Vous êtes sur la page 1sur 140

DESIGN OF HIGH-STRENGTH CONCRETE

MEMBERS : STATE-OF-THE-ART

Author:
A/Prof. Priyan Mendis, Reader, Dept. of Civil & Environmental Engineering,
The University of Melbourne
National Library of Australia cataloguing-in-publication entry

Mendis, P A (Priyantha Anuruddha)


Design of high strength concrete members: state of the art

ISBN 0 85825 808 0

1. High strength concrete. 2. Concrete construction.


I. Design of high strength concrete members: state of the art.

620.136

Published by Engineers Australia Pty Limited and distributed by EA Books, 2 Ernest Place,
Crows Nest, NSW 2065, phone 02 9438 1533, fax 02 9438 5934.
www.engaust.com.au
CONTENTS

1. Introduction 1-1 to 1-7

2. Engineering Properties of High-Strength Concrete 2- 1

2.1 Introduction 2- 1
2.2 Compressive strength 2- 1
2.3 Characteristic principal tensile strength 2- 7
2.4 Modulus of elasticity 2- 7
2.5 Shrinkage and creep in HSC 2-11
2.5.1 Shrinkage 2-11
2.5.2 Creep 2-14
2.6 Exposure classifications 2-15
2.7 Stress strain models 2-17
2.8 Fire resistance of high-strength concrete 2-20

3. Flexural Design Using High-Strength Concrete 3- 1

3.1 Introduction 3- 1
3.2 Deflections 3- 1
3.2.1 Flexural rigidity, EcIeff 3- 1
3.2.2 Recommendations for EcIeff 3- 3
3.2.3 Long-term deflections (Cl. 8.5.3.2 of AS3600) 3- 3
3.3 A rectangular stress-block model for high-strength concrete 3- 4
3.4 Minimum reinforcement 3- 9
3.4.1 Recommendation 3-10
3.5 Shear strength of high-strength concrete 3-11
3.5.1 Implications of using HSC to resist shear 3-11
3.6 Development length for bars in tension for fc<100MPa 3-13
3.7 Design Example 3-15

4. Design of High-Strength Concrete Columns 4- 1

4.1 Introduction 4- 1
4.2 Strength of high-strength concrete columns 4- 2
4.3 Interaction diagrams for HSC columns 4- 3
4.3.1 Computer generation of interaction curves for rectangular
Sections Spreadsheet INT_COL 4- 6
4.4 Laterally confined concrete columns-calculation of confinement
pressures 4- 8
4.4.1 Spread-sheets to calculate the confinement pressures 4-16
4.5 Ductility 4-16
4.5.1 Moment-curvature Relationship 4-17
4.6 Design of stirrups for high-strength concrete columns 4-21
4.6.1 Current Practices 4-21
4.6.2 Design Method-Detailed 4-25
4.6.3 Modifications Suggested for AS3600 4-28
4.7 Behaviour of slender high-strength concrete columns 4-30
4.7.1 Recommendations for slender columns 4-30
4.8 Transmission of high-strength concrete column loads through
slab 4-31
4.9 Design Examples 4-36
4A Appendix Development of Moment-Curvature Relationships 4-41

5. Punching Shear Resistance of High Strength Concrete Slabs 5- 1

5.1 Introduction 5- 1
5.2 Code Design Provisions 5- 2
5.3 Recommendations 5- 4
5.4 Design Example 5- 5

6. Design of Reinforced High-Strength Concrete Walls 6- 1

6.1 Introduction 6- 1
6.2 AS3600-94 Wall Design Method 6- 2
6.2.1 Background 6- 3
6.3 The American (ACI 318-95) Wall Design Method 6- 6
6.4 Comparison of The Code Wall Equations 6- 7
6.5 Recommendations For HSC Walls 6- 8
6.6 Application of Formulae in Core Walls 6- 9
6.6.1 Case Study 6-10
6.6.2 Case Study 6-13
6.7 Other Formulae 6-15
SPREADSHEETS
Chapter 3
BEAM_HSC - Flexural Design of HSC members includes Moment Capacity and Shear
checks

Chapter 4
i. INT_COL, Interaction Diagram for Column Sections
ii. CONFINE - Confinement pressures for Column Sections
iii. MOM_CURV Moment-Curvature Relationships for Beam and Column Sections
iv. SPAC_LIGS Ligatures for HSC Columns
FOREWORD
by Prof. Peter LeP. Darvall
Deputy Vice-Chancellor (Research and Development), Monash University

Advances in the cement and concrete technology in recent years have lead to substantial
increases in the compressive strength of concretes available at competitive prices. For investors,
this offers the hope of substantial savings in structural costs, and increased returns. For
designers, the prospect of using a material with uniformly enhanced properties is seductive.

But we should make progress carefully, and on the basis of the best research and analysis.
Pitfalls, as ever, await those who would assume that concrete design parameters and procedures
may be simply projected from existing knowledge and codes of practice into the realm of what
are substantially new materials.

This new book is thus a timely contribution to the information base for concrete design. It
synthesises recent advances in our knowledge of the properties of high strength concrete, and of
appropriate design methodologies. The explanations are clear for students, and the format
convenient for designers.

The author, Dr Priyan Mendis, is well qualified for this task. His fundamental laboratory and
analytical research has been internationally recognised in the leading professional journals. He
has had substantial experience with outstanding consultants, working at the leading edge of
concrete structural design. Finally, his extensive teaching experience has honed his skills in
presenting material in the best possible way.

This book is an important contribution to progress in the concrete building industry.


PREFACE

Many structures built now have at least some components constructed with high-strength
concrete (HSC). At present, the Australian Concrete Structures Code, AS3600 and other
major codes around the world cover only normal strength concrete. This publication provides
state-ofthe-art information on high-strength concrete and will become the basis of design in
HSC before the code is revised. This book is mainly written for practising engineers.
However it will also be useful for researchers, students and others involved in the
construction industry.

The book is divided into 6 chapters. Both approximate and rigorous methods are presented
here. Areas where research is particularly inadequate are noted. Chapter 1 provides a general
introduction including the various applications of high-strength concrete. Chapter 2 presents
some of the basic concepts in concrete technology that are essential for designers. Design of
flexural members is covered in Chapter 3. A design method for high-strength concrete
columns is presented in Chapter 4. Chapter 5 covers the punching shear resistance of HSC
slabs. The final chapter, Chapter 6 contains the design of walls. Several spread-sheets are
included with the book. It is believed that these programs, which simplify the design process,
will provide the reader with opportunities to explore and thereby optimise the design process.
The text assumes that the reader has an elementary knowledge of the concrete design process
and is reasonably familiar with AS 3600 relating to design considerations. As the theory,
background and the design examples in these Chapters are carefully covered in great detail
with step-by-step explanations, it should be relatively easy for the practising engineers to
understand and master the design of high-strength concrete members.

I wish to acknowledge the contributions of my former graduate students and research


associates, Raghu Pendyala, Senior Engineer at Meinhardt Ltd. (for Chapters 2 and 3),
Sujeeva Setunge, Senior Lecturer, RMIT (for Chapter 2), Joanne Pendyala, Engineer, Parks
Victoria (for Chapter 4) and Sam Fragomeni, Senior Lecturer, Griffith University (for Chapter
6) who helped me to put together the original document Changes to AS3600 to cover high-
strength concrete, which was published in 1994 by the Concrete Institute of Australia (Vic),
HSC sub-committee. In addition, I acknowledge the contribution of the reviewers of the
manuscript. I have incorporated many of their ideas into the final version as it is presented
here.
My consulting experience at Connell Wagner Ltd. and teaching/research experience at the
University of Melbourne and Monash University in Concrete Structures area were very
helpful in developing this book. Publication of this work was assisted by a publications grant
from the University of Melbourne. I would like to express my personal appreciation to Prof.
Peter LeP. Darvall, Deputy Vice-Chancellor (Research and Development) at Monash
University for encouraging me to commence research on high-strength concrete in 1980s and
writing the foreword for this book. Also I wish to express my sincere gratitude to Prof.
Graham Hutchinson, Head, Department of Civil and Environmental Engineering at the
University of Melbourne for his support and encouragement throughout the planning and
delivery of this book. Finally my deepest gratitude to my wife, Champika for her patience and
encouragement during the preparation of this book. It is appropriate to dedicate this book to
her.

Priyan Mendis
June, 2001
DESIGN OF HIGH-STRENGTH CONCRETE
MEMBERS : STATE-OF-THE-ART

Author:
A/Prof. Priyan Mendis, Reader, Dept. of Civil & Environmental Engineering,
The University of Melbourne

1-1
CHAPTER ONE

INTRODUCTION

The advancement of material technology and production methods has led to the development
of higher grades of concrete strengths. The use of high-strength concrete (HSC) elements for
concrete structures has proven most popular, with economy, superior strength, stiffness and
durability being the major reasons for its popularity. With future advances in concrete
technology, the definition of high-strength concrete will undoubtably change, however for the
purposes of this book HSC is considered to be concrete with a 28 day characteristic
compressive strength of at least 50 MPa. Presently, concretes with strengths up to 130 MPa
are used in many projects around the world [1-11]. These concretes can be produced using
conventional production procedures. Quality control is an essential part of the production of
high-strength concrete, hence full collaboration between the material and ready-mixed
concrete suppliers, engineers and builders is required for the effective application of this
valuable product. High-strength concrete will normally consist of not only Portland cement,
aggregates and water, but also superplasticizers and supplementary cementitious material.

A significant cost saving can be achieved by utilizing High-strength Concrete since member
sizes can be reduced. Table 1.1 presents typical applications of HSC in recently constructed
infrastructure. Structural elements and components where HSCs were used are highlighted
and also outlined are the advantages offered in design and construction by application of
HSC. Despite the benefits of using high-strength concrete, some structural engineers are
reluctant to use this material due to the lack of provisions in the design standards [11].

The main purpose of this book is to review and present the latest developments in the design
of high-strength concrete members up to 100 MPa. The previous Australian code, AS3600-
1994 [12], the new code, AS3600-2001 [13] and many other concrete codes around the world
cover only normal strength concrete (NSC). Extrapolating these design rules intended for

normal strength concrete to high strength concrete (HSC), 50 <fc<100 MPa, is not
appropriate. High-strength concrete is structurally a different material and rules applicable to
normal strength concrete are not always conservative when applied to high-strength concrete.
Due to the variations in fracture modes, microstructure and the differences brought about by

1-2
various additives like silica fume, fly ash, superplasticizers, etc., the empirical design rules

originally intended for concrete strengths f'c < 50 MPa, need to be re-evaluated. The main
concern regarding the use of high-strength concrete is the reduction in ductility with the
increase in compressive strength observed under uniaxial compression. This issue is
addressed in detail in this book.

Individual design rules given in national and international design codes need to be
experimentally verified and categorised with a view of taking further action. Although interest
in high-strength concrete has been around for some time (since the 70s) and a number of
research projects have been undertaken, some design issues are still not fully addressed. There
is also a need to consolidate the knowledge and make it readily available to the practitioners
for whom the research was intended in the first place.

Engineering properties are covered in Chapter 2. Other chapters address the design of flexural
members, columns, slabs and walls.

Table 1.1 Typical Applications of High-strength Concrete


Type of Applications Advantages of HSC
Structure
High Rise Columns in parking Strength, workability and
Buildings bays/high rise buildings [Fig. pumpability
1.1] Reduction of cross-sectional
High rise frames dimensions
Residential buildings Greater rigidity of the frames
Deck slabs and beams Overall economy and faster
Service cores [Fig. 1.2] construction, most cost
Shear walls and outriggers effective building material in
many cases
Structures with HSC result in a
larger rentable area
Shallower floor system leading
to reduced height of building
Increased punching shear
resistance in slabs
Precast/ Spun transmission poles Allows for a higher degree of
Prestressed Prestressed roof elements prestress [6]
industry Precast beam and column Reduction in weight; improved
elements handling
Prestressed piles and sheet Higher cracking loads
piles Reduced stripping time for
Precast tube elements for removal of formwork and

1-3
sewerage reduced time for transfer of
Tunnel lining prestress, reduced time for
Micro-tunneling segments lifting, etc.
Tiltup panels Improved durability in
aggressive environment
(sewerage and allied
applications)
Improved punching shear
resistance in slabs
Higher radial pressure capacity
Case studies show that HSC is
more cost effective than cast
iron hoops or normal strength
concrete in the case of tunnel
lining
Slabs and Residential slabs on ground Provides a low relative
Pavements Road/ Highway pavement humidity due to self desiccation
slabs to avoid moisture damage
Garage floors Shorter period of drying
Low maintenance needs;
improved durability
Better abrasion resistance
Overall a more favourable life
cycle cost
Bridges 17 - 35 m span beam girders Case studies show that precast -
and super-structure of pretensioned bridges are more
multiple span bridges, in economical than steel-concrete
precast, insitu and prestressed composite girders
HSC [Fig. 1.3] Durability
Submerged concrete bridges The use of HSC results in
for underwater pipelines smaller cross-sections and
Cable stayed bridges 150 - savings in foundations and
425m span material handling costs
Highway overpasses Ability to increase spans [6]
Repairs (strengthening of Low creep and shrinkage
bridge piers)
Marine Offshore oil production Increased strength, durability
structures platforms and buoyancy (in the case of
and other Floating bridge (pontoon) offshore platforms and
structures in Sea-wall panels pontoons)
contact with Slabs and walls of sludge Reduced wave and wind
water digesters loading by virtue of reduced
Water tanks dimensions
Tunnels Low shrinkage (reduced
shrinkage cracks)
Earthquake Frames Reduced inertial loads due to
resistant Walls reduced dimensions
design Enhanced ductility under
[14,15] flexure
Higher stiffness for sway
control

1-4
Fig. 1.1 The Petronas twin towers (worlds tallest building) at construction stage
(up to 80 MPa concrete was used for these columns) [16]

Fig. 1.2 A typical application of High-strength Concrete


(Core structure of Telecom Corporate Building in Melbourne under construction)

Fig. 1.3 Joigny Bridge in France (80 MPa concrete was used to construct the bridge)[3]

1-5
REFERENCES

1. Russell, H., Structural Design Considerations and Applications, Shah, S.P. and Ahmad,
S.H., eds. High performance concretes and applications / Edward Arnold Publishers,
UK, 1994.
2. Nagataki, S. and Sakai, E., Applications in Japan and SE Asia, Shah, S.P. and Ahmad,
S.H., eds. High performance concretes and applications / Edward Arnold Publishers,
UK, 1994.
3. Malier, Y, Ed. High-Performance Concrete - From Material to Structure, E&FN Spon,
Loandon, 1994.
4. CEB Bulletin No. 222, Application of High performance concrete, Lausanne, 1994.
5. Mendis, P.A. and Pendyala, R., "High-strength/High-performance Concrete in Australia
Design and Applications", Proceedings of the 4th World Conference on Utilization of
High-strength/High-performance Concrete, Paris, May, 1996, pp. 1581-1590.
6. Mendis, P., Nicholls, S. and Duffield, C., Optimum use of High-performance Concrete in
Prestressed Concrete Super-T Bridge Beams, PCI Journal, Vol. 45, No.3, May-June,
2000, pp. 56-65.
7. Day, K, HPC in SE Asia, Proceedings of a workshop on Technology, Design and
Applications of HSC/HPC at the University of Melbourne, Australia, Ed. P. Mendis, Feb.
1994.
8. Dinale, D., Observation in the use and placement of High performance concrete in the
manufacture of precast concrete elements, Proceedings of a workshop on Technology,
Design and Applications of HSC/HPC at the University of Melbourne, Australia, Ed. P.
Mendis, Feb. 1994.
9. Burnett, I, High performance Silica Fume Concrete Application, Proceedings of a
workshop on Technology, Design and Applications of HSC/HPC at the University of
Melbourne, Australia, Ed. P. Mendis, Feb. 1994.
10. Dean, B., Applications of High strength Concrete- Consultants Viewpoint, Proceedings
of a workshop on Technology, Design and Applications of HSC/HPC at the University of
Melbourne, Australia, Ed. P. Mendis, Feb. 1994.
11. Pendyala, R., Mendis, P.A. and Baweja, D., Towards the development of New Codes and
Standards to increase the Field Application of High-performance Concrete, Concrete 97
Conference. Adelaide, Australia, May, 1997, pp. 175-186.
1 2 .AS3600, "Concrete Structures", Australian Standard (until 2001 June), Standards
Association of Australia, Sydney, 1994.

1-6
13. AS3600-2001, "Concrete Structures", New Australian Standard, Standards Association of
Australia, Sydney, 2001.
14. Mendis, P.A., "Design Applications of High-strength Concrete in Seismic Regions",
Special Journal Publication on "High-strength Concrete in Seismic Regions", American
Concrete Institute, SP176, 1998, pp 437-443.
15. Panagopoulos, C., Mendis,P. and Portella,J., Seismic Performance of Frame Structures
with High-Strength Concrete and 500MPa Steel, Proceedings of Concrete 99
Conference, May 1999, pp. 585-597.
16. Mohamad, H., Choon, T., Azam, T. and Tong, S., The Petronas Tower The Tallest
Building in the World, Proceedings of the fifth world congress, Netherlands, 1995, pp.
321- 357.

1-7
CHAPTER TWO

ENGINEERING PROPERTIES
OF HIGH-STRENGTH CONCRETE

2.1 INTRODUCTION

High-strength concrete (HSC) is being commonly used for construction of tall buildings and
many other infrastructure projects in Australia and overseas. Advances in superplasticiser and
silica fume technology have greatly assisted in making this possible. The aim of this chapter is
to give an overview of the structural engineering properties and characteristics of HSC, in light
of the recent experimental and theoretical research and published results. There are other non-
structural benefits of using HSC, for example the improved durability of the material which is
a result of reduced porosity and the use of high-quality materials. However, the topic of
durability is not discussed here in detail.

Without any doubt, HSC differs in many respects from normal or ordinary strength concrete
(NSC). For the purposes of this book HSC is defined as concrete with compressive strength,

fc, in the range of 50 - 100 MPa. NSC is concrete with fc < 50 MPa.

2.2 COMPRESSIVE STRENGTH

Enhanced compressive strength is the most important of HSCs functional properties.


Admixtures such as silica fume or fly ash are not essential to the manufacture of high-strength
concrete, however, the incorporation of these mineral admixtures, particularly silica fume does
facilitate the process. The main reason for the spectacular increase in concrete strength in silica
fume concrete is the creation of a dense concrete matrix enabled by the uniformly distributed
fine silica fume particles in between larger cement particles. The use of superplasticisers and
good compaction by vibration aids in the densification process lead to the higher strength.
According to de Larrad and Malier [1], the microstructure of high-strength concrete is very

2-1
dense and amorphous and contains very little free water. It has a very low-porosity and lacks
the accumulation of lime crystals, as in the case of NSC. Theoretical work shows that
compressive strength should increase with the square of the compactness of the hardened
cement paste [1].

Test methods to determine the characteristic compressive strength of NSC are applicable to
high-strength concrete cylinders. The only exception being that in some circumstances the
limitations in the capacity of testing machines or other facilities may necessitate the use of
smaller cylinders of size 100(D)x200(H). Ting et. al. [2] showed that there is very little
variation in the tested strength of HSC when cylinders with smaller diameters are used, so
long as an aspect ratio of 2 or more is maintained to minimise the end restraining effect.
However, the Canadian Standard CSA23.1 [3] stipulates a reduction in concrete strength of
5% when cylinders with 100(D) x 200(H) are used.

High-strength concrete cylinders are more susceptible to end imperfections than normal
strength concretes due to the higher stresses in the concrete specimen. The ends have to be
smooth, parallel to each other and perpendicular to the axis of the cylinder. There are three
preferred methods of end preparation -
Smoothly ground ends
High-strength sulphur capping
Unbonded restrained rubber capping
More details and references regarding end preparation and the relative advantages and
disadvantages of these methods are described by Setunge [4].

Some of the important factors that affect the strength of HSC are:
Curing
Whilst there are conflicting reports on the relative importance of curing on strength gain of
HSC cylinders compared with NSC cylinders, curing procedures similar to those followed for
NSC members are sufficient and should be adopted. However, the possibility of self-
desiccation and high early shrinkage of HSC requires continuous moist curing for a longer
period.

2-2
De Larrad [5] reports that self-desiccation is probable in HSC and hence specimens cured in
water will absorb water, thus increasing the strength of the concrete. An opposite view is
expressed by some who argue that water evaporation from a NSC cylinder is greater than that
from a HSC cylinder. Therefore, the strength development of a NSC cylinder will be more
affected by deficient curing than the strength development of a HSC cylinder.

Studies by Aitcin [6] on curing of HSC show that HSC members have a delayed response to
strength gain. Aitcin suggests that due to the low permeability of high-strength concrete it
takes considerable time for water to penetrate the concrete and contribute to the hydration
process, hence longer periods of moist curing of HSC specimens is recommended.

Water/binder(w/b) ratio and cement content


HSC usually contains one or two mineral additives which are used as partial replacement for
cement. Therefore, the term water/cement (w/c) ratio used in reference to NSC is replaced by
w/b ratio, where the binder is the total weight of the cementitious materials
(cement+additives). The minimum w/b ratio for full hydration of cement pastes is
approximately 0.36 [7]. For NSC this limit is usually exceeded for workability requirements.
However, in the case of HSC, complete hydration is not essential for full strength to be
attained and therefore it can be made with w/b ratios less than 0.36. HSCs have been made
with w/b ratios as low as 0.2. However, high dosages of superplasticisers are required to
maintain workability [8]. Reporting on an experimental study, Patnaikuni and Patnaik [7]
suggest that a w/b ratio of 0.23 is an optimum value for maximum compressive strength of
very high-strength concrete mixes.

Setunge [9], while commenting on the work done by others, reports that increasing the cement
content in a mix beyond 500 kg/m3 does not significantly increase the compressive strength of
the mix. Other disadvantages of using a mix with cement content greater than 500 kg/m3 are
that the mix will be very sticky and difficult to handle and also concretes produced with high
cement contents may lead to thermal cracking of the members. Furthermore, cement is usually
the most expensive ingredient and hence mixes with high cement contents are usually not
economical. It is better to increase the binder content by using mineral admixtures rather than
increasing the cement content.
2-3
In an experimental study, Ting et al. [10] concluded that about 10% replacement of cement by
silica fume is about the optimum dosage. They found that at the optimum 10% level the
strength increased by approximately 20%. The use of ultra fine mineral admixtures, like silica
fumes, generate the higher strength due to the following two actions [11]:
a) Filler effect - the fine grains fill up the larger voids in between the cement particles, thus
reducing the water demand
b) Pozzolanic effect - The siliceous particles react with lime released by the cement in the
presence of water, to generate products of hydration.

Influence of mineral admixtures on compressive strength


As mentioned earlier, the incorporation of mineral admixtures such as silica fume, fly ash, slag
or rice-husk ash facilitates the manufacture of high-strength concrete. However, the
incorporation of these mineral admixtures is not mandatory.

Silica fume is a by-product of electric arc furnaces. These are used to manufacture ferro-silicon
and silicon metal alloys. Silica fume contains mainly non-crystalline silica in the form of very
fine particles (0.1 microns average diameter). Due to its finess as compared to other mineral
admixtures, silica fume is by far the most effective mineral admixture that can be used to
manufacture HSC as compared with other commercially available mineral admixtures. As
described later, it is essential to produce very high-strength concrete.

Fly ash is a by-product of the combustion of pulverised coal in thermal power plants. It is
removed as a fine dust by mechanical extractors, electrostatic precipitators or fabric filters.
Fly ash can be included into concrete either blended with cement or directly introduced as an
additional cementitious material at the concrete mixing plant. Typical applications are in
pumped or in superplasticised concretes, particularly where heat of hydration is considered to
be a problem.

Slags are by-products of the metallurgical industry. Slags that are most commonly used in
concretes for building and construction applications usually originate from iron blast furnace
facilities. These slags are glassy (amorphous) materials and are obtained by melt
2-4
waterquenching. In fresh concrete, slags tend to improve the workability of the concrete due
to their angular shape and smooth surface texture. However, their relatively slow reaction
rates [when compared to ordinary Portland cement (OPC)] cause increases in setting time and
bleeding tendencies. Slag cement concretes generally produce less heat of hydration than do
equivalent OPC concretes. Reductions in temperature rise and the associated likelihood of
thermal cracking can be reduced with slag inclusion into concrete, but such reductions only
become significant at higher slag replacement levels.

Superplaticizers
Superplasticizers are essential to produce good workable high-strength concrete. There are,
basically three principal types of superplasticizers:
(i) lignosulfonate-based (ii) melamine sulfonate and (iii) naphthalene sulfonate.
In general, a combination of the above types is used for high-strength concrete. The amount of
superplasticizer to be added to a mix is governed by the required workability.

Coarse Aggregates
As compared to NSC, the crushing strength of coarse aggregates has a significant effect on the
strength of a HSC mix. According to Setunge [3], higher strength aggregates do not necessarily
produce higher-strength concretes. A more desirable property is the compatibility of the
stiffness of the aggregates and the mortar. The ideal material will be crushed rock with low
stiffness and high strength. It is generally observed that smaller aggregates are desirable to
produce high strength concrete due to:
a) reduction in the water accumulating near the coarse aggregates
b) larger available surface area for bonding with cement matrix

For commercial applications, taking into account the economy of production, workability and
shrinkage and creep, well graded aggregates of 14-20mm size are recommended.

In another experimental study, Mak and Sanjayan [12] report that:


(a) the compressive strength of HSC is dependent on the type of aggregate used.
(b) gap graded aggregates yield higher strength but lower workability

2-5
(c) a higher percentage of sand increases the workability of the HSC mix but reduces the
compressive strength and
(d) 25% increase in strength was observed with a 8% replacement of cement by silica fume.

Aitcin [13], for the purposes of discussion divided high-strength concrete and very high-
strength concrete into five categories and discussed the relative importance of various factors
on the strength of concrete.

Category I - 50 -75 MPa - This grade of concrete can be manufactured using good quality but
generally used materials, existing production technology and a w/b ratio of about 0.40. There
is no particular need to incorporate mineral admixtures to attain this grade of concrete.
Superplasticisers may be used to achieve the required workability at the low w/c ratio and also
for slump recovery at the job site.

Category II - 75-100 MPa - To achieve this strength, high quality, generally used materials are
required. However, due to the very low w/b ratio of about 0.25 - 0.30, superplasticisers are
required to achieve adequate workability. The use of mineral admixtures are also strongly
recommended. The coarse aggregates have a significant influence on the resulting strength of
the concrete and they must be round or cubic in shape.

Category III - 100-125MPa - Very high quality materials, efficient mixing techniques and
stringent quality control are needed. The w/b ratio must be lowered to 0.22-0.25. If high
quality materials are not locally available they must be imported. High dosages of
superplasticisers are essential in conjunction with silica fume.

The other two categories refer to 125 MPa and beyond and will not be discussed here as they
are not commonly used, difficult to achieve in the field and beyond the scope of this book.

Manufacturers of high-strength concrete need to have in place a quality assurance programme


which controls the process from contract review through to onsite delivery [14]. Some typical
commercial mix proportions taken from Refs [15] and [16] are given in Table 2.1A. A typical
high-strength concrete mix used in Australia is given in Table 2.1B. This mix was used for the
2-6
53 storey Century Tower in Sydney. As seen the concrete was required to have an elastic
modulus of 45 GPa at 90 days in addition to the high strength.

2.3 CHARACTERISTIC PRINCIPAL TENSILE STRENGTH

The tensile strength of HSC is significantly greater than that of NSC, though to a lesser extent
than the compressive strength. Fracture surfaces are smooth, indicating the homogeneity of
the material. The densification of the concrete matrix and the aggregate-matrix transition zone
explains the improvement of the tensile strength. The indirect methods for evaluating the
tensile strength in accordance with the appropriate parts of AS1012- Methods of Testing
Concrete [17] are adequate for the tensile testing of HSC specimens.

According to AS3600 [18], the characteristic flexural tensile strength of concrete, fcf, is given
by Eq. (2.1).

f ' cf = 0.6 f ' c (2.1)


The ACI state-of-the-art report [19] recommends a higher value as given in Eq. (2.2).

f ' cf = 0.94 f ' c (2.2)


For high-strength concrete, this book recommends the use of the AS3600 equation.

2.4 MODULUS OF ELASTICITY

The modulus of elasticity (Ec) of HSC is dependent on parameters such as the volume of

aggregates, the modulus of the paste and the modulus of the aggregates [20]. Complex equations
incorporating these factors and as suggested by various researchers are reviewed by Setunge [4].
These equations are not easily or readily applied and are therefore not suitable for a design
office. Consequently AS3600 suggests a simplified, empirical formula to predict the elastic
modulus.

2-7
Table 2.1 A Examples of commercially produced high strength concrete mixes

M ix Number 1 2 3 4 5 6 7 8 9 10
water (kg/m3) 170 195 158 165 145 160 135 151 144 130
cement (kg/m3) 425 505 564 451 315 475 500 475 564 513
fly ash (kg/m3) - 60 - - - 59 - 104 - -
slag (kg/m3) - - - - 137 - - - - -
silica fume (kg/m3) 30 - - - 36 24 30 74 89 43
coarse aggregate (kg/m3) 1033 1030 1068 1030 1130 1068 1110 1068 1068 1080
fine aggregate 705 630 647 745 745 659 700 593 593 685
water reducer L/m3 4 0.98 - - 0.9 - - - - -
retarder L/m3 1.0 - 1.12 4.5 1.8 1.04 4.5 1.51 1.47 -
superplasticizer L/m3 3 - 11.6 11.2 5.9 11.6 14 16.4 20.1 15.7
1 5 1 5 2
w/cementious ratio 0.4 0.35 0.28 0.37 0.31 0.29 0.27 0.23 0.22 0.25

fc 28-day (M Pa) - moist cured 59 65 78.6 80 83 88.5 42.5 107 118.9 119

fc 91-day (M Pa) - moist cured 66 79 86.5 87 93 100.4 106.5 119.3 131.8 145

1. La Grande Arche, Paris 6. HSC mixture 2 in the Chicago area


2. Water Tower Place (1975) 7. Scotia Plaza, Toronto (1987)
3. HSC mixture 1 in the Chicago area 8. HSC mixture 3 in the Chicago area
4. Joigny Bridge, France (1989) 9. HSC mixture 4 in the Chicago area
5. La Laurentienne Building, Montreal (1984) 10.Two Union Square, Seattle (1988)

2-8
Table 2.1B - Specification Requirements for Concrete and Achieved Field Performance
for Century Tower Sydney
M ix Constituent Lower Columns and Core Level 9 Transfer Slab
or Property Specification Achieved Field Specification Achieved Field
Requirement Performance Requirement Performance
Strength Data fc - 80 M Pa 105 M Pa fc - 60 M Pa 72 M Pa
90 Day Ec (GPa) 45 48 - -
56 Day Drying 500 650 500 - 530
Shrinkage (mstrain)
Creep (AS 1012) 15 he/M Pa @ 3 - -
months
Temperature - - 10C reduction Liq. nitrogen cooled
Control in concrete temp Target - 14 2C at
time of discharge
Total binder - 640 kg/m3 - 440 kg/m3
Binder Type - SL cement , - SL cement, M icropoz
M icropoz & fly ash
& fly ash
Aggregate inclusion 10 mm max 10 mm max Penrith Penrith Crushed
Basaltic Crushed River Gravels River Gravels
aggregates*
Admixture inclusion Use of Water reducer/retarder Water red./retarder
superplasticisers and superplasticiser and superplasticiser
M ax. W:B - 0.25 to 0.28 -
Initial slump (mm) - < 30 - 60
Final slump (mm) - 120-180 - 120
(depending on
element)

*This table was taken from the document A guide for Designers and Specifiers CSR Constrcution
Materials, NSW Project Case Study 1

2-9
Setunge [9] has shown that the existing AS3600 formula of:
Ec = 0.43r1.5 f 'c 20% (2.3)

has the tendency to overestimate the elastic modulus of HSC.

Mendis et al. [21] proposed that the following expression is appropriate to predict the elastic
modulus of all grades of NSC and HSC concretes with various types of aggregates:

Ec = 0.43hr1.5 f 'c 20% (2.4)

where, h = 1.1-0.002fc 1.0

The term h has the value of 1 for NSC and is less than 1 for HSC. This formula was derived

by calibrating experimental results and comparing them with the widely used Carrasquillo et
al. [20] formula given in Eq. (2.5).

( )
1.5
r
Ec = 3320 f 'c + 6900 c (2.5)
2320
Fig. 2.1 shows a comparison between different formulae presented above. As can be seen, the
Carrasquillo et al. formula gives lower values at higher strengths.

55000
Elastic Modulus (MPa)

50000 AS3600

45000
Mendis et al.
40000
35000
30000
Carras quil o et al.
25000
20000
25 50 75 100
Concrete Strength (MPa)

Fig. 2.1 Comparison of Elastic Modulii calculated using different formulae


Poisson ratio retains the value of 0.2 - 0.25 as in NSC.

2-10
The nominal density of normal weight HSC also retains the current value of r = 2400 kg/m3 or

that determined by tests in accordance with AS1012.2.

2. 5 SHRINKAGE AND CREEP IN HSC

2.5.1 Shrinkage

Shrinkage in concrete comprises three distinct components viz. Plastic shrinkage, Autogenous
shrinkage and Drying shrinkage. Most codes and procedures do not differentiate between the
three types of shrinkage. The empirical procedures laid out in most codes determine the total
shrinkage at the end of a stipulated time or design period.

For HSC, which is expected to be produced with low w/b ratios and better aggregates, the
total shrinkage should decrease in comparison with NSC specimens. Because of the high paste
content in HSC, autogenous shrinkage is expected to be higher, and in some cases doubled [1].
However, since the specimens contain very little water, the drying shrinkage will be
significantly reduced. All experimental results point to the fact that the total shrinkage values
for HSC specimens are less than those obtained for NSC specimens.

Autogenous shrinkages occur with the loss of water used in hydration of cementitious
materials. This shrinkage occurs in the short term. Since the autogenous shrinkage is higher for
HSC, the early shrinkage in HSC is higher, and this aspect should be carefully considered by
the structural engineer. In particular, this behaviour is important in the design of
indeterminate structures consisting of many restraints at supports and also has implications
for prestress losses in prestressed concrete design. If adequate precautions are not taken,
shrinkage cracking in the structure is liable to occur. The structural engineer should consider
incorporating the support restraints after the high early shrinkage has occurred. Swamy [22]
reported that for HSC 20 - 50% of the total 700 day shrinkage occurred in the first 7 days. De
Larrard and LeRoy [23] have carried out measurements of autogeneous shrinkage of HSC
mixes and proposed a relationship between the composition of the mix and the autogeneous
shrinkage. Some results are presented in Table 2.2. As can be seen, HSC specimens show high
early shrinkage and a lower final shrinkage.

2-11
Table 2.2 Comparative shrinkages in NSC and HSC test specimens [23]

Mixes No. 1 2 3 4 5 6
Cement, kg/m3 350 450 456 453 453 421

Silica Fume, kg/m3 - - 36 36 36 42

Super-plasticizer, l/m3 - 4.5 7 6.6 3.6 7.9

Retarder, l/m3 - 0.9 0.5 0.5 0.5 -

Water, kg/m3 195 168 151 175 188 112

W/C ratio 0.56 0.37 0.33 0.39 0.42 0.27

fc (MPa) 40 78 94 83 74 101

Autogeneous Shrinkage 90 90 290 200 140 150


(at 1 yr)x10-6
Drying Shrinkage 290 90 120 190 260 110
-6
(at 1 yr)x10

Total Shrinkage 380 180 410 390 400 260


-6
(at 1 yr)x10

-6
Investigations by Yue and Taerwe [24] indicate that the lower limit of 600x10 is a

reasonable estimate of the basic shrinkage strain of HSC. Figs. 2.2(a) and (b) show the
comparison of shrinkage values predicted by three codes and experimental work. Shrinkage
strains determined by tests are approximately 30% below the values predicted by the
AS3600-1994 procedure [25].

2-12
0.0007

0.0006

0.0005

Shrinkage Strains
0.0004

0.0003

0.0002
CEB-FIP (1990)
NS3473
0.0001 AS3600
Expt. (Yue & Taerwe (1993))
0
0 200 400 600 800 1000 1200
Time Days

Fig. 2.2(a) Drying Shrinkage of an 80 MPa concrete kept at 50% RH and 230C [25]

0.0012

0.001

0.0008
Strain

0.0006

0.0004

0.0002

0
0 100 200 300 400 500
Time (Days)
AS 3600 CEB-FIP
NS 3473 BS 8110
Mix B, 103 MPa

Fig. 2.2(b) Drying Shrinkage of a 103 MPa concrete kept in an arid environment [25]

However, caution must be exercised in using the basic shrinkage strain values in design of
structural elements which are sensitive to differential settlement (like columns in tall
buildings). Every HSC mix has different characteristics and if the project is sufficiently
important and sensitive to shrinkage, appropriate tests and monitoring should be carried out
to determine the actual shrinkage values.

2-13
Gilbert [26] has suggested basic shrinkage versus concrete strength relationships for normal
and high-strength concrete. The AS3600-1994 [18] procedure assumed a tolerance of 30%

for the basic shrinkage strain, ecs.b, -6


of 700x10 . The basic shrinkage for NSC has been

-6
increased to 850x10 in the new AS3600-2001 [27] code. It is recommended that in the
absence of any other data, a value of 700x10-6 be used as the basic shrinkage strain for high-

strength concrete (fc > 50 MPa).

2.5.2 Creep

Creep is defined as the increase in strain under a sustained constant stress. Basic creep is the
creep encountered when no drying is involved and the drying creep is the additional creep that
occurs with simultaneous drying of concrete. Aggregates in concrete do not creep at the stress
levels encountered in concrete.

Creep decreases with an increase in strength of concrete. Thus, it is expected and experiments
confirm that HSC creeps to a lesser degree than NSC. A resulting advantage is that higher
proportions of the maximum stress can be endured by HSC for similar values of specific
creep.

Collins [28] and Persson [29] report that the single most significant factor that affects creep is

the stress level. The higher the stress level, the higher will be the creep strains. At 0.3fc,
50% higher creep strains were observed as compared to 0.2fc, and at 0.6fc the creep

strains were 3 times of those at 0.2fc.

Mendis et al.[21], while reviewing the work carried out by others, reported that a reduction in
w/c ratio in the HSC mix will lead to a reduction in creep.
The creep coefficient is defined as the ratio of total long term strain and initial elastic strain at

a particular level of stress. In AS3600 [18] values of creep co-efficient (fcc.b) are given for

concrete with compressive strengths up to 50 MPa and for a stress level of 0.4fc. Based on
an experimental and analytical study Setunge and Padovan [25] proposed basic creep factors

2-14
for HSC (Fig. 2.3). Gilbert [26] has suggested slightly higher creep factors (compared to
Setunge and Padovan) for high-strength concrete. Suggestions of Setunge and Padovan [25] and
Gilbert [26] are given in Table 2.3. It must be noted that the creep coefficients for concretes of
varying strengths depend on many factors other than the compressive strength, including mix
components and proportions, water/cement ratio, supplementary cementitious materials and
other additives. Basic creep factors suggested by Setunge and Padovan are recommended by
the author.

AS3600
5
Basic Creep Factor

Setunge & Padovan


4 (1997)
Other reported data
3
Proposed
2

0
0 20 40 60 80 100 120
Compressive Strength MPa

Fig. 2.3 Basic Creep Factor [25]

Table 2.3 Basic creep factors to extend AS3600 Table 6.1.8.1 to 100 MPa

Characteristic Compressive 20 25 32 40 50 60 80 100


strength, MPa
Basic creep factor 5.2 4.2 3.4 2.5 2.0 1.5 0.9 0.5
-Setunge and Padovan [25]
Basic creep factor 5.2 4.2 3.4 2.8 2.5 2.4 2.2 2.0
-Gilbert [26]

2.6 EXPOSURE CLASSIFICATIONS

It is noted in the commentary of AS3600 that carbonation and ionisation (an increase in the
reactive ion concentration such as chloride) are two factors that will influence durability in
terms of corrosion of steel. Exposure classifications set down in AS3600 and as given in Table

2-15
4.3 (of AS3600) are linked primarily to concrete strengths. The closer a structure is to sea
water, the higher the strength grade required for compliance. The use of concrete strength for
specification in this way is questionable and has been the subject of much debate for some
time. Recent data published for concrete in a marine environment (Exposure Classification C)
suggests that there are limitations associated with the use of compressive strength alone in
such specifications [30]. HSC consists of a more uniform microstructure and lower porosity

compared to NSC. This indicates a higher resistance to penetration of CO2 and Cl ions into
the concrete, thus reducing the corrosion of steel reinforcement. However, during production
of HSC, macrocracking due to plastic shrinkage, microcracking due to self-dessication and
thermal cracking may represent potential problems from a corrosion protection point of view.
HSCs are also being specified for a range of critical civil engineering structures where the
durability properties of the concrete are of paramount consideration and where design life
requirements of over 100 years are needed. It is therefore prudent to adopt the cover values
specified in AS3600 for 50 MPa concrete also for concrete strengths higher than 50 MPa.
However the designers may reduce the cover values according to the following table. It must

be noted that for fc< 70 MPa, the same values as given in AS3600 are recommended.

Table 2.4 Required Cover Values

Exposure Standard Formwork Rigid Formwork and Spun or Rolled


Classification and Compaction are Intense Compaction Members (Table
used (Table 4.10.3.2 are used (Table 4.10.3.5 of AS3600)
of AS3600) 4.10.3.4 of AS3600)
fc< 70 fc 70 fc< 70 fc 70 fc< 70 fc 70
MPa MPa MPa MPa MPa MPa
A1 20 20 15 15 10 10

A2 20 20 15 15 10 10

B1 25 20 20 15 15 10

B2 35 30 25 20 20 15

C 50 45 40 35 25 20

2-16
2.7 STRESS STRAIN MODELS

The stress-strain behaviour of HSC differs from that of NSC [Fig. 2.4]. The reasons for the
variance in behaviour are as follows:
For HSC, the stress-strain curve remains linear up to the highest value of the
stress/ultimate stress ratio. The internal microcracking that occurs in concrete as load is
applied is delayed until a large proportion of the ultimate load is reached. Therefore the
elastic response to compression is extended.
the strain at the peak stress increases with strength
the post peak branch becomes steeper as the strength increases (or reduced ductility). The
extensive ductility exhibited by lower strength concretes beyond maximum stress, caused
by the spreading of microcracks which form an interconnected network with large
redundancies to dissipate the energy is not observed in HSC.
the typical ultimate compressive strain decreases as the strength increases
the typical fracture surface of an ordinary concrete is rough. The fracture occurs along the
aggregate-matrix interface; the aggregates are not broken. For HSC, a typical fracture
surface is smooth and the cracks pass without any discontinuity through both the matrix
and aggregates. The fracture mode of HSC is more sudden than that for NSC [31].

Fig. 2.4 Uniaxial stress-strain curve for concrete [32]

2-17
It is important to know the details of stress-strain behaviour of HSC in order to determine the
full-range moment-curvature behaviour of HSC columns and beams. Various stress-strain
relationships for HSC have been proposed in the literature. A model developed at the
University of Melbourne is described below.

Modified Scott Model [32,33]


The Modified Scott Model has been based on the model for confined concrete developed by
Scott et al. [34] for NSC members. Mendis [35], based on an experimental and theoretical
investigation, recommended the Scott Model as being the most suitable to predict the full
range stress-strain behaviour of concrete beams. In order for the model to be applicable for
both NSC and HSC appropriate modifications were made to the Scott Model. The Modified
Scott model has been experimentally validated for HSC members [32,36].

The proposed Modified Scott Model is described by Eqs. (2.6) - (2.14) and Fig. (2.5).
For confined concrete:
The equation which defines the parabolic ascending portion of the curve is:

2 e e for
2
f = K f'c - e ecc (2.6)
ecc ecc

the equation for the linear descending portion is:

[
f = K f ' c 1 - Zm (e - ecc ) fres ] for e > ecc (2.7)

but not less than K f ' c fres


where,

fl
K =1+ 3 (2.8)
f 'c
and

0.5
Zm = Z 0
3 + 0.29 f ' c 3 h "
(2.9)
+ rs - 0.002 K
145 f c' - 1000 4 sh
2-18
f

Kf c

f c

Confined
Unconfined

K fc fres Zm Zm

fc f res
ec ecc 0.004 er e

Fig. 2.5 The Modified Scott Model

(
ecc = 0.24 K 3 + 0.76 ec ) = K *ec , and (2.10)

fres = Kf ' c ( 0.28 - 0.0032 f ' c ) 0 (2.11)

where:

fl = the confinement pressure (MPa) using the Mander method

outlined in Section 4.4.

rs = volumetric ratio of hoop reinforcement to concrete core

measured to outside of the hoops.

h'' = width of concrete core measured to outside of peripheral hoop

(mm).

sh = centre-to-centre spacing of hoop sets (mm).

4.26 f ' c
ec = , where Ec is the Modulus of Elasticity of concrete.
4 f ' c Ec

For unconfined concrete:


The equation which defines the parabolic ascending portion of the curve is:

2 e e
2
f = f'c - for e ec (2.12)
ec ec

2-19
the equation for the linear descending portion is:

[
f = f ' c 1 - Z ' m (e - ec ) > f ' res ] for e > ec (2.13)

but not less than f ' c fres


where:

0.5
Z' m = Z >0
3 + 0.29 f ' c (2.14)
- ec
145 f ' c - 1000
Z = 0.018fc + 0.55 (2.15)

3 + 0.29 f ' c
[Note: - e c 0, indicates a very steep slope beyond the ultimate
145 f ' c - 1000
load, hence a very large value (e.g. 1E5) should be allocated to Z ' m .

Application of the Modified Scott model for high-strength concrete columns is presented in
detail in Chapter 4.

2.8 FIRE RESISTANCE OF HIGH-STRENGTH CONCRETE

Fire resistance of concrete members is normally accomplished by structural adequacy and


insulation for a specified fire resistance period. Several researchers have concluded that
with the exception of spalling, which is defined as the detachment of pieces of concrete
when a concrete member is exposed to fire, there is no apparent reason to treat high-
strength concrete differently from lower strength concrete. The pieces can be large or small
and detachment can either occur explosively or pieces may dislodge and subsequently fall
(Fig. 2.6). Spalling can take place over the whole surface area of a member or in localised
areas [38]. The risk of spalling is higher in high-strength concrete due to the following
reasons:
(i) Low permeability of HSC retains the moisture inside the concrete resulting in a high
moisture content being present for prolonged periods.
(ii) Low porosity of HSC creates higher pore pressure.
(iii) HSC tends to be subject to higher compressive stresses than lower strength
concrete.

2-20
Sudden burst due to Water bound chemically and
the steam-pressure physically in cement paste

Fig. 2.6 Behaviour of HSC in Fire [37]

Conflicting observations have been reported in the literature on spalling of HSC. Some tests
have shown explosive spalling of HSC, while others have reported no difference to the
behaviour of NSC. In 1996, a comprehensive investigation was conducted at the National
Institute of Standards and Technology, on experimental and analytical studies on fire
performance of HSC. The key findings and a literature review are given by Phan [39]. Of
the ten materials test programs reviewed by Phan [39], five reported explosive spalling. Of
the five element test programs reviewed, three reported spalling. Also, it was observed that
explosive spalling did not occur for every specimen tested under identical conditions. The
reported temperature range when explosive spalling occurs is between 300C to 650C.
Concrete with dense pastes resulting from the addition of silica fume are more susceptible
to explosive spalling. HSC made with lightweight aggregate appears to be more prone to
explosive spalling than HSC made of normal weight aggregate concretes. Also, HSC
specimens heated at higher heating rates, and larger specimens are more prone to spalling
than specimens heated at lower rates and of smaller size.

Sanjayan [38] suggests the following methods to reduce the risk of spalling in HSC:
(i) Reduction of concrete covers (20 -25 mm) where possible;
(ii) Where it is not possible to provide small cover for structural steel, sacrificial steel
with small concrete covers (20-25 mm) may be provided to prevent spalling of
cover to the main steel bars. The sacrificial steel is not to be considered in structural
design for carrying loads; and

2-21
(iii) Use of fibre reinforcement to prevent spalling [see also ref. (37)]. The length of
fibres may be of importance. Short fibres, combined with a high fibre content may
even give adverse spalling effects.

According to the review by Phan [39], the material properties of HSC vary differently with
temperature as compared to those of NSC. The differences are more pronounced in the
temperature range of between 25C to about 400C, where higher strength concretes have
higher rates of strength loss than lower strength concretes. These differences become less
significant at temperatures above 400C. Compressive strengths of HSC at 800C decrease
to about 30% of the original room temperature strengths. The difference between the
compressive strength versus temperature relationships of normal weight and lightweight
aggregate appears to be insignificant, based on the limited amount of existing test data. The
tensile strength versus temperature relationships decreases similarly and almost linearly
with temperature for HSC and NSC. HSC mixtures with silica fume have higher strength
loss with increasing temperatures than HSC mixtures without silica fume. The failure of
HSC is more brittle than NSC at temperatures up to 300C. With further increase in
temperature, specimens exhibit a more gradual failure mode.

A summary of the current research projects on fire performance of HSC is given in the
proceedings of a workshop conducted by NIST [40]. This summary includes the research
programs at National Research Council of Canada and PCA (USA) on fire performance of
HSC columns and thermal properties of HSC. It was recognised by the workshop
participants that the amount of test data on fire-exposed HSC is insufficient relative to the
number of variables (concrete strength, aggregate types, test conditions, specimen size,
concrete density, concrete permeability, concrete porosity, heating rate, etc.). Also, it was
found that the variation of the stress-strain relationships of HSC with temperatures must be
established experimentally. Such relationships are not widely reported in the existing
literature but are essential for the development of constitutive models of HSC to predict
structural performance during a fire. A joint project is conducted by the University of
Melbourne and Monash University to suggest rules for fire design of high-strength concrete
walls and columns. These results will be published in the near future.

The following research needs have been identified.


Experimental studies:
Experimental studies should be designed to obtain a more complete body of data on the
fundamental behaviour of HSC at high temperature and, more importantly, to develop data

2-22
necessary for the development and validation of numerical models which can predict
moisture transport and the sudden spalling of HSC when subjected to fire. The experimental
study should consist of both materials testing (tests of fire-exposed HSC specimens) and
element testing (tests of fire-exposed HSC beams, columns, slabs, walls). The following
test variables should be considered for the materials test program:
test methods (with emphasis on the stressed tests)
aggregate type (siliceous, calcareous, lightweight)
water/cement ratio
addition of silica fume
moisture content and maturity at time of test
heating rate

The measurement should include compressive strength, tensile strength, modulus of


elasticity and compressive stress-strain relationships as functions of temperature. This data
will provide definable behavioural trends for the properties of fire-exposed HSC. Also, to
investigate the explosive spalling failure mechanism of HSC, measurements of internal
temperature, changes in permeability, porosity, and mass loss should be obtained.

The structural element test program should include columns with preload, beams and slabs.
Measurements of internal temperature should be obtained in addition of the furnace
temperature.

Analytical study:
The experimental data obtained from the materials studies can be used to develop a
material model to predict structural performance of HSC elements exposed to high
temperature. The basis for the model may be adopted from the established pore pressure
models, but including the measured properties (permeability, moisture content, thermal
conductivity, etc.) characteristic of HSC. Data obtained from element tests may be used for
model validation.

Development of Code Provisions:


The experimental and analytical data have to be synthesized and practical constitutive
models that are suitable for design purposes should be included in design codes to provide
guidance for fire design of HSC members.

2-23
REFERENCES

1. De Larrard, F and Malier, Y, Engineering Properties of Very High Performance


Concretes High-Performance Concrete - From Material to Structure,(Editor- Malier),
E&FN Spon, 1994, London, pp 85 -114.
2. Ting, E.S.K and Patnaikuni, I, Johanson, H.A and Pendyala, R.S. Compressive Strength
testing of Very High-strength Concrete 17th International Conference on Our World in
Concrete, 1992a, Singapore.
3. CSA 23.1 1994 Design of Concrete Structures, Canadian Standards Association,
Ontario, 1994.
4. Setunge, S., Engineering Properties of High-Performance Concrete, Proceedings of a
workshop on Technology, Design and Applications of HSC/HPC at the University of
Melbourne, Australia, Ed. P. Mendis, Feb. 1994.
5. de Larrard, F, A Mix Design method for High-Strength Concrete, proc. Conf. First
International Symposium on utilisation of High-strength Concrete, Stavanger, Norway,
1988.
6. Aitcin, P.C., High-performance Concrete, E & FN Spon, UK, 1998.
7. Patnaikuni I. and Patnaik, A.K., Design and Testing of High-Performance Concrete
International conference of Rehabilitation, Renovation, Repair of Structures,
Visakhapatnam, India, 1992.
8. Mak, S.L., Defining Performance and Basic Considerations for Materials and Mix
Design, Proceedings of a workshop on Technology, Design and Applications of
HSC/HPC at the University of Melbourne, Australia, Ed. P. Mendis, Feb. 1994.
9. Setunge, S. Engineering properties of High-Strength Concrete, Ph.D thesis, Monash
University, 1993.
10. Ting, E.S.K., Patnaikuni, I., Pendyala, R.S. and Johanson, H.A. Effectiveness of Silica
Fumes available in Australia to Enhance the strength of Very High Strength Concrete,
International conference on The Concrete Future, 1992(a), Kuala Lumpur.
11. de Larrard, F. A Mix Performance Method for High Performance Concrete, Engineering
Properties of Very High Performance Concretes High-Performance Concrete - From
Material to Structure, (Editor- Malier), E&FN Spon, 1994, London, pp 48 -62.

2-24
12. Mak, S.L. and Sanjayan, G., Mix Proportioning for Very High Strength Concrete,
Second National Structural Engineering Conference, Adelaide, October 1990.
13. Aitcin, P.C., How to Make High-Performance Concrete, Presented at the High-
Performance Concrete Seminar, National University of Taiwan, Taipei, 1992.
14. Munn, B., Manufacture, Procee Control and Quality Assurance, High-performance
Concrete - Technology, Design and Applications, Proceedings, Dept. of Civil and
Environmental Engineering, The University of Melbourne, 1994, Edited by P. Mendis.
15. Shah, S.P. and Ahmad, S.H., eds. High performance concretes and applications / Edward
Arnold Publishers, UK, 1994.
16. Malier, Y., ed. High performance concrete : from material to structure , E & FN Spon,
UK, 1992.
17. AS1012 Parts 10 and 11, "Methods of Testing Concrete", Standards Association of
Australia, 2000.
18. AS3600, "Concrete Structures", Australian Standard, Standards Association of Australia,
Sydney, 1994.
19. ACI Committee 363, "State of the Art Report on High Strength Concrete", ACI Journal
Proceedings, Vol.81, No.4, July-August 1984, pp 364-411.
20. Carrasquillo R.L., Slate F.O., Nilson A.H., " Microcracking and Behaviour of High
Strength Concrete Subjected to Short Term Loading", ACI Journal, Vol.78 (3), May-June
1981, pp 179-186.
21. Mendis, P.A, Pendyala, R.S. and Setunge, S. Requirements for High-Strength Concrete in
AS3600, High-Performance Concrete Sub-committee of the Concrete Institute of
Australia (Victoria), Melbourne, 1997.
22. Swamy, R.N., High-Strength Concrete- Materials, Properties and Structural Behaviour,
High-Strength Concrete Special publication, SP-87, ACI Detroit, 1986, pp. 119-145.
23. De Larrard, F. and Le Roy, R., The Influence of Mix Composition on the Mechanical
Properties of Silica-Fume High Performance Concrete, Fourth International ACI-
CANMET Conference on Fly Ash, Silica Fume, Slag and Natural Pozzolans in Concrete,
Istanbul, 1992.
24. Yue, L. and Taerwe, L.R. Empirical Investigations of Creep and Shrinkage of High-
Strength Concrete, Proc. of Conf. Utilisation of High-Strength Concrete, Lillihammer,
Norway, 1993, pp 1263 -1270.
2-25
25. Setunge, S. and Padovan C., Early Creep and Shrinkage in high-performance Concrete in
Arid environments, Proc. of Conf. Concrete 97, Concrete Institute of Australia, 1997,
Adelaide.
26. Gilbert, I., Serviceability Considerations and Requirements for High Performance
Reinforced Concrete Slabs, Proceedings HPHSC Conference, Perth, August, 1998, pp.
425-439.
27. AS3600-2001, Concrete Structures, New Australian Standard, Standards Association of
Australia, Sydney, 2001.
28. Collins, T.M., Proportioning High-strength Concrete to control Creep and Shrinkage,
ACI Materials Journal, Vol 86, No. 6, Nov-dec 1989 pp 576-580.
29. Persson, B.S.M., Self Desiccating high-strength Concrete Slabs, Proc. of Conf.
Utilisation of High-Strength Concrete , Lillehammer, Norway, 1993, pp 882-889.
30. Pendyala, R., Mendis, P.A. and Baweja, D., Towards the development of New Codes
and Standards to increase the Field Application of High-performance Concrete, Concrete
97 Conference. Adelaide, Australia, May, 1997, pp. 175-186.
31. Pliskin, L., High-Performance concrete - Engineering properties and Code Aspects,
High-Performance Concrete - From Material to Structure,(Editor- Malier), E&FN Spon,
1994, London,186-195.
32. Pendyala, R.S., The Behaviour of High-Strength Concrete Beams Ph.D thesis, The
University of Melbourne, 1997.
33. Mendis, P., Pendyala, R. and Setunge, S., Stress-strain Model to predict the full-range
Moment Curvature Behaviour of High-strength Concrete Sections, Magazine of Concrete
Research, UK, Vol. 52, No.4, 2000, pp227-234.
34. Scott B.D, Park, R. and Priestly, M.J.N Stress-Strain Behaviour of Concrete Confined
by Overlapping Hoops at Low and High Strain Rates, ACI Journal, 1982, 79, No. 1, 13-
27.
35. Mendis, P.A. Softening of Reinforced Concrete Structures Ph.D Thesis, Monash
University, 1986.
36. Kovacic, D., Design of High-strength Concrete Columns, M.Eng. Thesis, University of
Melbourne, 1995.
37. Breitenbucker, R., HSC C105 with increased fire resistance due to polypropylen fibres,
Proc. of Conf. Utilisation of High-Strength Concrete, Paris, 1996.
2-26
38. Sanjayan,G. , "Fire resistance of HSC", Proceedings of a workshop on Technology, Design
and Applications of HSC/HPC at the University of Melbourne, Australia, Ed. P. Mendis,
Feb. 1994.
39. Phan, L.T., Fire Performance of High-Strength Concrete: A Report of the State-of-the-
Art, NISTIR 5934, National Institute of Standards and Technology, Gaithersburg, MD,
1996, 105 pgs.
40. NIST Special Publication 919, Proceedings of International Workshop on Fire
Performance of High-Strength Concrete, Eds. L.T. Phan, N. Carino, D. Duthinh, E.
Garboczi, National Institute of Standards and Technology, Gaithersburg, MD, 1997, 164
pgs.

2-27
CHAPTER THREE

FLEXURAL DESIGN
USING HIGH-STRENGTH CONCRETE

3.1 INTRODUCTION

The Concrete Structures Code AS3600 2001 [1] only recognizes the use of concrete with
compressive strengths less than 50 M Pa. Due to the increased use of higher strength concrete, the
current AS3600 Code equations must be re-evaluated to also suit High-strength concrete.
Sections 3.2 through to 3.6 briefly review the current Code requirements for flexural strength
design of normal strength concrete (NSC) members and suggest rules for HSC members.

3.2 DEFLECTIONS

By using HSC, smaller cross sections with reduced dead load and longer spans can be
designed. For the smaller cross sectional members, excessive deflections may become a
problem. The efficient use of HSC in structures could be prevented by the large deflections
predicted by the current codes, for sustained loads.

3.2.1 Flexural rigidity, EcIeff

AS3600-2001 Cl. 8.5.3.1 recommends the Bransons formula given in Eq. 3.1 for calculating
the effective moment of inertia. Bransons formula [2] has been used to calculate the effective
moment of inertia since 1968.

M cr 3
I eff = I cr + ( I g - I cr )( ) Ig (3.1)
Ms
where, Ig = second moment of area of gross section

Icr = second moment of area of the cracked section

Ms = maximum bending moment at the centre of the section based on the

short term serviceability load

3-1
Mcr = cracking moment at midspan of the beam

Mcr = Z. fcf (3.2)

where, Z = section modulus of the uncracked section, referred to the extreme

fibre at which cracking occurs

fcf = characteristic flexural tensile strength of the concrete

f 'cf = 0.6 f 'c (3.3)

According to AS3600, the characteristic flexural tensile strength of the concrete is given by Eq.
(3.3). ACI state-of-the-art report [3] recommends a higher value as given in Eq. (3.4).

f 'cf = 0.94 f 'c (3.4)

For high-strength concrete, the AS3600 equation is recommended in this book. However, it
should be noted that the value of characteristic flexural tensile strength does not have a major
influence on long-term deflections.

Alternatively, a simplified equation to calculate the effective moment of inertia of reinforced


concrete rectangular sections is given in AS3600 (Eq. (3.5)).

I eff = 0.045bd 3 (3.5)

The above equation does not take into consideration the improved stiffness characteristics of
HSC members. Consequently, using deemed-to-comply provisions of AS3600 for slabs and

beams with the simplification I eff = 0.045bd 3 may give rise to a conservative and an

uneconomical design.

The ACI state-of-the-art report for high-strength concrete [3] recommends the use of the
Branson's formula for calculating the effective moment of inertia. However, Lambotte and

M cr
Taerwe [4] suggested that the exponent of may be higher than three for high-strength
Ms
concrete beams.

3-2
3.2.2 Recommendations for EcIeff

For determining the effective flexural stiffness, Ec I eff , it is recommended that AS3600s
Cl. 8.5.3.1 [Eqs. (3.1) and (3.5)] be adopted for calculating the effective second moment of
inertia of the member and the formula given in Section 2.4 be used to calculate the

characteristic modulus of elasticity, Ec .

3.2.3 Long term deflections (Cl. 8.5.3.2 of AS3600)

Almost all concrete structures are subjected to sustained stresses. The creep strains associated
with these stresses and the shrinkage strains have a significant effect on the long term
structural behavior.

According to AS3600, additional long-term deflection of reinforced concrete beams resulting


from creep and shrinkage is found by multiplying the immediate deflection caused by the

sustained load by a factor, kcs . The ratio of compression steel to tension steel is used in this
factor to account for the influence of compression steel. According to the ACI-1999 [5]

Asc
building code, the compression steel ratio and the time factor are included in the
bd
multiplying factor. AS3600 and ACI approaches are based largely on tests on low strength
concrete beams (20 MPa to 40 MPa) in the 1950s and early 1960s [6]. Code methods for
predicting long term deflections under sustained loads do not recognise concrete compression
strength as a significant variable.

It is now well recognised that HSC has a creep coefficient which is very much lower than that
of normal strength concrete (See Section 2.5.2). It follows that the long term deflections of
HSC beams should be significantly less than that of normal strength concrete elements. It
means that the predicted long-term deflection using the code method will be higher than the

actual values. The long-term factor ( kcs in AS3600) for high-strength concrete may be 40%-
65% less than that for normal strength concrete [4,7]. In high-strength concrete beams,
shrinkage deflection may also be less. However, the effect of compression steel in reducing the

3-3
long-term deflection in beams is less for HSC [8], which is not accounted for in the code
formulae.

Long-term deflections can be predicted with a reasonable amount of accuracy incorporating the
creep factors and the shrinkage strain values given in Chapter 2, from the age adjusted effective
modulus method (AEMM). This method is described in detail by Gilbert [9].

Recommendation for long-term deflections


By comparison of experimental results with the code formulae and the predicted values from
the age adjusted effective modulus method (AEMM), a revised formula for the long-term

factor Kcs was suggested [7]. The lower creep factors in HSC and the lower effect of

compression steel in reducing the long-term deflection are considered in this formula given
below.

A
K cs = m 2 - 1.2 sc m 0.8
Ast (3.6)

50
m= m1
fc'

3.3 A RECTANGULAR STRESS-BLOCK MODEL FOR HIGH-STRENGTH


CONCRETE

The equivalent rectangular stress block (RSB) has been the key concept behind the
development of the codified, simple ultimate strength theory. Ultimate strength design of
members subjected to axial load and bending is usually carried out using the rectangular stress
block. This stress block has been introduced into almost every national and international
concrete structures code since the concept was introduced by Hognestad et. al. [11]. The RSB
is defined by two parameters a (the parameter to determine the intensity of stress in the

stress block) and b (the ratio of the depth of the stress block to the depth of the neutral

axis). Fig. 3.1 illustrates the parameters a and b .

3-4
The rectangular stress block has been derived such that the magnitude and the location of the
compressive force C are the same when compared with the actual stress block for concrete
(refer Fig. 3.2).

Fig 3.1 Rectangular stress block

Fig. 3.2 Compressive stresses in concrete

Presently, most codes for concrete structures, including AS3600 [1], fix the value of a equal

to 0.85 and vary the value of b between 0.85 and 0.65 for concrete strengths between 28 MPa

and 50 MPa. These values have been determined from various investigations conducted by

researchers including Mattock et al. [12] and Winter [13]. The product ab is an indication of

3-5
the total compressive force on the section and b/2 is an indicator of the location of the

compressive force.

As explained in Section 2.7, the unconfined stress-strain curves for normal strength concrete
and high-strength concrete vary significantly. Referring to Fig. 2.4, higher strength concretes
have steeper ascending and descending portions and the curves tend to be more linear than
curves for lower strength concretes. Due to these and other differences in characteristics of
the unconfined stress-strain curves for normal and high strength concretes, the equivalent
rectangular stress block parameters need to be appropriately modified for HSC, especially for
members subject to an axial load component.

The need to revise the rectangular stress block to suit high-strength concrete has also been
recognised by the ACI committee 363 on High Strength Concrete [3] and the ACI committee
441 on Reinforced Concrete Columns [15].

The rectangular stress block suggested by Pendyala and Mendis [14] is outlined below. Other
suggested rectangular stress blocks are described in detail in Ref. [14].

Recommendation:
[The proposed rectangular stress block for HSC]
The proposed stress block parameters have been derived for concretes with

60 MPa f 'c 100 MPa . The derivation has a lower limiting value at 60 MPa, based

on the current AS3600 rectangular stress block and an upper limiting value at 100 MPa, based
on a triangular stress distribution. The proposed rectangular stress block has been validated by
using column and beam data from literature [14].

At the lower limit of 60 MPa, based on the current AS3600 rectangular stress block,

a = 0.85 and b = 0.65


and at an upper limit of 100 MPa, based on the triangular stress block,

a = 0.75 and b = 0.67


3-6
However, since both values of rectangular stress block parameters, a and b, should decrease

with increasing fc, b has been reduced to 0.60 for a concrete strength of 100 MPa.

Intermediate values of a and b are obtained by linear interpolation.

Thus, for 60 MPa f 'c 100 MPa


a = 0.85 - 0.0025(fc - 60) (3.7)

b = 0.65 - 0.00125(fc - 60) (3.8)

The proposed model has been formulated such that there is a continuity with the existing
AS3600 provisions for the rectangular stress block.

Ultimate Strength of a Singly Reinforced Section


The stress and strain diagrams for a rectangular cross-section at ultimate strength are shown in
Fig.3.3. It is assumed that the steel is yielding and the concrete is crushing. The ultimate

strain in the concrete at maximum moment capacity ecu, is taken as 0.003 for normal strength

concrete sections. The same value is recommended for high-strength concrete sections.

As C = T, a fc g ku db = Ast fy (3.9)

Ast f y rf y
Thus
ku = = (3.10)
afc'gdb afc'g

The depth to the neutral axis thus depends on the amount of reinforcement and the ratio of
steel and concrete strengths. Since ductility is greater for a lower percentage of reinforcement
(provided it is not so low that fracture of the steel occurs before the crushing strength of

concrete is reached), the ratio k u is low for high ductility. The Code specifies that a cross-

section may be assumed to be ductile where k u at ultimate strength is less than or equal to

0.40. If this value is exceeded the section is defined to be over-reinforced, or non-ductile.

3-7
0 .8 5 f'c
e =e
c cu
g C
k ud k ud
d ju d

A st
e e fy fy T
s sy
cr os s st rain st ress eq u iv alen t fo rce
sect ion st ress
Fig.3.3 Analysis of a Singly Reinforced Rectangular Section at Ultimate Moment

The ultimate moment capacity of the section, Mu, is

Mu = Cjud = Tjud = Ast fy jud (3.11)

now

Ast
M u = Ast f y d [1 - ]
2 bbdf 'c (3.12)
so that

Ast
M u = Ast f y d [1 - ]
2 bbdf 'c (3.13)

This is the basic formula for the ultimate moment capacity of a singly-reinforced section.

A capacity reduction factor,f, is applied to this formula to derive the effective ultimate

moment capacity, fMu with a value of 0.8 being assigned to f for bending.

The above formula also applies to T sections where the compression block lies within the

flange. If b is the width and Df is the thickness of a T section flange, the compression stress

Astfy
Df
block can be considered to be within the flange if afcbDf Ast fy, or af ' cb . If

it is outside the flange the concrete force has to be modified to include the web section.

3-8
The spread-sheet BEAM_HSC given in this book calculates the moment capacity

incorporating the suggested a and b factors.

Note on Ductility and Moment Redistribution:

Despite high-strength concrete being a more brittle material, high-strength concrete flexural
members exhibit greater ductility due to lower neutral axis depths [Eq.(3.10)]. Therefore same
moment redistribution factors [Cl. 7.6.8 of AS3600] can be maintained for high-strength
concrete flexural members. However as the new 500 MPa (N type) steel reinforcement is less
ductile compared to 400 MPa (Y type) steel, it is recommended to limit the moment
redistribution to 20%.

3.4 MINIMUM REINFORCEMENT

Minimum reinforcement is specified in various national and international codes to ensure that
at the inception of cracking the stresses are safely transferred to the steel and also to insure
that the crack widths are not excessive.

The minimum reinforcement is a function of the tensile strength of the concrete and since
tensile strength properties of concrete are different for normal and high strength concrete, the
minimum reinforcement requirement needs revisiting.

AS3600 prescribes the minimum reinforcement on the basis that the Mu be at least 20%

greater than Mcr. To this end AS3600-1994 suggests rmin = 1.4/fsy for beams. The revised
suggestion in the AS3600-2001 is given below.
2
D f D
rmin = 0.22 t ; If ft = 0.6 f 'c and = 1.15 is assumed
d fsy d

0.175 f 'c
then rmin = (3.14)
fsy

3-9
The ultimate strength in bending, Muo, at the critical sections shall not be less than 1.2 times

the cracking moment, Mcr, given by

Mcr = Z(fcf + P/Ag) + P.e (3.15)

where

Z= the section modulus of the uncracked section, referred to the extreme fibre at which the

cracking occurs

fcf = characteristic flexural tensile strength of the concrete

For the purpose of this clause, the critical section to be considered for negative moment shall

fM uo
be the weakest section in the negative moment region (i.e. where
M * is the least).

3.4.1 Recommendation

The formula given in AS3600-2001 (Eq. (3.14)) is not applicable to HSC due to the inherent
differences in material characteristics compared to NSC. The following formula is
recommended for HSC. For rectangular reinforced concrete cross-sections, this requirement
shall be deemed to be satisfied if minimum tensile reinforcement is provided such that

2
Astmin D f
= 0.25 t
bd d fsy (3.16)

For reinforced T-beams or L-beams where the web is in tension, b shall be taken as the width
of the web, bw. The above formula was derived by calibrating Eq. (3.14) to closely match

Nilsons [16] recommendations.

Note:

The minimum reinforcement depends on the cracking moment Mcr of the beam. Mcr in turn

depends on Ast, b, d, fc and fsy. The above equation is based on the work by Nilson [16]

after equating the cracking moment of a beam section to the resisting moment after cracking
and limiting the steel stress at cracking to 2/3 the yield stress [16].

3-10
The New Zealand code (1995) [17] determines the minimum area of tensile reinforcement to

Astmin 0.25 f 'c


be equal to bd
= for rectangular beams for all fc.
fsy

The Canadian standard [18] determines the minimum tensile steel reinforcement to be equal to

Astmin 0.20 f 'c


= for rectangular beams. The Canadian code has limited fc to a
bd fsy
maximum of 80 MPa. It can be seen that the New Zealand and Canadian provisions are
derivatives of Nilsons [16] work.

Based on an analytical procedure followed by numerical methods, Murugama et al.[19]


produced a chart to determine the minimum tensile steel requirement for different tensile stress

criteria just after cracking and for concretes with varying fc. Their recommendations are very
conservative compared to other suggestions.

3.5 SHEAR STRENGTH OF HIGH STRENGTH CONCRETE

Pendyala [20] has shown that especially when failure is governed by brittle or non-ductile
behaviour, such as the case with shear failure, extrapolating the existing normal strength design
rules to high-strength concrete will result in a design with reduced conservativeness or
possibly even an unsafe design.

3.5.1 Implications of using HSC to resist shear

There is a concern amongst researchers and designers that high-strength concrete may not be
strong in shear, as some of the mechanisms that function to resist shear, notably the aggregate
interlock mechanism, may be absent.

The shear resistance of reinforced concrete beams is dependent on the tensile strength of the
members. In high strength concrete, where the difference in strength of the concrete matrix and
the aggregates is comparatively less than in normal strength concrete, the fracture planes are

3-11
relatively smooth [20,22,23]. The smooth fracture planes reduce the shear resistance of
concrete by reducing the aggregate interlock between the fracture surfaces. Thus, there is
apprehension amongst designers and researchers who believe that there may be a significant
drop in the shear strength of HSC beams as compared to NSC beams.

A comprehensive literature survey by Pendyala [20] revealed that research and guidelines on
shear design for HSC beams were sorely lacking. There was enough evidence provided by past
researchers [21,22-25] to show that due to the absence of aggregate interlock, the present code
guidelines for shear including those contained in AS3600 may be unconservative or marginally
conservative. However, the lack of design guidelines has not prevented engineers from
designing and constructing HSC concrete structures including those members that carry shear
forces. As with all instances in the past, for lack of anything better, guidelines for design with
NSC have been simply extrapolated to design with HSC.

Shear design procedures contained in the present Australian concrete code, AS3600 and other
overseas codes are based on results of beam tests with concrete strength less than 41 MPa and
therefore their applicability to HSC beams needs to be thoroughly investigated before being
recommended for practice.

Shear strength of a reinforced beam excluding shear reinforcement


The shear strength of high-strength concrete does not proportionately increase with a higher

fc. The lack of aggregate interlock which resists a substantial portion of the post-crack shear

forces is the factor which reduces the shear strength of high-strength concrete. Unlike normal
strength concrete, high-strength concrete fracture surfaces are smooth with fracture passing

through aggregates. Pendyala[20] suggested that an additional factor b4 be used to compute

the shear strength of high-strength concrete beams,

where, b4 = 1 for fc < 50 MPa

= 1.25 - 0.005fc 0.9 for 50 MPa f 'c 100 MPa (3.17)

3-12
Therefore, referring to AS3600 Cl. 8.2.7.1, this clause would more appropriately read:
1
A f ' 3
Vuc = b1b 2 b 3b 4bv do st c (3.18)
bv do

Recommendation for Minimum Shear reinforcement

It is recommended that the minimum shear reinforcement Asv.min, provided in a beam shall be
given by:

bv s
Asv. min = 0.06 fc '
fsy. y (3.19)

Notes:

bv s
Asv. min = 0.35
The current AS3600 formula, fsy. y is independent of fc. As the

compressive and tensile strength increase, the cracking shear also increases. This increase in
cracking shear requires an increase in minimum shear reinforcement so that a brittle shear

failure does not occur at the onset of cracking. Thus Asv.min is now a function of fc as well to
take into account the increase in cracking shear.

The Shear strength of a beam with minimum reinforcement


To be consistent with the modified formula for minimum shear reinforcement [Eq. (3.19)], the

ultimate shear strength of a beam provided with minimum shear reinforcement, Asv.min, shall
be taken as,

Vu . min = Vuc + 0.1 fc' bv do (3.20)

These formulae are incorporated in Spreadsheet BEAM_HSC provided with the book.

3.6 DEVELOPMENT LENGTH FOR BARS IN TENSION FOR fc<100MPa

For all deformed bars in tension

The development length, Lsy.t to develop the yield strength fsy, of a bar in tension shall be

calculated as follows:

3-13
k1k2 fsy Ab
Lsy.t = 25 k1d b
(2 a + d b ) f 'c (3.21)

Notes:
Unlike normal strength concrete, where the bearing stresses are uniform over the entire length
of the splice, the bond stresses in HSC are non uniform and follow a triangular pattern (Fig
3.4), with peak stresses at the splice and gradually decreasing towards the end [26]. The high
intensity bearing stresses have the tendency to split the concrete at the tension splice leading

to a brittle fracture. Inclusion of a minimum amount of ligatures (for fc greater than 70 MPa)

over the length of the splice helps to distribute the bearing stresses uniformly and reduces the
splitting tendency due to high bearing stresses. As a guide, bars with minimum diameter of
10mm should be used to prevent fracture of the ligatures. Therefore it is suggested that

minimum amount of ligatures according to Eq. (3.19) should be provided for members with fc
> 70MPa. As an alternate to ligatures over the splice length, the cover concrete may be

increased, subject to other consideration of cover requirements.

Current design provisions of AS3600 [Eq. (3.21)] are based on empirical relationships
developed from tests on low-strength concrete. The results of a series of tests on high-strength
concrete, up to 100 M Pa, completed at the University of M innesota and 101 test results from
five other research studies were used by M endis and French [27] to review the existing
recommendations in AS3600 for design of splices and anchorage of reinforcement. It was
shown that AS3600 formula can be retained for concrete strengths up to 100 M Pa.

(a)

(b)

Fig. 3.4 Bond Stresses in (a) NSC (b) HSC

3-14
3.7 DESIGN EXAMPLE

Design flexural and shear reinforcement for the two span rectangular reinforced high-strength
concrete beam shown in Fig. 3.5.

f'c = 80 MPa fsy = 400 MPa


Dead Load, G = 65 kN/m (including beam self weight)
Live Load, Q = 35 kN/m

d/b 2

Length of clear span, Ln = 3.5 m


Centre-to-centre L = 4.0 m

Width of support is 0.5 m.


Note:
This design example is the same example presented by Darvall [28], except that high-strength

concrete (f'c = 80 M Pa) is used. The design is carried out according to the rules given in this
Chapter. Spreadsheet BEAM _HSC provided with the book follows these rules.

3.7.1 Design loads

3.7.1.1 Load Combination A


Live Load on one span only.
Design Loads
First Span: Dead Load Only

w = 1.0 G = 65 kN/m

Second Span: Dead Load and Live Load

w = 1.25G + 1.5Q = 133.8 kN/m (AS 1170.1 Cl. 2.2.2)

3-15
3.7.1.2 Load Combination B
Live Load on both spans.
Load Combination A
133.8 kN/m
65 kN/m

Load Combination B
133.8 kN/m

Fig.3.5 Design loads

3.7.2 Determination of load effects

Load cases:
(1) Live Load on both spans
(2) Live Load on this span
(3) Live Load on the other span

3-16
S.F.D

Critical Section 246 kN


for Shear at
distance d
267.6 kNm
(1)
(2)
(3)
B.M.D
150

177 kNm

Fig. 3.6 SDF and BMD envelopes

Peak Values Before Moment Redistribution


Negative: 267.6 kNm
Positive: 177.0 kNm

Design Values After 20% Moment Redistribution


Negative: 267.6 0.8 = (-) 214.1 kNm (AS 3600 Cl. 7.6.8)

Positive: 150 + 267.6 0.1 = (+)176.7 kNm

3.7.3 Design for flexure

Design Values:

M*(-) = - 214.1 kNm

M*(+) = 177 kNm

3-17
3.7.3.1 Reinforcement Calculations
Concrete Stress block parameters:

fc= 80 MPa
a = 0.85 - 0.0025(fc - 60) (Eq. 3.7)

b = 0.65 - 0.00125(fc - 60) (Eq. 3.8)

Ultimate moment capacity:

rfsy 2
M u = fsy r1 - bd
2af 'c
Since assumed moment redistribution is 20%, ensure ku 0.2 (AS3600 Cl.7.6.8)

fsy r
ku =
abf 'c
as ku 0.2,

0.8 0.625 80
r 0.2
500
r 0.016 (1.60 %)
Minimum Reinforcement
2
D f 80
rmin = 0.25 t = 0.25 x1.15 2 x 0.6 x = 0.0036 (0.36%) (Eq. 3.16)
d fsy 500
D
(assume = 1.15 )
d

3.7.3.2 Limit State Design for Strength

M u* fM u (f = 0.8)
214 kNm 0.8 M u
or M u 267.5 kNm
Beam Sizing

rfsy 2
M u fsy r1 - bd
2af 'c
3-18
0.016 500 2
267.5 10 6 500 0.016 1 - bd
2 0.8 80
267.5 10 6 7.5bd 2
bd 2 35.7 10 6
d
2
b
selectd = 405 mm
b = 225 mm
Area of Steel Reinforcement

Provide 2N24 + 2N20 bars, Ast =1533 mm2


Check reinforcement ratio:

1533
r= = 0.0168
225 405
Note: This ratio is slightly higher than the allowable value of 0.016 derived earlier to satisfy

the maximum limit of 0.2 for k u. However provision of compression reinforcement (top

reinforcement for mid span and bottom reinforcement for supports) will reduce the ku value.

Moment Capacity, for both positive and negative bending:

rfsy 2
M u = fsy r1 - bd
2af 'c
0.0168 500
M u = 500 0.01681 - 225 405
2

2 0.8 80
M u = 289 kNm 268 kNm \ o.k.

Beam Dimensions

d = 405 mm
b = 225 mm
Ast = 2N24 + 2N20 (placed in two rows)
Assume N12 ligatures
Clear Cover = 25 mm

3-19
Dmin = 405 + 24 + 10 + 25 = 464mm
bmin = 2 25 + 2 12 + 2 24 + 3 24 = 194mm < 225mm \ o.k.
Cross-section at critical section for negative bending moment:

3.7.4 Design for shear

Critical section for shear = 405 mm from supports (AS 3600 Cl. 8.2.4)

3.7.4.1 Design Shear

Vext* . sup port = 123 kN


Vint* . sup port = 246 kN

3.7.4.2 Design Using Modifications to AS3600 Provisions


Concrete Resistance
1
A f ' 3
Vuc = b1b 2 b 3b 4bv do st c (Eq. 3.18)
bv do
d
b1 = 1.11.6 -
1000
405
= 1.11.6 - = 1.31
1000

b2 = 1
b3 = 1
b 4 = 1.25 - 0.005 f 'c
= 1.25 - 0.005 80 = 0.85
However b 4 0 .90

Hence use
b 4 = 0.90
13
1533 80
Vuc = 1.31 0.9 225 405
225 405
= 119 kN
3-20
f Vuc = 0.7 119 = 83kN

Shear strength limited by web crushing

Vu.max = 0.2fcbd (AS 3600 Cl. 8.2.4)

= 0.2 x 80 x 225 x 405 N


= 1458 kN

fVu.max = 0.7 x 1458 = 1021 kN

Shear Strength with minimum reinforcement


'
Vu.min = Vuc + 0.1 f c bd (AS 3600 Cl. 8.2.9)

= 118 x 103 + 0.1 80 x 225 x 405 N


= 200 kN

fVu.min = 0.7 x 200 = 140 kN

Stirrup Spacing

Resistance provided by shear reinforcement

Asv fsy. f d
V * > fVu . min \ Vus = cot q v (AS3600 Cl. 8.2.10)
s
Asv = 226 mm (2 N 12 Stirrups )
246 - 140
q v = 30 + 15 = 31.80 o
1021 - 140
226 400 405 cot 31.80 o
Vus = N
s
74 10 6
= N
s
Resistance Required by shear reinforcement

V * fVu
246 kN 0.7Vu
Vu 351kN

3-21
Vus = Vu - Vuc (AS 3600 Cl. 8.2.2)
= 351 - 119 kN
= 233 kN
Stirrup Spacing for critical (internal support) condition

74 10 6
Vus = N
s
74 10 6
s= mm
233 10 3
s = 317 mm
Check Detailing
0.5D = 0.5 x 470 = 235mm < 317 mm (AS3600 Cl. 8.2.12.2)
Hence use N12@235mm.

Stirrup Spacing for other conditions (Mid span and End Support)
As Cl. 8.2.12.2 limits the spacing to 235mm, use N12@235mm as calculated for the internal
support condition.

3.7.5 Deflection calculations

Dtotal = D elastic + D shrinkage + D creep

or D total = D short -term + D long - term

3.7.5.1 Section Properties


Elastic Modulus

Ec = 0.43hr1.5 f 'c (Eq. 2.4)

h = 1.1 - 0.002 f 'c


= 1.1 - 0.002 80 = 0.94
Ec = 0.043 0.94 25001.5 80
= 45, 190 MPa

3-22
3.7.5.2 Long-term factor

A
kcs = m 2 - 1.2 sc m 0.8
Ast
(Eq. 3.6)

50 50
m= = = 0.625 1
f 'c 80
905
kcs = 0.625 2 - 1.2 0. 625 = 0.973
1533
3.7.5.3 Approximate Method
Check long-term deflection using C1 8.5.4, AS3600 (Deemed to comply method)

Fd.ef = (1+ kcs)g + (ys + kcsyl)q

= [2.973 65 + (0.7 +0.973 0.4)] 35

= 166.4 kN/m

k1 = 0.045, k2 = 1/185, D / Lef = 1/250

1
1 3

Lef 0 .045 225 45190


= 250
d 1
166.4
185

Lef
Allowable d = 12.3

Lef 4000
Actual
d = 405
= 9.9 < 12.3
Hence O.K.

3-23
3.7.5.4 Accurate Method

Gross section second moment of area, Ig at mid span:

y (n-1)Ast
470 405

Ast=1533 mm2

Es 200, 000
n= = = 4.43
Ec 45, 190
(n - 1) A
st = 3.43 1533 = 5252 mm 2
470
470 225 + 5252 405
y= 2 = 243 mm
470 225 + 5252
225 470 3
+ 225 470 (244 - 235) + 5252 ( 405 - 244)
2 2
Ig =
12
= 20.91 10 8 mm 4

Cracked Section second moment of area, Icr at mid span

y nAst

nAst = 4.43 1533 = 6785 mm 2


y2
225 + ( 405 - y ) 6785
y= 2
225 y + 6785
Solving this equation
y = 107mm

3-24
2
225 107 3 107
+ (225 113) + 6785( 405 - 107)
2
I cr =
12 2
= 6.94 10 8 mm 4

Cracking Moment at mid span

M cr = Z . f 'cf
20.91 10 8
Z= = 8605 10 3 mm 3
244
f 'cf = 0.6 f 'c = 5.4 MPa
M cr = 8605 10 3 5.4 = 46.2 kNm

Effective second moment of area, Ieff at mid span


3
M
I eff ( )
= I cr + I g - I cr cr
Ms
(AS 3600 Cl. 8.5.3.1)

46.2
3

= 6.94 + (20.91 - 6.94) 10 mm
6 4

103
= 8.2 10 8 mm 4
Value of Ms is obtained from the bending moment diagram (see Fig. 3.7).

As the amount of reinforcement is similar at the support, it is assumed that the Ieff at the

support is approximately equal to this value. Therefore the average value is equal to Ieff at

mid span.

3.7.5.4.1 Instantaneous Elastic Deflection D e

De is maximum when only one span is loaded

Short term load = G + y sQ (AS1170.1 Cl 2.4)

= 65 + 0.7 35 (AS1170.1-Table 2.2)

= 89.5 kN/m

3-25
89.5 kN/m
65 kN/m

103 kNm

BMD

3.97mm

Fig. 3.7 BMD and Deflection Shape

3.7.5.3 Long Term Deflections using (modified) AS3600 approach

Long term load = G + y lQ (AS1170.1 Cl 2.4)

= 65 + 0.4 35 (AS1170.1-Table 2.2)

= 79 kN/m
79 kN/m
65 kN/m

3.31mm

Fig. 3.8 Deflection Shape(Long-term loads)


Additional long-term deflection

D = kcs De.sus = 0.973 3.31 = 3.3 mm

3-26
Hence, total deflection Dtotal = 3.97 + 3.22 = 7.19 mm

D 7 .2 1
\ = < hence o.k. (AS 3600 Table 2.4.2)
L 4000 250

3.7.6 Detailing

C B A

4000
2N24+2N20 2-N24 2-N24

405
470

2-N24 2N24+2N20 2N24+2N20


225 225 225

SECTIO N AA SECTION BB SECTION CC

Fig. 3.8 Reinforcement details

3.7.7 Comparative advantages


Table 3.1 Comparison of HSC and NSC beams

DESCRIPTION HSC NSC COM M ENTS

'c 80 M Pa 25 M Pa

Ec 45191 M Pa 25280 M Pa Increase of 80%

b 225 mm 250 mm

d 405 mm 450 mm

bd 91,125 mm2 112,500 mm2 A saving of 23% in concrete area

3-27
Ast 1533 mm2 1533 mm2 no real saving in steel area

Shear Stirrups N12@235 N12@220

ku 0.172 0.266 A lower value of k u, for HSC for


(calculated with same steel amount, Hence greater
compression rft.) ductility.
Dtot 7.2 mm 7.8 mm HSC beams possess greater
stiffness overall and thus less
deflections.

Table 3.2 Comparison of HSC beams using 400MPa and 500MPa steel
(reducing steel rft. area to obtain similar moment capacities for 500MPa steel)

DESCRIPTION fsy = 400 M Pa fsy = 500 M Pa COM M ENTS

b 225 mm 225 mm
405 mm 405 mm
d
1810 mm2 1533 mm2 A saving of 16% in steel area
A st
275 MPa 289 MPa
Mu
ku 0.161 0.172 Less ductile
(calculated with
compression rft.)
Shear Stirrups N12@235 N12@235 no increase in shear resistance
Dtot 6.6 mm 7.2 mm less steel reinforcement lead to
greater deflection

Conclusions

Based on a simple design example, the following can be concluded:


1. There can be savings in the amount of concrete required to carry a given load if high-
strength concrete is used.
2. High-strength concrete flexural members are more ductile as evidenced by a smaller value
of ku, the neutral axis factor.

3. Increased overall stiffness.


The use of HSC can lead to increased overall stiffness due to:
3-28
(a) Improved tensile resistance
(b) Improved creep and shrinkage properties.
(c) Higher Elastic Modulus.
4. The use of 500 MPa steel reinforcement increases the flexural capacity, however it will
reduce the ductility of the member.

REFERENCES

1. AS 3600-2001, Concrete Structures", Standards Australia, Sydney, 2001.

2. Branson, D., Design Procedures for Computing Deflections, ACI Journal, Vol. 65, No.
9, Sept. 1968, pp. 730-742.

3. ACI Committee 363, State of the Art report on High-Strength Concrete(ACI 363R-92),
American Concrete Institute, 1992.

4. Lambottee, H. and Taerwe L.R., "Deflection and Cracking of High-Strength Concrete


Beams and Slabs", Proc. of High Strength Concrete - Second International Symposium,
ACI, Berkeley, California USA, SP-121, 1990.

5. ACI 318 "Building Code Requirements for Reinforced Concrete", American Concrete
Institute, Detroit, 1999.

6. Yu, W. and Winter, G., Instantaneous and Long-time Deflections of Reinforced Concrete
Beams under Working Loads, ACI Journal, Vol. 57, No. 1, July 1960, pp. 29-50.

7. Nicholls, S. and Mendis, P.A., Deflections of HSC Beams, Research Project Report,
1997.

8. Paulson, K., Nilson, A.H., Hover, K.C., Long-term Deflection of High-strength Concrete
Beams, ACI Materials Journal, Vol. 88, No. 2, Mar-Apr 1991 pp. 197-206.

9. Gilbert, I., Time effects in Concrete Structures, Elsevier Science Publishers, Amsterdam.

10. Mendis, P.A, Pendyala, R.S. and Setunge, S. Requirements for High-Strength Concrete in
AS3600, High-Performance Concrete Sub-committee of the Concrete Institute of
Australia (Victoria), Melbourne, 1997.

11. Hognestad, E., Hanson, N.W., and MacHenry, D., Concrete Stress Distribution in
Ultimate Strength design, ACI Journal, Proceedings, Vol. 52, Dec., 1955, pp. 455-479.

3-29
12. Mattock, A.H., Kriz, L.B., and Hognestad, E., Rectangular Concrete Stress Distribution
in Ultimate Strength Design, ACI Journal, Vol. 56, Feb 1961, pp. 875-928.

13. Winter, G., Safety and Serviceability Provisions in the ACI Building Code, ACI-CEB-FIP-
DCI Symposium, ACI Special Publication SP-59, 1979.

14. Pendyala, R. and Mendis P.A., "Rectangular Stress Block Parameters for High-strength
Concrete Members", Civil Engineering Transactions, Institution of Engineers, Australia,
Vol. CE39, No. 4, 1998, pp. 135-144

15. ACI committee 441 Reinforced Concrete Columns, State of the Art Report: High
Strength Concrete Columns (ACI 441), American Concrete Institute, 1997.

16. Nilson A.H., "Design Implications of Current Research on High Strength Concrete", SP87,
ACI Special Publication, American Concrete Institute, 1986, pp. 85-118.

17. New Zealand Concrete Structures Standard(NZS 3101), New Zealand Standards, 1995.

18. CAN3-A23.3, Design of Concrete Structures for Building, Canadian Standard


Association, Rexdale, Ontario, Canada, 1994

19. Murugama, H., Minimum amount of Tensile Reinforcement in Reinforced Concrete


Flexural Members with High-Strength Concrete, Proc. Concrete 95 Toward Better
Concrete Structures, Concrete Institute of Australia, Brisbane, pp. 329 -335,1995

20. Pendyala R.S.The Behaviour of High-Strength Concrete Beams Ph.D thesis, submitted
for examination, The University of Melbourne, August 1997

21. Mphonde, A.G. and Frantz, G.C. Shear Strength of High Strength Reinforced Concrete
Beams Research Report No. CE84-157, Department of Civil Engineering, The University
of Connecticut, Storrs, June 1984.

22. Konig, G., Grimm, R. and Remel, G. Shear Behaviour of Longitudinally Reinforced
Concrete Members of High-Strength Concrete, Darmstadt Concrete, Vol. 8, 1993, pp.
27-42

23. Elzanaty, A.H., Nilson, A.H. and Slate, F.O.Shear Capacity of Reinforced Concrete
Beams using High-Strength Concrete ACI Journal, Vol. 83, March- April 1986, pp. 290-
296.

3-30
24. Kong, P.Y.L and Rangan, B.V. Studies on Shear Strength of High Performance Concrete
Beams Research Report No. 2/97 march 1997, School of Civil Engineering, Curtin
University of Technology, Perth.

25. Johnson, M.K. and Ramirez, J.A.Minimum Shear Reinforcement in Beams with Higher
Strength Concrete ACI Structural Journal Vol. 86, No. 4, July-Aug 1989, pp. 378-382.

26. Azizinamini, A, et al., "Bond Performance of Reinforcing Bars Embedded in High-Strength


Concrete," ACI Structural Journal, v90 No.5, Sept.-Oct. 1993, pp. 554-561

27. Mendis, P.A and French, C.A, Development Lengths Of Reinforcement In High-Strength
Concrete, Accepted for publication, Civil Engineering Transaction, IEAust. 1997.

28. Darvall, LeP. P. Reinforced and Prestressed Concrete, Macmillan Co. of Australia, 1989.

3-31
CHAPTER FOUR

DESIGN OF HIGH STRENGTH CONCRETE


COLUMNS

4.1 INTRODUCTION

As previously noted, the use of high-strength concrete columns for both medium rise and high-
rise buildings has proven most popular with economy, superior strength and stiffness being
the major reasons for its prevalence. High strength concrete has been used for columns in
many parts of the world. Australia has taken advantage of the benefits of high strength
concrete through its widespread use on a number of buildings including 120 Collins Street,
Melbourne Central, The Telecom Building, Bourke Place, The Rialto Project and proposed
Eureka Tower in Melbourne, the R & I Tower in Perth and the Four Seasons Hotel, Century
Tower, Elarosa Development and proposed World Tower in Sydney. High-strength concrete
is an extremely cost effective material for the construction of columns. In addition to reducing
column sizes and producing a more durable material, high-strength concrete reduces the cost of
formwork.

The extrapolation of code formulae valid for normal strength concrete to concrete compressive
strengths greater than 50 MPa is a common design procedure, however it does raise some
concerns. One particular concern is the application of this approach to the design of high
strength concrete columns loaded predominantly axially, where it has been shown that there is
an apparent lack of ductile behaviour when stirrups are provided in accordance with the
Australian Concrete Code, AS3600 [1].

There are some state of the art reports on HSC columns (e.g. ACI 441 [2]) and codes of
practice which cover HSC (e.g. Norwegian Concrete Code, NS 3473 [3]). However, these
codes and reports do not provide simple and adequate rules to design HSC columns and
calculate the stirrup spacings to satisfy ductility requirements.

4-1
This chapter provides a method for the design of high strength concrete columns in terms of
the required spacing of lateral reinforcement in order to provide a similar level of ductility as
an equivalent column constructed with normal strength concrete would possess. In order to
achieve this, important concepts need to be addressed such as the choice of a measure of
ductile behaviour and a realistic high-strength concrete stress-strain model, as well as limiting
factors such as longitudinal steel buckling and lateral steel fracture. The main form of ductility
analysis adopted in this book is in terms of moment-curvature curves. The simple rules
suggested in this Chapter for the design of stirrup spacing for HSC columns can be used with

new D500N (fy =500 MPa) steel. Low ductility D500L bars are not recommended for high-
strength concrete columns.

4.2 STRENGTH OF HIGH-STRENGTH CONCRETE COLUMNS

The concrete strength determined by a standard cylinder test is known to be different than the
strength of unconfined concrete in a member. This discrepancy is explained by the differences
in size and shape and the concrete casting practice between a cylinder as compared to the
associated member. In normal strength concrete, the ratio of unconfined strength in the
member to standard cylinder strength is taken as 0.85, although a variation between 0.85 and
1.0 has been reported. Saatcioglu and Razvi [4] found that similar tests conducted on high-
strength concrete show approximately the same variation. However, experimental
observations on high strength concrete columns indicate that the cover concrete spalls off prior
to the attainment of column capacity [4]. Therefore, the maximum load is resisted by the core
concrete, including the contribution of the longitudinal reinforcement. The Canadian code
suggests a reduction of capacity to accommodate cover spalling. However columns with well
confined cores are able to resist loads higher than capacities based on gross cross-sectional area
and unconfined concrete strength. Those that do not have sufficient confinement are not able
to develop their unconfined capacities based on the total cross-sectional area.

Recommendation
If additional ligatures are provided as suggested in this chapter (see Section 4.6.3), the ultimate

strength in compression, Nuo, for a high-strength concrete column can be calculated using a

uniform concrete compressive stress of 0.85fc, as shown below:

4-2
Nuo = 0.85fc.(Ag - As) + Asfy (4.1)

where Ag = Gross area of the column


As = Area of Reinforcement

4.3 INTERACTION DIAGRAMS FOR HSC COLUMNS

The derivation of interaction diagrams for normal strength concrete columns is explained in
detail in text books [5]. The only difference in design of high-strength concrete columns is the
need for a more appropriate stress-block to calculate the compression force resisted by
concrete. This method is briefly described here.

In Fig. 4.1 Nu is the axial force applied at an eccentricity, e' from the centroid of tensile

reinforcement (or e from the section centroid). The co-existing ultimate moment about the

tensile reinforcement is M'u = Nue'.

as sumed
r ecta ngular stress ssc
Nu block
dsc Asc . a f 'c
e c = 0 .0 0 3
e C
e' Cc
gD kud b kud
D jud
d' sst
T=sstAst
e
s C c=(bkud)( .a f'c).b
plastic b
centroid As t C =sscA sc

Cross-sec tion S tra in (ku< 1) stress and force

Fig. 4.1 Basis of ultimate strength equilibrium equations for eccentrically loaded rectangular
columns

Using the Equivalent Rectangular Stress Block concept, the compressive force in the concrete
is:

4-3
Cc = a fc b kudb (4.2)

For fc < 60 MPa (existing recommendations in AS3600)

a = 0.85
(
b = 0.85 - .007 fc' - 28 )
where
0.65 b 0.85

For HSC, the rectangular stress block suggested in Section 3.3 or any other appropriate stress-
block can be used.

for 60 MPa f 'c 100 MPa


a = 0.85 - 0.0025( f ' c -60)
b = 0.65 - 0.00125( f ' c -60)
The force in the compressive steel is:

C = sscAsc (4.3)

The force in the tensile steel is:

T = sstAst (4.4)

The steel stresses have been expressed as ssc and sst since it is not determinable at this

stage of the analysis whether either or both of the steels have reached yield stress.

For force equilibrium, the internal forces must add up to Nu:

Nu = Cc + C - T (4.5)

Thus Nu = af'c b ku d b + ssc Asc - sst Ast (4.6)

Taking moments about the centroid of the tensile reinforcement:

Nu e' = Cc jud + C (d - dsc) (4.7)

4-4
Now jud = d(1 - 0.5bku). Using the expressions for Cc and C, we obtain:

M'u = Nue'

= a f'c b g ku d2 (1 - 1/2bku) + ssc Asc (d - dsc) (4.8)

and Mu = Nue = Nu(e' - d') = M'u - Nud' (4.9)

where Mu is the moment about the plastic centroid.

The point M uo (at A) on Fig.4.2 represents the ultimate strength in pure bending without

axial load and is found from beam formulae.

Fig. 4.2 A typical column interaction diagram

Nub and M ub (at B) are the axial load and moment capacity respectively at balanced failure,

i.e. when the bottom level of tensile steel has just yielded. The factor kub, which defines the

neutral axis position at balanced failure conditions is:

eu
kub = d = 0 .55 d (eu = 0 .003 ; if esy = 0 .0025 )
eu + esy

4-5
Notes:

1. esy=0.002 for 400 MPa steel and esy=0.0025 for 500 MPa steel.

2. When calculating Nuo, the maximum strain in the steel and the concrete of 0.0025 (Cl.

10.6.3 in AS3600) can be used. The strain at the maximum stress (fc) increases with concrete
strength (see Fig. 2.4). Therefore the present recommendations (in AS3600-2001) can be kept
without any modification.

4.3.1 Computer Generation of Interaction Curves for Rectangular Column


Sections- Spreadsheet INT_COL

It is obvious from the foregoing discussion that the hand computation of interaction curves is
tedious. Spreadsheet INT_COL generates interaction diagrams for rectangular column sections
(both normal strength and high-strength concrete) with up to 4 rows of non-prestressed
reinforcement. The spreadsheet INT_COL uses the Equivalent Rectangular Stress Block
[Section 3.3] to evaluate the concrete compressive force.

Steps in the analysis are as follows:

1. The ultimate strength of an axially loaded column (Nuo) is found using Eq. (4.1).

2. The moment capacity of the column when the axial load is zero (Muo) is calculated
using formulae for flexural members. The corresponding neutral axis depth and
maximum value of tensile strain in the extreme tensile concrete fibre are thus also
obtained.

3. A trial value of tensile strain in the extreme tensile concrete fibre is adopted and the
corresponding neutral axis depth evaluated. The steel and concrete forces are
calculated next. Nu is found using Eq. (4.5) together with Eq. (4.2) to evaluate the
concrete compression force. M u is calculated by summing moments about the plastic
centroid of the section. Hence, one pair of (Mu, Nu) co-ordinates is found.

4. The different (Mu, Nu) co-ordinates on the interaction diagram are found by
decreasing the value of tensile strain in the extreme tensile concrete fibre and
using 'Table Command' in Excel 5 to automatically fill a table of Nu and Mu
values corresponding to each particular strain value. These points are then plotted on a
graph to give the interaction diagram.

4-6
5. Strength reduction factors are calculated for each set of Mu and Nu using Table 2.3 of
AS3600 and the points on the reduced interaction curve are therefore obtained.
INT_COL can be used for both 400 MPa and 500 MPa steel.

Example: Interaction curves for a rectangular cross-section derived using INT_COL:

Fig. 4.3 Column Cross-Section


18000
16000 100 MPa
14000
Axial Force (kN)

100 MPa, fy=400 MPa steel


12000
80 MPa
10000
60 MPa
8000
6000 40 MPa
4000
2000
0
0 100 200 300 400 500 600 700 800 900 1000 1100
Moment (kN.m)

Fig. 4.4 Interaction Diagrams for Various Concrete Strengths


(500 MPa steel unless otherwise noted)

Fig. 4.4 shows how the interaction diagram developed using the rectangular stress block
changes with increasing concrete compressive strength. This diagrams also shows the
advantages of using high strength concrete for members with large axial forces. As can be seen,
the use of 500 MPa steel would not significantly change the capacity.

4-7
4.4 LATERALLY CONFINED CONCRETE COLUMNS-CALCULATION OF
CONFINEMENT PRESSURES

During the loading of a concrete column, a set of triaxial stresses are induced by the lateral
reinforcement which act upon the concrete core. The core is defined by the centreline perimeter
of the outer lateral reinforcement. The transverse reinforcement acts to confine the core under
the triaxial stresses when the stresses approach the uniaxial strength. Tests have shown that the
effective confinement of concrete can result in significant increases in both the strength and
ductility of a section.

During the loading of a column, stresses within the core increase up to the unconfined strength
of the concrete. At this stage the cover concrete becomes ineffective due to crushing, leaving the
core to carry the load. As the core approaches its uniaxial strength, the progressive internal
cracking produces very large transverse strains. Consequently, the core pushes against both the
transverse and longitudinal reinforcement, which in turn provide a reaction force, known as the
confining reaction. The confinement provided by the transverse reinforcement is known as
passive confinement.

The extent of change on the behaviour of the concrete is influenced by a number of factors such
as the volumetric ratio and configuration of lateral reinforcement, tie spacing, distribution of
longitudinal reinforcement and the characteristics of the steel.

It has been shown that the confining stress is not developed throughout the entire core due to
the effects of internal arching action of the stresses. The maximum transverse pressure from the
confining steel can only be exerted effectively on that part of the concrete core where the
confining stress has fully developed between the levels of reinforcement due to the arching
action. Hence, the effectively confined area that has a fully developed confining stress is less
than the core area (Fig. 4.5).

Aeff = keAcore (4.10)

where ke = confinement effectiveness coefficient.

4-8
There are many different versions of the development of ke (e.g. Fafitis and Shah [6]; M ander et

al.[7]; Sheikh and Uzumeri [8]).

For a circular section confined by circular reinforcement, the confinement is provided through
axial hoop tension which results in a continuous confining pressure. The internal arching action
is assumed to act in the form of a second degree parabola with an initial tangent slope of 45
degrees between the different levels of reinforcement.

Using M anders confinement model [7], the effectively confined core area can be defined as:

2
pd s2 s'
Aeff = 1 -
4 2d s (4.11)

while the concrete core given by:

pd s2
Acore =
4
(1 - r sl )
(4.12)

where rsl = volumetric ratio of the longitudinal steel in the confined core.

= total area of longitudinal steel / area of core

Therefore, the confinement effectiveness coefficient, ke, for a circular column is given by:

2
s'
1 -
Aeff 2d s
ke = =
Acore (1 - r sl ) (4.13)

4-9
Fig. 4.5 Confined core concrete in circular and rectangular sections

The behaviour of rectangular sections is similar to that of circular sections, although, rectangular
sections are much less efficient in confining the core area. The arching action (Fig. 4.5) occurs
between longitudinal bars at the point at which they are supported by the lateral reinforcement.

From the plan of a rectangular cross section, the effectively confined area is the core area minus
the sum of the parabolic areas:

bc d c -
n
( w' i )
2

i =1 6
The effective area midway between the lateral reinforcement is:

s' s'
1 - 1 -
2 bc 2d c
Thus, the effectively confined concrete becomes:

4-10

= bc d c -
( i ) 1 - s' 1 - s'
n w' 2
Aeff
i =1 6 2 bc 2d c

while the core is:

Acore = bc d c (1 - r sl )
Hence, the confinement effectiveness coefficient is:

n
2 s' s'
( w'i )
1
1 - 1 - 1 -
Aeff 6 bc d c 2 b 2 d c
=
i =1 c
ke =
Acore (1 - r sl )
For a square section, the equation is reduced to

n
2 s'
2

( w'i )
1
1 - 2 1 -
6d c 2d c
ke =
i =1

(1 - r sl ) (4.14)

A common assumption regarding normal strength concrete columns is that the ties reach their
yield point when either the peak axial load is reached, or even beforehand. Therefore, the
effective confining pressure for a square column can be derived as follows:

fr

Asy fyt Asy fyt


Fig. 4.6 Free-Body Diagram of Column Cross-Section

4-11
Letting fr . b . s = 2 Asy . fyt
where: fr = the confinement pressure,
b = the concrete core width,
s = the centre-to-centre tie spacing,
Asy = the cross-sectional area of the tie, and
fyt = the yield strength of the ties.
volume of lateral steel
rs =
Letting volume of core concrete
Asy . 4b
rs =
Then, b2 . s (for a square cross-section)

b 2 . s . rs
Asy =
So, 4b

2 b 2 . s . rs . f yt
fr =
And, 4 b2 s

f r = 0.5 r s f yt

When the effective confinement is introduced,

fr = 0.5 k e r s f yt (4.15)

Although this formula was derived for a square column section, it is also applicable for
rectangular or circular columns.

Example Calculation

Consider the following two sections:

4-12
(a) Section 1 (b) Section 2
Fig. 4.7a & b Two Column Cross-sections with Different Reinforcement Configurations

where: b (section width) = 400mm


d (section depth) = 400mm
concrete cover (to ties) = 40mm
N24 longitudinal bars
N10 ties @ 300mm
Variables Calculated:

ds = 400 - 2(40) - 10 = 310 mm ds = 400 - 2(40) - 10 = 310 mm

s = 290 mm s = 290 mm

wi = 400 - 2(40) - 2(10) - 2(24) = 252 mm wi = [(400 - 2(40) - 2(10) - 3(24)] / 2 = 114
mm

Ae = [(310)2 - (4/6)(252)2][1 - 290/(2x310)]2 Ae = [(310)2 - (8/6)(252)2][1 - 290/(2x310)]2

Ae = 15231.3 mm2 Ae = 22316 mm2

Acore = (310)2 - 4(p(24)2/4) = 94290.4 mm2 Acore = (310)2 - 8(p(24)2/4) = 92481 mm2

ke = 0.162 ke = 0.241

fyt = 500 MPa fyt = 500 MPa


Lateral Confinement

dt = 10 mm dt = 10 mm
pitch = 300 mm pitch = 300 mm

rs = volume of lateral steel (or tie perimeter x tie area) / core volume

4-13
310
tie perimeter = 4(310) mm tie perimeter = 4 ( 310 + )
2
rs = 0.34% rs = 0.58%

fr = 0.136 M Pa fr = 0.348 M Pa
This example shows the extra confinement provided when using Section 2.
As the confining pressure increases, so does the concrete strength and plastic deformation (see
Fig. 4.8).

Table 4.1 Summary of confinement pressure for different lateral configurations

400x400 section,

N24 long. reinforcement

40mm cover

N10@300 0.162 0.241 0.241 0.241


ke
N10@300
fr 0.136 0.204 0.306 0.348
N10@200
fr 0.348 0.354 0.778 0.886
N10@150
fr 0.578 0.588 1.294 1.472
N10@100
fr 1.056 1.076 2.366 2.692
N12@200
fr 0.513 0.524 1.139 1.296
N12@150
fr 0.854 0.870 1.894 2.155
N12@100
fr 1.561 1.593 3.465 3.944

4-14
Fig. 4.8 Stress-Strain relationship for a 60MPa concrete under different confining
pressures [9]

Researchers have shown (Setunge et al. [9]) the relationship between the enhancement in

strength due to confining pressure, fo, and a measure of ductility, (fr / f 'c), to be:

fo f
= 1 + 3 r for high strength concrete (4.16)
f'c f'c

fo f
= 1 + 4 r for normal strength concrete (4.17)
f'c f'c

where: fr = the confining pressure, and


fc = the uniaxial compressive strength.

Hence, the effectiveness of confinement in increasing the ultimate strength of concrete is less for
high strength concretes compared to normal strength concretes. In other words, the
effectiveness of ties placed in accordance with AS3600 [1] will diminish with increasing
compressive strength. Eqs. (4.16) and (4.17) can be used to obtain the enhanced concrete stress
due to confinement.

4-15
For the example presented in Fig. 4.7.

Section 1, fr = 0.136 M Pa; If fc = 80 M Pa.

The enhanced strength, fo = 80 + 3x0.136 = 80.41 M Pa.

Section 2, fr = 0.348 M Pa; If fc = 80 M Pa.

The enhanced strength, fo = 80 + 3x0.348 = 81.04 M Pa.

Note: Although the strength is only increased by a small percentage for this example, closer
stirrup spacings will increase the confinement pressure thus producing higher strength values.

4.4.1 Spread-sheets to calculate the confinement pressures

The EXCEL workbook labeled CONFINE provided with this book, can be used to calculate the
confinement pressures. The following spreadsheets are incorporated in CONFINE. They are
based on equations presented in the previous section.
CROSS_CONF Confinement pressure of rectangular/square columns with cross
stirrups (e.g. Fig. 4.7a Section 1)
DIA-CONF - Confinement pressure of rectangular/square columns with cross stirrups
(e.g. Fig. 4.7b Section 2)
SPIR_CONF Confinement pressure with spiral reinforcement
The effectiveness of confinement in a particular cross-section can be calculated using the above
spreadsheets.

4.5 DUCTILITY

Growing emphasis is being placed on the need for ductility of structures, especially those
constructed with high strength concrete. Designing for ductility ensures that the structures are
capable of large deformations at near-maximum load carrying capacity, in the extreme event of
such a load being applied. These large deformations give ample warning of failure and by
maintaining load carrying capacity, total collapse may be prevented altogether. To achieve the
desired level of ductility in column sections which are often subject to large axial forces,
designers need to make use of confining reinforcement [10]. Unfortunately, the amount of
confining reinforcement needed for a required level of ductility in column hinges is not given in
AS3600 [1]. The use of high strength concrete further increases the need for this omission to be
4-16
addressed. The safety margin that exists for normal strength concrete structures designed under
AS3600 must be maintained for higher strength concrete structures [11].

4.5.1 Moment-curvature Relationship

The ductility of a column hinge is characterised by the moment curvature relationship of the
column cross-section at the hinge location. In a beam segment subjected to end couples the
average strains vary linearly with depth, with the greatest variation occurring at wide deep
cracks. The end rotation q, depends on the length of the beam segment, hence it is not a

suitable measure of the local bending deformation. The ratio q/L however, defined as rotation

per unit length or curvature k, is independent of length and is a more appropriate measure of

the local bending deformation.

In practice, the curvature of reinforced and prestressed beams is calculated by measuring the
rotation over a particular length. In the absence of any other information the hinge length can
be approximately taken as 0.5 times the depth of the section.

M M

K a vge

C urv atu re, K

Fig.4.9 Variation of Curvature in a Constant Moment Region

Concrete sections characteristically soften, i.e. the bending moment capacity decreases with
increasing curvatures after the end of the plastic plateau. A typical moment- curvature curve is
given in Fig.4.10.

4-17
Various parameters which relate to the ductility of the concrete section are shown in Fig. 4.10
and include;

the yield curvature fy , defined as

f ' y Mi
fy =
M'y
where:

fy and My = the curvature and the corresponding moment calculated when either the

steel at the extreme tension face of the section is yielding, or when the
extreme concrete fibre compression strain reaches 0.002, whichever occurs
first, and

Mi = the maximum moment. It is assumed that once the first moment peak is reached, it is

not exceeded. This criterion can be checked by plotting the curve;

two values for the ultimate curvature (f0.8 and fu), defined as the curvature at which the

moment has reduced to 0.8 of its maximum peak, and the curvature which is obtained while
the moment is at the peak respectively are shown in Fig. 4.10. These ultimate curvatures can

then be used to calculate two different ductility parameters, fu / fy and f0.8 / fy. It is

recommended to use f0.8 / fy as the ductility parameter.

4-18
Fig. 4.10 Definitions of Parameters

In order to develop a moment-curvature curve for a particular section, an iterative procedure has
to be used. This procedure is explained in Appendix 4A. The spreadsheet program
M OM _CURV provided with the book calculates the moment-curvature relationship of a
rectangular cross-section, with and without axial forces.

M oment-curvature diagrams derived using spreadsheet M OM _CURV for the column section 1
in Fig. 4.7 (a) are shown in Figs. 4.11-4.13. Fig. 4.11 shows the reduction of ductility with
increasing axial loads. The moment-curvature curves for various concrete strengths are plotted in
Fig. 4.12. The axial load was kept constant at 5000 kN. These curves can be used to compare
the ductilities of columns with different strengths. As can be seen, the 100 M Pa column has a

low ductility ratio (f0.8 / fy) compared to lower strength columns.

4-19
-

800
3000kN
700

600 5000kN
Moment (kNm)

500
7000kN
400

300

200
0kN
100

0
0 0.00002 0.00004 0.00006 0.00008 0.0001 0.00012 0.00014

Curvature (rad/mm)

Fig. 4.11 Moment Curvature Diagrams for various applied axial loads
(Concrete strength = 80 MPa)

300 100MPa

80MPa
250
Moment (kN.m)

60MPa
200

40MPa
150

100

50

0
0 0.000002 0.000004 0.000006 0.000008 0.00001
Curvature (rad/mm)

Fig. 4.12 Moment Curvature Diagrams for various Concrete Strengths


(Axial load=5000 kN)

4-20
250

200
Moment (kN.m)

------------- 50mm - N16's


150 75mm - N12's

150mm - N12's
300mm - N12's
100

50

0
0 0.000002 0.000004 0.000006 0.000008 0.00001 0.000012

Curvature (rad/mm)

Fig. 4.13 Moment Curvature Diagrams for various tie spacings


(Axial load=5000 kN)

The reduction of ductility with increasing tie spacing is shown in Fig. 4.13.

4.6 DESIGN OF STIRRUPS FOR HIGH-STRENGTH CONCRETE


COLUMNS

M any formulae are available in codes, research papers and state-of-the-art reports to design
lateral reinforcement in normal strength concrete rectangular columns. Some code formulae (ACI
318 and AS3600) are reviewed here. A complete review is presented by M endis and Kovacic
[12]. The design rules suggested for normal strength concrete columns are unconservative for
high-strength concrete columns.

4.6.1 Current Practices

ACI 318-99
The design equations presented in the American Concrete Code, ACI 318-99, to calculate the
lateral reinforcement spacing of reinforced rectangular concrete columns can be classified into
two separate components. The first involves the general code provisions provided for structures

4-21
built in areas of very low seismicity, while the second relates to structures built in areas of
moderate to high seismic activity.

General Code Provisions (ACI 318-99)


The Formula
s = the lesser of:
16 longitudinal bar diameters,
48 transverse tie diameters, and the least cross-sectional dimension (4.18)
where: s = the spacing of transverse ties

Discussion
The formulation of the above equation is based on practical experience rather than research. The
main criteria which determine the spacing of transverse ties are the prevention of buckling of the
longitudinal reinforcement and the provision of adequate concrete confinement. The maximum
spacing is a function of the column dimensions, therefore the larger the column cross-section, the
larger the spacing. This takes into account the degree of penetration of the unconfined concrete
between the transverse steel, which is less significant for larger columns compared with smaller
columns.

Seismic Code Provisions (ACI 318-99 Clause 21.4.4.1)


The Formula

Ag f 'c
A sh = 0.3 s hc - 1
Ach fyh (4.19a)
or,

f 'c
A sh = 0.09 s h c
f yh (4.19b)
(whichever is lesser)

where: Ash = the area of lateral reinforcement (mm2),


s = the spacing of transverse ties (mm),

4-22
hc = the centre-to-centre cross-sectional dimension between ties (mm),
Ag = the gross cross-sectional area of the column section,

Ach = the concrete core cross-sectional area, from outer-to-outer of ties,


fc = the concrete compressive strength, and
f yh = the lateral reinforcement yield strength.
Discussion
Eq. 4.19(a) was derived on the basis that the additional load-carrying strength provided by the
spirals for concentrically loaded columns is equal to or greater than the strength lost when the
concrete cover spalls off. This principle has been with the American Concrete Code since 1963,
and the actual equation was derived in the 1930s. Eq. 4.19(b) was added for earthquake-
resistant construction, which governs for larger columns.

Eqs. 4.19(a) and 4.19(b) are a function of the concrete compressive strength. This is an
important relationship which takes into account the decrease in the effectiveness of transverse
reinforcement as the compressive strength is increased.

As a result of the common aim towards increasing ductility of normal strength concrete columns
in seismically active regions and high-strength concrete columns, coupled with the relationship
formulated between the spacing of lateral reinforcement and the concrete compressive strength,
designers have often opted to use the American Concrete Code seismic provisions in a more
general form to specify the required spacing of ties in high-strength concrete columns. The
strength of concrete being specified for columns is increasing with time and even though the
formulation of these equations has been based on a limited range of concrete compressive
strengths (such as from work dating back to 1929 by Richart et al. [13] using concrete
compressive strengths of just 18M Pa), the omission of any reference to the limitations of each
equation in the ACI code allows the engineer to extrapolate without due concern as to the
validity of their decision.

The Australian S tandards Code for Concrete S tructures (AS 3600)-present provisions
The Australian Code for Concrete Structures (AS3600) is very similar to the ACI Code with
respect to column detailing. Again, there are two separate components, the general code
4-23
provisions and the seismic provisions (Appendix A of AS3600). The discussion of the ACI
formulae is directly applicable to the AS3600 equations.

General Code Provisions (AS3600 Clause 10.7.3.3b)


The Formula

s = the lesser of: 15db, or Dc (4.20)

where: db = the diameter of the smallest bar in the column, and


Dc = the smaller column dimension.

The Seismic Code Provisions in AS3600 (Clause A11.3.2b) are similar to ACI equations [Eqs.
4.19(a) and 4.19(b)}.

Fig. 4.15 shows the stirrup spacing calculated by the aforementioned formulae for the cross-
section of the column given in Fig. 4.14. As seen from Fig. 4.15, ACI318-95 seismic formula
(only formula which has the concrete strength as a variable) gives unrealistic stirrup spacings for
higher concrete strengths. Hence a more rational design method is required for high-strength
concrete columns.

40 Cover
N24 Longitudinal Bars

400
N10

400

Fig. 4.14 Design Example - Column Cross-Section

4-24
Column Tie Spacing

Concrete Compressive Strength


120
ACI 318-99 General
100
ACI 318-99 Seismic
80 AS3600-General
(MPa)

60

40

20

0
0 50 100 150 200 250 300 350 400
Stirrup Spacing (mm)

Fig. 4.15 Stirrup Spacing in Accordance with Various Formulae

4.6.2 Design Method-Detailed

Defining the appropriate level of ductility for a concrete column may depend on the importance
of the structure or the level of earthquake risk in a particular area. The required amount of
ductility is achieved by providing closely spaced stirrups.
The design of lateral reinforcement in a column is based on three criteria.
(i). To prevent longitudinal steel from buckling.

An upper limit on the stirrup spacing should be placed to ensure that buckling of the
compression steel does not occur before failure of the column by hoop fracture or
crushing of concrete. In order to determine this upper limit, the buckling strain must be
determined. The buckling strain is calculated by an iterative method using an
appropriate model for the steel stress-strain relationship. The reinforcing bar is assumed
to buckle between the stirrups. An Euler buckling model was developed to simulate this
behaviour. This procedure is described in detail by Kovacic [14]. The strain at which
buckling of longitudinal steel occurs is independent of the compressive strength of the
concrete. Therefore the spacing requirements suggested for normal-strength concrete to
prevent buckling are applicable to high-strength concrete columns also [14].
(ii). To prevent fracture of lateral steel due to sudden concrete cover loss.

4-25
Unconfined cover concrete on high-strength concrete columns deteriorates at a higher
rate compared to that on normal-strength concrete columns. This may cause a transfer of
energy to the confining steel which may lead to the fracture of steel. In order to
investigate this, an energy balance method similar to that adopted by M ander et al. [15]
for normal strength concrete was adopted in a study conducted at the University of
M elbourne. It was shown that increasing compressive strengths from 50 to 80 M Pa
causes only a small reduction in the hoop fracture strain [14].
The basic energy balance equation is written as follows;

Usf = Ug - Uco (4.21)

where: Usf = the energy required for fracture of hoop steel at a cross-section,

Ug = the total energy absorbed by the column until fracture of hoop steel, and

Uco = the energy required for failure of an unconfined concrete column.


Using the integrals of the stress-strain curves of steel and concrete, the axial concrete

strain at which fracture of lateral steel occurs, ecu, can be determined. An elastic-plastic

stress-strain curve was assumed for both transverse and longitudinal steel and the
modified Scott model [see Section 2.7] was adopted as the concrete stress-strain model.
This procedure is explained in detail elsewhere [16].
(iii). To provide adequate confinement for ductility.
Adequate ductility for a high-strength concrete column is provided by comparing its
moment-curvature characteristics with that of a normal strength concrete column. The
level of ductility obtained for a 50 M Pa concrete column is therefore also ensured for a
similar column constructed with high-strength concrete.

To illustrate the application of this design method, the column cross-section shown in Fig. 4.14
was selected. For this column, AS3600 (applicable up to 50 M Pa) suggests a stirrup spacing of
360 mm. An 80 M Pa column was selected for comparison.

In order to determine the most appropriate stirrup spacing for the 80 MPa column in terms of
adequate ductility provision (equivalent to the standard 50MPa concrete column with 360mm

4-26
ligature spacing as specified by AS3600) as well as the prevention of both longitudinal steel
buckling and lateral steel fracture, a plot of the axial load level versus the ductility factor

f0.8/fy was developed, as shown in Fig. 4.16. As defined in Section 4.5.1, the ultimate

curvature f0.8 is the curvature at which the moment has reduced to 0.8 of its maximum peak

(see Fig. 4.10) and fy is the yield curvature.

The M -f curves were developed by entering different axial load levels into the spreadsheet

program M OM _CURV. The output data from the moment-curvature curves were then checked
to see if either the buckling or fracture strain had been reached before the curvature had reached

f0.8. The fracture strain had been reached for the 80M Pa column with a ligature spacing of

both 75 and 120mm when the axial load level was approximately 0.3 and lower, as shown in Fig.
4.16. The buckling strain was never reached in any of the columns.

From Fig. 4.16, it is clear that an 80M Pa column with an axial load level of 0.2 can reach the
same ductility level as a 50M Pa column designed under the current Australian Code provisions,
providing the lateral steel spacing is reduced from 360mm to 240mm. Similarly, an 80M Pa
column with an axial load level of 0.6 can reach the same ductility level as a 50M Pa column
designed under the current Australian Code provisions, providing the lateral steel spacing is
reduced from 360mm to 75mm. Kovacic [14] showed that the confinement is improved
significantly by providing 8 longitudinal bars (Fig. 4.17) compared to 4 bars in the section, due
to the larger effective core area.

4-27
0.6
50MPa - 360mm
80MPa - 240mm
0.5
80MPa - 120mm
80MPa - 75mm
0.4
Lateral Steel Fracture
P f c Ag

0.3
/ '

0.2

0.1

0
0 5 10 15 20 25 30 35 40
f 0.8 / f y

Fig. 4.16 Plot of Axial Load Level vs Ductility

The above procedure for determining the most appropriate stirrup spacing is quite lengthy and
complicated for design purposes. A simple formula incorporating similar features is suggested
in the next section.

N10s

400

N24s
40

400 400

Configuration A Configuration B

Fig. 4.17 Column Cross-Sections

4.6.3 Modifications Suggested for AS3600

A simple formula is proposed below to calculate the spacing of lateral reinforcement by


modifying the present requirements in the Australian Standard for Concrete Structures, AS3600.

4-28
This formula is based on analysing a number of typical columns [17] using the procedure
mentioned in the previous section.

The spacings of stirrups shall not exceed the smaller of values given below:

50<fc 80 M Pa* 80<fc 100 M Pa*


fc < 50 M Pa * *
(N /fcAg <0.3) (N /fcAg <0.3)
Dc, 15 db, 48 dt Dc, 15 db, 30 dt Dc, 10db, 24 dt

* For fc > 50 M Pa and N*/fcAg > 0.3, the spacings have to be reduced by a factor of 0.3/
*
(N /fcAg).
N* is the ultimate design axial load.
For fc > 50 M Pa, the recommended spacings are based on 12 mm, fyt = 500 M Pa stirrups. If

fyt
dt
other bar types are used the spacings have to be multiplied by a factor of .
500 12
Dc = the smaller column dimension
db = the diameter of the smallest longitudinal bar in the column
dt = the diameter of stirrups
fyt = yield strength of stirrups ( 500 M Pa)

As an example, for the column shown in Fig. 4.17 (Configuration A with fc = 100 M Pa), these
*
suggestions give a spacing of 150 mm with N10 (fyt = 500 M Pa) stirrups for N /fcAg equal to
0.4. With N12 stirrups the spacing is increased to 180 mm. The spreadsheet program
SPAC_LIGS provided with the book can be used to calculate these values.

D500N steel ligatures (fyt = 500 M Pa) are recommended for high-strength concrete columns.

D500L, low ductility steel should not be used as ligatures.

4-29
4.7 BEHAVIOUR OF SLENDER HIGH-STRENGTH CONCRETE COLUMNS

While the strength of a short column is governed solely by the strength of its cross-section, the
strength of a slender column is strongly influenced by its effective length. In most cases, the use
of high strength concrete in columns allows smaller cross-sections to be adopted, with the
implication being that slenderness effects are enhanced. This outcome means that consideration
and evaluation of slenderness effects in high strength concrete columns is more critical than in
normal strength concrete columns.

4.7.1 Recommendations for Slender Columns

The suitability of the moment magnification method suggested in ACI318 and AS3600 for high-
strength concrete slender columns has been reviewed in a paper written by the author [18].
From comparison with experimental results, it was shown that if the moment magnification
method suggested in ACI318 or AS3600 is used in conjunction with an interaction diagram
derived using an equivalent stress block suggested by the author (Section 3.3), then it is suitable
for HSC columns up to 100 M Pa. It was also shown [18] that the AS3600 formula predicts
more accurate results due to the better prediction of critical load, Nc , whereas ACI equations
give more conservative results. The spreadsheet INT_COL described in Section 4.3.1 can be
used to derive the interaction diagrams and moment magnification factors for HSC slender
columns.
Notes:
1. If more accurate results are required, the numerical method suggested in Ref. [18] can be
used to follow the load-deflection relationship of slender columns.

G
2. The long-term factor, bd (= ) used to calculate the buckling load for columns is
G+Q
conservative for high-strength concrete columns as the creep factor is less as described in
Section 2.5.2. However there is insufficient evidence to justify changing this factor.
Work is being currently conducted at the University of M elbourne to revise the long-
term factor.

4-30
4.8 TRANSMISSION OF HIGH-STRENGTH CONCRETE COLUMN LOADS
THROUGH SLABS

For economic reasons, columns in multistorey reinforced concrete buildings are often built with
high-strength concrete whereas slabs are built with concrete of weaker strength. In the preferred
method of construction, the columns are first cast up to the soffit of the slab they will support,
then the slab concrete is cast continuous through the columns. This process is repeated until the
last floor is cast. As a result, the axial load that is transmitted from a column above the floor
must pass through a layer of weaker concrete before reaching the column below the floor. A
design question arises as to what concrete compressive strength should be used in the design of
the column and the ways of improving the performance of the joint. Some buildings have already
been designed without considering this problem. There have been no systematic investigations
reported on the transmission of high-strength concrete column loads through lower strength
slabs, both locally and internationally, except for a study conducted by Ospina and Alexander
[19] at the University of Alberta, Canada. The limited amount of experimental studies
conducted on column-slab joints are given in the list of references [20-23]. None of the studies
covered the confinement of joint region by the stirrups and extra column reinforcement. The
major codes such as the Norwegian code [3], the CEB model code [24], Canadian Code [25], and
the New Zealand Code [26] or the ACI state-of-the-art reports on HSC columns do not provide
adequate design guidelines for the transmission of axial loads from high-strength concrete
columns through weaker slabs.

Design provisions in the American Concrete Code, ACI 318-95 [27] are based on the
experimental work carried out by Bianchini et al. [20] in 1960. These test results correspond to
relatively low slab strengths of 8.8 to 24.8MPa and column strengths of 15.8 to 56M Pa. Also
the slabs were not subjected to any loading. Design guidelines in AS3600 are based on ACI 318
recommendations.

Design provisions in ACI 318-95 suggest that when the column concrete strength does not
exceed by more than 40% the slab concrete strength, the design of the connection is based on the
column concrete cylinder strength. When this limit is exceeded, three solutions are presented.

4-31
The first solution refers to the construction procedure of puddling, in which column concrete is
placed within the slab-column connection region. Since both column and slab concrete should be
cast simultaneously, special care must be taken to avoid placement of the weaker concrete
within the joint region. The top surface of the puddled column concrete must extend at least
600 mm from the face of the column. Proper vibration of the column and slab concretes should
be accomplished to guarantee an optimum integration between both materials.

The second provision states that the axial load capacity of the column may be calculated based
on the lower intervening concrete cylinder strength (usually floor concrete) plus the addition of
vertical dowels and spirals to the connection. But no guidelines are given on how to design these
joints.

The third design recommendation is an equation to calculate the column effective compressive
strength as a weighted average of column and slab concrete cylinder strengths.

Selection of either the first or the second method has implications on ease of construction and
quality assurance. Presently, these methods have created serious problems for builders and
designers.

In the design of interior columns intersected by slabs of weaker strength, the joint concrete is
assumed to be confined and therefore capable of carrying compressive stresses in excess of the
cylinder strength of the slab concrete. This increased strength, defined as the effective
compressive strength, results from the lateral restraint provided to the joint by stresses acting in
the plane of the slab. The amount of restraint will vary according to the type of connection;
either interior, edge or corner (Fig. 4.18 ).

4-32
Normal-st rength
In terior Column concrete slab
(Hig h-st r ength )

Zer o Mom ent Border

Fig. 4.18 Typical Floor Slab

c
High-stren gth concrete column
Normal strength concrete slab

Fig. 4.19 Interior Column

Ospina and Alexander [19] have suggested the following formulae for interior and edge columns,
which are applicable up to 100 M Pa:

The resistance of the column in the joint region shall be based on an effective concrete

compressive strength, fce equal to:

a) for interior columns:

4-33

0.35 ' 0.25 '
fce = 1.4 -
'
fcs + fcc fcc'
h h

c c (4.22)

h
where shall not be taken less than 0.33. h and c are defined in Fig.4.19.
c
fcc is the concrete

strength of the column and fcs is the strength of the slab

fce' fce'
h
for ratios of 0.33 and 0.6. As
Fig. 4.20 shows the relationship between
fcs' and
fcs' c
h
seen at = 0.33, the present ACI 318 graph coincides with the above equation (graph
c

h
shown as HSC). However at = 0.6, both AS3600 and ACI318 equations predict
c

unconservative effective column strengths.

b) for edge columns

fce' = 1.4 fcs' fcc' (4.23)

c) for corner columns

Based on a study conducted recently Portella et al. [28] suggested the following formula [Eq.
(4.24)] for corner columns.

fce' = 1.3 fcs' fcc' (4.24)

Lee [29] extended Portellas work by including the effect of (h/c) ratio on the effective strength
of columns [Eq. (4.25)].

4-34
h
fce = 2.35 - 1.48 fcs fcc,
c
(4.25)
where fcs and fcc are the slab and column concrete cylinder strengths respectively. h is the

thickness of the slab and c is the column width.

4 h/c = 0.33
f'ce/f'cs

ACI318
3
AS3600
2 HSC

0
0 2 4
f'cc/f'cs

Fig.4.20(a) Effective strength of columns

4 h/c = 0.6
f'ce/f'cs

ACI318
3
AS3600
2 HSC

0
0 2 4
f'cc/f'cs

Fig. 4.20(b) Effective strength of columns

4-35
4.9 DESIGN EXAMPLES

Example 1

The column A-B shown in Fig. 4.21 (a) below has to carry a dead load of 1400 kN and a live
load of 1500 kN. The column cross-section is as detailed in Fig. 4.21 (b) and constant over the
entire height. The column is assumed to be pinned at the base and subjected to a uniaxial
moment of 200 kNm at the top. Design column A-B (fc= 80 MPa) assuming it is braced.

C
N24
N12

400

A 40

(cover to the
3m main rft.)
400
D D
Section D-D

B
(b)
(a)
Fig. 4.21 Column Details (Example 1)

Step 1
N* =1.25x1400 + 1.5x1500 = 4000 kN

N* 4000x10 3
= = 0.313
f ' c Ag 80x 400x 400
Design ligatures according to the rules given in Section 4.6.3.
Select N12 ligatures
Spacing = lesser of (400,15x24,30x12) = 360

4-36
0 .3 N*
Multiply by as 0.3
0.313 f ' c Ag
0 .3
Final spacing = 360x =345
0.313
Use N12 300 ligatures (any spacing up to 345 can be used).
Note:
1. The spreadsheet SPAC_LIGS provided with the book can be used to calculate these
spacings.

Step 2
Find the confinement pressure using spreadsheet CONFINE or use the method given in
Section 4.4.

fr = 0.52 MPa.
The enhanced strength is given by the following formula.

fo f
= 1 + 3 r
f 'c f 'c
fo = 80 + 3x 0.51 = 81.5 MPa.
Note:
The concrete strength can be taken conservatively as 80 MPa without calculating the enhanced
strength.

Step 3
Run INT_COL, Find Short/Slender and if applicable the magnified moment for the column.
The column is short. Hence design moment = 200 kNm. The point (200,4000) lies within the
s
inner curve. Hence 8N24 can be used.

(The outer curve shows the interaction diagram without capacity reduction factors)

4-37
14000
Column Interaction Diagr am

12000

10000
Axial Force, N (kN)
8000

6000

4000

2000

0
0 100 200 300 400 500 600 700
-2000
Mom e nt, M ( kNm )

Fig. 4.22

Example 2

An interior column (fc= 80 MPa) with a cross-section similar to Example 1 [Fig. 4.21(b)] has

to transmit an axial load through a weaker slab (fc= 32 MPa, slab thickness = 175 mm). Find

the effective design strength of the column.

h 175
= = 0.44
c 400

Now using Eq. 4.22 fi


0.35 ' 0.25 '
fce = 1.4 -
'
fcs + fcc fcc'
h h

c c

0.35 0.25
fce' = 1.4 - 32 + 80 fcc'
0.44 0.44

= 64.8 MPa

4-38
REFERENCES

1. AS3600: Concrete Structures Standard, Standards Association of Australia, 2001.

2. ACI Committee 441: State of the Report on High-strength Concrete Columns, American
Concrete Institute, 1997.

3. NS 3473 E Concrete Structures Design Rules, 4th Edition. Norwegian Council for
Building Standardization, Oslo, Norway, 1992.

4. Saatcioglu M. and Razvi S. R. Behaviour of Confined High-Strength Concrete Columns,


CSCE/CPCA Structural Concrete Conference, Toronto, Canada, 1993, pp. 37-50.

5. Darvall, LeP. P. Reinforced and Prestressed Concrete, Macmillan Co. of Australia,


1989.

6. Fafitis A. and Shah S.P. Lateral Reinforcement for High Strength Concretes Columns,
High Strength Concrete SP-87, ACI, Detroit, pp.212-232.

7. Mander J. B., Priestley M. J. N., and Park R. Theoretical Stress-Strain Model for
Confined Concrete, ASCE Journal of Structural Engineering, Vol. 114, No. 8, 1988, pp.
1804-1826.

8. Sheikh S.A. and Uzumeri S.M. Strength and Ductility of Tied Concrete Columns,
Journal of the Structural Division, ASCE, Vol.106 (5), pp.1079-1102.

9. Setunge, S., Attard, M. and Darvall, P.LeP. Ultimate Strength of Confined Very High-
Strength Concrete, ACI Structural Journal, Vol.90; No.563, pp.632-641.

10. Paulay T. and M.J.N. Priestley, Seismic Design of Reinforced Concrete and Masonry
Buildings, John Wiley & Sons, USA, 1992, 744pgs.

11. Attard M. M. and Mendis P. A. Ductility of High-strength Concrete Columns,


Australian Civil Engineering Transactions, Vol. CE35, No. 4. 1993, 295-306.

12. Mendis, P. and Kovacic, D., "Lateral Reinforcement Spacings for High-strength Concrete
Columns in Ordinary Moment-resisting Frames", Australian Structural Engineering
Journal, Vol. 2, Nos.2 & 3, 1999, pp. 95-104.

4-39
13. Richart F. E., Brandtzaeg A. and Brown R. L. A Study of the Failure of Concrete under
Combined Compressive Stresses, University of Illinois Engineering Experimental Station,
Champaign, Illinois, Bulletin No. 185, 1928.

14. Kovacic, D., Design of High-strength Concrete Columns, M.Eng. Thesis, The
University of Melbourne, 1995.

15. Mander J. B., Priestley M. J. N. and Park R., Seismic Design of Bridge Piers, Research
Report No. 84-2, University of Canterbury, New Zealand, 1984.

16. Mendis, P.A., Kovacic, D. and Setunge S., Basis for design of Lateral Reinforcement for
High-strength Concrete Columns, Structural Engineering and Mechanics Journal, Vol. 9,
No. 6, 2000, pp. 589-600.

17. Zois N., Stirrup spacings for High-strength Concrete Columns, Advanced Research
Project, The University of Melbourne, 1995.

18. M endis, P., "The behaviour of high-strength concrete slender columns, American Concrete
Institute Structural Journal, Vol. 97, No. 6, Nov-Dec 2000, pp. 895-901.

19. Ospina, C. and Alexander, S., "Transmission of High strength Concrete Column Loads
through Slabs, Dept. of Civil Engineering Report No. 214, University of Alberta, 1997.

20. Bianchini, A. C., Woods, R. E. and Kesler, C. E., "Effect of Floor Concrete Strength on
Column Strength," ACI Journal, Proceedings, Vol. 31, No. 11, 1960, pp. 1149-1169.

21. Gamble, W. L. and Klinar, J. D., Tests of High-Strength Concrete Columns with Inter,
ring Floor Slabs, Journal of Structural Engineering, ASCE, Vol. 117, No. 5, 1991, pp.
1462-1476.

22. Shu, C. and Hawkins, N. M., "Behavior of Columns Continuous through Concrete
Floors," ACI Structural Journal, Vol. 89, No. 4, 1992, pp. 405-414.

23. Kayani, M. K., "Load Transfer from High-Strength Concrete Columns through Lower
Strength Concrete Slabs, Ph.D. Thesis, Department of Civil Engineering, University of
Illinois, Urbana-Champaign, 1992, l l l pp.

24. CEB-FIP Model Code 90, Published by Thomas Telford, 1993.

25. CSA, A23.3-94, Design of Concrete Structures, Canadian Standards Association, 1994.

4-40
26. NZS 3101 Code of Practice for the Design of Concrete Structures, Standards Association
of New Zealand, Wellington, New Zealand, 1995.

27. ACI Committee 318: Building Code Requirements for Reinforced Concrete, Detroit,
American Concrete Institute. 1999.

28. Portella, J., Mendis, P., Baweja, D., Transmission of High-strength Concrete Column
Loads through Weaker Slabs, 16th Australasian Conference on the Mechanics of
Structures and Materials, Sydney Australia, Dec., 1999, pp. 139-144.

29. Lee, S.C., Effect of slab thickness and Column Rectangularity on Transmission of High-
strength Concrete Corner Column Loads through Weaker Slabs, M.Eng. Thesis, The
University of Melbourne, 2001.

30. Samra R. M., Ductility Analysis of Confined Columns, Journal of Structural Engineering,
ASCE, Vol.116 (11), 1990, pp.3148-3161.

APPENDIX 4A DEVELOPMENT OF MOMENT-CURVATURE


RELATIONSHIPS [SPREADSHEET MOM_CURV]

In order to develop a moment-curvature curve for a particular section, the following procedure
has been adopted:
The section is divided into a number of discrete laminae;

The extreme compressive strain, ecm, is incremented by a division of the maximum allowed

compressive strain (which is entered by the user);

The extreme tensile strain, et , is assigned a trial value;

The strain at the mid-depth of all laminae can now be calculated, assuming plane sections
remain plane;

The stress for each lamina is determined using the appropriate stress-strain relationships for
concrete, reinforcing steel and prestressing steel. (e.g. rectangular or modified Scott model
(16), Samra (30)]

4-41
If different arrangements of transverse reinforcement are to be used, the confinement pressure
must be entered manually. From the stresses above, the force for each lamina is computed;

To check the validity of the extreme tensile strain value which was chosen previously, the
sum of the laminae forces is calculated. By equilibrium, this sum of internal forces should
equal the externally applied axial force. If this is not the case, and the difference is
considerable, the extreme tensile strain is adjusted accordingly, and the whole process is
repeated.

Once the strain profile satisfies equilibrium, the moment, M, and curvature, f, values are

determined:

n
M = s ci Aci yci
i=1

e cm + e t
f=
D
as shown in Fig. 4A.1.

The extreme compressive strain is incremented to the next division, and the process is
repeated until the maximum allowable concrete strain is reached (which is entered by the
user).

Cover Core Concrete Cover Stress


ecm Core
Stress
M
y
y ci P
D Lamina of f Steel
area Aci Stress
et

Cross-Section Strain Profile Stress Profile Forces

Figure 4.A.1 Moment-Curvature Analysis of a Reinforced Concrete Column Section

4-42
CHAPTER FIVE

PUNCHING SHEAR RESISTANCE OF HIGH-


STRENGTH CONCRETE SLABS

5.1 INTRODUCTION

The use of high-strength concrete in reinforced concrete slabs is becoming popular in Australia
and other countries. Current design provisions of AS3600 and other major codes throughout the
world are based on empirical relationships developed from tests on low-strength concrete. In
this chapter, existing recommendations in design codes for punching shear failure of slabs are
reviewed. Design codes referred in this study are AS3600 [1] and CEB-FIP MC 90 [2]. In

AS3600 the punching shear strength is expressed as proportional to fc' . However in CEB-

FIP MC 90 punching shear strength is assumed to be proportional to


3 f'. It is shown that
c

the present provisions in AS3600 are applicable up to 100 M Pa.

The reinforced concrete flat slab system is a widely used structural system. Its formwork is
very simple as no beams or drop panels are used. However, the catastrophic nature of the
failure exhibited at the connection between the slab and the column has concerned engineers.
This area (Fig. 5.1) becomes the most critical area as far as the strength of flat slabs is
concerned due to the concentration of high bending moments and shear forces. The failure load
may be considerably lower than the unrestrained flexural capacity of the slab. The use of high-
strength concrete improves the punching shear resistance allowing higher forces to be
transferred through the slab-column connection. In spite of the wide use, only a few research
projects have been conducted on the punching shear resistance of high-strength concrete slabs.
The empirical expressions given in design codes are based on the experimental results from
slabs with concrete strengths between 15-35 MPa. Hence it is necessary to re-examine the
applicability of the present punching shear design methods for HSC slabs, using the published
data.

5-1
Column
Failure surface

Slab

Fig. 5.1: Punching failure surfaces of flat slab

5. 2 CODE DESIGN PROVISIONS


Most codes present formulae, where the design punching load is a product of a design nominal
shear strength and the area of a chosen control surface. Depending on the method used, the
critical section for checking punching shear in slabs is usually situated between 0.5 to 2 times
the effective depth from the edge of the load or the reaction. Influences of reinforcement, slab
depth and other parameters are customarily governed by the application of different
modification factors. The methods do not reflect the physical reality of the punching
phenomenon, but can, when properly calibrated, lead to reasonable predictions [3].

Generally the punching shear strength values specified in different codes vary with concrete
n
compressive strength fc and is usually expressed in terms of fc . In AS3600 (Cl. 9.2.3) the

punching shear strength is expressed as proportional to fc' . fc' is limited to 50 MPa in


the present code. The square-root formula in AS3600 is adopted from the ACI code [4]. ACI
provisions for punching shear are derived from Moes work on low strength concrete [5]. The

ultimate shear strength for slabs without prestress is given by Vuo = ud ( fcv )
where:

u= length of the critical perimeter, taken at a distance of d/2 from the column (mm)- see Fig.
5.2.

fcv =punching shear strength (MPa)

5-2
2
fcv = 0.171 + fc' 0.34 fc' (5.1)
bh
bh =ratio of longest column dimension to shorter column dimension

Fig. 5.2 - Control Perimeters specified in AS3600 and CEB-FIP codes

CEB-FIP MC-90 model Code [2] is also used by some engineers in Australia to design high-

strength concrete members. In MC-90 the punching shear resistance, Fsd is expressed as

(f )
1

proportional to ck
3
, Where fck is the characteristic compressive strength of concrete.

The highest concrete grade considered in MC90 is C80, which corresponds to fck equal to

80 MPa. Influences of reinforcement and slab depth are also considered in this design code.

Fsd = 0.12x (100 rfck ) u1d


1
3
(5.2)
where
200
x = 1+ is a size-effect coefficient
d
5-3
u1 = the length of the control perimeter at 2d from the column (Fig. 5.2)
r = rx ry
In the ultimate limit state the partial safety factor is 1.5. For the calculation of punching load
capacity

Eq. (5.2) is multiplied by 1.5, which gives Eq. (5.3).

Fsd = 0.18x (100 rfck ) u1d


1
3
(5.3)

In this study, measured concrete strength is taken as fck .

5.3 RECOMMENDATIONS

A comprehensive literature review revealed that only a few experimental studies have been

conducted on punching shear strength of high-strength slabs. These experimental results were

used [5] to check the validity of the punching shear strength formula given in AS3600 and

CEB-FIP MC-90.

It was found that only 2 out of 29 experimental capacities (Fig. 5.3) were below the AS3600
predictions with one result for a slab with a concrete strength of 108 M Pa. Therefore the
AS3600 formula [Eq. (5.1)] can be considered to be applicable up to 100 MPa. However the
ratios between observed and calculated loads clearly show that AS3600 is less conservative
for the HSC slabs and high scatter is found. As AS3600 provisions are similar to ACI
provisions, these conclusions are applicable to ACI 318-95. Generally CEB-FIP formula is
less conservative and may be unsafe for some cases. Therefore if the CEB-FIP code formula
[Eq. (5.2)] is used to calculate the punching shear strength, the concrete strength limit of 80
M Pa should be maintained.

5-4
1.80

1.60
Ptest/Ppred.

1.40
AS3600
1.20
CEB-FIP
1.00

0.80

0.60
40 60 80 100 120
Concrete Strength (MPa)

Fig. 5.3 - Ratios of experimental (Ptest) and predicted shear strengths (Ppred)

5.4 DESIGN EXAMPLE

c=400 mm

Square column
Slab (d=152 mm)

c+d
critical shear perimeter

Fig. 5.4 - Interior Column

For the design example shown in Fig. 5.4,

Punching shear strength Vuo = ud ( fcv )


where:

u= 4(c + d) = 2208 mm
fcv =punching shear strength (MPa)

5-5
2
fcv = 0.171 + fc' 0.34 fc'
bh
bh =ratio of longest column dimension to shorter column dimension (=1)

for fc' = 60 MPa

2
fcv = 0.171 + 60 0.34 60
1
= 2.63 MPa

Vuo = ud ( fcv ) = 884 kN


fVuo = 0.7 x 884 = 619 kN
Other values are given in Table 1. These values are plotted in Fig. 5.5. As seen there is a
significant increase in punching shear capacity with increasing concrete strengths.

Table 1: Punching Shear Capacity

fc' (MPa) fcv (MPa) Vuo (kN) fVuo (kN)


30 1.86 625 438

40 2.15 722 505

50 2.40 807 565

60 2.63 884 619

70 2.84 955 668

80 3.04 1021 714

90 3.23 1083 758

100 3.40 1141 799

5-6
Punching Shear Capacity (kN)
800

600

400
20 40 60 80 100
Concrete Strength (MPa)

Fig. 5.5 The Graph of Punching Shear Capacity Versus Concrete Strength

REFERENCES
1. AS3600: Concrete Structures Standard, Standards Association of Australia, 1994.
2. CEB-FIP State-of-the-art report on high-strength concrete, 90/1/1, Bulletin dInformation
No. 197, 1990.
3. ACI Committee 318: Building Code Requirements for Reinforced Concrete, Detroit.
American Concrete Institute. 1995. CEB-FIP M odel Code 1990, Thomas Telford Ltd.,
London, 1993.
4. M oe, J., Shearing Strength of Reinforced Concrete Slabs and Footings under Concentrated
Loads, Development Bulletin No. D47, Portland Cement Association, Skokie, 1961, 130
pp.
5. Ngo, T., Punching shear capacity of HSC slabs, Electronic Journal of Structural
Engineering, http://www.civag.unimelb.edu.au/ejse , Vol. 1, No. 1, 2001, pp. 52-59.

5-7
CHAPTER SIX

DESIGN OF REINFORCED HIGH-STRENGTH


CONCRETE WALLS

6.1 INTRODUCTION

In the past, concrete walls were designed in most structures for protection against the
exterior environment with little consideration to the capability of the wall as a structural
member. This was mainly due to the very low allowable design stresses for walls
specified in early versions of concrete codes.

Over the years, reinforced concrete walls have gained greater acceptance as load carrying
structural members from practising engineers. This is due to the increased research
undertaken on concrete walls and subsequent increases in allowable design stresses
incorporated in various concrete codes.

Reinforced concrete walls are now considered as important structural elements, with
major design codes allocating separate chapters for wall design. Reinforced concrete load
bearing walls can be used in various design situations. These include designing:
(i) walls to act as integral components in core wall systems,
(ii) shear walls to resist lateral loads such as wind or seismic effects acting in-plane,
and
(iii) external walls to resist combinations of horizontal and in-plane vertical forces.

The increased acceptance of tilt-up and other types of precast structures and a trend
towards reinforced concrete core walls in high-rise buildings is the reason for the
newfound popularity in the use of load-bearing reinforced high-strength concrete walls.
Enormous cost reductions can be obtained by the use of thinner high-strength concrete

6-1
walls. Thinner walls not only reduce the cost but also increase the net lettable space of a
building.

At present, wall design practice in Australia involves the use of the concrete standard
AS3600 [1], with some designers preferring to use the American and British codes for
guidance. These wall design methods are intended for load bearing walls where end
restraints are usually provided at top and bottom edges, and side edges are free. The
codes fail to recognise any contribution to load capacity from restraints on side edges,
which is a common feature of core walls. Consequently, there is a need to incorporate
this contribution in current wall design methods for a fully optimised design to be
realised. Development of a realistic design procedure for a wall panel requires a good
understanding of its strength and behaviour as influenced by its geometry, support
conditions, etc.

Initially this chapter will review the available methods for design of reinforced concrete
walls. The background to the derivation of the present AS3600 formula is given next.
Finally recommendations for design of HSC walls are presented.

6.2 AS3600 WALL DESIGN METHOD

The Australian concrete structures code AS3600 [1] gives a simplified procedure for the
design of walls. The ultimate design axial strength per unit length of a braced wall in
compression is given by the following formula.

fNu = f(t - 1.2e - 2ea)0.6f'c (N/mm) (6.1)

where

f = 0.6 for compression members.

t = thickness of the wall (mm).

e = eccentricity of the load measured at right angles to the plane of the wall (mm).

ea = (Hwe)2/2500t an additional eccentricity due to deflections in the wall (mm).

6-2
f'c = characteristic compressive cylinder strength of concrete (MPa).

Hwe = the effective height of a wall (mm).

The AS3600 equation applies to walls where the slenderness ratio, Hwe/t 30. If
the design axial force, N* 0.03f'cAg then Hwe/t 50. A practice sometimes
adopted in Australia is to use Hwe/t 20 when large axial loads are encountered.

The walls are subjected to a minimum reinforcement ratio (rw) of 0.0015 vertically and

0.0025 horizontally. Although f'c is restricted to the 20 - 50 MPa range, this formula is
being used to design walls with higher strengths.

6. 2.1 Background

Commentaries on AS3600 give little reference as to how the wall formula is derived. The
design equation is taken from the British standard [2], with small differences attributed
to the different load and material strength factors used [3]. BS8110 specifies the
following formula for the design of slender braced plain walls.

nw = 0.3(t - 1.2e - 2ea)fcu (6.2)

where fcu is the characteristic compressive cubic strength of concrete (MPa).

This British equation is based on the assumption of a simple rectangular stress-block

(see Fig. 6.1), with an average stress of lwfcu.

From Fig. 6.1 it is seen that: nw =lwfcu (t - 2e) (6.3)

6-3
nw
e

lw f cu

x = t - 2e

Fig. 6.1 Stress-block under ultimate conditions

The coefficient lw is dependent on the type of concrete and on the wall dimensions in

consideration. BS8110-97 allows lw to be equal to 0.3 for all conditions. In the

previous British code, CP110-72, lw varied from 0.3 to 0.5. BS8110-97 conservatively

uses the lower bound value. This lower bound value is applied to take account of the
increased difficulties of controlling the quality of concrete used in the finished wall [4].

To allow for additional deflection developed in slender walls (see Fig. 6.2), the total

eccentricity in the central region is taken as 0.6e + ea. Hence, Eq. (6.3) gives
nw = 0.3fcu[t - 2(0.6e+ea)]
= 0.3fcu(t - 1.2e -2ea) giving the result of Eq. (6.2).

6-4
e

Total eccentricity
taken as 0.6e + e a

Deflection
of wall, e a

Fig. 6.2 Eccentricity in a slender wall [4]


This expression is an empirical expression for the strength in terms of actual eccentricity

and the additional eccentricity ea. In BS8110, ea for a column is given as follows.

ea = b(1/2000)(Hwe/t)2t (6.4)

where b is taken as 0.8 for walls.

The only major differences between the AS3600 and BS8110 formulae are the strength
factors and the concrete compressive strength values used. If the readily assumed

relationship, f'c = 0.8fcu is used, Eq. (6.2) would yield:

nw = 0.375(t - 1.2e - 2ea)f'c.

And, if f = 0.6 is substituted in Eq.(6.1) then:

6-5
fNu = 0.36(t - 1.2e - 2ea)f'c, yielding almost identical results.

The slight difference may be attributed to the larger factored basic loads (n = 1.4DL +

1.6LL) used in BS8110. The AS3600 equation uses the factored basic load N* = 1.25DL
+ 1.5LL.

6.3 THE AMERICAN (ACI 318-99) WALL DESIGN METHOD

ACI318-99 [5] gives the equation for the design axial load strength of a wall as:

fPnw = 0.55ff'cAg[1 - (kH/32t)2] (6.5)

where

f = 0.7 for compression members.

k = 0.8 for walls restrained against rotation.

= 1.0 for walls unrestrained against rotation.


= 2.0 for walls not braced.

Ag = Lt, gross area of the wall panel section (mm2).

where L = horizontal length of the wall and t = thickness of the wall

H = height of the wall (mm).

The equation is based on one way action. (i.e. walls restrained top and bottom) and

applies to walls where H/t 25 or L/t 25 whichever is less for loadbearing walls.
The minimum thickness is 100mm. According to ACI 318-99 [5], the minimum

reinforcement ratios (rw) required are 0.0015 vertically for deformed bars > 16mm

diameter (0.0012 for deformed bars 16mm diameter or mesh) and 0.0025 horizontally
for deformed bars > 16mm diameter (0.0020 for deformed bars 16mm diameter or

mesh). If these criteria are not met, the design strength is reduced to,

6-6
fPnw = 0.45ff'cAg[1 - (kH/32t)2] (6.6)

where f = 0.65.

If Eq. (6.6) is used then H/t 24 and the minimum thickness is increased to 140mm.
The resultant load must be in the 'middle third' of the overall thickness of the wall for
the ACI formulae to be valid. This allows for a maximum eccentricity allowance of t/6.

6.4 COMPARISON OF THE CODE WALL EQUATIONS

As mentioned earlier, some designers in Australia use the ACI formula to design walls
[6]. The comparisons of the code formulae for two different eccentricities are plotted in
Figs. 6.3 (a) and (b). The following assumptions are made in the analysis.

1. e = t/6 for Fig. 6.3 (a) (maximum eccentricity allowed in ACI318) and e = t/20
for Fig. 6.3 (b).

2. Hwe = 1.0H (equivalent to k = 1.0), i.e. simply supported top and bottom.
3. f = 0.6 for Eq. (6.1) and f = 0.7 for Eq. (6.5).

With these assumptions the formulae can be rewritten in their design axial strength ratio
form as given in Table 6.1.
Table 6.1 Design axial strength ratios of code formulae

Code Design axial ratio Expression (e=t/6) Expression (e=t/6)


AS3600[1] fNu 0.288[1 - (1/1000) (H/t)2] 0.338[1 - (1/1175) (H/t)2]

f'ct
ACI318[5] fPnw 0.385[1 - (1/1024) (H/t)2] 0.385[1 - (1/1024) (H/t)2]

f'ct

6-7
0.5 0.5
Axial Load Ratio ACI318: e=t/6

Axial Load Ratio


0.4 0.4 ACI318:e=t/20
0.3 0.3
0.2 0.2 AS3600:e=t/20
AS3600: e=t/6
0.1 0.1
0 0
0 10 20 30 40 0 10 20 30 40
Slenderness, H/t Slenderness, H/t

(a) (b)

Fig. 6.3(a) and (b) Comparison of design axial load ratios (a) e = t/6 and (b) e = t/20

The ACI formula appears to give a higher design axial load for differing values of

slenderness when the eccentricity is t/6 (see Fig. 6.3a). For H/t >20 the ACI curve falls
off at a more dramatic rate than the AS 3600 curve. This dramatic reduction in capacity
would have been the reason for the lower slenderness ratio limit specified in the ACI code
(The maximum H/t in ACI318 is 25, whereas AS3600 may be used for cases up to H/t =
30). As seen from Fig. 6.3(b), for a lower eccentricity (e = t/20) the difference is not
significant. In some instances the AS3600 is superior as the eccentricity is included in the
design equation. It must be noted that although ACI318 formula gives higher capacities,
the load factors specified in ACI 318 are relatively higher. These load factors are given in
Table 6.2.

Table 6.2 Load Factors

Code DL LL Average
AS3600[1] 1.25 1.50 1.375

ACI318[5] 1.40 1.70 1.55

6.5 RECOMMENDATIONS FOR HSC WALLS

According to the AS3600 equation, a particular percentage increase in concrete strength


will result in the same percentage increase in wall strength. This observation is not the

6-8
case with the experimental results which indicate that a particular percentage increase in
concrete strength results in a much smaller increase in wall strength [7,8]. A simple
modification to the present formula was suggested [7-9] to take account of this non
proportional increase. It was considered that for a 40% increase in concrete strength
(from 50MPa to 80MPa) a 25% increase in wall strength is appropriate. This is an
acceptable assumption considering the average increase obtained in the testing. However

the f= 0.6 factor assured all the test results up to 80 MPa failed well above the

predicted design load. Therefore the present formula [Eq. 6.1] can be retained up to 80

MPa. This formula should be used with caution for f'c > 80 MPa, since no testing was

done for walls with strengths exceeding 80 MPa. An extrapolation of experimental


results suggests that the present formula can be used up to 100 MPa with the restriction

of Hwe/t 25.

For core walls and other walls with restraints on three or four sides, the design strength
can be further improved by considering the German Code effective height factors, which
consider the side restraints [see Eq. (6.8)].

6.6 APPLICATION OF FORMULAE IN CORE WALLS

A very common approach for design of core wall structures is to determine axial load
capacities of the wall elements in the core using the code design formulae and comparing
these with the loads acting on the structure. Initially the core box is divided into wall
elements and section properties are calculated. Then dead, live and wind loads are
determined for the calculation of design axial loads and moments (dead and live load
actions are determined via a load run-down of the structure which includes floor framing,
wall layout and thickness, various design live loads, compactus, etc. Wind actions are
determined from a computer program such as "Super-Etabs").

6-9
The ultimate stresses at the critical points (normally corner points) are determined by
the following expression:
(P/A M/Z)wind + (P/A)dead and live (or dead only) (6.7)

These stresses are compared with the capacities of individual wall elements. If these
conditions are not satisfied the core must be redesigned. Where the centroid of vertical
load is eccentric from the centroid of the core, consideration may need to be given to the
bending stress resulting from this eccentricity.

The incorporation of the above procedure in a spreadsheet is very useful in optimising


core wall thicknesses. This procedure is explained in detail by Mendis et al. [10]. When
the thicknesses are changed, new section properties are calculated and the program
checks the new conditions. If the design conditions are not satisfied the program issues a
warning and wall thicknesses are optimised until this stage is reached. The spreadsheet
allows for efficient design of concrete walls by the use of the AS3600, ACI or other
modified wall design formulae and by considering other aspects such as wind
directionality effects [10].

6.6.1 Case study

A tall concrete building designed in Melbourne, Australia is taken as an example to


illustrate the significance of the improved capacities obtained by the use of high-strength
concrete. It is a 47 storey building with a height above raft equal to 197m. The typical
cross-section of the building and the elevations are shown in Fig. 6.4. The core structure
was slightly modified to form a symmetrical cross-section. The modified section is
shown in Fig. 6.5.

Core box 2 was analysed using the spreadsheet for the purposes of this case study [10].
Initially, the modified AS3600 wall equation (Eq. 6.1) was used with an effective length
factor of 0.8. The wall thicknesses were then optimised for various values of concrete

6-10
strength. The resulting thicknesses are given in Fig. 6.6 (Note: A minimum wall
thickness of 200mm was incorporated. Thicker walls were rounded up or down to the
closest 25 mm). Fig. 6.6 indicates that a substantial saving of wall thickness and cost can
be obtained by using higher strength concrete. This is particularly evident in walls 2 and
3 where wall thickness can be reduced by large percentages by increasing concrete
strength.

Currently the AS3600 code allows an effective length factor k=0.75 to be adopted in the
design formula and a factor of k=0.8 is adopted in the ACI wall design formula when the
wall is braced on top and bottom against lateral translation and restrained against
rotation at one or both ends. The German code covers walls with more than two
restraints. The German Code effective length factors are:
k = 1.00, for walls restrained on two sides.
1 0.3, for walls restrained on three sides.
k =
1 + (H/3d)2
1
k = for H d
1 + (H/d)2
d
or for H > d for walls restrained on all four sides. (6.8)
2H
d = the distance of the free edge from the middle of the shear wall 'or' half the
centreline distance between shear walls.

Fig. 6.7 shows the resulting thicknesses if the German code effective length factors were
used along with the AS3600 formula [Eq.(6.1)]. This simulates a more realistic situation
since side restraints are taken into account. The main observation here is the amount of
reduction that can be obtained using appropriate effective length factors. It should also
be noted that in both cases [Figs. 6.6 and 6.7] the advantages of high strength concrete
seems to drop its effectiveness at concrete strengths between 80 to 90 MPa. The
optimum concrete strength may differ for different core sections.

6-11
y
North

20 Wind Direction

b= 53 m

h = 197 m

South Ele vation East Elevation

Fig. 6.4 Typical Cross-sections and Elevations

3 Core
Box
No.

2 2 4 3 4

1
wall numbers

1 6 5

Fig. 6.5 Typical Core Sections

6-12
Fig. 6.6 Change in Wall Thickness with Varying Concrete Strength
(AS3600 with k=0.8)

Fig. 6.7 Change in Wall Thickness with Varying Concrete Strength


(AS3600 with German k factor)

6.6.2 Cost savings

It is clear that significant cost savings can be achieved if a higher concrete strength and
more realistic design formulae are used.

6-13
A cost saving analysis has been carried out for the example given above to determine
which concrete strength gives the overall minimum wall cost.

The analysis takes into account both the increase in the net lettable area and the saving in
concrete material cost when higher concrete strengths are utilized. The 40 to 80 MPa
concrete strength range is considered. The cost benefit associated with the increase in net
lettable area is assumed to be $3500/m2 while the cost of concrete has been taken as
3 3
$250/m for 40 MPa concrete, increasing proportionally to $306/m for 80 MPa
concrete.

Cost savings for the walls analysed using AS3600 formula with k=0.8 are given in Table
6.3. Table 6.3 clearly indicates the benefits of using high strength concrete for wall
construction. A 40 MPa concrete strength is used as the basis for cost comparison. In
terms of material cost, HSC is most economical since its strength increases linearly more
quickly than its cost. Hence a wall of relatively smaller thickness and decreased cost
results with the use of HSC. The cost savings given in Table 6.3 are per storey height,
therefore if for example a 47 storey building was being considered, the cost saving would
be 47 times the figures given.

Table 6.3 Cost Savings per Wall for Increasing Concrete Strength.

Cost savings per wall as compared to


a 40MPa wall ($)
Concrete Wall 1 Wall 2 Wall 3 Wall 4 Total Savings
Strength (MPa) Per Core Box
50 824 3481 1344 1735 7384

60 797 4181 1269 3515 9762

70 770 6840 2076 3403 13089

80 743 7606 2309 4270 14928

6-14
As a comparison, the cost benefit for walls 2 and 3 was determined using the German
code effective length factors for walls supported on 4 sides [Eq. (6.8)], as opposed to
the ACI method of assuming an effective length factor k= 0.8.The results are given in
Table 6.4.

As expected, further cost savings are gained using the German code effective length
factors. Again, cost savings increase with concrete strength and Table 6.4 confirms that
80MPa is the most economical choice of concrete strength.

Table 6.4 Comparison of Cost savings using the German k factor and the k = 0.8 factor

Cost savings per wall as compared to a 40 MPa wall ($)


Concrete Wall 2 Wall 2 Wall 3 Wall 3
Strength (MPa) (German k factor) (k = 0.8) (German k factor) (k = 0.8)
50 4427 3481 1347 1344

60 6095 4181 1857 1269

70 8775 6840 2380 2076

80 9565 7606 2917 2309

6.7 OTHER FORMULAE

Apart from the AS3600 and ACI formulae, other researchers have suggested their own
design formulae. Most of these formulae are related to or are variations of the ACI
equation. These formulae are detailed in a paper published in the ACI journal [6]. The
large amount of research on the ACI wall design method has led to its design formula
giving more increased capacities than the AS3600 equation. This may be the reason why
some designers in Australia prefer to use the ACI method to design concrete walls.

The most notable research was that of Saheb [11]. He suggested improvements to the
current ACI formula by incorporating and investigating the effects of an aspect ratio
(H/L), slenderness ratio (H/t), thinness ratio (L/t) and reinforcement ratio (rv). He

6-15
suggested formulae for walls supported top and bottom (one-way) only and also for
walls supported on four sides (two-way).

The Saheb one-way equation includes the aspect, slenderness and reinforcement ratios.
Its design axial strength formula is of the following form:

Pu = 0.55fAg f'c[1 - (H/32t)2[1 + (fy/f'c - 1)rv] for H/L 2 (6.9a)

Pu = 0.55fAg f'c[1 - (H/32t)2[1 + (fy/f'c - 1)rv][1.2 - H/10L]

for H/L < 2 (6.9b)

The two-way equation is of similar form to the one-way equation but it also includes a
thinness ratio (L/t).

Pu = 0.67fAg f'c[1 - (L/120t)2][1 + 0.12(H/L)

valid when: 0.5 < H/L < 2.0 & L/t < 60 (6.10)

These equations differ to other wall formulae in that they depend on a number of
variables not considered before. In addition, the two-way formula could be advantageous
in core wall structures where walls are supported on all sides.

More economical sections can be obtained by further refining the design process, briefly
described in this chapter by conducting a rigorous analysis using an analytical method
involving beam-column stability methods. Fragomeni and Mendis [12] have developed
the computer program WASTAB to conduct this analysis. More details about the
program are given in a paper published in ASCE, Journal of Structural Engineering [12].

REFERENCES

1. AS3600 "Concrete Structures", Standards Association of Australia, North Sydney,


2001 (previous version 1994).

6-16
2. BS8110 "Structural Use of Concrete, Part 1 - Code of Practice for Design and
Construction", British Standards Institution, London, 1997.

3. Walsh, P. "Use of the Australian Standard for Concrete Structures", CSIRO, Inkata
Press, Melbourne and Sydney, 1988.

4. Rowe, R.E. et al. "Handbook on the Unified Code for Structural Concrete", Cement
and Concrete Association, London, 1972.

5. ACI318 - 99 "Building Code Requirements for Reinforced Concrete", American


Concrete Institute, Detroit, 1999.

6. Fragomeni, S., Mendis, P.A., and Grayson, W.R. "A Review of Reinforced
Concrete Wall Design Formulae", Journal of the American Concrete Institute, Sep.-
Oct. 1994, pp. 521-529.

7. Fragomeni, S. Mendis, P.A. and Grayson, W.R. "Axial Load Tests on Normal and
High Strength Concrete Wall Panels", Civil Engineering Transactions, The Institute
of Engineers, Australia, Vol. 36 No.3, Aug. 1994.

8. Fragomeni, S., "Design of normal and high-strength reinforced concrete walls", Ph.D.
thesis, The University of Melbourne, Australia, 1995.

9. Fragomeni, S. and Mendis, P.A. , "Design equations for normal and high-strength
concrete walls", Concrete in Australia, Vol. 23, No.3, Oct-Dec, 1997, pp 22-24.

10. Mendis, P.A. Karantzas, V. and Fragomeni, S. "Efficient Design of Concrete Cores",
Proceedings of the 18th conference on Our World in Concrete and Structures,
Singapore, August, 1993.

11. Saheb, S.M. "Strength and behaviour of Reinforced Concrete Wall Panels", Ph.D
Thesis, Indian Institute of Science, Bangalore, 1985.

12. Fragomeni, S. and Mendis, P.A., "Instability analysis of normal and high-strength
reinforced concrete walls", Journal of Structural Engineering, ASCE, Vol. 123, No. 5,
May 1997, pp. 680-685.

6-17

Vous aimerez peut-être aussi