, pp. 725-739.

R-6
Fardis, M. A. and Buyukozturk, O. (1979), "Shear Trnsfer Model for Reinforced
Concrete," ASCE Journal of Engineering Mechanics Division, Vol. 105,
No. EM2, pp. 255-275.

Farvashany, F. E. (2004), "Strength and Deformation of High Strength Concrete Shear


Walls," PhD Thesis, Department of Civil Engineering, Curtin University of
Technology, Curtin, Australia, 153 pp.

Feenstra, P. H., De Borst, R. and Rots, J. G. (1991a), "Numerical Study on Crack


Dilatancy I: Models and Stability Analysis," ASCE Journal of Engineering
Mechanics, Vol. 117, No. 4, pp. 733-753.

Feenstra, P. H., De Borst, R. and Rots, J. G. (1991b), "Numerical Study on Crack


Dilatancy II: Applications," ASCE Journal of Engineering Mechanics, Vol. 117,
No. 4, pp. 754-769.

Fenwick, R. C. (1966), "Shear Strength of Reinforced Concrete Beams," PhD Thesis,


University of Canterbury, Christchurch, New Zealand.

Fenwick, R. C. and Paulay, T. (1968), "Mechanism of Shear Resisting of Concrete


Beams," ASCE Journal of Structural Division, Vol. 94, No. ST10,
pp. 2325-2350.

Filippou, F. C., Popov, E. P. and Bertero, V. V. (1983), "Effects of Bond Deterioration


on Hysteretic Behaviour of Reinforced Concrete Joints," Earthquake
Engineering Research Center, University of California, Berkeley, California.

Fintel, M. (1991), "Shear Walls- An Answer for Seismic Resistance," ACI Concrete
International, Vol. 13, No. 7, pp. 48-53.

R-7
Foster, S. J. (1992), "The Structural Behaviour of Reinforced Concrete Deep Beams,"
PhD Thesis, School of Civil and Environmental Engineering, University of
New South Wales, Sydney.

Foster, S. J., Budiono, B. and Gilbert, R. I. (1996), "Rotating Crack Finite Element odel
for Reinforced Concrete Structures," Computers & Structures, Vol. 58, No. 1,
pp. 43-50.

Foster, S. J. and Gilbert, R. I. (1990), "Non-linear Finite Element Model For Reinforced
Concrete Deep Beams And Panels," Report No. R-275, University of
New South Wales, Sydney, Australia, 113 pp.

Foster, S. J. and Marti, P. (2002), "FE Modelling of RC Membrane using the CMM
Formulation," proceedings Fifth World Congress on Computational Mechanics,
Mang, H. A., Rammerstorfer, F. G. and Ebehardsteiner, J., ed., Viena, Austria.

Foster, S. J. and Marti, P. (2003), "Crack Membrane Model: Finite Element


Implementation," ASCE Journal of Structural Engineering, Vol. 129, No. 9,
pp. 1155-1163.

Foster, S. J. and Rangan, B. V. (1999), "Finite Element Modelling of HSC Squat Walls
in Shear," proceedings 16th Australasian Conference on the Mechanics of
Structures and Materials, Bradford, M., Bridge, R. Q. and Foster, S. J., ed.,
Sydney, Australia, pp. 115-120.

Franklin, J. C. (1970), "Nonlinear Analysis of Reinforced Concrete Frames and Panels,"


Structural Engineering and Structural Mechanics, University of California,
Berkeley.

Gambarova, P. (1980), "Shear Transfer by Aggregate Interlock in Cracked Reinforced


Concrete Subject to Repeated Loads (in Italian)," Report No. 1/79,
Politecnico di Milano, Milan, 43-70 pp.

R-8
Gopalaratnam, V. S. and Shah, S. P. (1985), "Softening Response of Plain Concrete in
Direct Tension," ACI Journal, Vol. 82, No. 2, pp. 310-323.

Gupta, A. and Akbar, H. (1983), "A Finite Element for the Analysis of Reinforced
Concrete Structures," International Journal for Numerical Methods in
Engineering, Vol. 19, pp. 1705-1712.

Gupta, A. and Rangan, B. V. (1996), "Studies on Reinforced Concrete Structural


Walls," Report No. 2/96, Curtin University of Technology, Curtin, Australia,
165 pp.

Gupta, A. and Rangan, B. V. (1998), "High-Strength Concrete (HSC) Structural Walls,"


ACI Structural Journal, Vol. 95, No. 2, pp. 194-204.

Gustafsson, P. J. (1985), "Fracture Mechanics Studies of Nonyielding Materials Like


Concrete," PhD Thises, Division of Building Materials, University of Lund,
Sweden.

Gylltoft, K. (1984), "Fracture Mechanics Model for Fatigue in Concrete," RILEM


Materials and Structures, Vol. 17, No. 97, pp. 55-58.

Hillerborg, A. (1980), "Analysis of fracture by means of the fictitious crack model


particularly for fiber reinforced concrete," Journal of Cement Composites, No. 2,
pp. 177-184.

Hillerborg, A., Modeer, M. and Petersson, P. E. (1976), "Analysis of Crack Formation


and Crack Growth in Concrete by Means of Fracture Mechanics and Finite
Elements," Cement and Concrete Research, No. 6, pp. 773-782.

Hofbeck, J. A., Ibrahim, I. O. and Mattock, A. H. (1969), "Shear Transfer in Reinforced


Concrete," ACI Journal, Vol. 66, No. 2, pp. 119-128.

R-9
Houde, J. and Mirza, M. S. (1972), "Investigation of Shear Transfer across Cracks by
Aggregate Interlock," Report No. 72-06, Ecole Polytechnique de Montreal,
Montreal, Canada, 82 pp.

Hu, H. T. and Schnobrich, W. C. (1990), "Nonlinear Analysis of Cracked Reinforced


Concrete," ACI Structural Journal, Vol. 87, No. 2, pp. 199-207.

Ingraffea, A. R. and Saouma, V. (1984), "Numerical Modelling of Discrete Crack


Propagation in Reinforced and Plain Concrete," in Fracture Mechanics of
Concrete: Structural Application and Numerical Calculation, Sih, G. and
DiTommaso, A., ed., Martinus Nijhoff, pp. 171-225.

Kabeyasawa, T. and Hiraishi, H. (1998), "Tests and Analyses of High-Strength


Reinforced Concrete Shear Walls in Japan," in SP-176: High-Strength Concrete
in Seismic Regions, French, C. and Kreger, M., ed., ACI, Detroit, Michigan,
pp. 281-310.

Kabeyasawa, T., Kuramoto, H. and Matsumoto, K. (1993), "Tests and Analyses of High
Strength Shear Walls," proceedings First Meeting of the Multilateral Project on
the of Use High Strength Concrete, Kyoto, Japan, pp. 1-26.

Karsan, I. K. and Jirsa, J. O. (1969), "Behaviour of Concrete under Compressive


Loadings," ASCE Journal of Structural Division, Vol. 95, No. 12,
pp. 2543-2563.

Kaufmann, W. and Marti, P. (1998), "Structural Concrete: Cracked Membrane Model,"


ASCE Journal of Structural Engineering, Vol. 124, No. 12, pp. 1467-1475.

Kupfer, H. B. and Gerstle, K. H. (1969), "Behaviour of Concrete under Biaxial


Stresses," ASCE Journal of Engineering Mechanics Division, Vol. 99, No. EM4,
pp. 852-866.

R-10
Kwak, H. G. and Kim, D. Y. (2004a), "Material Nonlinear Analysis of RC Shear Walls
Subjected to Cyclic Loadings," Elsevier Engineering Structure, Vol. 26,
pp. 1423-1436.

Kwak, H. G. and Kim, D. Y. (2004b), "Material Nonlinear Analysis of RC Shear Walls


Subjected to Monotonic Loading," Elsevier Engineering Structure, Vol. 26,
pp. 1517-1533.

Kwan, W. P. and Billington, S. L. (2001), "Simulation of Structural Concrete under


Cyclic Load," ASCE Journal of Structural Engineering, Vol. 127, No. 12,
pp. 1391-1401.

Lai, D. (2001), "Crack Shear-Slip in Reinforced Concrete Elements," MASc Thesis,


Department of Civil Engineering, University of Toronto, Toronto, Canada,
154 pp.

Laible, J. P., White, R. N. and Gergely, P. (1977), "An Experimental Investigation of


Seismic Shear Transfer Across Cracks in Concrete Nuclear Containment
Vessels," in SP-53: Reinforced Concrete Structures in Seismic Zones, ACI,
Detroit, Michigan.

Lefas, I. D. (1988), "Behaviour of Reinforced Concrete Walls and Its Implication for
Ultimate Limit State Design," PhD Thesis, Imperial College, University of
London, London, 330 pp.

Lefas, I. D. and Kotsovos, M. D. (1990), "Strength and Deformation Characteristics of


Reinforced Concrete Walls under Load Reversals," ACI Structural Journal,
Vol. 87, No. 6, pp. 716-726.

Lefas, I. D., Kotsovos, M. D. and Ambraseys, N. N. (1990), "Behavior of Reinforced


Concrete Structural Walls: Strength, Deformation Characteristics, and Failure
Mechanism," ACI Structural Journal, Vol. 87, No. 1, pp. 23-31.

R-11
Leonhardt, F. and Walther, R. (1966), Wandartige Trager, Bulletin No. 178, Wilhelm
Ernst and Sohn, Berlin, 159 pp.

Li, B. and Maekawa, K. (1987), "Contact Density Model for Cracks in Concrete,"
proceedings Computational Mechanics of Concrete Structures: Advances and
Applications, Delft, pp. 51-62.

Li, B., Maekawa, K. and Okamura, H. (1989), "Contact Density Model for Stress
Transfer Across Cracks in Concrete," Journal of the Faculty of Engineering,
Vol. 40, No. 1, pp. 9-52.

Lin, C. S. and Scordelis, A. C. (1975), "Nonlinear Analysis of RC Shells of General


Forms," ASCE Journal of Structural Division, Vol. 101, No. ST3, pp. 523-538.

Liu, T. C. Y., Nilson, A. H. and Slate, F. O. (1972), "Biaxial Stress-Strain Relations for
Concrete," ASCE Journal of the Structural Division, Vol. 98, No. ST5, pp. 1025-
1034.

Loeber, P. J. (1970), "Shear Transfer by Aggregate Interlock," M.E. Thesis, University


of Canterbury, Christchurch, New Zealand, 163 pp.

Lubell, A., Sherwood, T., Bentz, E. and Collins, M. P. (2004), "Safe Shear Design of
Large, Wide Beams," ACI Concrete International, No. 1, pp. 67-78.

Ma, S. Y. M., Bertero, V. V. and Popov, E. P. (1976), "Experimental and Analytical


Studies on the Hysteretic Behaviour of Reinforced Concrete Rectangular and T-
beams," Report No. EERC 76-02, Earthquake Engineering Research Center,
University of California, Berkeley, California.

Maekawa, K., Pimanmas, A. and Okamura, H. (2003), "Nonlinear Mechanics of


Reinforced Concrete," Spon Press, New York, 721 pp.

R-12
Maier, J. and Thurlimann, B. (1985), "Bruchversuche an Stahlbetonscheiben," Research
Report, Institue fur Baustatik und Konstruktion ETH, Zurich, 130 pp.

Mansour, M. and Hsu, T. T. C. (2005), "Behaviour of Reinforced Concrete Elements


under Cyclic Shear. II Theoretical Model," ASCE Journal of Structural
Engineering, Vol. 131, No. 1, pp. 54-65.

Marti, P. (2005), "Modelling of Structural Concrete," proceedings fib Symposium Keep


Concrete Attractive, Balazs, G. L. and Borosnyoi, A. B., ed., Budapest,
pp. 1217.

Marti, P., Alvarez, M., Kaufmann, W. and Sigrist, V. (1998), "Tension Chord Model for
Structural Concrete," Structural Engineering International, No. 4/98,
pp. 287-298.

Mattock, A. H. (2001), "Shear Friction and High-Strength Concrete," ACI Structural


Journal, Vol. 98, No. 1, pp. 50-59.

Mattock, A. H. and Hawkins, N. M. (1972), "Shear Transfer in Reinforced Concrete -


Recent Research," PCI Journal, Vol. 17, No. 2, pp. 55-75.

Mattock, A. H., Li, W. K. and Wang, T. C. (1976), "Shear Transfer in Lightweight


Reinforced Concrete," PCI Journal, Vol. 21, No. 1, pp. 20-39.

Menegotto, M. and Pinto, P. (1973), "Method of Analysis for Cyclically Loaded


Reinforced Concrete Plane Frames Including Changes in Geometry and
Nonelastic Behaviour of Elements under Combined Normal Force and
Bending," proceedings IABSE Symposium on Resistance and Ultimate
Deformability of Structures Acted on by Well-Defined Repeated Loads, Lisbon,
pp. 15-22.

R-13
Mau, S. T. and Hsu, T. T. C. (1986), "Shear Design and Analysis of Low-Rise
Structural Walls," ACI Journal, Vol. 83, No. 2, pp. 306-315.

Meyboom, J. (1987), "An Experimental Investigation of Partially Prestressed,


Orthogonally Reinforced Concrete Elements Subjected to Membrane Shear,"
MASc Thesis, Department of Civil Engineering, University of Toronto, Toronto,
Canada, 180 pp.

Milford, R. V. (1984), "Nonlinear Behaviour of Reinforced Concrete Cooling Tower,"


PhD Thesis, Department of Civil Engineering, University of Illinois, Urbana,
Illinios.

Mitchell, D. and Collins, M. P. (1974), "Diagonal Compression Field Theory - A


Rational Model for Structural Concrete in Pure Torsion," ACI Journal, Vol. 71,
No. 8, pp. 396-408.

Miyakawa, T., Kawakami, T. and Maekawa, K. (1987), "Nonlinear Behaviour of


Cracked Reinforced Concrete Plate Element under Uniaxial Compression,"
Proceeding of the JSCE, No. 374, pp. 249-258.

Mphonde, A. G. (1988), "Aggregate Interlock in High Strength Reinforced Concrete


Beams," The Institution of Civil Engineering, Thomas Telford, No. 85,
pp. 397-413.

Mufti, A. A., Mirza, M. S., McCutcheon, J. O. and Houde, L. (1970), "A study of the
Behaviour of Reinforced Concrete Elements," Report No. 70-5, McGill
University, Montreal, Canada.

Nawy, E. G. (2001), "Foundamentals of High-Performance Concrete," John Wiley &


Sons, USA, 441 pp.

R-14
Ngo, D. and Scordelis, A. C. (1967), "Finite Element Analysis of Reinforced Concrete
Beams," ACI Journal, Vol. 82, No. 2, pp. 162-169.

Nilson, A. H. (1968), "Nonlinear Analysis of Reinforced Concrete by the Finite


Element Method," ACI Journal, Vol. 65, No. 9, pp. 757-766.

Noguchi, H. (1985), "Analytical Models for Cyclic Loading of RC Members,"


proceedings Finite Element Analysis of Reinforced Concrete Structures, Meyer,
C. and Okamura, H., ed., Tokyo, Japan, pp. 486-506.

Oesterle, R. G., Arisizabal-Ochoa, J. D., Sihu, K. N. and Corley, W. G. (1984), "Web


Crushing of Reinforced Concrete Structural Walls," ACI Structural Journal,
Vol. 81, No. 3, pp. 231-241.

Oesterle, R. G., Fiorato, A. E., Johal, L. S., Carpenter, J. E., Russel, H. G. and Corley,
W. G. (1976), "Earthquake Resistant Structural Walls - Tests of Isolated Walls -
Phase I," Report to National Science Foundation, Portland Cement
Association(PCA), Skokie, Illinois, 315 pp.

Oesterle, R. G., Fiorato, A. E., Johal, L. S., Carpenter, J. E., Russel, H. G. and Corley,
W. G. (1978), "Earthquake Resistant Structural Walls - Tests of Isolated Walls -
Phase II," Portland Cement Association(PCA), Washington D. C., USA.

Okamura, H. and Maekawa, K. (1991), "Nonlinear analysis and constitutive models of


reinforced concrete," Gihodo Press, Tokyo, Japan, 182 pp.

Ortiz, M., Leroy, Y. and Needleman, A. (1987), "A Finite Element Method for
Localization Failure Analysis," Computational Methods in Applied Mechanical
Engineering, No. 90, pp. 781-804.

R-15
Palermo, D. (1998), "Testing of a Three-Dimensional Shear Wall under Cyclic
Loading," MASc thesis, Department of Civil Engineering, University of
Toronto, Toronto, Canada, 240 pp.

Palermo, D. (2002), "Behaviour and Analysis of Reinforced Concrete Walls Subjected


to Reversed Cyclic Loading," PhD Thesis, Department of Civil Engineering,
University of Toronto, Toronto, Canada, 331 pp.

Palermo, D. and Vecchio, F. J. (2002), "Behavior of Three-Dimensional Reinforced


Concrete Shear Walls," ACI Structural Journal, Vol. 99, No. 1, pp. 81-89.

Palermo, D. and Vecchio, F. J. (2003), "Compression Field Modeling of Reinforced


Concrete Subjected to Reversed Loading: Formulation," ACI Structural Journal,
Vol. 100, No. 5, pp. 616-625.

Palermo, D. and Vecchio, F. J. (2004), "Compression Field Modeling of Reinforced


Concrete Subjected to Reversed Loading: Verification," ACI Structural Journal,
Vol. 101, No. 2, pp. 155-164.

Park, R. and Kent, D. C., and Sampson, R. A. (1972), "Reinforced Concrete Members
with Cycling Loading," ASCE Journal of the Structural Division, Vol. 98,
No. ST7, pp. 1341-1360.

Paulay, T. and Loeber, P. J. (1974), "Shear Transfer by Aggregate Interlock," ACI,


Detroit, Michigan, 1-15 pp.

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


Masonry Buildings," John Wiley & Sons, US, 744 pp.

Petersson, P. E. (1981), "Crack Growth and Development of Fracture Zones in Plain


Concrete and Similar Materials," TVBM-1006, Lund Institute of Technology,
Lund, Sweden, 174 pp.

R-16
Pilakoutas, K. and Elnashai, A. (1995a), "Cyclic Behavior of Reinforced Concrete
Cantilever Walls, Part I: Experimental Results," ACI Structural Journal,
Vol. 92, No. 3, pp. 271-281.

Pilakoutas, K. and Elnashai, A. (1995b), "Cyclic Behavior of Reinforced Concrete


Cantilever Walls, Part II: Discussions and Theoretical Comparisons," ACI
Structural Journal, Vol. 91, No. 4, pp. 425-434.

Popov, E. P. and Ortiz, M. (1979), "Macroscopic and Microscopic Cyclic Plasticity,"


ASCE Journal of Engineering Mechanics Division, pp. 303-330.

Ramberg, W. and Osgood, W. R. (1943), "Description of Stress-Strain Curves by Three


Parameters," Report No. TN-902, National Advisory Committee on
Aeronautics (NACA), Washington, D. C.

Rangan, B. V. (1997), "Rational Design of Structural Walls," ACI Concrete


International, Vol. 19, No. 11, pp. 29-33.

Rashid, Y. R. (1968), "Analysis of Prestressed Concrete Pressure Vessels," Nuclear


Engineering and Design, Vol. 7, No. 4, pp. 334-344.

Reinhardt, H. W. (1984), "Fracture Mechanics of an Elastic Softening Material Like


Concrete," Heron, Vol. 29, No. 2, pp. 1-42.

Robinson, J. R. and Demorieux, J. M. (1977), "Essais de modeles d'ame de pouter en


double te," Annales de 1'Institute Technique du Batiment et des Traveaux
Publics, No. 172, pp. 77-95.

Rots, J. G., Nauta, P., Kusters, G. M. A. and Blaauwendraad, J. (1985), "Smeared Crack
Approach and Fracture Localisation in Concrete," Heron, Vol. 30, No. 1.

R-17
Sanez, L. P. (1964), "Equation for the Stress-Strain Curve of Concrete," ACI Journal,
Vol. 61, No. 9, pp. 1229-1235.

Seckin, M. (1981), "Hysteretic Behaviour of Cast-in-Place Exterior Beam-Column-Slab


Subassemblies," PhD Thesis, Department of Civil Engineering, University of
Toronto, Toronto, Canada, 266 pp.

Shah, S. P. and Winter, G. (1966a), "Inelastic Behaviour and Fracture of Concrete," ACI
Journal, Vol. 63, No. 9, pp. 925-930.

Shah, S. P. and Winter, G. (1966b), "Response of Concrete to Repeated Loading,"


proceedings International Symposium on the Effects of Repeated Loading on
Materials and Structural Elements, Mexico City.

Sharma, N. K. (1969), "Splitting Failure in Reinforced Concrete Members," Ph.D


Thesis, Department of Structural Engineering, Cornell University, Ithaca,
New York.

Shipman, J. M. and Gerstle, K. H. (1979), "Bond Deterioration in Concrete Panels


under Load Cyclic," ACI Journal, Vol. 76, No. 2, pp. 311-325.

Sinha, B. P., Gerstle, K. H. and Tulin, L. G. (1964), "Stress-Strain Relations for


Concrete under Cyclic Loading," ACI Journal, Vol. 61, No. 2, pp. 195-211.

Sittipunt, C. and Wood, S. L. (1993), "Finite Element Analysis of Reinforced Concrete


Shear Walls," Report No. 584, University of Illinois, Urbana, Illinois, 384 pp.

Sittipunt, C. and Wood, S. L. (1995), "Influence of Web Reinforcement on the Cyclic


Response of Structural Walls," ACI Structural Journal, Vol. 92, No. 6,
pp. 745-756.

R-18
Smith, G. M. and Young, L. E. (1955), "Ultimate Theory in Flexure by Exponential
Function," ACI Journal, Vol. 52, No. 3, pp. 349-359.

Stevens, N. J., Uzumeri, S. M. and Collins, M. P. (1987), "Analytical Modelling of


Reinforced Concrete Subjected to Monotonic and Reversed Loading,"
Report No. 87-1, University of Toronto, Toronto, Canada, 299 pp.

Suidan and Schnobrich (1973), "Finite Element Analysis of Reinforced Concrete,"


ASCE Journal of Structural Division, Vol. 99, pp. 2109-2122.

Taylor, H. P. S. (1970), "Investigation of the Forces Carried Across Cracks in


Reinforced Concrete Beam in Shear by Interlock of Aggregate," Report No.
42-447, Cement and Concrete Association, London.

Taylor, R. (1959), "A Note on the Mechanism of Diagonal Cracking in Reinforcement,"


Magazine of Concrete Research, Vol. 11, No. 31, pp. 151-158.

Thorenfeldt, D. T., Tomaszewicz, A. and Jensen, J. J. (1987), "Mechanical Properties of


High Strength Concrete and Application in Designe," proceedings International
Symposium on Utilisation of High Strength Concrete, Stavanger, Norway,
pp. 149-159.

Vecchio, F. J. (1989), "Nonlinear Finite Element Analysis of Reinforced Concrete


Membranes," ACI Structural Journal, Vol. 86, No. 1, pp. 26-35.

Vecchio, F. J. (1998), "Lessons From The Analysis of A 3-D Concrete Shear Wall,"
Structural Engineering and Mechanics, Vol. 6, No. 4, pp. 439-455.

Vecchio, F. J. (1999), "Towards Cyclic Loads Modeling of Reinforced Concrete," ACI


Structural Journal, Vol. 96, No. 2, pp. 193-202.Vecchio, F. J. (1999), "Towards
Cyclic Loads Modeling of Reinforced Concrete," ACI Structural Journal, Vol.
96, No. 2, pp. 193-202.

R-19
Vecchio, F. J. (2000a), "Analysis of Shear-Critical Reinforced Concrete Beams," ACI
Structural Journal, Vol. 97, No. 1, pp. 102-110.

Vecchio, F. J. (2000b), "Disturbed Stress Field Model for Reinforced Concrete:


Formulation," ASCE Journal of Structural Engineering, Vol. 126, No. 9,
pp. 1070-1077.

Vecchio, F. J. and Collins, M. P. (1982), "The Response of Reinforced Concrete to In-


Plane Shear and Normal Stresses," Department of Civil Engineering,
University of Toronto, Toronto, 332 pp.

Vecchio, F. J. and Collins, M. P. (1986), "The Modified Compression Feild Theory for
Reinforced Concrete Elements Subjected to Shear," ACI Journal, Vol. 83, No. 2,
pp. 219-231.

Vecchio, F. J. and Palermo, D. (2001), "NLFEARC: Look Both Ways Before


Crossing," in Finit Element Analysis of Reinforced Concrete Structures, Willam,
K. and Tanabe, T. A., ed., ACI, Farmington Hills, Michigan, pp. 1-14.

Walraven, J. (1980), "Aggregate Interlock- A Theoretical and Experimental Analysis,"


Ph.D Thesis, Faculty of Civil Engineering, Delft University of Technology,
Delft, Netherlands.

Walraven, J. (1981), "Fundamental Analysis of Aggregate Interlock," ASCE Journal of


Structural Division, Vol. 107, No. ST11, pp. 2245-2270.

Walraven, J. (1995), "Shear Friction in High-Strength Concrete," proceedings Progress


in Concrete Research, Delft University of Technology, Delft, Netherlands,
pp. 57-65.

R-20
Walraven, J. and Reinhardt, H. W. (1981), "Theory and Experiments on the Mechanical
Behaviour of Cracks in Plane and Reinforced Concrete Subjected to Shear
Loading," Heron, Vol. 26, No. 1a, pp. 5-68.

Walraven, J., Vos, E. and Reinhardt, H. W. (1979), "Experiments on Shear Transfer in


Cracks in Concrete. Part I: Descriptions of Results," Report No. 5-79-3, Delft
University of Technology, Delft, Netherlands.

Wang, T. and Hsu, T. T. C. (2001), "Nonlinear Finite Element Analysis of Concrete


Structures Using New Constitutive Models," Computers & Structures, Vol. 79,
pp. 2781-2791.

Warner, R. F., Rangan, B. V., Hall, A. S. and Faulkes, K. A. (1998), "Concrete


Structures," Longman, Melbourne, Australia, 974 pp.

White, R. N. and Holley, M. J. (1972), "Experimental Studies of Membrane Shear


Transfer," ASCE Journal of Structural Division, Vol. 98, No. ST8,
pp. 1835-1852.

Xu, C. (1991), "Analytical Model for Reinforced Concrete under Cyclic Loading," PhD
Dessertation, Department of Civil Engineering, Unversity of Illinois, Urbana,
Illinois, 164 pp.

Yankelevsky, D. Z. and Reinhardt, H. W. (1987a), "Model for Cyclic Compressive


Behavior of Concrete," ASCE Journal of Structural Engineering, Vol. 113,
No. 2, pp. 228-240.

Yankelevsky, D. Z. and Reinhardt, H. W. (1987b), "Response of Plain Concrete to


Cyclic Tension," ACI Materials Journal, Vol. 84, No. 5, pp. 365-373.

R-21
Yankelevsky, D. Z. and Reinhardt, H. W. (1989), "Uniaxial Behavior of Concrete in
Cyclic Tension," ASCE Journal of Structural Engineering, Vol. 115, No. 1,
pp. 167-182.

Yoshikawa, H., Wu, Z. and Tanabe, T. A. (1989), "Analytical Model For Shear Slip of
Cracked Concrete," ASCE Journal of Structural Engineering, Vol. 115, No. 4,
pp. 771-788.

Zhang, L. X. and Hsu, C. T. T. (1998), "Bahaviour and Analysis of 100 MPa Concrete
Membrane Element," ASCE Journal of Structural Engineering, Vol. 124, No. 1,
pp. 24-34.

Zienkiewicz, O. C. (1977), "The Finite Element Method," McGraw Hill, New York.

R-22
APPENDIX A

FINITE ELEMENT IMPLEMENTATION

A.1 Type of Element

The stiffness matrix for a particular element is determined using the strain displacement
matrix B as well as the constitutive matrix D xy and geometrical dimensions. Although

there has been a general trend among researchers that finite element analysis using
higher order elements are preferable to those using simple elements but in larger
numbers (Barzegar, 1995), the lower order elements have been efficiently utilised to
analyse reinforced concrete panels and shear walls (Vecchio, 1989, Foster, 1992, Ayoub
and Filippou, 1998, Vecchio, 1999, Wang and Hsu, 2001 and many others). In this
study the four-node isoparametric quadrilateral element (Q4) is adopted (Figure A.1).

Since the element is isoparametric, both coordinates ( x and y ) and displacements


( u and v ) of a point within the element are determined from the coordinates
( xi and yi ) and displacements ( u i and vi ) of nodes using the same shape functions as

follows:
3(1,1)

4(-1,1) η

ξ
(0,0)

y,v
1(-1,-1) 2(1,-1)

x,u

Figure A.1 – Four-node isoparametric quadrilateral element.

⎧ 4 ⎫ ⎧ 4 ⎫
⎧x⎫ ⎪
⎪ ∑ N i xi ⎪


⎧u ⎫ ⎪
∑ N i ui ⎪

1 1
⎨ ⎬=⎨ 4 ⎬ and ⎨ ⎬=⎨ 4 ⎬ (A.1)
⎩ y⎭ ⎪ ⎩v ⎭ ⎪
⎪ ∑ N i yi ⎪⎪ ⎪ ∑ N i vi ⎪⎪
⎩ 1 ⎭ ⎩ 1 ⎭

where i indicates the numbering of nodes as shown in Figure A.1 and N i is the shape

function of node i given by (Cook et al., 2002)

1
N1 = (1 − ξ )(1 − η )
4
1
N 2 = (1 + ξ )(1 − η )
4 (A.2)
1
N 3 = (1 + ξ )(1 + η )
4
1
N 4 = (1 − ξ )(1 + η )
4

The strain displacement matrix B is determined by

A-2
⎡ N 1, x 0 N 2, x 0 N 3, x 0 N 4, x 0 ⎤
⎢ ⎥
B=⎢ 0 N 1, y 0 N 2, y 0 N 3, y 0 N 4, y ⎥ (A.3)
⎢ N 1, x N 1, y N 2, x N 2, y N 3, x N 3, y N 4, x N 4, y ⎥⎦

in which N 1, x = ∂N 1 ∂x , and so on. Because the derivatives with respect to x and y are

not available directly, applying the chain rule for N i it can be shown that

⎧ N i ,ξ ⎫ ⎧ N i, x ⎫
⎨ ⎬ = J ⎨N ⎬ (A.4)
⎩ N i,η ⎭ ⎩ i, y ⎭

where J is the Jacobian matrix expressed by

⎡ x, ξ y,ξ ⎤
J =⎢ (A.5)
⎣ x,η y,η ⎥⎦

Frome Eqs. A.4 and A.5 the desired derivatives with respect to x and y are obtained by

⎧ N i, x ⎫ 1 ⎡ y,η − y,ξ ⎤ ⎧ N i ,ξ ⎫
⎨N ⎬ = ⎢ ⎨ ⎬ (A.6)
⎩ i, y ⎭ J ⎣− x,η x,ξ ⎥⎦ ⎩ N i,η ⎭

and subsequently B is determined.

A.2 Nonlinear Solution Procedure

Nonlinear structural problem can be solved using iterative solution techniques in


conjunction with the linear finite element analysis until the predefined convergence is
reached. There are three common iterative solution techniques including tangent
stiffness method, initial stiffness method and secant stiffness method (ASCE Committee
447, 1982). In the tangent stiffness method, commonly known as the Newton-Raphson
(NR) method, while the rate of convergence is quicker than other methods, the
structural stiffness matrix needs to be updated for each iteration leading to a high
A-3
computational time and cost which are especially important for large finite element
analyses. Furthermore it is evident that some numerical troubles may be encountered
during applying this technique particularly when a softening structure is analysed
(Bathe, 1996, Cook et al., 2002). On the other hand the initial stiffness method,
commonly known as the Modified Newton-Raphson (MNR) method, takes the distinct
advantage that the structure stiffness matrix need not be updated after each iteration.
This is, however, compromised by the slowest rate of convergence compared to other
methods. The secant stiffness matrix exhibits simpler formulation and higher numerical
stability, but slower rate of convergence than NR method. Although one criticism of this
method has been that it cannot be efficiently used to model response to general loading,
Vecchio (1999) and Palermo (2001) used it successfully to analyse the normal strength
RC shear walls under cyclic loading. It should be mentioned that to make a balance
between the factors described above, often a combination of initial stiffness method
with the tangent or secant stiffness method is used in which the structure stiffness
matrix is occasionally updated.

Here the secant stiffness method is adopted that can be used separately or in conjunction
with the initial stiffness method as shown in Figure A.2. Since the solution for RC
structures is generally path-dependent, an incremental displacement or load procedure
to achieve the total amount is applied. For each displacement or load step the iterative
cycles are applied until the satisfied convergence is reached. To check this, the
displacement change Δ u i and out-of-balance force R i are determined after each

iteration and used as described in the next section.

To overcome some of the numerical problems associated with relying on merely load or
displacement control the constant arc length method is used. This method can be
applied well using the secant stiffness as well as tangent stiffness technique. The arc
length is calculated at the end of each load or displacement increment.

A-4
Load
Pt + Δt

Ri
Pt

Pt - Δt
Δui

ui u i +1 Displacement

(a)
Load

Pt + Δt

Ri

Pt

Pt - Δt
Δu i

ui u i +1 Displacement

(b)

Figure A.2 - Nonlinear solution procedure; (a) secant stiffness method (b) combination
of secant and initial stiffness methods.

A-5
A.3 Convergence Criteria

In addition to the selection of the proper numerical nonlinear solution techniques, the
accuracy of the incremental iterative procedures also depends on the convergence
criteria. Since the iterative techniques at a given load step lead to only approximate
solutions to nonlinear structural analyses, the convergence criteria are required to define
a predetermined accuracy that must be reached such that the iteration can be terminated
and the analysis can advance to the next load step (Bathe, 1996). In general to monitor
the convergence, either the norm of imbalance forces or the norm of displacement
changes can be checked at the end of each iteration so that when Eq. A.7 is satisfied, the
iterative cycle may be considered complete (Foster and Gilbert, 1990)

ψi < ε ψ j
max
( j = 1,2,3,..., i ) (A.7)

In Eq. A.7, ε is the convergence tolerance and ψ i is expressed by

Force convergence: ψ i = R (u i )T . R (u i ) (A.8)

Displacement convergence: ψ i = Δ ui T .Δ ui (A.9)

where R (u i ) and Δ u i are the residual or out of balance forces and the displacement

changes at the iteration i , respectively, with respect to the previous iteration

(Figure A.2). The term ψ j is the maximum value calculated from first to current
max

iteration. Force and displacement convergence tolerance values may be different


depending on the nature of problem and the accuracy required.

Although occasionally because of localised force imbalances that have little effect on
overall structural response, the force convergence criterion is not easily satisfied, the
displacement criterion is usually not a satisfactory substitute. The displacement criterion
may ceased iterations only because convergence is slow, or may indicate convergence
when considerable force imbalance remains. Therefore it is recommended that the force

A-6
convergence criterion should be used as a supplement whenever the displacement
criterion is used (Cook et al., 2002). Also a limit is usually considered for the number of
iterations or for the computational time. If one of these limits is reached before
convergence, it may imply that the iteration or time is not adequate, that a too tight
convergence tolerance has been selected or that the solution process has encountered a
trouble (Cook et al., 2002).

A.4 Program RECAP

The model outlined in Chapter 3 is incorporated into the program RECAP, a


displacement-based finite element program developed by Foster and Gilbert (1990) at
the University of New South Wales for linear and nonlinear analysis of reinforced
concrete structures. The program name is an acronym for Reinforced Concrete Analysis
Program. Using a MACRO programming structure for inputting the data based on that
of Taylor (Zienkiewicz, 1977), the problem solution and outputting desired results,
RECAP has a great flexibility in the problem solution and output so that different
approaches and techniques may be employed to solve different classes of problems and
the output desired from one problem to another may be different. Some of main features
of RECAP may be summarised as follows:

1. No assumptions regarding to the maximum number of nodes, elements and


material sets

2. Mesh generation and node numbering

3. Analysis of large problems consisting many elements and nodes without losing
any efficiency in the solution of smaller problems

4. A range of element types in 1-D, 2-D and 3-D

5. Incrementally load and displacement control

A-7
6. Initial, tangential and secant stiffness iterative techniques

7. Arc length and line search techniques

8. Short-term and long-term effects

9. Different desired output and the form of output based on problem nature

A typical finite element analysis using RECAP will involve three stages. These stages
are described with the macro instructions RECAP, MACR and STOP. The RECAP
instruction starts the execution of the problem so that the problem control information
and data is loaded into and stored for the future use. Following the data input stage, the
solution stage begins with the macro command MACR after which the solution macros
line are placed. The final command is STOP, when executed stops the program. The
procedure for solution of nonlinear finite element analysis of RC structures using
RECAP is shown in Figure A.3. More information and details on the structure and
abilities of RECAP can be found elsewhere (Foster and Gilbert, 1990, Foster, 1992).

One feature which makes RECAP appropriate for this study is the capability to allow
installation of new finite element models easily. While the main tasks in the analysis
such as managing memory and databases, solving the equilibrium equations, forming
the stiffness matrix and printing the results are undertaken by the main system, the
major data on the element and material properties is provided from the element module
(Figure A.3) using subroutine arguments. Since these arguments are similar for all
element routines regardless of their type and complexity, a new element model or a new
material model can be added to the program without the modification of the main
program.

A-8
Start

Input data

Update No Yes Update


load Using displacement displacement
parameters control vector

Form structural stiffness


matrix

Yes

Update No Form out of balance force


Stiffness Matrix vectors
Element Module

Calculate structure nodal


displacements

Using Yes Yes


force convergence ψi < εψi max
criteria

No No

Using Yes Yes


displacement convergence ψi < εψi max
criteria
No No

No Max. No.
of iterations exceeded

Yes

Output results

Next
Yes
load or displacement step

No

Stop

Figure A.3 - Flow chart for nonlinear finite element analysis using RECAP.

A-9
APPENDIX B

RESPONSE OF LVDTS
B.1 Specimen SW1

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.1 – Specimen SW1 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.2 - Specimen SW1 response under the phase 2 of lateral loading.

B-2
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.3 - Specimen SW1 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.4 – Specimen SW1, LVDT L2.

B-3
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.5 – Specimen SW1, LVDT L3.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.6 – Specimen SW1, LVDT L4.

B-4
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.7 - Specimen SW1, LVDT L5.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.8 – Specimen SW1, LVDT L6.

B-5
B.2 Specimen SW2

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.9 - Specimen SW2 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.10 - Specimen SW2 response under the phase 2 of lateral loading.

B-6
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.11 - Specimen SW2 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.12 - Specimen SW2 response under the phase 4 of lateral loading.

B-7
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.13 - Specimen SW2 response under the phase 5 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.14- Specimen SW2, LVDT L2.

B-8
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.15 – Specimen SW2, LVDT L3.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.16 – Specimen SW2, LVDT L4.

B-9
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.17 – Specimen SW2, LVDT L5.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.18 – Specimen SW2, LDTV L6.

B-10
B.3 Specimen SW3

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.19 – Specimen SW3 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.20 - Specimen SW3 response under the phase 2 of lateral loading.

B-11
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.21 – Specimen SW3 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement of Top Slab (mm)

Figure B.22 - Specimen SW3 response under the phase 4 of lateral loading.

B-12
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.23 – Specimen SW3, LVDT L2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.24 – Specimen SW3, LVDT L3.

B-13
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.25 – Specimen SW3, LVDT L4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.26 – Specimen SW3, LVDT L5.

B-14
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.27 – Specimen SW3, LDTV L6.

B-15
B.4 Specimen SW4

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.28 – Specimen SW4 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.29 - Specimen SW4 response under the phase 2 of lateral loading.

B-16
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.30 – Specimen SW4 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.31 - Specimen SW4 response under the phase 4 of lateral loading.

B-17
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Strain

Figure B.32 – Specimen SW4, LVDT L2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.33 - Specimen SW4, LVDT L3.

B-18
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.34 - Specimen SW4, LVDT L4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.35 - Specimen SW4, LVDT L5.

B-19
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Diaplacement (mm)

Figure B.36 - Specimen SW4, LVDT L6.

B-20
B.5 Specimen SW5

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.37 - Specimen SW5 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.38 - Specimen SW5 response under the phase 2 of lateral loading.

B-21
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.39 - Specimen SW5 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.40 - Specimen SW5 response under the phase 4 of lateral loading.

B-22
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.41 – Specimen SW5, LVDT L2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24

Displacement (mm)

Figure B.42 - Specimen SW5, LVDT L3.

B-23
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.43 - Specimen SW5, LVDT L4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.44 - Specimen SW5, LVDT L5.

B-24
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.45 – Specimen SW5, LVDT L6.

B-25
B.6 Specimen SW6

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.46 - Specimen SW6 response under the phase 1 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.47 - Specimen SW6 response under the phase 2 of lateral loading.

B-26
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.48 - Specimen SW6 response under the phase 3 of lateral loading.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement of Top Slab (mm)

Figure B.49 - Specimen SW6 response under the phase 4 of lateral loading.

B-27
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.50 – Specimen SW6, LVDT L2.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.51 - Specimen SW6, LVDT L3.

B-28
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.52 - Specimen SW6, LVDT L4.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.53 - Specimen SW6, LVDT L5.

B-29
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-20 -16 -12 -8 -4 0 4 8 12 16 20

Displacement (mm)

Figure B.54 - Specimen SW6, LVDT L6.

B-30
APPENDIX C

LOADS OF MACALLOY BARS


C.1 Specimen SW1

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.1 – Specimen SW1, MacAlloy Bar 1.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.2- Specimen SW1, MacAlloy Bar 2.

C-2
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.3 - Specimen SW1, MacAlloy Bar 3.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.4 - Specimen SW1, MacAlloy Bar 4.

C-3
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.5 - Specimen SW1, MacAlloy Bar 5.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500
Lateral Load (kN)

Figure C.6 - Specimen SW1, MacAlloy Bar 6.

C-4
C.2 Specimen SW2

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.7 - Specimen SW2, MacAlloy Bar 1.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.8 - Specimen SW2, MacAlloy Bar 2.

C-5
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.9 - Specimen SW2, MacAlloy Bar 3.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.10 - Specimen SW2, MacAlloy Bar 4.

C-6
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.11 - Specimen SW2, MacAlloy Bar 5.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.12 - Specimen SW2, MacAlloy Bar 6.

C-7
C.3 Specimen SW3

250
Aial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.13 - Specimen SW3, MacAlloy Bar 1.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.14 - Specimen SW3, MacAlloy Bar 2.

C-8
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.15 - Specimen SW3, MacAlloy Bar 3.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.16 - Specimen SW3, MacAlloy Bar 4.

C-9
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.17 - Specimen SW3, MacAlloy Bar 5.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.18 - Specimen SW3, MacAlloy Bar 6.

C-10
C.4 Specimen SW5

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.19 - Specimen SW5, MacAlloy Bar 1.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.20 - Specimen SW5, MacAlloy Bar 2.

C-11
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.21 - Specimen SW5, MacAlloy Bar 3.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.22 - Specimen SW5, MacAlloy Bar 4.

C-12
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.23 - Specimen SW5, MacAlloy Bar 5.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.24 - Specimen SW5, MacAlloy Bar 6.

C-13
C.5 Specimen SW6

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.25 - Specimen SW6, MacAlloy Bar 1.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.26 - Specimen SW6, MacAlloy Bar 2.

C-14
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.27 - Specimen SW6, MacAlloy Bar 3.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.28 - Specimen SW6, MacAlloy Bar 4.

C-15
250

Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.29 – Specimen SW6, MacAlloy Bar 5.

250
Axial Load (kN)

200

150
-1500 -1000 -500 0 500 1000 1500

Lateral Load (kN)

Figure C.30 - Specimen SW6, MacAlloy Bar 6.

C-16
APPENDIX D

STRAINS OF EDGE ELEMENT REINFORCEMENT


D.1 Specimen SW1

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.1 – Specimen SW1, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.2 – Specimen SW1, Lateral Load-Strain Curve of SG8.

D-2
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.3 – Specimen SW1, Lateral Load-Strain Curve of SG9.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.4 – Specimen SW1, Lateral Load-Strain Curve of SG10.

D-3
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.5 – Specimen SW1, Lateral Load-Strain Curve of SG11.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.6 – Specimen SW1, Lateral Load-Strain Curve of SG12.

D-4
D.2 Specimen SW2

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.7 – Specimen SW2, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure D.8 – Specimen SW2, Lateral Load-Strain Curve of SG9.

D-5
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000
Strain (μm/m)
Figure D.9 – Specimen SW2, Lateral Load-Strain Curve of SG11.

D-6
D.3 Specimen SW3

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.10 – Specimen SW3, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.11 - Specimen SW3, Lateral Load-Strain Curve of SG8.

D-7
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)
Figure D.12 - Specimen SW3, Lateral Load-Strain Curve of SG9.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000
Strain (μm/m)
Figure D.13 - Specimen SW3, Lateral Load-Strain Curve of SG10.

D-8
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.14 - Specimen SW3, Lateral Load-Strain Curve of SG11.

D-9
D.4 Specimen SW4

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.15 - Specimen SW4, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.16 - Specimen SW4, Lateral Load-Strain Curve of SG8.


D-10
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000
Strain (μm/m)
Figure D.17 - Specimen SW4, Lateral Load-Strain Curve of SG9.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.18 - Specimen SW4, Lateral Load-Strain Curve of SG10.

D-11
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.19 - Specimen SW4, Lateral Load-Strain Curve of SG11.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.20 - Specimen SW4, Lateral Load-Strain Curve of SG12.

D-12
D.5 Specimen SW5

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.21 - Specimen SW5, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.22 - Specimen SW5, Lateral Load-Strain Curve of SG8.

D-13
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.23 - Specimen SW5, Lateral Load-Strain Curve of SG9.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.24 - Specimen SW5, Lateral Load-Strain Curve of SG10.

D-14
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.25 - Specimen SW5, Lateral Load-Strain Curve of SG11.

D-15
D.6 Specimen SW6

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.26 - Specimen SW6, Lateral Load-Strain Curve of SG7.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.27 – Specimen SW6, Lateral Load-Strain Curve of SG8.

D-16
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.28 - Specimen SW6, Lateral Load-Strain Curve of SG9.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.29 - Specimen SW6, Lateral Load-Strain Curve of SG10.

D-17
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure D.30 - Specimen SW6, Lateral Load-Strain Curve of SG11.

D-18
APPENDIX E

STRAINS OF WALL REINFORCEMENT


E.1 Specimen SW1

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2500 -2000 -1500 -1000 -500 0 500 1000 1500 2000 2500

Strain(μm/m)

Figure E.1 - Specimen SW1, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2500 -2000 -1500 -1000 -500 0 500 1000 1500 2000 2500

Strain(μm/m)

Figure E.2 - Specimen SW1, Lateral Load-Strain Curve of SG15.

E-2
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2500 -2000 -1500 -1000 -500 0 500 1000 1500 2000 2500

Strain(μm/m)

Figure E.3 - Specimen SW1, Lateral Load-Strain Curve of SG16.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000 5000 6000

Strain(μm/m)

Figure E.4 - Specimen SW1, Lateral Load-Strain Curve of SG17.

E-3
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2000 -1500 -1000 -500 0 500 1000 1500 2000

Strain (μm/m)

Figure E.5 - Specimen SW1, Lateral Load-Strain Curve of SG18.

E-4
E.2 Specimen SW2

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure E.6 - Specimen SW2, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.7 - Specimen SW2, Lateral Load-Strain Curve of SG14.

E-5
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.8 - Specimen SW2, Lateral Load-Strain Curve of SG16.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000

Strain(μm/m)

Figure E.9 - Specimen SW2, Lateral Load-Strain Curve of SG17.

E-6
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.10 - Specimen SW2, Lateral Load-Strain Curve of SG18.

E-7
E.3 Specimen SW3

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.11 - Specimen SW3, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.12 - Specimen SW3, Lateral Load-Strain Curve of SG14.

E-8
1400
1200
1000
800
600
Lareral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -4000 -2000 0 2000 4000 6000

Strain(μm/m)

Figure E.13 - Specimen SW3, Lateral Load-Strain Curve of SG15.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.14 - Specimen SW3, Lateral Load-Strain Curve of SG16.

E-9
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-4000 -3000 -2000 -1000 0 1000 2000 3000 4000

Strain(μm/m)

Figure E.15 - Specimen SW3, Lateral Load-Strain Curve of SG17.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-6000 -4000 -2000 0 2000 4000 6000

Strain(μm/m)

Figure E.16 - Specimen SW3, Lateral Load-Strain Curve of SG18.

E-10
E.4 Specimen SW4

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.17 - Specimen SW4, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.18 - Specimen SW4, Lateral Load-Strain Curve of SG14.

E-11
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.19 - Specimen SW4, Lateral Load-Strain Curve of SG15.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.20 - Specimen SW4, Lateral Load-Strain Curve of SG16.

E-12
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-8000 -6000 -4000 -2000 0 2000 4000 6000 8000

Strain(μm/m)

Figure E.21 - Specimen SW4, Lateral Load-Strain Curve of SG17.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-8000 -6000 -4000 -2000 0 2000 4000 6000 8000

Strain(μm/m)

Figure E.22 - Specimen SW4, Lateral Load-Strain Curve of SG18.

E-13
E.5 Specimen SW5

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.23 - Specimen SW5, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-2000 -1500 -1000 -500 0 500 1000 1500 2000

Strain ( m/m)
Figure E.24 - Specimen SW5, Lateral Load-Strain Curve of SG14.

E-14
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.25 - Specimen SW5, Lateral Load-Strain Curve of SG15.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.26 - Specimen SW5, Lateral Load-Strain Curve of SG16.

E-15
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.27 - Specimen SW5, Lateral Load-Strain Curve of SG17.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.28 - Specimen SW5, Lateral Load-Strain Curve of SG18.

E-16
E.6 Specimen SW6

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain (μm/m)

Figure E.29 - Specimen SW6, Lateral Load-Strain Curve of SG13.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.30 - Specimen SW6, Lateral Load-Strain Curve of SG14.

E-17
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000 5000

Strain(μm/m)

Figure E.31 - Specimen SW6, Lateral Load-Strain Curve of SG15.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.32 - Specimen SW6, Lateral Load-Strain Curve of SG16.

E-18
1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.33 - Specimen SW6, Lateral Load-Strain Curve of SG17.

1400
1200
1000
800
600
Lateral Load (kN)

400
200
0
-200
-400
-600
-800
-1000
-1200
-1400
-3000 -2000 -1000 0 1000 2000 3000

Strain(μm/m)

Figure E.34 - Specimen SW6, Lateral Load-Strain Curve of SG18.

E-19
APPENDIX F

DESIGN CODE CLAUSES FOR SHEAR WALLS

F.1 Australian Standard (AS 3600, 2001)

F.1.1 Strength in Flexural

The flexural strength of shear walls subjected to in-plane axial and lateral loads is
determined based on the requirements expressed for RC columns under bending
moment and axial compression as shown in Figure F.1. The main steps to calculate the
strength of shear walls in flexural are as follows:

1. The input data including the geometrical dimensions of wall cross-section, the
strength of concrete and reinforcement, reinforcement details and applied axial load is
determined.

2. Assuming a trial amount for d N (indicated in Figure F.1), the internal compressive
force due to the contribution of concrete is calculated by:

Fc = α f c'γ d N b f if γ d N ≤ t f (F.1)

Fc = α f c' [b f t f + t w (γ d N − t f )] if γ d N > t f (F.2)


α f c'
εc
γdN
dsi dN

Neutral Axis M

reinforcement layout
Web wall
Plastic Centroid
N
εsi

Cross-section Strain Stress

Figure F.1 - Flexural analysis of shear wall cross-section.

where b f is the width of flange in mm; t f and t w are the thickness of flange and web

wall in mm, respectively; f c' is the concrete strength in MPa and α and γ are two
parameters to determine the rectangular stress block of compressive concrete given as
(AS3600, 2001):

α = 0.85 − 0.004 f c' − 55 ; ( ) ( 0.65 ≤ α ≤ 0.85 ) (F.3)

γ = 0.85 − 0.008 f c' − 30 ( ) ( 0.65 ≤ γ ≤ 0.85 ) (F.4)

3. The stress in each layer of steel reinforcing bars, f si , is calculated from the
following formulation:

f si = Es ε c
(d si − d N ) ≤ f sy (F.5)
dN

F-2
where d si is the distance of reinforcing steel layer from the extreme compressive fibre

of concrete in mm; E s and f sy are the elastic modulus and yield stress of steel bars in

MPa, respectively, and ε c is the concrete strain at the crushing taken as 0.003. The

internal force resulted from all steel bars, Fs , is calculated using:

Fs = ∑ f si Asi (F.6)

where Asi is the area of seel bars in mm2 at ith layer.

4. Resulting axial force on the wall cross section, Ptrial , is calculated from the algebraic
summation of compressive and tensile forces as follows:

Ptrial = Fs + Fc (F.7)

If Ptrial equals to the applied axial load, the moment capacity of wall, M u , is
determined using the summation of the moments of the forces obtained previously
about the plastic centroid of wall cross-section. Otherwise, the steps 1 to 3 are
repeated until the Ptrial equals to the applied axial load.

F.1.2 Strength in Shear

The critical section for maximum shear is considered to be at a distance equal to the
lesser of 0.5 Lw and 0.5 H w where Lw and H w are the length and height of wall,

respectively. The shear strength of a shear wall, Vu , is given by:

Vu = Vuc + Vus (F.8)

where Vuc and Vus are the shear strengths contributed by concrete and shear

reinforcement, respectively. Vu is limited to a maximum value as follows:

F-3
Vu ≤ 0.2 f c' ( 0.8 Lw t w ) (F.9)

where t w is the thickness of wall. The contribution of concrete to the shear strength of

wall, Vuc , is taken as:

Hw
if H w / Lw ≤1 then Vuc = (0.66 f c' − 0.21 f c' ) 0.8 Lw t w (F.10)
Lw

if H w / Lw >1 then the lesser of the values determined from Eq. F.10 and from

⎡ ⎤
⎢ 0.1 f c'⎥
Vuc = ⎢0.05 f c' + ⎥ 0.8 Lw t w (F.11)
⎢ Hw
−1 ⎥
⎢⎣ Lw ⎥⎦

but in any case:

Vuc ≥ 0.17 f c' (0.8 Lw t w ) (F.12)

The value of Vus is determined from:

Vus = pw f sy ( 0.8 Lw t w ) (F.13)

where f sy is the yield stress of reinforcing steel bar and pw is the ratio of reinforcement

given as follows:

• For walls with H w / Lw ≤1 , pw is considered to be the lesser of the ratios of


transverse and longitudinal reinforcement.

• For walls with H w / Lw >1 , pw is the ratio of the transverse reinforcement.

F-4
The transverse reinforcement ratio should not be less than 0.0015. The minimum
amount of longitudinal reinforcement ratio depends on the shrinkage and thermal
conditions, degree of crack control and exposure conditions. The minimum
centre-to-centre spacing of parallel reinforcement is 3 db in which d b is the diameter of

reinforcing bar. In addition, the spacing should not exceed the lesser value of 2.5 t w or
350 mm.

F.2 American Concrete Institute Code (ACI 318, 2002)

F.2.1 Strength in Flexural

Similarly to AS 3600, the flexural strength of a shear wall in ACI 318 (2002) is
determined considering the wall as a compression member under a combination of axial
load and bending moment.

F.2.2 Strength in Shear

The shear strength Vu is determined using Eq. F.8 in which Vuc is the lesser of the
values computed as follows:

Vuc = 0.25 f c' ( 0.8 Lw t w ) + 0.2 N * (F.14)

or

⎡ ' N* ⎤
⎢ Lw ( 0.1 f c + 0.2 )⎥
L t ⎥ (0.8 L t )
Vuc = ⎢0.05 f c +
' w w
(F.15)
⎢ M * Lw ⎥ w w
⎢ − ⎥
⎣ V* 2 ⎦

F-5
but in any case Eq. F.9 applies. In Eqs. F.14 and F.15, N * , V * and M * are the axial

load, shear force and bending moment at the wall section, respectively, where N * is

M* Lw
negative for tension. When ( *
− ) is negative, Eq. F.15 does not applied. The
V 2
contribution of reinforcement to the shear strength of wall Vus is computed by

0.8 Lw
Vus = Avs f sy ( ) (F.16)
s

where Avs is the area of transverse reinforcement within a distance s . The shear

strength Vu should be less than a maximum value given as:

Vu ≤ 0.83 f c' ( 0.8 Lw t w ) (F.17)

The ratio of required shear longitudinal reinforcement should not be less than:

Hw
pl = 0.0025 + 0.5 ( 2.5 − ) ( pt − 0.0025) (F.18)
Lw

where pl and pt are the ratio of longitudinal and transverse shear reinforcement,

respectively. The amount of pl should be greater than 0.0025 and less than the required
transverse shear reinforcement. The spacing of transverse shear reinforcing steel bars
should not exceed Lw / 5 , 3 t w , nor 500 mm. The maximum spacing of longitudinal
shear reinforcement is similarly determined with the exception of first relation that
should be taken as Lw / 3 .

F-6
F.3 Sample Calculation

To show the calculation of shear and flexural strength of shear walls using the described
relationships, the solution for the specimen SW5 is given in the following.

a) Input Data for the Specimen SW5


Layers of flange steel reinforcement = 4
Layers of web steel reinforcement = 8

Concrete compressive strength ( f c' ) = 84.0 MPa

Flange steel reinforcement (N16) yield stress ( f sy ) = 535.0 MPa

Flange steel reinforcement (N16) young’s modulus ( Es ) = 204000 MPa


Longitudinal web wall reinforcement ratio = W8@100mm (each face)
Longitudinal web wall reinforcement ratio (pv) = 0.013
Web longitudinal steel reinforcement (W8) yield stress ( f sy ) = 498.0 MPa

Web longitudinal steel reinforcement (W8) young’s modulus ( Es ) = 179000 MPa


Transverse web wall reinforcement details = W6@167 mm (each face)
Transverse web wall reinforcement ratio (ph) = 0.005
Web transverse steel reinforcement (W6) yield stress ( f sy ) = 535.0 MPa

Flange Width (bf) = 375 mm


Flange Thickness (tf) = 100 mm
Web wall thickness (tw) = 75 mm
Length of wall (Lw) = 1000 mm
Height of wall (Hw) = 1000 mm
Applied axial load = 1200 kN
Height of applied lateral load = 1.100 m

b) Flexural Strength of the Specimen SW5


Using Equations F.3 and F.4, the rectangular stress block parameters are:

α = 0.85 − 0.004 (84 − 55) = 0.734 (0.65 ≤ α ≤ 0.85)

F-7
∴α = 0.738

γ = 0.85 − 0.008 (84 − 30) = 0.418 (γ ≤ 0.65)

∴ γ = 0.65

Using the trial and error process, the depth of the neutral axis in SW5 is found to be
129.8 mm. With this value the total compressive and tensile forces are –2841 kN and
1641 kN, respectively. This gives a resultant net axial force of about 1200 kN., which is
equal to the given applied axial load on the cross-section.

Summing moments about the plastic centroid gives the moment capacity of SW5 as
1897 kNm. The equivalent lateral load to cause failure is equal to:

moment capacity 1897 (kNm)


strength in flexural = =
height of lateral load 1.1 (m)
= 1725 kN

c) Shear Strength of the Specimen SW5

AS 3600 (2001)

The concrete component of the shear strength, Hw/Lw =1, therefore using Equation F.10:

⎛ 1000 ⎞
Vuc = ⎜ 0.66 × 83 − 0.21× × 83 ⎟ (0.8 × 1000 × 75)×10 −3
⎝ 1000 ⎠
= 246 kN

Also, checking the minimum value allowable from Equations F.12:

( )
Vuc min = 0.17 × 83 (0.8 × 1000 × 75)×10 −3 = 93 kN ( Vuc < Vuc min ∴OK)

F-8
For the steel component of the shear strength using Equation F.13 (Hw/Lw =1):

⎛ 56.5 ⎞
Vus = ⎜ × 535 ⎟ (0.8 × 1000 × 75)×10 −3 = 145 kN
⎝ 167 × 75 ⎠

From Equation F.8, the total shear strength of SW5 is:

Vu = 246 + 145 = 391 kN

Checking the shear strength value is under the maximum allowable, using Equation F.9:

Vu max = ( 0.2 × 83 ) ( 0.8 × 1000 × 75 )×10 −3 = 996 kN ( Vu < Vu max ∴OK)

ACI 318 (2002)

Longitudinal web wall reinforcement ratio = W8@100mm (each face)

The concrete component of the shear strength is the lesser of Equation F.14 and F.15:

( )
Vuc = 0.25 × 83 (0.8 ×1000 × 75)×10−3 + 0.2 ×1200
= 377 kN

⎡ ⎛ 1200 ×103 ⎞⎟ ⎤
1000⎜ 0.10 × 83 +
⎢ ⎜ × ⎟⎥
Vuc

(
= ⎢ 0.05 × 83 + ) ⎝

1000
1000 ⎞
75 ⎠ ⎥ (0.8 ×1000 × 75)×10−3

⎢ ⎜1100 − ⎟ ⎥
⎢⎣ ⎝ 2 ⎠ ⎥⎦
= 438 kN

Therefore,

Vuc = 377 kN

F-9
Also, checking the minimum value allowable from Equations F.12:

( )
Vuc min = 0.17 × 83 (0.8 ×1000 × 75)×10 −3 = 92.9 kN ( Vuc < Vuc min ∴OK)

For the steel component of the shear strength, using Equation F.16:

Vus = (56.5 × 535)


(0.8 ×1000) ×10−3 = 145 kN
167

From Equation F.8, the total shear strength of SW5 is:

Vu = 377 + 145 = 522 kN

Checking the shear strength value is under the maximum allowable, using
Equation F.16:

( )
Vu max = 0.83 × 83 (0.8 ×1000 × 75)×10 −3 = 454 kN ( Vu > Vu max )

Therefore,

Vu = Vu max = 454 kN

F-10
APPENDIX G

STRUT-AND-TIE MODEL

Rangan (1997) developed a simplified strut-and-tie model for the analysis and design of
shear walls subjected to inplane vertical and horizontal loads. The model is based on the
study undertaken by Gupta and Rangan (1996) on the HSC shear walls. This Appendix
includes an expression of the model associated with an example of its application.

For a shear wall under inplane vertical and horizontal forces, the shear strength, Vu is
calculated as:

⎛ N * ⎞⎟
Vu = t w d w ⎜ pl f sy + tan θ (G.1)
⎜ Ag ⎟⎠

where pl = Al/twLw; Al is the area of longitudinal steel in the wall on both faces in length
Lw; dw is the effective transverse length of the wall, equal to the transverse length of the
wall from centre-to-centre of the edge elements and taken as 0.8Lw for a wall with no
edge elements; N* is the design axial compressive force on the wall; Ag is the gross
cross-sectional area of the shear wall and θ is the strut angle given by:

dw
θ = tan −1 ( ) (G.2)
Hw
where Hw is the height of the wall. The strut angle θ is limited such that 30o ≤ θ ≤ 60o.
Similar to the design codes, Vu is limited to a maximum value given by:

k f ' t d sin θ cos θ


Vu ⋅ max = 3 c w w (G.1)
1.14 + 0.68 cot 2 θ

where k3 is the factor that accounts for the difference in compressive strengths of in-situ
concrete with the concrete test cylinder given by (Collins et al., 1993):

⎛ 10 ⎞⎟
k 3 = ⎜ 0.6 + ≤ 0.85 (G.2)
⎜ '⎟
⎝ f c⎠

Although Equation G.1 is in terms of the longitudinal steel, a minimum amount of


transverse reinforcement is required for an accurate prediction of the shear strength of
the wall. It has been suggested to follow the minimum amount set by AS3600 (Warner
et al., 1998).

Sample Solution

Using the model described above, in the following the shear strength of the specimen
SW5 is calculated. Input data for the wall SW5 is the same as that expressed in the
section F.3 with the addition of:
• Gross cross-sectional area of the shear wall (Ag) = 135000 mm2.

• Effective transverse length of the wall (dw) = 900 mm.

• Effective number of longitudinal N16 bars in each flange = 4.

• Total number of longitudinal W8 bars in web wall = 16.

• Longitudinal flange steel reinforcement (N16) yield stress (fsy) = 534.6 MPa.

G-2
The total area of longitudinal steel in the wall on both faces in length Lw is calculated as:

Al = 16 × 50.3 + 8 × 201.1 = 2413.6 mm 2

Therefore,

2413.6
pl = = 0.0322
75 × 1000

The effective yield stress for the longitudinal bars in length Lw is:

As web As flange 16 × 50.3 8 × 201.1


f sy eff = f sy web + f sy flange = × 498 + × 535 = 523 MPa
Al Al 2413.6 2413.6
Also, the strut angle from Equation G.2 is:

⎛ 900 ⎞
θ = tan −1 ⎜ ⎟ = 42
o
(30o ≤ θ ≤ 60o ∴OK)
⎝ 1000 ⎠

Now, using Equation G.1:

⎛ 1200 × 10 3 ⎞⎟
Vu = (75 × 900) 0.0322 × 523 +
⎜ tan(42) ×10 − 3 = 1564 kN
⎜ 3 ⎟
135 × 10 ⎠

To check Vu against the maximum allowable value, k3 factor is calculated from


Equation G.4:

⎛ 10 ⎞
k 3 = ⎜ 0.6 + ⎟ = 0.72 (k3 ≤ 0.85 ∴OK)
⎝ 83 ⎠

Then Vu max is determined from Eq. G.3 as:

(0.72 × 83 × 75 × 900)sin (42) cos(42) ×10 − 3 = 1013.4 kN


1.14 + (0.68 × cot 2 (42))
Vu ⋅ max = (Vu > Vu max)

Therefore the shear strength of SW5 is: Vu = Vu max = 1013.4 kN.

G-3

You might also like