Sei sulla pagina 1di 488

Behaviour of Prestressed Concrete Slab-Column Connections

by

Yadav Raj Khwaounjoo


B.Sc. (Engg.), M.Eng.

A thesis submitted as partial fulfilment


of the requirements for the degree of
Doctor of Philosophy

School of Civil and Environmental Engineering

The University of New South Wales

Sydney, Australia

February 2001
CERTIFICATE OF- ORIGINALITY

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

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

/~/'/
f 'hw-~
:t-1/•1/fl"'
Yadav Raj Khwaounjoo
:JJ rulivairvd ~ rrvlf ~~
Q/YL~
.}( 53 aluuLU/f'-- ~ y uba/l'VrVa :£0/X/rrUi
ACKNOWLEDGMENTS

The research presented in this thesis was undertaken in the School of Civil and
Environmental Engineering at the University of New South Wales, Australia.

The author is highly indebted to Dr. Stephen J. Foster for continuous guidance and
support he has given throughout this research. Dr. Foster showed special interest in
this research project and provided information, valuable suggestions, encouragement
during all phases of the research program. The author is also grateful to
Prof. R. Ian Gilbert for his suggestions, comments and supports throughout the period
of author's PhD study.

I would like to extend my appreciation to Prof. Somasundaram V alliappan for his


interest and support to the successful completion of my research work. I thank
Prof. Mark A. Bradford for the constructive criticism and suggestions during the
reviews of my research.

The computation work of this research was carried out in the computer laboratory of
the School of Civil and Environmental Engineering. The assistance of the computer
support staff is appreciated. The experimental programme described in this thesis was
undertaken at the Randwick Heavy Structures Laboratory. The successful completion
of the test programme was made possible by dedicative efforts of Messrs Chris
Gianopoulos, Tony Macken, Frank Scharfe, Ron Moncay, Paul Gwynne, and William
Terry.

Finally I wish to thank my wife Lila, daughter Prashannata and son Prashanna without
whose love, patience, encouragement and understanding this dissertation would never
have been completed.
ABSTRACT

In this study the behaviour of prestressed slab-column connections failing in punching


shear is investigated using both finite element and experimental modelling. Tested
were six anchorage pullout specimens and four one-third scale edge and comer
column connections.

The finite element program DIANA was used to analyse punching type problems
using 20-node brick elements to model the concrete and 2 or 3-node embedded bar
elements to model the reinforcing and prestressing steel. The model was calibrated
using a wide range of experimental tests and with the results of the pullout tests
undertaken in this study. It was concluded that the finite element models used in
DIANA give reasonable results provided that the studies are accompanied with
parameter sensitivity analyses. The softening slope of the concrete tension stress-strain
curve was shown to affect the flexural response of the slabs while the shear retention
factor was shown to be highly influential on the failure load for slabs failing in
punching shear.

In the main experimental investigation three one-third scale edge column connections
and one comer column connection were tested with three of the four specimens being
partially prestressed. Punching shear failure was observed in all of the specimens
tested with the strength and behaviour of the specimens influenced by the degree of
the applied prestress. The results of the experimental testing and subsequent numerical
modelling showed that the capacities of flat plate slab-column connections increase
approximately linearly with increasing slab prestress.

Shear force and moment data obtained from a wide range of experimental tests were
compared with predictions from five code models. There was a significant degree of
scatter between the test data and the predicted strengths with only the Australian code
giving a lower than five percent probability that a single specimen will fail below the
capacity given by the code model. The ACI code, Eurocode 2 and CEB-FIP Model
code fall slightly below the 95 percent characteristic level. The British code model
gave an unacceptably high probability of failure for the design of slabs failing in a
punching mode.
CONTENTS

Acknowledgments iv

Abstract v

Contents vi

Nomenclature xiii

CHAPTER 1 -INTRODUCTION 1-1

1.1 GENERAL 1-1

1.2 OBJECTIVES 1-4

1.3 SCOPE 1-4

1.4 ORGANISATION OF THESIS 1-5

CHAPTER 2- REVIEW OF LITERATURE 2-1

2.1 INTRODUCTION 2-1

2.2 EXPERIMENTAL STUDIES ON CONVENTIONAL SLAB-COLUMN CONNECTIONS 2-10

2.2.1 Interior Connections 2-10

2.2.2 Edge connections 2-22

2.2.3 Comer Connections 2-31

2.2.4 The Effect of Boundary Conditions 2-36

2.2.5 Tests on Multi-Panel Specimens 2-38

2.2.6 Tests of Connections with Shear Reinforcement 2-46

2.3 ENGINEERING MODELS FOR DESIGN OF SLAB-COLUMN CONNECTIONS 2-53

2.3.1 Models Based on Elastic Plate Theory 2-53

2.3.2 Beam Analogy Methods 2-55


vii

2.3.3 Models Based on Elasticity 2-58

2.3.4 Truss Model 2-62

2.3.5 Improved Rotational Model 2-66

2.3.6 Yield Line Analysis 2-67

2.4 PRESTRESSED CONCRETE SLAB-COLUMN CONNECTIONS 2-69

2.5 FINITE ELEMENT ANALYSIS OF SLAB-COLUMN CONNECTIONS 2-78

2.6 CONCLUSIONS 2-89

CHAPTER 3- FINITE ELEMENT MODELLING OF PUNCHING


PROBLEMS 3-1

3.1 MATERIAL MODELLING AND SOLUTIONPROCEDURE 3-1

3.1.1 Introduction 3-1

3.1.2 General Constitutive-Relationships 3-1

3.1.3 Concrete in Compression 3-3

3.1.4 Cracked Concrete 3-7

3.1.5 Reinforcing steel 3-14

3.1.6 Solution Procedure 3-14

3.2 SENSITIVITY OF MESH GRADING AND INTEGRATION 3-16

3.2.1 Introduction 3-16

3.2.2 Pullout of an Embedded Plate 3-16

3.2.3 Upper bound solution 3-26

3.3 INFLUENCE OF FLExURAL REINFORCEMENT ON PuNCHING STRENGTH 3-31

3.3.1 Introduction 3-31

3.3.2 Effect of Negative Moment Reinforcement on Punching Strength 3-32

3.3.3 Finite Element Results 3-35


viii

3.4 SENSITIVITY ANALYSESOFFINITEELEMENTMODELLINGPARAMETERS 3-40

3.4.1 Introduction 3-40

3.4.2 Sensitivity to the Concrete Tensile Strain at Zero Stress, Eu. 3-41

3.4.3 Sensitivity to the Shear Retention parameter, B. 3-43

3.4.4 Sensitivity to the Friction Angle, <j>. 3-44

3.4.5 Sensitivity to the Dilatancy Angle, 'If. 3-47

3.4.6 Sensitivity to the Compression Softening Slope, Ecf 3-49

3.5 CONCLUSIONS 3-49

CHAPTER 4- VERIFICATION OF THE FINITE ELEMENT MODEL 4-1

4.1 INTRODUCTION 4-1

4.2 REINFORCED CONCRETE BEAM OF MANSUR AND RANGAN (1978) 4-1

4.3 REINFORCED CONCRETE CONNECTION OF RANGAN AND LIM (1992) 4-11

4.3.1 Finite Element Modelling 4-11

4.3.2 Sensitivity Analyses 4-21

4.4 PRESTRESSED CONCRETE CONNECTIONS OF FOUTCH ET AL. ( 1990) 4-27

4.5 PRESTRESSED CONCRETE CONNECTION OF BURNS AND HEMAKOM (1977) 4-39

4.6 CONCLUSIONS 4-46

CHAPTER 5- PULLOUT TESTS 5-1

5.1 INTRODUCTION 5-1

5.2 EXPERIMENTAL PROGRAM AND SPECIMENS 5-4

5.3 TEST SET-UP AND INSTRUMENTATION 5-10

5.4 TESTS REsULTS AND DISCUSSIONS 5-13


ix

5.5 FINITE ELEMENT ANALYSES 5-22

5.6 CONCLUSIONS 5-26

CHAPTER 6- DESIGN AND CONSTRUCTION OF TEST SPECIMENS 6-1

6.1 INTRODUCTION 6-1

6.2 DESIGN PHILoSOPHY 6-1

6.3 ARRANGEMENT OF REINFORCEMENT 6-7

6.4 CONSTRUCTION OF FORMWORK, TEST SPECIMENS AND AsSEMBLY 6-14

6.5 BOUNDARY SUPPORTS 6-23

6.6 PRESTRESSING 6-26

6.7 LoADINGARRANGEMENTS 6-30

6.7.1 Specimens S1, S2 and S3 6-30

6. 7.2 Specimens S4 6-36

6.8 MATERIALPROPERTIES 6-36

6.8.1 Bonded Reinforcement 6-36

6.8.2 Prestressing wire 6-39

6.8.3 Ducts and Accessories 6-41

6. 8.4 Concrete 6-41

6.9 INSTRUMENTATION 6-44

6. 9.1 General 6-44

6.9.2 Measurement of Loads and Reactions 6-44

6. 9.3 Displacements 6-44

6.9.4 Strain Measurements 6-47

6.10 TESTINGPROCEDURE 6-47


X

CHAPTER 7- TEST RESULTS AND COMPARISON WITH CODE


MODELS 7-1

7.1 INTRODUCTION 7-1

7.2 FAILURELoADSANDREACTIONS 7-1

7.3 DEFLECTIONS 7-6

7.4 INDIVIDUAL TEsT REsULTS 7-11

7.4.1 Specimen S1 7-11

7.4.2 Specimen S2 7-14

7.4.3 Specimen S3 7-20

7.4.4 Specimen S4 7-27

7.5 DISCUSSIONOFRESULTS 7-27

7.5.1 Introduction 7-27

7.5.2 General Discussion 7-31

7.5.3 Comparison of Results with Code Models 7-36

7.5.4 Further analysis 7-56

7.6 CONCLUSIONS 7-65

CHAPTER 8- FINITE ELEMENT ANALYSES OF TEST SPECIMENS 8-1

8.1 INTRODUCTION 8-1

8.2 FINITE ELEMENT MODELLING 8-2

8.3 RESULTSOFFINITEELEMENTMODELS 8-7

8.3.1 Results of Model S1 8-7

8.3.2 Results of Model S2 8-18

8.3.3 Results of Model S3 8-25

8.3.4 Results of Model S4 8-32


xi

8.4 THE EFFECT OF PRESTRESS ON EDGE COLUMN CONNECTIONS 8-39

8.5 THE EFFECT OF PRESTRESS ON CORNER COLUMN CONNECTIONS 8-43

8.6 BEHAVIOUR OF SLAB-COLUMN CONNECTIONS WITH BONDED TENDONS 8-44

8.6.1 Results of Model S2-B 8-51

8.6.2 Results of Model S3-B 8-56

8.6.3 Results of Model S4-B 8-56

8.7 DISCUSSIONS AND CONCLUSIONS 8-56

CHAPTER 9- CONCLUSIONS 9-1

9.1 INTRODUCTION 9-1

9.2 CONCLUDINGREMARKS 9-1

9.3 REcOMMENDATIONSFORFU'ruRERESEARCH 9-6

9.3.1 Numerical Modelling 9-6

9.3.2 Experimental Investigations 9-7

9.3.3 Development of Design Models 9-8

REFERENCES R-1

APPENDIX A- DESIGN OF PROTOTYPE AND TEST SLABS A-1

A1 STRUCTURAL SYSTEMS A-1

A2 LoADINGS AND MATERIAL PROPERTIES A-1

A3 ANALYSIS AND DESIGN A-3

APPENDIX B- RELAXATION TEST AND PRESTRESS LOSSES B-1


xii

APPENDIX C- HORIZONTAL REACTION FOR SPECIMEN S1 C-1

APPENDIX D- CODE MODELS D-1

D.1 INTRODUCTION D-1

D.2 ACI 318-1999 D-1

D.3 AS 3600-1994 D-10

D.4 BS 8110: PART 1: 1997 D-12

D.5 CEB-FIP MODEL CODE 1990 D-15

D.6 EUROCODE 2-1992 D-19


NOMENCLATURE

A area of the critical section

a linear dimension of the critical shear perimeter measured parallel to the


direction of Mv *

Ast area of reinforcing steel

b width of the critical section

bo perimeter of the critical section

c cohesion

a parameter

equivalent cohesion

D overall depth of the slab

elasticity matrix

d effective depth

do diameter of the rigid plate/disk

d1 maximum diameter of the failure surface

de depth of a square column

De stiffness matrix corresponding to elastic strain

D~r cracking stiffness matrix for n-th crack

ncr overall cracking stiffness matrix

E, E 1 tangent moduli of elasticity for the ascending and descending branches


of concrete stress-strain relationship for tension

modulus of elasticity of concrete

crack traction vector in local coordinates

modulus of elasticity of the column of the model structure

tangent modulus of the steel stress-strain diagram in the post elastic


region

modulus of elasticity of steel


xiv

ex, ey distances of moment vectors in y and x directions

er local crack strain vector

f potential function

1; characteristic cylinder strength of the concrete.

!em cylinder strength of the concrete in MPa

fcp compressive strength of the in place concrete

feu characteristic cube strength of the concrete

/pe effective compressive stress in the concrete in each direction

/sy yield stress of the reinforcing bars

:ft tensile strength of the concrete

/yd design yield stress of the reinforcement

g plastic potential function

Gt fracture energy

h equivalent length of the mesh

overall depth

depth of conical part in the failure surface

first invariant of the stress tensor

moment of inertia of the column of the model structure

moment of inertia of the column of the test specimen

second invariant of the deviatoric stress tensor

properties of the critical section analogous to polar moment of inertia

K spring stiffness

material parameters for Nielsen et al. (1978) model

length of the column of the model structure

length of the column of the test specimen

direction cosines for the strain normal to the crack plane (n-axis)
relative to the global co-ordinate system
XV

m moment per unit length of the slab

unbalanced moment calculated at the centre of the column

ultimate unbalanced moment

Muo ultimate moment corresponding to zero shear

ultimate moment Mu calculated at the centroid of the critical section in


the x and y directions

unbalanced bending moment in the direction being considered

N transformation matrix

i-th component of transformation matrix N

N transformation matrix for n cracks


p projection matrix

Pftex flexural load capacity

shear force

failure load in pullout of the plate embedded in the concrete

q projection vector

r deviatoric stress vector

one of the component of Rx

reactions in x-, y- and z-directions

SSR shear stud reinforcement


{' traction vector across the crack

i-th traction vector

mode I normal traction

tscr ' tcr


t
mode II and mode ill shear tractions

u perimeter of the critical section

UJ perimeter

control perimeter

shear force transferred to the column centre at the collapse limit state
xvi

V, Vc shear stresses

nominal shear capacity

effective design shear force

vertical component of the prestress force passing through the column

total design shear force

design shear force transferred to the column

the maximum factored shear stress

ultimate shear force

ultimate shear strength for zero moment

w intensity of load

a parameter of the control perimeter

X length of the side of the perimeter parallel to the axis of bending

a threshold angle

tensioning softening parameters

shear retention factor

ratio of the largest overall dimension of the support Y to the overall


dimension X at right angles to Y

ratio of long side dimension to the short side dimension of the column
cracking strain of concrete

global crack strain

inelastic strain rate vector

i-th increment of global crack strain ecr

normal strain for the mode I crack

total global strain vector

If' global crack strain vector

elastic component of total strain vector e

inelastic (plastic) component of total strain vector e


xvii

strain in steel corresponding to the stress offryi

rate of plastic strain

{e} strain vector

q, angle of friction

strength reduction factor

initial angle of friction

partial safety factor

fraction of unbalanced moment transferred by flexure

'Ym partial safety factor for strength of materials

fraction of unbalanced moment transferred by shear

r~ shear strains for the mode IT cracks

,cr shear strains for the mode ill cracks


1 nt

equivalent plastic strain, state variable

rate of state variable

i·J a proportionality constant

f.1 Poisson's ratio

v effectiveness factor

(} angle between deviatoric stress vector and first principal stress


projected on the deviatoric plane or inclination of punching failure
surface with respect to the horizontal plane

p steel reinforcement ratio

ratios of tension steel in x and y directions

average intensity of effective prestress in concrete in MPa

average prestress without losses

global stress vector

a; (i =1, 2, 3) principal stresses


xviii

{ cr}, cr stress vector

O'cp average intensity of effective prestress in concrete in MPa

'rsd shear stress

length along the hydrostatic axis

dilatancy angle
CHAPTER 1 - INTRODUCTION

1.1 General

A flat plate floor system involves the direct transmission of forces from the slab to the

columns without the use of beams as shown in Figure 1.1. Reinforced and prestressed

concrete structures with flat plate construction are popular structural systems. This is

because of their architectural versatility, easier formwork construction and easier steel

and prestressing ducting arrangements leading to speedier construction, smaller

building heights and the ease of placement of services which ultimately leads to lower

overall construction costs. These structural systems are widely used in offices and car

park construction. Since the 1960's, increasing demand for longer spans has lead to

prestressing the slabs to overcome problems such as serviceability and inadequate

stiffness. The use of prestressing can effectively increase the stiffness of the member,

control cracking and deflections and may improve the punching resistance of the

slab-column connections, over that of conventionally reinforced connections.

Although considerable attention has been given to studying flat plate and flat slab

structural systems, uncertainties remain in understanding their structural behaviour

particularly for prestressed concrete systems. The three-dimensional stress field

developed in flat plates gives rise to complex load distributions in the vicinity of the

connection. High bending moments together with high shear forces may be developed

at these connections leading to a punching shear failure.

The design of slab-column connections is generally governed by shear strength

criteria. Figure 1.2 shows different types of shear problems in flat plate structures. The
1-2

Edge Corner
column column

Figure 1.1- Typical flat plate structural system.

..,X

Shear

D
ol
I
x~~D X
1Shear and moment
Or only shear transfer

D D D

beam shear failure

D D D

Note: x is the distance of critical section from the column face.


It is the function of the effective depth of the slab and
its value and the actual shape of critical sections
depend on the code considered.

Figure 1.2 - Types of shear problems in flat plate structures.


1-3

extent of the influence of the prestress in preventing punching failure remains unclear

especially for the connections with edge and comer columns. Limited experimental

and theoretical studies have been undertaken on edge column connections of post-

tensioned flat plate-column connections. No studies have been undertaken on the

behaviour o~ prestressed comer column _connections.

Different structural design codes use different models to incorporate the effect of

prestress in the design of slab-column connections. For example the American code

ACI 318 (1999) neglects the effect of prestress if the slab projection beyond the edge

column does not exceed four times the slab depth while CEB-FIP Model code (1993)

suggests a reference to specialist literature. The diversities in the design models used

by various codes highlight the need for further research in the field.

The research described in this thesis comprises of three main parts. Part one

investigates the appropriateness of using three dimensional finite element modelling

for the analyses of the punching type problems. In part two an experimental

investigation is undertaken on flat plate-column connections of edge and comer

columns. Finally in part three the finite element model is used to extend the findings of

the experimental investigation.

In the experimental programme four 1/3-scale flat slab-column connections were

tested, three edge column connections and one comer column connection. The

specimens are sized to be representative of a full scale structure and the slabs are

uniformly loaded to failure. Two of the edge column specimens and the comer column

specimen are prestressed with unbonded tendons, the aim being to study the influence
1-4

of prestressing on the behaviour of these connections. A series of plain concrete

pullout tests are also undertaken to provided data for the calibration of the finite

element model.

In part three ofthis thesis the fmite element program DIANA (1997) is used to extend

the range of the research with modelling of various levels of prestress on comer and

edge connections. Lastly the numerical model is used to determine the likely influence

of bonding of the tendons on the strength of prestressed connections.

1.2 Objectives

The objectives of this study are:

• To investigate the appropriateness of using 3D brick elements for the

modelling of slab-column connections failing in punching shear and reviewing

issues such as mesh sensitivity and material modelling.

• To design and undertake an experimental programme for testing to failure of

edge and comer column-slab connections that are punching shear critical.

• To study the effect of prestressing on the performance of slab-column

connections failing in punching shear.

1.3 Scope

This research is focussed on the punching shear behaviour of prestressed concrete

slab-column connections. Experimental and numerical models are used for the

investigation of edge and comer column connections under statically applied uniform

vertical loading. The outcomes of this research are applicable to the ultimate strength
1-5

design of flat plate slab-column connections failing in punching shear under gravity

load.

The effects of overhangs, holes and shear reinforcement are not dealt with in this

research. Also time dependent effects like creep and shrinkage, thermal effects,

dynamic effects, rate of application of load and size effects are beyond the scope of

this research.

1.4 Organisation of Thesis

This thesis is presented in nine chapters. In Chapter 2, literature on the behaviour of

slab-column connections is reviewed.

In Chapter 3, finite element modelling of punching type problems is investigated. The

effect of mesh grading and the sensitivity of different modelling parameters needed for

the fmite element modelling of punching type problems are studied.

In Chapter 4 the fmite element model described in Chapter 3 is verified using a

number of results from experimental studies. A range of studies was selected for a

wide variety of flexural, torsional and shear failures. The effects of boundary

conditions in the behaviour of these connections are also reviewed.

In Chapter 5 results of plain concrete pullout specimens are given. The tests give

further verification on data for the fmite element model.

Chapters 6 gives details of the main experimental programme and the results are given

in Chapter 7. Laboratory tests on four slab-column connections are described with


1-6

three of the specimens are edge column connections and one specimen is a comer

column connection. The results of the tests presented are loads, deflections, reactions

and the strains measured at key locations of a number of steel reinforcing bars. The

crack patterns and the failure mechanisms are discussed and the results are compared

with code models.

In Chapter 8, the fmite element models developed in Chapters 3 and 4 are used to

analyse of the experimental slab-column connections described in Chapter 6. The

fmite element modelling is used to gain further insight into the behaviour of the tested

slabs and to extend the range of observations.

Conclusions are given in Chapter 9.

The thesis has four appendices. In Appendix A the design details for prototype slab

used to size the model specimen are given. In Appendix B the prestressing losses in

the test slabs are calculated and the results of relaxation tests on the strand are given.

Appendix C provides the basis for the correction to the horizontal reaction measured

in specimen S 1 is presented. In Appendix D details of various code models are given

for the design of flat plate-column connections.


CHAPTER 2- REVIEW OF LITERATURE

2.1 Introduction

The earliest flat slabs are reported to have been built by Turner in the United States of

America in 1906. In Europe, Robert Ma.illart pioneered the construction of flat slabs

(Regan, 1981, Furst and Marti, 1997). Since this time numerous flat slabs have been

constructed and tested.

In flat slab concrete structures the shear force and a part of the unbalanced moment are

transferred from the slab to the columns at the slab-column interface. Transfer of shear

and moment for a typical internal slab-column connection is shown in Figure 2.1. Shear

stresses are developed in the critical section because of the shear force and the

unbalanced moment on each side of the connection.

For the case of a slab in one-way bending, the slab behaves as a beam. After the

formation of diagonal tension cracks in the vicinity of the critical section of the slab

(around the perimeter of the loaded area) the shear force is carried by a combination of

shear across the compression zone, aggregate interlock and dowel action. However, for

a slab in two-way bending, the nominal ultimate shear stress that can be developed at

the assumed critical section is much higher than in a beam or slab in one-way bending.

This increase in punching shear strength of two-way slabs is attributed to the

three-dimensional nature of the failure mechanism.


2-2

/
Vu~~
/ Mu1
Forces to be transmitted
by slab-column connection

Shear stresses on
~~!
~ andMux1
critical section due

Note:
Mu1 =Mux1 +Mux2
Mux1 and Mux2 are parts of
total unbalanced moment M
transmitted by shear and
flexure, respectively

Figure 2.1 - Typical load transmission in flat plate system.


(shown for an interior connection).

Under the action of load on a slab in two-way bending, the first crack is usually formed

by negative bending in the radial direction (van den Beukel, 1976, Park and Gamble,

2000). This crack is approximately circular and forms tangential around the perimeter

of the loaded area (Figure 2.2). Cracks radiating from the column are formed due to

negative bending moments in the tangential direction. The radial moment, which is the

moment causing the circumferential cracks, decreases rapidly away from the loaded

portion and, thus, to propagate the tangential cracks around the loaded area some

distance further from the first cracks, the load must be increased significantly. The

diagonal tension cracks in the slab normally originate near the mid-depth and are

analogous to web-shear cracks in beams. The opening of the diagonal tension


2-3

~-·l:--d/21
o~J
d/2-l t~l
= d slab effective depth

Vu

Figure 2.2 - Assumed critical section and failure mode for punching shear
failure of reinforced concrete slab-columns with concentrically
loaded columns (Park and Gamble, 2000).

cracks are dependent on the stiffness of the slab surroundlng the cracked region and on

the slabs ability to maintain the shear transfer (at high loads) by aggregate interlock and

dowel action. When punching shear failure occurs it is likely to be accompanied by

yielding of the negative moment slab reinforcement in the vicinity of the punch. The

diagonal tension (web shear cracks) are usually formed at about one-half of the load

required for punching shear failure. Equilibrium of the three-dimensional system can be

written in the generalised form for the horizontal forces acting on the section near the

critical diagonal tension crack (Figure 2.3). For the slab-column connection,
2-4

l:FHor: T1+T2 = C1+C2


C1 = T1 NOT REQUIRED

Figure 2.3 - Horizontal forces at the faces of diagonal tension cracks of a


reinforced concrete slab-column connection (ASCE-ACI
Committee 426 - 1974).

no unique value for the concrete compressive force C1 (refer Figure 2.3) at one face of

the column is obtained due to the possibility of redistribution to the surrounding

concrete. It may be thought that shear strength can be increased by providing an

increase in the slab tension reinforcement leading to a greater depth of concrete in

compression and increased dowel action. However, an increase in steel content in the

connection region has a limited effect on the shear strength (Park and Gamble, 2000).

The top cover concrete starts to split from the slab after a threshold value resulting in a

reduction in the dowel action and tensile force resisted by the steel.

The critical sections of the slab for both negative moment and punching shear are at or

close to the perimeter of the loaded area resulting in moment-shear interaction. The

characteristics of the failure mode and of the load-deflection curve for the slabs are

dependent on the reinforcement content. Based on small-scale tests by Dragosvic and


2-5

van Den (1974) (Figure 2.4) ductile flexural failure and brittle punching shear failure

are observed for slabs with both small and large slab steel ratios (refer Figure 2.5).

When a small amount of flexural steel is used punching shear failure was often observed

as a secondary failure after the slab reached its flexural strength limit.

Test data summarised in the ASCE-ACI Committee 426 (1974) report show that the

shear strength is independent of the ratio of tension reinforcement in the slab, p, and on

the yield strength of the steel, fsy· However, concentrations of top and bottom

reinforcement are recommended as the top steel improves the flexural behaviour of slab

in the service load range and the bottom steel acts as a suspension net holding the slab

to the column. Kinnunen and Nylander (1960) and Zaghlool and de Paiva (1973)

observed some increase in punching shear strength with an increase in flexural top

reinforcement but it is small and generally ignored. Once the cover to the top steel is

lost due to spalling, the top steel tears from the slab providing little to no shear carrying

capacity (Islam and Park, 1976, Pan and Moehle,1992).

Edge, comer and internal slab-column connections with non-uniform load, or unequal

spans on each side of the connection (see Figure 2.6) require the transfer of both shear

and unbalanced moment. Also any horizontal loading due to earthquake, wind, or other

sources give rise to a substantial unbalanced moment to be transmitted between


2-6

30mm thick, 475mm


diameter slab

Figure 2.4 - Outline of the specimens without punching shear reinforcement


used by Dragosvic and van Den (1974).

Experimental
Column
load at .
- - -Idealized
,___!_.....!.._.!...,__.:__. -
failure
vtest
• /
•/
Punching shear
failure
f
• /Flexure V
• / failure
/
Slab steel ratio

Column
load

Column
load

Central deflection Central deflection


Flexural failure with low Shear failure with high steel ratio

Figure 2.5 - Effect of slab reinforcement ratio on shear strength, ductility and
load deflection curves of slab-column connections with
concentrically loaded columns (Dragosvic and van Den 1974, Park
and Gamble, 2000).
2-7

Figure 2.6 - Typical slab-column connections and possible loadings.


2-8

the slab and the columns. In such cases, the transfer of unbalanced bending moment

causes the distribution of shear in the slab around the column to become non-uniform,

resulting in a reduction of shear capacity of the connection. The shear force and

unbalanced moment are transferred by combined bending, torsion, and shear at the

faces of the critical section in the slab as shown in Figure 2. 7. When the shear stress

developed at any section of the slab exceeds the shear capacity, the slab fails in diagonal

tension at the side where the high shear stress is developed resulting in a punching

failure. The location of the critical section and the proportions of the moments

transferred from shear and flexure depend on the model being considered. The different

code models are discussed in Appendix D of this thesis.

Flat slabs are susceptible to failure due to punching shear under the large early age

loads that may be imposed during construction. Generally these slabs are constructed

using proprietary flying or table forms and as a result the construction is relatively

rapid. Loads coming on the supporting slabs during construction can equal or exceed

the supporting slab's service load and load is transferred to the ground through the

immature slab and slab-column connections. In some cases slabs have been reported to

have failed during construction (Sbarounis, 1984, Lew et al., 1982).

Multistorey flat plate buildings have limited use as seismic-resistant structures without

the provision of other moment resisting and stiffening elements such as frames and

slab-column connections of flat plate structures are likely to be subjected to repeated

cyclic bending moments causing failure of the slab due to degradation of the shear
2-9

C2

c
Load transfer detail

¢
A

Critical sections

Note: MARx and VAB are x direction moment and shear for of side AB, similarly for the
other sides.

Figure 2. 7 - Typical load transfer mechanism (shown for a edge column


connection with unbalanced moments in the x andy directions).

capacity of the slab around the column. Such failure is highly brittle in nature and, as

such, the use of the flat plate construction is not recommended in high seismic regions

(ACI-ASCE Committee 352, 1989). Shear reinforcement have been found effective in

increasing the strength and in changing the failure pattern from brittle to ductile. The

amount of increase in strength and the extent of change in the failure mechanism

depend on the amount, type, and arrangement of the shear reinforcement.


2-10

2.2 Experimental Studies on Conventional Slab-Column Connections

Numerous experimental investigations have been conducted to study load transfer in

slab-column connections. Early tests reported (Elstner and Hognestad, 1956, Kinnunen

and Nylander, 1960, Moe, 1961, Dragosvic and van Den, 1974) were on simple

isolated specimens with the specimens extending close to the first line (nominal) line of

contraflexure (refer Figure 2.8). The effect of specimen size and boundary conditions

were not considered. Isolated small specimens were tested and observations were on

the basis of the results obtained. The design provisions in ACI-318 (1999) are based

largely on the results of empirical studies made using such models. It was observed by

Rankin and Long (1988) that these specimens do not represent the real structure well.

After the influence of specimen size and boundary conditions were better understood,

experimental models were improved to incorporate these effects into the test. ·In later

studies (Rangan and Hall, 1983, Robertson and Durrani, 1992 and Durrani et al., 1995)

statically indeterminate specimens were used for the investigations and sized to

incorporate the laws of similitude to simulate the behaviour of real sized slab-column

connections. In the following sections the literature on the tests conducted on various

types of slab-column connections are briefly reviewed.

2.2.1 Interior Connections

Elstner and Hognestad (1956) published a report on their tests on thirty-nine,

1.83 metre square, 152 mm thick slabs. The slabs were loaded through centrally located

column stubs. The variables considered were concrete strength, amount and

distribution of tension reinforcement, amount of compression reinforcement, the size


2-11

0
Second line of
First line of ----::----;-,...-----------1 contraflexure
contraflexure 1I
I
I
I
I
1
.----+- - - - - - 1 - - - Portionscovered in more
·..
Jl
· .1 I
recent tests
f, ' ,.,. ' t
+--------~-+-~----
Portions covered ------+--..1 · - l
in early tests __J J

:
-·1I
'I

1
_________________ j
:

Second line of

l~lt<--
UFkst Une of
contraflexure

Section X-X with contraflexure lines

Figure 2.8 - Slab-column connections showing the regions used in different


specimens tested in the past and present.

of the column support and loading conditions, the amount and the position of shear

reinforcement and loading eccentricity. In most of the cases punching failure was

reported after initial yielding of the reinforcement in the vicinity of columns.

Compression reinforcement and the eccentricity up to one-half of the column depth

were found to have no effect on the ultimate shear strength of the slab subjected to a

concentrated load through the column. However, the results showing that load

eccentricity has no effect on the punching shear strength was later contradicted in tests
2-12

undertaken by Moe (1961), Rangan and Hall (1983a, 1983b), Rangan (1990), and

Falamaki and Loo (1992) which showed clearly that the moment caused by load

eccentricity is a significant factor in the shear strength of slab-column connections.

For the connections without shear reinforcement Elstner and Hognestad (1956)

proposed the empirical relationship

V2 _ V 333 0.046
-+-- (2.1)
+'- Lbd+'
Jc 8 'Jc ~~ l/Jo

where b is the circumference of the loaded area (in inches); l/Jo is the ratio of the shear

capacity to flexural capacity; f~ is the cylinder strength (in psi); dis the effective depth

of the slab in inches; Vis the shear force (in pound); and v2 is ultimate shearing

strength (in psi).

Kinnunen and Nylander (1960) conducted a series of tests on 150 mm thick circular

slabs of 1840 mm diameter with flexural reinforcement but without shear

reinforcement. The slabs were loaded via a circular stub column located at the centre of

the specimens. The test variables were the type and quantity of flexural steel and the

ratio of the stub column diameter to slab thickness. They found an increase in the

punching capacity with an increase in the quantity of negative moment flexural

reinforcement. From their observations Kinnunen and Nylander developed a model to

determine the punching load-carrying capacity of slabs supported on columns.

Moe (1961) undertook an investigation to study the basic mechanisms of shear failure

by testing twelve 1.83 metre square slabs with thickness of 152 mm. The specimens
2-13

were simply supported allowing the comers to lift freely. The loads were applied at

different eccentricities through a centrally located square column stub. Holes were

provided in the vicinity of the column of the specimens to enable observation of crack

formation. During the test, cracks were observed at loads as low as 50 percent of the

ultimate load. Moe found the shear strength to be dependent on the flexure strength of

the slab. Concentration of flexural reinforcement in the narrow bands over column had

a significant influence on increasing the yield load of the slabs. However, the punching

resistance was not affected by the arrangement of the flexural reinforcement. Based on

the test results of 43 slab-column specimens, Moe developed a method for an ultimate

strength analysis for moment transfer in the slab-column connections under combined

shear and unbalanced moment. Moe assumed the critical section for shear to be at the

face of the column.

Moe found that the punching capacity is reduced with increasing eccentricities. Based

on the results of his tests, Moe concluded that approximately one-third of the total

unbalanced moment was transferred to the column by shear stresses and is independent

of the quantity of the flexural reinforcement. Moe also concluded that the shear

strength was proportional to .[1:, where J; is the compressive strength of the

concrete.

Kinnunen (1963) improved the mechanical model of Kinnunen and Nylander (1960) to

make the model applicable to slabs with two-way reinforcement but without polar

symmetry. This modification was based on the further assumption that the

reinforcement cutting across the shear crack carries a part of the shearing force. The
2-14

modified model is as shown in Figure 2.9. In the modified mechanical model shown in

Figure 2.9, r is the radius, h effective depth, y distance from the soffit of the slab to the

root of the shear crack at failure, z distance from the neutral surface to the top surface

of the slab in the middle part of a panel in a flat slab, P the column load, R1, R2, R3 and

R4 the force resultants from the reinforcement and T the oblique compression force in

the imaginary conical shell.

The Kinnunen (1963) model was found to be in good agreement with experimental

results for the slabs with ring and radial reinforcement. However, for the slab with two

way reinforcement the average calculated punching load and average angle of rotation

of the slab portion outside the crack at failure by punching were smaller than the

observed values by about 20 percent and 40 percent, respectively. These discrepancies

were attributed to dowel action. In the study of Kinnunen the effect of reinforcement

cutting across the shear crack was taken into consideration in transmitting a part of the

column load to the slab. The resultant force consisted of two components (i) membrane

force and (ii) a dowel force. This revised model gave a better agreement with the

experimental results for the slabs with two-way reinforcement although it is complex.

Hanson and Hanson (1968) conducted seventeen tests on slab-column connections

including one exterior column connection. Their main aim was to study the behaviour

of moment transfer at interior columns. The variables used in their investigation were

the loading arrangement and the location of voids adjacent to the column. Hanson and

Hanson used their test results to evaluate the design methods proposed by Di Stasio
2-15

p~.f
2n

.... .... ....

.. .... ... ..
. . .. ... . .
~

.. . .. 1.

Note: R2 = R3 for the connections with shear reinforcement and R3 = 0 for the
connections without shear reinforcement.

Figure 2.9 - Modified mechanical model of Kinnunen and Nylander (1960)


(Kinnunen, 1963).
2-16

and Van Buren (1960), Moe (1961), the ACI-ASCE Committee 326 (1962)

recommendations and those given in ACI 318 (1963). A modified version of the model

proposed by ACI-ASCE committee 326 (1962) was found to estimate the strength of

slab-column connections well.

Stamenkovic and Chapman (1974) tested internal, edge and comer columns with

varying axial load and moment applied to the loading columns. In all 52 tests were

undertaken on 914 mm square by 76 mm thick slabs. Two column sizes were used,

127 mm square columns and 152x76 mm rectangular column with the column

extending above and below the slab. The parameters considered were the type of

loading (vertical or horizontal and combined vertical and horizontal), the column

location and the column shape. The slabs were supported along the four edges using tie

rods spaced at 165 mm centres.

Stamenkovic and Chapman developed a set of equations based on Moe's (1961)

method to calculate the strength at the column head under axial force and bending.

Interaction curves were presented for all types of connections under combined vertical
'
and horizontal loading. For internal connections Stamenk.ovic and Chapman proposed a

straight line interaction formula as shown in Figure 2.10 and given by

V M
-+-=1 (2.2)
Vu Mu

where V and Mare the failure shear force and moment, respectively, Vu is the shear

capacity for zero moment and Mu the flexural capacity for zero shear.
2-17

0 M Mu

Figure 2.10- Moment-shear interaction diagram for internal column


connections (Stamenkovic and Chapman, 1974).

Dragosavic and Van Den (1974) conducted tests on 24 small scale slab-column

connections representing concentric interior slab-column connections and a further 24

representing eccentrically loaded slab-column connections including edge and comer

column connections (Figure 2.4). The test results showed that the punching shear

strengths are not significantly affected by the reinforcement content.

Van Den (1976) proposed an analytical model for predicting the punching shear

strength of different slab-column connections. The effects of the eccentricity in the edge

and comer column were included by introducing an eccentricity factor based on the

geometrical properties of the connection. The model was compared with the test results

ofDragosavic and van Den (1974).

Hawkins et al. (1989) tested 36 interior slab-column connection specimens. Slab

thickness, concrete strength, concrete type, column rectangularity, stirrups in the


2-18

integral beam and slab reinforcement pattern were the main variables of the test

programme. The 1/3 scale specimens consisted of 2.1 m wide square slabs supported

on a central 305 mm square column. Hawkins et al. found that concrete splitting and

dowel effects for the slab reinforcement passing through the column reduced the

moment transfer capacity. Also shown was that the rotational stiffness of connections

decreased continuously with the increase in load and the equivalent column stiffness

procedure of ACI 318-1983 was satisfactory only in the linear range. However, after

the propagation of cracks due to shear and torsion, the rotational stiffness was

observed to be directly proportional to the amount of reinforcement within a width of

two times the slab thickness on either side of the column. The ACI 318-1983 model

gave reasonable results for the strength of the connections provided that flexural

reinforcement within a width of 1.5 times the slab thickness on either side of the

column did not exceed 0. 7 percent. The model yielded conservative results when the

reinforcement ratios exceeded 0. 7 percent.

Based on the Kinnunen-Nylander model, Broms (1990a) developed a method to

calculate the shear strength and deflection for slab-column connections. Broms

hypothesised that the failure occurs when the concrete in compression near the column

is distressed by either a high circumferential strain or a high radial stress. The

compression zone of the slab near the column face was considered to be under biaxial

compression. Considering this biaxial compression of the concrete, punching was

assumed to occur when the stress in the conical shell reached the value 10 percent

higher than the uniaxial strength of the concrete. The critical strain of 800 J..l£ was

assumed as the strain for developing critical macro-cracks parallel to the compression
2-19

direction. These cracks enabled the shear crack to propagate through the compression

zone of the slab. The method was compared with data from a number of studies and

was shown to give a reasonable correlation with the test data for interior flat

plate-column connections under different loadings.

Based on the tests of eight isolated internal square slab-column connections with

different flexural as well as shear reinforcement, Broms (1990b) concluded that with

the use of a combination of bent bars and multiple U-stirrups a good ductile

performance can be achieved in slab-column connections. However, ordinary shear

reinforcement in the form of stirrups enclosing only the tension flexural reinforcement is

ineffective in improving ductility. Broms tests showed that openings in the slab inside

the critical perimeter do not effect on the strength or stiffness of the slab.

Alexander (1990) tested 12 isolated interior flat plate-column connections. The test

specimens consisted of a 155 mm thick by 2750 mm square plate slab with

200 mm x 200 mm x 200 mm column stubs above and below the plate. Four of the

specimens were reinforced with steel fibres. The main parameters considered in the test

were the boundary restraints, fibre content, clear cover to the reinforcing steel and the

spacing of the reinforcement through the column. From the results of the tests,

Alexander observed that the truss model proposed by Alexander and Simmonds (1986)

required modification to describe the behaviour of a slab-column connection with the

modified model referred to as the bond model. The bond model was compared against

116 test results and was shown to correlate well with the experimental data. With the
2-20

inclusion of steel fibres in the concrete mix the cracks in the specimen were reduced

and both strength and ductility were improved.

Pan and Moehle (1992) tested four slab-column sub-assemblages under combined

gravity and lateral loads. Substantial reduction in strength, ductility and drift capacity

was observed under bi-directional loading compared to uni-directional loading. For

small gravity loads, the ductility and load carrying capacity were significantly improved

when lateral loads were applied. However, with the increase in gravity load significant

reductions in the shear capacity and the lateral deformation performance were

observed. Bottom reinforcement in the slab, which passed through the column,

improved the post punching behaviour of the joint. Finally, the post punching shear

strength was found to be dependent on the detailing of the top and bottom bars in the

slab. Similar observations are given by the ASCE-ACI Committee 426 Report (1974),

Park and Gamble (2000), Durrani et al. (1995) and Gardner and Shao (1996).

Chana and Desai (1993) tested a total of 14 specimens of two different geometry (see

Figure 2.11) to study the effect shear reinforcement on punching of the interior

slab-column column connections. Also investigated was membrane effects. The

specimens dimensions tested were 3 m x 3 m x 0.228 m and 9 m x 9 m x 0.250m. All

the specimens failed in brittle punching shear mode. The maximum crack widths before

failure were reported to be approximately 0.3 mm and 0.15 mm for the small and large

specimens, respectively, indicating the beneficial effect of compressive membrane

action in controlling the crack widths. Based on the test results a number of design

recommendations were made.


2-21

t-0.4L-j r-o.4L---I
~ !
¥ I
f
I. I
1.5L
f
.I
I

type I type II

Figure 2.11- Specimens outline (Chana and Desai, 1993).

Farhey et al. (1995) studied the effects of attaching steel plates in the repair of the

damaged slab-column connections. The plates were bolted and glued to the slab. They

concluded that the method is effective in resisting both lateral and vertical loads. The

response of the repaired connection was found to be better than the original.

Chow and Selna (1995) developed a method for modelling the seismic response of

non-ductile flat-plate buildings with interpolated non-linear curvature zones adopted to

fit with the experimental data of others. From their analyses Chow and Selna concluded

that although flat-plate systems can withstand moderate earthquakes, they would be

subjected to the large drift causing heavy damage to nonstructural elements. The

primary mode of failure of the structure was predicted to be the punching of the slab.

Shear stresses in the slab that accompany the yield moment were sufficient to cause low

cycle fatigue failure with vertical collapse of the slab induced by punching failure due to

sway. However, with continuous bottom reinforcement in the slab the likelihood of this

mode of failure is reduced (also reported by ASCE-ACI Committee 426, 1974, Durrani

et al., 1995, Gardner and Shao, 1996).


2-22

Menetrey (1998) conducted tests on 12 by 120 mm thick octagonal slab specimens

resting on rectangular hollow steel tubes to study the flexural and punching shear

failure of the flat plate column connections. A vertical downward load was applied to

the slabs via a 120 mm diameter steel tube at the slab centre. It was observed that the

load carrying capacity of the connection decreased with the increase of punching crack

inclination. Providing one ring reinforcement in the slab in addition to the minimum

orthogonal reinforcement controlled the inclination of the punching crack. The angle of

inclination of the punching crack decreased with an increase in the diameter of the ring

reinforcement. It was concluded that by changing the punching crack inclination angle,

which was achieved by providing the ring reinforcement, the mode of failure could be

changed in these connections.

Dechka et al. (2000) proposed a shear friction based mechanical model to determine

the capacity of the slab-column connections. The model was calibrated using 91 test

data on isolated interior slab-column connections with and without shear reinforcement.

The model was shown to give a reasonable prediction of the strength of the

connections.

2.2.2 Edge connections

In their 17 tests, Hanson and Hanson (1968) tested one edge column connection. The

specimen consisted of a 1.22 m x 1.14 m x 0.07 m slab with a symmetrically placed

150 mm square column at the edge of one of the longer sides. The slab was reinforced

with 1.67 % of reinforcement in each direction at both the top and bottom of the slab.

The slab is reported to have failed in punching shear. Hanson and Hanson
2-23

recommended distributing the unbalanced moment in the proportions of 60 % as the

flexure and 40 % by the eccentricity of shear about the centroid of the critical section.

Zaghlool (1971) tested 11 edge column connection specimens. All specimens were

1.83 m x 0.97 m x 0.15m. Square columns of 178 mm, 267 mm and 356 mm were

used. The slab edges without the column were rested on the test frame through

12.5 mm thick neoprane pads. The loads were applied to the test specimens through the

column stubs. The parameters considered in the tests were the moment shear ratio, slab

reinforcement content and the column width to effective depth ratio. From the results

of their tests Zaghlool concluded that the punching of the column through the slab at

ultimate was a secondary phenomenon because it occurred following the destruction of

the compression zone due to the combined action of flexure and shear. It was

concluded that besides the concrete compressive strength, the quantity of reinforcement

in the vicinity of the column, the ratio of column size to the effective depth of the slab

and the ratio of the bending moment to the shear force significantly influence the

capacity of exterior columns.

Stamenkovic and Chapman (1974) investigated the effect of axial load and moment

transfer in edge column connections. They tested 12 edge column connection

specimens with a square column of 130 mm. In their tests, the edge adjacent to the

column was kept free with the other three edges supported along their length using

12.5 mm and 38.1 mm diameter bars. The outline of the typical edge column specimen

used by Stamenkovic and Chapman is shown in Figure 2.12.


2-24

r--915-----1 r--915-----1
130mm square
column
l t lJ915
p
M-P j -l,_,r--_M
130mm square
column

T p -----1"''""1 25.4mm thld< mo1:r p -----1"''""1 25.4mm thick steel plate

T
76

788 788
'-----,_[76
--rT
l 130mm square
column
........... ~p
FRONT ELEVATION
l 130mm s;;::-T
column

"""''"""~P
SIDE ELEVATION
·--·- T

(a) Moment in x-direction (b) Moment in y-direction

Figure 2.12- Outline of Stamenkovic and Chapman (1974) specimens for the
edge column connections (a) Moment in x-direction; and
(b) Moment in y-direction.

Stamenkovic and Chapman modified the expression they developed for internal column

connections to model edge column connections. Under horizontal loading the

unbalanced moments are assumed to be transferred from the column to the slab by the

combined action of bending and torsion of the vertical slab-column interfaces normal to

the moment axis. For edge connections under normal moments (developed from

horizontal forces perpendicular to free edge, refer Figure 2.12a) a circular interaction

relation was developed and is shown in Figure 2.13. For edge connections subject to

tangential moments caused by horizontal loading parallel to the free edge (refer

Figure 2.12b) the straight line interaction formula shown in Figure 2.13 was proposed

and is similar to that for interior connections.


2-25

- - Edge column with normal moment


and corner columns

':!... +~ = 1 -...--+--- Internal column and edge column


Vu Mu with tangential moments

0 M/Mu

Figure 2.13- Interaction curves proposed by Stamenkovic and Chapman (1974).

Kanoh and Yoshizaki (1979) tested 8 edge slab-column connections to measure the

torsional shear stress on the side faces of slab-column connections. Using a beam

analogy similar to that of Park and Islam (1976), Kanoh and Yoshizaki concluded that

the ultimate torsional shear stress capacity is 0.636 .[1: MPa. This is considerably

greater than was given in ACI 318 ( 1977) with the increase in the strength attributed to

restraint of the critical section of the slab. The forces at the critical section of the slab

are shown in Figure 2.14. To calculate the ultimate strength of the slab-column

connections transferring the moment and shear, Kanoh and Yoshizaki proposed the

following equations

(2.3a)

(2.3b)

(2.3c)
2-26

Column

Figure 2.14- Forces acting at the critical section (Kanoh and Yoshizaki, 1979).

where Mo is the moment transferred to column; Mt is the flexural strength at the critical

section; Ms is the moment transferred by the shear applied at the critical section; Mt is

the torsional moment at the critical section; Mtest is the measured moment transferred to

column about the centroidal axis of column; Vo is the maximum shear capacity of the

connection; Vu is the ultimate vertical shear stress; Vtest is the measured shear force

transferred to column; and Ac is area of critical section.

Rangan and Hall (1983a, 1983b) tested five half-scale models of edge panels of a flat

plate floor systems. They increased the size of the specimens over those previously

tested by moving the boundary to the second line of contraflexure from the connection

(refer Figure 2.8). Two of the five models tested had a spandrel beam of

300 mm x 250 mm along the free edge. In the other four specimens a spandrel section
2-27

was provided within the depth of the slab over a width of 250 mm. The spandrel

contained closed ties at 60 or 70 mm centres. Rangan and Hall showed that most of the

floor load entered the column via the spandrel, even in case of the slabs without the

spandrel beams. Thus, the strength of the spandrel in combined shear and torsion was

found to be critical in the strength and behaviour of slab-column connections.

From their tests on edge-column-slab connections, Rangan and Hall (1983b) observed

considerable redistribution of moments throughout the slab with an increase in the

applied load. The restraint to the spandrel against longitudinal expansion considerably

increased its torsional strength. Once cracking occurred in the connection, the stiffness

decreased significantly compared to that of the uncracked state. Based on their test

data and the concept of lateral restraint put forward by Onsongo and Collins (1972),

Rangan and Hall developed expressions for the moment and shear transfer in the

spandrel faces of the column.

The AS 3600 (1994) model for the design of the flat slabs against punching shear

strength is based on the results of Rangan and Hall (Rangan and Hall, 1983a and

Rangan, 1987). The physical model used to develop these models is as shown in

Figure 2.15, where C1 and C2 are the column size normal and parallel to the free edge,

respectively; bt and bz are the corresponding widths of the critical section of the slab

and are given by b1 = C1 + d/2 and b2 = C2 + d; dis the effective depth of the slab; V,
V1 and V2 are shear forces transferred to column centre and at the front or at a back

face of the critical section, respectively; M* and M1 are the unbalanced moment

transferred to the column section and the yield moment of the slab steel at the front or

at back of the critical section; and T1 is the torsional moment at a side of the critical
2-28

SLAB

Front face of
critical section

Figure 2.15 - Physical model showing load transfer (Rangan, 1987).

section. The critical section was assumed to be at a distance of d/2 from the face of the

column. It is assumed that the shear stress at the critical section is uniformly

distributed. In Rangan's (1987) model 40 percent of the unbalanced moment is

transmitted to the slab by shear unlike the ACI 318-1999 model, where the proportion

of the unbalanced moment transferred by shear is a function of the column size and the

effective depth of the slab. Rangan compared the model against 104 different test

results of interior, edge and comer column connections. The average test to calculated

ratio was 1.6 with the standard deviation of 0.4 and coefficient of variation of 25 %.
2-29

Thus the design equations of AS3600-1994 are seen to be conservative. Rangan (1987)

stated that shear strengths measured in the laboratory studies were significantly

influenced by the boundary conditions adopted for the test specimens giving such

conservative results.

Rangan (1990a) used the model developed by Rangan (1987) to compare the 117 test

data covering wide range of interior, edge and comer slab-column connections. The

average of the test to calculated value was shown to be 1.59 with a coefficient of

variation of 25 percent. It was claimed that AS3600-1988 was simpler to use and a

useful altemativ~? to the ACI building code model for the design of punching shear in

flat slabs and plates.

Rangan (1990b) tested four specimens to study the effect of spandrel beams, closed ties

and different concentrations of slab steel on the moment and shear transfer strength in

flat-plate floors in the vicinity of edge columns. It was observed that shallow spandrel

beams with nominal closed ties are effective in preventing the punching shear failure

with the closed ties in the spandrel strip increasing the shear strength. Top

reinforcement provided within the width of column strip plus the adjacent half middle

strips controlled the transfer of bending moment to the column and was only slightly

affected by the presence of the torsion strips with closed ties. Rang an further concluded

that the ratio of unbalanced moment to the shear transmitted to the column significantly

influenced the punching shear strength.

El-Salakawy et al. (1998) undertook an experimental study of eight slab-column edge

connections with and without openings and shear reinforcement in the slabs and loaded
2-30

under high moment to shear ratios. The basic dimensions of the specimens are as shown

in Figure 2.16. All the specimens except those with stud shear reinforcement

r-1020-----j

25Dmm square
column
l
Slab

Simply
TI J1540

Column supported

H Slob

ELEVATION

Figure 2.16 - Edge connection specimens of El-Salakawy et al. (1998).

failed in punching shear mode, whereas the two specimens with shear studs failed in

flexure and punching-flexure. The openings in the slabs had a significant effect in

reducing the capacity of the connections with the greatest effect for slabs with higher

moment to shear ratios. The shear reinforcement significantly improved the ductility

and the strength of the connections.

El-Salak.awy et al. (1999) tested a further three full-scale slab-column edge

connections. Together with the El-Salak.awy et al. (1998) study five specimens tested

had square openings located in the neighbourhood of the connections. The main

parameters of the study were the locations and size of the openings in the slabs. The

size and location of the slab openings influenced the capacities of the connections. The
2-31

larger the openings the lower the strength of the connections. The connection capacity

also reduced when the opening was at the front of the column compared to similar

sized openings located at the sides of the column.

2.2.3 Corner Connections

Zaghlool et al. (1970) tested four specimens of comer column connections with the

slabs resting on square columns located at the comers. All specimens consisted of a

3.05 m x 3.05 m x 0.165 m slab (see Figure 2.17) supported by 140 mm or 165 mm

square columns. Three specimens failed in shear and one failed through a support. A

linear shear moment equation was proposed similar to that proposed by Moe (1961).

However, the test data of Zaghlool et al. suggested a factor of 0.04 instead of 0.33 for

the moment reduction factor as proposed by Moe for the interior slab-column

connections. Similarly, the test results compared best with the moment-shear

interaction model of Hanson and Hanson (1968) with a moment reduction factor of

0.04 instead of 0.33.

Zaghlool et al. (1970) developed a model for calculating the strength of comer column

connections considering the axial force transferred to the column alone. The model is

based on the principal diagonal tensile strength of the failure cone with an angle of

inclination of the failure plane of 45 degrees. The mean test to calculated strength for

his test results was found to be 1.03 with the standard deviation of 0.09. In the model

the effect of dowel action and aggregate interlock were neglected.

Zaghlool (1971) and Zaghlool and de Paiva (1973) reported tests on eleven comer

column connections. All specimens consisted of a 150 mm thick by 1.07 m square slab.
2-32

Square columns of different sizes 178 mm, 267 mm and 356 mm were used. As in the

380 mm square load plate

----
/t

Reactions

Figure 2.17 - Test specimen (Zaghlool et al., 1970).

edge column connections, in these tests the parameters studied were the slab

reinforcement content, the moment to shear ratio and column size to slab effective

depth ratio. The edges of the slabs without the column were rested on the test frame

through 12.5 mm thick neoprane pads. The load was applied to the specimens

symmetrically through the column stubs using hydraulic loading systems. Zaghlool and

de Paiva (1973) reported that the straight line shear hypothesis adopted in the 1971

ACI Code to find the strength of the column-slab connections gave extremely

conservative results compared to the results of their tests. Punching of the column

through the slab at failure was considered to be a secondary phenomenon because it


2-33

occurred following the destruction of the compression zone due to the combined action

of flexure and shear. It was also concluded that besides the concrete compressive

stress, the amount of reinforcement in the vicinity of the column, the ratio of column

size to the effective depth of the slab and the ratio of the bending moment to the shear

force significantly influenced the capacity of exterior connections.

Stamenk:ovic and Chapman (1974) tested six comer column connection specimens with

the main variable being the ratio of applied axial moment. The arrangement of the

comer column specimens of Stamenk:ovic and Chapman is shown in Figure 2.18. From

their test results, Stamenk:ovic and Chapman proposed an interaction curve for comer

column connections based on a modification of their internal column interaction

relationship (Figure 2.13).

Walker and Regan (1987) tested eleven comer column specimens with the slab resting

on four columns. A typical specimen is shown in Figure 2.19. The bottom of the

columns were connected by steel ties (see Figure 2.19) with pinned ends. The tie forces

were determined using the strain gauges fitted to the rods. Except the specimen SC6,

the specimens contained both positive and negative moment reinforcement in

orthogonal directions. In specimen SC6 the negative moment reinforcement was

provided in the diagonal directions. Most of the specimens failed in punching shear.

Comparisons of the results with ACI 318-1983 provided a good assessment of the

column moments. The results of specimen SC6 showed that the special arrangement of

negative reinforcement increased the stiffness of the slab and had some influence in the

behaviour of comer slab-column connection.


2-34

r----915~

130mm square
l_l 915

column
p ~p

T P ......---irzzll25.4mm thkk •tee1~


76
788 I

l 130mm square
column
lzzzi- ~ p
ELEVATION
-.
T

Figure 2.18 - Comer column connection specimens used by Stamenkovic and


Chapman (1974).
2-35

,.. 2000-------~, s 80

y---t-;:=====::::=:::;~ t
8 0

250 250
I I 5oo I 5oo I 5oo I I
Column size:
SC8, SC9, SC11: 160mmx160mm
ELEVATION SC12: 300mmx300mm

PLAN
(SC8 AND SC9)

250
300
300
300
250

250 250
500 500 500
I I I I I I
PLAN
(SC11 AND SC12)

Figure 2.19- Specimens SC8-SC12 (Walker and Regan, 1987).


2-36

Hammill and Ghali (1994) tested five full scale reinforced concrete flat plate

connections with comer columns subjected to shear and moment transfer. Three of the

specimens were without any shear reinforcement while the remaining two contained

shear reinforcement. The specimens used consisted of 1075 mm wide square slab by

150 mm thick and 250 mm square column. Hammill and Ghali reported that the

Canadian code (CAN3-A23.3-M84, 1984) and the ACI 318 (1989) code are

conservative for slab-column connections without shear reinforcement. Analyses by

Hammill and Ghali showed that the equation developed by Elgabry (1991), based on

the linear finite element analysis, is more realistic than the code models for comer

columns. Finally, they suggested that a nonlinear finite element analysis is required to

obtain a better understanding of the behaviour of comer slab-column connections.

2.2.4 The Effect of Boundary Conditions

Rankin and Long (1988) observed that the results of the tests using the conventional

specimens with the slabs extending only to the first line of contraflexure

(refer Figure 2.8) lead to erroneous conclusions. In such conventional models the

influence on the results of the boundary conditions was significant. Rankin and Long

conducted tests of 27 conventional and 17 full panel slab-column specimens to failure

to verify this behaviour with the reinforcement content and span to depth ratio of the

slabs varied.

Vertical deflection of reinforced concrete slabs is known to be accompanied by an

outward expansion of the slab at the boundaries and if this expansion is restrained the
2-37

load carrying capacity of the slab is increased by arching (also referred to as

compression membrane action). Rankin and Long observed that in the punching of

interior slab-column connections, the slab beyond the nominal line of contraflexure

(that is beyond the first line of contraflexure, refer Figure 2.8) provides lateral restraint.

Consequently the punching capacity of full panel slab-column connection specimens is

enhanced by 30 to 50 percent compared to the punching strength of the specimens with

the slabs extending up to nominal line of contraflexure. Rankin and Long proposed a

series of equations to account for arching effects.

Kuang and Morley (1992) tested reinforced concrete slab-column connections of

different thickness, reinforcement contents and lateral restraints. A concentrated load

was applied through the columns with all slabs failing in a punching shear mode. The

variation in the reinforcement content was found to have little effect on the ultimate

punching load of the heavily reinforced specimens, but it was effective for lightly

reinforced specimens. Johanson's (1962) yield line theory was found unsuitable in

predicting the punching load capacity for highly restrained and/or lightly reinforced

slabs. Compression membrane actions were developed depending on the degree of end

restraint with the action enhancing the punching shear capacity.

Kuang and Morley also observed that serviceability conditions such as deflections and

crack widths were affected by the degree of the boundary restraint. As may be expected

deflections were reduced when a compression membrane action develops. Crack widths

in the lightly reinforced slabs were larger than the more heavily reinforced slabs. Also

the cracks were fine and large in numbers in the slabs that were highly restrained.
2-38

Kuang and Morley concluded that compression membrane forces play a significant role

in controlling the formation of cracks in the slab.

2.2.5 Tests on Multi-Panel Specimens

The performance of flat slab structures subjected to seismic loading has attracted

increasing attention especially because of the poor performance of this type of structure

under dynamic loading. In early studies the testing was related specially to the

performance of individual interior or exterior connections (determinate structures)

subjected to a combination of moment and shear. Rangan and Hall (1983a, 1983b), and

Robertson and Durrani (1992) showed that the tests studied did not take into account

the restraining effect of the boundaries and the shift of contraflexure lines with the

deterioration of the structure as the test progresses, both of which affect the strength.

The load carried to individual columns is found to be significantly influenced by the

redistribution of forces within the slab. Given that the behaviour of slab-column

connection is non-ductile, redistribution of slab forces may be significant in overloading

the column.

Regan (1978, 1981) conducted a series of tests comprising of different types of single

and multi-panel specimens with interior, exterior and comer columns. Regan tested 12

specimens supported by comer columns, 18 slabs supported by edge columns, 3

statically indeterminate slabs with interior columns and 2 slabs mixed with wall and

edge column supports.


2-39

Gosselin (1984) tested four slab-column connections consisting of exterior and interior

panel extending laterally to the centre lines between the column. The overall size of the

slabs was 3 m x 2 m x 0.063 m. The exterior and the interior columns were of

225 mm x 150 mm and 225 mm x 225 mm, respectively, and drop panels were

provided in the column regions. The columns were post-tensioned to simulate a gravity

load equivalent to a stress of five storeys and combinations of gravity and lateral load

were applied to the specimens.

Lamb (1984) tested four one third scale specimens of reinforced concrete slab-column

connections with and without the spandrel beams. Of the four specimens two were

constructed with the spandrel beams. The specimens consisted of interior and exterior

columns extending above and below the slab a distance equivalent to mid-height of the

column used in the scaled prototype. The tests were undertaken by applying

combinations of lateral and gravity loads over a number of cycles. Of the specimens

without edge beams, only one specimen failed due to punching.

Robertson and Durrani (1992) investigated the behaviour of interior slab-column

connections as a part of an indeterminate system under gravity and reversed cyclic

lateral load. The panel dimensions were 2.9 m by 2.0 m and the slabs were 115 mm

thick as shown in Figure 2.20. Robertson and Durrani reported that an increase in

gravity load at the interior slab-column connection caused significant reduction in the

capacity of the connection to transfer unbalanced moment. A reduction in the stiffness

was also reported with the increase in the gravity load and was attributed to accelerated

cracking in the slab around the connection. Robertson and Durrani (1990)
2-40

12900
Lateral
0

"
L(')
..-

0
~--254mm square columns
"
Slab width = 2.0m
Reactions

Figure 2.20 - Details of a typical test specimen of Robertson and Durrani (1992).

tested altogether nine specimens including two isolated (one interior and one exterior

column connections) specimens.

Falamaki and Loo (1992) tested to failure a series of nine half-scale models of

reinforced concrete flat-plate structures. They developed their specimens from the flat

plate structural system shown in Figure 2.21 with their model representing two

continuous panels of a building floor. The models consisted of six columns including

three slab-column connections included with a spandrel beam or torsion strip. The

columns consisted of rectangular steel sections with stiffness matched to that expected

from concrete columns. The steel columns were connected to the concrete slab via steel

plates and bolts. Their aim was to study the ultimate behaviour of reinforced concrete

flat-plates with spandrel beams of different depths. The variables considered


2-41

Z•Q

.... ~ ....
~

X
I
X
c
·a
~
Jl

.....
"";:::... ;:::...

Figure 2.21 - Structural system used for the development of specimen by


Falamaki and Loo (1992).

were the strength and stiffness of the spandrel beam, the slab reinforcement ratio in the

vicinity of the columns, the volume of closed ties, the width of the edge columns and

the loading patterns.

Based on their test data, Falamaki and Loo concluded that the failure mechanisms of

the comer and edge connections were greatly influenced by the distribution of slab

flexural reinforcement through the column regions and the strength of the spandrel

beam or torsion strip.

Whilst the tests of Falamaki and Loo (1992) is a study in the area, the use of steel

section connected to the slab via bolts must be questioned. As has been discussed
2-42

previously in this thesis, the boundary conditions are of vital importance in the

interpretation of the experimental results. In this respect the moment transfer to the

column in concrete slab-column construction is likely to be influenced by the

connection detail. The tests of Falamaki and Loo are for a different floor construction

system and should be interpreted with this in mind.

Durrani et al. (1995) tested four two-bay slab-column sub-assemblies designed and

detailed for gravity load as was traditionally used in the 1950's and 1960's. The tests

were carried out under earthquake type loading with the aim to study the behaviour of

slab-column subassemblies of the buildings constructed in 1950's and 1960's. With the

increase of gravity load, the interior connections were more vulnerable in punching

shear than the exterior connections due to the lack of continuous bottom reinforcement

through the columns. The moment transfer capacity of the slab at the interior

connections was found to depend on the amount of, and the detailing of, the top and

the bottom reinforcement.

Durrani et al. observed that under combined lateral and gravity loading the drift

capacity was reduced with an increase in the gravity load. Cracking of slab under the
I

gravity load decreased the initial lateral stiffness of the slab-column sub-assemblies

under lateral loading. As may be expected, sub-assemblies with spandrel beams were

found to have larger initial stiffness compared to those without spandrel beams. The

reduction in stiffness was greater in the exterior connection in the negative moment

region. With the loss of stiffness in the exterior connection, the gravity shear was
2-43

redistributed to the adjacent interior connection, hence, averting a punching failure in

the exterior connections.

The behaviour of the interior connection of the Durrani et al. (1995) tests generally

agreed with the eccentric shear model of ACI 318 (1989). The shear strength of the

connections decreased with the increase of lateral drift under cyclic loading. Durrani et

al. concluded that the moment transfer capacity of the exterior connections was

typically characterised by the combination of bending and torsional action. Hence a

good prediction could be made using the equivalent beam method, that is by combining

the flexural strength of the slab over the width and depth of the columns and adding the

torsional capacity of the spandrels.

Gardner and Shao (1996) tested to failure a half scale two-bay by two-bay (9 columns)

slab with span lengths of 2743 mm (see Figure 2.22). The support system consisted of

7 columns of 254 mm square and two circular columns of 254 mm diameter (one edge

column and one comer column). In their test arrangement Gardner and Shao used

supplementary supports as shown in Figure 2.23. However, because of the use of the

supplementary supports the true failure load could not be obtained for the external

columns. Based on the strength enhancement model proposed by Shehata (1990) and

the CEB (1990) size effect expression for the interior connections, Gardner and Shao

proposed that the strength of the connections be given by

(2.4)
2-44

5867

254
D
D=254mm/
2 43

254mmx254mm
Columns
5867 [] D

~--27431--~~--2743,--~
127 254

Figure 2.22 - Dimensions of Gardner and Shao (1996) specimens.

~------5867------~

•· Ill
Ill

II II II Ill Ill
5867 CJ• Ill··'·
Ill

95mmx95mm supplementary supports


at the clear distance of 1 27mm 2743
from the edge of the column

II II II II II
127
2743 2743
127 254

Figure 2.23 - Positions of supplementary supports (Gardner and Shao, 1996).


2-45

where Vu is average shear stress (MPa); Vc is a nominal shear stress of the slab (MPa);

Vu is the shear force (N);fyis yield stress of the reinforcement (MPa); u is the length of

the critical perimeter (mm); dis average effective depth of tension reinforcement (mm);

p is average of the top and bottom reinforcement content ratio in the slab; /em is mean

cylinder strength of the concrete (MPa).

Although Gardner and Shao's investigation focused mainly on the behaviour of the

interior column, for comer and edge connections they suggested that designers should

follow the ACI linear interaction formula or to use simple expressions independent of

eccentricity ofload and steel content (similar to BS 8110, 1985). From the test it was

observed that the interior slab-column connection was more critical in punching shear

than edge or comer columns. Similar results were found in the study of Durrani et al.

(1995) and were attributed to the redistribution of forces from external columns to the

internal column. It was concluded that as punching shear is associated with high

moment regions it can be classified as a flexure-shear phenomena. They remarked that a

strut and tie model can be used to explain better the behaviour but did not give any

further details on this aspect. Gardner and Shao concluded that isolated punching tests

can be used to study the behaviour of the interior slab-columns connections but did not

discuss boundary conditions.

Dechka and Dilger (2000) tested 2 by two-span continuous slab-column frame

specimens. One specimen was designed for gravity load only whereas the second

specimen was designed for combined seismic and gravity load. Both specimens had a

150 mm thick by 5 m wide by 10.25 m long slabs. Each specimen had 3 columns of 250
2-46

mm square (1 interior and 2 edge columns) projecting 1.5 m above and 1.5 m below the

slab. Shear studs were provided in both specimens. The test results showed that shear

studs were effective in increasing not only the punching shear strength of these

connections under seismic load, but also in improving the ductility of the system. Shear

studs along with other reinforcement maintains a high level of concrete integrity to

allow the structure to form a plastic hinge at the connections, thereby making the

connections capable of dissipating energy during seismic loading.

2.2.6 Tests of Connections with Shear Reinforcement

It has been reported~ the literature (Brown and Dilger, 1994, ACI 408.2R, 1992) that

concrete shear strength under cyclic moment transfer is less than the strength under

monotonic loading. Cyclic moment transfer causes a widening and propagation of

cracks and a degradation of the concrete shear resistance. Hence, slab-column

connections subjected to cyclic loading may require shear reinforcement, which may

not be required under monotonic loading. Tests have been conducted using mainly four

types of shear reinforcement (shown in Figure 2.24), conventional closed ties, bent up

bars, rolled steel sections and shear stud reinforcement (SSR).

Islam and Park (1976) carried out tests on concentrically loaded square columns with a

rectangular slab with and without shear reinforcement. The tests were performed under

gravity and lateral loads. The main conclusions arising from the tests were:

• Specimens without shear reinforcement failed suddenly with the splitting of the

concrete along the bars at the top of the slab on the side of the column where

the shear was greatest.


2-47

1-1-
I LL
I u
1-t-

{~}

Closed ties
0
Bent up bars

I I

Rolled steel section Shear stud reinforcement (SSR)


shear head

Figure 2.24 - Different types of shear reinforcement used in slab-column


connections.
2-48

• Cranked bars used for the shear head reinforcement did not increase the

ductility of the slab but marginally increased the strength.

• Closed stirrups in the slab around the longitudinal bars passing through the

column significantly improved the ductility under cyclic unbalanced moment in

the inelastic range.

• Two legged and four legged stirrups performed equally well and were effective

in torsion and flexure and improved the strength and ductility of the system.

• The provision of closed ties formed a small in-built beam inside the slab

allowing the connection to behave similar to that of beam-column systems.

Elgabry and Ghali (1987) reported the test results of five full-scale reinforced concrete

flat-plate connections with interior columns. All the specimens were subjected to

shear-moment transfer and with the exception of one specimen stud shear

reinforcement was used in the column heads. The test results showed that the

stud-shear reinforcement was effective in increasing not only the shear strength of the

connections, but also increased the ductility of the connections.

Elgabry and Ghali (1990) described the design and detailing of different flat

plate-column connections with shear stud reinforcement according to the

ACI 318-1989 model. Based on their test results a new design method was

recommended.

Dilger and Cao (1991) reported that the reversed cyclic loading of slab-column

connections reduces the capacity to transfer the unbalanced moment. This observation

was further verified in the study of Megally and Ghali (1994). Megally and Ghali
2-49

recommended that slab-column connections must be capable of withstanding a drift

ratio of 1.5 percent without punching failure. Generally in seismic zones lateral load

resisting structural elements, such as shear walls, should be combined with flat plates to

ensure the lateral drift ratio does not exceed 1.5 percent.

Mortin and Ghali (1991) tested six full scale edge slab-column connection using SSR.

The use of SSR was effective in improving the shear strength by about 30 to 50 percent

and the ductility was also significantly improved. Deflections at failure were found to be

double that of companion tests on panels without shear studs. It was observed that SSR

could change the failure mode from a brittle punching shear failure to a more

acceptable ductile :flexural failure.

The effect of the type of shear reinforcement and the shear reinforcement ratio on the

punching shear strength of monolithic slab-column connections was studied by Yamada

et al. (1992). Yamada et a1. tested 13 slab-column specimens. The slabs were supported

by concentrically located column stubs and were tested under symmetrically distributed

monotonic loading. Yamada et al. used hat and hook type shear reinforcement (shown

in Figure 2.25). The hat type reinforcement was prefabricated into a cage and welded

to two horizontal straight bars. The hook type shear reinforcement consisted of

individual bars with a 180 degree hooks at each end. The ratios of the shear

reinforcement varied from 0.0 to 1.98 %. The use of both types of shear reinforcement

showed a superior performance in the post peak region of the load-deflection response.

In addition a significant increase in the punching shear resistance with the hook type

shear reinforcement was reported. This increase was due to the good anchorage of the

shear reinforcement.
2-50

ottom flexural
reinforcement

Hat type shear reinforcement

H ok type boro-
rro

Every node
lL Flexural barsl!.VEvery second node

Hook type shear reinforcement


rJ
Hooks

Figure 2.25 - Shear reinforcement types used in the specimens of


Yamada et al. (1992).

Rangan and Lim (1992) reported tests of three large scale edge slab-column

connections with one specimen containing no shear reinforcement and two specimens

containing SSR. The tests consisted of 4950 mm by 3170 mm by 110 mm thick slabs

with a 250 mm square columns located at the middle of the free (4950 mm) edge.

The slab was loaded via 24 point loads evenly distributed across the slab simulating a

uniform loading. The SSR was effective in increasing the punching strength and the

ductility of the concrete slab-edge column connection with the greatest capacity from

the specimen with the highest quantity of SSR.

Megally and Ghali (1994) reported on the design of slab-column connections in seismic

zones based on the analyses of various experimental studies. Significant improvement in

ductility and drift capacity was reported with the use of shear stud reinforcement. It
2-51

was concluded that SSR is better than the conventional stirrups in improving the

response of slab-column connections.

Brown and Dilger (1994) tested four interior flat slab-column connection specimens

containing SSR. The specimens consisted 150 mm by 1900 mm square slabs with

250 mm square columns at the centre protruding 700 mm from each face of the slab.

The parameters investigated were the size of the shear reinforced zone and the

distribution of the main steel in the slab. The specimens were subjected to shear due to

gravity loading combined with reversed cyclic lateral displacements imposed at the

columns to represent seismic loading. Two specimens failed around the column after

the SSR yielded and the other two failed outside the shear reinforced zone before the

yielding of the SSR. All the specimens had a ductile failure mode with the final storey

drift ratio of more than 5 percent. It was concluded that the shear stress resisted by the

concrete in shear zone reinforced with SSR was about 0.15.fj[, where f~ is the

cylinder strength of the concrete.

Lim and Rangan (1994, 1995) reported tests of nine large scale reinforced concrete

slabs to further study the effectiveness of SSR. Of the nine specimens seven specimens

were edge column connections, which includes the three specimens reported in Rangan

and Lim (1992) and two specimens were comer column connections. The four edge

column connection specimens not reported in Rangan and Lim (1992) consisted of a

4950 mm by 3120 mm, 110 mm thick slab with a column located at the middle of its

free (4950 mm) edge. Rectangular columns of different sizes were used in these

specimens. The comer column connection specimens consisted of a 2.60 m x 3.17 m,


2-52

110 mm thick, slab with a 250 mm square column at one comer. It was observed that

the use of SSR was effective in increasing the shear strength of the slab in the

slab-column connections as was found in the earlier studies ofMortin and Ghali (1991),

Rangan and Lim (1992) and Megalli and Ghali (1994).

Rangan and Lim's tests showed that the punching shear failure with SSR is initiated as

a combined torsion-shear failure in the torsion strips with closer spacing of the SSR

providing a better performance of the connections.

Lim and Rangan modified the truss theory of Simmonds and Alexender (1987) to

incorporate the effect of SSR in slab-column connections. The average ratio of test to

calculated shear strength was 1.02 with the coefficient variation of 10 percent. Lim and,

Rangan concluded that the strength of the slabs correlated well with the truss theory.

Gomes and Regan (1999a, 1999b) used short offcuts of steel 1-beams in the twelve

200 mm thick slab specimens representing interior slab-column connections. The

punching load was increased from 560 to 1227 kN with the use of shear reinforcement.

The mode of failure was also changed from a brittle punching mode near the column to

a ductile failure mode with failure outside of the shear reinforcement region when shear

reinforcement was added. Gomes and Regan (1999b) proposed a theoretical model

based on the results of the tests of Kinnunen and Nylander (1960), Andersson (1963)

and Shehata (1985).

Megally and Ghali (2000b) recommended a design procedure for slab-column

connections under earthquake load. A minimum amount of shear stud reinforcement


2-53

was recommended to ensure a ductile behaviour of the connection. They further

concluded that shear capitals are not as effective as providing stud shear reinforcement

in avoiding a punching failure during an earthquake.

2.3 Engineering Models for Design of Slab-Column Connections

Many analytical investigations have been conducted to study the behaviour of

slab-column connections. Elastic plate theory, beam analogy, linear variation of shear

stress model, truss model, finite element methods are but some of the approaches used

(Aalami, 1972, Park and Islam, 1976, Simmonds and Alexander, 1987). These models

are briefly discussed below.

2.3.1 Models Based on Elastic Plate Theory

Based on elastic plate theory, Mast (1970) concluded that the straight line shear

distribution hypothesis, which considers the linear effect of the unbalanced moment in

calculating the punching capacity of the slab-column connections, underestimated the

torsional effect in punching shear. Hence, the model is only suitable for the design of

connections having columns with large depth to width ratios. Mast developed an

analogy for behaviour of the edge column connections based on the behaviour of the

interior column connections. His analogy assumes that for edge column connections

only the supporting column width is significant in carrying the reaction. Based on this

assumption various equations to find the stresses due to the applied moment at the edge

were developed. Mast compared the results of his analytical model with the

experimental data of edge-column connections of Hanson and Hanson (1968). The

stresses determined using Mast's theory did not agree with the experimental results for
2-54

rectangular columns having side dimensions cl > c2, where c] is the side dimension

perpendicular to the axis of bending and c2 is the dimension parallel to the axis of

bending (see Figure 2.26).

Aalami (1972) presented a method of obtaining the lower and upper bounds to the

rotational stiffness of flat floor plates at the column-slab connection. The method is

based on the elastic theory of isotropic thin plates. In the lower limit the column

stiffness does not add to the bending stiffness of the floor. For the upper limit the

column is assumed to make the slab-column interface infinitely stiff. Aa1ami showed

that the moment-rotation stiffness contributed by a floor plate is a function of the

column dimensions. He concluded that the rotational stiffness of the torsion strips is not

greatly influenced by the span or the boundary conditions. This, however, contradicts

the findings of other investigators (Rangan 1987, Rankin and Long 1988) and is in

violation of the main assumption adopted in the beam strip model

Based on elastic plate theory and on Levy's equation, Pecknold (1975) proposed an

expression for the effective width of the interior panel of column-slab connection

systems for various column sizes and slab aspect ratios. The results from the proposed

method compared well with numerical results using 16 degrees of freedom plate

bending elements.

Wong and Coull (1980) developed an influence co-efficient method to determine the

effective width and coupling stiffness of a floor slab in a laterally loaded flat
2-55

Free edge of the slab

Column

Figure 2.26 - Column dimensions relative to the direction of moment Mx used by


Mast (1970).

plate-column structure using the classical thin plate theory based on Levy's method of

analysis (Timoshenko and Woinowsky-Kreiger, 1959). They claimed the method to be

better than the finite element solution in terms of computational effort. Pavlovic and

Poulton (1991) developed an approximate closed form solution for computing the

effective slab width of the flat plate-column structure under vertical and lateral load

again using Levy's method of analysis (Timoshenko and Woinowsky-Kreiger, 1959).

Pavlovic and Poulton compared their results with the methods of Wong and Coull

(1980) and Pecknold (1975) and claimed it to be more accurate and efficient.

2.3.2 Beam Analogy Methods

Various beam analogies have been proposed to study the behaviour of slab-column

connections (Hawkins and Corley, 1971, Park and Islam, 1976). In the beam analogy

method the slab is idealised as beams running over the column from two directions at
2-56

right angles (see Figure 2.27) and the strength of the connection is taken to be that of

the beams. In Figure 2.27, C1 and C2 are the column dimensions and dis the effective

depth of the slab. The beam analogy transforms a three-dimensional problem into

two-dimensions and provides a convenient method for the design of connections

involving the transfer of shear and unbalanced moment from the slabs to the columns.

The slab strips adjacent to the column and parallel to the column edges act as beams

under the action of flexure, torsional moment and shear force similar to that shown in

Figure 2. 7. The strength of each beam depends on the combinations of the stress

resultants and the strength of the section with respect to these stress resultants.

Assuming a sufficiently ductile system the method gives a lower bound solution to the

capacity of the connections. A detailed summary of the various beam analogy models

proposed is given in Park and Gamble (2000).

Hawkins and Corley (1971) developed an ultimate strength design procedure for

internal and external slab-column connections using the beam analogy. In their model

the strength of the edge connection is governed by either combined flexure and torsion

or combined shear and torsion. The applied moment is initially resisted at the front face

of the column. The flexural strength of this face in negative bending is nearly exhausted

before significant moments are distributed via the side faces by torsional action.

Hawkins and Corley assumed that once the shear on the front face is close to its

ultimate value, additional shear is transferred to the column via the side faces until a

failure condition is reached.


2-57

Idealised beam
Column~

"D
+
~
[I'
I ~. . ...·lrI
G1' -
T
Slob

INTERIOR COLUMN CONNECTION

, . C1 +d/2 ·I
Ideolised beam
Slob
1 Idealised beam
1-

. ·. llI 1
"D
+
s C::l :~C1 s
+
fj
t-
. .l
O.Z,
• •
ctt
• r
·l -
T lt"column
T Column
Slob
CORNER COLUMN CONNECTION

EDGE COLUMN CONNECTION

Figure 2.27 - Beam analogy for the design of edge column connections (Park and
Gamble, 2000).

Park and Islam (1976) developed a beam analogy for interior slab-column connections.

They assumed that sufficient ductility is available for bending, torsion and shear at the

critical faces to allow the development of the ultimate capacities. In their model the

critical section was located at one half of the effective depth from the outside of the
' -
column. The unbalanced bending moment and the shear forces transferred by the

connection were expressed in terms of flexural moment, torsion and the shear force

transferred at each face of the critical section. An interaction diagram plotting values of
2-58

the ultimate moment (Mu) and ultimate shear force (Vu) based on the equations derived

gave slightly improved results to that of the ACI (1971) model. Park and Islam

extended their approach to cases with shear reinforcement. The shear reinforcement

considered was bent up bars, closed ties and structural steel column heads.

Other beam analogy studies include those ofZaghlool (1971) and Kanoh and Yoshizaki

(1979). These studies have been discussed in Sections 2.2.2 and 2.2.3 of this thesis.

2.3.3 Models Based on Elasticity

DiStasio and Van Buren (1960) proposed a working stress method of analysis for the

strength of the slab-column connections under combined shear and unbalanced

moment. Di Stasio and Van Buren assumed that the shear stresses at the critical section

vary linearly with the distance from the centroidal axis of the perimeter, where the

critical section was taken at the distance h-1.5 inches from the face of the column

(h =total depth ofthe slab in inches). In considering punching shear, DiStasio and Van

Buren limited the maximum vertical shear stress to 0.0625 J; when the critical section

is assumed directly adjacent to the column periphery.

Moe (1961) had also used the concept of linear variation of shear stress to develop a

moment-shear interaction equation. Moe used a reduction factor of 1/3 to transfer the

unbalanced moment in the form of shear stress. Moe recommended a limiting vertical

shear stress to allow the flexural failure of the connections to govern the ultimate

strength rather than shear failure.


2-59

In the ASCE-ACI Committee 326 (1962) report, based mainly on the studies of

DiStasio and Van Buren (1960) and Moe (1961), the concrete strength, the magnitude

of bending moment near the column and the relative size of column compared to slab

thickness were recognised as the main parameters governing the punching shear

strength of slabs. ASCE-ACI Committee 326 (1962) recommended a limiting shear

stress of 4.[1{(d I 2 + 1) (psi) on a critical section which is at the periphery of the

column, where J; is the compressive strength of the concrete and d effective depth of

the slab. Based on the evaluation of 25 test results the Committee further recommended

that at d/2 from the face of the column the shear stress to be limited to 4.[1{ . The

ASCE-ACI (1962) suggested that the fraction of the unbalanced moment to be

transferred by shear be taken as K =0.2.

Hanson and Hanson (1968) also developed a model based on the linear variation of

shear stress. They found a better correlation between the test and theoretical results

when the value of the moment reduction factor was taken asK = 0.4.

The American Concrete Institute gave design recommendations for the first time in its

1971 edition of the code (ACI 318-1971). For rectangular columns the moment

reduction factor (K) given in ACI 318-1971 was

(2.5)

where C1 and C2 are the dimensions of the column measured in the direction of the

moment and in the direction normal to the moment, respectively and d is the effective
2-60

depth of the slab. For a square column (C1 = C2) equation 2.5 gives K = 0.4 as

recommended by Hanson and Hanson (1968). In the 1971 edition of the code a limit

was placed on the shear stress of 4..Jf: (psi) at the critical section.

ASCE-ACI Committee 426 published a comprehensive report on the state of art on the

shear strength of slab-column connections in 1974. Design methods for the shear of

reinforced concrete slab-column connections without shear reinforcement and with the

transfer of shear and unbalanced moment were introduced in ACI 318 (1977). The

design shear model adopted in ACI 318-1977 took the critical section at a distance of

d/2 from the column. The shear stress on a critical perimeter was taken to vary linearly

with the distance from the centroidal axis of the critical section and that the shear stress

is developed because of the shear force and a component of the unbalanced bending

moment. The fraction ~ of the unbalanced moment transferred by shear is given by

(2.6)

The remainder of the unbalanced moment is carried by flexure at the front face of the

column. The nominal shear strength of the connection is limited by

Vc =(2+~).[1: ~ 4..Jl: (psi) (2.7)


f3c

where f3c is aspect ratio of the column.


2-61

Recommendations for the design of slab-column connections in monolithic reinforced

concrete structures was reported by ACI-ASCE Committee 352 in 1989

(ACI 352.1R-89). The recommendations were limited to columns with an aspect ratio

less than 4 and with conventionally reinforced connections without prestress or shear

reinforcement. The recommendations were also limited to slab-column connections

constructed in regions of low seismic risk. The reduction factor rv recommended was as
given by equation 2.6 and a simplified approach based on the shear force was given for

the design of comer and the edge connections transferring moments perpendicular to

the slab edge. Guidelines for the detailing of reinforcement in slab-column joints were

given in the report.

Megally and Ghali (1999) proposed a model to design all type of slab-column

connections for shear. The model is based on the linear eccentric shear model of ACI

318-1995. In this model equations for the moment reduction factors Yvx and ')Ivy were

proposed in terms of the projections of the critical section on the principal axes of the

critical section, instead of the width of the critical section as used in ACI 318 (1995).

The proposed modifications are based on the results of the linear finite element analyses

of Elgabry and Ghali (1996). A further proposed change is the replacement of terms lx

and ly, which are defined as the terms analogous to polar moment of inertia, by the

moment of inertias lx and ly of the critical section about the x and y axes, respectively.

Design procedures with some examples were presented for a range of slab-column

connections.
2-62

2.3.4 Truss Model

Alexander and Simmonds (1986) and Simmonds and Alexander (1987) developed a

three-dimensional truss model to determine the ultimate strength of slab-column

connections. A mechanism was proposed to explain the punching phenomenon and

punching was described as failure of the slab to confine the concrete compression

forces out of the plane of the slab. The model was calibrated with the results of 48

interior slab-column connection tests under shear load and verified against the test data

of 43 edge column-flat slab connections without shear reinforcement reported by 8

investigators. The model was shown to predict well the failure of edge column-flat slab

connection without shear reinforcement with an average experimental to predicted

strength of 1.09 for slabs with no unbalanced moment and standard deviation of 0.11.

Similarly the mean for 23 shear-moment tests was 1.0. The truss model for the edge

column connection is shown in Figure 2.28.

In the eight tests isolated interior column connections reported by Alexander (1990),

Alexander and Simmonds (1992a) observed that in most of the specimens anchorage

failure of many reinforcing bars occurred before the actual punching failure although

visual inspection suggested a punching failure. It was also observed that arched struts

(see Figure 2.29b) form to carry the compression from the beam strips to the column

support, rather than straight struts (see Figure 2.29a) as was assumed in the earlier

works of Simmonds and Alexander (1987).

Strain measurements in the reinforcing steel in the Alexander (1990) tests showed a

curved arch profile (as shown Figure 2.30) rather than the straight line compression
2-63

Strut angle cx1

Uplift

Figure 2.28 - Truss model for edge slab-column connections


(Alexander and Simmonds, 1986).
2-64

bar

jd jd

1----Column face ---column face

(a) Compression strut model

jd

1 - - - Column face ---column face

(b) Test measurements defining compression strut

Note: re, n, ro =distances ofthe reinforcement forming the compression strut


jd = lever arm

Figure 2.29 - Struts assumed and observed by Alexander (1990) and Alexander
and Simmonds (1992b) (a) compression strut model; and (b) test
measurements defining compression strut.
2-65

Curved compression strut

jd

1---Column face

Figure 2.30 - Curved compression strut model of Alexander and


Simmonds (1992b ).

profile assumed in the model Alexander and Simmonds (1986). Therefore, Alexander

and Simmonds (1992b) modified the 1986 model to better describe the curved strut

profile observed in the tests. The model was developed by combining the radial arching

action effect with the critical shear stress on the critical section of the connections (also

referred to as beam-action shear). The bond strength of the reinforcement is considered

a limiting factor for beam-action shear for punching shear failure. The punching

capacity of the slab is determined by summing the contribution of all radial strips, with

the radial strips as shown in Figure 2.31. The total uniformly distributed load in the

radial strip 2w for the length lis determined from the primary shear, where wand l are

intensity of load for half of the strip and the loaded length of the radial strip,

respectively. Three different approaches were proposed to find the value of w, which

has a significant influence on the strength calculation. The model was compared with

115 data from eight investigations. The best mean test strength to predicted strength

was obtained with the value of w determined using the ACI code model. With this

model the mean of the test data to model prediction was 1.29 with a standard deviation

of0.16.
2-66

Lines of zero shear

Remote end of
radial strip

Reinforcemnt
direction

of
radial strip

Figure 2.31 - Layout of radial strips (Alexander and Simmonds, 1992b ).

Alexander (1999) further developed the bond model given in Alexander and Simmonds

(1992b). The loaded length of each radial strip was treated as a deep beam with arching

action, while the quadrant of the two-way slab was treated as for beam action

delivering an internal shear to the side faces of the supporting radial strip. The model

was described as a strip model. The comparison of the strip model with the

ACI 318-1989 code model suggested that ACI model was conservative with increasing

reinforcement ratios. Comparing the strip model with experimental data showed that, in

general, the model performed well. The use of this model for the design of interior

slab-column connection showed the lateral distribution of flexural reinforcement to be

an effective way to ensure adequate shear strength of the connections.

2.3.5 Improved Rotational Model

Shehata and Regan (1989) proposed a model for axisymmetric reinforced concrete

slab-column systems under concentrated loads. The model is based on test observations
2-67

and numerical analyses. The slab was divided into rigid segments, with each segment

bounded by two radial crack lines (refer Figure 2.32). The angle of inclination of the

internal crack surface was taken as 20° and the concrete in compression at the column

face was taken to be in a plastic state. The proposed model describes the slab behaviour

beyond the cracking stage and calculates both the punching and flexural capacities of

the slab. Comparing the results of their model with experimental data, Shehata and

Regan concluded that the model correlated well with the data.

2.3.6 Yield Line Analysis

In flat plate structures it is preferable that flexural failure takes place before the shear

failure at the slab-column connections. The flexural capacity of the slab-column

connection can be determined from the yield line analysis. Under combined shear and

unbalanced moment there are primarily two yield line patterns to be investigated. These

are 1) local mechanism in the slab forming fans of yield lines surrounding the column

and 2) folding type collapse mechanism with the yield lines extending the length or

width of the structure.

Gesund and Kaushik (1970) examined different punching shear tests to study the extent

of the validity of yield line theory in the outcomes of these tests. Expressions were

derived to find the strengths of the connections based on yield line pattern. The

arithmetic mean of the P.YieldlineiPtest calculated for 106 slab-column specimens was

found to be 1.02 with the standard deviation of 0.25, where PYieldline and Ptest are load

carrying capacity of a slab predicted by yield line theory and the load at which the
2-68

~------------------

(d-x4
a~
dO P 2n

Figure 2.32 - Punching failure model and forces involved


(Shehata and Regan, 1989).
2-69

specimen was reported to have failed, respectively. Gesund and Kaushik developed a

non-dimensional parameter, Q , to indicate the mode of failure of the slab. The

non-dimensional parameter is given by,

(2.8)

where p is the reinforcement ratio; h is the yield stress of reinforcing steel; d is the

effective depth; f~ is the compressive strength of concrete; b is the column perimeter;

and B is the perimeter of slab supported on one column.

For a value of Q < 2 bending capacity controls the strength of the slab-column system.

For 2 : : ; Q : : ; 4 the strength is controlled by either flexure or shear and for Q > 4 the

punching shear strength controls the design. It was recommended that the

non-dimensional parameter should be limited to Q < 2 to ensure a flexural failure.

2.4 Prestressed Concrete Slab-Column Connections

Post-tensioned prestressed concrete slabs are often used for the flat slab construction.

The design of a post-tensioned slab must ensure that the concrete stresses are not

excessive at service loads, the slab must be strong enough to resist the factored design

loads and the deflection of the slab should be within allowable limits.

Typical flat slabs have inherent ductility in flexure. A two-way slab is highly capable of

redistributing local loads by finding alternative load paths. Tests on real structures

(Ockleston 1955, Vecchio and Collins 1990, Collins and Mitchell 1991) indicate that
2-70

post-tensioned slabs have beneficial effect of compression membrane action which

significantly increases their strength.

Early tests on prestressed flat plates were undertaken on square slabs resting on

supports without monolithically cast columns. Scordelis et al. (1956) conducted a test

on a 127 mm thick, 4.6 m square slab with 10 by 19 mm diameter straight tendons in

each direction. The tendons were arranged to give a uniform prestress of 2.8 MPa. The

slab was supported on roller supported base plates at its comers and a ductile flexural

failure was reported.

Scordelis et al. (1959) tested a one-third scale model of a continuous prestressed

concrete flat slab with a parabolic cable profile. The slab was 75 mm thick and

consisted of two bays in each direction. The overall dimension of the square slab in

each direction was 4.6 m. The slab was supported on nine supports with rocker and

roller arrangements at each support to allow for sufficient rotations and horizontal

movements without any restraint. The average prestress was 1.1 MPa. After extensive

flexural cracking the slab failed at an applied load of 17.3 kPa with the centre support

punching through the slab at failure angle of approximately 45 degrees. This was the

first time punching shear was identified in a test series on prestressed slabs.

Since the early tests of Scordelis et al. (1956) a number of similar studies were

undertaken (Scordelis et al., 1958, Gerber and Bums, 1971, Smith and Burns, 1974).

However, tests undertaken in the earlier period did not have the column monolithically

cast with the flat plate. Some of these tests were for the simulation of lift slabs, which
2-71

are slabs cast on the ground and lifted up the columns, placed in position to the final

height and locked into place using a shear key inserted in the column.

Gerber and Bums (1971) reported tests on ten prestressed concrete slab specimens.

These specimens consisted of four cast in place and six lift slab specimens. All slabs

were 3.7 m square and 180 mm thick representing a half scale model of an interior

prototype flat slab. The slabs rested on 300 mm square columns. The total lengths of

the columns were 1380 mm including a projection of 600 mm above the slab. The

objectives of the tests were to study the shear and flexural capacities and to investigate

the effect of various types of reinforcement and the arrangement for crack control and

load carrying capacity. All the specimens were reported to fail in shear followed by

flexural failure with the yielding of the bonded reinforcement. The bonded

reinforcement played a significant role in the control of crack growth and in increasing

the ductility. Passing post-tensioning tendons over the column made the connection

capable of transmitting load to the column after the failure of the concrete adjacent to

the column. Analyses of the results by Gerber and Bums show that the ACI 318 (1963)

was conservative in calculating the shear transfer to the column.

Smith and Bums (1974) tested three one-third scale post-tensioned flat plate specimens

to failure. The slab consisted of a 2. 74 m square plate with a 200 mm square column

stub at the middle. The slabs were 70 mm thick with parabolic prestressing cables with

an average prestress of 2.2 MPa. The tests were conducted under symmetrically applied

line loads placed at 600 mm from the centre, as shown in Figure 2.33, via a

whiffle-tree. The main aim of the study was to investigate the effect of adding
2-72

channel sections

2 00

1200

I. 2700
.. I

c====L=in=e==·lo=a=d==~.-==~==L=in=e=~=lo=a=d==~~
t

Figure 2.33- Loading system for the specimens of Smith and Burns (1974).
2-73

bonded reinforcement to prestressed concrete flat plate systems. The shear capacity, the

flexural strength and general behaviour of the structures were analysed. All specimens

were reported to have failed in a combination of flexure and shear with the final mode

of failure being punching shear. In all tests the failure surface outlined a truncated

pyramid. The failure surface extended conically from the perimeter of the column at the

bottom surface of the slab to the perimeter on the top surface of the slab. The slope of

the surface was observed to be significantly less than 45 degrees. The inclusion of

non-prestressed bonded reinforcement increased the stiffness, ductility and load

carrying capacity of the slabs.

Bums and Hemakom (1977) conducted a series of multiple panel tests to study the

behaviour of prestressed concrete flat plate structures up to collapse for a range of load

cases. The safety of prestressed concrete slabs against punching shear failure was also

investigated. The variables studied were the distribution of cracking, the strength gain

due to inclusion of bonded non-prestressed reinforcement and the stress increase in the

unbonded tendons with the increase in load.

Trongtham and Hawkins (1977) report the results of tests on five unbonded

post-tensioned flat plate slabs. Four of these slabs represented interior slab-column

connections and one a typical exterior slab-column connection. The model slab

representing the exterior connection was loaded incrementally with shear and moment.

From their study Trongtham and Hawkins concluded that the ACI 318-77 approach

could be applied for calculating the capacity of connections transferring moment as well

as shear.
2-74

Hawkins (1981) reported tests on six full-scale unbonded post-tensioned prestressed

slab-column subassemblages. Out of the total six specimens tested, four were to

simulate interior column connections and one each to represent edge and interior lift

slab-column connections. The sizes of all interior column specimens and the edge

column specimen were 3.96 mx2.13 mx0.15 m and 2.13 mx 2.16 mx0.15 m,

respectively. All the specimens except the lift slab specimen consisted of 356 mm

square columns. The main objective of the study was to determine the strength and

stiffness of prestressed concrete plate-column connections transferring moments.

Bonded reinforcement was also provided in all specimens.

Hawkins reported that in their tests all the interior column specimens failed in punching

shear. The bonded reinforcement at the edge column connection had yielded before the

failure. Based on the test results, Hawkins recommended the provision of bonded

reinforcement at the discontinuous edge of prestressed concrete exterior column

connections when the shear stress in the critical section exceeds 0.17 .fj{. Hawkins

also recommended that tendons bundled through the column, or over the lifting collar,

increase the moment transfer capacity of interior column connections.

Cooke et al. (1981) published the results of tests on 12 prestressed concrete slabs, nine

with unbonded tendons and three with bonded tendons. The span to depth ratio was

varied in the study. The results were compared with the predicted values using

Pannell's method (Pannell, 1969) for calculating the flexural strength of slabs having

taken into account the effect of the span to depth ratio with a wide scatter of results

observed. The reasons for the scatter were attributed to the differences in the loading

arrangements, the duct material used for the tendons, the deviation of actual cable
2-75

profile from the proposed profile and inaccuracies in measuring techniques for the

prestress. Because of such uncertainties, they recommended to use the conservative

ACI 318 (1963) model to estimate the ultimate strength in prestressed concrete

structures.

Sunidja et al. (1982) published a report on the response of prestressed concrete

plate-edge column connections. They carried out tests to failure on four, two-third

scale, models of unbonded post-tensioned two way edge slab-column connections. The

specimens contained non-prestressed reinforcement as recommended by the

ACI-ASCE committee 423 (1974). Each of the specimens consisted of a 1.52 m square

prestressed concrete slab by 100 mm thick. A 305 mm square column was located at

the middle of one edge of the slab with the top and bottom ends of the column held in

position. The overall length of the column was 2.03 m. The specimens were reinforced

with 11 by 9.5 mm diameter tendons with an average prestress level of 1.65 MPa in the

slab. All the tendons were grouped in a narrow band over the column line in one

direction while the tendons in the orthogonal direction were uniformly distributed

(see Figure 2.34).

In the Sunidja et al. tests the closely spaced tendons were assumed to create beams in

one direction while the :flat plate acted as a one-way slab between these beams in the

transverse direction. In the first two specimens the banded tendons were perpendicular

to the direction of the exterior edge of the slab, while in the other two specimens the

bonded tendons were parallel to the exterior edge of the slab. The loads were applied

using a whiffle-tree at four points lying in a line parallel to the free edge and opposite
2-76

2-f_ 1520
-;Io
-t- 405
1 h _J_ 1
L I 15o 152~--- X
f-J I
-+5 280
-----.!-
1030
115
+
2_j

Figure 2.34 - Typical tendon arrangement for Sunidja et al. (1982) tests
(shown for specimens S3 and 84).

to the column. The distance of the line from the column was varied to vary the moment

to shear ratio transferred from the slab to the column. The results showed that the

ultimate bending moment transmitted to the column and the ductility of the specimen

decreased as the shear between the slab and column increased. The cracking load and

the strength of the connection were increased when the prestressing tendons were

concentrated in the vicinity of the edge columns.

Sunidja et al. (1982) compared the strength of their experimental connections with

different models such as the ACI 318-1977 model, the modified beam analogy of Park
2-77

and Islam (1976), the yield line method and with finite element analyses. The beam

analogy proved the best approach for calculating the strength of the tested prestressed

concrete slab-edge column connections. With the increase of shear transmission from

the slabs to the columns the ductility, as well as the ultimate bending moment

transmitted to the column, decreased. The crack development and its effect on ultimate

behaviour of an edge column connection were found to be a complex three-dimensional

problem.

Sunidja et al. remarked that punching remains a problem with no simple solution and

recommended that further research be undertaken with models that take into account

the three-dimensional nature of punching shear behaviour. Sunidja et al. concluded that

additional data and new experiments were required to understand the true behaviour of

such connections. They stated that:

"The development of finite element techniques that allow

three-dimensional cracking would provide a very powerful tool for

investigating the punching shear phenomena. "

Martinez-Cruzado et ai. (1994) conducted tests on four interior, two edge and two

comer column post-tensioned slab-column connections under simulated biaxial

earthquake loading. The tendons in the slab were banded in one direction and uniformly

distributed in the orthogonal direction. The slabs were reinforced with mild steel bars

satisfying the ACI 318-1989 minimum requirements for bonded reinforcement. It was

observed that the ACI 318-1989 and the Park and Islam (1976) models were

conservative for the design of interior connections. Repaired connections were also
2-78

tested and the efficacy of various retrofit and repairs strategies were identified. Using a

number of repair technique the retrofitted connections were found to be stronger than

the original connections.

Based on their test data, Loo et al. (1995) developed formulae for the punching shear

strength of post-tensioned flat plates with comer and edge columns with and without

spandrel beams. The shear strengths measured from the laboratory tests were compared

with various predicted strengths from various design models. Loo et al. claimed that

their procedure gave satisfactory and consistent results while the British standard

procedure was described as being conservative and the Australian and ACI approaches

as being unsafe.

2.5 Finite Element Analysis of Slab-Column Connections

Many investigations have been carried out to analyse the behaviour and the strength of

slab-column connections using finite element procedures. However, in the majority of

these studies the use of finite element method was limited to elastic section analysis due

to limitations on computational capacity. With the exponential growth in computing

power through the 1990s, 3D non-linear modelling of concrete structures has become a

powerful tool in analysis and design. In the case of slab-column connections, non-linear

modelling can supplement experimental studies to provide a better understanding of

load transfer mechanisms and behaviour of connections. In this section, finite element

studies undertaken to date on the load transfer from slabs to columns are summarised.
2-79

Long (1967) reported the results of theoretical analyses of punching shear problems of

slabs having flexural reinforcement in two directions and for slabs without shear

reinforcement. Elastic theory was used to formulate the finite elements with the failure

load determined based on an orthogonal shear criterion for the concrete. Several

approximations and correction factors were employed in the analyses. Materson and

Long (1974) carried out further studies on Long's work using finite element analysis

for plate bending. However, the analyses were limited to slabs in which the flexural

reinforcement yielded before the punching failure.

Mehrain and Aalami (1974) evaluated the lower and upper limits to the rotational

stiffness of the junction of the slab and the column of square flat slabs with centrally

located square columns. They conducted finite element elastic plate analyses of a

quarter of a plate using 12 by 12 rectangular elements. The results were comprehensive

and effectively expressed the rotational stiffness behaviour of flat slabs. Furthermore,

Mehrain and Aalami concluded that the boundary conditions parallel to the direction

considered did not significantly affect the slab stiffness and hence might be neglected in

computing the slab stiffness in the equivalent frame method.

Sunidja et al. (1982) carried out a linear finite element analysis of one of their

prestressed slab specimens. Sunidja et al. used four node flat rectangular elements

capable of handling flexural and membrane forces. Each node had six degrees of

freedom consisting of three translations and three rotations. A finer mesh was used in

the immediate area around the column and a coarser mesh away from the column. The

prestressed forces were defined to be uniformly distributed along the outer edge of the
2-80

elements. Since the analysis was limited to a linear analysis, only qualitative information

could be obtained about the behaviour of the tested specimen. Sunidja et al. stated that:

"As more sophisticated elements are developed that allow cracking due

to a three-dimensional stress state, this technique may be a very

powerful tool for investigating the punching shear phenomenon. "

It was also concluded that punching shear is a complex three dimensional problem

without simple solutions.

Barzegar et al. (1991) conducted a parametric study on edge slab-column connections

using 3D linear elastic finite element analyses. The pattern of one of the finite element

meshes used by Barzegar et al. is shown in Figure 2.35. As the column size was

increased, the connection became more rigid and the load was found to concentrate

more near the edge of the slab and comer of the slab-column interface. Barzegar et al.

also concluded that in the elastic range, the transfer mechanism of unbalanced moment

from the column to the floor slab may be assumed to be independent of slab aspect

ratio and of the size of the columns.

Zhou and Jiang (1991) applied the non-linear finite element method to analyse an

axisymmetric reinforced concrete slab with a concentrically located circular column

stud. The failure criterion and the equivalent stress-strain relation adopted for the

concrete was that of Ottosen (1977,1979). Nonlinearity of concrete under different


2-81

(!)

-
E
c
.....
Symmetry lines
+ -'
c ~----,,____ X
~
c 2u square
>
·:;
o-
w

Plan of the slab-column connection

-.--

.0

- ... ·-
Plane of symmetry
PLAN
Colu mn

Plane of symmebw

Plane of symmetry

FRONT ELEVATION SIDE ELEVATION

Mesh pattern

Figure 2.35- Plan of the slab-column connection analyses (Barzegar et al., 1991).
2-82

stress levels and post peak characteristics were incorporated using a secant modulus

formulation. The steel reinforcement was assumed to be elasto-plastic and perfect bond

between the steel and the concrete was assumed. Both radial and circumferential

reinforcement were considered as bar elements, with the radial bars along the sides of

the triangular elements and the circumferential reinforcement at the nodes. Although it

is said that bar elements were used to model the circumferential reinforcement, it is

more appropriate to say that point elements were used.

Zhou and Jiang modelled the slab-column connection of Kinnunen and Nylander (1960)

using the finite element mesh shown in Figure 2.36. Displacement control was used to

solve the problem with sufficiently small displacement increments used to allow for

stress redistribution within the slab. The predicted crack pattern and ultimate load were

close to the experimental results. The punching shear failure was concluded to be due

to the interaction of normal (compressive) and shear stresses over the failure surface. A

sketch of the analysed stress distribution pattern along the punching failure surface is

shown in Figure 2.37.

Guan and Loo (1994) implemented a non-linear finite element model for the failure

analysis of slab-column connections in the reinforced concrete flat plate structures.

Eight-node degenerate Mindlin shell elements were used having 5 degrees of freedom

per node (3 translation and 2 rotations). Shear deformations were not considered in the

study. The failure behaviour of the slab was studied using the three-dimensional

constitutive laws and the four-parameter failure criterion of Ottosen (1977). Each layer

of reinforcing bars was represented by a smeared continuous layer of an equivalent


2-83

~------------920------------~

~-----------865-----------.~

0
LO

450

Figure 2.36- Dimension and finite element mesh (Zhou and Jiang, 1991).

Applied load

Figure 2.37 - Sketch of calculated stress distribution over a conical surface at


ultimate load (Zhou and Jiang, 1991).

thickness with perfect bond being assumed. The results of the analyses implemented for

the comer column specimens of Zaghlool and de Paiva (1973) and for other slabs were

in good agreement with the experimental observations.

Guan and Loo (1995) modelled a plane stress problem using a layered finite element

model with transverse shear deformation. Cracking was modelled in the concrete using
2-84

a tension cut off model Cracking was taken to occur when a specified principal tension

was reached with the cracks forming in the direction normal to the principal tension

stress. After the crack, the principle stresses are released and the remaining stresses are

distributed. A maximum of two sets of orthogonal cracks were allowed to form at each

sampling (Gauss) point. Comparisons of the results using the deflections and the crack

pattern for the reinforced concrete slab were concluded to be satisfactory. The

proposed method was claimed to be capable of predicting the transverse and punching

shear failure of these slabs.

Guan and Loo (1996a, 1996b) used a 3D non-linear finite element model to study

punching shear failure behaviour of reinforced concrete flat plates with spandrel beams

or torsion strips. Five degrees of freedom (3 displacements and 2 rotations) elements

were used and a strain-hardening plastic material model (Owen and Figueiras, 1984)

used to represent the concrete. The reinforcing steel was smeared through layers within

the mesh. Shear deformations associated with the Mindlin hypothesis were also

considered. The closed ties were simulated in the analysis considering the contribution

of out-of-plane shear reinforcement. Thus, a full interaction between the spandrel beam

and the adjoining slab was considered.

Guan and Loo (1996b) compared 11 half-scale reinforced concrete flat plate specimens

and 4 single slab-column specimens. The results were compared with the

semi-empirical model of Loo and Falamaki (1992) as well as different design code

procedures. It was concluded that the Loo and Falamaki model is accurate and more

reliable than the code models.


2-85

Elgabry and Ghali (1996a, 1996b) proposed a new model for the calculation of the

transfer of shear stresses developed at the critical section of slab-column connections.

In developing this model linear 3D finite element analyses were used to fix the

unbalanced moment distribution factor for the moment to be transferred by the

eccentricity of vertical forces. Isoparametric Mindlin rectangular elements with eight

nodes on the middle surface were used for the finite element analyses. Each node had

six degrees of freedom (3 translational degrees of freedom and 3 rotations) with both

shear and bending deformations considered in the element formulation. The finite

element mesh shown in Figure 2.38a was used to model the comer column connection

shown in Figure 2.38b. In Figure 2.38 cy , cy and L are the dimensions of column in

the x and y directions and centre to centre distance between columns, respectively. The

nodal translation in the direction normal to the slab (w) and rotations ( Oy and By)

were prescribed for the column. Finer elements were used in the vicinity of the columns

in all models.

Elgabry and Ghaliexamined a total of 189 critical sections for interior, edge and comer

columns with different slab aspect ratios. The distance between the critical section and

the column face was varied between half the effective depth of the slab to eight times

the effective depth. From the analyses the elastic values for the component of

unbalanced moment to be transferred by shear was determined and denoted as 'Yv.elastic·

To arrive at more realistic results, however, a reduction factor of 0.15 was applied to

the value for distribution factor rv. that is rv =0.15rv,elastic. Curves were selected to
envelop the results of the finite element analyses for the plots of 'Yv verses
2-86

®y
wL®x
Nodal degree of freedom

Comer column Edge column


Free edges Q

"' F

.,"'~ ®y-=0_
.,"
E
"-

H
J I
- G
Edge column Interior column

(a) (b)

Figure 2.38 - Elgabry and Ghali (1996) (a) Finite element mesh and boundary
conditions for corner slab-column connections; and (b) top view
of flat plate showing corner panel.

l/lx where ly and lx are the length of the projections of critical section perimeter on the

principal axes. For internal columns ~ was determined to be a function of the shape of

the critical section. These results indicated that ACI 318-95 and CAN3-A23.3-M94

method used to calculate ~ at d/2 from column faces for all cases is incorrect.

Megally and Ghali (1996) developed a nonlinear finite element model to study the

fraction of unbalanced moment transferred as a shear (~) to the column by using

20-node isoparametric brick elements. The element had three translational degrees of
2-87

freedom at each node. The concrete constitutive model used was capable of

considering concrete cracking, crushing and post cracking shear degradation. Perfect

bond was used between the concrete and steel with an elastic-plastic stress-strain

model Based on their parametric studies it was shown that the equation developed by

Elgabry and Ghali (1996b) for %, combined with the linear equation given by

CSA-A23.3-94, was suitable to calculate the maximum shear stress distribution in edge

column connections.

Loo and Guan (1997) developed a non-linear layered finite element model to analyse

cracking and punching shear failure of reinforced concrete flat plates with spandrel

beams or torsion strips. In the model the full interaction between the spandrel beams

and the adjoining slab was considered. The transverse shear deformations associated

with the Mindlin hypothesis were also considered. A plasticity based incremental

stress-strain model (Owen and Figueiras, 1984) with smeared cracking was adopted for

the concrete with the contribution of the out of plane shear reinforcement and the

tension stiffening effect included.

Loo and Guan compared their results of the numerical model with test results and a

good correlation was claimed. It was also reported that the performance of the

proposed analyses was satisfactory and consistent for flat plates with either spandrel

beams or torsion strips. The approach of ACI 318-1989 and BS8110-1985 were

reported to overestimate the punching strengths to different degrees.

Polak (1998) examined the applicability of non-linear layered finite element modelling

based on the shell formulation, on the punching shear problem of reinforced concrete
2-88

slabs. The elements used had five degrees of freedom at each node (3 translations and 2

rotations). The uncracked concrete was treated as a linear elastic material and the

model formulation was based on the compression field theory of Vecchio and Collins

(1986). A smeared rotating crack approach was adopted. The model was implemented

on the slabs of Yamada et a1. (1992) and Elstner and Hognestad (1956) failing in shear

and/or flexure. Although the finite element model predicted the behaviour well, the

actual strength of the slab was underestimated by about 20 percent. This was attributed

to an underestimation of concrete contribution in its shear carrying capacity.

Hueste and Wight (1999a, 1999b) developed a 2-dimensional nonlinear finite element

model for investigating the behaviour of interior slab-column connections under

dynamic loading. The model could account the loss of rotational stiffness caused by the

punching damage as well as predicting the punching shear failure. The model was

applied to a four-story reinforced concrete framed office building that was damaged

during the Northridge earthquake. It was observed that the inclusion of punching shear

failure in the model modified the overall response of the building in terms of drift,

fundamental period and base shear distribution.

Megally and Ghali (2000a) observed that the fraction of unbalanced moment

transferred between the column and slab given as~ in ACI 318-1995 is adequate when

the critical section is considered at d/2 distance from the face of the column, where d is

the effective depth of the slab. Further, it was concluded that the expressions for ~

developed by Elgabry and Ghali (1996), using linear finite element analysis are

satisfactory for other critical sections.


2-89

Megally and Ghali (2000a) undertook a study on the behaviour of slab-column

connections using nonlinear finite element analyses. The concrete was modelled with

20-node isoparametric brick elements with three degrees of freedom per node. The

constitutive relation was based on the classical incremental theory of plasticity relating

the increments of plastic strain to the state of stress. Tension softening and shear

retention were included in the model for the cracked concrete. The steel reinforcement

was modelled using bar sub-elements embedded within the concrete elements.

Ozbolt et al. (2000) implemented a three dimensional nonlinear finite element model

based on the microplane model with a smeared crack approach to analyse interior

slab-column connections. The parametric studies showed that fracture energy and the

reinforcement ratio significantly influenced the punching load. However, the influence

of concrete compressive and tensile strengths were relatively small.

2.6 Conclusions

Numerous investigation have been undertaken to develop an understanding of the

behaviour of different types of flat plate-column connections. These studies have been

concentrated mainly on conventionally reinforced connections with most studies

focussing on laboratory tests with some analytical modelling. The size of the test

specimens varied from the very small (Dragosavic and van den Beukel, 1974) to the

relatively large multi-panel specimens (Robertson and Durrani, 1990, Gardner and

Shao, 1996).
2-90

In many studies the test specimens are developed as scale models of a prototype

structure ignoring size effects. However, the scale effects are yet to be verified. The

bu1k of the research has been for gravity loaded structures with a few more recent

studies considering both gravity and lateral loading. Besides the theoretical approaches

based on plate theories and beam analogy some research has been undertaken using

linear, and to a lesser extent, non-linear finite element modelling.

Code models have been developed mostly from research on interior column

connections with significantly fewer studies for edge and the comer column

connections. Similarly, studies on prestressed concrete flat plate-column connections

are limited in number. The literature review has highlighted a lack of data on

prestressed edge connections and no studies were found on the testing of prestressed

comer column connections.

Most of the previous studies are based on the laboratory tests on scaled specimens.

From these studies and other analytical investigations it is well understood that the

complex behaviour of all types of slab-column connections is truly a three-dimensional

problem. In some investigations the necessity and importance of development of

three-dimensional finite element-based models to study punching shear behaviour were

highlighted. However, only a few studies have been carried out using finite element

modelling.
CHAPTER 3 - FINITE ELEMENT MODELLING OF PUNCHING PROBLEMS

3.1 Material Modelling and Solution Procedure

3.1.1 Introduction

Different material models are used to represent the behaviour of the concrete and the

reinforcing steel at various stress states. In this thesis the model developed by Rots

(1988) and implemented in DIANA version 6.2 (1997) is used for numerical

modelling. A smeared tension cracking model for the concrete with a plasticity model

for compression is used and is discussed briefly in the following sections.

3.1.2 General Constitutive Relationships

For small displacements, without an appreciable change in geometry, th~ total strain e

can be decomposed into an elastic component(~) and an inelastic (plastic) component

(i'). That is,

(3.1)

The flow theory of plasticity is implemented to describe the elasto-plastic material

behaviour. The total stress at any time (t) is a function of the total strain e and it is also

a function of stress and strain history. The stress and strain history of the material is

taken into account implicitly by introducing an internal parameter tc, which is

governed by a specific evolutionary law. Here K'represents an equivalent plastic strain

of the concrete at different levels of stress. The elastic relationship between the total

stress a and the elastic strain ~ is given by


3-2

(3.2)

where

T
<1 = [axx <Yyy <Yzz <Yxy <Yyz azx] ; (3.3a)

ee = ~ix e
Eyy ee
zz ri;, e
ryz r:X] (3.3b)

The material elasticity relationship, D, is given by

E* E*
1 0 0 0
E* 1 E* 0 0 0
E(l- J.l) E* E* 1 0 0 0 (3.4)
D=
(1 + J.l)(1- 2f.l) 0 0 0 G* 0 0
0 0 0 0 a* 0
0 0 0 0 0 a*

(1- 2J.l)
where E is the elastic modulus, J1 the Poisson's ratio, E* = f.l and a*
(1- J.l) 2(1- J.l)

The yield condition governs the state of stress at the initiation of plastic flow. The

yield condition is a fu~ction of the stress vector and the internal state parameter (a

hardening parameter). If the value of the yield condition is less than zero, the system is

in an elastic state. A state with the yield condition greater than zero is not admissible

for rate-independent plasticity. Therefore, the yield condition is given by

f(o;7()=0 (3.5)

The inelastic strain rate vector is a function of the state of stress and the j-th plastic

potential functions gi =gi (o; 7() (j =l,n) and is given by

·P _ ~ 1 ag;
e -""A,·-
j=l J aa
(3.6)
3-3

The proportionality constant (ij) is a scalar such that ij ~ 0. If the yield function is

less than zero, no plastic flow will occur and ij = 0.

The evolution of the internal state parameter 7C is given as a function of the stress

vector cr and the plastic strain rate vector eP, ie. 7C = 7C(a,eP). Using the consistency

condition j =0 the relationship between the stress rate vector and the strain rate

vector is derived as

(3.7)

Concrete .possess a nonlinear volume change during hardening. Experimental results

indicate that under compressive loadings, inelastic volume contraction occurs at the

initiation of yielding and volume dilatation occurs at about 75 to 90 percent of the

ultimate stress (Chen and Han, 1987). Inflection points are usually observed and this

sort of behaviour violates the associated flow rule. A plastic potential function g other

than the loading function/is therefore required to define the flow rule. In such cases a

non-associated flow rule f '# g is adopted. For f '# g the tangent stiffness matrix. on

the right hand side of equation 3.7 becomes asymmetric.

3.1.3 Concrete in Compression

The compression behaviour of the concrete was modelled using a nonlinear

constitutive law. The uniaxial compression curve, shown in Figure 3.1, is represented

by a cohesion-equivalent plastic strain. Effects of confmement provided by triaxial

compressive stresses (if any) are modelled using the Drucker-Prager yield criterion.

The general yield condition of this model in principal stress space is


3-4

Stress

t:u = C:X1 ~r

Figure 3.1- Stress-strain curve for concrete in uniaxial compression and tension.

(3.8)

and P is the projection matrix, where

-1 -1 0 0 0
2
-1 2 -1 0 0 0
-1 -1 2 0 0 0
P= (3.9)
0 0 0 6 0 0
0 0 0 0 6 0
0 0 0 0 0 6

The scalars ar and j3 are given by


2sin if> (1() {3, - 6 cos l/>o
a
f
=3-sinl/>(1() o- 3-sinlf>o (3.10)

where if> is the angle of internal friction, l/>o the initial angle of internal friction,

c cohesion, c(t() equivalent cohesion. 1(, an internal state variable is an equivalent

plastic strain.
3-5

The Drucker-Prager yield condition with tension cut-off is shown in Figure 3.2. In this

figure it is shown that the deviatoric stress vector r ( r = ~2 J 2 ) makes an angle (J with

the first principal stress in the state of stress projected on the deviatoric plane. The

length along the hydrostatic axis (g) is given by~ = I d..f3 . The first invariant of the
stress tensor, /1 and the second invariant of the deviatoric stress tensor, hare given by

where Oi (i=1, 2, 3) are principal stresses and r', f are values of r and ~

corresponding to the tensile strength of the concrete, ft.

After yielding of the concrete a non-associated flow rule is used with the plastic

potential (g) given by,

(3.12)

with

(3.13)

and where 1fl( r<) is the dilatancy angle required to model volume changes due to shear

distortions. The dilatancy angle represents the ratio of plastic volume change over

plastic shear strain. For three-dimensional structures under low levels of confmement

lf/(7<) was found to affect the load carrying capacity (Vermeer and de Borst, 1984) but

its effect vanishes with high confining pressures and ranges from 0 to 20 degrees. The

relationship between the friction angle (l/J) and the dilatancy angle (lfF) is taken as
3-6

(a) (b)

Figure 3.2 - Drucker-Prager yield condition with tension cut-ofT (a) 1t-Piane

b) Meridian planes (} =0° and (} =n /3.

(3.14)

The relationship between the state variable and the rate of plastic strain in the minor

principal direction ( ef ) is given by

(3.15)

For the case of uniaxial compression, the equivalent cohesion, c, is

_ 1-a
1 (3.16)
c = fcp f3o

where fcp is the uniaxial compression strength of the concrete and CXJ and f3o are as

defmed by equation 3.10.


3-7

For concrete, the angle of internal friction is reported to be within the range of 30 to 35

degrees (Vermeer and de Borst, 1984), although in some cases angles as high as

l/J = 37 o have been used (Nielsen et al., 1978). In this thesis the angle of internal

friction was taken as l/J = 30 o • The cohesion-plastic strain curve for the concrete

(pre-peak) is based on the concrete stress-strain curve generated using Hognestad's

(1951) stress-strain model with a strain in the concrete at the peak uniaxial

compression stress of 0.002. In this thesis the dilatancy angle was taken as

lfl = 12.6°, as recommended by Vermeer and de Borst (1984).

3.1.4 Cracked Concrete

The multi-directional smeared crack model of de Borst and Nauta (1985) and

Rots (1988) is used for the cracked concrete. The total strain e is decomposed into an

elastic strain e and a crack strain ~r. that is

(3.17)

The strain vectors in equation 3.17 relate to the global coordinate axes and for the

three dimensional configuration they have six components. The global crack strain

vector ~r is given as

ecr _
-
r~xx
cr ecr
yy
ecr
zz
ycr
xy
ycr
yz
ycr] T
zx (3.18)

where x, y, and z refer to global coordinate axes. The crack stress-strain laws are

defmed with the local coordinate system aligned with the crack plane as shown in

Figure 3.3. The crack traction vector ecr in local coordinates is


3-8

Figure 3.3 - Local coordinate system and traction across a crack.

ecr _ f cr ycr ycr] T (3.19)


- ~nn ns nt

where e~~ is the normal strain for the mode I crack and r~~, r~[, respectively, are the

shear strains for the mode ll and mode m cracks. The relationship between local crack

strain ecr and global crack strain ecr is

(3.20)

where N is the transformation matrix

z2 lxly lz lx
X
m2 mxmy mzmx
X
n2 nxny nznx
N= X (3.21)
2lxmx lxmy+lymx lzmx +lxmz
2mxnx mxny+mynx mznx +mxnz
2nxlx nxly+nylx nzlx +nxlz

and where lx, mx and nx are the direction cosines for the strain normal to the crack

plane (n-axis) relative to the global co-ordinate system (similarly the direction cosines
3-9

with the subscript y refer to the s-axis and those with the subscript z to the local

t-axis).

The traction across the crack is defined by a vector r and is expressed as

(3.22)

where t~' is the mode I normal traction and t;' , tf are mode II and mode III shear
tractions, respectively. The relationship between the global stress and the local

tractions are given by

(3.23)

In the case of the multi-directional cracks that simultaneously occur at a sampling

point, the crack strain ecr is further decomposed as the sum of the global primary

crack ef' and the global crack strain increment owing to the secondary cracks, that is

e cr =e1cr +e2cr + ..... (3.24)

Each (fixed) crack is assigned its own local crack strain vector ef', its own traction

vector tf' and its own transformation matrix Ni according to equations 3.19 to 3.22.

The relationship between the crack-strain and traction for the multi-crack system is

(3.25)

where l' = ~f' t~' tg'] T ; D~2 :::


Df2 l
3-10

T
and ecr = ~f' e~r ...]

Finally the relationship between global stress a and global strain£ is given by

(3.26)

where De is stiffness matrix corresponding to elastic strain (Rots, 1988). The cracking

stiffness vcr is the corresponding matrix assembled for all the cracks formed at a

sampling point with Dcr for the n-th crack denoted as D~r and given by

EEt
0 0
E-Et
vcr- {3 E
n - 0 0 (3.27)
1 - {3 2(1 + Jl)
0 0 __L E
1- {3 2(1 + Jl)

where E and Et are the tangent moduli of elasticity for the ascending and descending

branches of the concrete stress strain relationship for tension. The Poisson's ratio is

given by Jl and {3 is a shear retention factor to account for the effects of aggregate

interlock and dowel shear. The transformation matrix for n cracks is

(3.28)

For concrete in tension a linear tension cut-off model was used with smeared cracking

added to the plasticity model discussed above. A crack arises if the major principal

tensile stress (a1) exceeds either the uniaxial tensile strength of the concrete or if

(3.29)
3-11

where C13 is the minor principal stress and f cp is the strength of the concrete. Multiple

cracking at a gauss point can occur provided that the principal tensile stress exceeds

the maximum stress condition and the angle between the principal tensile stress and an

existing crack exceeds a threshold angle a..

With the use of a threshold angle the number of possible crack formations at a point is

controlled between the fixed single crack model (a = 90 degrees) and the rotating

crack model (a= 0 degree). In this investigation a value of a= 15 degrees was used.

The uniaxial stress-strain curve used for modelling the concrete is shown in

Figure 3.1. In the case of plain concrete, crack propagation and fracture are primarily

dependent on the behaviour of the material in tension. It has been well established that

this response is primarily controlled by the formation of micro-cracks (Rots et al.,

1985). Upon loading, a limited numbers of such cracks can develop anywhere in the

specimen but if, at a point in the specimen, the tensile stress reaches the limiting

strength (/t) all additional deformation due to micro-cracking localises within the

fracture zone. In the fracture zone, the stress gradually decreases while the strain

increases. This phenomenon is known as tensile strain softening.

After cracking the stiffness of the reinforced concrete structure is reduced. However, a

concrete block between two adjacent cracks is capable of resisting tensile forces by the

bond between the concrete and the reinforcement. Due to the tension stress between

cracks caused by the influence of the bonded reinforcement a higher stiffness is

observed in the reinforced concrete as compared with the effect of the reinforcement

alone. This is referred to as tension stiffening.


3-12

A number of models have been developed to consider the tension stiffening effect. The

first approach to modelling tension stiffening was introduced by Scanlon (1971).

Scanlon used a modified stress-strain curve for the concrete in tension by assigning a

component to the descending branch to account for the retained stiffness upon

cracking. Gilbert and Warner (1978) artificially increased the stiffness of the

reinforcing bars by modifying the stress-strain response of steel This method

produced satisfactory results in different reinforced concrete slab analysed by Gilbert

and Warner (1978). However, its application in analysing orthogonally reinforced

concrete panels was not always as successful (Hu and Schnobrich, 1988). Lin et al.

(1975), Cope (1986) and Prakhya and Morley (1990) developed various approaches

based on experimental observations to incorporate the tension stiffening effect. Others

such as Petersson and Gustavasson (1980), Bazant and Oh (1983, 1984) and Sih and

DiTommaso (1985) developed models using a fracture energy approach.

In this investigation the tension softening and tension stiffening effects in the cracked

concrete were simulated using the simple linear tension-softening model, as shown in

Figure 3.1. The slope of the descending branch is adjusted to allow for tension

softening and tension stiffening using the model of van Mier (1987). The strain for

zero stress, post cracking ( eu) is expressed in terms of the cracking strain (ecr) such

that

eu =a1ecr =2Gf jh ft (for unreinforced concrete sections) (3.30a)

eu =alecr =az!yjEs (for reinforced concrete sections) (3.30b)

where Bcr = ft / Ec ; Gt is the fracture energy; h is the crack band width (equivalent
length of the mesh); ft is the tensile strength of the concrete and Es and Ec are moduli
3-13

of elasticity of steel and concrete, respectively. For solid elements h is calculated as

the cube root of the volume of the elements.

Different investigators have selected various shapes for the softening response of

concrete in tension and different strain levels for eu . In most cases, the selection of the

multiplier az depended on the experience of the analyst with the specific problem. For

example, Lin (1973) used a value of about az = 5, Gilbert and Warner (1978) and

Al-Manaseer and Phillips (1987) adopted a value of az = 10. Abdel-Rahman (1982)

assigned values between az = 10 and az = 25 and van Mier (1987) used

a1 =az Ec fsy where a2 is in the range of 0.3 to 0.8. Barzegar and Schnobrich (1990)
Es ft

presented different expressions for different problem types for ultimate cracking strain

using a nonlinear tension stiffening model similar to that used by Scanlon (1971). In

summary the tension stiffening effect is considered in different ways by different

investigators based on the experience of the investigator.

In this study, equation 3.30 is used with a2 = 0.5 for reinforcing bars with

fsy $; 550 MPa and a2 = 0.4 for bars with fsy > 550 MPa. Sensitivity tests for the

influence of eu are presented in Chapter 4.

The development of cracks in concrete reduces the shear stiffness but (not to zero)

because of aggregate interlock (Suidan and Schnobrich 1973, Channakeshava and

Iyengar 1988, Barzegar 1989). This effect is taken into account by the introduction of

a shear retention factor (/3). In this study a constant shear retention factor is used with
3-14

f3 = 0.2 unless stated otherwise. Sensitivity testing for the influence of f3 is presented

in Chapter 4.

3.1.5 Reinforcing steel

A bi-linear or tri-linear elasto-plastic model was adopted for the conventional and

prestressing reinforcement as shown in Figures 3.4a and 3.4b, respectively. In

Figure 3.4b, fsyi is the steel stress corresponding to the strain of Bsyi and Ep is the

tangent modulus of the steel stress-strain diagram in the post elastic region. Perfect

bond was assumed between the concrete and all conventional reinforcement and for

the prestressing steel when grouted prestressing systems were used.

3.1.6 Solution Procedure

Many solution procedures are available to solve the nonlinear finite element problems.

The solution procedure adopted is important as it influences the time required for

solution, disk space and the accuracy of the results. In nonlinear problems, the solution

procedure has to be carefully chosen in order to get the results efficiently without

losing accuracy of solution.

In this study, a modified Newton Raphson iteration procedure was used with the

stiffness matrix updated at the start of each load step. Load control was used in the

early (elastic) part of the analyses and load-displacement control via an updated

normal arc-length procedure was used in the latter part of the analyses. Except where

mentioned specifically convergence was set at 1 percent of the energy norm with a

maximum of 50 iterations for any one load step.


3-15

Stress, a

(a) Elasto-plastic bi-linear stress-strain curve

Stress, a

(b) Elasto-plastic tri-linear stress-strain curve

Figure 3.4 -Material model for steel (a) Elasto-plastic bi-linear stress-strain
curve; and (b) Elasto-plastic tri-linear stress-strain curve.
3-16

3.2 Sensitivity of Mesh Grading and Integration

3.2.1 Introduction

The accuracy of the solutions obtained using finite element modelling depends on

many factors, with mesh grading, the element aspect ratio and the integration scheme

being some of the most important. The effects of the element aspect ratio and the

number of Guassian integration points, both in plan and through the section, for

punching problems are discussed in this chapter. The analyses were carried out using

20-node solid isoparametric elements with Gaussian integration on a 2 x 2 x 2 or

3 x 3 x 3 quadrature.

3.2.2 Pullout of an Embedded Plate

To study the effects of aspect ratio on punching type problems a plate of uniform

thickness was modelled under a centrally applied load. Meshes of different aspect ratio

are considered. The slab, shown in Figure 3.5, consisted of a concrete plate of 1.7 m

square in plan and with a 100 mm square rigid plate embedded at a depth of 100 mm

from top surface of the slab. The depth of concrete was taken to be large compared to

the depth of the plate from the top and, thus, the base was modelled as restrained in the

z-direction. The rigid plate action was modelled using constant displacement for the

nodes representing the plate and it was taken that the plate is not bonded to the

concrete. The problem was investigated using 20 node isoparametric brick elements

with the aspect ratios shown in Figure 3.6. One quarter of the plate was modelled due

to symmetry. The meshes (shown in Figure 3.6) were developed to investigate the

effect of increasing the mesh size away from the critical zone. Details of the meshes

are given in Table 3.1.


3-17

y
j----850--t-f-----850-----j
... .. .. .... .. p

. .. .. ...

q~adre~~t :.us~d.
jfor. Ft' ·~nafysj~

I 1700 ~
....~ ·. :::.....
''..:.. X I
--T- oo- _j
L
1
Faoo----j r-aoo
100
SECTION 1-1

~ 50,•.------------------800----------------~~
<..: (8 @ 100)
.....
Q)
E I
~~~----+---~~---+-----r----+-----~---+----~
Ul

b ~~~--*-~~~?r-*~~~~--*-~~~?r~~~7
Q)
:§ p
Constant
displacement
TYPICAL SECTION FOR FE MODEL (Mesh 1E2)

Figure 3.5 - Square concrete mass under concentric load.


3-18

Mesh 1A2 Mesh 1B2

Mesh 1C2 Mesh 1D2

..< >-

>..

~ X ~
><><
~~'I: ::t:-.
~~ ~ w
/'
~
/'

~
Mesh 1E2 Mesh 1F4

Figure 3.6 - Finite element meshes used to study the effects of aspect ratio.
3-19

Table 3.1 - Properties of the different FE models having 2 and 4 elements through
the thickness.

ASPECT RATIOS NEAR THE


CRITICAL AREA*

Maximum Minimum
Model Total Total Total
Nodes Elements DOF x/y Max(x, y)/z x/y Min(x, y)/z

1A2 152 18 366 8 8 1 1

1B2 245 32 604 4 4 1 1

1C2 356 50 890 2 2 1 1

1D2 497 72 1260 2 2 1 1

1E2 1040 162 2694 2 2 1 1

1F4 5976 1734 16506 1 2 1 2

Note: * x/y is the aspect ratio of an element in the plan of the slab.
Max(x ,y)/z and Min(x,y)/z are the maximum and the minimum aspect ratios
through the thickness respectively.

For all the analyses undertaken in this study, the concrete is modelled with an in-situ

uniaxial compressive strength offcp = 25 MPa, a tensile strength offt = 1.65 MPa and

an initial modulus of elasticity of Ec =25300 MPa. The values of the tension softening

parameter a 1 (as defmed by equation 3.30) were calculated for each model and are

given in Table 3.2. The fracture energy was based on the CEB-FIP (1993) model code

recommendation and is taken as 85 Nm/m2 assuming maximum aggregate size of

32 mm for plain concrete. The other parameters needed to define the concrete material

laws were taken as


3-20

Table 3.2 -Values of <Xt of eu for different models.

Model 1A2 1B2 1C2 1D2 1E2 1F4

<X.t 10.5 13.3 15.4 15.4 15.4 40.0

¢=30° lfl= 12.6 ° a= 15 °


J-l =0.2 [3= 0.2 eo= 0.002

where J-l is the Poison• s ratio and other parameters are as defmed in Section 3.1 of this

thesis.

The results of the finite element analyses are presented in Table 3.3 and the load

versus deflection curves are shown in Figure 3.7. The responses in the elastic region

are similar for all cases, however. with the propagation of cracks the aspect ratio

adopted in the critical region significantly influences the results. In models 1C2 and

1D2, with larger meshes in the region away from the critical zone, the pattern of

load-deflection curves are similar to that of the uniformly meshed model 1E2. In all

cases the responses of the two layer mesh slabs are seen to be in the same range as

predicted by the refined 4 layered mesh 1F4 showing that some of the other meshes

are somewhat stiffer than mesh 1F4. The ultimate load for the two layered meshes are

close to each other with the difference between the extremes being about 10 percent of

the value predicted by mesh 1F4. It is concluded that away from the critical region the

mesh has little influence on the failure load and the load versus deflection response.

However, there is some effect on the post cracking behaviour with the increase in the
3-21

300

250

-,-
z
~
200

150 -+-Mesh 1A2


m -11- Mesh 1B2
0
..J
100 ........_Mesh 1C2
-a-Mesh 102
-o-Mesh 1E2
50
---Mesh 1F4

0
0 0.2 0.4 0.6 0.8 1 1.2
Deflection (mm)

Figure 3.7 - Load- deflection curves for modei1A2 to 1F4 (3 x 3 x 3 Gaussian


quadrature).

Table 3.3 - Ultimate load from the numerical analyses for the models with two or
four elements along the thickness.

Model 1A2 1B2 1C2 1D2 1E2 1F4

Ultimate load (kN) 249 217 212 240 233 218

number of elements through the thickness. It is to be noted here that the values of a.1

were based on equation 3.30 and a large variation occurs between analyses due to the

varying mesh sizes. Taking a.1 = 20 for mesh 1F4 reduces the ultimate load from

218 k:N to 188 kN and thus the fracture load is sensitive to the value of a.1 adopted.
3-22

The computation time required for nonlinear three dimensional analyses is an

important factor in the mesh sizing. The CPU times for each of the models is given in

Table 3.4. The analyses were performed on a DEC Alpha running UNIX with a single

300 MHz processor and 128 megabytes of RAM. Reviewing the analytical results

together with the CPU time, models 1C2, 1D2, and 1E2 are found to be the most

efficient. Of these three models, model1D2 is the cheapest with a minium of sacrifice

in the accuracy of the results. This model is chosen for further investigation.

Models similar to 1D2 were discretised with 1, 3 and 41ayers through the thickness of

the slab. The models are designated as 1D1, 1D3, and 1D4 and are shown in

Figure 3.8. Details of meshes 1D1 to 1D4 are given in Table 3.5 and the values of a1

used are presented in Table 3.6.

The ultimate loads obtained for slabs 1D1 to 1D4 are tabulated in Table 3.7. Except

for mesh 1D 1 the failure loads are in the same range with the maximum difference of

less than two percent. A variation of 22 % was obtained between modellD1 and 1D2.

The load versus deflection behaviour is plotted for each case in Figure 3.9 where it is

shown that similar results are obtained for meshes 1D2, 1D3, and 1D4. It is concluded,

therefore, that two elements through the thickness of the slab are sufficient to obtain

accurate results.

To study the influence of the order of integration on the results of finite element

analyses models 1D2 and 1E2 were analysed using 2 x 2 x 2 Gaussian quadrature

(designated as 1D2-2 and 1E2-2) and 3 x 3 x 3 Gaussian quadrature (designated as

1D2-3 and 1E2-3).


3-23

Mesh lDl Mesh 1D2

Mesh 1D3 Mesh 1D4

Figure 3.8 - Finite element meshes used to study the effects of different division
through the thickness of the slab.
3-24

350

300
250
--,
~ 200
as 150
0 --Mesh 101 - 1 layer
...1
-+--Mesh 1D2 - 2 layers
100 _..,_Mesh 103 - 3 layers
~Mesh 104 - 4 layers
50

0
0 0.2 0.4 0.6 0.8 1
Deflection {mm)

Figure 3.9 - Comparison of the load versus deflection curves for the lD mesh
with one to four layers through the thickness.

Table 3.4 - CPU times for Models 1A2 to 1E2 and 1F4
(3 x 3 x 3 Gauss point integration).

CPU time (s) for


Total Linear analysis Nonlinear analysis
Model DOF Absolute Relative to 1A2 Absolute Relative to 1A2
1A2 366 3.3 1.0 225 1.0
1B2 604 4.5 1.4 346 1.5
1C2 890 6.1 1.8 543 2.4
1D2 1260 7.8 2.4 704 3.1
1E2 2694 18.8 5.7 4312 19.1
1F4 16506 548 166 27160 121
3-25

Table 3.5 - Properties of models with different numbers of elements through the
thickness.

ASPECT RATIOS NEAR THE


CRITICAL AREA*

Maximum Minimum
Model Total Total Total
nodes elements DOF x/y Max(x, y)/z x/y Min(x, y)/z

1D1 315 36 754 2 1 1 0.5

1D2 497 72 1260 2 2 1 1

1D3 679 108 1766 2 3 1 1.5

1D4 861 144 2272 2 4 1 2

Note: * x/y is the aspect ratio of an element in the plan of the slab.
Max(x ,y)/z and Min(x,y)/z are the maximum and the minimum aspect ratios
through the thickness for the , respectively

Table 3.6 - Values of a1 of eu for different models.

Model 1D1 1D2 1D3 1D4

€lt 12.1 15.4 17.7 19.3

Table 3.7 - Ultimate load from the numerical analyses with different
numbers of elements through the thickness.

Model Ultimate load (kN)

1D1 293

1D2 240

1D3 243

1D4 242
3-26

The load versus deflection curves for mesh 1D2-2 and 1E2-2 are compared in

Figures 3.10 and 3.11 with the corresponding curves obtained using the 3 x 3 x 3

Gaussian integration (that is meshes 1D2-3 and 1E2-3). The figures show that the

order of integration scheme has no significant effect on the numerical results. The

ultimate loads for each analysis are given in Table 3.8.

The CPU time for each of the analyses are shown in Table 3.9 and it is seen that

reducing the integration order from 3 x 3 x 3 to 2 x 2 x 2 reduced the CPU time by a

significant amount. Thus, it is concluded that the reduced 2 x 2 x 2 integration gives

adequate accuracy and efficient solution times.

3.2.3 Upper bound solution

The model studied in Section 3.2.2 of this thesis is similar to that of punching shear in

plain concrete. In order to assess the validity of the model the results are compared

with the plasticity solution of Nielsen et al. (1978).

Nielsen et al. proposed a plasticity model for the pullout of circular discs in concrete.

The failure surface is formed by a solid of revolution, with the remainder of the slab

remaining rigid. The failure surface that gives the lowest failure load P, as shown in

Figure 3.12, is given by

r = do + x tan¢ foro:::;; x:::;; ho (3.3la)


2

(3.3lb)
3-27

300

250

--
z
~

"C
200

150
as
0
..J
100
~Mesh 102- 3 x 3 x 3 Gaussian integration
50 -+-Mesh 102- 2 x 2 x 2 Gaussian integration

0
0 0.2 0.4 0.6 0.8
Deflection (mm)

Figure 3.10 - Load versus deflection curves for modei1D2 with different
integration schemes.

300

250

--
~
"C
200

150
as
0
..J
100
~Mesh 1E2 - 3 x 3 x 3 Gaussian integration
- Q - Mesh 1E2 - 2 x 2 x 2 Gaussian integration
50

0
0 0.2 0.4 0.6 0.8 1
Deflection (mm)

Figure 3.11- Load versus deflection curves for slab 1E2 with different integration
schemes.
3-28

do

Figure 3.12 - Failure surface (Nielsen et al., 1978).

Table 3.8 - Ultimate loads for slabs 1D and 1E with different sampling points.

Model Ultimate Load (kN) Gaussian Quadrature


102-2 208 2x2x2
102-3 240 3x3x3
1E2-2 250 2x2x2
1E2-3 233 3x3x3

Table 3.9 - Solution time for slabs 1D and 1E for different integration
quadrature.

CPU time (seconds)


Model TotalDOF Linear analysis Nonlinear analysis

102-2 1260 5.8 215


102-3 1260 7.8 704
1E2-2 2694 14.4 724
1E2-3 2694 18.8 4310
3-29

wherec = ~a 2 -b2 , do is the diameter of the rigid plate; ¢ is the angle of internal

friction for the concrete; and ho is the depth of conical part in the failure surface. The

function generated by equation 3.31 consists of a catenary curve, joined by a straight

line and is constrained by the boundary conditions

a= do +hotan¢ (3.32a)
2

!!_=tan¢ :::} b = a tan¢ (3.32b)


c ~l+tan2¢

dt = a cosh h - ho + b sinh h - ho (3.32c)


2 c c

The pullout load, P, is calculated by minimising the function

P HV
2
fcp [ ho\do+hotarup
( co~ +lc(h-ho )+l
)1-sincp ~d1
(d12Vl2) 2 J-m ((dl2)1-a2J~~
1-c--ab (3.33)

where v is the effectiveness factor; f cp is the compressive strength of the in-situ

concrete; d1 is maximum diameter of the failure surface; and l and m are the material

constants and expressed as

l=1-(k-1).1L (3.34a)
fcp

m = 1- (k + 1) ft (3.34b)
fcp

where

k = 1+sin¢ (3.34c)
1-sin cp
3-30

The ratio J;jcp is notionally based on the ratio of the tensile strength to the

compressive strength of the concrete, but, in reality is a calibration factor used to

correlate the dimension d1 with experimental observations. Nielsen et al. (1978) found

that a value 400 :s;; C1(o.t :s;; 1000 gives the best fit with the experimental data. In

equation 3.33 the efficiency factor, v, is introduced to account for the imperfect

assumption that concrete is a perfectly plastic material and to calibrate the model with

experimental data. Based on their observations assuming a square root dependence on

concrete strength, Nielsen et al. (1978) proposed that

v p;-.
= 4.22 < 10 (3.35)

In verifying the model with experimental data, Nielsen et al. showed a reasonable

mean strength correlation but with a coefficient of variation as high as 21 percent. The

coefficient of variation, however, was generally smaller when individual series were

considered.

Nielsen et al. (1978) also recommended a simplified relationship to calculate the

average shear stress through the depth of the section taking (} = 26° to the soffit of the

slab (refer Figure 3.12). This leads to

p
-r =----;------ (3.36)
n(do+2h)h

If a lower bound to the failure surface is taken at(} =45° (as per ACI 318, 1999 and

AS 3600, 1994) then the shear stress is given by


3-31

p
't' = ---,-----.,.- (3.37)
tr(do+h)h

With an angle of internal friction of t/J =30°, taking do =}n de where de is the depth
of a square column (that is taking an equivalent circular area) and with -r =0.33..[1;,

h = 100 mm and de = 100 mm, by equations 3.36 and 3.37 the bounds of the ultimate

load are

110 kN ~ Pu ~ 162 kN

With the Nielsen et al. refilied model given by equations 3.31 to 3.35, and with the

bounds of {t
Jcp
= 400 and 1000, the failure load for the pullout of a 100 mm square

plate at the depth of 100 mm into the concrete are

184 kN~ Pu ~ 193 kN

The fmite element results (given in Table 3. 7) are of the order of 25 percent higher

than the model proposed by Nielsen et al. (1978) and suggest that taking a 45° failure

surface (as given by ACI 318 and AS 3600) is overly conservative.

3.3 Influence of Flexural Reinforcement on Punching Strength

3.3.1 Introduction

Dragosavic and van den Beukel (1974) and Park and Gamble (2000) established that

the punching shear strength of flat slabs is not sensitive to the reinforcement content.

In this section the influence of the reinforcement content on the punching shear
3-32

predicted by the finite element modelling is assessed. The effect of the integration

scheme on the results of the finite element models is also considered. For this purpose

the behaviour of a concentrically loaded reinforced concrete slab, resting on the central

square column, with varying amounts of steel is investigated. The results are compared

with the results obtained from the punching predictions of the Nielsen et al. (1978)

model and with flexural theory.

3.3.2 Effect of Negative Moment Reinforcement on Punching Strength

To study the effect of flexure-shear interaction a 150-mm thick slab resting on a

square concentric column (shown in Figure 3.13) is considered. The slab is uniformly

loaded and is reinforced with 21 bars at 100 mm spacing in each direction. The depth

to the centroid of all reinforcements (in both directions) was taken as 40 mm from the

top surface.

The slab, shown in Figure 3.13, was modelled using 363 by 20-node isoparametric

solid elements using the mesh shown in Figure 3.14. One quarter of the slab is

modelled due to symmetry. The reinforcing bars were modelled as embedded line

elements within the concrete elements. The CEB-FIP Model Code (1990) states that

dowel effects are not significant when only flexure reinforcement is used due to the

thin layer of concrete above the steel. The finite element formulation used here does

not allow for any dowel shear. The finite element model, Model 2, consisted of 2064

nodes for the concrete with 44 additional nodes used to model the 22 embedded

reinforcing bars in the concrete and include the two ~ area bars at the lines of

symmetry. The bottom face of the central element was restrained against vertical

movement to simulate a stiff column support.


3-33

1r'--1050---+---1050---,r{

l
I

~ ~ ....
21 Bars in each directions
~~~~~~~~
'-------..,.-t-.,r------40-~ X
1

*
150T

.j'-100
COLUMN REACTION

Figure 3.13- Reinforced slab resting on square column.

z
y X

Plane of symmetry Plane of symmetry

Mesh for Model2

Figure 3.14- FE mesh for investigating flexure-shear interaction of reinforced


concrete slab-column connections -Model 2.
3-34

The material properties used for the concrete were

fcp= 25 MPa /t= 1.65 MPa Ec = 25.3 GPa.

a1 = 15.3 /3= 0.2 l/J = 30 0

1jl= 12.6 ° a= 15 ° J1 = 0.2

eo =0.002

An elastic plastic stress-strain model was used to model the reinforcing steel with

perfect bond assumed between the reinforcing steel and the concrete. The yield

strength of the bars was taken as 400 MPa and the elastic modulus as 200 GPa.

Analyses were undertaken for reinforcement ratios of p = 0.0045, 0.0101, 0.0140,

0.0182, 0.0235, 0.0285 and 0.0318 with the ratio being calculated using the effective

depth of 110 mm. The governing yield line pattern is as shown in Figure 3.15. The

flexural strength of the slab is found to be

Pflex =6.36m (3.38)

where m is flexural strength of a unit width and is approximately given by

m =0.9 p fsy d 2 ; p =Ast I bd ; and fsy, Ast, b, and d are the yield stress of the

reinforcing bars, the bar area, the width of the panel and the effective depth from the

extreme compression fibre to the centroid of tension force, respectively. In

determining the punching shear capacity of the slab, the dowel effect was not

considered and, hence, the punching shear capacity is not a function of amount of

flexural reinforcement.
3-35

1~-"~•f---1 oso----r---1 oso~------~·~1

/
I
/
2100

~-----210D---------~

- - Negative moment yield line


-------- Positive moment yield line

External boundary: Free on all sides

Figure 3.15 - Failure pattern for yield line method.

3.3.3 Finite Element Results

The ultimate loads determined for different percentages of longitudinal steel are

presented in Table 3.10. With the increase in the reinforcement ratio the ultimate load

is also increased but at a decreasing rate with the increasing reinforcement content.

The load deflection curve for each of the analyses (with numerical integration over 27

Gauss points) is presented in Figure 3.16. In the elastic region the load versus

deflection responses of the slab are similar, but significant variation is observed in the

post-cracking stage with the post-cracking response being a function of the

reinforcement content.

At the punching shear limit as predicted by the Nielsen et al. (1978) model, the slab

fails in a non-ductile mode and the volume of reinforcement has only a small influence
3-36

400 -0.45%
--+- 1.01 o/o
350 ___._ 1.40 o/o
300 ------ 1.82 o/o
2.35 o/o
z 250 --8-

-
~
,
ca
200
---+-- 2.85 o/o
--&-- 3.18 o/o

.9 150
100
50
0
0 5 10 15 20 25
Deflection of symmetrical edge node (mm)

Figure 3.16 - Load- deflection curves for different reinforcement contents.

Table 3.10- Ultimate load from the FEM for Model 2.

Reinforcement Ratio, p (%) Ultimate Load (kN) Failure Mode

0.45 179 Flexural

1.01 271 Flexure Shear

1.82. 287 Punching shear

2.85 320 Punching shear

3.18 336 Punching shear


3-37

on the failure load. The load versus strain in the reinforcement near the column for the

different analyses is shown in Figure 3.17. The figure shows that ductile failures, as

indicated by yielding of reinforcement, occurred for the lower reinforcement contents.

The shear capacity of the slab can be obtained using the model of Nielsen et al. (1978),

as discussed in Section 3.2.3 of this thesis. The flexural capacity is obtained using

yield line theory for the critical yield line pattern as shown in Figure 3.15. The failure

load for the slab versus reinforcement content is plotted in Figure 3.18 and it is shown

to compare favourably with the theoretical model.

For the slab with reinforcement ratios of 0.0101, 0.0182 and 0.0318 further analyses

were undertaken using integration on a 2 x 2 x 2 Gaussian quadrature. The load versus

deflection results and load versus maximum steel strain for these analyses are shown

in Figures 3.19 and 3.20. Ultimate loads are presented in Table 3.11. The figures show

that the order of integration has little influence on the results of the finite element

analyses. The CPU time required for the analyses is given Table 3.12 and it is seen

that the integration on the reduced 8-point integration is 30 percent more time efficient

than that for the analyses on the full27-point quadrature.

Table 3.11 - Ultimate load for the analyses with different Gauss sampling points.

Reinforcement Ultimate load (kN) Gaussian Quadrature


Ratio, p (%)
1.01 281 2x2x2
271 3x3x3
1.82 305 2x2x2
287 3x3x3
3.18 376 2x2x2
336 3x3x3
3-38

-11-0.45%
400
-+- 1.01%
350 Yield Strain .........,_ 1.40%
~ 1.82%

300 ~2.35%
--+- 2.85%
-250
z
-
~

, 200
as
0
~ 3.18%

...J 150

100

50
0
0 5 10 15 20 25
Maximum strain in reinforcement over the column (E-03)

Figure 3.17 - Load- strain developed in steel for various reinforcement contents.

400
350
• • •
300

z 250 Punching shear capacity

-
,ca 200
~

0
(Nielsen et al., 1978)

...J 150
100
Flexural capacity (Yield line theory)
50 • FE results

0
0 0.5 1 1.5 2 2.5 3 3.5
Reinforcement content in one direction (%)

Figure 3.18 - Ultimate load versus reinforcement content.


3-39

400
350
300
-250
-
~
"0
C\1
0
200
-s- 1.01 o/o(2 2 2)
---- 1.01 o/o (3 3 3)
...I 150 ~ 1.82 o/o (2 2 2)

100 --+-- 1.82 o/o (3 3 3)


~ 3.18 o/o(2 2 2)
50 _._ 3.18 o/o (3 3 3)
0
0 5 10 15
Deflection of symmetrical edge node (mm)

Figure 3.19 - Load deflections curves found from 3 X 3X 3 and 2 x 2 x 2 gaussian


integration rule.

400
350

300

z 250
-
~

"0
C\1
0
200 -s- 1.01 o/o (2 2 2)
---- 1.01 o/o (3 3 3)
...I 150
~ 1.82 o/o (2 2 2)

100 --+-- 1.82 o/o (3 3 3)


~ 3.18%(222)
50 _._ 3.18 o/o(3 3 3)

0
0 1 2 3 4 5 6
Maximum strain in reinforcement over the column (E-03)

Figure 3.20 - Load versus steel strain curves found using 3 x 3 x 3 and 2 x 2 x 2
integration rules.
3-40

Table 3.12 - CPU time used for nonlinear analyses with 2 x 2 x 2 and 3 x 3 x 3
Gaussian integration.

Reinforcement CPU time (seconds)


Ratio, p(%)
2 x 2 x 2 Quadrature 3 x 3 x 3 Quadrature

1.01 3558 4278

1.82 3325 5187

3.18 3910 5648

3.4 Sensitivity Analyses of Finite Element Modelling Parameters

3.4.1 Introduction

The behaviour of cracked reinforced concrete structures is complex. While micro and

fracture models are becoming more widely available to study the mechanics of

concrete cracking, they require intensive computing resources. At present, these

models have limited applications to large scale concrete structures. Other less

numerically demanding, smeared and discrete crack models are available. The 3D

element in DIANA is such a model, but it requires the calibration of a number of

parameters. The element response is modelled by using factors for tension stiffening,

compression softening and shear retention. In this study, a linear tension softening

model was used, as discussed in Section 3.1.4. The tension softening model depends

on three factors, the tensile strength of the concrete ft, the initial elastic modulus, Ec,

and ultimate cracking strain of the concrete for zero stress, eu . Of these parameters eu

is selected using van Mier's (1987) equation (equation 3.30). In the analyses
3-41

performed in this study a shear retaining factor of f3 = 0.2 was used unless stated

otherwise.

3.4.2 Sensitivity to the Concrete Tensile Strain at Zero Stress, eu.

With the development of cracks, the stiffness of the concrete reduces with the rate of

stiffness degradation given by the slope of the tension softening curve. In reinforced

concrete structures, an increase in the slope of the concrete constitutive model over

that of concrete in uniaxial tension is needed to account for the tension stiffening

effects.

In this study the concrete tension strain at zero stress is taken as

(3.39)

where ecr is the cracking strain and is taken as e = %c .The uniaxial tension strength
is taken as ft =0.33..[1; wherefcp is the in-situ concrete strength (ACI-318, 1999).

To study the sensitivity of the finite element modelling to eu the reinforced concrete

plate modelled in Section 3.3.2 with p = 2.85 percent was analysed for different values

of a1. All other parameters were unchanged and the shear retention factor was taken as

f3 = 0.2.

The results of the analyses are shown in Figures 3.21 and 3.22 where the load versus

deflections and load versus maximum steel strain, respectively, are shown. The

parameter a1 is found to have significant effect in the deflection response of the slab
3-42

450
400
350
-300
z ---Aipha1 = 31
~ 250 ~Aipha1 =24
"D
as 200 ~Aipha1 =15
0
-+-Aipha1 =8
...I
150
_._Aipha1 =2
100
-ir-Aipha1 =0
50
0
0 2 4 6 8 10
Deflection (mm)

Figure 3.21 - Load versus deflection curves for model 2 with p = 2.85 % for
different values of a1.

450
400
350
-300
z
~ 250 ---Aipha1 = 31
"D
as ~Aipha1 = 24
0 200
...I
150
~Aipha1 =15
-.-Aipha1 =8
100 _._Aipha1 =2
50 -ir-Aipha1 =o
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain for central steel near the column (E-3)

Figure 3.22 - Load versus strain in steel near the midspan for model 2 with
p = 2.85 % for different values of a1.
3-43

and on the failure load. The development of steel strains are also affected by the

selected values for a1 in the range from cracking of concrete to the point where the

tension stress in the concrete is zero. The load versus strain curve converge to a

common line once the tension stiffness of the concrete is zero.

3.4.3 Sensitivity to the Shear Retention parameter, {3.

With the development of cracks, the shear stiffness of the concrete is reduced but not

. to zero due to the effects of dowel shear and aggregate interlock. The shear retention

factor f3 is used to calibrate the shear capacity for cracked concrete, however, its exact

calibration is a function of the aggregate type, the concrete strength, crack widths and

spacing and the amount of reinforcement crossing the cracks. As such, a realistic

mathematical model for this factor is yet to be determined.

Although the importance of this factor is recognised in the nonlinear finite element

simulation of the concrete structures, no consensus is found amongst researchers on a

mathematical model Some studies have taken a constant value, while others use a

variable factor based on the tension strain. In studies by van Mier (1987),

Channakeshava and Iyengar (1988), Barzegar (1989) constant values for f3 were used
varying between f3 = 0.1 and f3 = 0.5. Other studies such as those by Al-Manaseer and
Phillips (1987), and Foster (1992) used a variable shear retention factor model In this

section the sensitivity of the finite element results is examined for different values of

{3. The other parameters were kept constant.

The reinforced plate subjected to punching shear shown in Figures 3.13 and 3.14 were

used for the study with a reinforcement content of p =2.85 percent. The results of the
3-44

analyses for load versus deflection and load versus maximum strain in the

reinforcement are plotted in Figures 3.23 and 3.24, respectively. The results of the

analyses show that the trend of the load deflection response is not affected by {3,

however, the ultimate load is influenced by the value of f3 selected. The observations

from the load versus strain response in the reinforcement, shown in Figure 3.24, are

similar to those for the load versus deflection response. Thus the value of f3 adopted,

while not influencing the overall stiffness behaviour of the slab, has significant effect

on the failure load.

3.4.4 Sensitivity to the Friction Angle, ;.

An angle of internal friction of 30 to 35 degrees (Vermeer and de Borst, 1984) is

commonly used for concrete. However values as high as 37 degrees

(Nielsen et al., 1978) have been adopted. In this section sensitivity analyses are

undertaken to determine the influence of¢ on the finite element results. Values of

¢ = 25, 30 and 35 degrees were used for the parametric study.

As for the previous sensitivity studies the reinforced concrete plate shown in

Figures 3.13 and 3.14 with p = 2.85 percent were used for the analyses. The other

parameters were taken as

az = 15.3 f3 =0.2 ljl= 12.6 °


a= 15 ° J1 = 0.2 eo= 0.002
The results of the analyses for load versus deflection and load versus strain in the

reinforcement are presented in Figures 3.25 and 3.26, respectively. The analyses show
3-45

500
450
400

--
~
"D
350
300
250
---Beta = 0.99
~Beta= 0.5

-+-Beta = 0.2
"'
0
..J 200 -Beta=0.1
150 __.__Beta = 0.05
-.tr- Beta = 0
100
50
0
0 2 4 6 8 10 12
Dflection (mm)

Figure 3.23 - Load versus deflection curves for model 2 with p = 2.85 % for
different values of {J.

500
450
400

-,- 350
~ 300
250
-.Beta = 0.99
~Beta =0.5

"'
0 200
..J
-+-Beta = 0.2
150 -+-Beta = 0.1
~Beta= 0.05
100
-A-Beta= 0
50
0
0 0.5 1 1.5 2

Strain in central steel near column (E-3}

Figure 3.24 - Load versus strain in the steel near the midspan for model 2 with
p = 2.85 % for different values of f3.
3-46

400
350
300
--
z.:.= 250
"D 200
m --Phi = 35 Degrees
0
..J 150 -a- Phi = 30 Degrees
........,_Phi = 25 Degrees
100
50
0
0 2 4 6 8
Deflection (mm)

Figure 3.25 - Load versus deflection curves for model 2 with p = 2.85 % for
different values of ;.

400
350
300
z.:.= 250
-
"D
m
0
200
--Phi = 35 Degrees
..J 150
-a- Phi = 30 Degrees
100 ........,_Phi = 25 Degrees

50
o~~~~~~~~~~~~~~~~~~~~~~

0 0.2 0.4 0.6 0.8 1 1.2


Strain for central steel near the column (E-3)

Figure 3.26 - Load versus strain in steel near the midspan for model 2 with
p =2.85 % for different values of tfl.
3-47

that for the range of ifJ examined the general response of the slab is unaffected. There is

a small influence, however, on the failure load.

3.4.5 Sensitivity to the Dilatancy Angle, 1/1.

In the non-associated plasticity model the dilatancy angle lfl is always less than the

friction angle ifJ and reduces to zero for concrete subject to high confming pressure.

Vermeer and de Borst (1984) adopted a value of 0° ~ lfl ~ 20°. A value of 12.6° was

used in this study as recommended by Vermeer and de Borst. For the sensitivity test

the analyses was carried out for values of lfl = 1°, 12.6° and 18°. The other parameters

were fixed at

0
az = 15.3 /3= 0.2 ifJ = 30.

a= 15 ° J1 =0.2 eo=0.002

The reinforced concrete plate given in Figures 3.13 and 3.14 was again used for the

analyses with p = 2.85 percent. The results of the analyses for load versus deflection

and load versus strain in the reinforcement are presented in Figures 3.27 and 3.28,

respectively.

The analytical results show that the trend of load deflection and the load versus strain

response in the reinforcement plot are not affected by lfl, however, the ultimate loads

are influenced to some extent, for lflless than 12.6°. Thus the value of lfl adopted,

while not influencing the pattern of general response of the slab, has some marginal

effect on the failure load.


3-48

350

300

250
--,i 200
_._Psi= 18.0 Degrees
ca -a-Psi= 12.6 Degrees
0 150
..J --+-Psi = 7.0 Degrees
100

50

0
0 1 2 3 4 5 6 7
Dflection (mm)

Figure 3.27 - Load versus deflection curves for model 2 with p = 2.85 % for
different values of Vf.

350
300

--,i 250
200
ca 150 _._Psi= 18.0 Degrees
0
..J -e- Psi = 12.6 Degrees
100 --+-Psi = 7.0 Degrees

50
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain in central steel near column (E-3)

Figure 3.28 - Load versus strain in steel near the midspan for model 2 with
p = 2.85 % for different values of Vf.
3-49

3.4.6 Sensitivity to the Compression Softening Slope, Ect·

The slope of the softening branch of the concrete stress-strain curve, Ec1, is given by

Ecf =~c. In Hognestad (1951) model Ec" =- fcm and for /em= 25 MPa this gives
':1 :1 0.012

Ecf = -2083 MPa and ( = -12.

The sensitivity test of the compression softening slope Ecf in the results of the

nonlinear fmite element analyses was conducted using Model 2 with a reinforcement

content of 2.85 percent. The other parameters were set at

a1 = 15.3 ~=0.2 l/J =30. 0


a= 15 ° J1 = 0.2 Eo= 0.002

The study was undertaken for ( =-24, -12, and -6. The results of the analyses for load
versus deflection and load versus strain in the reinforcement are presented in

Figures 3.29 and 3.30, respectively. No effect is observed in the results for varying Ecf·

3.5 Conclusions

In this chapter, mesh grading is shown to have significant influence on the results of

the fmite element analyses. The integration scheme adopted is important in terms of

the solution time required for the analysis but has little effect on the overall results.

The finite element model predictions are in reasonable agreement with the closed form

plasticity model of Nielsen et a1. (1978) and in close agreement with the
3-50

350
300

--,
~ 200
250
-a-Zeta= 24

co 150 -Zeta=12
0 -tr-- Zeta = 6
...1

100
50
0
0 2 4 6 8
Deflection (mm)

Figure 3.29 - Load versus deflection curves for model 2 with p = 2.85 % for
different values of '·

350
300

-,-
~ 200
250

co 150 ~Zeta=24
0
...1
-Zeta=12
100 -tr--- Zeta = 6

50
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain in central steel near column (E-3)

Figure 3.30 - Load versus strain in steel near the midspan for model 2 with
p = 2.85 % for different values of'·
3-51

Nielsen et al. 's simplified model with (} =26° . The numerical model also suggests

that the ACI 318 and AS 3600 codes are conservative in taking the failure plane at an

angle of (} = 45° to the surface of the slab.

The results of the finite element models studied in Section 3.2.2 of this chapter reflect

the variability in the results of finite element analyses. Considering all the aspects

discussed above, including the results of the plasticity model of Nielsen et al. (1978),

it is concluded that the model 1D2-2 is the best compromise for efficiency and

accuracy for the study of punching type failures. That is, a fme mesh within the critical

region with a maximum plan aspect ratio of two and two elements ·through the

thickness of the slab is sufficiently accurate and efficient for solving punching

problems. Further, the reduced 2 x 2 x 2 Gaussian integration scheme gives similar

results to the full integration, that is integration on a 3 x 3 x 3 Gauss quadrature. The

reduced integration saves storage and considerably reduces the solution time without

significantly sacrificing accuracy. Away from the punching zone the aspect ratio of the

mesh may be increased without loosing accuracy.

Two modes of failure are clearly seen from the model, that of flexure and that of shear.

Once the punching limit is reached increasing the quantity of steel leads to only a

small increase in the capacity of the slab. This observation is made, however, in the

absence of any dowel effects.

The results of the finite element modelling for a reinforced concrete flat slab in

punching shear show that using a reduced 8 point integration scheme gives a good

solution without sacrificing accuracy. Thus, as for the plain concrete slab analysed in
3-52

Section 3.2.2 of this chapter the use of a reduced integration scheme is appropriate in

the numerical modelling of flat slabs.

The results of the fmite element model are consistent with the observation of

Dragosavic and van den Beuk.el (1974) in that the negative moment reinforcement

does not significantly influence the punching shear strength. The finite element results

are also consistent with the theoretical model for strength. It is concluded, therefore,

that the fmite element formulation is capable of predicting punching failure in

reinforced concrete slabs.

The sensitivity tests of the different parameters give important insights to

understanding the finite element results. The shear retention factor f3 and the

descending slope of the tension stress-strain law (as given by a1) were observed to

have a significant influence on the results of the finite element analyses. The other

parameters have only minor effect on the results. The commonly used values may be

adopted for all these parameters in the fmite element modelling of punching shear type

problems. The value of f3 = 0.2 for the shear retention factor is found to be suitable.

The value of a1 must be carefully chosen to represent the tension stiffening effect,

which is a complex phenomenon involving the interaction between concrete and

bonded reinforcing steel.


CHAPTER 4- VERIFICATION OF THE FINITE ELEMENT MODEL

4.1 Introduction

To verify the generality of the finite element model, described in Chapter 3 and

implemented in the finite element program DIANA (1997), a number of experimental

studies have been numerically simulated. The problems selected comprised of a beam

failing in flexure, shear and torsion to more complex systems like slab-column

connections. The specimens selected were based on the level of details provided in the

experimental study reports. The aim was to test the model against a range of

applications failing in shear and torsion modes including slab-column connections

failing in punching shear.

4.2 Reinforced Concrete Beam of Mansur and Rangan (1978)

Mansur and Rangan (1978) tested a series of floor beams framing into spandrel beams.

The beams were subjected to combined flexure, shear, and torsion. Beam SB-2 is

selected for the analyses. The specimen consisted of a floor beam spanning 3.0 metres

and attached to the mid-span of the spandrel beam of 1.5 metres span. Both the floor

beam and the spandrel beam had a depth of 300 mm and a width of 180 mm. Details

of the specimen and the loading arrangements are shown in Figure 4.1 and the details

of the steel reinforcement arrangements are shown in Figure 4.2. The properties of

reinforcing steel are given in Table 4.1.

The finite element mesh used for the analysis is shown in Figure 4.3 with one half of

the specimen modelled on one side of the line of symmetry. The model consisted of
4-2

I.P.=Inflexion point

Spartdrel Bearn

(a)

OUTLINE OF TEST ASSEMBLY SYMMETRICAL HALF SECTION USED FOR FE MODELING

(b)

Figure 4.1 - (a) Idealisation of the spandrel and floor beam system within the
frame; (b) outline of test assembly and symmetrical half portion
used for FE modelling.
4-3

-
90
I• 9 @ 85 c(_c • 1

I ~ D
2-10 mm dia.
~
-~
6.3 mm dia. stirrups
2-16 mm and 1-12 mm dia.

r~~ ~
,...- 750

SPANDREL BEAM SB - 2
Lines of symmetry

0
1 0 3@100 J
I
21 @ 130 c/c J
2-12 mm dia.
V:l
~~
3000
~~
L.
~
J
I 150
[D
g
--1
6.3 mm dia. stirrups
2-20 mm dia.
J. 180

FLOOR BEAM

Refer Mansur and Rangan (1978) for joint reinforcement detail


REINFORCEMENT DETAILS FOR T-BEAM

Figure 4.2- Specimen detail of the T -beam (Mansur and Rangan, 1978).

Table 4.1 - Material properties of reinforcing bars (Mansur and Rangan, 1978).

Nominal Bar Diameter

Material Property 6.3mm lOmm 12mm 16 mm 20 mm

Actual Bar diameter (mm) 6.60 10.2 11.7 15.3 20.8

Yield strength (MPa) 540 289 488· 502 510

Ultimate strength (MPa) 506 442 597 605 608

Young's modulus (GPa) 202 202 201 202 200


4-4

Central 2 elements under UDL

Elements under UDL

Plane of symmetry

1
Surface fully restrained

Figure 4.3 - Finite element model of specimen SB-2 (Mansur and Rangan, 1978).

992 nodes and 148 elements to model the concrete with additional 401 nodes for the

reinforcing steel modelled as embedded bar elements. The ends of the spandrel beam

were fully restrained against vertical translation and twist. The material properties

used for the concrete elements were

!em = 44.8 MPa ft=4.0MPa Ec=29.0GPa

a1 = 13.1 /3= 0.2 l/> = 30°

lff= 12.6° a= 15° f.l =0.2

eo=0.002

In Figures 4.4 to 4.7 the results of the finite element analyses are compared with the

experimental results for deflections, twists and torque. The load versus deflection

curves of the analytical results generally compare well with the experimental results

(Figure 4.4) but give a slightly stiffer response than that of the experiment.
4-5

180
160
140

--
z
~
120
100 --Floor beam (Experimental)
D.
,; 80 ---Floor beam
ca
0 ------- Spandrel beam (Experimental)
...1 60 ~Spandrel beam
40
20
0
0 10 20 30 40
Deflection (mm)

Figure 4.4 - Load versus deflection curves of floor and spandrel beams.

180
160
140
z 120
.lll:
-Q. 100 --Experimental

ca 80
"C
-+-FE
0
...1 60
40
20
0
0 5 10 15 20 25 30
Average angle of twist of spandrel beam (E-3 Radian/m)

Figure 4.5 - Load versus Twist curves.


4-6

180
160
140
--
z
~

a. 100
120

,ca 80
0
..J 60 -Experimental
40 -+-FE

20
0
0 5 10 15 20
Torque, T (kN m)

Figure 4.6 - Load versus torque curves.

18
16

-E
z~
14
12
-
1-
o)
10
8
..
::s
C"
0
1-
6 -Experimental
-+-FE result
4
2
0
0 5 10 15 20 25 30
Average angle of twist of beam (E-3 Radian/m)

Figure 4.7 - Torque versus twist curves for spandrel beam.


4-7

The load versus average angle of twist in the spandrel beam, shown in Figure 4.5, and

the torque developed in the spandrel beam, shown in Figure 4.6, both compare well

with the experimental data. In Figure 4.7 the torque versus the angle of twist at the

centre of spandrel beam are compared for the finite element and the experimental data.

While the cracking torque was slightly higher for the numerical model, the general

trend is in reasonable agreement.

Strains in the top and the bottom reinforcing bars in the floor beam were monitored at

the locations shown in Figure 4.8. The results of the load versus strain are shown in

Figure 4.9 and a good correlation is observed between the finite element and

experimental data. Strains near the middle of the bottom longitudinal bars and in the

stirrups near the mid-span of the spandrel beam were measured at the locations shown

in Figure 4.8. The finite element results are compared with the experimental data in

Figures 4.10 and 4.11 for the longitudinal bars and the stirrups, respectively. The finite

element results for the longitudinal bars compare well with the experimental data. The

results for the stirrups show a reasonable overall trend but are significantly different

before cracking of the section.

The relatively high strains in the gauged stirrup at low loads in the spandrel beam

(shown in Figure 4.11) and the stiffer response of the sections than that evidenced in

the load displacement curves, indicate that the beam may have been subject to some

pre-cracking before testing. On the whole, however, it is concluded that the finite

element model gives a good prediction to the overall response of the floor-spandrel

beam system.
4-8

-Gauge points

Figure 4.8 - Locations of gauge points for measuring strain in steel


(Specimen SB-2 Mansur and Rangan 1978).
4-9

180
160
140
--
z 120
~

a. 100
'ti'
as 80
0 -Experimental
..J 60 -+-FE
40
20
0
0 5 10 15 20
Strain {E-3)

(a) Bottom bar

180
160
140
--
z~ 120
a. 100
'ti'
as 80 -Experimental
0 -+-FE
..J 60
40
20
0
0 2 4 6 8
Strain {E-3)

(b) Top bar

Figure 4.9- Load versus steel strain curves for floor beam (a) Bottom bar
b) Top bar.
4-10

180
160
140
z 120
.llli::
-100
a.
-c
ca 80 -Experimental
0 -+-FE
..J 60
40
20
0
0 1 2 3 4 5 6 7
Strain (E-3)

Figure 4.10 - Load versus steel strain curves for the bottom longitudinal bars of
the Spandrel beam.

180
160
140
z 120
.llli::
-100
a.
-c
ca 80
0 -Experimental
..J 60
-+-FE
40
20
0
0 2 4 6 8 10 12
Strain (E-3)

Figure 4.11 - Load versus steel strain curves for the stirrup of the spandrel beam.
4-11

4.3 Reinforced Concrete Connection of Rangan and Lim (1992)

4.3.1 Finite Element Modelling

Lim and Rangan (1992) tested three slab-column connections for punching shear with

two of the three specimens containing shear studs at the connections to increase the

punching shear strength of the slab. Slab 1 contained no shear connectors and was

analysed in this study.

Slab 1 consisted of a 4950 by 3170 mm slab with a column of 250 mm square located

at the middle of its free edge. The specimen details and the arrangement of the steel

reinforcement are shown in Figure 4.12. The base of the column was pin supported.

The slab edge adjacent to the column was free while the opposite edge and the side

edges rested on 1-beams. The slab was restrained against lifting at the comers by

placing a beam across all top comers of the slab. However, no direct connection was

made from the supports to the slab and, thus, the slab was free to slip horizontally

against the boundaries. The equilibrating horizontal forces for the horizontal reaction

measured at the base of column could then only have come from friction of the slab

against the 1-beams, or from restraint developed at the comers where the slab was

prevented from lifting, or both. The slab was loaded via 24 equal point loads evenly

distributed across the slab simulating a uniform loading.

The finite element mesh used for the analyses is shown in Figure 4.13 and the

boundary conditions are shown in Figure 4.14. One half of the slab-column connection

was modelled due to symmetry (Figure 4.13). The slab-column connection was

modelled using 504 by 20-node isoparametric brick elements. The reinforcing steel
4-12

~---3170-~.
1B
I

A
i 1 I 10 W8 @ 220 mm
h-1--1-- ·- --- -- foo-l--
9 W8 @1 ?5 mm
I I
I ~ I
I I
I
~ --- I Top bars
---
~

~~--r---- --t- f.-·--


250X250
I
I 0
--
l.()
,......
~
I
I
Bottom bars
COLUMN '¢ I
I
@
E~~
~

4950 I I
I @ IX) I
I I
xl I IX)
S:
S:
I
I
I ~
I
N
~ iI
I
I - I
1-1--'---- --- '"-- f--'-- 13 W8 @ 220 mm
IJ. I I 3 W8 @ 125 mm
D C 16 W8 @ 220 mm

3170
, .. .. , 110

~___1
W8 @ 140 mm

r
800
w 175 mm
nnn>b"I
Simple support with
frictional restraint only

L j IE
SUPPORT
I--
R6 @ 00 mm
4 Y20

250
COVER=15 mm

SECTION X - X

Figure 4.12 -Specimen Slab 1 (Rangan and Lim, 1992).


4-13

~ X

(a) PLAN
y

i
I"""' ~ X

(b) ELEVATION

(c) PERSPECTIVE VIEW

Figure 4.13 - Finite element model of Slab 1 (Rangan and Lim, 1992).
(a) Plan, (b) Elevation and (c) Perspective view.
4-14

EDGE RESTRAINED IN Z

EDGE RESTRAINED IN Z

1 1

L 125
C. --. ---------~-------------------------------
COLUMN
J
---X
~
z
I ·...... . ..
Spring with -~
stiffness, K S1mple support
-------11- X
Note: K=O, 800, 2400, Infinity N/mm

SECTION 1 -1

Figure 4.14 -Boundary conditions used in the fmite element model.


4-15

was modelled using 78 embedded bar elements. To model the boundary friction the

comers of the slab (marked A and Bin Figure 4.14) were restrained in the x, y, and z

directions and the fixity of the column in x was replaced with a spring of variable

stiffness. These boundary conditions being statically equivalent to fixing the column

base in the x-direction and applying horizontal springs at the comers. The stiffness of

the spring at the base of the column was then tuned to simulate different boundary

conditions.

The recorded mean cylinder strength at the time of testing was 25 MPa. In the finite

element modelling, the in-situ strength of the concrete was taken as fcp = 0.9 fcm . The

material parameters for concrete were taken as:

fcp = 22.5 MPa ft = 1.7 MPa Ec=26.0GPa

a1 = 12. /3= 0.2

l/J =30 ° 1fl =12.6° J.l = 0.2

eo=0.002

The properties of the reinforcing bars used are given in Table 4.2.

The spring at the base of the column (in the x-direction) was modelled using

stiffnesses of K =0 N/mm (free), K =800 N/mm, K =2400 N/mm and K = = (fully

restrained). The stiffness K =800 N/mm was selected to match the measured horizontal

load in the column under elastic conditions.


4-16

The calculated failure loads from each of the analyses are presented in Table 4.3. The

load versus mid-panel deflection and load versus vertical column reaction are given in

Figures 4.15 and 4.16, respectively. The vertical reaction versus the horizontal

x-direction reaction is plotted in Figure 4.17.

Table 4.2 - Material properties of reinforcing bars.

Material Property 6mmDia.Bar 8 mmDia. Bar 20 mm Dia. Bar

Area As (mm2) 28 50.0 314

Yield strength (MPa) 250 516 415

Ultimate strength (MPa) 250 516 415

Young's modulus (GPa) 200 173 200

Table 4.3 - Failure load for different base spring stiffnesses.

Stiffness, K (N/mm) Failure Load (kPa) Mode of Failure

0 39 Flexure

800 32 Flexure/Punching Shear

2400 25 Punching Shear

00 19 Punching Shear

Experimental 22 Punching Shear


4-17

45
40
C\1 35
..e
z 30
e
"C
2s --+-Experimental

"'
.2 20 --o- K = 0.0 Nlmm

s0 15 -e-- K = 800 Nlmm

..... 10 -+-- K = 2400 Nlmm


"""'1!l- K =INFINITY
5
0
0 10 20 30 40 50 60 70
Deflection {mm)

Figure 4.15 - Load deflection curves.

Figure 4.16- Load shear force curves.


4-18

-e- Experimental
160
---Q-K =0.0 N/mm
140 --e- K =800 N/mm
~K =2400 Nlmm
120 --tr- K =INFINITY
_100
z
-
.lll::

a:
N
80
60
40
20
0
0 10 20 30 40
Rx (kN)

Figure 4.17 - Vertical reaction {Rz) versus the horizontal reaction (Rx) at the base
of the column.

With the increase of stiffness, the load carrying capacity of the system is found to

decrease. The transfer of load through the connection depends on the combined effect

of the shear and the moment interaction. With the increase of spring stiffness the

bending moment transferred to the connection increases causing a reduction in overall

load carrying capacity.

The mid-span deflections for the finite element analyses and the experimental slab are

shown in Figure 4.15. A reasonable correlation is seen between the experimental data

and the finite element model for the slab. Figure 4.15 also shows that the slab

deflections are not sensitive to the spring stiffness. The main variance between the
4-19

numerical deflections and those recorded in the laboratory is in the pre-cracked elastic

range with negligible deflections being measured in the laboratory.

The vertical reaction at the base of the column is plotted against the total slab load in

Figure 4.16. The figure shows that the vertical reaction in the column and hence, the

shear transferred to the slab is not sensitive to the spring stiffness.

The greatest sensitivity to the boundary conditions is seen in the development of base

shear at the column, shown in Figure 4.17. With K =0, that is with no lateral restraint
at the base of the column, no unbalanced moment occurs at the column-slab joint.

With infinite spring stiffness the horizontal reaction and consequently the unbalanced

moment at the connection is significantly greater than that measured in the laboratory.

The best match is with K = 800 N/mm. In the load versus the horizontal reaction of the

column the influence of boundary conditions is clearly demonstrated.

In addition to influencing the column base shear and, hence, the unbalanced moment

in the column, the spring stiffness affects the strains in the negative moment

reinforcing steel. Figure 4.18 shows the strain in the short span central top steel at the

edge of the column. The strain is small when the lateral restraint is zero and increases

with increasing spring stiffness. The maximum strain in the short span bottom

reinforcement, near the mid-span, is shown in Figure 4.19. The reinforcement for the

spring stiffnesses of K = 0.0 N/mm and K = 800 N/mm were found to have yielded

before the advent of punching failure. This indicates some load redistribution occurred

in the slab prior to failure. No strain measurements on the steel were recorded in the

experimental study.
4-20

~K=o.o
YIELD STRAIN -EI- 1<::800 Nlmm
~ 1<::2400 Nlmm
....-tr- K:: INFINITY

-0.5 0.5 1.5 2.5 3.5 4.5 5.5 6.5


Strain (E-3)

(a) Central top reinforcement bar

~ 1<::0.0 Nlmm
-e- K=SOO Nlmm
--+- K=2400 Nlmm YIELD STRAIN
....-tr- K=INFINITY

0 0.5 1 1.5 2 2.5 3 3.5


Strain (E-3)

(b) Bar immediately adjacent to the column

Figure 4.18 - Maximum strains developed in the short span reinforcement.


4-21

Figure 4.19 - Comparison of maximum strains developed in the short span


central bottom reinforcement (mid-span) for varying degrees of
lateral restraint.

This study shows the influence of boundary conditions in the laboratory test results.

The boundary conditions have a significant effect on the moment transferred from the

slab to the column in flat plate and flat slab structures. This in tum affects the capacity

of the joint and puts into question how well the experimental model simulates a real

structure. This study also showed that the slab deflections and the transfer of vertical

load to the column are not significantly influenced by the boundary conditions.

4.3.2 Sensitivity Analyses

In modelling the punching type problems the behaviour of the cracked concrete is

important in obtaining accurate and reliable results for the failure load and failure

mechanism. In this respect, the sensitivity analyses presented in Chapter 3 show that
4-22

the parameters a1 and f3 are of fundamental importance. Model 2 (K = 800 Nlmm) was

chosen for further sensitivity analyses for the concrete tension parameters a1 and the

shear retention factor /3.

Sensitivity to a1

With the development of cracks, the stiffness of the concrete reduces. This reduction

and its rate is dependent on Eu (refer Figure 3.1). The ultimate uniformly distributed

load on the slab predicted from the finite element analyses with a constant shear

retention factor of f3 = 0.2 and for different values of the parameter a 1 are given in

Table 4.4 and shown in Figures 4.20 and 4.21. The failure load, the displacement of

the slab and the shear to moment ratio at the junction of the column and the slab are

significantly influenced by the slope of the descending branch of the tension stress-

strain curve for the concrete. The development of strain history in one of the critical

bars (central top bar in the short span) is presented in Figure 4.22. The parameter a1 is

seen to be influential in the strain history of the steel and the failure mode.

Sensitivity to J3

For fully cracked plain concrete the shear stiffness becomes small after the onset of

cracking. However, in reinforced concrete the interaction of the reinforcement

crossing the cracks on the influence of aggregate interlock is significant. Although the

effect of shear stiffness of cracked concrete is acknowledged as important in the finite

element modelling community, no consensus is found amongst researchers on a

satisfactory mathematical model. The model with K=800 N/mm is analysed with

a 1 = 15 for different values of /3.


4-23

40
35
N- 30
.e~
- 25
' C 20
as
----Experimental
_._Aipha1 =25
.2
15 -+-Aipha1 =15
~
1- 10 ---tr-Aipha1 =5
5 ......e-Aipha1 =0
0 --~~~~~~~~;-~~~~~~~~,_~~~~

0 10 20 30 40 50
Deflection (mm)

Figure 4.20 - Load versus deflection curves for the Rangan and Lim (1992) slab
with different values of a1•

160
140
120
-100
z
-
~
N
a:
80
60
----Experimental
_._Aipha1 25=
-+-Aipha1 =15
40 ---tr-Aipha1=5
20 ---+-Aipha1 =0

0
0 5 10 15 20 25 30
M(kNm)

Figure 4.21 - Shear versus moment at the column slab connection for the Rangan
and Lim (1992) slab with different values of a1o
4-24

40

-e- Alpha1 = 25
__._Aipha1 = 15
=5
"""'1!1-- Alpha1
-e- Alpha1 =0

-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4


Strain {E-3)

Figure 4.22 - Load versus strain for the central top bar for the short span for the
Rangan and Lim (1992) slabl fordifferent values of a1o

Table 4.4 - Ultimate load for different values of a1o

Model Value of al Ultimate Load (kPa)

2-A 0 21.7

2-B 5 23.5

2 15 32.4

2-C 25 34.4

5 Experimental 22.0
4-25

The effect of the shear retention factor on the slab deflections, the distribution of

vertical to horizontal loads at the base of the column, the development of strain in the

top reinforcement at the centre of the column and on the failure loads are given in

Figures 4.23 to 4.25 and Table 4.5, respectively. The analyses show that the shear

retention factor does not influence the results of the load versus displacement; the

distribution of horizontal to vertical reaction and, hence, the shear to moment ratio at

the slab-column intersection; or the development of strain with increasing applied load

in the negative moment reinforcement in the vicinity of the connection. However, the

value of f3 chosen has a significant influence on the failure load. The failure load

varied by up to a factor of two depending on f3 (see Table 4.5).

Table 4.5 - Ultimate load for different values of f3.

Model Value of~ Ultimate Load (kPa)

2-D 0.01 20.0

2 0.20 32.4

2-E 0.50 40.5

2-F 0.90 45.7

5 Experimental 22.0
4-26

50
45

.e-
40
C\1
35

-,~ 30
ca 25
.S! 20
-+-Experimental
=0.9
---e...- Beta
--b:- Beta =0.5
s 0 15 -+-Beta =0.2
1-
10 --e- Beta = 0.01
5
0
0 20 40 60 80
Deflection (mm)

Figure 4.23 - Load versus deflection curves for the Rangan and Lim (1992)
slab-column connection with different values of {3.

200
180
160
140
-120
z
e1oo -+-Experimental
N
a: 80 - - 0 - Beta = 0.9

--b:- Beta =
0.5
60 -+-Beta 0.2 =
40 --e- Beta 0.01 =
20
0
0 10 20 30 40 50 60
Rx(kNm)

Figure 4.24 - Vertical versus horizontal reactions at the column for the Rangan
and Lim (1992) slab-column connection with different values of {3.
4-27

50
45 Yield limit
-40
N

.e
35

-
~30
1125
0
~Beta=0.9

---&--- Beta = 0.5


20
-+-Beta = 0.2
~ 15
~Beta =0.01
1- 10

5
0
0 2 4 6 8 10
Strain {E-3)

Figure 4.25 - Load versus strain for the central top bar for short span for the
Rangan and Lim (1992) slab1 for different values of f3.

4.4 Prestressed Concrete Connections of Foutch et al. (1990)

Foutch et al. (1990) tested four post-tensioned slab-edge column connections. The

specimens were based on a two-third scale model of a prototype structure. The

specimens are identical in the external dimensions and had similar boundary

conditions. The outlines of the specimens with the loading points are shown in

Figure 4.26. The arrangement of the bonded non-prestressed reinforcement in the

specimens is shown in Figure 4.27. The locations of the tendons for specimens S 1 and

S2 are shown in Figure 4.28a. The location of the tendons for specimens S3 and S4 are

shown in Figure 4.28b.


4-28

1520 Half of the specimen


used for FE modeling

Note:
D: Distance of point
of application P
A: Central edge point

VALUE OF D (mm)
Specimen D
PLAN
S1 1070
points
S2 610
for whiffle tree
S3 610
/////// S4 305
/......_
Steel Plate 300X3 OOX25

300X300 Column

A_L
2030 T1 00
D
p
Load applied th rough
whiffle tree
'V"
LABOR ATORY FLOOR /\.

ELEVATION

Figure 4.26 - Outline with key points for the Foutch et al. (1990) specimens.
4-29

1520
1 ...
2 13
L
1 128 '1
• 47 77
L77
128 ~~ !1-
~I-~------1~5~2~0------~-~~~

'
13 SECTION 2-2
52 100
1245
II I I
100
Bors used for slob: #3 (9.5mm) deformed bos.

- Strain gouge position end number

0
0
N

0
0 Stirrups #3 Deformed Bors
N

2030
0
0
N
1
0
0
N

0
0
N
II)
Ol
8 #6 Deformed Bars
D
SECTION 3-3
LABOR ATORY FLOOR
SECTION 1-1

Figure 4.27 - Arrangement of non-prestressed reinforcement


(Foutch et al., 1990).
4-30

1520 1520

=i-1
.3 45

.3 @75
1 r-h 1 1 15 1 h _J_ 1

L 152':J 2 @75----X L I 15o 152~--- X

]'j
i-1-J 11 5 f-J
.3 @75

.3 45 260
_j_
150 405 405 405 10.30
75 75 115
2_J 2_j

(a) Specimen Sl and S2 (b) Specimen S3 and S4

z 1520

1066

_l_
!--:---:=====~======== :;:zss~s1 ~ X

SECTION 1-1
z
1219 1520

102
l
T
-rn I
~
51
25.5
25.5 .. y
183* 180* 245 .245 180* 183*

610
-
~~ ~

610

SECTION 2-2

NOTE: * - Assumed values

Figure 4.28 - Arrangement of tendons (Foutch et al., 1990).


4-31

The properties of the concrete, reinforcement and tendons used in the test are

presented in Tables 4.6 to 4.8. The tests were carried out using whiffle tree to apply

four vertical loads arranged symmetrically at a distance of 1.07, 0.61, 0.61 and

0.305 metres from the front face of the column for specimens S 1, S2, S3 and S4

respectively (see Figure 4.26).

Table 4.6 - Concrete properties used for the FE model (Foutch et al., 1990).

Material Property Specimen S1 Specimen S2 Specimen S3 Specimen S4

fcp = 0.9 !em (MPa) 45.3 38.5 37.9 43.4

.ft (MPa) 2.34 2.15 2.14 2.29

a(Degrees) 15 15 15 15

</>(Degrees) 30 30 30 30

VI (Degrees) 12.6 12.6 12.6 12.6

aI 17.6 17.4 17.5 17.5

J.l 0.2 0.2 0.2 0.2

f3 0.2 0.2 0.2 0.2

eo 0.002 0.002 0.002 0.002

Ec (GPa) 33.0 30.0 30.0 32.0


4-32

Table 4.7 - Material properties of reinforcing steel (Foutch et al., 1990).

Property Bar#3 Bar#6

Area As (mm2) 71 284

Yield strength (MPa) 501 501

Ultimate strength (MPa) 835 793

Young's modulus (GPa) 200 200

Ultimate strain (%) 0.045 0.046

Table 4.8 - Material properties of tendons, 3/8 inch - 7 Wire Strand.

Description Values

Area As (mm2) 51.6

Yield strength (MPa) 1900

Ultimate strength (MPa) 1900

Young's modulus (GPa) 195

The specimens were modelled using the mesh shown in Figure 4.29 with one half of

the specimen modelled due to symmetry. The finite element model had 276 20-node

solid isoparametric concrete elements with 1755 nodes. The bonded reinforcement and

the prestressing tendons were modelled using embedded elements defined by 120

nodes for specimens Sl and S2 and 114 nodes for specimens S3 and S4. The

reinforcement was placed in the same locations for all specimens. Perfect bond was

assumed between the bonded non-prestressed reinforcement and the concrete. The

pinned supports were applied at the top and at the bottom of the column, matching that
4-33

-- - ~- --x
1 ..
1524 .. 1
Plane of symmetry
Column
PLAN
z z
~~ ~ IT-
- -

r- r-

2 030 -x 20 30 y

- -

- -

~~~ ../. ~762


1524

FRONT ELEVATION SIDE ELEVATION

(a) Plan and elevations showing boundary conditions

(b) 3D view

Figure 4.29 - Finite element model of the slab-column (a) Plan and elevations
showing boundary conditions; and (b) 3D view.
4-34

of the experimental test set up. The Hognestad (1951) model was used to model the

concrete compressive stress strain diagram with the concrete strength taken as

fcp = 0.90 !em. where !em is the mean cylinder strength of the concrete at the time of test.

The concrete tension strength was taken as ft


.. =0.33~ fern MPa.

Figures 4.30 to 4.33 show the load versus displacement of the slab edge (shown as

point A in Figure 4.26) for slabs S 1 to S4, respectively. In Figures 4.34 to 4.37, the

load versus strain in the steel reinforcement is compared for different bonded

reinforcement at the locations shown in Figure 4.27. In all cases the numerical model

is in good agreement with the measured response of the slab-column connection. The

ultimate loads are given in Table 4.9.

Table 4.9- Comparison of ultimate loads.

Specimen Load Ultimate Load (kN)


Eccentricity
(mm) Experimental Finite Element

Specimen S1 1070 58.0 61.0

Specimen S2 610 50.0 55.0

Specimen S3 610 40.0 44.0

Specimen S4 305 36.0 42.0


4-35

60

50

-40
z~
-
"C
as
0
30
...1
20 - Experimantal
--FE
10

0
0 10 20 30 40 50 60
Edge deflection at A (mm)

Figure 4.30 - Load versus deflection curves for Specimens Sl.

100
-Experimental
90
--FE
80

--
z
~
70
60
50
"C
as
0 40
...1
30
20
10
0
-1 9 19 29 39 49 59
Edge deflection (mm)

Figure 4.31 - Load versus deflection curves for Specimens S2.


4-36

80
70
60
z 50
-Jill::
"C
ca
0
40
..J 30 -Experimental
-4-FE
20
10
0 --~~~~~~~~~~~~~~~~~~~~~~

0 10 20 30 40 50 60
Edge deflection (mm)

Figure 4.32- Load versus deflection curves for Specimens S3.

160
140
120

--
zJill:: 100
"C 80
ca
0
..J 60
40 -Experimental
-4-FE
20
0
0 10 20 30 40
Edge deflection (mm)

Figure 4.33 - Load versus deflection curves for Specimens S4.


4-37

60

50

--z
Jill::
40

30
"'0
as
0
...1
20 ~Experimental: Gage 6,7
-+--Experimental: Gage 8
-11- FE: Gage 6
10 -+-FE: Gage 7
-.-FE: Gage 8

0
0 5 10 15 20
Strain (E-3)

Figure 4.34 - Load versus reinforcement strain history for Specimen Sl.

100
90
80

-c
z
70
60
50
"'0
as
0 ~Experiment: Gage 4
...1 40
-+--Experiment: Gage 5
30 --11- Experiment: Gage 6
-11- FE: Gage 4
-+--FE: Gage 5
-.-FE: Gage 6

-1 4 9 14
Strain (E-3)

Figure 4.35 - Load versus reinforcement strain history for Specimen 2.


4-38

80
70
60

-,-
z
~
50
40 -a- Experiment: Gage 6
as __..._Experiment: Gage 7
0
...1 30 --t1- Experiment: Gage 8
-----FE: Gage 6
20
~FE:Gage7

10 ......._FE: Gage 8

0
0 5 10 15 20 25 30
Strain (E-3)

Figure 4.36 - Load versus reinforcement strain history for Specimen S3.

160
140
120

-z 100
-
,as
~

0
80
-a- Experiment: Gage 5
__..._Experiment: Gage 6
--t1- Experiment: Gage 8
...1 60
---FE: Gage 5
40 ~FE:Gage6

......._FE:Gage8
20
0
0 5 10 15 20
Reinforcement strain (E-3)

Figure 4.37 - Load versus reinforcement strain history for Specimen S4.
4-39

4.5 Prestressed Concrete Connection of Bums and Hemakom (1977)

Bums and Hemakom (1977) conducted tests on a 1/3 scale specimen of prestressed

concrete slab-column system. The specimen consisted of a 74 mm thick slab resting on

16 short columns representing a 3-bay frame structure in each direction (shown in

Figure 4.38). Bonded reinforcing bars were used only in the negative moment regions

(see Figure 4.39) with all positive moment reinforcement being unbonded prestressed

tendons. The arrangements of the tendons were as shown in Figure 4.40. The slab was

uniformly loaded and unloaded a number of times, with 16 different loading patterns,

before being tested to destruction. In this study only the uniform load wad modelled.

For the purpose of the analysis the overhang portion of the slab was neglected and one

quarter of slab was modelled with symmetry. The cable profile is also taken as

symmetrical. The finite element mesh used is shown in Figure 4.41. The slab was

modelled using 1282 twenty-node solid isoparametric elements with 7520 nodes. The

bonded reinforcement was modelled as two-node embedded bars and the straight and

curved tendons were modelled as embedded two-node and three-node unbonded bars,

respectively. An additional 216 nodes were used to define the bonded bars and 310

nodes for the unbounded tendons.

The parameters used to define the concrete model were

fcp = 30.6 MPa ft= 1.92MPa Ec=29.0GPa

a1 = 14.0 f3 =0.2 l/J = 30°

a= 15° 1/f= 12.6° J.l = 0.2

eo= 0.002
4-40

. :L
. G~ . .. .. c2 Portion of the specimen
u u

. ... .. . . .
4
4
4
• 4
for FE modeling
.
,4
.
.. .. .·• . . . . Span/h = 44
4.
0
co 4 •
co
0 0
v
. c4'l!J .. .c3. Prototype - 9 .144 m span
4

I
4
: 4
,[

. . . " • Slab 1-1/3 sc ale model


• 4
• I
I
0
0
0
0
- 4681'\
70 % Tendons on column strip
0 0 0 [

ALL COLUMNS ARE 200 MM X 200 MM


F/A = 2.24 MPo
0 0 0 [

PLAN
I. 10,000 .I
---17621----3048
v
·I· 2048
·I· 3048----11--102

r--

i
1"1{ K· K 1 ~25 mm
bearing
ball

v
r'1
ELEVATION

Figure 4.38 - Plan and elevation of slab-column connection with the portion used
for f"mite element analyses for the Bums and Hemakom (1977)
specimen.
4-41

0
0
0
o~m
each way
/,
*/•
~Line
5 Bars each

of
Column used in ';nalyses

D
8-9.5 mm dia.
bars
Ties: 6 mm dia.

SECTION X-X
(assumed)
0

+
762

NOTE:
Length of each bar if not mentioned = 1.524 m
Reinforcement layout symmetrical about diagonal

Figure 4.39 -Arrangement of bonded reinforcement


(Burns and Hemakom, 1977).
4-42

@ 254 mm c/c
@ 108 mm c/c

u u L.:J
"'
tO

a if ~
1-
NOTE:
~
~ Areo of tendons 23.22 mm2
a= ~ 0 = 188 mm
v
0 ;! 0 ~ c
a= f=
0 X 0
,.... X
0

t J
0 ~
0 a= ~
v
0
0 0 0 [
;!
f=
1404 ~~~ ~
II
ll
a
ll127oll
14D4J 1
a
111 ?oil
1L 1401411
a a
IL75
a
@
@
254 mm c/c
108 mm c/c

v
0 0 0 0 [
;!

10,000

Bonded bars

SECTION X-X

Figure 4.40 - Tendon arrangements (Burns and Hemakom, 1977).


4-43

"
c

£Q)

E
E
4E 0~
0
Q)
c
0
0::

u ---x
Plane of symmetry

4670

:!===F ~~~~~~4::e:=E670~~i~
FRONT AND SIDE ELEVATIONS

Figure 4.41- Finite element model of Burns and Hemakom (1977) slab-column
(Plan, elevations and 3D view with boundary from below).
4-44

with the in situ concrete strength was taken as 0.9fcm· The reported cylinder strength

was 34.0 MPa. The properties of the non-prestressed reinforcement and tendons used

in the test are given in Tables 4.10 and 4.11.

The locations of main points to measure the deflection (point C) and strain in steel

(gauge 1 and 2) are shown in Figure 4.42. Figure 4.43 compares the load versus

mid-panel deflection response at C (see Figure 4.42) from the finite element model

with the experimental data. A good correlation is observed. The failure load obtained

in the finite element model was 9.8 kPa and compares well with the experimental

result of 9.9 kPa. Calculated from the strains measured in the steel of the columns, the

reaction in column C4 at failure was reported to be 150 kN. In the finite element

model the load in column C4 at failure was 159 kN, within 6 percent of the

experimental result.

Table 4.10 -Material properties of reinforcing steel (Burns and Hemakom 1977).

Bars

Material Property #2 #3

Area As (mm2) 32 71

Yield strength (MPa) 380 280

Young's modulus (MPa) 200 200


4-45

Coordinates of gauging points


1: (1364, 4664, 53) mm
2: (1364, 1524, 53) mm

Point C for deflection measurement


C: (3050, 3050, 74) mm

0
co
<.0
..q-

L---f----------------------------~~----x

01 .. 4680 ----~

Figure 4.42 - Location of strain points for finite element analysis of Burns and
Hemakom (1977) slab.

12

-
"'e
.....
z
10

-
~

"C
ca
.2
8

6
"C
.S! 4 --Experimental
c.
c. --FEM
<C
2

0
0 50 100 150
Deflection at C (mm)

Figure 4.43 - Load deflection curves for the Burns and Hemakom (1977) slab.
4-46

Table 4.11 - Material properties of 3/8 inch -7 wire strand tendons


(Burns and Hemakom, 1977).

Property Value

Area As (mm2) 23.2

Yield strength (MPa) 1725

Young's modulus (GPa) 200

No strain histories were gauged in the test. The maximum strain developed at key

reinforcement locations in the finite element analyses are shown in Figure 4.42. The

load versus strain plots in the bonded reinforcement for the finite element model at

these locations are shown in Figure 4.44. The plots show that these bars had yielded

before failure. In the test the slab was reported to have failed by punching at the

interior column connection C4. The finite element results predict that the slab failed in

flexure followed by punching failure at the interior column C4.

4.6 Conclusions

In this chapter finite element models were developed for beams and slab-column

connections reported to have failed in shear-torsion or punching shear mode from four

experimental studies. The findings of the investigation presented in Chapter 3 were

incorporated in developing these models. The overall observation is that finite element

modelling is able to compute the failure loads and failure modes of the experiments.

Measured data such as load versus deflections for example, were shown to correlate

well for the numerical and the experimental results.


4-47

16
14
-.e
C'll

z 10
12
~Strain location 1

-~

"D
ca
.2
8
.........._Strain location 2

6
s0
1- 4 Yield limit
2

-1 4 9 19
Steel strain (E-3)

Figure 4.44 - Load versus strain curves for bonded bars in the negative moment
region of the slab.

From this study it is concluded that the finite element models used in DIANA (1997)

give reasonable results within a tolerable range. However, the studies must be

accompanied with parameter sensitivity analyses. In particular the softening slope of

the stress-strain curve for the concrete in tension was shown to dramatically effect the

flexural response of the slabs and the shear retention factor was shown to highly

influence the failure load, particularly for slabs failing in punching shear. In

conclusion the model can be used for studying the behaviour of slabs failing in

punching shear provided that it is appropriately calibrated and verified.


CHAPTER 5 • PULLOUT TESTS

5.1 Introduction

As discussed in Chapter 3, the shear retention factor f3 and the tension stiffening

parameter a1 are two important parameters required for calibration of the finite
element modelling of the concrete structures. These parameters are investigated below

by studying the results of a series of anchor pullout tests.

A number of experimental and analytical studies have been undertaken to study the

failure mechanism of pullout of bars and anchors embedded in concrete. According to

Stone and Carino (1983) and Hellier et al. (1987), pullout tests were originally

proposed by Soviet scientists in 1934. Experimental work and numerical modelling

(Malhotra, 1975, Stone and Carino, 1983, Stone and Carino, 1984, Yener and Li,

1991, Yener 1994) have been used to calibrate the results of pullout tests with the

strength of the concrete. Klinger and Mendonca (1982), Ichihashi et al. (1985) and

DeVries et al. (1999) are just some of the many experimental studies undertaken on

pullout of anchors in concrete.

Figure 5.1 shows a schematic diagram of a typical pullout test of an embedded plate.

Stone and Carino (1983) carried out two similar pullout tests using large-scale

specimens having apex angles (2a, refer Figure 5.1) of 54 and 70 degrees. Figure 5.2

shows the test set-up used in these studies. Based on discontinuities in the load-strain

histories observed in the embedded gauges, Stone and Carino concluded that pullout

failure occurs in three phases:


5-2

~--------d~------~

d2 =Disk size p
d, =Int. dimension of
reaction ring
h =Embedment depth
2cx =Apex angle
t
ring

\4
\
Idealized failur~\ <l
4
surface \ <l

<l
<l
4
4
<l
<l
d
<l
4
4
<l 4
<l

Figure 5.1 - Schematic representation of a typical axisymmetric pullout test.

I) Initiation of a circumferential crack near the upper edge of the disk at


approximately 113 of ultimate load, P u·

II) Completion of circumferential cracking from the disk edge to the reaction ring at
approximately 2/3 of p u•

III) Shear failure of the concrete matrix and degradation of aggregate interlock
beginning at 0.8Pu·

Stone and Carino showed that the increase in apex angle changes the failure surface

from idealised conic frustum shape to flatter curved surface.

Stone and Carino (1984) conducted a two-dimensional linear analysis of their

laboratory tests, which showed a good agreement with the observed data up to the
5-3

A: VSL ES-31 PULLING HEAD For specimen 1:


B: JACKING HEAD Apex angle = 2cx =70 degrees
Embedment depth h = 254mm
C: RAM CLUSTER

D: TOP RETAINER RING

E: BOTTOM RETAINER RING For specimen 1:

F: RAM PEDESTAL Apex angle = 2cx =54 degrees


Embedment depth h = 343mm
G: FRAME STRUT
H: REACTION RING
J: STEM
K: DISK
L: FIXED END ANCHOR

Apex angle = 2 ()(

All dimensions are in mm.

\.l,.__j__jj_~~~>portir>g
bars @ 150 c/ c
AROUND CIRCUMFERENCE OF HOOP

1370 Diameter

2030 Square

Figure 5.2 -Test set-up used by Stone and Carino (1983).


5-4

initiation of cracking. Using an axisymmetric finite element model Hellier et al. (1987)

studied the behaviour of pullout specimens similar to that shown in Figure 5.1 for an

apex angle of 70 degrees. They concluded that the ultimate load carrying mechanism

is primarily due to aggregate interlock across the failure surface.

Based on elastic-plastic-fracture finite element analyses of Yener and Li (1991) and

Yener (1994), Yener (1994) state that in the post micro-cracking stage stresses are

continuously redistributed with crack propagation and the formation of new cracks

until failure. The residual strength in the pullout test is a consequence of the crushing

of the concrete in the vicinity of the support ring.

In this study a large apex angle is used in order that the failure mechanism is not

influenced by the boundary conditions. Under these conditions failure of the anchor by

pullout is similar to that of pure punching shear where no moment is introduced into

the connections.

5.2 Experimental Program and Specimens

A study of the literature of punching failure of slab shows that the failure cone forms

at a slope of less than 45 degrees with the horizontal (Regan, 1981, Hammill and

Ghali, 1994, Walker and Regan, 1987, Alexander and Simmonds, 1986). In this study

the apex angles were kept large to ensure a minimal influence on the punching cone

from the remote boundary. Two test series were undertaken in this study with concrete

compressive strengths of nominally 20 MPa and 40 MPa. Each test series consisted of

three specimens of different sizes and with different anchorage embedment depths (h)

(see Figure 5.3). The apex angle was between 130 and 144 degrees. The specimens are
5-5

t
1 ..
D2
I
..I
~ 1DOX100X20
1
steel plate

.v
at the centre
I
1 1

1t
D2 --
J
---x
11:<~

L so--il ..
~
01
I
I

.. If--so
Rig

Reactive farce p
!~ ~
<3
h <3~
----r
(h+250)
Steel
;1 ~_)
Plate
250
;1 ___i

.-I•t---- D2 -------i·~l
SECTION 1-1

Specimen h D1 D2 Apex Angle


(mm) (mm) (mm) (Degrees)

PXX-50 50 730 900 144

PXX-100 100 1000 1200 130

PXX-150 150 1500 1700 132

Figure 5.3 - Details for pullout test specimens.


5-6

designated as PXX-YYY where XX is the nominal concrete strength and YYY is the

embedment depth of the plate. For example P20-1 00 was cast with a nominal concrete

compressive strength of 20 MPa and a plate embedment of 100 mm.

The pullout load was applied via a 20 mm thick by 100 mm square plate, embedded at

various depths and connected to a 2 m long, 29 mm diameter Macalloy bar pulling

rod, as shown in Figure 5.4. Three embedment depths were tested, 50, 100 and

150 mm. For the specimens with 100 mm and 150 mm embedment depths a net of

wire mesh of 4 mm diameter was placed at a depth of 30 mm from the top surface

(see Figure 5.4). The function of the wire mesh was to avoid any accidental damage to

the specimen during curing. No damage was observed to the surface of any specimen

at the location of Macalloy rod entering into the specimen. The size of wire mesh in

plan was small as not to have any influence on the failure. A typical detail of the wire

mesh is shown in Plate 5.1. In all specimens, the pulling rod rested on the base of the

formwork. The rod was enclosed in a greased 40 mm diameter PVC pipe below the

plate to ensure no bond occurred between the concrete and the Macalloy bar below the

anchor plate. The formwork for all specimens was made of 18 mm thick plywood.

Once all parts were in place for the required position the assembly was braced. The

wire mesh was held in position using plastic bar chairs or 6 mm diameter steel wires

projected from the bracing. The assembly is shown in Figure 5.4 and Plate 5.2.

The concrete used in the study was provided by a local ready-mix concrete supplier.

The specifications for the concrete were a maximum aggregate size of 10 mm; a slump

of 80 mm and a target mean cylinder strength of either 20 or 40 MPa.


5-7

"I D3/2 "I D3/2


" "
Pulling rod
~

..........._
~ '100

L~
f---100~
OOX100X20 mm

29 mm Macalloy bar
concentric to the plate

SECTION 1-1

WIRE MESH DETAIL

force Reactive force

t <1

L
(h+250) <!

Wire<1 mesh
~
'-----;;----t-1 OOX1 OOX20 mm steel plate
<!
---~..._-.:r---+-PVC pipe coated with
lubricant inside and

..
I~ - - - D 2 - - - . . -..l l outside both

Note: D3 = 200 mm and 280 mm for specimens PX:X-100 and PX:X-150, respectively.
No mesh was provided for specimens PXX-50.

Figure 5.4 - Cross-sectional detail of the pullout specimens.


5-8

Plate 5.1 -Arrangement of bearing plate and wire mesh.


5-9

Plate 5.2- Set-ups ready before the concrete pour.

The actual strengths at the time of testing are given in Table 5.1 together with the

concrete tension strength obtained from split cylinder tests (tested in accordance with

AS 1012.10-1985) and the measured elastic modulus of the concrete. The concrete

compressive strength was measured using 150 mm diameter by 300 mm high cylinders

tested in accordance with AS 1012.9-1986.

The concrete was poured carefully so not to disturb the assembly and an electric

needle vibrator was used to consolidate the concrete. A steel trowel was used for

finishing the top surface of each specimen. For each pour ten large cylinders (150 mm

diameter x 300 mm height) and fifteen small cylinders (100 mm diameter x 200 mm

height) were cast to measure the properties of the concrete. One day after the concrete

was placed, the specimens were covered by hessian and plastic with the hessian
5-10

watered twice daily for a minimum of 7 days. The cylinders were striped from their

moulds after 24 hours and cured with the test specimens.

Table 5.1 - Concrete properties for different sets.

Set Age w fcm It Ec


(Days) (kg/m3) (MPa) (MPa) (GPa)

P20-YYY 58 22.2 24.1 2.9 24

P40-YYY 212 23.0 41.7 4.0 32.5

5.3 Test Set-up and Instrumentation

The reaction rig was fabricated of rectangular tubular section having external size of

100 mm by 50 mm and was placed on top of the concrete blocks concentric with the

central pullout rod. The reaction rig was connected to a 1000 kN capacity hydraulic

jack as shown in Figure 5.5. The load was measured using a 500 kN capacity load cell.

In test series P20 displacements were measured at four points along a central line. In

the second test series (P40) displacements were measured at three symmetrical points

on top of the concrete blocks. These points and their positions are presented in

Figure 5.5. Linear variable differential transducers (LVDT's) were used to measure

displacement with the transducers connected to a data acquisition control unit. The

reading accuracy of the transducers was 0.01 mm and their maximum travel ranges

were 100 mm. A typical arrangement for the test set-up is shown in Plate 5.3. The

output of all electronic measurements was channelled through a HBM signal amplifier

and then to Labview ( 1999) via a data acquisition control unit.


5-11

-Reaction rig-

1 1
L _j

2_j

LVDTS FOR SERIES P20-YYY LVDTS FOR SERIES P40-YYY

Macalloy bar Macalloy


Nut Nut
Loadcell
. - - - - - - MS plates - - - - - , 1
Hydraulic Jack t (at top and bottom) Hydraulic Jack

"" - - 1 girders - - - - - - - M
!!
"
!!
r
(h+250)
L_
~
Concrete b)ock
'---Lf--- Reaction rig ----+-'---;;---'
C;ncr~te
17
..
block

(h+250)
_j_
SECTION 1-1 SECTION 2-2

Note:
a=85mm for h=100mm For clarity stiffeners
a=120mm for h=100mm are not shown.
a=150mm for h=150mm 01 is as shown in Figure 4.3.
b=100mm for all h = embedment depth

Figure 5.5 - Test set up and gauging points.


5-12

Plate 5.3 - Test set up showing L VDTs and steel girders resting on reaction rig.
5-13

5.4 Tests Results and Discussions

The jack was slowly pressurised to apply the load at the rate of approximately

0.3 kN/minute. Thus depending on the specimen the testing times varied from Y2 hours

to 2 hours. The failure took place suddenly in all tests as expected and the dimensions

of the failure cones were measured and are given in Figures 5.6 and 5.7. Some of the

typical failure cones are presented in Plate 5.4. In all cases the failure occurred well

away from the boundaries and, thus, the boundary did not influence the results.

The peak load and the corresponding displacement at point A (see Figure 5.5) for all

tests are given in Table 5.2. The angle of the failure cone was significantly less than

45 degrees in all specimens. The plan dimensions of the top of failure surface are

presented in Figures 5.6 and 5.7 for series P20 and P40, respectively. The load

deflection curves at different points for series P20 are plotted in Figures 5.8 to 5.10

and those for P40 in Figures 5.11 to 5.13.

The experimental results are compared with the Nielsen et al. (1978) and

ACI 318-1999 models as shown in Table 5.3. The ACI model (ACI 349-1990) used to

determine the pullout load P u is given by Pu = 0.33..JI: Ac , where J; is the

characteristic cylinder compressive strength in MPa and Ac is the area defined by the

projected area of a 45 degrees cone radiating from the edge of the anchor plate. The

in-situ compressive strength was taken as 0.9 fcm . For the Nielsen et al. model the

lower bound was taken for ft = fcp/1000 and the upper bound ft = fcp/400 (see

Section 3.2.3). The angle of friction was taken as l/J = 30° and the efficiency factor

was taken as 0.91 for the P20 specimens and 0.69 for the P40 specimens, as obtained

from equation 3.35. Table 5.3 shows that the ACI model gives a safe prediction of the

failure load while the Nielsen model errs on the non-conservative.


5-14

Failure line

Edge of concrete block

Specimen Plate embedment Dimension A Dimension B


depthh (mm) (mm) (mm)

p 20-50 50 530 630


p 20-100 100 870 650

p 20-150 150 750 1000

Figure 5.6- Plan dimensions of failure cone for test series P20-YYY.
5-15

c
L

o Central
Central plate
plate

Failure line Failure line

Edge of concrete block

Specimen Plate Dimension Dimension Dimension Dimension


embedment A B c D
depth h (mm) (mm) (mm) (mm) (mm)

P40-50 50 550 720 650 605

P40-100 100 870 650 1070 505

P40-150 150 720 850 930 680

Figure 5.7- Plan dimensions of failure cone for test series P40-YYY.
5-16

(a) Specimen P 20-100

(b) SpecimenP 40-50

Plate 5.4 - Failure cones.


5-17

(c) Specimen P 40-100

(d) Specimen P40-150

Plate 5.4 (Continued)- Failure cones.


5-18

80
70

-
~50
:; 40
60 LVDTB LVDTA

~ 30
20
10
0
0 0.2 0.4 0.6 0.8 1
Displacement (mm)

Figure 5.8 - Load versus displacement curves for specimen P 20-50.

140 LVDTA
120

-,-
z
~
100
80
ca 60
0
..J
40
20
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Displacement (mm)

Figure 5.9- Load versus displacement curves for specimen P 20-100.


5-19

250 ~LVDTC
LVDT LVDTA
LVDTB
200
--
~ 150
i0 100
...1

50

0
0 0.1 0.2 0.3 0.4 0.5
Displacement (mm)

Figure 5.10- Load versus displacement curves for series P 20-150 specimens.

120
LVDTC
100 LVDTB

--
z
~

'0
80
60
tiS
0
...1 40
20
0
0 0.1 0.2 0.3 0.4 0.5
Displacement (mm)

Figure 5.11 - Load versus displacement curves for series P 40-50 specimens.
5-20

180
160
140 LVDTB
-120
z
,
c1oo
ca 80
LVDTA

0
..J 60

40
20
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Displaceemnt {mm)

Figure 5.12- Load versus displacement curves for specimen P 40-100.

350
300

-~ 250

-
,ca
0
200
150
..J
100
50
0
0 0.1 0.2 0.3 0.4
Displacement {mm)

Figure 5.13- Load versus displacement curves for specimen P 40-150.


5-21

Table 5.2 - Failure load, corresponding displacements and the average slope of
the cone.

Specimen Failure Load Pu Displacement A Average slope of the


(kN) (mm) failure cone (}
(Degrees)

P20-50 73 0.337 10

P20-100 129 0.248 15

P20-150 207 0.304 19

P40-50 107 0.326 9

P40-100 171 0.268 15

P40-150 306 0.253 20

Table 5.3- Comparison of peak load with ACI 318-1999 and Nielsen et al. (1978)
models.

Specimen Failure Load Pu (kN)

ACI Nielsen et al. (1978)


318 -A-
Expt
Experimental Lower LB Upper UB
bound -- bound --
Expt Expt
(Expy) (A) (LB) (UB)

P20-50 73 39 0.53 49 0.67 59 0.80

P20-100 129 103 0.80 150 1.16 173 1.34

P20-150 207 190 0.91 287 1.39 326 1.57

P40-50 107 52 0.49 64 0.60 77 0.71

P40-100 171 135 0.79 198 1.16 228 1.33

P40-150 306 250 0.82 378 1.23 428 1.40


5-22

5.5 Finite Element Analyses

The primary aim of the pullout tests was to calibrate the shear retention factor f3 and
the tension stiffening factor a1 for the problems failing in punching shear. All

specimens were modelled using 20 node solid isoparametric elements with one quarter

of the specimen modelled with symmetry. The effect of the rigid plate action was

modelled using constant displacement for the nodes representing the plate and it was

assumed that the plate was not bonded to the concrete. The planes of symmetry were

modelled using the appropriate boundary conditions and the top nodes of the elements,

where the reaction rig was placed, were restrained against the movements in the

vertical direction (see Figure 5.14). DIANA (1997) was used to carry out the finite

element analyses.

The finite element meshes used to model the specimens are shown in Figure 5.15. The

meshes are designated FE-50, FE-100 and FE-150 and correspond to a plate

embedment depth of 50, 100 and 150 mm, respectively. The number of nodes and the

elements contained in the meshes are given in Table 5.4. The in-situ strength of the

pullout concrete block was taken as fep = 0.9 fern and the tensile strength was taken as

ft = 0.33~ fern . Material properties used for the concrete elements are

For P20 specimens:

fep = 21.7 MPa .ft = 1.62 MPa Ec=20.0GPa

f3 = 0.2 t/> =30° a= 15°

Vf= 12.6° f.l =0.2 eo= 0.002


5-23

For P40 specimens:

fcp = 37.5 MPa .ft=2.13MPa Ec= 25.0GPa

[3 = 0.2 tfJ =30° a= 15°

/.l =0.2 eo= o.oo2


0
1JI = 12.6

The tension stiffening parameters a1 for the different models of P20 and P40 are given

in Table 5.5.

Table 5.4 - Nodes and element numbers for finite element meshes.

Model Number of Nodes Number of Elements

FE-50 1136 196

FE-100 1449 256

FE-150 2189 400

Table 5.5 - Tension stiffening parameters for different pullout models.

Specimen Parameter a1

P20-50 7.0

P20-100 8.0

P20-150 9.3

P40-50 12.3

P40-100 12.3

P40-150 10.6
5-24

y y

t
1· 12po~
~~~=zz:zz;JJ-IPortion used for I
I / FE modelling I
I
I

;
12 00 f- - --f<eelll~l--·
- - - -- X I

L50
'--1 OOX 1OOX20 steel
~~~g~~~U plate at the centre

---1 f--- 1000 --I f-- 50


I

I
~'-- ..._ ___ ..___ X

PLAN

Note:
Plane with nodes restrained
X-direction
Plane with nodes restrained
1
350
in Y -direction
- - - - - Plane with nodes restrained
in Z -direction
L y X

600---1
ELEVATION

Figure 5.14 - Typical finite element mesh showing boundaries


(shown for model FE-100).
5-25

y X y X

Model FE-50 Model FE-1 00

Model FE-150

Figure 5.15 - Three dimensional views of meshes used for different models.
5-26

The peak loads of the fmite element analyses are compared with the experimental

results in Table 5.6. It is seen that the finite element predictions are higher than the

experimental results in all cases but are mostly within 10 % of the experimental

results. The load versus deflection curves are plotted in Figures 5.16 to 5.21 for

different model of series P20-YYY and P40-YYY, respectively. The responses

predicted by the finite element analyses are seen in these figures to be within a

reasonable range. Typical contours of displacement in vertical directions are plotted in

Figure 5.22. It is observed that the values of the parameters {3 = 0.2 and az as given
Table 5.5 are generally suitable for the modelling of these problems.

5.6 Conclusions

For large apex angle the pullout in the plain concrete takes place with the average

angle of the failure cone much less than 45 degrees to the horizontal. This agrees with

the findings of Stone and Carino (1983). The ACI model with the effective stress area

taken at 45 degrees from the outside of the anchorage plate gave conservative

predictions of the failure load while the Nielsen et al. (1978) model was

non-conservative.

The finite element model gave reasonable prediction given the limitation of the model

in analysing fracture type problems. The finite element analyses verify that the values

of {3 = 0.2 and a 1 as given by van Mier (1987) are suitable for modelling of pullout

and punching type problems.


5-27

90
80
70
-60
~50
"C
40
_.! 30
-Experimental
20 -FE
10
0
0 0.5 1 1.5
Displacement (mm)

Figure 5.16 - Load versus deflection curves at point A for specimen P20-50.

160
140
120
z~ 100
:;- 80
~ 60 -Experimental
40 -FE
20
o~~~~~~~~~~~~~~~~~~~~

0 0.1 0.2 0.3 0.4 0.5 0.6


Displacement (mm)

Figure 5.17- Load versus deflection curves at point A for specimen P20-100.
5-28

250

200

--
~ 150
'C -Experimental
.9 100 -FE

50

o~~~~~~~~~~~~~~~~~;-~~~

0 0.1 0.2 0.3 0.4 0.5 0.6


Displacement (mm)

Figure 5.18 -Load versus deflection curves at point A for specimen P20-150.

140
120

--
z
~

'C
100
80
ca 60 -Experimental
0
..J
-FE
40
20
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Displacement (mm)

Figure 5.19 - Load versus deflection curves at point A for specimen P40-50.
5-29

250

200

--
~
"C
150
ca
.9 100 -Experimental
-FE
50

o~~~~~~~~~~~~~~~~~~~~~

0 0.1 0.2 0.3 0.4 0.5 0.6


Displaceemnt (mm)

Figure 5.20- Load versus deflection curves at point A for specimen P40-100.

400
350
300
z~ 250
:;;- 200
ca
.9 150
100 -Experimental
50 -FE
0
0 0.2 0.4 0.6 0.8
Displacement (mm)

Figure 5.21 - Load versus deflection curves at point A for specimen P40-150.
5-30

1.937E-3
l.S41E-3
.744E-3
.647E-3
.55E-3
'.453E-3

L.x .356E-3
l.259E-3
1.162E-3
1.649E-4
I

(a) P20-50

1.74E-3
t.G63E-3
'.597E-3
.511E-3
.434E-3
f .35BE-3
1.291E-3
$,285E-3
1.12BE-3
1.516E-4
I

(b) P20-150

Figure 5.22- Typical contour plots of vertical displacements (a) P20-50;


(b) P20-150; (c) P40-100; and (d) P40-150 (View of a quadrant
from top with the plate at the bottom left corner shown,
displacements in metres).
5-31

1.612E-3
1.549E-3
.4B6E-3
.423E-3
.359E-3
.2S6E-3
~.233E-3
%.17E-3
1.187E-3
1.43BE-4
I

(c) P40-100

i.!BSE-2
1.S47E-3
.834E-3
.721E-3
.SBBE-3
'.4S6E-3
~ .383E-3
l.27E-3
i.157E-3
1.442E-4
I

(d) P40-150

Figure 5.22 (Continued) - Typical contour plots of vertical displacements


(a) P20-50; (b) P20-150; (c) P40-100; and
(d) P40-150 (View of a quadrant from top with the
plate at the bottom left corner shown,
displacements in metres).
5-32

Table 5.6 - Comparison of peak loads between experimental and f"mite element
results.

Peak Load P u (kN)

Specimen FE!Expt
Experimental Finite element

p 20-50 73.0 80.2 1.10

p 20-100 129.0 140 1.09

p 20-150 207.0 227 1.10

p 40-50 107.0 120 1.12

p 40-100 171.0 207 1.21

p 40-150 306.0 337 1.10


Mean 1.12
Standard deviation 0.04
Coefficient of variation 3.68%
CHAPTER 6 - DESIGN AND CONSTRUCTION OF TEST SPECIMENS

6.1 Introduction

The test specimens were designed as a one-third-scale model of the prestressed flat

plate building shown in Figure 6.1. The portions of the slab at the edge and at the

comer column were modelled.

Four specimens were constructed and tested to failure with the primary variable being

the level of prestress. Specimens S 1, S2 and S3 were designed as edge panels while

Specimen S4 was design as a comer panel. Specimen S 1 was a control specimen and

had no prestressing. Specimens S2, S3 and S4 had unbonded prestressing in both the

longitudinal and transverse directions.

6.2 Design Philosophy

The design of the slab of the prototype and the models was based on the analyses of

the frames using the equivalent frame method and using the simplified method of AS-

3600 (1994). The worst case of the results from these analyses was considered to

design the slab. Using the simplified method of AS-3600 it is assumed that the column

strip and the middle strip take 70 and 30 percent of the load, respectively. The

specimens were developed as a 113 scale model of the prototype. The design of the

prototype and the model were carried out for a live load of 3.0 kPa. All the designs

were done with the concrete having the characteristic strength of 25 MPa and with

yield strengths of the conventional reinforcement and prestressing steel of 450 MPa

and 1250 MPa, respectively. A 10 mm clear cover was used in the model slabs. To

give bias towards the punching shear failure the flexural reinforcement was increased

by 50 percent in all slabs over that required for the design moments.
6-2

0 0 0 0

-77~
Region undbr
stud~!
0 0 0 0

L __ Lj All columns 600x600

0 0 0 0
1
J
L,2ooo •I• 12000 I I I
12000 I I8 12ooo----J

TYPICAL FLOOR PLAN

o---- p;~u~n-i:ind~rl
stuay l
-----------"

SECTION 1-1

Figure 6.1 - Prototype Oat plate building.

The design of the prototype and the model slabs are given in Appendix A. For slabs

S 1, S2 and S3 the slabs were designed for an average prestress of 0.0, 1.0 and 2.0 MPa

respectively. For the corner panel slab the average prestress was 2.1 MPa.

The columns of all specimens were overdesigned so as not to fail before the slab. A

200 mm x 200 mm section was used for the test slabs and contained eight 16 mm
6-3

diameter 400 grade hot rolled deformed bars. A 6 mm diameter diamond arrangement

was used for the column tie reinforcement and the ties were spaced at 80 mm centres.

Details of the reinforcement arrangements for the test specimens are presented in

Section 6.3.

The boundary of the slab was taken at the second point of the contraflexure away from

the design connection as recommended by Rangan and Hall ( 1983) and Rankin and

Long (1988). The normalised bending moment and shear force distributions for the

model and the specimen slabs with edge and comer column (based on the frame

analyses) are shown in Figures 6.2 and 6.3, respectively. The size of the column was

determined to simulate the stiffness of the prototype column framing into the

connection. Further details of the column sizing are given in Appendix A.

The dimensions of the edge connection specimens S 1, S2 and S3 are shown in

Figure 6.4. The slabs were 3565 x 4900 mm and 100 mm thick. The 553 mm long

200 mm square column was cast monolithically with the slab at the middle of the long

edge. The three sides of the slab away from the column support were supported on the

!-sections of the testing frame with a bearing width of 100 mm. To carry the

equilibrating horizontal load needed to develop a moment in the slab-column

connection, the slab along the long edge opposite the connection was attached to the

supporting frame via shear connectors (refer Section 6.5). The column base rested on a

load cell mounted within a steel box connected to the load frame.

The specimen for the comer column connections (Specimen S4) consisted of a

3565 mm x 2550 mm by 100 mm thick slab, as shown in Figure 6.5. The 553 mm
6-4

2000 1565 435


1.27
,l'\
/
Key:
--Specimen
-------- Model
00

NORMALISED B.M. I End with column


in the specimen

3565 435

NORMALISED S.F.

(a) Longitudinal direction

Key:
--Specimen
----------Model
NORMAliSED B.M
End with column
I in the specimen

NORMALISED S.F.

(b) Transverse direction

Figure 6.2 - Normalised bending moment and shear force diagrams comparing
the prototype slab design with the specimen design for edge
column specimens Sl, S2 and S3.
6-5

2000 1565 435


1.25
1.0
0.9

'
,//' Note:
Specimen

~ 0.7
-~ Model
End with column
in the specimen
00

NORMALISED B.M.

3565 435

~--------~0010-----------~

NORMALISED S.F.

(a) Longitudinal direction

2550 450

Key:
--Specimen
-------- Model
NORMALISED B.M. End with column
in the specimen

2550 450

'-------.JOOD-------'

NORMALISED S.F.

(b) Transverse direction

Figure 6.3 - Normalised bending moment and shear force diagrams comparing
the prototype slab design with the specimen design for the corner
column specimen 84.
6-6

~--------------3565--------------~____t

200X200 mm
COLUMN
SECTION 1-1
Hinge~ sup:~ L 100

200

ylll
A 35 65-----•-'1 B
-r----- '=-=~=--=~=--=~=--=~=--=~=--=~=--=~-=-=~-=-=~1=,
,,
,,II
II,,

200X200 mm
-COLJJK4N_________ XI
4900

,,_j
:I,,
II
100 Ji'
~ 100
------------=-=:-.=..=:-.=..=:-J
D
L-,, c

49eo
" "
. . "
~
6
Simpl e support Simple su
~ 200mmX200mm

1 .. 2350 -- ~
Column
'--200
2350 -I
SECTION 11-11

Figure 6.4 - Dimension of slab and column of specimens Sl, S2 and S3.
6-7

Ill
y z
~----3565----~1

-, }---------2550--------~
----l!r----1 00
:I
!i
0

~
• •...

Hinged support
.
H1nged support
~ x::t
100
I
l ~200 X 200
!!JI
?.!"" COLUMN
mm :I
I'
I X
200 -----t -ll
2350 - - - - - - - - - - - 1..
0 SECTION 11-11
II _-:j

z
~-------------3565------------~~ I100

E
6
O • H;ogod '"PPOrt
1--200 SECTION 1-1
• Hinged support Xl

Figure 6.5 - Dimensions of slab of specimen S4.

long by 200 mm square column was cast monolithically with the slab at the comer of

the free edges. The two sides of the slab opposite the column connection were

connected to the testing frame via shear studs (refer Section 6.5). The bearing widths

of the slab on the supporting edges were 100 mm.

6.3 Arrangement of Reinforcement

Details of the bonded reinforcement are given in Figure 6.6 and Table 6.1 for

specimens S1, S2 and S3 and in Figure 6.7 for specimen S4. The corresponding bar

bending schedules for ~ese specimens are shown in Figures 6.8 and 6.9. In all slabs

the top reinforcement normal to the hinge supported edges were cut off to 80 mm from

the edge of the slab.


6-8

y
3565
I
A B

~~~to~
4900 r--X
IL
0
I
_j
A

D
I I
c
D c
8

(a) Bottom reinforcement

y
3565
I
A 8
I ~
H
r--"'- •·f--------1----
I I
G
I I
I I
I I
I
I
~ I
I
I I{) I
200x200 I N I
COLUMN I I
@ 1--X
4900
0~ :
I

IL --r- ~~
I >- ------r--- _j
I
I
I
I I
~

I I
I E
I I
I I
I I
I I
I I
D c
F

For slob:
CLEAR COVER = 10 mm (Top and bottom)
CLEAR COVER = 25 mm (Sides)

(b) Top reinforcement

Figure 6.6 - Arrangement of bonded reinforcement for specimens Sl, S2 and S3.
6-9

z
. 3515
I I 80 mm Studs
@ 2 00 mm c/c
1 _l

II
r
L653
II
jll
___L:::,_
I X
~

L A
R6 @ 80 mm (2 Sets)
8 Y16
For sl ob:
CLEAR COVER = 10 mm (Top and bottom)
CLEAR COVER 25 mm (Sides)

For column:
CLEAR COVER 20 mm (Side)
SECTION 1-1

200
r
l
l--2oo__j
SECTION 11-11

(c) Sections

Figure 6.6 (Continued)- Section detail of bonded reinforcement for specimens


Sl, S2 and S3.
6-10

J. 3565

I
! 9 Y10 @ 275 mm
- I
~- ~-- ~----- .... -~-l 7 Y10 @ 160 mm

I I 10 Y10 @ 110 mm
2 550 : l : 9 Y10 @ 275 mm
IL
I
I
I
:
I
_jl
200x200
COLUMN
r'J
- ~
- - ----:----
~ c
T
X
0 10 Y10 @ 80 mm
6 Y10 @ 285 mm
--Top bars 200 Y16
7 Y10 @ 125 mm
Bottom bars

PLAN
6 Y10 @ 280 mm

1I--
'-----------'
___j
200
SECTION 11-11

Totol 13 Y10 _j_


80 mm Studs

G
X @ 200 mm c/c
!QQf
II L653 Y10

L 2 Sets)

CLEAR COVER
CLEAR COVER
10 mm (Top and bottom)
25 mm (Sides)

SECTION 1-1

Figure 6.7- Bonded reinforcement for specimen S4.


6-11

BAR BENDING SCHEDULE (Not to the scale)

Clear cover = 10 mm for slab and 20 mm for column


Clear cover on sides = 25 mm (Minimum)

1. Bottom bars along X-direction


300
60 ~----------------------------------~60
3505
2. Bottom bars along Y-direction

60 60
4840
3. Top bars along X-direction (a) Bar away from column (b) Bar within column region
3455
(a) 60
300
3455
(b)
400

4. Top bars along Y-direction

4800

5. Column lonitudinal bars 6. Lateral ties

155

Figure 6.8 -Bar bending schedule for specimen Sl, S2 and S3.
6-12

BAR BENDING SCHEDULE (Not to the scale)

Cover = 10 mm for slab and 15 mm for column ties


Cover on sides for slab = 25 mm

1. Bottom bars along X-direction:


100

2. Bottom bars along Y-direction:


100
50 c==
2490

3. Top bars along X-direction:

soc== 3455
100

4. Top bars along Y-direction:

soc== 2440
100

5. Top bars over column:

2440/3455

6. All column bars: Same as for other specimens

Figure 6.9 - Bar bending schedule for specimen 84.


6-13

Table 6.1 -Bonded reinforcement detail for specimens Sl, S2 and S3.

Bonded Reinforcement Details

Bar Location Bar Mark Sl S2 and S3

A 26Y10@60 26W8@60

Bottom Bars B 20Yl0@170 16W8@210

c 6Y10@200 6Y10@200

D 7Y10@340 10WB@240

E 13Y10@125 13Y10@125

Top Bars F 10Y10@320 16W8@200

G 12Y10@90 13Y10@90

H 10Y10@240 11W8@220
6-14

The prestressing ducts and the tendons were arranged as per the design given in

Appendix A with small variations to avoid conflicts with the bonded reinforcement.

The tendon arrangements for slabs S2, S3 and S4 are given in

Figures 6.10, 6.11 and 6.13 and are summarised in Table 6.2. The details of the cable

profiles for tendons in specimens S2 and S3 are shown in Figure 6.12. The

reinforcement is arranged in all specimens using the design outcomes presented in

Appendix A.

6.4 Construction of Formwork, Test Specimens and Assembly

The formwork for the slabs consisted of 18 mm thick timber plywood to form the base

of the slab and 100 mm x 50 mm timber for the edges. The plywood was rested on the

adjustable rectangular hollow section steel beams that were propped at regular

intervals (refer Figure 6.14). All specimens were constructed in place on the testing

frame, shown in Plate 6.1, which was fabricated as a stiff I-beam frame tied to the

laboratory reaction floor.

First the conventional bonded reinforcement was cut and bent to the required shapes

and sizes. The column reinforcement cage was prepared and placed in position with

the load cell assembly. The bottom layer of bonded reinforcement was laid in position

followed by the top layer of the bonded reinforcement. All the crossing points of

bottom x- and y- direction bars were tied. Next the ducts and the prestressing wire

were placed and tied in position. The test assembly and reinforcement prior to the

concrete pour are shown in Plates 6.2 to 6.5 for slabs S 1 to S4, respectively.
6-15

85516851

420
640

200x200
column
640
5
362~ /
/
v- Central cable
3mm off

--x
388
375
640

640
420

~1-~----3565-------~1
Effective Prestresses:

X-direction wires: 860 MPa


Y-direction wires: 990 MPa

Figure 6.10 • Tendon arrangement for specimen 82.


6-16

y
! 1150
2365 - - - j
7@320 mm I

120~ ~

rE
ro
0
E
0
CX)
0
1'- N

!-~¢ 0

""'""
h
I-'
X
~§ 4900

ro E
0
1'- 0
..- N

l~
.. I
~1--- 3565 --.....-~
Effective Prestresses:

X-direction wires: 990 MPa


Y-direction wires: 1050 MPa

Figure 6.11 - Tendon arrangement for specimen S3.


6-17

1 0
y

Lgso--~1~.,__-15oo--~,__-15oo--~~-gso_J
L~~---------------------4900---------------------_J~
z
CABLE PROFILE ALONG LONG SPAN
·~

J ------
-vi
18 r
'.J

1.
----- 1(0

-
~ X

CABLE
2000

PRO~LE
3565
ALONG SHORT SPAN
1565
:I
Note: Not to the scale

Figure 6.12- Cable profile for specimens S2 and S3.


6-18

y =~
l~-oo~------3565------- _[205

J
0 L, 8@125
I = _____j 23BOe------'•l
7@340
t_,, '
_j L 120
TENDON LAYOUT PLAN

356~
I
r-18
_j X __j
0 -;QJ
200n 156

~ SECTION 1-1

z Note: Sections not to the scole

rso 255n r-1R

t 0
y
150n 105o--

~ SECTION II - II

Effective Prestresses:

X-direction wires: 1000 MPa


Y-direction wires: 1000 MPa

Figure 6.13 - Arrangement and profile of tendons for specimens S4.


6-19

Form work
(18mm thick plywood)

Timber section bolted to


1-section to support
Edge 100x50mm timber form work
expandable beam system
bolted to steel sections

I Lc:~,~ti1 ~on--be_a_m--~~------~------~~~------~~------~n
2400 connected to frame

Laboratory floor

Figure 6.14 - Cross-section of the formwork.

Table 6.2 - Prestressing details of specimens.

Width of the Number of Effective Average


~
......
u
SlabL
(m)
Prestressing
wires
prestress
(MPa)
Prestress, Pet/A
(MPa)
Q.)
0..
(/.)
X-Dir. Y-Dir. X-Dir. Y-Dir. X-Dir. Y-Dir. X-Dir. Y-Dir.

S2 3.565 4.900 9 5 860 990 0.80 0.70

S3 3.565 4.900 22 15 990 1050 2.27 2.25

S4 3.565 2.550 11 15 1000 1000 2.20 2.15


6-20

Plate 6.1 - Slab testing frame.

Plate 6.2- Specimen Sl before concrete pour.


6-21

Plate 6.3 ·Specimen S2 before concrete pour.

Plate 6.4 ·Specimen 83 before concrete pour.


6-22

Plate 6.5- Specimen 84 before concrete pour.

All the specimens were cast with commercial ready mix concrete with an electric

needle vibrator used to compact the concrete. The surface of the slab was levelled

using wooden screeds and the surface was finished using steel trowels. Concrete

cylinders and prisms were cast with the slabs as control specimens.

The control specimens were stripped the day followed the concrete pour, wrapped in

hessian and kept with the slab specimens. One day after the concrete was placed the

surface of the slab was covered with the wet hessian and plastic and, with the

exception of specimen S4, the hessian was watered twice daily for a minimum of 7

days. Slab S 1 was stripped after 30 days and slab S2 and S3 after 7 days. Slab S4 was

cured for 6 days before the formwork was stripped.


6-23

6.5 Boundary Supports

The boundary supports were chosen to simulate that of the prototype structure. For

hinged edges the movement of the slab in horizontal direction was restrained by using

the shear connectors welded to the top flange of the !-beams of the testing frame, as

shown in Figure 6.15. The shear connectors were 78 mm long, 13 mm diameter with

7 mm thick by 15 mm diameter circular heads. The connectors were spaced at 200 mm

centres and an 8 mm diameter reinforcing bar was placed behind the connectors

towards the edge of the slab. In specimens S 1, S2, and S3 the connectors were welded

along the long edge opposite to the connection. In specimen S4 the shear connectors

were welded along both of the edge supports. The shear connectors provided an

equilibrating horizontal reactive force to the horizontal reaction developed at the base

of the column.

The base of the column was connected to a 20 mm thick steel plate with the

longitudinal column reinforcement welded to the plate. The column base plate was

fitted to a 115 mm long, 70 mm diameter steel boss which was connected to the load

cell assembly. In specimens S1, S2 and S3 the edge column specimens were

symmetrical about the x-axis and no horizontal support was provided to the short

edges of the slab. Thus, no horizontal y-direction reaction (refer Figure 6.4) could be

maintained by the column support. Pins were provided along y-direction to maintain

the column in position while the load cells were assembled on both sides of the boss to

measure the horizontal x-direction reactions. The base of the boss sat on an assembly

with the load cell that measured the vertical reaction in the column. The load cells and

pins were co_nnected to a box support via adjusting screws which were tightened to

hold the assembly in position.


6-24

f ~Rx
--l
200mm X 200mm
Column

l----200
Rz
DETAIL A

t
Studs @ 200 Welded
to 1-Section
1-Beam of
test frame

DETAIL A-HINGED SUPPORT DETAIL FOR SLAB

Figure 6.15 - Details of hinged support connection.

The column support assembly for specimen S 1 was designed as a hinged support

capable of measuring the vertical and the horizontal (x-direction) reactions. The steel

section attached with the load cell measuring the vertical reaction was fitted to a

circular groove at the base of the box, as shown in Figure 6.16. Post analysis of the

specimen S 1 results, however, showed that while the vertical reaction measurement

was as expected a significant portion of the horizontal reaction was transferred to

testing frame via the base connection of the vertical load cell and, thus, was lost to the

horizontal load cell measuring system. The problem was that since the steel section

holding the load cell to measure the vertical section was fitted in the groove, the lower
6-25

..

(a) Vertical section

1 :Central vertical steel bass


2:25mm Die. ball bearings
3:Vertical wall of box
4:Verticol stiffeners
5:Bolts
6:Bose plate of box
7:Lood cells to measure horizontal
reactions
B:Steel sections
9:Steel pins

l=!~;;;f"'o---oY16
Long. bars
t--it---oW6 (2Sets)

base plate
X
115mm long 70mm dia. boss
fitted to 105x 105x25mm plate

all bearings connection slots

(b) Plan (c) Column base connection

Figure 6.16 - Support assembly for specimen Sl.


6-26

portion of the assembly could not move freely in horizontal direction and a part of the

horizontal reaction developed was transmitted to the base. The column support

assembly was modified for specimens S2 and S3. For specimen S2 and S3 the vertical

load cell was connected to a roller system, as shown in Figure 6.17 and Plate 6.6.

The column support assembly used for specimen S4 is shown in Figure 6.18 and

Plates 6.7 and 6.8. The assembly was modified to that used for the edge column

connection tests to account for the lack of reaction symmetry in the corner panel test.

Load cells were added to the assembly to measure the support reactions in both the x

and y directions (refer Figure 6.5). The roller support at the base of the vertical load

cell (used in specimens S2 and S3) was replaced with 35 by 25.4 mm diameter high

strength steel ball bearings. The ball bearings were positioned uniformly around the

base of the assembly and located using an 8 mm thick timber plate with the ball

bearings placed in oversized holes drilled in the timber plate.

All the load cells placed to measure the horizontal and vertical column reactions of all

specimens except a load cell used to measure the vertical reaction in specimen S 1 were

of 200 kN capacity. The load cell used to measure the vertical reaction in specimen S 1

was of 500 kN capacity.

6.6 Prestressing

Specimens S2, S3 and S4 were prestressed to different degrees in both the x and the y

directions. In all cases the slabs were post-tensioned the day before the test to

minimise the loss of prestress due to relaxation. Twenty millimetre thick steel bearing

plates of 80 mm x 90 mm were placed at each end of the tendons with jacking from

one end only.


6-27

..

Figure 6.17 - Support system at the column base for the specimen S2 and S3.

Plate 6.6 - Load cell and pin arrangement for specimens S2 and S3.
6-28

..
5 0

-14--20 mm thick Steel Plate on which


---""'"'n .n Long. bars ore welded

~~~t~~tj--25 mm thick plate over 35-25mm


diameter hightensile ball bearings
(Refer plan for detail arrangements of bearings)

'----tsox welded to top of I Beam

(a) Vertical section

1:Central vertical steel bass


2:25mm Dia. ball bearings
3:Vertical wall of box
4 Vert1cal st1ffeners
S·Balts
6:Base plate of box
7:Load cells to measure horizontal

~ ~
reactions
8:Steel sections
9.Steel pins
225

-r-i-6:····•
• ••••••••
...._- 25 4mm diameter
boll bear,ngs

•••••
245 • • • 27mm hales

X
••••••
••••• L
8mm thick l1mber
plate

1451-

(b) Plan (c) Schmatic plan of ball bearing assembly

Figure 6.18- Support system at the column base for the specimen S4.
6-29

Plate 6.7 - Arrangement of ball bearings at the bottom of specimen S4 column.

Plate 6.8 - Load cell arrangement for specimen S4.


6-30

The required load for prestressing was determined with allowances for the loss due to

draw-in and relaxation. The extents of these losses were determined by carrying out

relaxation tests on the tendons in a similar environment to that of the slabs. Details of

the relaxation test and the calculation of losses are given in Appendix B. A 200 kN

capacity hydraulic jack was used to prestress in the tendons. The jack was placed

between the prestressing yoke and an anchorage plate, as shown in Figure 6.19, and a

load cell placed between the plates and connected to a HBM amplifier. The oil was

pumped to the jack until the required load was reached. During the process the wedges

of the prestressing grip were tapped gently, regularly and evenly into the barrel. Once

the desired prestressing load was applied the wedges were tapped into position.

The tendons were prestressed individually starting with the wire farthest from the

column. The jacking and nonjacking ends were alternated along each side of the slab

to distribute the prestress uniformly through the slab. The prestressing sequence is

shown in Figure 6.20 with the numbering in Figure 6.20 shown at the jacking end.

6.7 Loading arrangements

6.7.1 Specimens Sl, 82 and 83

The use of a whiffle-tree to apply a distributed loading has been successfully used in

similar tests (Bums and Hemakom,1977, Lim and Rangan, 1992). Specimens S1, S2

and S3 were tested using twenty four concentrated loads applied via three identical

hydraulic jacks placed in parallel through the three sets of whiffle-trees. Each of the

jacks had a capacity of 270 kN and a maximum stroke of 150 mm. Three outlets from
6-31

Anchorage Plate
(80x90mm)

\_Barrel and
Dead end Jacking end wedge

200mm long prestressing yoke


Slab

Figure 6.19 - Prestressing set up.

the hydraulic pump regulated the three jacks with identical pressure in each of the

lines. The jacks were calibrated for pressure versus load in a Shimadzu Universal

Testing Machine with the results shown in Figure 6.21. The maximum variation

between the jacks is less than 3.0 percent. The arrangement of the loading points on

the edge panel slabs S 1, S2 and S3 are shown in Figure 6.22. The whiffle-tree

arrangement shown in Figure 6.23 and Plate 6.9.

The cables from the whiffle-tree were attached to the loading points of the specimen

slab through the 20 mm holes using 20 mm diameter PVC pipes cast into the slab. The

load was applied to the top surface of the slab via 20 mm thick by 100 mm square

bearing plates. The arrangement of the cable system to apply load on the slab via the

loading tree is shown in Figure 6.24. A load cell was placed in one branch of the

whiffle-tree (see Figure 6.23) at the middle tier to measure of load applied to the slab.

The load cell measured 116 of the total applied load.


6-32

y y
0 G)
®®:5® @ ® @)
0 CD
6

@
@ 10
Q
Column
@
'-.../
X ~2
16
14
18
20
@ - - - -- -- --- x(z2)_

~
0 -
-c0
® 5
1
@ 3
G) 9

® 5

~@® @@

(a) Specimen S2 (b) Specimen S3


y

H+H++H~~~~~-+-++~
~+H+H+H~-+-+-+~~~
H+H++H~~~~~~~+~

(c) Specimen S4

Note: The number inside the circle indicates the sequence of prestressing and are

shown on the jacking ends of the tendon.

Figure 6.20 - Sequence of prestressing.


6-33

250

200
-z
-
~

"C
ca
0
150

100
--Jack#173
-----Jack#174
..J ········ Jack#178
50

0
0 10 20 30 40 50
Pressure (M Pa)

Figure 6.21 - Calibration of jacks.

---="'--
J+ck~
11)

+ + """

200X200 mm column
+ + + +
+ + + +
0 4700
+ + + +
1 00 mm wide bearing on
steel 1-section frame all + + + +
round the edges except 0
the one with the column
+ + + + 0
11)

NOTE: ----'-'r
Refer other drawings for details
of studs in 1-section and other
details.
565~ 800 ~ 800 1 800 1600
3565

Figure 6.22 - Arrangement of loading point in the slab (Specimens Sl, S2 and 83).
6-34

~--Cables attached
to slab loading
points

sections

Rod attached to
hydraulic jack

Note:- Not to the scale

(a) Schematic drawing of a set of the whiftle-tree for specimens Sl, S2 and S3.

..
200x200mm
column

Barrel and wedge

930x 130x25mm
ms plates welded
to RHS ot top
and bottom
RHS 203x 102x6.3

Barrel and wedge

(b) Front elevation of whiftle-tree system for specimen Sl, S2 and S3.

Figure 6.23 - Whiftle-tree system.


6-35

Plate 6.9 - The loading system (Specimens Sl, S2 and S3).

1OOmm squa
20mm thick plate

<1

4
<1
<1

Figure 6.24 - Section through the slab showing connection of slab to the
whiftle-tree.
6-36

6.7.2 Specimens S4

Specimens S4 was tested using the sixteen concentrated loads applied via two identical

hydraulic jacks placed in parallel through the two sets of whiffle-trees. The

arrangement of loading points for the slab S4 is as shown in Figure 6.25. As for the

edge panel specimens, holes were cast in the slab to fit the loading cables. The loading

system for specimen S4 was similar to that used for the other specimens and is shown

in Figure 6.26. A load cell was located in the middle of one of the whiffle-trees

measuring IA of the total load. The arrangement of the whiffle-tree to apply the load in

the specimen is shown in Plate 6.10.

6.8 Material Properties

6.8.1 Bonded Reinforcement

Specimen S 1 contained only conventional reinforcing bars while all other specimens

contained both bonded deformed bars and unbonded prestressing wires. The bonded

reinforcement consisted of hard-drawn deformed bars of 8 mm and 10 mm nominal

diameter. The longitudinal bars of the column consisted of 400 grade hot rolled

deformed bars of 16 mm nominal diameter while the ties in the column were

fabricated from 6 mm diameter hard drawn wire. Tensile tests on the bars were

undertaken using an Instron Universal Testing Machine with the average results of

three specimens of each sized steel bars given in Figure 6.27 and Table 6.3.
6-37

100 mm wide bearing ----------1r------,


l?nlzi•~1 ~~6~5~~1~7Z2~1~80~0~~~1
7
on steel !-section frame ~
all round the edges except o o
those with the column F2 + + + +
0
L[)
L[)
~-~ ~cks ~
+ + +
N
NOTE: + + + +
Refer other drawings for details -~
of studs in !-section and other ~ + + + +
~OOX200 mm column L[)
details. tO
N)t .,. X
+: Load points
0
565 600

3565

Figure 6.25 - Arrangement of loading points in specimen S4.

800

Tub u Ia r-L---< iubular sections


1600
sections

Rod attached to
hydraulic jack

Note: Not to the scale

Figure 6.26 - Schematic drawing of a set of the whiffle-tree for specimen S4


(Looking towards y-direction).
6-38

(a) Looking towards y-direction

~-········1..~-·:t>
l.ft• ~-
; ........ * ........ .

(b) Looking towards the column along the x-direction

Plate 6.10 - The loading system used in specimen S4.


6-39

800
700
'iS 600
~ 500
;; 400 -------- Y16

-e
U)
3oo
en 200
100
0 ~~~~~~~~~~~~~~~~~~~~~~

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07


Strain

Figure 6.27 - Stress versus strain curves for bonded reinforcement.

Table 6.3 - Reinforcing steel properties.

Bar Diameter Area /sy Es


(mrn) (mrn2) (MPa) (Gpa)

W8 7.3 41.5 585* 240


YlO 10.1 80.5 675* 225
Y16 16.0 201 450 200
W6# 6.0 28.3 520 200
Note: * 0.2 % Proof stress

6.8.2 Prestressing wire

Specimens S2, S3 and S4 were prestressed in two directions. An 8 mrn diameter

carbon steel spring wire, conforming to AS 1472-1991 and marked as S8 was used for

the prestressing tendons. The stress-strain curve for the steel was obtained using an

Instron Universal Testing Machine and the average of the three results are given in

Figure 6.28 and Table 6.4.


6-40

1400

-ca
1200
c. 1000

-.
==0 800
0
Cl)
600
u; 400
200
0
0 0.005 0.01 0.015 0.02 0.025
Strain

Figure 6.28 - Stress versus strain curve for prestressing wire.

Table 6. 4 - Properties of prestressing wire.

Wire Diameter Area /sy Es Pareaking

(mm) (mm2) (MPa) (GPa) (kN)

S8 8.1 51.0 1030# 200 74.0

Note: #0.2% Proof stress


6-41

6.8.3 Ducts and Accessories

Corrugated plastic conduit, normally used for the electric wiring, was used as ducts to

house the prestressing wire in the specimens. The outside diameter of the duct was
16 mm and the clear inside diameter was 13 mm. A section of the duct is shown in

Figure 6.11.

Two types of barrels and wedges (yv2185M and W999L manufactured by CCL

System Ltd., United Kingdom) were used to grip the 8 mm diameter prestressing wire
and for applying the load to the slab via 12.7 mm diameter tendons connected to the
whiffle-tree systems and pulled down from the reaction floor. All the reinforcement

including the ducts for the prestressing were kept in position by using PVC and steel
bar chairs.

6.8.4 Concrete

Ready mix concrete delivered by a local supplier was used for all test specimens. The
specifications for the concrete were a maximum aggregate size of 10 mm, a slump of

80 mm and a target 28-days concrete strength of 25 MPa for specimen S 1 and 32 MPa

for specimens S2, S3 and S4. The target mean cylinder strength at time of testing was
25 MPa. The actual strengths at the time of testing are given in Table 6.5.

Seventeen cylinders of 150 mm diameter and 300 mm high and twenty cylinders of

size 100 mm diameter and 150 mm height were cast along with each specimen. The

cylinders were used to monitor the concrete strength gain with time for prestressing

and testing. The prestressing was done the day before the test when the concrete

strength was approximately 25 MPa. Specimen S4, however, showed a rapid strength

gain and the slab was prestressed at a concrete strength of approximately 37 MPa with

the slab tested the following day.


6-42

Plate 6.11 - Corrugated plastic ducting.

On the day of testing of the slab a minimum of six cylinders were capped with a

sulphur-ash capping compound and tested for the compressive strength. The tests were

undertaken in accordance with to AS 1012.9-1986 applying the load at a rate

corresponding to 20 MPa per minute. Two or three 150 mm x 300 mm cylinders were

tested for indirect tension, in accordance with AS1012.10-1985, at a loading of

1.5 MPa per minute. For specimens S 1, S3 and S4 three cylinders were tested in

uniaxial compression to obtain the stress-strain properties of the concrete. The load

was applied sat a slower rate of strain than for the standard tests, with the testing time

being about 7 minutes and the specimens gauged over a 150 mm interval. The stress-

strain plots of the concrete are presented in Figure 6.29. Three flexural prism tests

were conducted for the specimen S4 with the average modulus of rupture being 4.5

MPa. A summary of the material properties for the concrete is presented in Table 6.5.
6-43

40 4
35
-30
ca
a. 25
:E
,
';" 20
! 15
ti)
10
5
0
0 0.0005 0.001 0.0015 0.002 0.0025 0.003
Strain

Figure 6.29 - Stress versus strain relations for concretes in uniaxial compression.

Table 6.5 -Concrete properties for different specimens.

Specimen Age r /em !t Ec


(Days) (kg/m3) (MPa) (MPa) (GPa)

S1 41 2250 24.1 3.07 20.2


S2 15 2380 24.9 2.75 -
S3 10 2320 26.3 3.07 22.2
S4 9 2700 37.5 2.02 33.0
6-44

6.9 Instrumentation

6.9.1 General

The applied load, reactions, deflections and steel strains were measured in all tests. In

this section details of the instrumentation used in the test are given. The output from

the electronic measuring devices was filtered through a signal conditioning amplifier

and recorded using Labview (1999).

6.9.2 Measurement of Loads and Reactions

Load cells of 200 kN capacity were mounted in the whiffle-trees as discussed in

Section 6.7 and used to measure the applied loads. The horizontal and vertical

reactions at the base of the column, as shown in Figure 6.30, were measured using 200

kN or 500 kN load cells and mounted as discussed in Section 6.5.

6.9.3 Displacements

Displacements were measured using linear variable differential transducers (LVDT' s)

at five points for slabs S 1, S2 and S3 and three points for slab S4. The locations of the

LVDT's are given in Figure 6.31. The displacement transducers were linked to the

main data acquisition control unit and recorded using Labview (1999). The reading

accuracy of all transducers was 0.01 mm and their maximum travel range was

100mm.

For specimen S 1 mechanical dial gauges were installed at the top comers of the slab to

measure uplift and to the column base assembly unit to measure any horizontal

movement of the testing frame. Negligible movement (less than one division) was

recorded in any of the gauges. The dial gauges used were sensitive to 0.00254 mm

(0.000 1").
6-45

,___,
II
3565
~
---------- ----- --":..---:..-,
_L.-1•e------3565------'1
I

R~flum:'ll;:::.==R=x=z====,O=Oq~
I•

I
653] fRz
SECTION 1-1

J
200X200mm
It
~OLUMN
48 0 Rx1 r--' RJQ 47 00
00
~2550---1[
1
!i
I•
H
I•
:I
ii
·· colu~n T
J -so
I
653I Rz
: ---------- --=---=-..J
I SECTION 11-11
11--

(a) Specimens Sl, S2 and S3

Ill 1-4---

X
column
Rx1 Rx 2 100
200X200 mm 653 SECTION 1-1
COLUMN

xz X z

2550---l yr100
PLAN column
Ryz Ry,
653 Rz SECTION 11-11

(b) Specimen S4

Figure 6.30 - Reactions measured at the column support (a) specimens Sl, S2 and
S3; and (b) specimen S4.
6-46

Coordinates(mm):

B:A:
!1758,0,0)
1758, 1200,0)
c. 1758,-1200,0)
D: 50, 1200,0)
E: 50,-1200,0)
~~OL~M~-----tA-------- X

z
(o.o.o3~rr: ======::J....___,..
_ x

(a) Speciemns Sl, S2 and S3

Coordinates(mm):
A: ~1700,1300,0)
,._c B: 1700,50,0)
C: 50, 1300,0)

~
~LC~OL~U_MN__~~B______-L---X
z

( O,O,O~~f-;======:::::J._-- .. X

(b) Specimen S4

Figure 6.31 -Locations of LVDT's (a) specimens Sl, 82 and 83;


and (b) specimen 84.
6-47

6.9.4 Strain Measurements

Strains were measured using electronic strain gauges at key locations on the bonded

reinforcement. The locations of the strain gauges for the slab are shown in Figure 6.32.

Approximately 20 mm length of the top of the bars was grinded to make a plane

surface for fixing the gauge. Care was taken to grind the projected rib of the bar

without causing any appreciable loss of cross sectional area. The gauges were glued to

the surfaces of the reinforcing bars and were covered with a silicon compound. The

gauges were connected to a HBM signal conditioning amplifier which was connected

to the data acquisition system. In some cases the strain gauges were damaged during

concrete pour. Table 6.6 lists gauges that were damaged (X) and operational (./) at

the time of testing.

6.10 Testing Procedure

A small load (less than 0.5 kPa) was applied to each slab to set the loading system and·

test the electronic measuring equipment. The load was then released, the gauges

zeroed, and the test started. The load was applied slowly by applying pressure in the

jack in the increments of approximately 0.7 kPa on the slab. The test was paused for

approximately two to five minutes between each load step. Measurements were

recorded electronically via the data acquisition unit. In specimen S 1 the top surface of

the slab was checked between load steps for surface cracks and marked as appropriate.

No surface cracking was measured for the prestress slabs due to safety considerations.

The test was continued beyond the punching shear failure for specimens Sl. However,

for specimens S2, S3 and S4 the tests were stopped as soon as the punching shear

failures occurred. The time taken to test each specimen was approximately 2 to 3

hours.
6-48

y y

Coordinates(mm):
1: 221, o. 75)
2: 221, 125, 75)
3: 221, -125, 75) 8
4: 115, 121, 85)
5: 115, -121, 85)

2 COLUMN 6
o~5eE==========~-x
(0,0,0,)3
0
(0,0,0,) 7
X

Coordinates(mm):
6: !1637, 30, 25)
7: 1637,. -30, 25) 9
8: 225, 1563, 15)
9: 225, -1563, 15)

TOP BARS BOTTOM BARS


z

(o,o,o3~ • X

(a) Specimen Sl

y y

Coordinates(mm}:
1: ~221, -10, 75~
2. 179, -10, 75
3 r21. 125. 75) 9
4 221, -125, 75)
COLUMN 5 115, 121. 85)
6 115, -121, 85)
2-<:: 3 COLUMN 7
06 X 0 X
(0,0,0,) 4 (0,0,0,) 8
Coordinates(mm}:
7, r637. 30. 24)
8: 1637, -30, 24) 10
15~
9: 230, -1563, 15
10: 230, -1563,

TOP BARS BOTTOM BARS


z

(o,o,o3~ • X

(b) Specimen S2

Figure 6.32 - Locations and numbering of strain gauges in different specimens


(a) specimen Sl; (b) specimen S2; (c) specimen S3; and
(d) specimen S4.
6-49

y y

Coordinates(mm):
1: (230, 0, 75)
2: (179, 0, 75)
3: !230.0, 125.0, 75.0) 9
4: 230, -125, 75)
COLUMN 5: 120, 120, 85)
6: 115, -120, 85)
2-...<:; 3 COLUMN 7
06 X 0 X
(0,0,0,) 4 (0,0,0,) 8
Coordinates(mm):
7: t1637, 30, 24)
8: 1637, -30, 24) 10
9: 230, -1563,15)
10: 230, -1563, 15)
TOP BARS BOTTOM BARS
z

(o,o,o~r.r:======~•--'X
(c) Specimen S3

Coordjnqtes(m m) · y
y
1: ~221, 30, 75)
h----------, 2: 221, 155, 75)
3: 30, 221, 85)
4: (190, 221, 85)

Coordingtes(mm)· 8
5: ~1637, 30, 25) 7
6: 1637, 110, 25)
7: 1463, 750, 25) COLUMN 6
~~~~5~~~~~x
8: (140, 1258, 15)
corner with TOP BARS BOTTOM BARS
column

(0,0,0~~1":;•=======---· X

(d) Specimen S4

Figure 6.32 (Continued) • Locations and numbering of strain gauges in different


specimens (a) specimen Sl; (b) specimen S2;
(c) specimen S3; and (d) specimen S4.
6-50

Table 6.6 -List of working strain gauges.

Gauge Specimen
No.
Sl S2 S3 S4

1 X ./ ./ ./

2 X ./ ./ X

3 ./ ./ ./ ./

4 ./ ./ ./ ./

5 X ./ ./ ./

6 ./ ./ ./ ./

7 ./ ./ ./ ./

8 ./ ./ ./ X

9 X X ./ --
CHAPTER 7- TEST RESULTS AND COMPARISON WITH CODE MODELS

7.1 Introduction

The test results of the slab-column connections are given in this chapter. The measured

failure load and the reactions developed at the base of the column are discussed in

Section 7.2 together with plots showing the development of the reactions at the

column support with the increase of load. The load versus displacements are given in

Section 7.3. Individual test results are discussed in Section 7.4 with observed crack

patterns and load versus steel strain graphs.

In the figures presented in this chapter, the total load is taken as the sum of the applied

load, the self weight load and the load due to the whiffle-tree arrangement. For slabs

Sl, S2 and S3 the weight of each of the whiffle-trees was 3.68 kN and corresponds to

pressure of 0.63 kPa (3 whiffle trees per slab). In slab S4 the weight of each whiffle-

tree was 7.35 kN which is equivalent to a pressure of 0.81 kPa (2 wiffle trees for the

slab). The self weight component of the slab deflection, reactions and steel strain was

determined from the measured density of the concrete and the initial linear-elastic

response of the slab under the applied load.

7.2 Failure Loads and Reactions

The edge column specimens rested on I-beams on three sides and the slabs were

capable of carrying loads higher than the punching shear failure load when considered

as a slab supported on three sides. For specimen S I the test was continued beyond the

punching shear failure. For specimens S2, S3 and S4 the tests were stopped once the
7-2

punching failure had occurred. The discussions in this section are limited to

observation up to the punching shear failure. However, for specimen S 1 the plots for

displacements and reactions are presented covering the whole of the test including

beyond the punching failure.

As discussed in Section 6.5, for specimen S1 the load cells placed to measure the

horizontal reaction at the column base recorded only a part of the total reaction. To

correct the horizontal reaction a linear finite element analysis (given in Appendix C)

was undertaken. The finite element model showed the ratio of the horizontal to vertical

reaction to be 0.5. In the graphs that follow both the recorded and corrected horizontal

reactions are presented.

The punching shear failure loads and the corresponding reactions in the vertical and

horizontal directions at the base of the column are presented in Table 7.1. The

horizontal reaction at the time of failure was lower than the maximum values, which

were developed a few load steps before the punching failure. The values of the

maximum horizontal reaction Rx-Max are tabulated in Table 7.2 along with the vertical

reaction Rz-Max and the total load at the corresponding load step. The reaction histories

for specimen S 1 are given in Figure 7.1 with the measured reaction in horizontal

direction designated as the R~ and the corrected reaction as Rx. The load versus

vertical reaction is also shown in Figure 7.1 with the vertical reaction shown as Rz.

In Figures 7.2 and 7.3 the load versus vertical and horizontal reactions up to the

punching shear failure are presented. The corrected horizontal reaction (see Figure 7.3)

of specimen S 1 is compared with the measured horizontal reactions for specimens S2


7-3

and S3 which were of similar configurations and concrete strengths. Figure 7.3 shows

Table 7.1- Punching shear failure loads and reactions.

Concrete Total Failure Load Vertical Horizontal


Cylinder (including self weight) Reaction Reaction
Specimen Strength (MPa) (kPa) Rz {kN) Rx {kN)

S1 24.1 18.5 96.5 43.3 *

S2 24.9 21.7 108.2 49.3

S3 26.3 25.5 135.6 53.9

*Note: Corrected values (refer Sections 6.5 and 7.2).

Table 7.2- Reactions and loads corresponding to the maximum horizontal


reaction measured Rx-Max•

Concrete Maximum Horizontal Corresponding Corresponding


.....~ Cylinder Reaction Measured Vertical Reaction Total Load
(.)
Q) Strength Rx-Max {kN) Rz-Max {kN) (kPa)
0..
tl) (MPa)

S1 24.1 43.1* 92.5 16.4


'
S2 24.9 49.6 105.8 21.1

S3 26.3 54.4 133.6 25.1


*Note: Corrected values (refer Section 6.5 and 7.2).
7-4

20
18
ca 16
~ 14
-12
i
.2 10
"i 8
0 6
1-
4
2
o~~~~~~~~~~~~~~~~~~~~~~

0 20 40 60 80 100 120
Reaction (kN)

Figure 7.1- Load versus vertical reactions (Rz), corrected horizontal reaction
(RJ and measured horizontal reaction ( R~) for specimen Sl.

30
S3

-
ca
25

-
~ 20
"D
ca 15
.2
"i 10
0
1-
5

0
0 50 100 150
Vertical reaction (kN)

Figure 7.2- Vertical reactions for specimens Sl, S2 and S3 up to the


punching failure.
7-5

30
S3
25
-ca

-
c.
~

'C
20
ca 15
.2
"i
0 10
.....
'Corrected reaction
5
0
0 10 20 30 40 50 60
Horizontal reaction (kN)

Figure 7.3- Load versus horizontal reactions for specimens 81,82 and 83 up to
the punching failure.

that the correction to the horizontal reaction of slab S 1 appears reasonable.

Tables 7.1 and 7.2 show that for slabs S1, S2 and S3 the failure load increases with an

increase in the prestress.

Specimen S4 failed in a punching mode when the total applied load (including self

weight and weight of the loading system) was 353.6 kN corresponding to a pressure of

39.0 kPa. The maximum total reactions measured in x- andy-horizontal directions for

the corner specimen S4 were 37.3 kN and 24.2 kN, respectively. The vertical reaction

at failure was 94.0 kN. However, corresponding to the maximum horizontal reaction

in x-direction of 37.3 kN, the horizontal reaction in y-direction, vertical reaction and

the total load on the slab were 23.9 kN, 86.3 kN and 33.4 kPa respectively. The load

versus reactions for slab S4 are shown in Figure 7 .4.


7-6

50
45

-
~
as
D.
40
35
30
"D
as 25
.2
a; 20
15 15
.....
10
5
0
0 20 40 60 80 100
Reaction (kN)

Figure 7.4 - Load versus vertical reaction (Rz) and the horizontal reactions in
the x- andy-directions (Rx and Rv) for specimen S4.

7.3 Deflections

For slabs S 1, S2 and S3 deflections were measured at the five locations shown in

Figure 6.31 a. For slab S4 the deflection was measured at three points on the slab,

shown in Figure 6.31 b. The load versus deflection for the complete test of the

specimen S1 is shown in Figure 7.5. For specimen S1 the test was carried on beyond

the punching failure. However, the discussions are limited only to the behaviour up to

the punching shear failure.

The load versus deflection curves up to punching shear failure for specimen S 1 are

shown in Figure 7.6. In Figures 7.7 and 7.8, the load versus deflection are plotted for

specimens S2 and S3, respectively. For slab S4 the load versus deflections are given
7-7

25 Flexural failure for slab supported on 3-sides


~ ~
Punching shear faHure

20

-ca

-8
0.
.lll::
"C
15 LVDTA

~ 10
1-

0 20 40 60 80 100
Deflection (mm)

Figure 7.5- Load versus deflections at A, Band C for specimen Sl


(refer Figure 6.31).
7-8

20 LVDTB

-
: . 15
LVDTA

-

' tJ
~ 10
'ii
....0 5

0+-~~~~,_~~~~~~~~~~~~-L-L~

0 10 20 30 40
Deflection (mm)

(a) Deflections at A, B and C

-16
20
18 -
:.

14
-12
i
- 8
10
'ii
0 6 -LVDTE
.... 4 -LVDTD
2
OT-~~~~~~,_~~~~~~~~~~~~~~

0 2 4 6 8 10 12
Deflection (mm)

(b) Deflections atE and D

Figure 7.6- Load versus deflections for specimen Sl up to punching failure.


7-9

25
LVDTC LVDTA

-
ca
D.
20

c"C 15
ca
.2
'i 10
0
1-
5

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

(a) Deflections at A, B and C

25

-20
ca
D.
,cca 15
0
'i 10
0
1- -LVDTD
5 -LVDTE

0
0 1 2 3 4 5 6
Deflection (mm)

(b) Deflections at D and E

Figure 7.7- Load versus deflections for specimen 82.


7-10

30

25

-ca 20
-
D..
.lll:
' tJ
ca
.2
15

"i 10
0 -LVDTA
1- -LVDTB
5

0
0 5 10 15 20 25
Deflection (mm)

(a) Deflections at A and B

30

25
-ca

-
D..
.lll:
' tJ
ca
.2
20

15
"i
0 10 -LVDTD
1-
-LVDTE
5

0
0 1 2 3 4 5
Deflection (mm)

(b) Deflections at D and E

Figure 7.8 -Load versus deflections for specimen 83.


7-11

in Figure 7.9. In Figures 7.5 to 7.9 the deflections are plotted against the total load,

which includes an allowance for the deflections due to the slab self weight and weight

of the whiffle-tree. The self weight was calculated using the measured concrete

densities and deflections were determined from the initial linear response of the slab

under the applied load.

In Figure 7.10 the load versus deflections at point A are plotted for slabs S 1, S2 and

S3. The deflections decreased with the increase of prestressing.

7.4 Individual Test Results

7.4.1 Specimen Sl

Specimen S 1 failed suddenly in a punching shear mode at an applied load of 270.6 kN

and corresponds to a pressure of 15.5 kPa (excluding self weight). However, since the

slab was resting on 1-beams on three edges it could carry further load. The specimen

was loaded further until an applied load of 338.4 kN was reached with the pressure of

19.4 kPa (excluding self weight).

The development of cracks was monitored during the test. Flexural cracks first

appeared at the top of the slab, in the vicinity of the column when the applied load was

183 kN (30.5 kN in the load cell). Cracking on the slab soffit was more difficult to

observe because of the whiffle-tree arrangement, however, no cracks could be seen up

to Yz of the failure load. The crack pattern for the top and the bottom surfaces of the

slab at the end of the test are shown in Figure 7 .11. Plate 7.1 shows the crack pattern

in the negative moment region at an applied load of 261 kN (43.5 kN on load cell).
7-12

50
45
LVDTB
_40
l. 35
c3o
"D
8 25
=as 20
'5 15
1- 10
5
o~~~~~~~~~~~~~~~~~~~~~

0 10 20 30 40 50 60
Deflection (mm)

Figure 7.9 -Load versus deflections for specimen S4.

30

-
as
25

-
~ 20
"D
as 15
.2
Sl

'ii 10
'5
1-
5
0
0 10 20 30 40
Deflection (mm)

Figure 7.10 - Load versus deflections at point A for specimens Sl, S2 and S3.
7-13

Punching shear failure


(See Plate 5.1)

Distributed flexural
cracks in the negative
moment region

(a) Top surface of the slab (b) Bottom surface of the slab

Figure 7.11- Crack patterns for specimen Sl after test.

Plate 7.1- Cracks in the vicinity of the column top for specimen Sl.
7-14

The punching shear failure surface is plotted in Figure 7.12. At the soffit of the slab

the failure started from the junction of the column and the slab. The measurements of

the punching cone as observed at the top and at the free edge of the slab are drawn in

Figure 7.12. The average angle of failure cone with the horizontal plane varied from

13 to 18 degrees.

The strain histories of different main tensile reinforcement at critical points (refer

Figure 6.32a) are given in Figures 7.13 and 7.14 for the working gauges. The strain

plots show some yielding of the central top bars at the punching failure. Figure 7.14

shows that yielding did not occur in the positive moment reinforcement and, thus,

sufficient reinforcement was provided to the slab to prevent a flexural failure before

the punching failure.

7.4.2 Specimen S2

Slab S2 failed suddenly (with a loud bang) when the applied load was 324.6 kN

(excluding self weight). The corresponding total pressure applied to slab was 18.6 kPa

(excluding self weight). The slab failed in punching shear near the column connection.

The failure surface started from the junction of column and the soffit of the slab.

Fewer flexural cracks were observed in specimen S2 than for specimen S 1. This is

attributed to the influence of prestress in controlling the crack development. For safety

reasons cracks were not marked during the test. The crack pattern, at failure on the top

and the bottom surfaces of the slab are presented in Figure 7.15 and shown in

Plate 7 .2. The failure surface was measured and is presented in Figure 7 .16. The

punching failure cone formed an average angle of between 13 and 24 degrees with the

horizontal plane.
7-15

Interface between the punching


X
surface and top slob surface:

line

I

I 0
y
(mm) (mm)
X

507
-Y X
(mm) (mm)
0 507
206 507 .300 507
t .350
415
415
427
.300
0
525
525
.300
0

"
0
N

0 0
0 0
I"") I"")

1.. 415 • 10· 525


..I Punching failure surface projected
{a} DETAIL - A TOP PLAN on the edge surface:
y z -Y z
(mm) (mm) (mm) (mm)
z 100 0 100 0
150 .32 125 20
175 40 150 28
225 67 175 49
250 77 200 51
275 85 225 54
.300 90 250 57
.375 91 275 62
400 94 300 62
415 100 325 70
350 74
400 77
425 75
475 80
500 82
(b) DETAIL - A FRONT VIEW 525 100

Figure 7.12- Cracks near the column; (a) crack detail near the column at slab
top and; (b) punching cone as seen at the free edge of the slab of
specimen Sl.
7-16

20
18
-16
:. 14 Yield limit
c"C 12
8 10
1i 8
0 6
1- -Gauge2
4 --Gauge4
2
0
0 1 2 3 4 5
Strain (E-3}

Figure 7.13- Load versus strain histories for x- andy-direction top bars of
specimen Sl.
7-17

-Gauge&
-Gauge7

0 0.2 0.4 0.6 0.8 1 1.2 1.4


Strain (E-3)

(a) X-direction reinforcement

20
18
- 16
:.
~
14
- 12
i0 10
- 8
1ii
0 6 -GaugeS
..... 4
-Gauge9
2
o~~~~~~-L-L-L~-+~~~~~~~L-L-~

0 0.5 1 1.5 2
Strain (E-3)

(b) Y-direction reinforcement

Figure 7.14- Load versus strain histories for bottom bars of specimen Sl
(a)X-direction reinforcement and; (b) ¥-direction
reinforcement.
7-18

Punching failure
surface
(See Plate 5.2)

(a) Top surface of the slab (b) Bottom surface of the slab

Figure 7.15 - Crack pattern at failure for specimen S2.

Plate 7.2 - Failure crack at the top surface of the slab for specimen S2.
7-19

ck line

Y'---L---J\Lm~/'7,.-~

- "
Detail A

Interface between the punching surface


and the top slob surface:

y X -Y X
(mm) (mm) (mm) (mm)
0 440 0 440
70 385 120 440
160 450 210 410
240 330 335 405
375 160 375 300
525 20 455 200
525 0 500 100
520 20
520 0

Punching failure surface projected


on the edge surface:
y X -Y X
(mm) (mm) (mm) (mm)
100 0 100 0
150 37 150 48
200 40 200 53
250 71 250 56
300 85 300 67
350 85 350 84
400 85 400 84
450 82 450 84
500 93 500 90
525 100 520 100

Figure 7.16- Cracks near the column; (a) crack detail near the column at slab
top and; (b) punching cone as seen at the free edge of the slab of
specimen S2.
7-20

The strain histories for the working strain gauges attached to the top and bottom

reinforcement (refer Figure 6.32b) are shown in Figures 7.17 and 7.18. None of the

bars had yielded at the point of failure indicating that the punching event occurred

before the flexural capacity of the slab was reached.

7.4.3 Specimen S3

Specimen S3 failed in punching shear with a sudden bang. The applied load at failure

(excluding self weight) was 391.2 kN with the corresponding total pressure on the slab

(excluding self weight) being 22.4 kPa.

The crack patterns developed at the top and the bottom surfaces of the slab are

presented in Figure 7.19. Because of the prestress effect these cracks could not be seen

easily. For reasons of safety no crack patterns were recorded until after failure of the

specimen and release of the applied load. Cracks developed at the top of the slab near

the column and are shown in Plate 7.3. The failure surface of the slab is shown in

Figure 7.20. At the soffit of the slab the failure surface started from the junction of

column and the slab, similar to that for specimens S 1 and S2. Due to the influence of

the prestress anchorage plates, the crack that formed on the free edge created a stepped

cone rather than a continuous cone. The average angle of the failure surface to the

horizontal was 26° and is steeper than that for specimens S 1 and S2.

The strain histories for the tension reinforcement (refer Figure 6.32c) are presented in

Figures 7.21 and 7.22 for the top and bottom reinforcement, respectively. None of the

steel had yielded before the punching failure.


7-21

25
Gauge3
Gaugel

-c
ca
0..
15
20

"D
ca
.2
1i 10
0
1-
5

0
0 0.5 1 1.5 2 2.5
Strain (E-3)

(a) X-direction reinforcement

25

i 20
~
-"D 15
8
1i 10
0 -Gauges
1- 5
-Gauge&

0 0.5 1 1.5 2
Strain (E-3)

(b) Y-direction reinforcement

Figure 7.17- Load versus strain histories for top bars of specimen 82.
7-22

25

l20
~
- 15
i
: 10
0 -Gauge7
1- 5 -GaugeS

0 0.5 1 1.5 2
Strain (E-3)

(a) X-direction reinforcement

25

-20
ca
D..
c"C 15
ca
.2
"i 10
0
1- -Gauge9
5

0
0 0.2 0.4 0.6 0.8 1
Strain (E-3)

(b) Y-direction reinforcement

Figure 7.18 - Load versus strain histories for bottom bars of specimen S2.
7-23

Punching failure
urface (See Plate 5.3)
.

(a) Top surface of the slab (b) Bottom surface of the slab

Figure 7.19- Crack patterns at failure for specimen S3.

Plate 7.3 - Failure surface at the top of specimen S3 in the region of the column.
7-24

..... - "
Detail A

crack line

I

Interface between the punching surface


and the top slab surface:

327 322 y X -Y X
(mm) (mm) (mm) (mm)
0 403 0 403
327 420 322 412
0 297 0 342 226
N
'<t 304 0

y
0
\ . 297 .. \ • 304 .. \

(a) DETAIL A - TOP PLAN

z
Punching failure surface projected on the
edge surface:
y z -Y z
(mm) (mm) (mm) (mm)
100 0 100 0
255 12 255 13
304 100 297 100
(b) DETAIL A - FRONT VIEW

Figure 7.20- Punching cone for specimen S3 (a) at the top surface above the
column; and (b) at the free edge of the slab.
7-25

30

-
ca
~ 20
25

,-ca
15
.2
'iii 10
0
1-
5
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Strain (E-3)

(a) X-direction reinforcement

30

-
ca
25

,-
~ 20
ca 15
.2

-
'iii 10
0
1-
5
-GaugeS
-Gauge&

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Strain (E-3)

(b) Y-direction reinforcement

Figure 7.21 -Load versus strain for the top reinforcement of specimen S3.
7-26

30

-as
25

-
c.
~

"'C
20
as 15
.2
'iii
0 10 -Gauge7
1-
5 -GaugeS

0
0 0.2 0.4 0.6 0.8 1 1.2
Strain (E-3)

(a) X-direction reinforcement

30

-
as
25

-
~
"'C
as 15
.2
20

'iii
0 10
1- -Gauge9
5 -Gauge10
0
0 0.2 0.4 0.6 0.8
Strain (E-3)

(b) Y-direction reinforcement

Figure 7.22 - Load versus strain for the bottom reinforcement in specimen S3.
7-27

7.4.4 Specimen S4

Specimen S4 failed at an applied load (excluding self weight) of 322 kN. The

corresponding total pressure on the slab (excluding self weight) was 35.4 kPa. The

slab failed in a brittle punching shear mode and was accompanied with a loud bang.

The punching shear surface is shown in Plate 7.4 and is mapped in Figure 7.23. The

failure surface started at the junction of the support column and the slab soffit. The

measurements of the punching cone along the top and along the free edge of the slab

(Figure 7.23) give average slopes of between 18 and 24 degrees.

The strain versus total applied load (including self weight) for the working strain

gauges (located as shown in Figure 6.32d) are presented in Figures 7.24 and 7.25 for

the top and bottom reinforcement, respectively. The strain histories

(Figures 7.24 and 7.25) show that the top reinforcement along the long edge had

yielded before failure. The negative moment reinforcement along the short edge and

the positive moment reinforcement had not yielded before failure indicating a reserve

of flexural capacity.

7.5 Discussion of Results

7.5.1 Introduction

The test results of specimens S 1, S2 and S3 show that increasing the prestressing force

gives an increase in the punching load carrying capacity of the connections. In the

same way the results of specimen S4 show that prestress has a significant influence on

the behaviour of comer column connections. In the following sections direct and

indirect comparisons are presented to show the influence of prestress on the punching

shear strength of slabs.


7-28

Plate 7.4 - Failure as viewed from the top of specimen S4.


7-29

lnclined crack line

I
IZ

----
Detail A
fi
>----
X

crack line

Interface between the punching


surface and top slab surface:
X y
(mm) (mm
520 0
500 20
0 500 440
0 450 440
Ill
--------, <1'
0
~ 342
264
370
420
2D0x200 J
0 500
Column 1
I
X
~1. 520
·I
(a) Detail A Top plan

z z
520

y
[!] 0 0
.[!] Ee

(b) Detail A - YZ Face (c) Detail A - XZ Face

Punching failure surface projected


on the edge surface:
X z y z
(mm) (mm (mm) (mm)
100 0 200 0
300 10 270 3D
430 30 300 60
480 85 450 90
520 100 500 100

Figure 7.23- Crack details near the column of specimen S4 (a) at the top surface;
(b) on the face of the YZ edge; and (c) on the face of the XZ edge.
7-30

45
40 Gaugel
-35
ca
,-ca 25
c.
~
30

..2 20
1ii
0 15
1- Yield limit
10
5
0
0 1 2 3 4 5 6
Strain (E-3)

Figure 7.24- Load versus strain histories for X- andY-direction top bars of
specimen S4.

45
40
35
Tot
al 30
loa
d 25
(kP 20
a)
15
10
5
0
0 0.5 1 1.5 2 2.5 3
Strain (E-3)

Figure 7.25- Load versus strain histories for X-direction bottom bars of
specimen S4.
7-31

7.5.2 General Discussion

The theoretical moments and shear forces obtained from the equivalent frame model

and the simplified method of AS 3600 (1994), for the model structure shown in

Figure A.2, are presented in Tables 7.3 and 7.4 together with the measured values. In

Table 7.5 the shear force to bending moment ratios are given for these models together

with the measured values. The bending moments and shear forces calculated in

Tables 7.3 and 7.4 are for a design load equivalent to the measured failure load. While

the specimens were sized to attract a similar moment to shear ratio to that of a multi-

bay flat slab systems, this clearly was not achieved. Table 7.4 shows that the measured

shear forces were generally 15 percent lower than the design values while Table 7.3

shows that the moment in columns was approximately half of the code values.

Table 7.1 and Figure 7.26 show that the load carrying capacity for specimens S2 and

S3 increased by 15 % and 35 % with the prestress of 0.8 MPa and 2.2 MPa,

respectively, compared to the non-prestressed slab Sl. Thus, the prestress has

significant effect in increasing the punching shear capacity of slab-column

connections.

The tests show that the number and the spread of cracks were better controlled with

increasing amounts of applied prestress. The damage of the top surface of the slab at

failure was observed to be over a smaller conical region with the increase of prestress.

That is, the average slope of the failure cone increased with increasing prestress. The

slope of the failure surface measured from the horizontal varied from 13 to 26 degrees

for specimens S 1 to S3. The increase in average slope can be attributed to the
7-32

30

S3
25

20

-"'
,-
a.
..llil:

15
"'
.S!
"ii
0
1-

10

5
Deflection at Point A
(refer Figure 6.31a)

0 10 20 30 40
Deflection (mm)

Figure 7.26- Load versus deflection at the mid panel for specimens Sl, 82 and
S3.
7-33

Table 7.3- Comparison between the theoretical and measured bending moments
at the column connections for different specimens.

Equivalent Simplified Method Measured


Failure Load Frame Method (As 3600-1994) BM
Specimen (kPa) BM BM at Failure
(kNm) (kNm) (kNm)

S1 18.5 44.1 41.6 26.1 *

S2 21.7 51.8 48.8 29.7

S3 25.5 60.9 57.4 32.5

S4 (X-Dir) 39.0 58.2 43.6 22.6

S4 (Y-Dir) 39.0 48.1 32.7 14.6

*Note: Corrected value (see Sections


. 6.5 and 7.2)

Table 7.4 - Comparison between the theoretical and measured load transferred to
the column for different specimens.

Failure Equivalent Frame Simplified Method Measured


Load Method (As 3600-1994) SF
Specimen (kPa) SF SF at Failure
(kN) (kN) (kN)

S1 18.5 101 111 96.5

S2 21.7 118 130 108

S3 25.5 139 153 136

S4 39.0 110 117 94.0


7-34

Table 7.5 - Comparison of shear (V) to moment (M) ratios.

£_ ~-t)
M

Specimen Equivalent Frame Simplified Method Measured


Method (AS 3600-1994)
S1 2.3 2.7 3.7 *
S2 2.3 2.7 3.7
S3 2.3 2.7 4.2
S4 (X-Dir) 1.9 2.7 4.2
S4 (Y-Dir) 2.3 3.6 6.4
*Note: Based on corrected bending moment (see Sections 6.5 and 7.2)

confinement effects of the prestress. Regan (1981 ), Hammill and Ghali ( 1994), Walker

and Regan ( 1987) and Alexander and Simmonds ( 1986) all reported failure cones of

lower than 45 degrees. For the comer column specimen S4 the angle of failure cone

was approximately 18 degrees. The slopes of the punched cones observed in different

tests in this and the previous studies indicate that the ACI 318-1999 and AS 3600-

1994 models underestimate the size of the critical perimeter. The load transferred to

the edge column increased with the increase of prestress although the size of the slab

forming the punched cone decreased.

Figure 7.27 compares the crack pattern formed in the non-prestressed comer

slab-column connection tests of Zaghlool and de Paiva (1973), Stamenkovic and

Chapman (1974) and Lim and Rangan (1994) with that of the column of comer-

column specimen S4. The failure cone of specimen S4 is similar to that of the edge

column connections and to that used in the CEB-FIP Model Code 1990 (1993) rather

than that of the previously reported non-prestressed connections.


7-35

··..
·····... ·······. . ~Crack line

··...

------------, '• ..

I ··..

I
~--~---------····~·----~----x

(a)Typical crack pattern reported in RC connections

...... / C r a c k lines
............ .
'•..
··... ·····

(b) Crack pattern in specimen S4

Figure 7.27- Comparison of the crack pattern developed at the top of the slab for
comer columns (a) reinforced concrete specimen reported by
Zaghlool and de Paiva (1973), Stamenkovic and Chapman (1974)
and Lim and Rangan (1994); and (b) prestressed specimen S4.
7-36

7.5.3 Comparison of Results with Code Models

Different design codes have different models for predicting the punching shear

strength of flat plate slab-column connections. For the purpose of the comparison the

strength is normalised against the shear strength for zero moment (Vuo) and the

moment for zero shear (Mu 0 ). The normalised equations for different code models are

presented in Appendix D. All strength reduction factors and partial safety factors for

the material are taken as unity. Where the cube strength lfcu) is used in the code model

the cube strength is taken as feu = 1.25 fern , where !em is the compressive cylinder

strength of the concrete.

The results of the tests are compared against various code models (see Appendix D) in

the following sections. Also included are the test data from other investigators with

1) all specimens included in the data pool with any mode of failure and; 2) only those

specimens included where the capacity of the slab was governed by punching shear

failure. That is where the load on the slab could not be increased for increasing

displacements due to limitations imposed by flexural considerations. In previous

studies, however, specimens failing in a punching mode before reaching their flexural

capacity are not always clearly reported. For the purpose of the following comparisons

only specimens clearly identified as failing in punching before the formation of a

flexural mechanism and specimens with average reinforcement content of greater than

one percent are included in the second data set. A total of 40 specimens including 7

prestressed concrete connection specimens are included for the comparison of the test

results of edge column connections. Details of the data are given in Tables 7.6 and 7. 7.

A total of 34 test results including one prestressed concrete specimen are


7-37

Table 7.6- Test data for reinforced concrete edge column connections.

Reference Specimen Average Mux Vu Mode of


p(%) (kNm) (kN) Failure#

Hanson and Hanson (1968) D15 1.67 10.5 12.0 punching

Zaghlool (1971) Z-IV(l) 1.23 45.0 122 punching

Z-V(1) 1.23 85.0 215 punching

Z-V(2) 1.65 94.0 247 punching

Z-V(3) 2.23 104 268 punching

Z-V(4) 1.23 81.0 0.00 punching

Z-V(5) 1.23 0.00 279 punching

Z-V(6) 1.23 88.0 117 punching

Z-VI(l) 1.23 107 265 punching

Stamenkovic and V/E/1 1.17 0.00 75.0 punching


Chapman (1974)
CJ./E/1 1.17 5.60 73.0 punching

CJ./E/2 1.17 9.20 54.7 punching

CJ./E/3 1.17 10.0 25.0 punching

CJ./E/4 1.17 8.80 11.0 punching

M.JE/2 1.17 8.30 0.00 punching

Regan et al. (1979) SE1 1.15 39.5 198 punching

SE2 0.249 34.0 192 not reported

SE4 1.15 30.5 152 punching

SE5 0.962 38.5 164 not reported

SE6 0.534 27.5 149 not reported

Note:# reported mode of failure. Mode of failure for some tests are not known from
the available literature and mentioned as "not reported" in these tables.
7-38

Table 7.6 (Continued) - Test data for reinforced concrete edge column
connections.

Reference Specimen Averagep Mux Vu Mode of


(%) Failure#
(kNm) (kN)

Regan et al. ( 1979) SE7 1.14 31.7 129 punching

Continued .... SE8 0.960 33.7 136 not reported

SE9 0.577 35.7 123 not reported

SE10 0.577 36.0 114 not reported

SEll 0.577 39.5 138 not reported

Rangan (1990) B 0.785 27.9 108 punching

c 0.408 9.35 21.9 punching

Mortin and Ghali (1991) JS1 0.775 60.5 141 punching

JS4 1.040 60.3 141 punching

Rangan and Lim (1992) Slab 1 0.450 19.9 108 punching

El-Salakawy et al. (1998) XXX 0.854 37.5 125 punching

HXXX 0.854 45.8 69.4 punching

This Study S1 1.13 29.2 96.5 punching

Note:# reported mode of failure. Mode of failure for some tests are not known from
the available literature and mentioned as "not reported" in the table.
7-39

Table 7.7- Test data for prestressed concrete edge column connections.

Reference Specimen Average Average Mux Vu Mode of


p(%) PIA Failure#
(MPa) (kNm) (kN)

Hawkins (1981) 2 0.780 3.09 50.8 139 flexure

Sunidja et al. S1 0.875 3.49 66.8 54.9 flexure


(1982)
S2 0.875 2.22 62.7 82.6 punching

S3 0.875 2.18 50.3 66.2 flexure

S4 0.875 0.741 53.4 117 punching

This Study S2 0.774 2.21 29.7 108 punching

S3 0.774 1.10 32.5 136 punching

# reported mode of failure


7-40

used for the comparison of test results of corner column connection specimens. The

details of these corner column connections are given in Tables 7.8 and 7.9. Only test

results of flat plate edge and corner column specimens without any shear

reinforcement (including ties in the spandrel) and without holes in the slab adjacent to

the column are considered.

7.5.3.1 Comparison withACI 318-1999

The normaslised moment-shear load path for specimens S1, S2 and S3 of the present

tests are compared with the ACI 318 (1999) model in Figure 7 .28. The figure shows

that the edge column specimens have a similar load path. In Figure 7.29 the

normalised load path for the corner column connection, specimen S4 is presented.

Compared to the edge column connections the slope of the interaction line for the

corner column connection is seen to be flatter and this is attributed to the effect of the

biaxial moment.

The test data given in Tables 7.6 to 7.9 for the edge and corner column connections are

plotted against the ACI 318 (1999) model, in a normalised form, in

Figures 7.30 to 7.33. Figure 7.30 shows the test data for the edge column connections

for the full data set. Figure 7.31 compares the data for the limited data set. Both

Figures 7.30 and 7.31 show a similar scatter in the data compared to the design model.

Similarly, Figures 7.32 and 7.33 show the scatter of test data for corner column

specimens for the full and limited data sets, respectively. Again the scatter in the data

is similar for the two sets. The scatter plots show that a large variation exists between

the ACI model and the observed data.


7-41

Table 7.8 - Test data for reinforced concrete corner column connections.

Reference Specimen Average Mux Muy Vu Mode of


p(%) Failure#
(kNm) (kNm) (kN)

Zaghlool and de I 1.47 8.76 6.67 105 punching


Paiva (1970) *
m 1.47 16.1 16.0 99.0 punching

IV 1.47 15.8 22.0 118 punching

Zaghlool (1971) Z-II(5) 1.23 0.00 0.00 149 punching

Z-ll(5)* 1.23 0.00 0.00 138 punching

Z-II(1) 1.23 38.5 38.5 138 punching

Z-II(6) 1.23 38.9 38.9 82.3 punching

Z-II(4) 1.23 28.3 28.3 0.00 punching

Z-II(1) 1.23 38.5 38.5 138 punching

Z-II(2) 1.23 53.4 53.4 177 punching

Z-II(3) 1.23 58.0 58.0 178 punching

Z-1(1) 1.23 19.2 19.1 74.3 punching

Z-II(1) 1.23 38.5 38.5 138 punching

Z-ID(l) 1.23 52.7 52.7 180 punching

Stamenkovic and V/C/1 1.17 0.00 0.00 27.2 punching


Chapmen (1974)
C/C/1 1.17 6.24 0.00 24.9 punching

C/C/2 1.17 6.38 0.00 15.9 punching

C/C/3 1.17 6.18 0.00 7.96 punching

C/C/4 1.17 5.63 0.00 3.59 punching

M/C/1 1.17 4.56 0.00 0.00 punching

# reported mode of failure


* Out of the data available for 4 symmetrical columns, data for the critical column is
considered here.
7-42

Table 7.8 (Continued) - Test data for reinforced concrete comer column
connections.

Reference Specime Average Mux Muy Vu Mode of


n p{%) Failure#
(kNm) (kNm) (kN)

Walker and SCI 0.645 25.2 25.2 81.5 ductile punching


Regan (1987)
SC2 0.415 24.0 24.0 74.8 punching

SC3 0.834 31.6 31.6 74.2 ductile punching

SC4 0.645 16.7 16.7 63.8 ductile punching

SC5 0.930 18.8 18.8 82.2 punching

SC6 0.760* 25.4 25.4 79.0 punching

SC7 0.930 27.6 27.6 82.2 punching

SC8 0.275 4.70 4.70 33.0 flexure punching

SC9 0.563 5.90 5.90 33.0 flexure punching

SC11 1.42 4.60 2.20 33.0 punching

SC12 0.869 12.7 8.90 36.8 punching

Hammill and NHl 1.47 43.0 43.0 147 punching


Ghali (1994)
NH2 1.47 40.2 40.2 139 punching

NH4 1.47 33.0 33.0 0.00 punching

# reported mode of failure


*: Reinforcement in diagonal direction

Table 7.9 - Test data for prestressed concrete comer column connections.

Reference Specimen Average Average Mux Muy Vu Mode of


p (%) PIA (kNm) (kNm) (kN) Failure
(MPa)

This Test S4 0.910 2.21 22.2 14.0 87.8 punching


7-43

1.5

1
S3
ACI 318-1999

0.5

0 0.5 1 1.5
MufMuo

Figure 7.28 - Comparison of load path for edge column connection specimens Sl,
S2 and S3 with ACI 318-1999.

1.5

1
ACI 318-1999

0 0.5 1 1.5 2
MJMuo

Figure 7.29 - Comparison of load path for comer column connection specimens
S4 with ACI 318-1999.
7-44

- ACI318 • 1999 • Speelman S1


+ Speelman S2 A Specimen S3
+ Rangan et aL (1992) 0 Sunldja et aL (1982)
2 <> Hawkins (1981) ll Regan et aL (1979)
o Stamenkovlcet aL (1974) X Zaghlool (1971)
)I( Hanaon et al. (1968) - Rangan (1990)
- Mortin et aL (1991) e EI-Salakawy et aL (1998)

1.5

0
:I
0
2;,1 A
> 0
X >C
0.5 •
- ~0 *
X
0

0 0.5 1 1.5 2
MJMuo

Figure 7.30- Comparison of test results with ACI 318-1999 model for edge
column connections (all data).

-ACI318·1999 • Speelman S1
+ Speelman S2 A Specimen S3
e EI-Salakawy et al. (1998) + Rangan et aL (1992)
2 - Mortin et aL (1991) - Rangan (1990)
0 Sunldja et al. (1982) ll Regan et aL (1979)
0 Stamenkovic et aL (1974) X Zaghlool (1971)
)I( Hanaon et aL (1968)

1.5

g
0
2;,1 A
> 0
llll X
~
>C
0.5 •
- _o X
0
Ql:

)I( 0

0 0.5 1 1.5 2
MJMuo

Figure 7.31- Comparison of test results with ACI 318-1999 model for edge
column connections (limited data set).
7-45

-ACI318 -1999 • Specimen S4


0 Hanvnlll et al. (1994) ¢ Walker et 81.(1987)
O Stamenkovlc et al. (1974) X laghlool (1971)
2 )I( laghlool et al. (1970)

1.5

0 1 2 3
MJMuo

Figure 7.32- Comparison of test results with ACI 318-1999 model for corner
column connections (all data).

-ACI318- 1999 • Specimen S4


0 Hanvnlll et al. (1994) ¢ Walker et 81.(1987)
o Stamenkovic et al. (1974) X laghlool (1971)
2 )I( laghlool et al. (1970)

1.5
0

~ 1

0
0.5 X
~ ~ o"'r:JI x
¢ d«>x ¢

0 1 2 3

Figure 7.33- Comparison of test results with ACI 318-1999 model for corner
column connections (limited data set).
7-46

7.5.3.2 Comparison with AS 3600-1994

The normalised load path for specimens S 1 to S4 are compared with the AS 3600

model in Figures 7.34 and 7.35. The normalised load paths for the edge columns are

similar to those of the ACI 318 model (see Figure 7.28). However, the normalised path

for comer column connection specimen S4 (see Figure 7 .35) is steeper than that

obtained using ACI 318 model.

The test data are plotted against the AS 3600 model in Figures 7.36 to 7.39.

Figure 7.36 shows the distribution of test data for edge column connections for all

data. Figure 7.37 shows the selected data set for edge connections. Similarly, the test

data for the comer column connections are plotted for all data against the code model

in Figure 7.38 and the selected data for the comer connection specimens failing in

punching shear are shown in Figure 7.39. There is a significant difference between the

code model and the experimental data for both data sets for both the edge and comer

column connections. Figures 7.36 to 7.39 show that the code model errs on the

conservative for the edge connections and generally for the comer connections,

although some data falls on the non-conservative side of the interaction curve.

7.5.3.3 Comparison with BS 8110: Part 1: 1997

Unlike the American and Australian code models, in BS 8110 the reinforcement

content is a factor in determining the strength of the connection. The shear strength is

zero for slabs without flexural reinforcement.

The British code BS 8110 (1997) refers the designer to specialist literature for the

design of prestressed slab-column connections. In Figures 7.40 and 7.42 the test
7-47

g
.~
>

0 0.5 1 1.5

Figure 7.34- Comparison of load path for edge column connection specimens S1,
S2 and S3 with AS 3600-1994.

1.5

1
g
.; S4

0.5

o~~~~~+-~~~~~~~~~

0 0.5 1 1.5
M,*/Muo

Figure 7.35 - Load path for corner column connection specimens S4 with
AS 3600-1994.
7-48

-AS 3600 ·1994 • Specimen S1


+ Specimen S2 A Specimen S3
+ Rangan et al. (1992) 0 SunldJa et aL (1982)
0 Hawkins (1991) 6 Regan et aL (1979)
o Stamenkovic et aL (1974) X Zaghiool (1971)
2 ::K Hanson et al. (1968) - Rangan (1990)
- Mortin et aL (1991) e EI-Salakawy et aL (1998)

1.5 0

0 0
:I

.~
>

0.5
1
6
6
+
~--
6
6

6
6
6
X • •
0 X
l •
0

0

0
::KcP

0
0 0.5 1 1.5 2
Mv"!Muo

Figure 7.36 - Comparison of test results with AS 3600-1994 model for edge
column connections (all data).

-AS 3600·1994 • Specimen S1


+ Specimen S2 A Specimen S3
e EI-Salakawy et al. (1998) + Rangan et al. (1982)
- Mortln et al. (1991) - Rangan (1990)
D SunidJa et aL (1982) 6 Regan et aL (1979)
2 0 Stamenkovic et aL (1974) X Zaghlool (1971)
::K Hanson et aL (1968)

1.5 0

0
g
.~
~~·
1
>

0.5 ox 0

0
::Ko

0 0.5 1 1.5 2
Mv.,Muo

Figure 7.37 - Comparison of test results with AS 3600-1994 model for edge
column connections Oimited data set).
7-49

2
-AS 3600-1994
o Hammill et al. (1994)

¢
SpecimenS4
Walker et al. (1985)
o Stamenkovic et al. (1974) X laghlool (1971)
::K laghlool et al. (1970)

1.5
::1( X
0 ::1( X

..~
> 1
>tr.)Q 0
¢ 0
¢ ¢
~ ¢
0.5 ¢¢0 Xo
¢ ¢

0 0.5 1.5 2

Figure 7.38- Comparison of test results with AS 3600-1994 model for corner
column connections (all data).

-AS 3600-1994 • Specimen S4


2 0 HammUI et al. (1994) <>Walker et al. (1985)
0 Stamenkovic et al. (1974) X 2'aghlool (1971)
)K laghlool et 81.(1970)

1.5
)K X
0 )K X
:I

.c:: 1
~XJ
> 0
<> 0
X<> <>
<>
X
0.5 <>

0 0.5 1.5 2

Figure 7.39- Comparison of test results with AS 3600-1994 model for corner
column connections (limited data set).
7-50

data for the edge and corner column connections are plotted. The test data for the

prestressed connections in this study are compared to the code model together with the

data from other test series. Although BS 8110 (1997) does not strictly apply to the

design of prestressed specimens the prediction of strength for the test specimen is

consistent with the general data scatter. Figures 7.40 to 7.43 show that the British code

predicts well the mean of test data but considerable scatter exists.

For columns subjected to flexure but without axial force, equation D.l3

(see Appendix D) suggests that shear transfer via the slab is not a design

consideration. Thus there is poor correlation with experiments where no axial force

exists in the column. For example in specimen M.JE/2 by Stamenkovic and Chapman

(1974) (refer Figure 2.12) the slab-column connection is under the action of moment

without any axial force in the columns. By the code the capacity of this specimen is

not limited by shear capacity of the section. Other examples include specimen Z-V(4)

of Zaghlool (1971) and the corner column connection specimens, slab M/C/1 of

Stamenkovic and Chapman (1974), Z-ll(4) of Zaghlool (1971) and specimen NH4 of

Hammill and Ghali (1994).

7.5.3.4 Comparison with CEB-FIP Model Code 1990 (1993)

The relative load path for the edge column specimens S 1, S2 and S3 and corner

column specimen S4 are plotted in Figures 7.44 and 7.45, respectively. The relative

load paths are steeper than those for the ACI 318 and the AS 3600 models.

The non-dimensionalised ratios of shear and moment for different test series are

plotted against the CEB-FIP model in Figures 7.46 to 7.49 for edge and corner column
7-51

-BS8110·1997 • Speelman S1
+
Speelman S2 A Specimen S3
e EI-Salakakwy et aL (1998) + Rangan at aL (1992)
- Martinet aL (1981) - Rangan (1990)
400 D Sunldja at aL (1982) <> Hawklna (1981)
A Regan at al. (1979) o Stamankavic at al. (1974)
X 2'aghlool (1971) :1: Hanaon at aL (1968)
350

300 X
z X X
c
.c
250 X X X
'&
e200
..
u;
: 150
.c
0
100
:1:
50

0
0 50 100 150 200 250 300 350 400
v.,(kN)

Figure 7.40- Comparison of results for edge columns with BS 8110-1997


(all data).
- BS 8110 • 1997 • Specimen S1
+ Specimen S2 A Specimen S3
• EI-Salakawy et al. (1998) + Rangan at aL (1992)
- Martin at aL (1991) - Rangan (1990)
400 D Sunldjaet aL (1982) A Ragan at aL (1979)
o Stamankavic at aL (1974) X 2'aghlool (1971)
:1: Hanaon et aL (1968)
350

300
z
c.c 250
a,
c
2! 200
.
u;
ftl
Gl 150
.c
0
100

50

0
0 50 100 150 200 250 300 350 400
V811 (kN)

Figure 7.41 -Comparison of results for edge columns with BS 8110-1997


(limited data set).
7-52

300 -898110-1997
• Specimen S4
0 Hanvnill et al. (1994)
250 <> Walker et al. (1987)

-
z
~200
0 Stamenkovic et al. (1974)
X Zaghlool (1971)
)I( Zaghlool et al. (1970)

-e
.s::.
0)

150 X
X

-;;
...
: 100
.s::.
CJ)

50

0
0 50 100 150 200 250 300
V8 ,,(kN)

Figure 7.42 ·Comparison of results for corner columns with BS 8110-1997


(all data).

300 -BS 8110-1997


• Specimen S4
0 Hanvnill et al. (1994)
250 <> Walker et al. (1987)

-
z
~200
0 Stamenkovic et al. (1974)
X Zaghlool (1971)
)I( Zaghlool et al. (1970)
.s::.
a, X

-e...
0

: 100
150 X

.s::.
CJ)

50

0
0 50 100 150 200 250 300
Vett (kN)

Figure 7.43 ·Comparison of results for corner columns with BS 8110-1997


(limited data set).
7-53

S3
1.5

g
~ 1

0.5

0 0.5 1 1.5 2

MJMuo

Figure 7.44- Comparison of load path for edge column connection specimens S1,
S2 and S3 with CEB-FIP Model Code 1990 (1993).

1.5

~ 1

0.5

0 0.5 1 1.5 2
MJMuo

Figure 7.45 - Comparison of load path for comer column connection specimens
S4 with CEB-FIP Model code 1990 (1993).
7-54

-CEB-FIP 1990 • Speelman S1


2 + Speelman S2 A Specimen S3
e EJ.Salakawy et aL (1998) + Rangan et aL (1992)
- Mortln at. AL (1991) - Rangan (1990)
0 Sunldja et al. (1982) <> Hawkins (1991)
I!. Regan et aL (1979) 0 Stamenkovic et al. (1974)
X Z'aghlool (1971) lK Hanson et al. (1968)

1.5
[!,.

0
.(l
[!,.
X •
~
0

1 aa_, •o
~X'
0

0.5

0 0.5 1 1.5
MJMuo

Figure 7.46- Comparison of results with CEB-FIP model Code 1990 for edge
column (all data).

-CEB-RP 1990 • Specimen S1


2 + Specimen S2 A Specimen S3
• SJ.Salakawy et al. (1998) + Rangan et al. (1992)
- Mortln et al. (1991) - Rangan (1990)
0 Sunldja et aL (1982) I!. Regan et al. (1979)
0 Stamankovlc et aL (1974) X Zaghlool (1971)
lK Hanson et aL (1968)

1.5 •
0

+
[!,.
X •
s
0
:I [!,.
1 -~ .0
[!,.

0.5 x•

0 0.5 1 1.5
MJMuo

Figure 7.47- Comparison of test results with CEB-FIP Model Code 1990 for edge
column (limited data set).
7-55

1.8
X
1.6

1.4
0
X "'
0
~ Xo
1.2
0
0
:I 1 oo8 0

~ 0 0
0 0 0
>0.8 X
X
-CEB-FIP 1990
0.6 • Specimen S4
0 Hammill et al. (1994)
0.4 0 Walker et al. (1987)
0 Stamenkovic et al. (1974)
X Zaghlool (1971)
0.2
0 lK Zaghlool et al. (1970)

0
0 0.5 1 1.5
MJMuo

Figure 7.48- Comparison of test results for comer column with CEB-FIP Model
Code 1990 (all data).

1.8
X
1.6

1.4 X "'
0
~ Xo
1.2
0
0 0 0 0
:I 1
~
>0.8
- CEB-FIP 1990
0.6 • Specimen S4
0 Hammill et al. (1994)
0.4 0 Walker et al. (1987)
0 Stamenkovlc et al. (1974)
0.2 X Zaghlool (1971)
lK Zaghlool et al. (1970)

0
0 0.5 1 1.5
MJMuo

Figure 7.49- Comparison of test results for corner column with CEB-FIP Model
Code 1990 (limited data set).
7-56

specimens. A wide scatter in the data is seen for both the edge and the comer column

specimens both for the full and limited data sets. For a number of test specimens from

different test series the model errs on the non-conservative and the general trend is

poor.

7.5.3.5 Comparison with Eurocode 2 (1992)

Figures 7.50 and 7.51 compare the Eurocode 2 model for edge column specimens with

the full and selected data sets, respectively. Figures 7.52 and 7.53 compare the

Eurocode 2 model for comer column specimens with the full and selected data sets,

respectively.

As for the British Standard BS 8110, Eurocode 2 does not make any allowance for

specimens where flexure is passed from the column to the slab in the absence of axial

force (see Figure 2.12). Therefore, for specimens without an axial force component the

comparison between the test data and the Eurocode 2 model is poor. With this

exception the model mostly errs on the conservative.

7.5.4 Further analysis

In all ~ases the comparison of the experimental data with the design codes show a

significant amount of scatter. In Tables 7.10 and 7.11 the means and the standard

deviations are given for the test data compared with each of the code models for the

edge and comer columns, respectively. In Table 7.12 the means and standard

deviations are given for all data for both the comer and edge connections.
7-57

- Eurocode 2 (1992) • Specimen S1


+ Specimen S2 • Specimen S3
e EI-Salakawy at al. (1998) + Rangan at al. (1992)
- Mortln at al. (1991) - Rangan (1990)
0 Sunldja at al. (1982) <> Hawkins (1981)
6. Regan et al. (1979) 0 Stamenkovlc et al. (1974)
X Zaghlool (1971) * Hanson at al. (1968)

400
350
300
z25o
.Ill:
"';:'
'Q
200 xx
X
..: 150
100
50
o~~~~~~~~~~~~

0 50 100 150 200 250 300 350 400

v.d (kN)

Figure 7.50 - Comparison of test results for edge column connection with
Eurocode 2 (all data).

-Eurocoda 2 (1992) • Specimen S1


+ Specimen S2 • Specimen S3
e EI-Salakawy at al. (1998) + Rangan et al. (1992)
- Mortln at aL (1991) - Rangan (1990)
0 Sunklja at al. (1982) 6. Regan at al. (1979)
0 Stamenkovlc at al. (1974) X Zaghlool (1971)
::t: Hanson et al. (1968)

400
350
300
z25o
.Ill:
"';:'
'Q
200 xx
X X
..: 150 &<
100
50
o~~~~~~~~~~~~

0 50 100 150 200 250 300 350 400

Vad (kN)

Figure 7.51 - Comparison of test results for edge column connection with
Eurocode 2 Oimited data set).
7-58

- Eurocode 2 (1992)
400 • Specimen S4
0 Hammill et al. (1994)
350 <> Walker et al. (1987)
o Stamenkovlcet al. (1974)
300 X zaghlool (1971)
:::K zaghlool (1970)
-250

-;!
~
.... 200
~~
DO
:::K X
150 X
• )I(
100

50
o~~~~~~~~~~~~~~~LWY

0 50 100 150 200 250 300 350 400


v.d (kN)

Figure 7.52 - Comparison of test results for corner column connection with
Eurocode 2 (all data).

-Eurocode 2 (1992)
400 • Specimen S4
0 Hammill et al. (1994)
350 <> Walker et al. (1987)
0 Stamenkovlc et al. (1974)

300 X Z&ghlool (1971)


:::K Z&ghlool et al. (1970)

--
~
250

200 Do
J 150
3E~ :::K X
X
.)I(

100

50

0
0 50 100 150 200 250 300 350 400
vsd (kN)

Figure 7.53 - Comparison of test results for corner column connection with
Eurocode 2 (limited data set).
7-59

Table 7.10- Summary of analyses for the experimental data compared to design
codes for edge column connections.

Data Set I DataSet IT

Code Mean so Mean so


ACI318 1.41 0.34 1.49 0.35

AS 3600 1.63 0.38 1.71 0.39

BS 8110 1.03 0.39 1.00 0.37

CEB-FIP 1990 1.39 0.37 1.38 0.41

Eurocode 2 1.56 0.62 1.61 0.67


Note: Set 1: All data included.
Set 2: Limited to specimens reported to have failed in punching shear before
yielding of flexural reinforcement and specimens with greater than 1.0 percent
of reinforcing steel

Table 7.11- Summary of analyses for the experimental data compared to design
codes for corner column connection.

DataSet I Data Set IT

Code Mean SD Mean so


ACI318 1.55 0.51 1.57 0.55

AS 3600 1.49 0.46 1.53 0.48

BS 8110 1.06 0.34 1.06 0.36

CEB-FIP 1990 1.72 0.63 1.70 0.67

Eurocode 2 1.36 0.44 1.39 0.47


Note: Set 1: All data included.
Set 2: Limited to specimens reported to have failed in punching shear before
yielding of flexural reinforcement and specimens with greater than 1.0 percent
of reinforcing steel
7-60

Table 7.12 - Summary of analyses for the experimental data compared to design -
codes for all data of edge and corner column connections.

DataSet I DataSet II

Code Mean SD Mean SD

ACI318 1.45 0.40 1.51 0.42

AS 3600 1.59 0.45 1.64 0.48

BS 8110 1.04 0.37 1.03 0.37

CEB-FIP 1990 1.54 0.53 1.54 0.58

Eurocode 2 1.47 0.55 1.50 0.59

For the British Standard BS 8110, the mean measured shear strength to the capacity

calculated from the code (Ve.u) is 1.03 and 1.06 for the edge and comer columns,

respectively. When all the data for the edge and comer column connections are

considered it is 1.04. While the best fit is with BS 8110, the standard deviation is high

at 0.37. This implies that a significant amount of data lies on the non-conservative side

of the design model. The high standard deviation is of concern and requires a high

partial safety factor if the overall safety of the structure is to be maintained. For the

test data used in this study, BS 8110 gives a significantly improved prediction than

either the ACI 318 or AS 3600 models.

The Australian code of practice, AS 3600, has one of the largest mean experimental to.

theory ratios at 1.63 and 1.49 for edge and comer column connections, respectively.
7-61

Figures 7.36 and 7.38 show that no points fall below the theoretical strength for edge

connections and few data points fall below the theoretical strength for the comer

column connections. The mean and standard deviation of the combined edge and

comer connections for all data are 1.59 and 0.45, respectively. Assuming a normal

distribution of data there is a 91 percent level of confidence that the design model

would predict a conservative result.

The scatter of data for the ACI model is similar to that for the AS 3600 model but

inspection of Figures 7.30 and 7.36 show that more data points fall below the

theoretical limit than for the Australian code model. The mean of all data for the ACI

model is 1.45 with a standard deviation of 0.40. The confidence level of any one point

falling on the theoretically safe side of the interaction line (again assuming a normal

distribution) is 87 percent.

The Eurocode 2 model is of a similar form to the British model but with considerably

fewer data points falling on the non-conservative side. The mean experimental to

theoretical ratio for the combined data of edge and comer connections is 1.47; higher

than both the British and American codes but lower than the Australian code. The

corresponding standard deviation for the Eurocode 2 model of 0.55 gives a confidence

level of 80 percent. Comparing Tables 7.10 and 7.11 shows that there is a significant

variation in the mean and standard deviation between the edge and comer data sets.

The CEB-FIP (1993) model has one of the lower means for edge column specimens

(Table 7.10) but has a high standard of deviation of 0.37. The mean and standard

deviation for comer column connections are 1.72 and 0.63, respectively, and are the
7-62

highest of all models. An inspection of Figures 7.46 and 7.48 also shows that a number

of data points fall below the theoretically safe line.

The results of the statistical analysis carried out for the connections reinforced with

conventional reinforcement and with prestressed concrete are shown in

Tables 7.13 and 7.14, respectively. Except for the British code the means of the

experimental data to code ratio increased with a greater scatter when prestress is added

to the slab. It is also seen (Tables 7.10 to 7.15) that the means are generally higher

when the specimens not identified as punching failure are excluded from the data pool.

The generally poor predictions, however, of all code models together with the high

degree of scatter in the data suggests that the mechanics of punching shear are not well

represented by the code models. Further research is required into the load transfer

mechanism between the supporting columns and flat plate slabs.

The analysis of results discussed above do not consider the effect of material partial

safety factors or strength reduction factors in determining the probability of failure.

Table 7.12 is reproduced as Table 7.15 with the effects of the strength reduction

(or partial safety) factors included in the calculations. It is seen that BS 8110 yields the

most economical design while the CEB-FIP (1993) model the most conservative. The

confidence levels for the data set IT with and without the effect of the safety factors are

compared in Table 7 .16.

The calculations on the probability of a single data point falling below the theoretical

strength for data set IT with the material factors excluded is as high as 4 7 percent for

the British code (BS 8110) and down to 9 percent for the Australian code (AS 3600).
7-63

Table 7.13 - Summary of analyses for the experimental data compared to design
codes for edge column connections considering only non-prestressed
concrete connections.

DataSet I Data Set II

Code Mean SD Mean SD

ACI318 1.39 0.31 1.45 0.32

AS 3600 1.63 0.38 1.69 0.41

BS 8110 1.0;4 0.39 0.97 0.37

CEB-FIP 1990 1.33 0.34 1.29 0.36

Eurocode2 1.56 0.61 1.55 0.67

Table 7.14- Summary of analyses for the experimental data compared to design
codes for edge column connections considering only prestressed
concrete connections.

Data Set I Data Set II

Code Mean SD Mean SD

ACI318 1.53 0.44 1.80 0.38

AS 3600 1.60 0.36 1.82 0.29

BS 8110 0.95 0.39 1.19 0.34

CEB-FIP 1990 1.64 0.40 1.91 0.32

Eurocode 2 1.57 0.65 1.97 0.57


7-64

Table 7.15 - Summary of analyses for the experimental data compared to design
codes for all data of edge and corner column connections
considering the effect of t/1 and Ym as appropriate.

Data Set I DataSet IT

Code Mean SD Mean SD

ACI 318 (q, = 0.85) 1.71 0.47 1.78 0.49

AS 3600 (q, = 0.7) 2.27 0.67 2.34 0.69

BS 8110 (rm = 1.25) 1.30 0.46 1.29 0.46

CEB-FIP 1990 (YM = 1.5) 2.31 0.80 2.31 0.87

Eurocode 2 (rc = 1.5) 2.20 0.83 2.25 0.89

Note: ifJ strength reduction factor and Ym. ~. and Yc are partial safety factors for the
materials depending on the code.

Table 7.16- Confidence level for data Set II with q, or Ym set equal to unity and
with the values set by corresponding code.

Code With t/1 or y equal to With t/1 or r set to


1.0 appropriate value

ACI318 (ifJ=0.85) 88.7 94.2

AS 3600 (q, = 0. 7) 90.8 97.5

BS 8110 (rm = 1.25) 53.2 73.2

CEB-FIP 1990 (rM = 1.5) 82.4 93.5

Eurocode 2 (rc = 1.5) 80.2 92.1

Note: ifJ strength reduction factor and Ym. ~. and Yc are partial safetys for the materials
depending on the code.
7-65

When material strength reduction factors (or partial safety factors) are included in the

calculations only the Australian code model gives a less than five percent probability

of failure. The ACI, Eurocode and CEB-FIP models give probabilities of a failure

between six and eight percent. The probability that an edge or comer column

connection designed to the British code (BS 8110) is high at 27 percent.

7.6 Conclusions

All specimens tested in this study failed in a punching shear mode and the failure

surface started from the intersection of the column and slab soffit. The slopes of the

failure surface with reference to the horizontal plane were between 17 and 26 degrees.

The primary conclusion from the tests is that prestressing has an influence on the

strength of slab-column connections. The addition of prestress increases the load

carrying capacity of the connections and relatively larger unbalanced moments can be

transferred to columns. The formation of cracks and the extent of damage are also

controlled by the addition of prestress into the concrete slab. The failure mode,

-however, remains brittle.

A comparative study of different test results for non-prestressed and prestressed

slab-column specimens, using various code models, shows a large scatter in the data

and generally poor predictions. However, for the most part the codes give conservative

predictions with the exception of BS 8110 which shows a good prediction of the mean

of the test data although, as for the other models, the scatter is high.

Calculation on the probability that an edge or comer column connection designed to

different code models fails before the code requirement show that only the Australian
7-66

code gives a lower than five percent probability of failure, although the ACI,

Eurocode 2 and the CEB-FIP model code are not significantly greater than five

percent. The British code, however, gives a high probability of failure.


CHAPTER 8 - FINITE ELEMENT ANALYSES OF TEST SPECIMENS

8.1 Introduction

In Chapter 3 of this thesis different material models, mesh patterns and aspect ratios

were investigated to obtain optimum results for the analysis of the punching type

problems. The model and the finite element program DIANA were tested and verified

in Chapter 4 using a number of experimental studies. These studies represented a

range of problems such as simple flexure to the more complex punching shear in flat

plate-column connections including prestressed concrete connections. In Chapters 6

and 7, the testing of four flat plate-column connections was discussed. These

specimens were developed from a one-third-scale model of the prototype flat plate

structure. Three of the specimens represented connections with edge columns while

the fourth specimen represented a comer column connection. Except for specimen S 1

the specimens were prestressed in two directions with the amount of prestress being

varied. A number of evenly distributed point loads were applied to the top surface of

the slabs simulating a uniformly distributed load. All the specimens failed in a

punching shear mode.

In this chapter, the experimental slabs reported in Chapter 6 are modelled using the

finite element program DIANA as discussed in Chapter 3. The concrete was modelled

using 20-node 3D isoparametric brick elements with the three degrees of freedom at

each node. The steel was modelled using embedded bar elements.

In addition to the numerical modelling of the test specimens, further analyses of the

slab-column connections are undertaken to study parameters not investigated directly

as a part of the experimental testing program such as bonding of the prestressing steel.
8-2

8.2 Finite Element Modelling

The finite element meshes developed for slab-column connections S 1, S2 and S3 are

shown in Figures 8.1 and 8.2. One half of the specimens were modelled using

symmetry. The finite element models consisted of a total of 2137 nodes and 344

isoparametric brick elements to model the concrete slab and column.

The reinforcement, including the prestressing wires, were modelled as bar elements

embedded in the concrete with an additional nodes used to define the embedded bars.

Straight reinforcing bars were modelled with two additional nodes per element while

the prestressing wires with parabolic prof:tles were defmed using three nodes per

element. The position of the reinforcement in the finite element model was as for the

test specimens. The specimens S2 and S3 differ only in the content of prestressing

wires and the intensity the applied prestress. Specimen S 1 had a different arrangement

of bonded reinforcement to that of specimens S2 and S3 and had no prestressing steel.

The basic difference between the models is the reinforcement arrangement and

content. Details of the nodes and elements used for the reinforcement for the different

specimens are given in Table 8.1.

The fmite element mesh used to model specimen S4 is shown in Figures 8.3 and 8.4.

The mesh consisted of 2799 nodes and 460 brick elements to model the concrete and

110 embedded bar elements to model the bonded reinforcement and prestressing

wires. As for slabs S2 and S3, in slab S4 2-node bar elements were used to represent

the straight bars and 3-node element for the parabolic cables. Details of the finite

element mesh are given in Table 8.1.


8-3

0 -x
Column - 3515

z
3515

Hinged slab
Hinged column edge
centre

FRONT ELEVATION

z
2400

Simply supported
Hinged edge

SIDE ELEVATION

Figure 8.1 - Plan and elevation showing the finite element mesh for edge column
models Sl, S2 and S3.
8-4

X
y

Plane of symmetry
0

Figure 8.2 - 3D view of f"mite element mesh for models Sl, S2 and S3.

Table 8.1 -Finite element model data for different models.

Model Number of Nodes Bonded Prestressing wire


20-node for reinfrocement
concrete concrete
elements elements Elements Nodes Elements Nodes

S1 344 2137 87 192 - -


S2 344 2137 93 204 10 30

S3 344 2137 93 204 26 78

S4 460 2799 84 200 26 78


8-5

25

[_ -x
Column
.Jo 3515
NOTE:
ml Load points

z
3515

Hinged slab
base
Hinged column
centre
FRONT ELEVATION

z
2500

Hinged slab
base
Hinged column
centre
SIDE ELEVATION

Figure 8.3 - Plan and elevation showing the finite element mesh for model S4.
8-6

X
y

Figure 8.4 - 3D view of f"mite element mesh for model S4.

For the isoparametric concrete elements numerical integration was undertaken on a

3 x 3 x 3 gauss quadrature and a modified Newton Raphson iteration procedure was

implemented to solve the problem. In the early part of the analyses (the linear range)

load control was used while load-displacement control via an updated normal arc-

length procedure was used in the latter part of the analyses. Convergence was set at 1

percent of energy norm with a maximum of 50 iterations for any one load step.

The concrete was modelled using a cohesion-plastic strain curve based on the concrete

stress-strain relationship of Hognestad (1951). The in-situ compressive strength was

taken as 0.9fcm and tensile was taken as 0.33~ fcm , where !em is the mean cylinder

compressive strength of the concrete. The modulus of elasticity used for the concrete

(Ec) was taken to match the measured linear response of the experimental tests. The

other parameters were taken as determined in Chapters 3 to 5 of this thesis. The

concrete properties used for the finite element modelling are given in Table 8.2.
8-7

A bi-linear or tri-linear elasto-plastic model (refer Figures 3.4a and 3.4b) was adopted

for the conventional and prestressing reinforcement. The values of material parameters

used are given in Tables 8.3 and 8.4 for the conventional steel and prestressing wires,

respectively.

8.3 Results of Finite Element Models

The ultimate load carrying capacities of the specimens obtained from the tests and the

numerical analyses are given in Table 8.5 and are in close agreement with that of the

experimental results. The exception is specimen S3 in which the finite element model

predicted a 20 percent higher failure load than that measured. The mean of the

numerical failure loads to the test failure loads was 1.07 with a standard deviation of

0.08.

The results of the individual tests are described below. The load versus deflections,

steel strains and reactions are compared with those of the experimental study and

concrete tension strains transverse to the failure planes are presented.

8.3.1 Results of Model Sl

The finite element model for the specimen S 1 predicted a punching failure load close

to that measured in the test (see Table 8.5). The deformed shape of the slab and the

contour plot for the vertical deflections at failure are shown in Figure 8.5. The load

versus deflection plot for the measured slab deflection (see Figure 6.31a for details) is

given in Figure 8.6. A reasonable correlation between the numerical model and test is

seen in the load-displacement history although the deflections show the finite element

model to be stiffer than the experimental slab in the post-cracking region.


8-8

Table 8.2- Concrete properties used in the FE models.

Material Property Model S1 ModelS2 ModelS3 ModelS4

fcp = 0.9 !em (MPa) 21.7 22.4 23.7 33.8

It (MPa) 1.62 1.65 1.69 2.02

Ec (GPa) 20 24 20 25

a(degrees) 15 15 15 15

l/J (degrees) 30 30 30 30

1f1 (degrees) 12.6 12.6 12.6 12.6

a1 (column) 13.9 17.2 14.6 13.9

a 1 (slab) 14.8 15.8 12.9 14.9

J.l 0.2 0.2 0.2 0.2

f3 0.2 0.2 0.2 0.2

Eo 0.002 0.002 0.002 0.002

Table 8.3- Material propertieS for conventional reinforcing bars.

Bar Area /sy Es


(mm2) (MPa) (MPa)

W6 25.5 520 200

ws 44.2 585# 240

YlO 80.5 675# 225

Y16 201 450 200

#
Note: 0.2 % Proof stress
8-9

Table 8.4 - Material properties for prestressing wires.

Size Area /sy Es Esv Ep /syJ


(mm2) (MPa) (GPa) (J.!E) (GPa) (MPa)

8 51.0 1030# 200.0 5150 21.2 1280


Note: 0.2 % Proof stress

Table 8.5 - Comparison between model and experimental ultimate load.

Specimens Model (J) Experimental (2) (1)


(kPa) (kPa) (2)

S1 18.5 18.5 1.00

S2 22.3 21.7 1.03

S3 30.7 25.5 1.20

S4 40.1 39.0 1.03


Mean 1.07
Standard Deviation 0.08
8-10

(a) 3D view

l-.1~1E-2
l-.32~E-2
-.5B7E-2
-.69E-2
-.873E-2
~ -.1B6E-1
l-.12~E-1
l-.142E-1
B-.16E-1
1-.179£-1

(b) Contour plot (View from top with the column at bottom left corner)

Figure 8.5 - Deformed shape of specimen Sl at failure (a) 3D view; and


(b) contour plot (deflection in metres).
8-11

0 10 20 30 40
Deflection (mm)

(a) Deflections at A, B and C

20
18
_16
l14
~
-12
"tJ
8 10
- 8 --Experimental E
"iii
0 6 · · ·······Experimental D
1- 4 --FEEandD
2
0 ~~~~~~~~~~~~~~~~~~~

0 2 4 6 8 10 12
Deflection (mm)

(b) Deflections at D and E

Figure 8.6 - Load versus deflections for specimen Sl.


8-12

In Figure 8. 7 the load is plotted against horizontal and vertical reactions at the base of

the column. The analytical model gave a similar result to that measured up to the

initiation of the punching failure. As the punch took place, however, the reactions on

the test slab started to decrease and this was not seen in the finite element model.

The load versus steel bar strains at the points measured in the slab (see Figure 6.32a)

are plotted in Figures 8.8 and 8.9 for the top and bottom bars, respectively. These plots

show a good correlation between the finite element model and the experimental test.

As measured in the test a top bar in the x-direction located at the centre of the column

yielded before the punching failure (see Figure 8.8a). The finite element model

indicates, however, that the adjacent bars had not yielded before the punching failure.

Therefore the slab retained residual flexural capacity and failed in a true punching

failure mode.

The experimental observations showed the failure surface to be at an average angle of

18 degrees to the horizontal .. In the finite element modelling for Slab 1, the major

principal strains across the experimentally observed failure plane were obtained and

are plotted in Figure 8.10 for increasing load. The figure shows that first cracking

occurred at approximately 40 percent of the failure load. At 60 percent of the failure

load cracking had occurred across a significant portion of the failure surface and it can

be said that the failure crack was initiated at between 40 and 60 percent of the failure

load. This observation is consistent with the experimental observations of Moe ( 1961)

whose experimental tests showed that the failure crack was initiated at about 50

percent of the failure load. Thus the results of cracking strains output in the finite

element model are consistent with those observed in the laboratory. The compressive
8-13

strain along the junction of the slab soffit and the column are shown in Figure 8.11.

The magfiitude of the principal tensile strain along this failure line is as high as 7600

Jl£, almost 100 times the concrete cracking strain. The magnitude of the maximum

compressive strain was 4890 J1£ occurring at the internal corner of the column but

reduces quickly away from the corner of the column. This indicates only local

crushing of the concrete near the slab-column connection. The principal tensile strain

contours are shown in Figure 8.12 and show the high tensile strains along the failure

surface.

20
18

-ftS
16
14
-
a..
~

"C
ftS
.2
12
10
"i 8
0 6
1-
4
2
0
0 20 40 60 80 100 120
Reaction (kN)

Figure 8.7 -Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen Sl.
8-14

20
18
_16
£f14
~ Yield limit
~2
~10
- 8
"i -Experimental Gauge 2
15 6 --FEGauge1
..... 4
~FEGauge2
2
o--~~~~~~~~~~~~4-~~~-+~~~~

0 1 2 3 4 5
Strain (E-3)

(a) X-direction reinforcement

20
18

-
: . 14
c"D 12
16

CIS 10
.2
-"i 8
0
t-
6
4
-Experimental Gauge 4
--FEGauge4
2
0
0 0.5 1 1.5 2
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.8 - Load versus strain histories for x- and y-direction top bars of
specimen Sl (a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-15

20
18

-ftl
16

-,
c.
~
14
12
ftl 10
.2
'i 8 --Experimental Gauge 6
0 6 · · ·······Experimental Gauge 7
1-
4
• FE Gauges 6 and 7
2
0
0 0.5 1 1.5
Strain {E-3)

(a) X-direction reinfrocement

20
18
_16
.f 14
c12
i0 10
- 8
'i --Experimental Gauge 8
0 6
1- 4 - - - - -Experimental Gauge 9
2 - - F E Gauges 8 and 9
o--~~~~~~~~-+~~~~4-~~~~

0 0.5 1 1.5 2
Strain {E-3)

(b) Y-direction reinforcement

Figure 8.9 - Load versus strain histories for bottom bars of specimen Sl
(a)x-direction reinforcement; and (b) y-direction reinforcement.
8-16

At failure load
Maximum strain = 7656 J.l.l:.

At 82 % of the failure load


Maximum strain = 2170 JLE

A
B At 70 % of the failure load
Maximum strain = 901 j.J.f:.

A~
A~
At about 60 % of the failure load
Maximum strain = 462 JLE

B At 50 % of the failure load


Maximum strain = 196 JJ.I!:.

A~B At 40 % of the failure load


Maximum strain = 132 j.J.f:.

A~ At about 30 % of the failure load


Maximum strain = 75 JLE

A~
B SCALE:
t-----1 : 1491 j.J.f:.
SLAB
AB: Failure plane
_.4

.11 . .4 ;

COLU_MN.
:.1 •
4 .

Figure 8.10- Principal tensile strain along failure line for specimen Sl on the
vertical surface near the central symmetrical vertical plane.
8-17

PLANE OF SLAB SOFFIT

Plane of SCALE:
symmetry f--------1 :4890 Ji-E

Figure 8.11 - Principal compressive strain at the slab and column junction for
specimen 81.

1.121E-l
2.189E-l
.964E-2
.839E-2
.714E-2
1 .589E-2
1.~65E-2
1.J4E-2
i.215E-2
1.985E-J
I

Figure 8.12 - Contour of principal tensile strain at the column region of the finite
element model for specimen 81.
8-18

8.3.2 Results of Model S2

The results of the finite element modelling of slab S2 gave a punching failure at a load

of 6 percent higher than that of the experiment. The deflected shape of the model at

the point of failure and a contour plot of the vertical displacements are shown in

Figure 8.13. The load versus deflection histories for different key points in the slab are

plotted in Figure 8.14. The deflections obtained from the finite element model

compare reasonably well with the test results although the model slab was stiffer than

that of the experiment.

The load versus horizontal and vertical reactions are given in Figure 8.15. The figure

shows a very good correlation between the numerical and the test results.
8-19

(a) 3D view

L.t32E-2
1-.32E-2
t-.SBBE-2
-.696E-2
-.883E-2
-"-.187E-l
l-.126E-1
•-.HSE-1
1-.163E-1
I-.182E-1

(b) Contour plot (View from top with the column at the bottom left corner)

Figure 8.13 - Deformed shape of specimen S2 at failure (a) 3D view; and


(b) contour plot (displacement in metres).
8-20

25

a;- 20 Experhnental A
c.
~
-"C 15
~
"i 10
0
..... 5

0 10 20 30 40
Deflection (mm)

(a) Deflections at A, B and C

25

-20
ftl
c.
=- 15
"C
ftl
.S!
"i 10 -Experimental D
0 -Experimental E
..... 5
--FEDandE

0
0 1 2 3 4 5 6
Deflection (mm)

(b) Deflections at D and E

Figure 8.14- Load versus deflections for specimen S2.


8-21

25 FERx FERz

-
ca 20

,-
a.
~
15
ca
.2
"i
10
0
.... 5

0
0 50 100 150
Reaction (kN)

Figure 8.15 -Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S2.

Load versus strain histories for the reinforcing bars gauged in the experiment (see

Figure 6.32b) are shown in Figures 8.16 and 8.17 for the top and bottom bars,

respectively. These plots show a reasonable correlation between the finite element

model and the experimental test although the bottom x-direction reinforcement

indicates the finite element slab is stiffer than that of the experiment. This is consistent

with the deflection observations. The strain histories show the maximum strains in the

reinforcing bars were well below the yield strain.

The distribution of principal tensile strain along the experimentally measured line of

failure is plotted in Figure 8.18. The maximum principal tensile strain along this

failure line was 8000 JlE, 120 times larger than the cracking strain. The compressive

strains at the slab soffit close to the column junction are plotted in Figure 8.19. For
8-22

FE Gauge 1
25

-c
ca
a.
20

15
"CJ
ca
.2
"i 10
0
.....
5 Yield line

0
0 0.5 1 1.5 2 2.5 3 3.5
Strain (E-3)

(a) X-direction reinforcement

25

fti' 20
~
- 15
~ 10 --Experimental Gauge 5
~
..... 5
- - Expermental Gauge 6
- F E Gauges 5 and 6

0 0.5 1 1.5 2
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.16- Load versus strain histories for x- andy- direction top bars of
specimen S2 (a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-23

25

-20
ca

,-ca 15
~
~

.S!
"i 10 - Experiemntal Gauge 7
0 -Experimental Gauge 8
..... 5 --FE Gauges 7 and 8

0
0 0.5 1 1.5 2
Strain (E-3)

(a) X-direction reinfrocement

25

-20
:.
c"'C 15
!
"i 10
0 -Experimental Gauge 9
..... 5 --FEGauge9

0 ~~~~~~~~~~~~~~~~~~~~~

0 0.2 0.4 0.6 0.8 1 1.2


Strain (E-3)

(b) Y-direction reinforcement.

Figure 8.17 -Load versus strain histories for bottom bars of specimen S2
(a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-24

SCALE:
1----------l : 1856f.U; ~COLUMN
Crocking strain = 70 f.-LC 4

Figure 8.18 - Principal tensile strain along failure line for specimen S2 on the
vertical surface near the central symmetrical vertical plane.

PLANE OF SLAB SOFfiT

5240 J.t.e(Maximum)

Plane of Scale:
symmetry • '• -: •• : ~ 1-------l :5240 j.tt:
......

Figure 8.19 - Principal compressive strain at the slab soffit and column junction
of specimen 82.
8-25

simplicity the directions of compressive strains are not considered in the plot. The

maximum compressive strain was 5240 J.le at the comer of the column but reduces

quickly away from the comer of the column indicating only local crushing of the

concrete. The contours of the principal tensile strain in the region of the connection are

plotted in Figure 8.20. The distribution of the principal tensile strains obtained from

the finite element model show that the failure of the specimen is due to punching.

8.3.3 Results of Model 83

The results of the finite element model of slab S3 gave a failure load of 30.7 kPa, 20

percent higher than the experimental failure load. The deformed shape of the finite

element model and the vertical displacement contours are shown in Figure 8.21. The

load versus deflection graphs comparing the results of finite element model to the

experimental results are presented in Figure 8.22. Although the trends of the curves

obtained in the finite element model are similar to the experimental results, the finite

element model is stiffer and cracked later than was observed in the test.

Plots of the applied load versus horizontal and vertical reactions are given m

Figure 8.23. The reactions obtained from the numerical model are in good agreement

with the test data.

The numerical and experimental load versus strain histories at the locations shown in

Figure 6.32c, are compared in Figures 8.24 and 8.25 for the top and the bottom

reinforcing bars, respectively. As for the deflections the strains from the finite element

model indicate that the numerical model was stiffer and cracked later than the test

slab.
8-26

i.171E-1
n.lS<E-1
.137E-1
.119E-1
.192E-1
l.a«-z
!.677E-2
1.504E-2
i.331E-2
:.159E-2

Figure 8.20 - Contour of principal tensile strain at the column region of the finite
element model S2.
8-27

(a) 3D view

1-.126E-2
1-.295E-2
-.465E-2
-.635E-2
-.8B4E-2
f-.974E-2

l . ~. )( t-.114E-l
1-.131E-1
1-.HSE-1
~-.165E-1

(b) Contour plot (View from top with the Column at bottom left corner)

Figure 8.21 - Deformed shape of specimen S3 at failure (a) 3D view and


(b) contour plot (deflection in metres).
8-28

35
30
-
~
.:.:::
25
:; 20
as
.2 15 -Experimental A
"i -Experimental 8
0 10 ---FE A
1-
-+-FE 8
5
0
0 5 10 15 20 25
Deflection {mm)

(a) Deflections at A and B

35
30
-
a.as 25
.:.:::
:; 20
as
.2 15
"i -Experimental D
0
1- 10 -Experimental E
-----FE D and E
5
0
0 1 2 3 4 5 6
Deflection {mm)

(b) Deflections at D and E

Figure 8.22 - Load versus deflections for specimen 83.


8-29

35

-
: . 25
30

-,
.:..::
ca
.2 15
20

"i
0 10
1-
5
0
0 50 100 150 200
Reaction (kN)

Figure 8.23 - Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S3.
8-30

35
30
-t.
~
25
:; 20
ca
.2 15 -Experimental Gauge 1
"i --Experimental Gauge 2
15 10
t- -FEGauge1
5 -+-FE Gauge 2

0
0 0.5 1 1.5 2 2.5 3
Strain (E-3)

(a) X-direction reinforcement

35

-t.
~
25
30

:; 20
ca
.2 15
"i -Experimental Gauge 3
15 10 -Experimental Gauge 4
t-
-FE Gauges 3 and 4
5
0
0 0.5 1 1.5 2 2.5
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.24- Load versus strain histories for x- andy- direction top bars of
specimen S3 (a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-31

35

-
:. 25
~
30

:;- 20
8 15
"iii -Experimental Gauge 5
0 10
1- --Experimental Gauge 6
5 -FE Gauges 5 and 6
o--~~~~-r~~~~~~~L-~~+-~~~~~

0 0.5 1 1.5 2
Strain (E-3)

(a) X-direction reinfrocement

35

-
: . 25
~
30

:;- 20
ftS
,g 15
-Experimental- Gauge 7
"iii
0 10 --Experimental- Gauge 8
1-
5 -FE Gauges 7 and 8

0
0 0.2 0.4 0.6 0.8 1 1.2
Strain (E-3)

(b) Y-direction reinforcement.

Figure 8.25 - Load versus strain histories for bottom bars of specimen S3
(a)x-direction reinforcement; and (b) y-direction
reinforcement.
8-32

The principal tensile strains along the measured failure line are plotted in Figure 8.26.

The maximum principal tensile strain was 8400 p,e, over 100 times the cracking strain

of the concrete. The principal compressive strains near the junction of slab soffit and

the column are plotted in Figure 8.27. The maximum compression strain was 4770 J.le

at the corners of the slab but, as for slabs S 1 and S2, reduces quickly away from the

corner. The principal tensile strain contours in the critical region of the connection are

shown in Figure 8.28.

8.3.4 Results of Model 84

The finite element model for slab S4 failed in punching shear at the load of 40 kPa and

compares favourably with the experimental result of 39 kPa. Figure 8.29 shows the

displaced shape and displacement contours obtained from the finite element analysis.

The load versus deflection plots for the points shown in Figure 6.31 b are given in

Figure 8.30. The numerical model predicted a stiffer response than measured in the

test. The earlier cracking observed in the experimental tests combined with the earlier

degradation of stiffness, compared to the finite element results, could be due to the

effect of the disturbance caused by the prestressing ducts. The duct may act as crack

inducers and with the unbonded tendons there is some loss of cross-sectional area.

In Figure 8.31 the load versus horizontal and vertical reactions are plotted. Some

differences between the finite element results and the experimental test are seen near

the failure load, otherwise the results show a reasonable correlation. The load versus

strain histories for the bar locations, shown in Figure 6.32d, are presented in

Figures 8.32 and 8.33 for the top and bottom reinforcement, respectively. The strain

histories determined from the finite element modelling are generally stiffer than those

measured.
8-33

LJ"¥
SCALE:
1----------l : 2080 P,l:

Crocking strain = 80 J..LE .


COLUMN

.ll ~

Figure 8.26 - Principal tensile strain along failure line for specimen S3 on the
vertical plane near the central symmetrical vertical surface.

PLANE OF SLAB SOFRT

4770 p.t:(Maximum)
#.--:-1--....,....":-"~~-J

Plane of SCALE:
symmetry 1--------l : 4 77 0 P,l:

Figure 8.27 - Principal compressive strain at the slab soffit and column junction
for specimen S3.
8-34

1.14SE-1
l.133E-1
.117E-1
.1B2E-1

t:~~=~
l.559E-2
L4B5E-2
I.251E-2
1.977E-3
I

Figure 8.28 - Contour of principal tensile strain at the column region of the finite
element model for specimen S3.
8-35

(a) 3D view

•-.137£-2
i-.32E-2
-.593£-2
-.SBSE-2
-.BSBE-2
~- .185E-l
l-.123E-l
1-.H2E-l
i-.!SE-1
1-.178E-l
I

(b) View from top (column located at the bottom left corner)

Figure 8.29 - Deformed shape of specimen S4 at failure (a) 3D view; and


(b) contour plot (displacement in metres).
8-36

Experimental C
50
45

-c
ca
a.
40
35
30
"C
ca 25
.2

-
'ii
0
1-
20
15
10
5
0
0 10 20 30 40 50 60
Deflections {mm)

Figure 8.30 - Load versus deflections at A, B and C for specimen S4.

Experimental Ry
50
Experimental Rx
45

-c
:.
40
35
30
FERy

i0 25
-
'ii
20
15 15
1- 10
5
o~~~~-r~~~-+-L~~~~~-L~+-~~~~

0 20 40 60 80 100
Reaction {kN)

Figure 8.31 - Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S4.
8-37

45
40
m 35
-i
~ 30

.2 20
25
- Experiemntal Gauge 1
'ii 15
0 -FEGauge1
..... 10 Yield limit
5 -+-FE Gauge 2
0
0 1 2 3 4 5 6
Strain (E-3)

(a) X-direction reinforcement

45
40
m 35
-
~ 30
"C
~ 20
25
-Experimental Gauge 3
1i 15 -Experimental Gauge 4
0
..... 10 --.--FE Gauge 3
5 -FEGauge4
o--~~~-r~~~-+~~~~~~~~+-~~~~

0 0.5 1 1.5 2 2.5


Strain (E-3)

(b) Y-direction reinforcement

Figure 8.32 - Load versus strain histories for x- and y- direction top bars of
specimen S4 (a) x-direction reinforcement; (b) y-direction
reinforcement.
8-38

45
40
Ci 35

-"'
~
"tJ
.2
30
25
20 -Experimental Gauge 5
"i 15 -Experimental Gauge 6
0
..... 10 -+--FE Gauge 5
5 ----FE Gauge 6
0
0 0.5 1 1.5 2 2.5 3
Strain (E-3)

(a) X-direction reinfrocement

45
40
Ci 35

-"'
~
"tJ
.2
30
25
20 -Experimental Gauge 7
"i 15
0
..... ----FE Gauge 7
10
5
0
0 0.5 1 1.5
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.33 - Load versus strain histories for bottom bars of specimen S4
(a)x-direction reinforcement; and (b) y-direction
reinforcement.
8-39

The distributions of principal tensile strains along the experimentally measured lines

of failure near the free edges of the :XZ and YZ-planes are shown in

Figures 8.34 and 8.35. The maximum principal tensile strain was 11970 Jl£, 145 times

larger than the concrete cracking strain. The principal compressive strains developed

near the junction of the column and the slab soffit are shown in Figure 8.36. The local

maximum compression strain near the internal comer of the column was 5045 Jl£.

Away from the column comer the compression strain reduced rapidly indicating only

local crushing of the slab near the comer of the column. The principal tensile strain

contours are shown in Figure 8.37 further highlighting the high strain region in the

vicinity of the connection.

8.4 The Effect of Prestress on Edge Column Connections

To cover the full range of prestress used in practice, a model with an average prestress

of 3 MPa was developed with a separate design being undertaken. As for the other

specimens the slab was designed for a factored design load of 14.75 kPa. All

parameters including the concrete properties and bonded reinforcement were kept as

for model S3. The new model with the prestress of 3.0 MPa is designated as

Model SE-3.0. The failure load obtained from finite element analysis is compared with

the finite element analyses of the other edge column connection models in Table 8.6.

The deflections at point A (see Figure 6.31a) on the slab are compared in Figure 8.38.

The failure loads against the average prestress applied to the slab are plotted in

Figure 8.39. The figures show an almost linear trend of increase in the load carrying

capacity of the slab with the increase in prestress.


8-40

SCALE:
1------1 1009 """ COLUMN
Crocking strain = 80 JJ.£

Figure 8.34 - Principal tensile strain along failure line for specimen 84 near the
free XZ-surface.

SCALE:
1------1 2335 p,e COLUMN

Crocking strain = 80 p,e

Figure 8.35 - Principal tensile strain along failure line for specimen 84 near the
free YZ-surface.
8-41

PLANE OF SLAB SOFFIT

COLUMN
SCALE:
:5045 f..U;

Figure 8.36 • Principal compressive strain along the slab soffit and column
junction for specimen S4.

I.HJE-1
f.128E-1
.ll~E-1
.SS2E-2
.MGE-2
~. 781E-2
t555E-2
l.~1E-2
1.26~E-2
1.119E-2
I

Figure 8.37 - Contour of principal tensile strain at the column region of the finite
element model S4.
8-42

40
--s1-o.o
35 -+-S2-0.7
-+-S3-2.2
-ca
a.
30
25
__._SE-3.0

,-
~

ca
20
0
...I
15
10
5
0
0 5 10 15 20 25

Deflection at A (mm)

Figure 8.38 - Load versus deflection plots for deflection at point A for finite
element results of edge column connections with varying
amounts of prestress.

Table 8.6 - Comparison of the failure loads for edge column connection models
with different prestress.

FE Model Average Prestress Failure load


(MPa) (kPa)

Sl-0.0 0 18.5

S2-0.7 0.7 22.9

S3-2.2 2.2 31.4

SE-3.0 3.0 34.3


8-43

40
-35
cu

-
fl. 30
~

; : 25
,;; 20
cu
.2 15
1a
0 10
1-
5
0
0 1 2 3 4
Average prestress, (P/A)Avg (MPa)

Figure 8.39 - Load versus average prestress applied to slabs of edge column
connections (results off"mite element models).

8.5 The Effect of Prestress on Corner Column Connections

To study the influence of prestress on the behaviour of corner column connections the

finite element model of a corner slab-column specimen similar to specimen S4 was

developed by designing tendons for a factored design load of 14.75 kPa (as for the

other models). The models are designated as S4-X.X, where X.X is the average

prestress in MPa applied to the slab. For example, Model S4-2.1 refers to model S4

with the average applied prestress in slab of 2.1 MPa. With the exception of the

arrangement and the intensity of the applied prestress, the amount and arrangement of

conventional bonded reinforcement, the geometry and boundary conditions and

concrete material properties for all S4-X.X models (see Figures 8.3 and 8.4) were as

for the experimental slab S4.


8-44

The failure loads for slab S4 with the different amount of prestress in slab are

compared in Table 8.7. The load deflection curves are compared in Figure 8.40 and

the failure load versus the applied prestress are plotted in Figure 8.41. As for the edge

column connections the load carrying capacity increases approximately linearly with

increasing prestress.

The horizontal reactions in x- and y-directions and the vertical reaction developed at

the base of the column are compared in Figures 8.42, 8.43 and 8.44, respectively. The

figures show that the prestress has only minor effect on the distribution of the

reactions. Strains in the x- and y-direction top bars at the locations shown in

Figure 6.32d are plotted in Figures 8.45 and 8.46, respectively. The strain histories of

x-direction bottom bars are shown in Figure 8.47. In no case had the steel yielded

indicating that the failure load was not limited by the flexural strength of the slab.

8.6 Behaviour of Slab-Column Connections with Bonded Tendons

While the use of unbonded tendons in construction is common practice in many

countries, in Australian practice it is mandatory to grout tendons in prestressed

concrete structures. Finite element modelling was used to investigate the influence of

bonding of the prestressing steel on the behaviour of the edge and comer column

specimens.

The numerical models studied in the previous sections were modified with the

prestressing wires bonded to the concrete. The models are designated as S2-B, S3-B

and S4-B. All properties other than the bonding of the prestressing wires, including the

geometry and boundary conditions, were as for the previous models. The finite
8-45

50
45

-~ 35
40

c"D 30
ca 25 -s4-o.o
.2
"iii 20 -+-S4-1.0
0
..... 15 ......_84-2.1
_._S4-3.0
10
5
0
0 5 10 15 20
Deflections at A (mm)

Figure 8.40 - Load versus deflections at point A for slab model S4 with different
intensity of average prestress.

Table 8.7 - Comparison of the failure loads for corner column connection models
with different levels of prestress.

F.E. Model Average Prestress Failure load


(MPa) (kPa)

S4-0 0 18.6

S4-1 1.0 26.5

S4-2 2.1 40.1

S4-3 3.0 47.7


8-46

60

"'5o
a.
.!II:
-40
;::
"ti
ca 30 W =10.0 {P/A)Avg + 17.9
.2
as 20
0
1- 10

0+-~~~~,_~~~~-r~~~~-+~~~~~

0 1 2 3 4
Average prestress, (P/A)Avg (MPa)

Figure 8.41 - Finite element results of failure load versus average prestress for
comer column connection model S4.

50
45

-ca
40

-
a.
.!II:

i
.2
35
30
25 --84-0.0 MPa
as 20
-+-84-1.0 MPa
0 15
1- 10 -lr- 84-2.1 MPa
5 --e-- 84-3.0 MPa
0
0 10 20 30 40 50 60
Reaction (Rx) (kN)

Figure 8.42 - Load versus x-direction horizontal reaction (Rx) at the column
support for different comer column models.
8-47

50
45
-as
a.
40

,-
~
35
30
as 25 --54-0.0 MPa
.2
"i 20 -+--54-1.0 MPa
0 15
..... 10 _.....54-2.1 MPa

5 __._ 54-3.0 MPa


0
0 10 20 30 40
Reaction (Ry) (kN)

Figure 8.43- Load versus y-direction horizontal reaction (Rv) at the column
support for different corner column models.

50
45
-a.
as
40

-,
~

as
.2
35
30
25 - - 54-0.0 MPa

"i 20 -+--54-1.0 MPa


0 15 __.__ 54-2.1 MPa
..... 10
__._ 54-3.0 MPa
5
0
0 20 40 60 80 100 120
Reaction (Rz) (kN)

Figure 8.44 -Load versus vertical reaction (Rz) at the column support for
different comer column models.
8-48

60

-
ca
50

-
~ 40
' tJ
ca 30
.2 -84-0.0 Gauge 1
"i 20 -+--84-1.0 Gauge 1
0 __..._ 84-2.1 Gauge 1
1- Yield limit
10 ---- 84-3.0 Gauge 1
0
0 1 2 3 4 5
Strain (E-3)

60 ~ 84-0.0 Gauge 2
-+--84-1.0 Gauge 2

-ca
50 -&- 84-2.1 Gauge 2
-e- 84-3.0 Gauge 2

-
~ 40
' tJ
ca 30
.2
"i 20
0
f- Yield limit
10

0
0 1 2 3 4
Strain (E-3)

Figure 8.45 - Load versus strain histories for x- direction top bars for
comer column connection with different amount applied
prestress.
8-49

60

-ca
50

-
~ 40
"C
ca 30
.2

-
ca
0
.....
20
10
--- 54-0.0 Gauge 3
~54-1.0 Gauge 3
.....,.__ 54-2.1 Gauge 3
--- 54-3.0 Gauge 3
0
0 0.5 1 1.5 2 2.5 3
Strain (E-3)

60

-
ca
50

-
~ 40
"C
ca 30
.2 -s- 54-0.0 Gauge 4
ca 20 ~54-1.0 Gauge 4
0 --&- 54-2.1 Gauge 4
.....
10 -e- 54-3.0 Gauge 4

0
0 0.5 1 1.5 2 2.5
Strain (E-3)

Figure 8.46 -Load versus strain histories for y- direction top bars for
corner column connection with different amount applied
prestress.
8-50

60

-
ca
50

-
~ 40
"C
ca 30
.2
'iii 20 - - 84-0.0 Gauge 5
0 ~84-1.0 Gauge 5
1-
10 .....,._ 84-2.1 Gauge 5
---*-- 84-3.0 Gauge 5
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain (E-3)

60

-
ca
50

-
~ 40
"C
ca 30
.2
'iii 20 -e- 84-0.0 Gauge 6
0 --.-84-1.0 Gauge 6
1-
10 -tr- 84-2.1 Gauge 6
--e- 84-3.0 Gauge 6
0
0 0.2 0.4 0.6 0.8 1 1.2
Strain (E-3)

Figure 8.47 - Load versus strain histories for x-direction bottom bars of corner
column connections with different applied prestress.
8-51

element mesh is shown in Figures 8.1 and 8.2 for models S2-B and 83-B. Similarly the

finite element mesh for the model S4-B is as shown in Figures 8.3 and 8.4. With the

exception of a1 the material parameters used were as for the unbonded tendon

analyses. The slope of the softening branch of the concrete tension stress-strain law

(represented by a1) was modified to allow for additional tension stiffening effect due

to bonding of the prestressing wires. The values of a1 are calculated from

equation 3.30 and are given in Table 8.8.

The failure loads for the finite element analyses are compared in Table 8.9. The effect

of bonding the prestressing wires was to increase the failure load with the largest

relative increase for the specimen with higher levels of prestress. For specimens S3-B

and S4-B with 2.2 and 2.1 MPa of prestresses, respectively, the increase in the failure

load for the bonded compared to the unbonded prestress was 20 and 14 percent,

respectively.

8.6.1 Results of Model 82-B

The load versus deflection at point A on the slab (refer Figure 6.31a) is plotted in

Figure 8.48 for the bonded and unbonded analyses. The figure shows that overall

responses of the specimens are similar for both the bonded and unbonded models. The

load versus reactions are plotted in Figure 8.49. The distributions of the reactions are

not affected by bonding of the tendons.

Finally, the strain histories are compared for the bonded and unbonded models in

Figures 8.50 and 8.51 for the top and bottom reinforcement, respectively. As for the

deflections and reactions there is little difference between the bonded and unbonded

prestressing systems.
8-52

Table 8.8 - Values of a1 used for modelling of the bonded tendons.

aJ S2-B S3-B S4-B

a 1 (Column) 17.2 14.6 13.9

a1 (Slab) 19.8 16.1 18.6

Table 8.9 - Comparison of failure loads for the unbonded and bonded finite
element models.

Average Failure Load (kPa) B


-
Prestress u
Model (MPa) Unbonded Bonded
Tendons(U) Tendons(B)

S2 0.7 22.9 24.3 1.06

S3 2.2 30.7 36.8 1.20

S4 2.1 40.1 45.6 1.14


8-53

30

25
-
ca

-
~20
"C
~15

~10 ----- Unbonded


1- -a-Bonded
5

0
0 5 10 15 20 25
Deflection at A (mm)

Figure 8.48 - Load versus deflections for specimen S2.

30

25
-
-
:~. 20
"C
ca 15
.2
- - Rx (Unbonded)
'i
0 10 -e- Rx (Bonded)
1- -+- Rz (Unbonded)
5 -+- Rz (Bonded)

0
0 20 40 60 80 100 120 140
Reaction (kN)

Figure 8.49 - Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S2.
8-54

30

25
(i

-
~ 20
"0
ftS
.2
15
Yield limit

a; --Gauge 1 (Unbonded)
'5 10 -a- Gauge 1 (Bonded)
1- -+-Gauge 2 (Unbonded)
~Gauge 2 (Bonded)
5 -+--Gauges 3 and 4 (Unbonded)
---&---Gauges 3 nad 4 (Bonded)
0
0 1 2 3 4
Strain (E-3)

(a) X-direction reinforcement

30
_25
ftS

-
~20
"0
!15
-Gauges 5 and 6 (Unbonded)
"i 10
'5 -a- Gauges 5 and 6 (Bonded)
1-
5
0
0 0.5 1 1.5 2
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.50- Load versus strain histories for top bars of specimen 82
(a)'x-direction reinforcement; and (b) y-direction
reinforcement.
8-55

30

25
-CIS

-
c.
,
~
20
CIS
.2
15
'iii
0 10
1- ----Gauges 7 and 8 (Unbonded}
5 -s- Gauges 7 and 8 (Bonded}

0
0 0.2 0.4 0.6 0.8 1
Strain (E-3)

(a) X-direction reinfrocement

30

-
CIS
25

-,
c.
~

CIS
.2
20

15
'iii
0 10 ----Gauges 9 and 10 (Unbonded}
1-
5 -s- Gauges 9 and 1o (Bonded}

0
0 0.2 0.4 0.6 0.8 1 1.2
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.51 - Load versus strain histories for bottom bars of specimen S2
(a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-56

8.6.2 Results of Model S3-B

The load versus deflection plot for the bonded model S3-B is compared with the

unbonded model S3 in Figure 8.52. The reactions at the column supports are shown in

Figure 8.53 and the strain histories for the top and the bottom reinforcement are

compared in Figures 8.54 and 8.55, respectively. The bonded and unbonded systems

have a similar overall response but with an increase in failure load for the bonded

prestress connections over that of the unbonded system.

8.6.3 Results of Model S4-B

The load versus deflection plots for points A and D of the slab (refer Figure 6.31 b) are

shown in Figure 8.56. The reactions are presented in Figure 8.57 and the strain

histories for the top and bottom main bars are compared in Figures 8.58 and 8.59,

respectively. As for slabs S2 and S3, the response of slab S4-B is similar to that of the

unbonded finite element models.

8.7 Discussions and Conclusions

In this chapter the finite element model DIANA was used to analyse the experimental

test slab-column connections. The model was calibrated and verified in Chapters 3 to 5

of this thesis. The finite element model generally simulated the overall response of the

test specimens well although some variance occurred in some of the measured versus

analytical data. In particular the redistribution of horizontal reactions measured in test

specimens S 1 and S4 that occurred towards the end of the tests were not seen in the

finite element modelling. The fact that this redistribution was not also observed in the

experimental results of specimens S2 and S3 may indicate that at high loads some

locking of the pin base occurred. Alternatively, some load redistribution may have

resulted as failure of the connection became imminent.


8-57

40
35
Ci' 30
a.
c 25
, 20
"'
.2

-
'ii 15
0
1- 10
------ Unbonded
-a-Bonded

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

Figure 8.52 - Load versus deflections for specimen S3.

40
35
Ci' 30
a.
c 25
,
"' 20 - Rx {Unbonded)
.2
-e- Rx (Bonded)
'ii 15
0 --+-- Rz (Unbonded)
1- 10 ~ Rz (Bonded)
5
0
0 50 100 150 200 250
Reaction (kN)

Figure 8.53 - Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S3.
8-58

40
35
Ci' 30
c.
,c
as 20
25 Yield limit

.2 -Gauge 1 (Unbonded)

-
"i 15
0
1- 10
-e- Gauge 1 (Bonded)
-+-Gauge 2 (Unbonded)
~Gauge 2 (Bonded)
.......,.._Gauges 3 and 4 (Unbonded)
5 --b- Gauges 3 and 4 (Bonded)
0
0 1 2 3 4
Strain (E-3)

(a) X-direction reinforcement

40
35
-c
c.as 30
,as 25
20
.2
"i 15 -Gauges 5 and 6 (Unbonded)
0
1- 10 -e- Gauges 5 and 6 (Bonded)

5
0
0 0.5 1 1.5 2 2.5
Strain (E-3)

(b) Y-direction reinforcement

Figure 8.54- Load versus strain histories for x- andy- direction top bars of
specimen S3 (a) x-direction reinforcement;.(b) y-direction
reinforcement.
8-59

40
35
"i' 30
c.
~ 25
'§ 20
'ii 15
'5
1- 10 ---Gauges 7 and 8 (Unbonded)
5 -e- Gauges 7 and 8 (Bonded)
o--~--~-L--~~~--~~--~~--~~~--~~

0 0.5 1 1.5
Strain {E-3}

(a) X-direction reinfrocement

40
35
"i' 30
c.
~ 25
"D
ca 20
.2
'ii 15 --Gauges 9 and 10 (Unbonded)
'5
1- 10 -e- Gauges 9 and 1o (Bonded)
5
0
0 0.5 1 1.5
Strain {E-3}

(b) Y-direction reinforcement

Figure 8.55 - Load versus strain histories for bottom bars of specimen S3
(a) x- direction reinforcement; and (b) y-direction
reinforcement.
8-60

50
45
40
-
"' 35
a.
~ 30
'a
25
"'
.S! - - A (Unbonded)
'ii 20 -e- A (Bonded)
0 15 -+- B (Unbonded)
1-
10 ~B(Bonded)

5
0
0 5 10 15 20 25
Deflections (mm)

Figure 8.56 - Load versus deflections at A and B for specimen S4.

50
45
-40
:.
~
35
-30 - - Rx (Unbonded)
i0 25 -e- Rx (Bonded)
- 20 -+- Ry (Unbonded)
'ii ~ Ry (Bonded)
0 15
1- 10 __.__ Rz {Unbonded)
--&- Rz (Bonded)
5
o~~~~~~~~~~-r~~~~~~,_~~~

0 20 40 60 80 100 120
Reaction (kN)

Figure 8.57- Load versus horizontal (Rx) and vertical (Rz) reactions at column
support for specimen S4.
8-61

50
45
-m
40

-
ll.
~

"D
m
.S!
35
30
25 ---Gauge 1 (Unbonded)
"i
20 -e- Gauge 1 (Bonded)
0 15 Yield limit --+-Gauge 2 ( unbonded)
1- 10 -+--Gauge 2 (Bonded)
5
0
0 1 2 3 4 5 6
Strain {E-3)

(a) X-direction reinforcement

50
45

-m
40
35
-
ll.
~

"D
m
.S!
30
25 ---Gauge 3 (Unbonded)
20 -e- Gauge 3 (Bonded)
"i
0 15 --+-Gauge 4 (Unbonded)
1- -+--Gauge 4 (Bonded)
10
5
0
0 0.5 1 1.5 2 2.5
Strain {E-3)

(b) Y-direction reinforcement

Figure 8.58 - Load versus strain histories for top bars of specimen S4.
(a) x-direction reinforcement; and (b) y-direction reinforcement.
8-62

50
45
-40
:.
.lll::
35
-30
"tS -+-Gauge 5 (Unbonded)
~ 25
- 20 -e- Gauge 5 (Bonded)
'ii --Gauge 6 (Unbonded)
0 15
1- 10 --+-Gauge 6 (Bonded)
5
o.-~--~~~~,_~--~-L--~~~--~~--~~

0 0.5 1 1.5
Strain {E-3)

(a) X-direction reinfrocement

50
45
-40
:. 35
c"tS 30
~ 25
--Gauge 7 (Unbonded)
'ii 20
0 15 -e- Gauge 7 (Bonded)
1- -+-Gauge 8 (Unbonded)
10
--+-Gauge 8 (Bonded)
5
0
0 0.2 0.4 0.6 0.8 1
Strain {E-3)

(b) Y-direction reinforcement.

Figure 8.59 - Load versus strain histories for bottom bars of specimen S4
(a) x-direction reinforcement; and (b) y-direction
reinforcement.
8-63

The difference between the measured and the predicted failure load was higher for the

specimen S3 compared to that of the other specimens and the stiffness degradation of

specimen S4 was greater than that produced by the finite element model. In these

specimens the closeness of ducting may have formed weakness in the slabs leading to

earlier cracking of the specimens than predicted by the finite element modelling.

The numerical models were used to study the effect of the level of prestress on the

comer column connections. Slabs were analysed for zero to 3.0 MPa of average

prestress. As for the experimentally observed edge connections, the level of prestress

was found to influence the load carrying capacity of the comer connections. For

3.0 MPa of prestress the increase in capacity of the connection was 85 percent above

that of the non-prestressed slab-column connection. The capacity of the connections

increased approximately linearly with increasing levels of prestress.

Finally, the finite element models were used to access the differences between bonded

and unbonded prestressed slab systems. It was found that bonding the prestressing

tendons to the concrete did not influence the distribution of reactions, the overall

deflection response or the development of steel strains with load. Higher punching

failure loads, however, were obtained for the bonded prestressed slab system over that

of the unbonded system with the greatest load increase being for the slabs with the

highest level of prestress.


CHAPTER 9 - CONCLUSIONS

9.1 Introduction

In this study the ultimate behaviour of flat plate slab-column connections has been

investigated with the focus being on prestressed concrete edge and comer slab-column

connections. The load path in these structures is three-dimensional and with the

non-linear material behaviour the response of the slabs under load is complex.

The research described in this thesis comprises of three parts; part one was an

investigation into the suitability of using finite elements to model punching shear

failure in flat plate slab-column connections using 3D solid brick elements. In part two

an experimental programme was established to study the effects of prestress on the

behaviour of flat plate-column connections. Part three dealt with the numerical

modelling of the test specimens with unbonded and bonded tendons. The effect of

prestress level on the behaviour of these connections was investigated.

The work undertaken in this thesis has identified a number of areas where the research

remains deficient. Recommendations for further study are given in this chapter.

9.2 Concluding Remarks

In this study an investigation into the development of efficient finite element models

to analyses punching type failures was undertaken. Analyses of a plain concrete slab

panel under a pullout load were conducted using DIANA (1997) to study the effects of

mesh grading and the Gaussian quadrature used for numerical integration. The failure

loads obtained from the finite element model were compared with the ACI code model
9-2

and with the plasticity model of Nielsen et al. (1978) for anchorage pullout out

strengths. The ACI 318 was found to be conservative while the failure loads obtained

from the Nielsen et al. model compared well with the numerical results.

Numerical tests on the finite element mesh grading showed that the mesh grading has

a significant effect on the results of the analyses. It was observed that the mesh in the

critical region requires an aspect ratio not greater than 2 while away from the critical

zone the mesh aspect ratio may be increased. Two elements through the thickness of

the slab were found to be sufficient for modelling up to the peak load. Numerical

integration over a 2 x 2 x 2 Gaussian quadrature gave good numerical results without

lose of accuracy when compared to analyses using a 3 x 3 x 3 quadrature. The

computational time and the disk space consumed for the analysis with 2 x 2 x 2

scheme is significantly less than that for the 3 x 3 x 3 scheme.

The behaviour of an isolated internal reinforced concrete slab-column was analysed to

investigate the effect of flexural reinforcement quantity on the behaviour of slabs

subject to a punching loading. Two distinct modes of failure were observed from the

model, that of flexure and that of punching shear. The primary failure mode is one of

flexure until a critical quantity of flexural reinforcement is used. Beyond this point

punching shear governs the failure. For increasing reinforcement content beyond the

critical point the failure load was independent of the quantity of flexural

reinforcement. While the effect of dowel action is not included in the numerical model

the results are consistent with the experimental observations of Dragosvic and

van Den (1974).


9-3

Sensitivity tests of the different parameters needed to model slabs using the 3D brick

element available in DIANA were carried out to provide an insight into the effect of

the parameters on the interpretation of the finite element results. The shear retention

factor f3 and the slope of the descending branch of the tension stress-strain law of

concrete (given by a1) were observed to have a significant influence on the results of

the finite element analyses. Other parameters such as friction angle, dilatancy angles,

slope of the descending branch of concrete stress-strain law were observed to have

only minor effects in the numerical results. The value of f3 = 0.2 for the shear retention

factor was found to give the most consistent numerical results for a wide range of

experimental tests on specimens failing in shear or torsion modes. The value of a1

must be carefully chosen to represent the tension stiffening effect. The values of

different parameters recommended by Vermeer and de Borst (1984) and by van Mier

( 1987) were found to be suitable for the finite element modelling of punching shear

type problems.

An experimental program and subsequent numerical modelling of a series of pullout

tests are reported. The pullout tests were undertaken to further verify the shear

retention factor (/3) used in the numerical modelling. Six large square rectangular

prism blocks were tested with two different concrete strengths and with an anchor

plate embedded at three different depths. The test set-up was such that the boundaries

of the slabs were located far away from the anchor plates to elin_rinate boundary

effects. Finite element analyses of the specimens simulated the experimental

behaviour generally well but with some exceptions. The shear retention factor f3 =0.2
was confirmed to be suitable for the numerical modelling of these problems.
9-4

A major component of this study was testing of four slab-column specimens

developed as one-third scale models of a prototype building. Of the four specimens

tested, three specimens represented edge flat plate-column connections and one a

corner flat plate-column connection. The primary variable was the level of prestress in

the slab.

All four specimens tested failed in a punching shear mode with the failure surface

emanating from the intersection of the column and slab soffit. The prestress was found

to have an influence on the load carrying capacity of the connections. The load

carrying capacity increased with the increase in the intensity of the prestress applied to

the slab. In all cases the failure was brittle.

In the edge column connections tested, the angle of the failure surface to the slab soffit

increased with increasing levels of prestress. This indicates that confinement of the

concrete across the failure surface has an effect on the angle of the failure surface. The

average failure angle was in the range of 17 to 26 degrees depending on the type of

connection and the intensity of the applied prestress. This is significantly less than the

45 degree failure angle assumed in the Australian and ACI code design models.

The test results were compared with the analytical shear to moment ratios calculated

from the,.equivalent frame and simplified code methods given in AS 3600 (1994) for

the load equal to the failure load measured in the tests. The shear force calculated from

these methods were close to the measured values but significant differences occurred

for the moments. Thus, the dimensional similitude model used to size the specimens

did not give the desired moment/shear ratio. This is possibly due to the zero vertical
9-5

displacement boundary condition used in the test at the second line of contraflexure.

Perhaps a better way of sizing of the laboratory specimens would be to use finite

element modelling to match the moment/shear ratios between the laboratory model

and the prototype structure.

Shear force and moment data obtained from a range of experimental tests were

compared with predictions from five international codes of practice. There was a

significant degree of scatter between the test data and the predicted strengths.

Statistical analyses of the experimental failure load to the code model predictions were

undertaken for a wide range of test data. The British code model (BS 811 0) gave the

best fit for the mean of the test data but with a large scatter of results and low

confidence level as to the probability of failure being greater than the design strength.

The Australian code model (AS 3600) showed a poor correlation with the measured

data but with a similar standard of deviation to that of other codes of practice, thus,

giving the highest confidence level that the slab capacity will exceed the design

strength.

In Chapter 8, the results of numerical analyses on the specimens were presented. The

finite element model was able to simulate the behaviour of the specimens reasonably

well. Further detailed investigations on the behaviour of these connections were

carried out using the numerical modelling. The finite element studies further showed

that the specimens had failed in a punching shear mode.

On verification of the finite element models, the models were modified to investigate

the effect of bonding of the prestress and to further study the influence of the prestress
9-6

on these connections. The connections with bonded prestressing reinforcement were

found to be marginally stronger than the unbonded connections. The finite element

model for the prestressed comer-column connections showed that prestressing in two

directions has a significant influence on the strength of the connection.

In summary, in this study the effect of prestress on the load carrying capacity of flat

plate-column connections was investigated. For both edge and comer column

connections the load carrying capacity of the connection increases approximately

linearly with increasing prestress. The influence of bonding the prestressing wires to

the slab was investigated using numerical modelling. The results showed that bonding

of the prestressing tendon gives a slight increase in the strength of the connection. The

relative increase in the capacity of the connection for the case of bonded tendons

compared to unbonded tendons increases with increasing levels of prestress.

9.3 Recommendations for Future Research

From the research undertaken in this thesis a number of areas for further studies have

been identified. These fall into three broad categories:

• numerical modelling

• experimental investigation, and

• development of design models.

Details of further research needed in each of these areas are given below.

9.3.1 Numerical Modelling

Numerical models of punching problems need to be made more realistic to the

physical problem and the following aspects should be considered:


9-7

• Taking the value of shear retention factor as a constant lacks physical


justification and this assumption does not appear realistic. That is, the shear
retention factor lacks fundamental physical definition. Studies are needed to
determine the mechanism of shear retention as a function of crack growth or
other fracture properties.

• The tension stiffening model needs rationalisation and the tension stiffening
and tension softening effects need to be separated. Modelling of tension
stiffening is essential in order to obtain the correct displacements in the slab but
it is not rational that elements along the punching failure surface have added
tension strength beyond cracking.

• Two 20-node brick elements through the thickness gave the satisfactory results
for studying punching shear strength in thin slabs. However, refinement of the
mesh through the thickness is needed if post failure behaviour is to be
investigated or for modelling thick slabs.

9.3.2 Experimental Investigations

fu relation to the experimental investigation of the slab-column connection the

following are noted for further study:

• The effect of prestress is significant in determining the capacity of slab-column


connections. However, in all tests undertaken in this study the ductility was
poor. Therefore, investigations on the methods of increasing the ductility of
prestressed connections may be considered of some importance. As for
non-prestressed concrete connections the provision of shear reinforcement in
the form of ties, shear studs or other mechanical means requires investigation.

• fu this study one moment shear ratio in the connection was tested. Further tests
are needed varying the stiffness of the column and slab to obtain data over the
full range of moment shear interaction possible. Linear finite element analyses
may be used to develop these specimens.
9-8

• There is a lack of experimental data on prestressed comer column connections.


Further research is needed on the topic.

9.3.3 Development of Design Models

Despite ninety years of research on the design of slab-column connections, no one

theory dominates the literature. The fact that design models vary from country to

country is perhaps the best evidence that the fundamentals of load transfer between

slabs and their supporting columns are yet to be properly understood. For the models

investigated in this study the scatter in the test data to predicted results is significant.

Further investigations are needed to obtain an improved understanding of the load

transfer mechanism in slab-column connections. Only then can a more realistic design

model be developed.
REFERENCES

Aalami, B., (1972). "Moment-rotation relation between column and slab", ACI
Journal, Vol. 69, No.5, May, pp. 263-269.

Abdel Rahman, H. H., (1982). "Computational models for the nonlinear analysis of
reinforced concrete flexural concrete slab systems", Doctoral Thesis, University
College of Swansea, Swansea, United Kingdom.

ACI 318 (1963). "Building code requirements for reinforced concrete (ACI 318-63)",
American Concrete Institute, Farmington Hills, 144 pp.

ACI 318 (1971). "Building code requirements for reinforced concrete (ACI 318-71)",
American Concrete Institute, Detroit.

ACI 318 (1977). "Building code requirements for reinforced concrete (ACI 318-77)",
American Concrete Institute, Farmington Hills, 103 pp.

ACI 318 (1983). "Building code requirements for reinforced concrete (ACI 318-83)",
American Concrete Institute, Detroit.

ACI 318 (1989). "Building code requirements for reinforced concrete and commentary
(ACI 318-89/318R-89)", American Concrete Institute, Detroit, 353 pp.

ACI 318 (1995). "Building code requirements for structural concrete (ACI 318M-95)
and commentary (ACI 318RM-95)", American Concrete Institute, Farmington Hills,
371 pp.

ACI 318 (1999). "Building code requirements for structural concrete (ACI 318M-99)
and commentary (ACI 318RM-99)", American Concrete Institute, Michigan, 391 pp.

ACI 349 (1990). "Code requirements for nuclear safety related concrete structures",
American Concrete Institute, Michigan, 130 pp.

ACI 408.2R (1992). "State-of-the-art report on bond under cyclic loads", American
Concrete Institute, Detroit, 32 pp.
R-2

ACI Committee 421, (1992). "Abstract of: shear reinforcement for slabs", ACI
Structural Journal, Vol. 89, S56, September-October, pp. 587-589.

ACI Committee 435, (1991). "State of art report on control of two-way slab
deflections", ACI Structural Journal, Vol. 88, S53, July-August, pp. 501-514.

ACI-ASCE Committee 326, (1962). "Shear and diagonal tension: Part 3 -slabs and
footings", Journal of American Concrete Institute, Vol. 59, No.3, March, pp. 352-396.

ACI-ASCE Committee 352, (1989). "ACI352.1R-89: Recommendations for design of


slab-column connections in monolithic reinforced concrete structures", American
Concrete Institute, Detroit, Michigan, 22 pp.

ACI-ASCE Committee 423 (1974). "Tentative recommendations for prestressed


concrete flat plates", Journal of American Concrete Institute, Vol. 71, No.2, February,
pp. 61-71.

Alexander, S. B., (1990). "Bond model for strength of slab-column joints", Doctoral
Thesis, University of Alberta, Canada, 228 pp.

Alexander, S. D. B., and Simmonds, S. H., (1986). "Shear-moment transfer in


slab-column connections", Structural Engineering Report No. 141, Department of
Civil Engineering, University of Alberta, Edmonton, Alberta, July, 94 pp.

Alexander, S. D. B., and Simmonds, S. H., (1992). "Bond model for concentric
punching shear", ACI Structural Journal, Vol. 89, No.3, May-June, pp. 325-334.

Alexander, S. D. B., and Simmonds, S. H., (1992). "Tests of column-flat plate


connections", ACI Structural Journal, Vol. 89, No. 5, September-October, 1992,
pp. 495-502.

Alexander, S., (1999). "Strip design for punching shear", The Design of Two-Way
Slabs, ACI SP-183, pp. 161-179.
R-3

Al-Manaseer, A. A., and Philips, D. V., (1987). "Numerical study of some


post-cracking material parameters affecting nonlinear solutions in RC deep beams",
Canadian Journal of Civil Engineering, Vol. 14, pp. 655-666.

Andersson, J. L., (1963). "Punching of concrete slabs with shear reinforcement",


Meddelande Nr. 47, Institutionen for Byggnadsstatik, Kungliga Tekniska Hogskolan,
Stockholm, Switzerland.

AS 1012.10 (1985). "Method for the determination of the indirect tensile strength of
concrete cylinders ('Brazil' or Splitting test)'', Standards Australia, Sydney, Australia.

AS 1012.9 (1986). "Method for the determination of the compressive strength of


concrete specimens", Standards Australia, Sydney, Australia.

AS 1472 (1991). "Carbon steel spring wire for mechanical springs", Standards
Australia, Sydney, Australia, 9 pp.

AS 3600 (1994). "Concrete structures", Standards Australia, Sydney, Australia,


154 pp.

AS 3600 Supplement 1 (1994). "Concrete structures-commentary", Standards


Australia, Sydney, Australia, 165 pp.

ASCE-ACI Committee 426, (1974) "The shear strength of reinforced concrete


members- slabs", Journal of Structural Division, ASCE, Vol. 100, No. ST8, August,
pp. 1543-1591.

Barzegar, F., (1989). "Analysis of RC membrane elements with anisotropic


reinforcement", Journal of Structural Engineering, Vol. 115, No. 3, March,
pp. 647-665.

Barzegar, F., and Schnobrich, W. C., (1990). "Post-cracking analysis of reinforced


concrete panels including tension stiffening", Canadian Journal of Civil Engineering,
Vol. 17, pp. 311-320.
R-4

Barzegar, F., Behle, R., and Foroozesh, M., (1991). "Moment transfer and slab
effective widths in laterally loaded edge connections." ACI Structural Journal,
Vol. 88, No.5, September-October, pp. 615-623.

Bazant, Z. P., and Oh, B. P., (1983). "Spacing of cracks in reinforced concrete",
Journal of Structural Engineering, ASCE, Vol.109, No.9, September, pp. 2066-2085.

Bazant, Z. P., and Oh, B. P., (1984). "Numerical simulation of progressive fracture in
concrete structures: Recent developments", Proceedings, International Conference on
Computer Aided Analysis and Design of Concrete Structures, Pineridge Press,
Swansea, 1984, pp. 1-18.

Broms, C. E., (1990), "Punching of flat plates- a question of concrete properties in


biaxial compression and size effect," ACI Structural Journal, V. 87, No.3, May-June,
pp. 292-304.

Broms, C. E., (1990), "Shear reinforcement for deflection ductility of flat plates," ACI
Structural Journal, V. 87, No.6, November-December, pp. 696-705.

Brown, S., and Dilger, W. H., (1994). "Seismic response of flat plate column
connection", Annual Conference of the Canadian Society for Civil Engineering,
Winnipeg, June, pp. 388-397.

BS 8110 (1997). "Code of practice for design and construction (BS 8110: Part D",
British Standard Institution, London, 126 pp.

Bums, N. H., and Hemakom, R., (1977). "Test of a scale model post-tensioned flat
plate", Journal of Structural Division, ASCE, Vol. 103, No. ST6, June, pp. 1237-1255.

CAN3-A23.3-M84, (1984). "Design of concrete structures for buildings", Canadian


Standards Association, Rexdale, 281 pp.

CEB-FIP Model Code for Concrete Structures 1990 (1993). Comite Euro-Intemational
du Beton/Federation International de la Precontrainte, Lausanne, Switzerland.
R-5

Chana, P., and Desai. S., (1993). "Design fro provision of resistance against punching
shear", Proceedings: Concrete 2000, Dundee University, September, Vol. 1,
pp. 815-825.

Channakeshava, C., and Iyengar, K. T. S. R., (1988). "Elasto-Plastic cracking analysis


of reinforced concrete", Journal of Structural Engineering, Vol. 114, No. 11,
November, pp. 2421-2438.

Chen, W.F., and Han, D. J., (1988). "Plasticity for structural engineers",
Springer-Verlag, New York, 606 pp.

Chiang, Y. J., and Lee, Y. L., (1994). "Evaluation of modeling accuracy of 20-node
solid elements by statistical factorial design", Computers & Structures, Vol. 52, No. 6,
pp. 1309- 1314.

Chiang, Y. J., and Tang, C., (1995). "Accuracy assessment to applying 20-node solid
elements to pressurized composite shells", Finite Elements in Analysis and Design,
Vol. 20, pp. 219-231.

Chow, H, and Selna, L. G., (1995), "Seismic response of non-ductile flat-plate


buildings", Journal of Structural Engineering, Vol. 121. No. 1, January, pp. 115-123.

Collins, P. M., and Mitchell, D., (1991). "Prestressed concrete structures",


Prentice-Hall, Inc., New Jersy, 1991, 766 pp.

Cooke, N., Park, R., and Young, P., (1981). "Flexural strength of prestressed concrete
member with unbonded tendons," Journal of the Prestressed Concrete Institute,
Vol. 26, No.6, November-December, pp. 52-80.

Cope, R. J., (1986). "Nonlinear analyses of reinforced concrete slabs", Computer


Modelling of RC structures, Pineridge Press, Swansea, pp. 3-43.

CSA 23.3 (1994). "Design of Concrete Structures", Canadian Standards Association,


Rexdale, Canada, 220 pp.
R-6

de Borst, R., and Nauta, P. (1985). "Non-orthogonal cracks in a smeared finite element
model", Engineering Computations, March, Vol. 2, pp. 35-46.

DeVries, R. A., Jirsa, J. 0., and Bashandy, T., (1999). "Anchorage capacity in
concrete of headed reinforcement with shallow embedments", ACI Structural Journal,
Vol. 96, No.5, September-October, pp. 728-736.

Dechka, D. C., and Dilger, W. H., (2000). "Full scale continuous slab-column frame
subjected to reverse cyclic loading", American Concrete Institute, 2000 Spring
Convention, San Diego, USA, March.

Dechka, D. C., Loov, R. E., and Dilger, W.H., (2000). "Prediction of punching shear
capacity by shear friction", American Concrete Institute, 2000 Spring Convention, San
Diego, USA, March.

DiStasio, J., and Van Buren, M.P., (1960). "Transfer of bending moment between flat
plate floor and column", Journal of the American Concrete Institute, Vol. 57, No. 3,
September, pp. 299-314.

DIANA-Release 6.2 (1997). TNO Building and Construction Research, Delft, The
Netherlands.

Dilger, W., and Cao, H., (1991). "Behaviour of slab-column connections under
reversed cyclic loading", Proceedings of the 2-nd International Conference of
High-Rise Buildings, China.

Dragosvic, AM. and van den Beukel, A., (1974), "Punching shear", HERON, vol. 20,
No. 2,48 pp.

Durrani, A. J., Du, Y., and Luo, Y. H., (1995). "Seismic resistance of nonductile
slab-column connections in existing flat slab buildings", ACI Structural Journal,
Vol. 92, No.4, July-August, pp. 479-487.
R-7

Elgabry, A. A., (1991). "Shear and moment transfer of concrete flat plates", Doctoral
Thesis, Department of Civil Engineering, The University of Calgary, Alberta, Canada,
February, 226 pp.

Elgabry, A. A., and Ghali, A., (1987). ''Tests on concrete slab-column connections
with stud-shear reinforcement subjected to shear-moment transfer", ACI Structural
Journal, Vol. 84, S46, September-October, pp.433-442.

Elgabry, A. A., and Ghali, A., (1990). "Design of stud-shear reinforcement for slabs",
ACI Structural Journal, Vol. 87, S35, May-June, pp. 350-361.

Elgabry, A. A., and Ghali, A., (1996). "Moment transfer by shear in slab-column
connections", ACI Structural Journal, Vol. 93, ST 2, March-April, pp. 187-196.

Elgabry, A. A., and Ghali, A., (1996). "Transfer of moments between columns and
slabs: Proposed code revisions", ACI Structural Journal, Vol. 93, No. 1,
January-February, pp. 56-61.

El-Salakawy, E. F., Polak, M. A., and Soliman, M. H., (1998). "Slab-column edge
connections subjected to high moments", Canadian Journal of Civil Engineering,
Vol. 25, pp. 526-538.

El-Salakawy, E. F., Polak, M. A., and Soliman, M. H., (1999). "Reinforced concrete
Slab-column edge connections with openings", ACI Structural Journal,
January-February, Vol. 96, No. 1, pp. 79-87.

Elstner, R. C., and Hognestad, E., (1956). "Shearing strength of reinforced concrete
slabs", Journal of American Concrete Institute, Vol. 28, No. 1, July, pp. 29-58.

Eurocode 2 (1992). "Design of concrete structures- Part 1: General rules and rules for
buildings", European Committee for Standardization, Brussels, 274 pp.

Falamaki, M., and Loo, Y. C., (1992). "Punching shear tests of half-scale reinforced
concrete flat-plate models with spandrel beams", ACI Structural Journal, Vol. 89, No.
3, May-June, pp. 263-271.
R-8

Farhey, D. N., Adin, M. A., Yankelevsky, D. Z., (1995). "Repaired RC


flat-slab-column subassemblages under lateral loading", Journal of Structural
Engineering, Vol. 121, No. 11, November, pp. 1710-1720.

Foster, S. J., (1992). "The structural behaviour of reinforced concrete deep beams",
Doctoral Thesis, School of Civil Engineering, The University of New South Wales,
Sydney, Australia.

Foutch D. A., Gamble, W. L., and Sunidja, H., (1990). "Tests of post-tensioned
concrete slab-edge column connections", ACI Structural Journal, Vol. 87, No. 2,
March-April, pp. 167-179.

Furst, A., and Marti, P., (1997). "Robert Maillart's design approach for flat slabs",
ASCE Journal of Structural Engineering, Vol. 123, No.8, August, pp. 1102-1110.

Gardner, N. J., and Shao, X., (1996). "Punching shear of continuous flat reinforced
concrete slabs", ACI Structural Journal, Vol. 93, No.2, March-April, pp. 218-228.

Gerber, L. L., and Burns N. D., (1971). "Ultimate strength tests of post-tensioned flat
plates", PCI Journal, Vol. 16, No.6, November-December, pp. 40-58.

Gesund, H., and Kaushik, Y. P., (1970). "Yield line analysis of punching failures in
slabs" International Association for Bridge and Structural Engineering, Zurich,
Vol. 30-1, pp. 41-60.

Gilbert, R. 1., (1979). "Time dependent behaviour of structural concrete slabs",


Doctoral Thesis, School of Civil Engineering, The University of New South Wales,
361 pp.

Gilbert, R. 1., and Warner, R. F., (1978). ''Tension stiffening in reinforced concrete
slabs", ASCE Journal of Structural Division, Vol. 104, ST12, December,
pp. 1885-1900.
R-9

Gomes, R. B., and Regan, P.E., (1999). "Punching resistance of RC flat slabs with
shear reinforcement", ASCE Journal of Structural Engineering, June, Vol. 125, No. 6,
pp. 684-692.

Gomes, R., and Regan, P., (1999). "Punching strength of slab reinforced for shear with
offcuts of rolled steel !-section beams", Magazine of Concrete Research, April,
Vol. 51, No.2, pp. 121-129.

Gosselin, D., (1984). "The behaviour of reinforced concrete slab-column structures


with drop panels subjected to gravity and lateral loading", Department of Civil
Engineering, Royal Military College of Canada, Canada, 184 pp.

Guan, H., and Loo, Y. C., (1994). "Layered finite element method of analysis of
reinforced concrete flat plates", International Conference on Computational Methods
in Structural and Geotechnical Engineering, Hong Kong, December 12-14,
pp. 984-989.

Guan, H., and Loo, Y. C., (1995). "On the improvement of layer-type finite element
model", Proceeding Fifth East Asia-Pacific Conference on Structural Engineering and
Construction, Gold Coast, Australia, pp. 73-78.

Guan, H., and Loo, Y. C., (1996). "Numerical analysis and computer visualization of
punching shear failure of reinforced concrete flat plates: Part I Methodology", Third
Asian-Pacific Conference on Computational Mechanics, Seoul Korea, pp. 259-264.

Guan, H., and Loo, Y. C., (1996). "Numerical analysis and computer visualization of
punching shear failure of reinforced concrete flat plates: Part II Applications", Third
Asian-Pacific Conference on Computational Mechanics, Seoul Korea, pp. 265-270.

Hall, A. S., and Rangan, B. V., (1983). "Forces in the vicinity of edge columns in flat
slab floors", Magazine of Concrete Research (London), Vol. 35, No. 122, March,
pp. 19-26.
R-10

Hammill, N., and Ghali, A., (1994). "Punching shear resistance of comer slab-column
connections", ACI Structural Journal, Vol. 91, No. 6, November-December,
pp 697-707.

Hanson, N. W., and Hanson, J. M., (1968) "Shear and moment transfer between
concrete slabs and columns", Journal of the Portland Cement Association, Research
and Development Laboratories, Vol. 10, No. 1, January, pp. 2-16.

Hawkins, N. M., (1981). "Lateral load resistance of unbonded post-tensioned flat plate
construction", PCIJoumal, Vol. 26, No.1, January-February, pp. 94-116.

Hawkins, N. M., and Coreley, W. G., (1971). "Transfer of unbalanced moment and
shear from flat plate to column", Cracking, Deflection, and Ultimate Load of Concrete
Slab Systems, ACI Special Publication, SP- 30, American Concrete Institute, Detroit,
1971, pp. 147-176.

Hawkins, N. M., Bao, A. and Yamazaki, J., (1989). "Moment transfer from concrete
slab to columns", ACI Structural Journal, November-December, Vol. 86, No. 6,
pp. 705-716.

Hawkins, N. M., Bao, A., and Yamazaki, J., (1989). "Moment transfer from concrete
slabs to columns", ACI Structural Journal, Vol. 86, S70, November-December,
pp. 705-716.

Hellier, A.K., Sansalone, M., Carino, N.J., Stone, W. C., and Ingraffea, A. R., (1987).
"Finite-Element analysis of the pullout test using a nonlinear discrete cracking
approach", Cement, Concrete and Aggregates, American Society of Testing and
Materials, Vol. 9, No. 1, pp. 20-29.

Hemakom, R., (1975). "Strength and behavior of post-tensioned flat plates with
unbonded tendons", Doctoral Thesis, The University of Texas, Austin, December,
pp. 272.
R-11

Hognestad, E., (1951). "A study of combined bending and axial load in reinforced
concrete members", University of lllinois Engineering Experimental Station, Bulletin
Series No. 399, November, 128 pp.

Hu, H. H., and Schnobrich, W. C., (1988). "Nonlinear analyses of plane stress state
reinforced under short time monotonic loading", Civil Engineering Studies,
SRS No. 539, University of lllinois, Urbana, April.

Hueste, M. B. D., and Wight, J. K., (1999). "Modelling punching shear failures in
static and dynamic nonlinear analyses using a macro model", American Concrete
Institute, 2000 Spring Convention, San Diego, USA, March.

Hueste, M. B. D., and Wight, J. K., (1999). "Nonlinear punching shear failure model
for interior slab-column connections", ASCE Journal of Structural Engineering,
September, Vol. 125, No.9, pp. 997-1008.

Ichihashi, 1., Shibata, H., and Nomura, T., (1985). "Some test results of the strength of
embedded plates and or bolts in concrete", Transaction of the International Conference
on Structural Mechanics in Reactor Technology, Amsterdam, pp. 453-458.

Islam, S., and Park, R., (1976). "Tests on slab-column connections with shear and
unbalanced flexure", Journal of Structural Division, ASCE, Vol. 102, No. ST3, March,
pp. 549-568.

Jiang, D. H., and Zhou, K. R, (1991). "Imitation of punching failure for reinforced
connect slabs in finite elements method", Proceeding of Asian Pacific Computational
Conference on Mechanics, December, 221-226 pp.

Johansen, K. W., (1962). "Yield-Line theory", Cement and Concrete Association,


London, 181 pp.

Kanoh, Y., and Yoshizaki, S., (1979). "Strength of slab-column connections


transferring shear and moment", ACI Journal, Vol. 76, No.3, March, pp. 461-478.
R-12

Khwaounjoo, Y. R., Foster, S. J., and Gilbert, R. I., (1999). "Influence of boundary
conditions on punching shear behaviour of flat plate-column connections", 16th
Australian Conference on the Mechanics of Structures and Material, ACMSM16,
Sydney, December, pp. 145-150.

Kinnunen, S., (1963). "Punching of concrete slabs with two-way reinforcement",


Nr. 198, Transactions of The Royal Institute of Technology, Stockholm, Sweden.
107pp.

Kinnunen, S., and Nylander, H., (1960). "Punching of concrete slabs without shear
reinforcement", Nr. 158, Institutionen for Byggnadsstatik, Kungliga Tekniska
Hogskolans, Stockholm, 112 pp.

Klinger, R. E., and Mendonca, J. A., (1982). "Tensile capacity of short anchor bolts
and welded studs: a literature review", ACI Journal, Vol. 79, No. 27, July-August,
pp. 270-279.

Kuang, J. S., and Morley, C. T., (1992). "Punching shear behavior of restrained
reinforced concrete slabs", ACI Structural Journal, Vol. 89, No. 1, January-February,
pp. 13-19.

Lamb, J. W., (1984). "Moment transfer and joint stiffness in reinforced concrete flat
plate-column connections", Department of Civil Engineering, Royal Military College
of Canada, Canada, 184 pp.

Lew, H. S., Carino, N. H., Fattal, S. G., and Batts, M. E., (1982). "Investigation of
construction failure of Harbour Cay Condominium in Cocoa Beach, Florida", Building
Science Series No. 145, National Bureau of Standards, Washington, D. C., August,
135 pp.

Lim, F. K., and Rangan, B. V., (1994). "Strength of concrete slabs with stud shear
reinforcement in the vicinity of edge corner columns", Research Report No. 1/94,
School of Civil Engineering, Curtin University of Technology, April, 180 pp.
R-13

Lim, F. K., and Rangan, B. V., (1995). "Studies on concrete slabs with stud shear
reinforcement in vicinity of edge and comer columns", ACI Structural Journal,
Vol. 92, No.5, September-October, pp. 515-525.

Lin, C. S., (1973). "Nonlinear analysis of reinforced concrete slabs and shells",
Doctoral Thesis, UC_SESM 73-7, University of California at Berkeley.

Lin, C., and Scordelis, A. C., (1975). "Nonlinear analysis of RC shells of general
form", Proceedings ASCE, Vol. 101, ST3, March, pp. 523-538.

Lin, T. Y., and Bums, N.H., (1981). "Design of prestressed concrete structures", John
Wiley & Sons, 646 pp.

Long, A. E., "Punching failure of reinforced concrete slabs", Doctoral Thesis, Faculty
of Applied Science and Technology, Queen's University of Belfast, May 1967,
210pp.

Loo, Y. C., and Falamaki, M., (1992). "Punching shear strength of reinforced concrete
flat plates with spandrel beams", ACI Structural Journal, Vol. 89, S36, July-August,
pp. 375-383.

Loo, Y. C., and Guan, H., (1997). "Cracking and punching shear failure analysis of
RC flat plates", Journal of Structural Engineering, Vol. 123, No. 10, October,
pp. 1321-1330.

Loo, Y. C., Chiang, C. L., and Guan, H., (1995). "Performance of prediction method
for punching shear strength of post-tensioned concrete flat plates", 14th Australian
Conference on the Mechanics of Structures and Materials, Hobart, Australia,
pp. 130-135.

Luo, Y. H., and Durrani, A. J., (1995). "Equivalent beam model for flat-slab buildings-
part ii: exterior connections", ACI Structural Journal, Vol. 92, No. 2, March-April,
pp. 250-257.
R-14

Luo, Y. H., Durrani, A. J., Bai, S., and Yuan, J., (1994). "Study of reinforced detail of
tension bars in frame comer connections," ACI Structural Journal, Vol. 91, No. 48,
pp. 486-496.

Luo, Y. H., Durrani, A., and Conte, J., (1995). "Seismic reliability assessment of
existing r/c flat-slab buildings", Journal of Structural Engineering, Vol. 121, No. 10,
October, pp. 1522-1530.

Malhotra, V. M, (1975). "Evaluation of the pull-out test to determine strength of


in-situ concrete", Materials and Constructions, Vol. 8, No. 43, January-February,
pp. 19-31.

Mansur, M. A., and Rangan, B. V., (1978). "Torsion in spandrel beams", Journal of
Structural Division, ASCE, Vol. 104, ST7, pp. 1061-1075.

Martinez-Cruzado, J. A., Qaisrani, Allah-N, and Moehle, J.P., (1994)."Post-tensioned


flat plate slab-column connections subjected to earthquake loading", Proceedings:
Earthquake awareness and mitigation across the nation/fifth US National Conference
on Earthquake Engineering, July 10-14, Chicago, lllinois, pp. 139-148.

Mast, P. E., (1970). "Plate stresses at columns near the free edge", ACI Journal,
Vol. 67, No. 11, November, pp. 898-902.

Materson, D. M., and Long, A. E., (1974). "The punching strength of slabs, a flexural
approach using finite elements", Shear in Reinforced Concrete, ACI Special
Publication, SP-42, American Concrete Institute, Detroit, pp. 747-768.

Megally, S., and Ghali, A., (1994). "Design considerations for slab-column
connections in seismic zones", ACI Structural Journal, Vol. 91, No. 3, May-June
1994, pp. 303-314.

Megally, S., and Ghali, A., (1996). "Nonlinear analysis of moment transfer between
columns and slabs", 1st Structural Specialty Conference, May 29-June 1, Edmonton,
Alberta, Canada, pp. 321-332.
R-15

Megally, S. and Ghali, A., (1999). "Design for punching shear in concrete", The
Design of Two-Way Slabs, ACI SP-183, pp.37-66.

Megally, S., and Ghali, A., (2000). "Punching of concrete slabs due to column
moment transfer", ASCE Journal of Structural Engineering, February, Vol. 126, No.2,
pp. 180-189.

Megally, S., and Ghali, A., (2000). "Avoiding punching failure of slabs in
earthquakes", American Concrete Institute, 2000 Spring Convention, San Diego, USA,
March.

Mehrain, M., and Aalami, B., (1974). "Rotational stiffness of concrete slabs", ACI
Journal, Vol. 71, September, pp. 429-435.

Moe, J., (1961). "Shearing strength of reinforced concrete slabs and footings under
concentrated loads", Bulletin No. D47, Portland Cement Association, Research and
Development Laboratories, Skokie, lllinois, April, pp. 1-130.

Menetrey P., (1998). "Relationship between flexural and punching failure", ACI
Structural Journal, July-August, Vol. 95, No.4, pp. 413-419.

Mortin, J.D., and Ghali, A., (1991) "Connection of flat plates to edge columns," ACI
Structural Journal, Vol. 88, No.2, March-April, pp. 191-198.

Nielsen, M. P., Braestrup, M. W., Jensen, B. C., and Bach, F., (1978). "Concrete
plasticity", Specialpublikation udgivet af Dansk Selskab, October, 129 pp.

Noor, A. K., and Babu.ska, 1., (1987). "Quality assessment and control of finite
element solutions", Finite Elements in Analysis and Design 3, pp. 1-26.

Ockleston, A. J., (1955). "Load test on a three storey reinforced concrete building in
Johannesburg", Structural Engineer, London, Vol. 33, No. 10, October, pp. 304-322.
R-16

Onsongo, W. M., and Collins, M. P., (1972). "Longitudinally restrained beams in


torsion" Publication No. 72-07, Department of Civil Engineering, University of
Toronto, May, 35 pp.

Ottosen, N. S., (1977). "A failure criterion for concrete," Journal of Engineering
Mechanics Division, ASCE, Vol. 103, No.4, pp. 527-535.

Ottosen, N. S., (1979). "Constitutive model for short-time loading of concrete",


Journal of Engineering Mechanics Division, ASCE, Vol. 105, No. 1, pp. 127-141.

Ozbolt, J., Vocke, H., and Eligehausen, R., (2000). ''Three-dimensional numerical
analysis of punching failure", American Concrete Institute, 2000 Spring Convention,
San Diego, USA, March.

Pan, A. D., and Moehle, J. P., (1992). "Experimental study of slab-column


connections", ACI Structural Journal, Vol. 89, No. 6, November-December,
pp. 626-638.

Pannell, F. N., (1969). "The ultimate moment of resistance of unbonded prestressed


concrete beams", Magazine of Concrete Research, Vol. 21, No. 66, March, pp. 43-54.

Park, H., (1999). "Numerical study on RC flat plates subjected to combined axial and
transverse load", Structural Engineering and Mechanics, Vol. 8, No.2, pp.137-150.

Park, R., and Gamble, W. L., (2000). "Reinforced concrete slabs", John Wiley &
Sons, Inc., New York, 716 pp.

Park, R., and Islam, S., (1976). "Strength of slab-column connections with shear and
unbalanced flexure", Journal of Structural Division, ASCE, Vol. 102, No. ST9,
September, pp. 1879-1991.

Park, R., and Islam, S., (1976). "Strength of slab-column connections with shear and
unbalanced flexure", Journal of Structural Diyision, ASCE, Vol. 102, No. ST9,
September, pp. 1879-1901.
R-17

Pavlovic, M. N., and Poulton, S. M., (1985). "On the computation of slab effective
widths", ASCE, Journal of Structure Division, Vol. 111, No.2, February, pp. 363-377.

Pecknold, D. V., (1975). "Slab effective width for equivalent frame analysis", ACI
Journal, Vol. 72, No.4, April, pp. 135-137.

Petersson, P. E., and Gustavasson, P. J., (1980). "Numerical methods in fracture


mechanics", Pineridge Press, Swansea, pp. 707-719.

Polak, M. A., (1998). "Modelling punching shear of reinforced concrete slabs using
layered finite elements", ACI Structural Journal, January-February, Vol. 95, No. 1,
pp. 71-80.

Prak:hya, G. K. V., and Morley, C. T., (1990). "Tension stiffening and moment-
curvature relation of reinforced concrete elements", ACI Structural Journal, Vol. 87,
No.5, September-October, pp. 597-605.

Rangan, B. V., (1987). "Punching shear strength of reinforced concrete slabs", Civil
Engineering Transaction Institution of Engineers, Australia, Civil Engineering,
Vol. CE29, No.2, April, pp. 71-78.

Rangan, B. V., (1990). "Punching shear design in the new Australian standard for
concrete structures", ACI Structural Journal, Vol. 87, S15, March-April, pp. 140-166.

Rangan, B. V., (1990). "Tests on slabs in the vicinity of edge columns", ACI
Structural Journal, Vol. 87, No.6, November- December, pp. 623-629.

Rangan, B. V., and Hall, A. S., (1983). "Forces in the vicinity of edge columns in flat
plate floors (VI-tests on R. C. models)", UNCIV Report No. R-203, The University of
New South Wales, Sydney, Australia, January, 240 pp.

Rangan, B. V., and Hall, A. S., (1983). "Moment and shear transfer between slab and
edge column", ACI Journal, May-June, Vol. 80, No.3, pp. 183-191.

Rangan, B. V., and Lim, F. K., (1992). "Test on concrete slabs with stud shear
reinforcement in the vicinity of edge columns", Research Report No. 1192, School of
R-18

Civil Engineering, Curtin University of Technology, Perth, Perth Australia, May,


140 pp.

Rankin, G. I. B., and Long, A. E., (1988). "A rational approach to punching at interior
slab-Column Connections", Proceedings of American Concrete Institute Annual
Convention, Florida, March, pp. 45-55.

Regan, P. E., (1978). "Design of reinforced concrete slabs", CIRIA Project


Record 220,0ctober, London.

Regan, P. E., (1981). "Behaviour of reinforced concrete flat slabs", CIRIA Report 89,
February, London, 89 pp.

Regan, P. E., Walker, P. R. and Zakaria, K. A. A., (1979). "Tests of reinforced


concrete flat slabs", CIRIA Project No. RP 220, Polytechnic of Central London,
United Kingdom, 217 pp.

Robertson, I. N., (1997). "Analysis of flat-plate structures subjected to combined


lateral and gravity loads", ACI Structural Journal, November-December, Vol. 94,
No.6, pp. 723-729.

Robertson, I. N., and Durrani, A. J., (1991). "Gravity load effect on seismic behavior
of exterior slab-column connections", ACI Structural Journal, Vol. 88, S28, May-June,
pp. 255-267.

Robertson, I. N., and Durrani, A. J., (1992). "Gravity load effect on seismic behavior
of interior slab-column connections", ACI Structural Journal, Vol. 89, No. 1,
January-February, pp. 37-45.

Roddeman, D. G., and Jansen L. F., (1993). "An a priori geometry check for a single
isoparametric finite element", Computers & Structures, Vol. 47, No. 1, pp. 69-72.

Rots, J. G., (1985). "Bond-Slip simulation using smeared cracks and/or interface
element", Delft University of Technology, Netherlands, November, 56 pp.
R-19

Rots, J. G., (1988). "Computational modeling of concrete fracture", Doctoral Thesis,


Delft University of Technology, The Netherlands, 134 pp.

Rots, J. G., Nauta, P., and Kusters, G. M.A., and Blaauwendraad, J. (1985) "Smeared
crack approach and fracture localization in concrete", HERON, Vol. 30, No. 1, 48 pp.

Sbarounis, J. A., (1984). "Multistory flat plate buildings", Concrete International:


Design and Construction, Vol. 6, No.2, February, pp. 70-77.

Scanlon, A., (1971). "Time dependent deflection of reinforced concrete slabs",


Doctoral Thesis, University of Alberta, Edmonton, 174 pp.

Schnobrich, W. C., (1977). "Behavior of reinforced concrete structures predicted by


the finite element method", Computers & Structures, Vol. 7, pp. 365-376.

Scordelis, A. C., Lin, T. Y., and Itaya, R., (1959). "Behaviour of a continuous slab
prestressed in two directions", Journal of the American Concrete Institute, Vol. 31,
No.6, September, pp. 441-459.

Scordelis, A. C., Lin, T. Y., and May, H. R., (1958). "Shearing strength of prestressed
concrete lift slabs", Journal of the American Concrete Institute, Vol. 30, No. 4,
October, pp. 458-506.

Scordelis, A. C., Pister, K. S., and Lin, T. Y., (1956). "Strength of a concrete slab
prestressed in two directions", Journal of the American Concrete Institute, Vol. 28,
No.3, September, pp. 241-256.

Shehata, I. A. E. M., (1985). "Theory of punching in concrete slabs", Doctoral Thesis,


Polytechnic of Central London, London.

Shehata, I. A. E. M., and Regan, P. E., (1989). "Punching in R. C. slabs", Journal of


Structural Engineering., ASCE, 115, No.7, July, pp. 1726-1740.

Sih, G. C., and DiTommaso, A., Eds. (1985). "Fracture mechanics of concrete:
Structural application and numerical calculation", Martinus Nijhoff Publishers,
Dordrecht, 276 pp.
R-20

Simmonds, S. H., and Alexander, D. B.S., (1987). "Truss model for edge column-slab
connections", ACI Structural Journal, Vol. 84, July-August 1987, pp. 296-302.

Smith, S. W., and Bums N. D., (1974). "Post-tensioned flat plate to column
connection behavior", PCI Journal, Vol. 19, No.3, May-June, pp. 74-91.

Stamenkovic, A., and Chapman, J. C., (1974). "Local strength at column heads in flat
slabs subjected to a combined vertical and horizontal loadings", Proceedings,
Institution of Civil Engineers, London, Part 2, June, pp. 205-232.

Stone, W. C., and Carino, N. J., (1983). "Deformation and failure in large-scale
pullout tests", ACI Journal, Vol. 80, No. 46, November-December, pp. 501-513.

Stone, W. C., and Carino, N. J., (1984)."Comparison of analytical with experimental


internal strain distribution for the pullout test", ACI Journal, Vol. 81, No. 1,
January-February, pp. 3-12.

Suidan, M., and Schnobrich, W. C., (1973). "Finite element analysis of reinforced
concrete", ASCE Proceedings, Journal of Structural division, Vol. 99, STlO, October,
pp. 2109- 2121.

Sunidja, H., Foutch, D. A., and Gamble, W. L., (1982). "Response of prestressed
concrete plate-edge column connections", Civil Engineering Studies, Structural
Research Series No. 498, University of Dlinois, Urbana, and March, 232 pp.

Swamy, R.N., and Andriopoulos, A. D., (1974). "Contribution of aggregate interlock


and dowel forces to the shear resistance of reinforced beam with web reinforcement",
SP 42.6 ACI Publication on Shear in Reinforced Concrete, Detroit, pp. 129-166.

Timoshenko, S. P., and Woinowsky-Krieger, S., (1959). "Theory of plates and shells",
McGraw-Hill International Book Company, New York, 580 pp.

Trongtham, N., and Hawkins, N. M., (1977). "Moment transfer to columns in


unbonded post-tensioned prestressed concrete slabs", Report SM 77-3, Department of
Civil Engineering, University of Washington, Seattle, Washington, October.
R-21

van den Beukel, A., (1976). "Punching shear at inner, edge, and comer columns",
HERON, vol. 21, No.3, 30 pp.

van Mier, J. G. M. (1987). "Examples of non-linear analysis of reinforced concrete


structures with DIANA", HERON, Volume 32, No.3, 147 pp.

Vecchio, F. J., and Collins, M.P., (1990). "Investigating the collapse of a warehouse",
Concrete International: Design and Construction, Vol. 12, No. 3, March 1990,
pp. 72-78.

Vermeer, P. A., and de Borst, R., (1984). "Non-associated plasticity for soils, concrete
and rock", HERON, Vol. 29, No.3, 65 pp.

Walker, P.R., and Regan, P. E., (1987). "Comer column-slab connections in concrete
flat plates", ASCE Journal of Structural Engineering, April, Vol. 113, No. 4,
pp. 704-720.

Wong, Y. C. and Coull, A., (1980). "Effective slab stiffness in flat plate structures",
Proceedings of Institution of Civil Engineers, Part 2, 69, London, September,
pp. 721-735.

Yamada, T., Nanni, A., and Endo, K., (1992). "Punching shear resistance of flat slabs:
influence of reinforcement type and ratio", ACI Structural Journal, Vol. 88, No. 4,
September-October, pp. 555-563.

Yener, M., (1994). "Overview and progressive finite element analysis of pullout tests",
ACI Journal, Vol. 91, No. S6, January-February, pp. 49-58.

Yener, M., and Li, G. C., (1991). "Progressive finite element fracture analysis of
pullout concrete", Journal of Structural Engineering, Vol. 117, No. 8, Auglist,
pp. 2351-2371.

Zaghlool, E. R. F., (1971). "Strength and behaviour of comer and edge column-slab
connections in reinforced concrete flat plates", Doctoral Thesis, Department of Civil
Engineering. University of Calgary, Alberta, 366 pp.
R-22

Zaghlool, E. R. F., and de Paiva, H. A. R., (1973). "Tests of flat-plate corner


column-slab connections", Proceedings ASCE, Vol. 99, ST3, March, pp. 551-572.

Zaghlool, R. F., de Paiva, H. A. R., and Glockner, P. G., (1970). "Tests of reinforced
concrete flat plate floors", Proceedings of American Society of Civil Engineers,
Journal of Structural Division, March, pp. 487-507.

Zerbe, H. E., and Durrani, A. H., (1990). "Seismic response of connections in two-bay
reinforced concrete frame subassemblies with a floor slab", ACI Structural Journal,
Vol. 87, S40, July-August, pp. 406-415.

Zerbe, H. E., and Durrani, A. J., (1990). "Seismic response of connections in


indeterminate r/c frame subassemblies", Report No. 40, Department of Civil
Engineering, Rice University, Houston, Texas, April, 216 pp.

Zhou, K. R., and Jiang, D. H., (1991). "Imitation of punching failure for reinforced
connect slabs in finite elements method", Proceeding of Asian Pacific Computational
Conference on Mechanics, December, pp. 221-226.
APPENDIX A- DESIGN OF PROTOTYPE AND TEST SLABS

A.l Structural Systems

The flat plate-column structural system shown in Figure A.l was considered as a

prototype structure for the design of the experimental specimens. A one third scale

model (see Figure A2) was generated from the prototype structural system shown in

Figure A 1 to develop the test specimens. The columns were 600 mm square and the

slab 300 mm thick in the prototype structure and were reduced to 200 mm square

columns and a 100 mm thick slab in the model structure.

A.2 Loadings and Material Properties

A live load of 3 kPa and a total dead load of 8.2 kPa were considered for the design of

the prototype structure. The dead load included the self weight of the slab plus a 1 kPa

superimposed dead load. In the design the load factors of 1.25 and 1.5 were used for

the dead load and the live load, respectively. The design load of 14.75 kPa was used

for both the prototype and model structures.

A concrete of strength of 25 MPa was used for both the prototype and model

structures and the yield strengths of the conventional reinforcing bars and prestressing

tendons were assumed to be 450 and 1250 MPa, respectively. The modulus of

elasticity of the concrete was taken as 20 GPa and for the conventional and

prestressing reinforcement as 200 GPa.


A-2

c c c

c c c

All columns 600x600

c c c

J
L12000 .I I 12000 I I I 12000 I I I 12ooo----J

TYPICAL FLOOR PLAN

-
4( po
-
f'U
l:: -
bo
:--
Portion
stuoy
unde~

__j
4( po
- / \v ~
SECTION 1-1

Slab thickness = 300 mm

Figure A.l -Prototype flat plate slab-column connection structural system.


A-3

D D D D

Region un+r
s~
D D D
o
L_~ All columns 200x200

D D D D
1
L! 0 77-T7
Region und r
study"
J
~000 4000 I. 1000

TYPICAL FLOOR PLAN

1~ ~0

?U
[
I~
Portion unde~
jStUOy __j
1 :JO
1:' ;>0

v ~
SECTION 1-1

Slab thickness = 100 mm

Figure A.2 -Model flat plate slab-column connection structural system.

A.3 Analysis and Design

The design stress resultants for the prototype and the model structures were

determined using the simplified method outlined in AS 3600-1994 and using the

equivalent frame method for the model frames. The results from both methods were

compared and the design was carried out considering the worst case from these

analyses. In the model slabs, the flexural moment reinforcement was increased by a

factor of 1.5 to give a bias towards the punching shear failure.


A-4

Two dimensional frame analyses were carried out for the internal and edge frames in

the x- and y-directions of the model structure. The width of the internal frame was

taken equal to the centre to centre distance of the adjacent panels, while for the edge

frame the effective width was taken from the edge of the slab to the centre of the

panel. The bending moment diagrams for the design regions of these frames are shown

in Figure A3. In Figure A4 the bending moment diagrams are given for the analyses

using the simplified design method of AS 3600. The bending moments shown in these

figures are the total bending moments and were distributed to column and middle

strips in the proportion of75 percent and 25 percent, respectively.

The specimens were developed with the slabs extending to second line of

contraflexure as discussed in Section 3.2 of this thesis. The length of the column for

the specimen slab-column connections (see Figure A5) was set to give the correct

stiffness to the model slab. The flexural stiffness of the column of the 1/3 scale

4
slab-column connection is 2 E4nim , where Em, Im and Lm are the modulus of

elasticity, moment of inertia and length of the column for the model structure. The

flexural stiffness of the column of the slab-column connection for the test specimen is

3
Es Is where E8 , Is and Ls are modulus of elasticity, moment of inertia and length of
Ls

the column for the test specimen, respectively. Equating these stiffnesses, the required

length of the column for the specimen is

(Al)
A-5

48

Note:

End with column


In the specimen
~-----------400o-----------~

Not to the scale


BMD FOR X-DIRECTION INTERNAL FRAME

10.5 Units:
BM (kN)
SF (leN)

"~: : : : : :"'
24

::-----=:::::::::: I
~a...--------------41000f-------------
BMD FOR X-PIRECTION EXTERNAL FRAME

(a) X-direction frames

Key:
End with column
I in the specimen
Y QIRfCTIQN INTERNAl SpAN EDGE FRAME Not to the scale
Units:
BM {KNm)
SF (kN)

Y-QIREC]ON ENQ SpAN EQGE FRAME

(b) Y-direction frames

Note: Total bending moments are shown. These are distributed to column and middle
strips.

Figure A.3 - Bending moment diagrams from the analyses of the model frame.
A-6

Note:

~----------~ooo-----------~
I End with column
in the specimen
Diagrams - Not to the scale
Units:
8MD FOR EDGE COLUMN STRIP
BM (kNm)
SF (kN)

~-----------400o-----------~

8MD FOR MIDDLE STRIP

74.7kNm
24.
\

8MD FOR INTERNAL COLUMN STRIP

(a) X-direction frames

Key:
End with column
I In the specimen
Y-Q!RFC]ON INJFRNA! FQGE CO! liMN SJBIP Not to the scale
Units:
BM (kNm)
SF (kN)

V QIRECTIQN INIFRNAI MIQQI E Sm!p

(b) Y-direction frames

Figure A.4 - Bending moment diagrams for the model frames from the simplified
design method (AS 3600 - 1994).
A-7

600mmx600mm
column

200mmx200mm
column OO
1
[
T

1/3 scale slab-column Specimen slab-column

Prototype slab-column

Figure A.5 - Size of column and slab for different cases.

Using Lm = 1.35 min the equationAl gives Ls = 0.51 m. Considering laboratory and
other constraints the length of the column for the laboratory specimens was taken as

Ls =0.60m.

The fmal geometry of the test specimens and the positions of the conventional and

prestressing reinforcement are given in Chapter 6 of this thesis. The maximum

reinforcing steel ratio in the design slabs came to 1.1 percent which was then increased

by 50 percent to I. 7 percent to give a bias in the design to a punching failure.

Table A 1 gives typical average flexural reinforcement contents of slabs designed by

other investigators. The table shows that the quantity of steel used in this study is

typical of that used in studies on the punching shear strength of plate slabs.
A-8

Table A.l • Reinforcement contents of some typical flat plate slabs tested

Reference Specimen Connection p (%)

Hanson and Hanson (1968) D15 Edge 1.67

Zaghlool (1971) Z-V(3) Edge 2.23

Zaghlool (1971) Z-V(2) Edge 1.65

Stamenkovic and Chapman (1974) V/E/1 Edge 1.17

Zaghlool et al. (1970) I,ill,N Comer 1.47

Zaghlool (1971) Z-IT (5) Comer 1.23

Stamenkovic and Chapman (1974) V/C/1 Comer 1.17

Walker and Regan (1987) SCll Comer 1.42

Hammill and Ghali (1994) NH1. NH2 and NH4 Comer 1.47

Alexander and Simmonds (1992) P19S75 Interior 1.23#

Alexander and Simmonds (1992) P19S50 Interior 1.76#

#Note: Maintained this reinforcement content within a distance from the face of the
column equal to the effective depth of the slab.
APPENDIX 8 - RELAXATION TEST AND PRESTRESS LOSSES

Two types of prestressing losses were considered in the tests, these being the loss of

prestress caused by the anchorage slip and loss due to relaxation of the tendons. Tests

for evaluating the prestress losses were carried out for spans of 3.82 m for the edge

column specimens and 2.44 m for the comer column specimen. The set up used for

these tests is shown in Figure B.l. A load cell was placed between the anchorage plate

and the barrel-wedge system to regularly monitor load in the prestressing wire. The

output terminal of the load cell was connected to an HBM amplifier and a data

acquisition unit.

To test for the losses the load was applied gradually and the wedges were tapped in the

barrel gently and uniformly. Once the load reached the desired level, the wedges were

taped to fit tight in the barrel, the pump was released and the wire was cut. The loss

due to anchorage draw in was obtained from the change in the load measured at the

cutting of the wire from the jack. Three tests were undertaken for each span. Load

readings from the amplifier were taken regularly for two days to determine the

relaxation losses. The averages of the three tests for each series are shown in

Table B.l. The average room temperature during the test was recorded to be 20

degrees with the tests undertaken in the same environment as the test slabs.

The load to be applied to the prestressing wires in each test was determined by adding

the load given in Table B.2 to the required effective prestressing force. Further, a load

of 1 kN was added to the first two cables prestressed in each slab to compensate for

the effect of elastic shortening of the slab.


B-2

plate
Anchorage Plate
(100x100mm)

\_Barrel and
Dead end Jacking en d wedge

200mm long prestressing yoke


255x255mm steel sections L = 3820 mm for series I
fixed ta the laboratory floor L = 2440 mm far series II

Figure B.l - Test set up for loss evaluation.

Table B.l - Prestressing losses.

Average Average Loss Due to Relaxation (kN)


Anchorage
Series Length Slip Loss 24Hours 48 Hours
(m) (kN)

I 3.82 5 4 6

IT 2.44 8 2 3

Table B.2 - Amount of additional load added to the effective prestress.

Span (m) 3.57 4.90 2.55

Additioanlload (kN) 10 7 14
APPENDIX C- HORIZONTAL REACTION FOR SPECIMEN S1

In the post analysis of the test data of specimen S1 a problem with the experimental

set-up for the measuring of the horizontal reaction became evident. The support details

and support reactions for specimen S 1 are detailed in Figure C.1. In the test

arrangement the connection of the load cell measuring the vertical load allowed for an

accidental transfer of horizontal load into the testing frame and this component was

unmeasured.

The horizontal reaction Rx was split into two components Rh1 and Rh2 , as shown in

Figure C.1c. The arrangement of the load cells measured Rh1 only and Rh2 was not

measured. Using a linear finite element model the component of the horizontal

reaction Rh2 is determined. The mesh used in the finite element analyses consisted of

20-node 3D brick elements and is shown in Figure C.2. The finite element program

DIANA (1997) was used for the analysis with the elastic modulus taken as

Ec = 20.0 GPa.

From the finite element analysis the ratio of the vertical to horizontal reaction was

calculated and the missing component determined. The finite element analysis gave a

ratio of the total horizontal reaction to vertical reaction of 0.5. The ratio of the

measured component of the horizontal load Rht to the vertical reaction Rz is shown in

Figure C.3 up to the point of punching failure. The measured ratio is almost constant

at 0.28 and supports the calculated correction factor being a constant. Further, in

Figure 7.3 the corrected horizontal reaction for specimen S 1 is plotted with the

measured horizontal reactions for specimens S2 and S3, that were of similar geometry.

The figure shows that the corrections based on the finite element model appear to be

reasonable.
C-2

I")
0
<0

(a) Reactions at the column support

mm thick Steel Plate on which


Long. bars ore welded

mm die. boll bearings

welded to top of I beam

(b) Arrangement of load cells showing the source of error

I")
0
"'

R,

(c) Ditribution of horizontal reactions in the specimen Sl

Figure C.l - Support system showing the source of error and actual distribution
of horizontal reaction.
C-3

X
y

Plane of symmetry

Figure C.2 -Mesh for finite element analysis of specimen Sl.

0.35
0.3
0.25

-N
a:,.. 0.2
.c
a: 0.15
0.1
0.05
0
0 20 40 60 80 100 120
Rz(kN)

Figure C.3 - R%z versus Rz. plot for specimen Sl.


APPENDIX D- CODE MODELS

D.1 Introduction

Different codes use different models for the design of punching shear. In this appendix

a brief review of these models is presented. Altogether five codes were considered in

this study and these are:

• ACI 318-1999

• AS 3600-1994

• BS 8110-1997

• CEB-FIP Model Code 1990 (1993)

• Eurocode 2-1992.

A non-dimensional moment-shear interaction expression is given for the American,

Australian and CEB-FIP model codes and the test results were compared in Chapter 7

of this thesis using these relationships.

D.2 ACI 318-1999

The eccentric shear model for the design of flat plate-column connections given in

ACI 318 (1999) is based on many studies including those by DiStasio and Van

Buren (1960), Moe (1961) and Hanson and Hanson (1968). The ACI code defines the

critical section as the section with its perimeter b0 to be minimum with the perimeter

taken as not closure than d/2 to the columns, where b0 is the perimeter of the critical
D-2

section and d is the effective depth of the slab. The critical perimeters for edge and

· comer column connections are as shown in Figures D.l and D.2, respectively.

For exterior rectangular columns in a two-way flat slabs, where any portion of the

column cross section is closer than 4 times the slab thickness from the free edge, the

nominal shear capacity Vc is the smallest of

(D.la)

V =(asd+ 2].Jf:bod (D.1b)


c bo 12

(D.1c)

where f3c is the ratio of long side dimension to the short side dimension of the column;

as is 30 for edge columns and 20 for the comer columns; and J; is the characteristic

cylinder strength of the concrete.

ACI 318 ( 1999) treats prestressed concrete slab-column connections similar to

non-prestressed members provided that one or more of the following conditions apply:

• The minimum distance from the column to the discontinuous edge is less than

4 times the slab thickness.

• J; s; 35.0MPa.
• 0.9 ~fpc~ 3.5 MPa, where /pc is the effective compressive stress in the

concrete in each direction.


D-3

Column section

Mux
ex
(C1 +d/2)
1----Free edge

Figure D.l - Critical section and location of moments for edge column
connections.

section

uy
Y----x cc2 *-d/2)
Free edges

(C1 +d/2)

Figure D.2 - Critical section and location of moments for corner column
connections.
D-4

The ACI code takes a fraction of the unbalanced moment, 'W Mu. of the total factored

unbalanced moment, Mu, as being transferred to the slab by flexure. The remaining

moment Yv Mu is transferred as a shear linearly varying around the critical section as

shown in Figures D.3 and D.4 for edge column connections and in Figure D.5 for

corner column connections. The total shear stress developed at the critical section is

the sum of the shear stress developed due to part of the factored unbalanced moment,

Mu, and the uniform shear stress due to the factored shear force, Vu. The maximum

factored shear stress is written as

(D.2)

where A is the area of the critical section; ex and ey are the eccentricities of moment

vectors in y and x directions, respectively, from the critical faces as shown in

Figures D.l and D.2 for edge and corner column connections; and lx and ly are

properties of the critical section analogous to polar moment of inertia about the y and

x directions, respectively.

The terms in equation D.2 are as defined in Figures D.l and D.2 and equations D.3

and D.4 for edge and corner connections, respectively. The moments Mux and Muy are

the moment Mu calculated at the centroid of the critical section in the x and y

directions, respectively.

For edge column connections:

(D.3a)
D-5

Vu, M ucx: Forces to be transmitted


by slab-column connection

~y Shear stresses on
critical section due Vu
/ and Mua

Span
be in

~X
Note:
Mucx =Mua +Mub

Figure D.3 - Shear moment transfer in edge slab-column connections.


D-6

Critical section

Free

u=critical perimeter=(2C1 +Cz +2d)


A=Area of critical section=u d

Jx .Jy =Properties of assumed critial section


II analogous to polar moment of inertia

Vu _ Vu + l'x Mux ex
- A Jx

Figure D.4 - Shear stress distribution due to Vu. and "{vy Muy for a typical column
(Shown for the edge column).
D-7

Shear stresses on
critical section due
Mux1 and Muy1

Free edge
"--x
Note:
Mucx =Mux1 +Mux2
Mucy =Muy1 +Muy2 Vu ,Mucx ,Mucy :Forces to be transmitted
by slab-column connection

Figure D.S - Shear moment transfer in corner slab-column connections.


D-8

A=du (D.3b)

(D.3c)

(D.3d)

[(CI+d /2)d3 +(Ct +d /2)3d] +(C +d)d 2 +2(C +d/2)d[(q +d/2) ex]2 (D.3e)
Jx 2 ex 1
6 2

(D.3t)

1
Yvx =1- 2 Ct+d/2 (D.3g)
1+3 C2+d

(D.3h)

For corner column connections:

(D.4a)

A=du (D.4b)

(D.4c)

(D.4d)
0-9

1
rvx = 1- 1+ -2 q +d t2
I----,,----::-
(0.4g)
3 Cz+d

1
Yvy =1- --=:2---;=== (0.4h)
1+- Cz+d
3 Ct+d/2

For rectangular column connections without shear reinforcement in the slab, the

governing equation is

(0.5)

where Vu is the maximum factored shear stress, ifJ is the strength reduction factor and

Av = ifJVc (0.6)
.,., n bod

The interaction between shear and moment is given by

Vu Mu
--+-..!:!:..- 1 (0.7a)
if' Vuo ifJ M uo

For comer column connections M u / M uo ,is taken as

(0.7b)

In equations D.7a and D.7b, Vu and Mu are the strength limit state shear force and

moment at the centroid of the column and Vuo and Muo are, respectively, the shear

capacity for zero moment and moment capacity for zero shear. The suffixes X and Y

denote the directions of the moment being considered. The relationships for Vuo and

Muo are given in equation 0.8 below. Since in ACI 318 (1999) the moments are
D-10

calculated at the centroid of the critical section, the terms in the moment-shear

interaction equation D.7a are modified to reflect the eccentricities x andy about they-

and x- axes, respectively. In this study all the code comparisons are made by

considering the moment and shear at the centroid of the slab-column intersection.

f.J:/3
Vuo (D.8a)
(-1 _Yv ex X y)
Yry 'y
ly
Ac lx

f.J:/3
Muo (D.8b)
rvx exflx

[M uo ]y = f.J:/3jJ (D.8c)
rvx eX X

Finally, it is worth highlighting that the moments [Mu]x and [Mu]r are the moments at

the centroid of the column about y and x-axes, respectively, and Mux and Muy are the

moments calculated at the centroid of the critical section in the x and y directions,

respective! y.

D.3 AS 3600-1994

In AS 3600-1994, the design punching shear capacity of reinforced and prestressed

concrete slabs with rectangular columns and without shear reinforcement in the slab is

limited by

(0.9)

where Vuo is the ultimate shear strength for zero moment and is given by

(D. lOa)
D-11

with

fcv =0.17(1+ jh)~fcm ~ 0.34~fcm (D. lOb)

and where v* is the factored design shear force; l/J is the strength reduction factor; u is

the length of the critical shear perimeter defined by a line parallel to the edge of the

column at a distance of d/2 from the column face; d is the effective depth of the slab

averaged around u; a is the linear dimension of the critical shear perimeter measured

parallel to the direction of M/; M/ and v* are, respectively, the unbalanced bending
moment in the direction being considered and shear force transferred at the centroid of

the column at the collapse limit state; f3h is the ratio of the largest dimension of the

support (Y) to the overall dimension (X) at right angles to Y ( f3h =Vx ,Y ~ X ;refer
Figure D.6); rJcp is the average intensity of effective prestress in concrete in MPa and

f~ is the characteristic compressive cylinder strength of the concrete in MPa.

Rearranging equation D.9 we get

(D.lla)

Taking M uo =Vuo Safu and substituting into equation D.lla gives the interaction

equation

v* M*
--+ v 1 (D.llb)
l/J Vuo l/J M uo

For comer column connections the term Mu vM


uo
is written as
D-12

Boundary of effective
area of column
r
Boundary of effective
area of column

r-----l~/2
[
r- ---- d/2
I ~--..,...-,...--t---' I
..
I
I
I I
I - - - Critical shear ----l . :
I
l \ \
:
I
perimeter :
I ..
<I
I
I
I
y

\ I I I
L --- J
X ~2

Figure D.6 - Critical shear perimeter used in AS 3600 (1994).

_v
M*_ _ [ _ v_
Muo Muo X
M*] +[M*] _v_
Muo y
(D.11c)

where the suffixes X and Y indicate of moments in X and Y directions, respectively.

Substituting equation D.lOa in the equation forM uo we write the ultimate moment

capacity for zero shear as

(D.12)

D.4 BS 8110: Part 1: 1997

The shear perimeter considered in BS 8110 code is different from those used in

American and Australian codes. The first shear perimeter recommended to be checked

is at a distance of l.Sd from the face of the column. Successive perimeters are

considered at 0.75d intervals for designing shear reinforcement. For corner columns

and edge columns with the bending about an axis parallel to the free edge (see

Figure D.7) the effective design shear force (Veff) is calculated to be 1.25 times the

design shear force transferred to the column Vr. that is


D-13

y
4
Free edges Column ~ ~Cdtical section
Column

~1 ==--=-----r~
t~
1_ _

Mux
a ritical sections a
a=(C 1 + 1 .5d) a=(C 1 + 1.5d)
b=(C2 + 1.5d) b= (C2 +3d)
1---Free edge

(a) Comer column (b) Edge column

Figure D.7- Critical section with moments (BS 8110-1997).

(D.13)

For an edge column with a bending moment Mt equal to Muy. about an axis

perpendicular to the edge (as shown in Figure D.8), the design effective shear Veff is

given by

Veff =l't (1.25 + 1.5M1 ) (D.l4)


Vrx
where x is the length of the side of the perimeter parallel to the axis of bending. The

shear stress (v) is given by

Veff
v=-- (0.15)
ud

where u is the perimeter of the critical section and d is the effective depth of the slab.

For a slab without any shear reinforcement v must be less than Vc where
D-14

y
4
Column : rCritical section

!c,i
! Muy ___ X Ib

a
a=(C 1 + 1.5d)
b= (C2 +3d)
~----Free edge

Figure D.S - Edge column with moment about the axis perpendicular to the free
edge.

v = 0.79 (lOOAs tx' (400 f(fcu )X' (D.16a)


c rm bd ) d 25 )

100Ast :s:; 3.0 (D.16b)


bd

400 ~1.0 (D.16c)


d

feu S40MPa (D.l6d)

and where Ym is the_ partial safety factor for strength of materials; Ast is the area of

tension reinforcement in the critical section; b is the width of the critical section; dis

the effective depth and feu is the characteristic cube strength of the concrete. For

feu< 25 MPa, the (ja/25) 113 term in equation D.l6a is taken as unity.
D-15

D.S CEB-FIP Model code 1990

The approach taken by the CEB-FIP model code 1990 (1993) differs completely in

form from the American and Australian codes but has some similarity with BS 8110-

1997. For edge column connections with the moment normal to the free edge the

CEB-FIP Model code assumes a shear stress of 'isd on the perimeter u1 as shown in

Figure D.9a. The shear stress ('isd) is given by

(D.17)

where Psd and Msd are shear force and unbalanced moment calculated at the centre of

the column; Vp is the vertical component of the prestress force passing through the

column within the distance of D/2 from the column, D being the overall depth of the

slab; K is a coefficient depending on the ratio of column dimensions C1 and C2 (for

square columns K=0.45); and W1 is a parameter of the control perimeter such that,

(D.l8)

where dl is an elementary length of the perimeter and e is the distance of dl from the

axis about which the moment Msd acts.

For an edge connection

(D.19)

For edge connections with the loading at the interior of the slab and with loading

eccentricity in the direction perpendicular to the slab edge, the shear stress is
D-16

and O.SC1

(a) Perimeter u1 (b) Perimeter ui


Figure D.9 - Critical perimeter for edge columns in the CEB-FIP Model
Code 1990 (1993): (a) perimeter u1 (b) perimeter ui.

assumed to be uniform on the perimeter ui as shown in Figure D.9b. The design shear
stress is given by

(D.20)

and the shear capacity is given by

(D.21)

where YMis a partial safety factor;fck is the characteristic cylinder strength; and

~ :::: 1 +.pff with din mm (D.22)

The control perimeter u; is given by


(D.23)
0-17

Letting -rRd =-rSd and denoting Vu =Psd - Vp and M u =M Sd equation 0.20 is

written as

(0.24a)

Further, substituting equation 0.21 into 0.24 and taking

(0.24b)

(0.24c)

leads to the interaction equation

YM Vu + YM Mu 1 (0.25)
Vuo Muo

For corner column connections the CEB-FIP Model Code assumes a shear stress of 'rsd

on the perimeter u1 as shown in Figure O.lOa. However, provided the eccentricity of

loading is towards the interior of the slab, the uniform shear distribution takes place

along the perimeter ui as shown in Figure O.lOb and the shear stress developed is

given by

(0.26)

The suffixes X and Y denote the moment in x and y directions, respectively. The

symbols are as defined as above for the case of edge column connections and the shear

perimeter ui and W1 are given by


D-18

Slab free edges Slab free edges


Column Column

T Flesser of 1.5d ond 0.5c 2


i 2d

perimeter ~-~iX ntml pedmetec

Lesser of 1.5d and 0.5C1

(a) Perimeter u1 (b) Perimeter ui


Figure D.10 - Critical perimeter for the comer column in CEB-FIP Model
Code 1990 (1993).

(D.27)

(D.28a)

(D.28b)

For the comer column connections the interaction diagram is given by equation D.25

with

Mu -[Msd] [Msd] (D.29)


Muo- Muo x + Muo y
As for the BS 8110 ( 1997) model, the amount of tensile reinforcement is one of the

factors controlling the shear strength. The shear strength is zero for the slabs without

tensile reinforcement.
D-19

D.6 Eurocode 2-1992

In Eurocode 2 ( 1992) the design model for slabs without shear reinforcement is

limited to columns having an aspect ratio of C1/C2 of less than 2. For edge and comer

columns the Eurocode 2 model for punching shear assumes the critical perimeter to be

at a distance of 1.5d from the faces of the column (shown in Figures 0.11 and D.12),

provided that the perimeter is less than 11d. For slabs where the perimeter exceeds

lld, the control perimeter is defined as shown in Figure D.13. For slabs without shear

reinforcement the shear force on the critical section is given by

(D.30)

where Vsd is the total design shear force, u is the critical perimeter, {3 is a coefficient to

take into account the effects of eccentricity of loading and is taken as 1, 1.5, 1.4 and

1.15 for concentric loading (no eccentricity), comer columns, edge columns and

interior columns, respectively.

The design resistance is given by

(D.31)

where 'rRd is the basic resisting shear strength of the concrete and is taken as

0.0525 f% (D.32)
rc ck

d is the average of the effective depths of the slab in the x and y directions and rc is a
partial safety factor. The other parameters used in equation D.31 are
D-20

Column

perimeter

1-----Free edge
Figure D.ll - Control perimeter of Eurocode 2 for edge column.

edges
Column

~o~~rol
1
...... ..:..?.<J. perimeter

1.5d

Figure D.12 - Control perimeter of Eurocode 2 for comer column.

, .sd -L_. \=. . . . . . 1


Where,
c,
l
1...............................\

c,, ~
l 2C2
;..--~-.,.

' C22/2 S.6d-C22.


4
• 4ColurJJn
• 4 .4

\_i C22/2 C£2. ~ ~ C22..8d


C,~---- ~
\ .....L...
•...•-.....•." --.....•.-.• - - - ·•---'4
...•..

Cn/2 Contcol pedmeter


C1 >C2

Figure D.13 - Control perimeter of Eurocode 2 for column with u > lld.
D-21

k = 1(1.6-d) ~ 1.01 (where dis in metres) (0.33)

~ 0.9CTpc
P1 = Plx Ply+ < 0.15 (0.34)
Acfyd

where Pix, Ply are the ratios of tension steel in x and y directions, respectively, CTpc is

the average prestressing force resulting from the initial prestress without losses and /yd

is the design yield stress of the reinforcement. As for BS 8110 (1997), Eurocode 2

does not provide guidelines for the transfer of unbalanced moment into the column

without shear.

Potrebbero piacerti anche