Anda di halaman 1dari 147

Influence of Concrete Strength on the

Behaviour of Bridge Pier Caps

by
Gavin MacLeod
March, 1997

Department of Civil Engineering and AppIied Mechanics

McGill University
Montreal, Quebec

Canada

A thesis submitted to the Faculty of Graduate Studies and Research in partid fulfilment of the

requirements for the degree of Master of E n g i n e e ~ g

Gavin MaCLeod, 1997

National Librciry

Bibliothque nationale

du Canada

Acquisitions and
Bibliographic SeMces

Acquisitions et
services bibliographiques

395 Wellington Street


OttawaON KlAON4

395. nie Wellington


OttawaON KlAON4
Canada

Canada

The author has granted a nonexclusive licence allowing the


National Library of Canada to
reproduce, loan, distribute or seIl
copies of this thesis in microform,
paper or electronic formats.

L'auteur a accord une licence non


exclusive permettant la
Bibliothque nationale du Canada de
reproduire, pr?er, distniuer ou
vendre des copies de cette thse sous
la fome de micro fi ch el^ de
reproduction sur papier ou sur format
lectronique.

The author retains ownership of the


copyright in this thesis. Neither the
thesis nor substantid extracts fiom it
may be printed or otheniise
reproduced without the author's
permission.

L'auteur conserve la proprit du


droit d'auteur qui protge cette thse.
Ni la thse ni des extraits substantiels
de celle-ci ne doivent tre imprims
ou autrement reproduits sans son
autorisation.

ABSTRACT
Two full-sale rernforced concrete bridge pier caps were constructeci and tested to
investigate the influence of concrete strength on their behaviour. The arnount of uniformly
distributeci reinforcement required for crack control at service load Ievels was aiso varied in order
to investigate the suitabili~of current design approaches for these disturbed regions. in addition,
strut-and-tie modeIs, refined strut-and-tie models and non-Iinear finite element analyses are used
CO predict

the comp!ete behaviour of the test specirnens.

Deux chapiteaux de pont grandeur relle en bton arm ont t construits et tests pour
tudier I'infiuence de la rsistance du bton sur leur comportement. La quantit d'armature
distribue uniformment, ncessaire pour contrler les fissures sous charges de service, a t
varie pour dterminer si les approches de conception actuelles conviennent pour ces structures
spciales. De plus, des modles bielle et tirant simple, des modies bielle et tirant plus dtaills
et des analyses non-linaires par lments finis sont utiliss pour prdire Ie comportement complet
des spcimens d'essai.

ACKNOWLEDGEMENTS
The author would Iike to thank Professor Denis Mitchell for his cornpetent supervision,

support and encouragement throughout this research programne. The author would also like to
express his gratitude to Dr. WiiIiam Cook for his advice and assistance durhg this programme.
The efforts of Marek Pnykorski, Ron Sheppard, John Bartczac and Darnon Kiperchuk

in preparing the experiments are gratefully acknowtedged. The author would also like to thank
Homayoun Abrishami, Arshad Khan, Stuart Bristowe, Glenn Marquis, Peter McHarg and PierreAlexandre Koch for their contributions in the construction and testing of the specimens.
The financiai support provided by Concrete Canada, a Network of Centres of Excellence
Pro-

hnded by the Minister of State, Science and Technology in Canada, is greatly

appreciated.

iii

TABLE OF CONTENTS
ABSTRACT
RSUM

..

II

ACKNOWLEDGEMENTS

LIST OF FIGURES
LIST OF TABLES
LIST OF SYMBOLS
1.

INTRODUCTION
Introduction .
Disturbed Regions

Previous Research on Strut-and-Tie Models


Design Using Strut-and-Tie Models

AC1 Design Approaches for Disturbed Regions

1S. 1 AC1 Provisions for Deep Beams

1.5.2 AC1 Provisions for Brackets and Corbels

Experiments on Deep Beams, Corbels and Pier Caps


1.6.1

DeepBearns

1.6.2

Corbels

1.6.3

PierCaps

Detailed Analysis Procedures

.
.

1.7.1

Refined Strut-and-Tie ModeIs

1.7.2

Non-Linear Finite EIement Analysis

1.8.1

.
Compressive Strength .

1.8.2

Flexure and Axial Loads

1.8.3

Minimum Reinforcernent for Flexure and Shear


Strut-and-Tie Provisions

High-Performance Concrete

1.8.4

Crack Widths and Crack Spacing


Research Objectives

..

iii

viii

2.

EXPERIMENTAL PROGRAMME
2.1

Details of Specimens

2.2

Material Properties

.
.

2.2.1

Concrete

2.2.2

Reinforcing Steel

2.3

Test Setup and htnimentation

2.4

Testing Procedure

3.1

Load-Deflection Responses

3.2

3.1.1

Specimen CAPN

3.1.2

Specimen CAPH

39

39

.
.
.

43
46

48

51

76

43

Development of Strains
3.2.1

Specimen CAPN

3-2.2 Specimen CAPH


3.3

4.

Development of Cracking
3.1.1

Specimen CAPN

3.1.2

Specimen CAPH

COMPARISONS AND ANALYSES OF RESULTS .


4.1

Cornparison of Responses of Normal- and High-Strength Concrete

4.2

76

dredictions of Results .

83

4.2.1

AppIicability of Plane-Sections Anaiysis

83

4.2.2

Simple Strut-and-Tie Models

83

.
4.2.4 Non-Linear Finite Element Analysis Using Program FIELDS .
.
.
Estimates of Crack Widths

85

4.2.3

4.3

5.

Specimens

Refined Strut-and-Tie Models .

CONCLUSIONS

REFERENCES

APPENDK - EXPERIMENTAL DATA .

88
101

103

105

LIST OF FIGURES
Typical forms of cap beams and pier caps used in bridge construction .
Examples of disturbed regions .
Strut-and-tie modelling of a deep beam with a direct support and a tension
hanger support .
Influence of principal tensile strain, E , , on compressive strength of
diagonally cracked concrete

Compressive strength of sut versus orientation of tension tie passing


through strut

Provisions for brackets and corbels

Applicability of stmt-and-tie mode1 for predicting series of bearns testeci by


Kani

Crack control reinforcement required with assurnption of straight-tine

.
compressive struts
Investigating the effect of distributed reinforcement on deep beams

Evaiuating stresses at Gauss points in quadraterai element


Determining average concrete tensile stress, f,,,from suain,
Investigating stress condhon at crack interface
Influence of concrete strength on shape of stress-strain curve

E,

Crack width parameters


Side-face cracks controlled by skin reinforcement
Test simulation of cantilever cap beams
Specimen details
Concrete properties

Typical stress-strain resopnses of reinforcing bars


Specirnen CAPN under the MTS testing machine

Different bearing details of specimens CAPN and CAPH

LVDT locations
Strain gauge locations and crdck width lines of measurernent

.
54

Strains in bottom bar of CAPN tension fie, determined from strain gauges

.
.

St.4ns distributed reinforcement of CAPN, determined from strain gauges

58

Loaddeflection responses of speciniens

56

Longitudinai strains from LVDTs at mid-height and at the level of the tension
tie of CAPN

Responses of CAPN-A rosettes A6 and A7

Responses of CAPN-B rosettes B6 and B7

Strains in bottom bar of CAPH tension tie, determined from m a i n gauges

Strains distributed reinforcement of CAPH, determined from strain gauges

Longitudinal strains frorn LVDTs at mid-height and at the level of the tension
tie of CAPH Respomes of CAPH-A rosettes A6 and A7

Responses of CAPH-B rosettes B6 and B7

Development of cracks in CAPN


CAPN after failure

Development of cracks in CAPH


CAPH after failure

Comparison of Ioad-deflection responses of specimens

FIexural crack widttis measured at the level of the tension tie in specimens

Diagonal crack widths measured at mid-height of specimens

infiuence of distributed reinforcernent ratio on crack control

S imple strut-and-tie mode1 for specimen CAPN


Simple strut-and-tie mode1 for specirnen CAPH

Refined strut-and-t ie models for specirnen CAPN


Refined stnit-and-tie models for specimen CAPH
Predicted load-deflection responses of specirnens
Predictions of deflected shapes of specimens at maximum predicted loads
Predicted strains and stresses in specimen CAPN at a load of 920 kN .
Predicted strains and stresses in specimen CAPH at a load of 920 kN .
Predicted strains and stresses in specimen CAPN at a load of 2280 kN
Predicted strains and stresses in specimen CAPH at a Ioad of 2280 kN
Predicted strains and stresses in specimen CAPN at a load of 4980 kN
Predicted strains and stresses in specimen CAPH at a load of 5340 kN
Predictions of stress development in main tension ties of specimens

Comparison of predictions of stress in main tension ties at general yield

vii

LIST OF TABLES
1.1

Effective stress levels in struts .

1.2

Effective stress levels in nodal zones

2.1

10

Mix design for 35 MPa concrete

44

2.2

Mix design for 70 MPa concrete

2.3

Concrete properties

46

2.4

ReUrforcing steel properties

48

4.1

Cornparison of strut-and-tir: predictions with measured loads at general


-

88

98

102

102

yielding
4.2

Cornparison of refined stmt-and-tie predictions and non-linear finite element


predictions with rneasured loads at generai yielding

4.3

Cornparison of predicted and measured crack widths and principal tensile


strains in the main tension ties of specimens

4.4

Cornparison of predicted and rneasured diagonal crack widths and principal


tensile strains at mid-height of the specimens

Readings from vertical LVDTs used to determine the deflection of


specimen CAPN
Readings fiom LVDTs located at the Ievel of the main tension tie in
specimen CAPN-A

Readings from LVDTs located at the level of the main tension tie in
specimen CAPN-B

Readings from LVDTs Iocatd at mid-height of specimen CAPN-A

Readings from LVDTs located at rnid-height of specirnen CAPN-B


Readings from LVDTs rosettes located in end A of specirnen CAPN

Readings from LVDTs rosettes Iocated in end B of specimen CAPN


Strains fiom strain gauges located in end A of specirnen CAPN
Strains Crorn strain gauges located in end B of specimen CAPN
Readings from vertical LVDTs used to determine the deflection of
specimen CAPH
Readings from LVDTs located at the level of the main tension tie in
specimen CAPH-A

Readings from LVDTs located at the level of the main tension tie in
specirnen CAPH-B

.
viii

.
.

A. 13

Readings fiom LVDTs located at rnid-height of specimen CAPH-A

124

A. 14

Readings from LVDTs loated at mid-height of specimen CAPH-B

125

A. 15

Readings from LVDTs rosettes located in end A of specimen CAPH

A. 16

Readings from LVDTs rosettes located in end B of specirnen CAPH

.
.

.
.
.

127

A. 17

Strains fiom strain gauges located in end A of specimen CAPH

128

A. 18

Strains from strain gauges located in end B of specimen CAPH

130

126

LIST OF SYMBOLS
r?iaximum aggregate size

shear span
effective area of concrete surrounding each reinforcing bar
area of effective embedment zone of concrete where reinforcing bars can influence crack
widths
effective cross-sectionai area of concrete compression strut
area of reinforcement required to resist moment, Mu,in corbel
area of horizontal stirrup reinforcement in corbel
area of reinforcernent required to resist horizontal rensile force, N,,, in corbel
area of primary tension reinforcement
area of reinforcing steel

minimum area of flexural reinforcement


area of reinforcing steel in main tension tie
area of shear reinforcement perpendicular to axis of member within a distance s
area of shear friction reinforcement
area of shear reinforcement parallel to axis of member within a distance s2
width of member
width of tension zone of member
minimum effective web width within depth d
clear concrete cover
distance from extreme compression fibre to neutral axis
force in compression strut
distance from extreme compression fibre to centroid of tension reinforcernent
depth of compression strut
diarneter of reinforcing bar
distance from exueme tension fibre to centre of closest bar
modulus of elasticity of concrete
modulus of eIasticity of reinforcing steel
modulus of elasticity of tension tie reinforcernent
concrete stress
compressive strength of concrete
concrete cracking stress, equal ro E,E,
limiting compressive stress in concrete compression stnit

average principai tensile stress in concrete


average principal compressive stress in concrete
mdulus of rupture of concrete
average stress in reinforcing steel
calculted stress in reinforcement at specified Ioads
stress in reinforcing steel across crack
limiting compressive stress of diagonaily cracked concrete

overall depth of beam


height of effective embedrnent of tension tie
distance of main tension reinforcement from neutral axis
distance of extreme tension fibre from neutrai axis
post-peak decay t e m for stress-suain relationship of concrete
coefficient that characterizes bond properties of reinforcing bars used in CEB-FIP crack
width expression
rtmforcing bar location factor used in development length expression
coefficient to account for strain gradient used in CEB-FIP crack width expression
reinforcement coating factor used in development le@

expression

reinforcing bar size factor used in deveIopment length expression


length of bearing
development length of reinforcement
straight embedrnent length
clear span
cracking moment
factored mom~ntat a section
factored moment resistance at a section
design ultimate moment
curve fitting factor for stress-strain relationship of concrete
applied axial tension
horizontal tensile force
spacing of shear reinforcement parallel to axis of member

maximum spacing between longitudinal reinforcing bars


mean crack spacing

mean spacing of diagonal cracks


spacing of shear reinforcement perpendicular to axis of member
force in tension tie

nominai shear strength provided by concrete


shear stress at crack interface
limiting shear stress dong crack
nominal shear strength at a section
nominal shear strength provided by shear reinforcement
factored shear force at a section
average crack width. equal to

E,S,

characteristic crack width, equal to 1 . 7 ~ ~

mean crack width, equai to e#,,,


maximum crack width
limiting crack width parameter, equal to f;@A) '13
ratio of average stress in rectangular compression block to concrete strength
factor accounting for strain gradient, equaI to b l h ,
ratio f depth of rectangular compression block to depth to neutral axis
density of concrete
shear strain
yield deflection
ultimate deflection
compressive strain
strain in concrete at peak compressive stress
strain in concrete caused by stress
suain in concrete at cracking
strain in reinforcing steel
strain in reinforcing bar at crack location
horizontai tensile strain
suain in ydirection
principal tensile strain
largest tensile strain in effective embedment zone
principal compressive strain
smdlest tensile strain in effective embedment zone
angle of principal compressive main from horizontal
smallest angle between compression stmt and tension tie crossing suut
effective coefficient of shear fiction
factor to account h r influence of high-strength concrete, eqiial to 0.55 + 1.2Wylf;
reinforcement ratio of primary tension reinforcement, equal to AJbd
xii

Pef

AiAccf

Pw

Aibwd
capacity reduction factor, taken as 0.85 for shear
material resistance factor for concrete

matend resistance factor for prestressing steel

6,

material resistance factor for reinforcing steel

xiii

CHAPTER 1

INTRODUCTION

1.1

Introduction
Figure 1.1 shows some typical forms of cap bearns and pier caps used in bridge

construction. Although there is a variety of forms for these types of elements, this research
programme will examine a pier cap of f o m shown in Fig. l.l(b), which is also representative
of the cantilever portions of the cap beam shown in Fig . 1.1(a). This chapter first reviews the
behaviour and design of disfurbed regions, highlighting the use of strut-and-tie modeIs for design.
.4 review of experimental work carried out on deep beams, corbels and pier caps is presented to

provide background information on the behaviour of disturbed regions which are similar to those
investigated in this research programme.

1.2

Disturbed Regions
Regions of a member in which the "plane-sectionsn assurnption is appropriate are

sometimes referred to as B-regions (where B stands for beam or Bernoulli hypothesis). Other
regions of a member where the strain distribution is significantly non-linear are referred to as Dregions, or disturbed regions (Schlaich et ai. 1987). This non-Iinear distribution of strains is due
to a complex interna1 flow of stresses adjacent to abrupt changes of cross-section or the presence
of concentrated loads or reactions. The two maid design assumptions, that plane sections remain
plane and that the shear stress can be assumed to be unifonn over the nominai shear m a , are no
longer valid in disturbed regionsSeveral examples of disturbed regions are shown in Fig. 1.2. where the flow of
compressive stresses is shown by dashed lines, and tende ties are indicated by solid lines.
Figure 1.2(a) shows how the presence of a support reaction intempts the uniform diagonal
compression field in a s h p l y supported "slendernbeam with stimps. The flow of compressive
stresses fan into the support causing a disturbed region near that location. Figure 1,2(b) shows
a deep beam subjected to concentrated loads. Because of the complex flow of stresses in this
member, the entire member is a disturbed region. The flow of forces from the top of the beam

(a) Continuous cap beam

(b) Cantilever cap bearn

(c) Deep-water pier cap

Figure 1.1 Typical foms of cap bearns and pier caps used in bridge construction

unifom
field

fan

/'

/'

' tension tie

(a) Simply supported beam


compressive
Suut

\'--

tension tie

(b)Deep beam

CO mpressive

stnlt

(c)Corbet

Figure 1.2 Examples of disturbed regions

to the reaction areas delineates concentrated compressive stresses as shown. The resisting
mechanism, consisting of the flow of compressive stresses and the presence of the tension tie,
resembles a tied arch.

The corbel shown in Fig. 1.2(c) is a D-region characterized by

concentrated compressive stresses flowing from the bearing areas to the column. The horizontal
components of these diagonal compressive stresses must be equilibrated by tension in
reinforcement which is well-anch~redat the outer edges of the bearing areas.

1.3

Previous Research on Strut-and-Tie Models


Truss models have been used since the turn of the century for the design of slender

reinforced concrete beams (Ritter 1899, Morsch 1909). These early truss models had a
compression chord and tension chord with diagonal compressive stmts, typically assumed to act
at 45". These tmss models formed the basis of code developrnents in Europe and North America
for the design of slender beams. Recently, renewed interest has been generated in t m s models
as a design tool, not only for siender beams, but also for the design of disturbed regions.
A strut-and-tie model provides a simple toof for the design of disturbed regions, that is,

regions having a complex flow of stresses. The flow of forces in a disturbed region is ideaiized
using compressive stnits to represent the concentrated compressive stresses and reinforcement to
represent the tension ties (see Fig. 1.2). Figure 1.3 illustrates the development of a strut-and-tie
model for a deep beam with a direct support and a tension hanger support. The first step in
design is to sketch the flow of compressive stresses, in the form of compressive stmts, from the
location of the applied loads to the support regions. The next step is to sketch the tension tie
reinforcement required to complete the strut-and-tie model (see Fig. 1.3(a)). The shaded areas
in Fig. 1.3(a) where compressive struts and tension ties meet are referred to as nodal zones.
These nodal zones are regions of multidirectionaily stressed concrete. In order to examine the
equilibriurn of the model, an ideaiized truss model is created as shown in Fig. 1.3(b). The
dashed lines represent the centreline of the compressive struts, and the solid lines are located at
the centroid of the tension tie reinforcement. The nodes of the idealized t m s occur at the
intersections of the compressive struts and tension ties in the idedized t m s model. One of the

main advantages of using strut-and-tie models is that the flow of forces can be easily visualized
by the designer. Scme experience is required to determine the most efficient strut-and-tie model
for any given situation, as no unique solution exists. As this is a lower-bowid solution technique,
al1 solutions will give conservative resulrs provided that equilibriwn is satisfied, applicable stress
limits are not exceeded and the reinforcement is capable of developing the required stress.
Schlaich and Schafer (1984) and Schlaich et al. (1987) suggest choosing the geometry of a sirut-

- 4 -

- 4 -

(a) Strut-and-tie rnodel

(b) Tniss idealization

Figure 1.3 Strut-and-tie modelling of a deep beam with a direct support


and a tension hanger support

and-tie model such that the angle of each compression diagonal is within

* 15" of the angle of

the resultant of the compressive stresses obtained from an elastic analysis. While this approach
gives some guidance in choosing the model geometry, it should be noted that considerable
redistribution of stresses may occur afier cracking.
Once the geometry of the strut-and-tie model has been chosen, forces in the t m s
members can be found from equilibrium. The required arnount of reinforcement for each tension
tie can then be determined while ensuring that this reinforcement is anchored in such a way to
transfer the required tension to the nodal zones of the truss. The dimensions of a concrete
compressive strut m u t be made large enough such that the calculated stress in the stmt is less
than its h i t i n g stress.
Considerable research has been carrieci out on limiting stresses in concrete compressive
struts and the influence of anchorage details on the dimensions of these stnits. Thrlirnan et al.
(1983) and Marti (1985) concludeci that the stress in the stmts be limited to 0.60fCf, while

Ramirez and Breen (1991) suggest a compressive stress limit of 2 . 4 9 K (in MPa units).
Schlaich et al. (1987) proposed stress limits for the struts which depend upon the stress conditions
and the angle of cracking associated with the stmt (see Table 1.1).
Vecchio and Collins (1986) developed expressions for the modifieci compression field
theory which accounted for the strain softening of diagonally cracked concrete (see Fig. 1.4).
The limiting compressive stress, f-.

where: fCt
1

is given as:

concrete compressive strength,

= principal tende strain.

The following strain compatibility equation provides a means of detennining the principd tensile

strain, E , in diagonally cracked concrete:

where:

E,

= horizontal tensile strain,

= principal compressive strain,

= angle of the principal compressive strain from horizontal.

---

I I -

Conditions of Strut

1
1

--

Effective
Stress Level

by
-.
Schlaich er

Undisnubed and uniaxial state of


compressive stress that may exist for

al.

prismatic struts

( 1987)

Tensile strains ancilor reinforcement


perpendicular to the axis of the strut may
cause cracking parailel to the strut with
nonnal crack width
Tensile strains andior reinforcement at
skew angles to the axis of the strut may
cause skew cracking with normal crack

II width

1 Skew cracks with extraordinary crack 1


width (expected if modelling of the stmts
departs significantly from the theory of

elasticity's flow of interna1 stresses)


-

Uncracked uniaxiaily stresseci struts of

MacGregor
( 1997)

Stmts cracked longitudinally due to


bottle-shaped stress fields with sufficient
transverse reinforcement
Struts cracked longitudinally due to
bottle-shaped stress fields without
transverse reinforcement

Il

Struts in cracked zone with transverse


tensions From transverse reinforcement

Table 1.1 Effective stress Ievels ii stmts (adaptai from Schlaich et al. 1987,
and MacGregor 1997)

(a) Average concrete compressive stress,f;_,from strains E, and E,

(b) Reduction in compressive strength with increasing values of E,

Figure 1.4 Infiuence of principal tensile strain, E,, on corvpressive strength of diagonally
cracked concrete (Vecchio and Coilins 1986)

Figure 1.5 Compressive strength of stmt versus orientation of tension tie passing
through stmt (Collins and Mitchell 1986)

In order to apply the strain softening expression to diagonal compressive struts, for use
in a stmt-and-tie model, Collins and Mitchell (1986) gave the foilowing expressions for the
limiting compressive stress in the struts:

where: f,

f:
1

where:

Iirniting compressive stress in the strut,

= concrete cylinder strength,


=

principal tensile strain.

principal compressive strain in the strut, taken as 0.002,

es

= strain in the tension tie crossing the strut,

= smallest angle between the stmt and the tension tie crossing the strut.

The variation of the compressive strength, f,, of a strut as a function of the angle, 8,, between
the strut and the tension tie passing through the strut is shown in Fig. 1.5.

It is also necessary to Iimit the compressive stress in the nodal zones of the strut-and-tie
model. The maximum compressive stress Iimits in nodal zones depend on the different straining
and confinement conditions of these zones. Figure 1.3 iIlustrates three types of nodes identifieci

as follows:
1

CCC - nodal zone bounded by compressive struts and bearing areas only,

2.

CCT - node with a tension tie passing through it in only one direction, and

3.

CTT - node with tension ties passing through it in more than one direction.

The two nodal zones in Fig. 1.3 located under the bearing areas at the top of the deep beam are
examples of CCC-nodes, that is, each node is bordered by a bearing area and two compressive
struts. The node above the direct support at the right end of the beam is an example of a CCTnode since it contains one principal tension tie passing through the zone. The node at the indirect
support, located at near the bottom-left corner of the deep beam show. in Fig . 1.3 is an example
of a CTT-node since two tension ties pass through it. These nodal zones must be chasen large

enough to ensure that stresses do not exceed the applicable stress limits. Table 1.2 outlines the
effective stress limits for nodal zones as determineci by several researchers.

Effective

Conditions of Nodal Zone

Proposed

Stress LeveI

CCC-nodes

O. 85f:

Collins and

CCT-nodes

O. 75f;

CTT-nodes

0.60fc

Mitchell
(1986)

Nodes where reinforcement is anchored


in or crossing the node
Nodes bounded by compression stnrts

1 .O v2 fcf (')

MacGregor

and bearing areas


Nodes anchoring one tension tie

O=

Nodes anchoring tension ties in more

than one direction


Table 1.2 Effective stress levels in nodal zones (adapteci from Collins and Mitchell 1986,
Schlaich et al. 1987, and MacGregor 1997)

1.4

Design Using Strut-and-Tie Modeis


The CSA Standard A23.3 "Design of Concrete Structures for Buildings" (CSA 1984,

1994), the Ontario Highway Bridge Design Code (OHBDC 199l), the Canadian Highway Bridge

Design Code (CHBDC 1997, currently under development) and the AASHTO LRFD
specifications (AASHTO 1992) have adopted the strut-and-tie design methods developed by
Collins and Mitchell (1986). In cornparing different codes, it is important to realize that the
Canadian standards use material resistance factors (@, for concrete. t$s for reinforcing steel and
4p for prestressing steel), while the U.S. codes use capacity reduction factors, 9, which depend

on the type of action- Both the CSA Standard and the CHBDC use the s a m material resistance

The CSA Standard uses 9, = 0.60, while the


CHBDC uses 9, = 0.75. In this section reference will be made to the requirements of the
Canadian Highway Bridge Design Code. The CHBDC states that strut-and-tie models shall be
considered for the design of deep footings and pile caps or other situations in which the distance
between centres of applied load and the supporting reaction is less than twice the cornponent
thickness.
factors for steel with t$$ = 0.85 and

#p = 0.90.

The design procedure, with reference to the relevant code requirements, is sumrnarized

in the following steps1.

Sketch the strut-and-tie model, assuming straight compression struts to mode1 the flow
of forces from the points of application of the loads to the supports (see Fig. 1.3fa)).

2.

Choose the size of each bearing such that the limiting compressive stress of the adjacent
nodal zone is not exceeded. The nodal zone stresses are limited to 0.85 4, fCf for a CCCnode, 0.75 &, fCt for a CCT-node, and 0.604,f; for a CTT-node.

The tension tie

reinforcement must be distributed over an effective area of concrete such that the force
in the tension tie does not exceed the appropriate stress limit, given above, times this
effective area.
For example, the bearing area of the direct support on the right end of the deep beam
shown in Fig . 1.3(a) is adjacent to a CCT-node, so the area of the bearing plate, 1, b,

must be chosen Iarge enough to ensure that the factored reaction force of the support does
not exceed 0.75 6,fer k b. in addition. the reinforcernent making up the tension tie musc
be detailed such that the effective area surroundhg the bars (defmed as that area having
the same centroid as the tension tie, that is, h,b) is large enough such that the tension in
the tie does not exceed 0.75 d,f: hab.
3.

Draw the t m s rnodel ideaiizing the strut-and-tie mode1 (see Fig. 1.3(b)). All nodes are
tocated at the intersections of the lines of action of suuts, tension ties, applied loads and

bearing reactions.
In order to detennine the line of action of compressive stnits, it is necessary to determine
the dimensions of the struts, such that the compressive stress lirnits are ~ o exceeded.
t
Since the design of deep beams usuaily commences by considering equilibrium at the
location of maximum moment, it is useful to realize that the depth of the horizontal stmt,
da, cm be found by :

where C = T = @&As, of the main tension tie.


The line of action of this strut is located a distance of dJ2 from the top surface of the
beam (see Fig. 1.3). The CCT-node at the right-hand direct support of the deep beam
in Fig. 1-3 is located at the point of intersection of the lines of action of the vertical
bearing force, and the horizontal tension tie. The CTT-node located at the left-hand

hanging (indirect) support in Fig . 1-3 is located at the inters~ctionof the centroids of the
horizonta1 and vertical tension ties.
4.

Calculate the factored forces in the tniss mcmbers (compression struts and tension ties)
through static equilibriurn.

5.

Choose tension tie reinforcement such that the calculated tension force, T, in each tie
does not exceed

A,,

wheref, and A, are the yield stress and cross-sectionai area of

the tension tie reinforcement. Distribute this reinforcement over an effective area of
concrete at least equal to the force in the tie divided by the stress limit of the nodal zone
which anchors it. Figure 1.3(a) shows the effective anchorage ara, h,b, of the CCTnode of the deep beam.
6.

Check the development of the reinforcement. For example, the tension tie reinforcernent
in the deep beam shown in Fig. 1.3(a) m u t be anchored over the length lb so that it is
capable of resisting the caiculated force in the tension tie, T, at the inner edge of the
bearing .

7.

Check the compressive stresses in the stmts. The dimensions of the strut shall be large
enough to ensure that the caiculated compressive force in the strut does not exceed
&faA,,

where fa and A, are the Iimiting compressive stress and effective cross-

sectional area of the strut, respectively.

Equations 1.3 and 1.4 give the limiting

compressive stress in the strut. The iimiting compressive stress, f,, decreases as the

principal tensile strain, E , , increases. The principal strain, E , increases as the angle, O,,
between the strut and the tension tie passing through the strut decreases (see Fig. 1.5).
It is necessary to determine the area, A,,

of each strut. For example, the area of the

diagonal stmt at the intersection with the nodal zone at the right-hand end of the deep
beam in Fig. 1.3(a) equals (lbsins+ h,cos,)b.

This stress must not exceed the limiting

compressive stress in the stmt, f,, as detennined by Eq. 1.3 and 1.4. In caiculatingf,,
the strain in the tension tie passing through the strut,

E,,

rnay be taken as the factored

force in the tie divided by q5,AstEsr, where E,, is the modulus of elasticity of the tension
tie reinforcement.
8.

Choose crack control reinforcement. The CSA Standard and the CHBDC require the
inclusion of uniforrnly distributed reinforcement in the horizontal and vertical directions,
having minimum reinforcement ratios of 0.002 and 0.003, respectively, in order to
control cracking. The maximum spacing of this u n i f o d y distributed reinforcement is
300 mm.

1.5

AC1 Design Approaches for Disturbed Regions

The Amencan Concrete Institute Standard 318-95 "Building Code Requirements for
Structurai Concreteu (AC1 1995) has separate provisions for the design of deep beams and for
the design of brackets and corbels. These two different design approaches are discussed below.

1.5.1

AC1 Provisions for Deep Beams


The AC1 Standard requires that:

where: Vu

factored shear force at the section considered,

Vn

= nominal shear suength at that section, and

= strength reduction factor, taken as 0.85 for shear.

The nominal shear strength is defined as:

where: V,

v~

= nominal shear strength provided by the concrete, and

= nominal shear suength provided by the shear reinforcement.

The AC1 Code defines a member as as deep beam when ZJd is Iess than 5 (where I, is
the clear span measured from face to face of supports). The shear strength, Vn, for deep flexural

memtrers shall not be taken greater than:

V,, r 1 E b , d for i J d < 2


3

At the upper lirnit of IJd = 5, V, 5 0.83 @b,d

(the same for ordinary beams).

The criticai section for shear shall be taken at a distance of O. 15I, from the face of the
support for uniformiy loaded beams, and at a distance of 0.50a (where a is the shear span) from
the support face for beams with concentrated loads, but shall not be taken at a distance greater
than d from the face of the support. Unless a more detaiied caiculation is made in accordance

with Eq. 1.10, V ,shall be computed as follows:

The nominal shear strength provided by the concrete may also be calculated as:

where:

p,

AJb,d

except that the value of the first bracketed term shall not exceed 2.5. and V, shall not be taken
greater than 0 . 5 0 g b W d . Muis the factored moment occurring simuitaneously with Vu at the
critical section defined above.

Where the factored shear force, Vu, exceeds the concrete

resistance, @ V,. shear reinforcement shall be provided to satisw Eq. 1.6 and 1.7, where

V' shall

be computed by:

where: A,

= area of shear reinforcement perpendicular to the flexural tension reinforcement

within a distance S. and


A,

area of shear reinforcement parallel to the flexural tension reinforcement

within a distance s2, and

spacing of shear reinforcement in a direction parallel to the flexuraI tension

reinforcement, and
$2

spacing of shear reinforcement in a direction perpendicular to the flexurai

tension reinforcement.

In addition, the area of shear reinforcement, A,, shall not be less than 0.00 15bwsand s shall not
exceed d/5,nor 500 mm, and A, shall not be less than O.ME5 b,s2 and s2 shalI not exceed 6/3,
nor 500 mm. The shear reinforcement calculated for use at the critical section shall also be used
throughout the span.

1.5.2

AC1 FVovisious for Brackets and Corbels

The AC1 Code limits the applicability of the design provisions for brackets and corbels
to cases where the shear span-to-depth ratio, ald, is not greater than unity. These brackets and
corbels tend to act as t r w e s or deep beams rarher than flexural memben (see Fig. 1.6).
According to the Commentary, the upper limit of uniy for ald is specified because, for larger

shear span-to-depth ratios, the diagonal tension cracks are less steeply inclined and the use of
horizontd stirrups alone as shown in Fig. 1.6@)may not be suitable. Also, the specified method
of design has only been validateci experimentally for ald not exceeding unity , and for a factored
horizontal tensile force, N,

not greater than the factored shear force, Vu. lt is assumed that a

corbel may fail by shearing dong the column-corbel interface, by yielding of the main tension
tie, by crushing or splitting of the compression strut, or by Iocalized shearing or bearing failure
under the loading plate. In order to limit the size and shape of the corbel, it must have a

minimum depth at the outside edge of the bearing area of 0.5d. This limit is specified so that
a premature failcre will not occur due to the propagation of a major diagonal tension crack from
below the bearing area to the outer sloping face of the corbel. The section at the face of the
support shall be designed to resist a shear. Vu,a moment [V,a + Nuc(h-d)]. and a horizontal
tensile force, N,,, simuItaneously. In al1 design calculations, 4 is taken equd ro 0.85 since the
behaviour of corbeis is predominantiy controlled by shear.
The design procedure is suIIlfnanzed in the folIowing steps:
Select the initial geometry of the corbel ensuring that the shear span-to-epth ratio, ald,
does not exceed unity, and that the minimum depth at the outside edge of the bearing
area is 0.5d. Also, the shear strength, Vn,shall not exceed 0.2f:bWdnor 5.5bwd (in
Newtons).
Caiculate the area of shear friction reinforcement, A* across the shear plane necessary
to resist the applied shear force, Vu, as:

where L( = 1.40 for normal weight concrete.


Determine the area of reinforcement, A/, required to resist the moment at the face of
support of the corbel at the level of the p:timary reinforcement.

It is necessary to

estimate the distance ( h - d ) from the top face of the corbel to the centroid of the main
tension reinforcement. The design uitimate moment, Mu,to be resisted is:

tension tie

plane

(a) Strut-and-tie action of corbel

4', (primav reinforcement)

stirmps orties

(b)Detailing of corbel
Figure 1.6 Provisions for brackets and corbels

The area of reinforcement, Al. necessary to resist Mu shall be caiculated in accordance


with the flexural provisions of Clauses 10.2 and 10.3 (AC1 1995) using a capacity
reduction factor, #, of 0.85.
4.

Cdculate the area of reinforcement, A,, required to resist the horizontal tensile force,

N,, frorn:

where # is taken as 0.85. The value of N, shail not be taken less than 0.2 Vu,uniess
speciai provisions are made to avoid tensile forces.
Caiculate the total area of prirnary tension reinforcement, A,, frorn:

Provide a minimum area of primary tension reinforcement. Ensure that p = A,lbd is at


least equal to 0.04Cf,' If, )
Calculate the total cross-sectionai area of horizontal stirrup reinforcernent, A,, as:

Distribute this reinforcement uniformly within two-thirds of the effective depth of the
corbel adjacent to the primary tension reinforcement.

Experiments on Deep Beams, Corbek and Pier Caps


This section provides a bief surnmary of some of the experimental studies that have been
carried out by other researchers investigating the performance of deep beams, corbels and pier

caps.

Figure 1.7 illustrates the marner in which the shear strength reduces as the shear span-todepth ratio, ald, increases. This series of tests were carried out by Kani in the 1960's and are
reported by Kani et al. (1979). The beams in this senes had the same flexural reinforcement and
no shear reinforcement. The two main variables were the shear span and the size of the bearing
plates. Aiso shown in this figure are the predicted capacities (Collins and Mitchell 1991) using
the modified compression field theory and stmt-and-tie models. This figure demonstrates that
for beams with ald less than about 2.5 strut-and-tie rnodels give more accurate predictions. The
1995 AC1 Code provides special provisions for deep flexural rnernbers with clear span-to-depth
rstios, IJd, tess than 5, that is beams with ald l e s than 2.5 (see Section 1.5.1).
Franz and Niedenhoff (1963) used photoelastic mode1 studies to investigate the stresses

in isouopic homogeneous deep bearns before cracking. These beams had a uniformly distributed
load applied dong their top surface and were sirnply supported. Franz and Niedenhoff found that
the srnailer the span-to-depth ratio, the more pronounced the deviation of stress distribution from
that assurned by the Bernoulli hypothesis. For beams with a spart-to-height ratio of one, the
extreme fibre tensile stress can be more than twice that predicted by traditional engineering beam
theory. It has also been demonstrated that the flexural lever a m for the elastic solution is las
than 0.67h, which corresponds to that for a slender beam (Park and Paulay 1975). Furthemore,
the interna1 iever arm for very deep beams does not significantly increase after cracking. For
very deep beams, Franz and Niedenhoff also found that the depth of the tension zone near the
bottom of the beam is relatively srnall (roughly 0.251 thick).
Franz and Niedenhoff (1963) also tested reinforced concrete pier cap specimens which
when inverteci resemble simply supported deep bearns . Two different reinforcing bar layouts
were investigated, one which had concentrateci reinforcing bars representing a tension tie and the
other contained bent-up bars for the main tension reinforcement.

The specimen with the

horizontal bars had a capacity which was 23 % higher chan that with the bent-up bars due to the
larger arnount of tension tie reinforcement at the inside edge of the bearing.
Leonhardt and Walther (1966) carried out experirnents on deep bearns to investigate the
influence of detailing of the reinforcement and the influence of type of loading. They made the
following conciusions and recomrnendations:
1.

The main tension tie reinforcement should be distributed over a depth of 0.25 h -0.051
from the bonom (tension) face, for cases where h 5 f.

152 x 76 x 9.5 mm plate

rn 1 5 2 ~ 1 5 2 ~ 2 5 m r n p l a t e
152 x 229 x flmm plate

strut-and-tie rnodel-\

= 372 MPa
max agg. = 19 mm
d = 538 mm
b = 155 mm
A, = 2277 mm2

sectional mode1

Figure 1.7 Applicability of strut-and-tie model for predicting series of beams tested
by Kani (adapted from Collins and Mitchell 1991)

2.

At least 80% of the maximum calcuiated force in the main tension tie reinforcement
should be developed at the b e r face of the supports.

3.

Small diameter bars of mechanicd anchorages should be used as the main tension tie
reinforcement to prevent premature anchorage faiiure,

4.

A minimum web reinforcement ratio of 0.2% in both directions, as in reinforced concrete


walls, is adequate to conuol cracking. This reinforcement shouid be provided in the
form of s d l diameter bars.

5.

Near the supports, closely spaced horizontal and vertical bars of the same size as the web
reinforcement should be provided.
In their tests of sirnply supported deep beams, the location of the application of load was

varied. For the case of a point load applied on the top surface at midspan, the load path
resembles a ~k3-arch. When a unifonnly distributed load was suspended from the bottom of the
beam instead of being applied to the top (compression) face of the beam, a more severe loading
condition was created. For this case, the load must first be transferred by vertical or inclined
tension reinforcement up to the compression region of the beam before it can be transferred to
the supports by means of tied-arch action. Therefore, verticai stimps must be provided to satisQ
this force requirement as well as to control cracks.
Rogowsky et al. (1986) carried out tests on 7 simply supported and 17 two-span
continuous deep beams.

Variables included: the shear span-to-depth ratio, the flexural

reinforcement ratio, the amount of vertical stirrups and the amount of horizontal web
reinforcement. Two main types of behaviour were observeci. Near failure, beams without
vertical stirrups or with minimum verticai stirmps approached tied-arch action regardless of the
arnount of horizontal web reinforcement present. These beams failed suddenly with Iittle plastic
defonnation, while those with large amounts of vertical stirrups failed in a ductile manner.
The AC1 Code provisions for deep beams (see Section 1S. 1) have been developed based
solely on past experiments of single-span, simply supported deep beams Ioaded on their top
(compression) face. Rogowsky et al. (1986) concluded that the AC1 Code expressions gave
conservative results for the simply supported beams and the continuous beams with large amounts
of vertical reinforcement tested. However, these expressions proved unconservative for the
continuous beams without web reinforcement and for those containing only horizontai web
reinforcement. They determined that the AC1 Code predictions were unconservative due to the
fact that they are based on an incorrect mechmical mode1 for the shear strength of deep beams .
RogowsIq and MacGregor ( 1986) proposed the use of stnit-and-tie models as a more rational and
consistent means of analyzing single- and multiple-span deep beams .

Franz and Niedenhoff (1963) carrieci out photoelastic experiments on corbels having a
shear span-to-depth ratio, ald, of less than 1.0. These experimental studies of the elastic response
of corbels indicated that:
1.

The tensile stress dong the top edge of the corbel is almost constant between the bearing

area and the face of the column.


2.

The compressive stress flowing in from the bottom of the corbel into the c o l m are
almost parailel and resemble a compressive strut.

3.

Rectangular corbels exhibited a nearly stress free zone at the outer-bottom corner of the
corbel.
Franz and Niedenhoff developed a simple truss analogy based on their observations of

the stress trajectories. In addition, they gave the following detailing recommendations:
1.

The prirnary tension reinforcement should be anchored at the outside face of the corbel.
They recornmended providing the main tension reinforcement in the form of closed
hoops .

2.

A minimum amount of compression reinforcement, with appropriate ties to prevent

buckling, should be placed parallel to the compression face of the corbel.


3.

A minimum amount of uniforrniy distributed reinforcement, having an area of at Ieast

25% of that provided by the primary tension reinforcement, should be provided.


In 1964, Kriz and Raths tested 195 corbels, of which 124 were subjected vertical Ioads
alone and 7 1 others were loaded verticaily and horizontally (Kriz and Raths 1965). The variables
studied in these tests included size and shape of the corbel, amount of main tension tie
reinforcement and its detailing, concrete strength, ratio of shear span to effective depth, and ratio
of horizontal to vertical loading. Kriz and Raths gave the foltowing recomrnendations:
1.

The ratio of the arnount of main tension reinforcernent to the gross cross-sectional area

2.

of the corbel should not be l e s than 0.004 in order to control cracking.


A cross bar should be welded to the main tension tie reinforcement near each end in
order to provide proper anchorage (see Fig. l.b(b)).

The size of this cross bar should

be at least equal to the largest bar used in the main tension tie reinforcement, and it
should be located as near to the outer face of the corbel as cover requirements permit.
3.

Closed horizontal stimps shouid be provided having an area not less than 50% of that
provided by the main tension tie reinforcement. These stirnips shouId be uniforxnly
distributed throughout the upper two-thirds of the effective depth of the corbel.

4.

The total depth of the corbel at the outer edge of the loading plate shoitid be at least
equal to one-haif the depth of the corbel at the colurnn face.

5.

The outer edge of the bearing plate shouid be at least 50 mm from the outer face of the
corbel.

6.

When corbels are designed to resist horizontal forces, the steel bearing plates should be
welded to the main tension tie reinforcement to transfer the horizontal force directly to
these bars (see Fig . I .6(b)).

7.

Bearing stresses at ultimate load should not exceed OS$.


Mast (1968) inuoduced the "shear-friction" concept for the design of corbels. His goal

was to develop a simple rational approach based on physical models of behaviour which could

be used in the design of a number of different concrete connections. The approach assumes
nurnerous failure planes for which reinforcement m u t be chosen to prevent failure dong these
planes.
The shear-friction concept assumes that a crack interface has some roughness and hence.
as shear is applied, the defocmations include not oniy some shear displacernent dong the crack

interface, but also some widening of the crack.

The crack opening causes tension in the

reinforcement crossing the crack which is balanced by compressive stresses in the concrete across
the crack interface. The shear on the interface is assumed to be related to the compression across
the interface by a coefficient of friction,

p,

which depends on the roughness of the interface

surface. The nominal shear capacity is thus given as:

In detennining the shear strength, Mast made the following assumptions:


1.

The reinforcement crossing a crack is sufficientiy anchored such that the bars can yield.

2.

The cohesive strength of concrete is negligible.

3.

The effective coefficient of shear friction, p, depends on the surface roughness but is
independent of concrete strength.
This concept was applied to test data reported by Kriz and Ratlis where the shear span-to-

depth ratio, ald, was less than or equal to 0.7 and where the reinforcement had yielded. The
shear-friction concept gave reasonably conservative strength predictions for both vertical and
combineci vertid and horizontal loading cases.
Mattock et al. (1976) tested 28 reinforced concrete corbels subjected to vertical and
horizontal loading. The variables included in these tests were: the ratio of shear span to effective

depth, the ratio of horizontal to vertical load, the maunts of main tension tie and distributed
reinforcement, concrete strength and type of aggregate.
The design procedure first introduced in the 1971 AC1 Code was based on the research
of Kriz and Raths (1965) with later modifications to include the design procedure developed by
Mattock (1976)and Mattock et al. (1976). This approach is still in use today (see Section 1-5.2).

1.6.3

Pier Caps
Al-Soufi (1990) carried out an experimental investigation which involveci testing six

reinforced concrete pier caps. Parameters which were varied in these specimens included: the
geometry of the pier caps, the amount and distibution of unifonnly distributed reinforcement,
and the anchorage details of this reinforcement.

He made the following conclusions and

recornmendations:
After yielding of the main tension tie reinforcement, yielding spreads to the distributed
reinforcement.
The unifonniy distributed reinforcement contributes significantly to the strength and plays

a key role in controlling cracks.


Standard 90" end hooks rnay be provided to anchor the reinforcement of the main tension
tie provided that these bars can develop their yield force at the inner edge of the bearing
plates.
The unifonnly distributed horizontal reinforcement may be provided in the form of Ushaped stirrups properly lap spliced over the central region of the pier cap.

The

uniformly distnbuted vertical reinforcement rnay be provided in the f o m of closed


stirrups or lap-spliced U-shaped stirnips.
The column reinforcement, which was extended into the pier cap, provided additional
horizontal (column ties) and vertical reinforcement in the central region of the pier cap.
This additional reinforcement controlled cracks and provided some confiement for the
lap splices of the uniforrnly distriburai horizontal reinforcement.
Stnit-and-tie models provided a usehl tool for evaluating the strength of these pier caps,
while non-linear finite element analysis provided a means of predicting behaviour at service load
Ievels (see Section 1-7.2).

1.7

Detailed Analysis Procedures


This section discusses more detailed anaiysis procedures, including refined stnit-and-tie

modelling, and non-Iinear finite element analysis.

1.7.1

Refined Strut-and-Tie Modds


Simple strut-and-tie models typically assume that the compressive struts can be

represented by straight lines between loading and support bearhg areas, and usuaily ignore the
contribution of uniformiy distnbuted reinforcement. More refmed strut-and-tie models attempt
to include the effects of bulging and curving compressive stnits due to the presence of tensile
stresses in the concrete and uniformiy distributecl reinforcement. Accounting for the presence
of unifonnly distributed reinforcement also increases the total amount of tension tie
reinforcement, and hence the strength of the member.
Figure 1.8 shows a simply supported deep bearn with a concentrated load applied on its
top surface. In Fig. 1.&a), the flow of principal compressive and tensile stresses are indicated
with dashed and solid lines, respectively. The diagonai compressive struts buige between the
loading point and the supports due to the presence of tensile stresses in the concrete. A possible
strut-and-tie mode1 which acccJunts for this bulging action of the struts is presented in Fig. 1.8(b).
Design procedures have typically adopted a simpler assumption of straight compression struts in
combination with a minimum arnount of reinforcement uniforxniy distributed in the horizontal and
vertical directions as s h o w in Fig. 1.8(c). This uniformly distributed reinforcernent serves to
control cracking in disturbed regions (see Section 1.4).
In deep bearn design, unifonnly distributed reinforcement corresponding to geornetnc
ratios of 0.002 to 0.003 is usually provided in the horizonta1 and vertical directions. Marti
(1985) investigated the role of this reinforcement in controllhg cracks, perrnitting redistribution
of stresses after cracking and increasing the strength of the member. Figure 1.9 shows one-half
of a deep bearn which is subjected to a concentrated load applied on its top surface and with
simple supports at its ends. It is assumed that this beam has u n i f o d y distributed reinforcemnt
in the verticai direction ody. Figure 1,9(a) illustrates the arching and fanning of the compressive
stresses in the member. The presence of the unifonnly distributed vertical reinforcement perniits
the main compression stmt to curve, thus forming an arcb, and also provides anchorage for the
fanning compression s m t s near the bottom of the beam. Figure 1.9(a) shows the variation of

the force in required in the main tension tie reinforcement. The main tension tie has a maximum
force at the centre-line of the beam (point d) corresponding to the tensile force, T. In the region

crack

cantrol
sted

(a) Flow of principal


stresses

(b) Required tension


ties

VV

I I

(c) Assumption of straight


compressive stnits

Figure 1.8 Crack control reinforcement required with assumption of straight-line


compressive stnits (adapted from Schlaich et al. 1987)
effective distnbuted
vertical reinforcement

lie representing distributed

vertical reinforcement

(a) Arch and fan action

(b) Strut-and-tie model

Figure 1.9 lnvestigating the effect of distributed reinforcernent


on deep beams (adapted from Marti 1985)

between points b and d, that is where uniformly distributed vertical reinforcement is present, the
force in the longitudinal reinforcement changes as shown in Fig. 1.9(a). in the regions between
points a and b and between d and e the force in the main tension tie remains constant. As
pointed out by Marti (1985), the vertical distributed reinforcement allows curtailment of some
of the longitudinal tension tie reinforcement. Figure 1.9(b) shows the idedized strut-and-tie
model for this deep beam containing uniformly distributed vertical reinforcement. The uniformiy
distributed vertical reinforcement has been idealized as a tension tie at the centre of zone

This idealkition permit5 the compressive arch and


compressive fanning to be represented as shown in Fig. 1.9(b). Also shown is the variation of
force in the longitudinal reinforcement as predicted by this refined strut-and-tie model.
containing vertical reinforcement.

While uniformiy distributed vertical reinforcement reduces the demand on the main
tension tie reinforcement close to the support region, its presence does not result in increased
member strength. This is due to the fact that the strength is controlled by the conditions at
midspan. The presence of uniforxniy distributed horizontal reinforcement assists the main tension
tie reinforcement and resuIts in increased strength.
Although the simple strut-and-tie models are very useful in design and give conservative
strength predictions, for detailed analysis of the strength a more refmed stmt-and-tie model,
including both the vertical and horizontal distributeci reinforcement, gives more accurate strength
predictions.

As was mentioncd in Section 1.2, it is not appropriate to design disturbed regions with

the usual beam theory assumptions. Elastic finite element analysis may be used to determine the
stresses in a reinforced concrete member prior to cracking, however this type of analysis may not
be appropriate for predicting stresses in a cracked member as significant redistribution of stresses
occurs after cracking. In order to predict the full response (including post-cracking response) of
reinforced concrete members a computer program, FIELDS, was developed (Cook 1987, Cook
and Mitchell 1988) which combines two-dimensional non-linear finite etement analysis with the
compression field theory (Collins and Mitchell 1980, 1986, and Vecchio and Collins 1986).
Triangular and quadrilateral elements are used to mode1 the reinforced concrete member.
To account for significant non-linearities which may arise within a finite element, up to four-byfour Gauss quadrature may be chosen for an element. Figure 1.10 filustrates the method used
to evaluate stresses correspondhg tcb a state of strain at each Gauss point (Cook and Mitchell

(a) Element with 4 x 4 quadrature

(b) State of strain at a single Gauss point

( c ) Deterrnining stresses at a Gauss point corresponding to a strain state

Figure 1. I O Evaluating stresses at Gauss points in quadrilateral elernent


(Cook and Mitchell 1988)

1988). n i e principal tensile strain, cl, the principal compressive strain.

direction, E - ~ ,the strain in the y-direction,

E,,

the strain in the x-

5, and the principal compressive strain direction, B.

are inter-related by the requirements of strain compatibility (see Fig . 1.1O@)).


The average steel stresses, f, andf,, at a Gauss point can easily be detennined by using

the stress-strain relationships of the reinforcing steel. However. the average stresses in the
cracked concrete, f,,and f,, are not as easy to determine. The average principal campressive
stress, f,, is not only a huiction of the principal compressive strain, q,but is also dependent on
the principal tensile strain, c l . As

E,

increases f, decreases; this effect is known as strain

softening. Combining the Iimiting compressive stress for cracked concrete developed by Vecchio
and Collins (1986) and a parabolic concrete stress-strain curve gives the compressive stress-strain
relationship for cracked concrete (see Fig. 1.4(a)) as:

where:

and

c:

strain in the concrete at peak compressive stress.

After cracking, the principal tensile stress in the concrete varies from zero at a crack location to

a maximum between cracks. Figure 1.11 shows the average principal tensile stress. fcl, plotted
against the principal tensile strain, E , (Vecchio and Collins 1986) as:

where: E,

initial tangent modulus of the concrete,

E c ~

strain in the concrete at cracking, and

Lr

= concrete cracking stress = Ececr.

Figure 1.12 shows that the average principal tensile stress may be lirnited by yielding of
reinforcement or by sliding dong the crack interface (Vecchio and Collins 1986). Between the

Figure 1.11 Determining average concrete tensile stress,^, , from strain, E,


(Vecchio and Collins 1986)

(a) Cracked reinforced


concrete element

(b) Transrnitting shear


across crack interface

(c) Average stresses


between cracks

(d) Stress condition at


crack interface

Figure 1.12 Investigating stress condition at crack interface


(Vecchio and Collins 1986)

cracks, the concrete and the steel are assumed to have average values of stress (see Fig . 1.12(c)),

while at a crack the tensile stress in the concrete is zero, the steel stress is a maximum, and a
shear stress v, may exist at the crack interface (see Fig. l.l2(d)). An approximate expression
for the shear stress limit dong a crack has been developed (Vecchio and Collins 1986) based on
the interface shear transfer tests conducted by Walraven (1981).

This expression can be

simplifieci as:

where: w
a

= average crack width in


=

mm,

maximum aggregate size in mm.

Note that this is an empirical expression and that stresses are expressed in MPa units. The
average crack width c m be assurneci to equal the average crack spacing tirnes e,. Since the stress
States shown in Fig. 1.12(c) and (d) are statically equivalent, it is possible to determine whether
yielding of the reinforcenent across the crack (Le. ,
f
interface (i.e. v, equals v,-)

or f,, equals 4)or sliding at the crack

will result in a value off,, less than thac given by Eq. 1-20.

Examples of the application of non-linear f ~ t element


e
analysis applied to deep beams.
corbels, dapped end beams and anchorage zones are given by Cook and Mitchell (1988) and
Collins and Mitchell (1991).

1.8.1

Compressive Strength
Advances in concrete technoIogy over the past two decades have resulted in the

availability of ready-mixed concrete with compressive strengths as high as 100 MPa in several
North American cities. There is a need to investigate whether the design proceciures, developed
for use with normal-strength concretes, are applicable to the full range of high-strength concrets
currently available (Collins et al. 1993).
The traditional parabolic stress-strain curve for normal-strength concrete recommended
by Hognestad (1957) can be expressed as:

This formula provides a reasonable approximation to the stress-strain curve for nomial-strength
concrete. However , as concrete strength increases, the compressive stress-strain cuve is near
linear over the rising branch, and exhibits greater initial stiffkess and decreased ductility (see Fig .
1-13). The parabota is too " rounded" to accurately represent this increased linearty and the more
bnttle post-peak response of very high-strength concrete.
Thorenfeldt er al. (1987) introduced a post-peak decay term, k, to the stress-strain
relationship developed by Popovics (1973) such that it could be applied to wide range of concrete
strengths. This resuited in the following expression:

= compressive stress,

where: fc

= maximum compressive stress,

fc'
Cc
I

= compressive strain,
= compressive strain when fc reaches fCt,

= cuve fitting factor, as n becomes higher. the rising portion of the curve

becomes more Iinear, and

= post-peak decay term. taken as 1 when

&

E:

is Iess than 1, and taken greater

than 1 when e C E,' exceeds 1.


Equation 1.23 gives fc as a function of

tc and

involves four constants, namely. fct, rct, n

and k. While these four constants can al1 be determineci from actual cytinder stress-strain curves,
in rnany design situations, only the cylinder strength, f:, is laiown. Collins and Porasz (1989).

and Collins and Mitchell (1991) suggest that for J e C r > 1:

Figure i.13 Influence of concrete strength on shape of stress-strain curve

where Er

= modulus of elasticity of concrete.

The 1994 CSA Standard gives an expression for E, (in m a ) , for concrete with y, between 1500
and 2500 kg/m3, as follows:

where y,

= concrete density (in kg/m3)

From the value off,' the four constants in Eq. 1.24 through 1.27 can be used to develop
the stress-strain relationship given in Eq. 1.23.

1.8.2

Nexure and Axiai ha&


The new rectangular stress block factors, a land

PI, of

the 1994 CSA Standard are

suitable for a wide range of concrete compressive strengths. It is assurneci that a concrete stress
of al$cfc'is uniformiy distributed from the extreme compression fibre into the member a
distance of 8, c, where c is the distance of the neutrai axis from the extreme compression fibre.
These factors now depend on the concrete cylinder strength as follows:

The new factors are intended to account for both the signifiant shape change in the stress-strain
cuve as the concrete strength increases and the difference between the cylinder strength and the

in-situ concrete strength.

1.8.3

Minimum Reinforcement for FIexure and Shear

The 1994 CSA Standard requires a minimum arnount of fiexural reinforcement in order
to give adequate reserve of strength after cracking and hence provide a ductile response. The
1994 CSA Standard requires that one of the foiiowing three provisions be satisfied:
1.

The factored moment resistance, M,, must be such that:

2.

Except for slabs and footings, provide a minimum area of flexural reinforcement, A
,,
as foIlows:

where: b,
3.

= width of tension zone of member.

The minimum reinforcement requirements given above may be waived provided that the
factored moment resistance, M,, is at least one-third greater than the factored moment,

Mr.
The 1994 CSA Standard also requires a minimum arnount of shear reinforcement which
is dependent on concrete strength. An increase in the concrete compressive strength leads to an
increase in the tensile strength, which in turn results in an increase in the cracking shear. This
increase in the cracking shear requires an increase in the minimum shear reinforcement in order
to ensure that the shear strength exceeds the cracking shear. The 1994 CSA Standard requires

a minimum area of shear reinforcement, A,, as follows:

where: b,
s

=
=

minimum effective web width,


spacing of shear reinforcement.

This requirement, together with the maximum spacing limits for shear reinforcement, is intendeci
to control inclineci cracking at service load levels.

1.8.4

SW-and-Tie Provisions
The provisions for strut-and-tie models in the 1994 CSA Standard are the same as those

in the 1984 CSA Standard. The stress limit for a concrete strut. 0.85+,fc'. remains a linear

function of the concrete cylinder strength. MacGregor (1997) has introduced a factor. v,, to
account for the infiuence of high-strength concrete (see Tables 1.1 and 1.2).
The 1994 CSA Standard and the 1997 CHBDC provisions for suut-and-tie models require
minimum arnounts of uni forrnly distributeci reinforcement which are independent of concrete

strength (see Section 1.4).

1.9

Crack Widths and Crack Spacing

Concrete cm only withstand srnall tensile strains before it cracks. As these cracks do not
form at equai spacings, crack widths may Vary in size. and it is therefore appropriate to define
the mean crack width, w,,,, as:

where: s,,,
cf

= mean crack spacng , and


=

strain in the concrete caused by stress.

The characteristic crack width (the width which only 5% of the cracks will exceed), w,, is
approximated by the CEB-FIP Model Code (CEB 1990) as w, = 1.7 w,. The CEB-FIP Model
Code (CEB 1990) gives the following expression for the mean crack spacing:

where: c
S

= clear concrete cover,


= maximum spacing between longitudinal reinforcing bars, but shall not be taken

as greater tiian 15d, (where d, is the diameter of the reinforcing bar),


Pd

= Ai*c,g

As

= area of steel considered to be effectively bonded to the concrete,

4.4

area of the effective embedment zone of the concrete where the reinforcing

bars c m influence crack widths (see Fig. 1.14(a)),

15 d,

4= shaded area

(a) CEB-FIP expression


/

neutral axis

shaded area

\tension
face

(b) Gergely-Lutz expression


Figure 1.i4 Crack width parameters

neutral
axis
skin reinforcernent

cross-section

elevation

Figure 1.15 Side-face cracks controlled by skin reinforcement

= coefficient that characterizes bond properties of reinforcing bars,

ki

= 0.4 for defonned b a s


=

0.8 for plain bars

= coefficient to account for the strain gradient.

k2

= 0.25 (t,

+ c 3 / 2 e,. where e,

and c2 are the largest and srnallest ten

in the effective embedment zone, respectively.


The Gergely-Lutz expression (Gergely and Lutz 1968) estirnates the maximum crack
width as:

where: 6

= factor accounting for strain gradient,

= 1.0 for unifom strains


= h2/h, for varying strains, where h,

is the distance of the main tension

reinforcement from the neutral axis and 4 is the distance of the extreme tension
fibre from the neutral axis
S.

cr

= strain in a reinforcing bar at a crack Iocation,

= effective area of concrete surrounding each bar, taken as the total concrete

distance frorn the extreme tension fibre to the centre of the closest bar, and

area in tension, which has the sarne centroid as the tension reinforcement,
divided by the number of reinforcing bars (see Fig. 1.14(b)).

In the Gergely-Lutz expression, the strain in the reinforcement at a crack is taken as:

where: N

ES

= applied axial tension, and


=

modulus of elasticity of steel.

The CEB-FIP Mode1 Code, (CEB 1990) lirnits crack widths to 0.30 mm for structures
exposed to both frost and de-king conditions. The 1995 AC1 Code and the 1994 CSA Standard
require the calculation of a crack width parameter, z , to detennine if the crack widths would be
within acceptable limits. This crack width parameter is based on the Gergely-Lutz expression
(see Eq. 1.35) and is given as:

where: f,

= calculated stress

in reinforcement at specified loads, may be taken as 0.64.

The 2-factor is lirnited to 30,000 Nlrnrn for interior exposure and 25,000 Nlmm for exterior
exposure. These limits correspond to maximum crack widths of 0.40 and 0.33 mm, respectively.
If epoxy-coated reinforcement is used the CSA Standard requires multipIication of the limiting
crack width parameter, z, by a factor of 1.2, based on the research of Abrishami et al (1995).
Figure 1.15 illustrates the requirement for skin reinforcement in the 1995 AC1 Code and
the 1994 CSA Standard for members with an overall depth, h, exceeding 750 mm. The required
longitudinal skui reinforcement shall be unifonniy distributed dong the exposed side faces of the
member over a depth of O S h - 2(h -6)fiom the principal reinforcement (see Fig. 1.15). The total
area of such reinforcement shall be p&,

where A, is t h e sum of the area of concrete in suips

dong each exposed side face, each strip having a height of 0.5h - 2(h -4 and a width of twice the
distance from the side face to the centre of the skin reinforcernent but not more than hdf the web
width. The minimum amount of skin reinforcement shall be such that p, equals 0.008 or 0.0 10
for interior or exterior exposure, respectively. The maximum spacing of this skin reinforcement

is 200 mm.

Research Objectives
The objectives of this research programme are:

to study the complete behaviour of full-scale reinforced concrete cantilever cap


bearns ,
to investigate the suitability of current design approaches for these disturbed
regions,
to compare the predicted responses using simple strut-and-tie models, refined
strut-and-tie models arid non-linear finite element analyses,
to investigate the influence of concrete strength on the behaviour of large
cantilever cap beams, and
to study the amount of uniformly distributed reinforcement required for crack
control at service load Ievels.

Cl3APTER2

EXPERIMENTAL PROGRAMME

Two full-scaie cantilever cap beams were constructeci and tested in order to study their
complete responses. These test specimens are representative of the cantilever portions of
continuous cap beams and of cantilever cap beams as shown in Fig. 2.1. These cantilever cap
bearns were designed using the strut-and-tie approach of current codes (CSA 1994, CHBDC 1996
and AASHTO LFRD L994). The arnounts of uniformly distributed horizontal and vertical

reinforcement were varied in order to study their influence on crack conrol at service load levels.
The geometry of the cantilever cap beams was chosen after snidying a number of
drawings of typical cap beams (see Fig. 1.1). The loads at each bearing location for the
prototype bridge investigated were a service dead load of 460 kN and a service dead plus Iive
plus impact Ioad of 1140 kN.

2.1

Details of Specimens
Cap beam specimen CAPN was cast with normal strength concrete (design fc' = 35 W a ) ,

while specimen CAPH was constmcted with high-performance concrete (design fc' = 70 MPa).
Both specimens have identicd geometries. As shown in Fig. 2.2, each cap beam is 3350 mm
long, 750 mm wide and has cantilevers which extend 1300 mm from the faces of the 750 mm
square columri. The depth of each cantilever is 900 mm at its end, increasing to 1100 mm over
a distance of 625 mm from the end.
The reinforcement for both specimens was identicai, with epoxy-coated bars used
throughout to conform to the requirements of the Canadian Highway Bridge Design Code
(CHBDC 1996) for corrosive environrnents. The main tension tie reinforcement was provided
by two layers of reinforcement, each containing 5 No.25 bars, with a clear vertical spacing of

35 mm. One layer of 5 No.25 bars served as the compression steel. The square colurnn was

reinforceci with 12 No.25 bars and confued by sets of 3 No.10 colurnn ties spaced at 300 mm
(see Fig. 2.2). The specimens had crack control reinforcement ratios of 0.18% and 0.30% in
cantilever ends A and B, respectively (see Sections A-A and B-B). In end A, this reinforcement

(a) Prototype continuous cap beam

(b) Prototype cantilever cap bearn

(c) Test setup

Figure 2.1 Test simulation of cantilever cap beams

7 4 No. 10
/double stirrups
(vertical
distributed
reinforcement)
s = 300

double stirrups
(vertical
distributed
reinforcement)
s =l?S

2'
- NO. 15
U-shaped bars
(horizontal
distnbuted
reinforcement)
s = 295

4 No. 15U-shaped bars


(horizontal
distributed
reinforcement)
s =l?

Section B-B

Section A-A

10 -No.25
(tension
reinforcement)

12 No. 25
(column bars)

3 -NO. I O
(colurnn ties)
s = 3G0

Notes:
dimensions in mm
minimum cover = 50 mm

Figure 2.2 Specimen details

was provided by 4 double No.10 stirrups spaced at 300 mm in the vertical direction, and 2

U-shaped No. 15 bars spaced at 295 mm in the horizontal direction. These spacings were reduced
to 175 mn for the 7 vertical double stirnrps and 177 mm for the 4 horizontal crack control bars
in end B. A minimum cover of 50 mm was maintained &oughout the specimens.
Ninety-degree end anchorages with free end extensions of 300 mm beyond the bend were
provided on al1 the No. 25 bars used for the main tension tie reinforcement. In order to fully
develop the reinforcement (fy = 400 m a ) , the code (CHBDC 1996) requires straight embedment
lengths, l&, of 430 mm and 304 mm beyond the hooks for specirnens CAPN and CAPH,
respectively. The tension development length, ld, of the No. 25 bars in CAPN is determined as:

where:

bar location factor, taken as 1.O.

k,

k2

= coating factor, taken as 1.2 due to the epoxy-coated bars.

k,

= bar size factor, taken as 1.0 because bars are larger than No. 20.

Likewise, ld = 782 mm for the No. 25 bars of CAPH. The stress in the bar that can be
developed by the hook is [(Il06 - 430)/1106] * 400 MPa = 244 MPa for specimen CAPN and

[(782 - 304)/782] * 400 MPa = 244 MPa for CAPH. Knowing the geometry of the bend and
the placement of the bearing pads (see Fig. 2.2). the avaiIable straight bar embedment length to
the inner edge of the bearing plate is 286 mm for the bottom Iayer and 226 mm for the bottom
layer of bars for CAPN. Therefore, the bottom Iayer of bars in CAPN is capable of developing
a stress of 244 MPa + 286/1106 * 400 MPa = 348 MPa, while the top bars can develop 326
MPa. SimiIarly the bottom and top bar layers of CAPH can develop 371 MPa and 340 MPa,
respectively. These calculations assume that the bond stress is uniform over the development
tength, which results in a linear build-up of stress aiong l& A1thoug.h stresses greater than 400
MPa are expected during testing, these smaller embedment lengths were provided to investigate
the beneficial effects of the compressive bearing stress on the bond strength.
The horizontal distributed steel was lap spliced in the central regions of the specirnens
where additiod confinement is provided by the column ties which are typically continued into
the cap bearn. Without considering the beneficial effects of the confinement provided by these

column ties, the required lap splice length is calculateci as 1.3 1, (CHBDC 1996). where:

Hence the required lap splice Iength for specimen CAFN is 1.3 x 531 mm = 690 mm.
Sirnilarly, the required lap length for specimen CAPH,having a design compressive strength of
70 MPa is 488 mm. Full development of the horizontal bars was therefore achieved in the
central region of the cap beam. Over the constantdepth porons of the cap beam the vertical
uniformiy disuibuted reinforcement was provided by No. 10 double closed stimps (see Fig. 2.2).
Because of the changing depth of the cap beam near its ends, it was necessary to use double Ushaped spliced stimps over the tapered portions of the specimem. The required lapsplice length
for these U-shaped stimps, using Eq. 1.1, is 460 mm for CAPN and 340 mm for CAPH. A
conservative value of 460 mm was used for both specimens.

2.2

Material Properties

2.2.1

Concrete
Both specimens were cast with ready-mix concrete. The specified concrete strength for

CAPN was 35 MPa with a water to cernent ratio (wlc) of 0.40 and a maximum aggregate size
of 14 mm. The high performance concrete of CAFH had a specified strength of 70 MPa, a wlc
of 0.28, and a 10 mm maximum aggregate size. Mix designs are presented in Tables 2.1 and
2.2, and the slump and air content measurernents taken upon delivery are shown in Table 2.3.
The test specimens, together with the controI cylinders and flexural beams, were covered with
wet burlap and plastic sheeting a few hours after casting, and were kept moist. The test
specimens and the control specimens were stripped of their formwork 4 days after casting and
kept in the sarne air-cured conditions of the laboratory. The compressive strengths were
detennined fiom the resuits of testing 3 standard, 150 m m diameter by 300 mm long, concrete
cylinders, and the splitting tensile strengths were taken as the average fiom 3 Brazilian tests on
150 mm 4 by 3 0 mm cylinders. in addition, 3 flexural beam tests were used to determine the
average modulus of rupture. These flexural beam specimens measured 150 x 150 x 600 mm and
were subjected to third-point loadrg over a span of 450 mm. A sumrnary of the results of the
cylinder and beam tests are presented in Table 2.3. Representative compressive stress-strain
curves for the 35 MPa and 70 MPa concretes are shown in Fig . 2.3(a). In addition, shrinkage
strains were determinecl from externaily applied strain targets on concrete beam specimens
measuring 100 x 100 x 400 mm. The strain targets were placed on these shrinkage specimens
24 hours after casting. The shrinkage strains determineci From these measurements are shown

Ir

cernent

415 kg/m3

10

fine aggregate

Lafarge St.-Gabriel

10-14 mm limestone

coarse aggregate

757 kg/m3

1003 kg/m3

water (fotal)

167 kg/m3

total density

2342 kg/m3

I
I
I

superplasticizer

1
1
1

air-entraining agent

Rheobuild 1 0 0

retarder

pozzolith 1 0 0 ~ ~

water reducer

Pozzolith 2ON
Micro- Air

1
1
1

1346 mL
330 mL
2.7 L

375 mL

Table 2.1 Mix design for 35 MPa concrete

- --

cernent

1
I

480 kg/m3

lOSF

IL

fine aggregate

Lafarge St.-Gabriel

850 kg/m3

coarse aggregate

10 m m lirnestone

1015 kg/m3

1
1

water (total)

135 kg/m3

totai density

2480 kg/m3

water reducer

Pozzolith 2ON

1630 mL

superplasticirer

Rheobuild 1 0 0

L3.0 L

retarder

Pozzolith lXR

780 rnL

Table 2.2 Mix design for 70 MPa concrete

in Fig . 2.3(b). It is interesthg to note that the 70 MPa concrete exhibited considerably higher
shrinkage strains in the first few days after casting than the 35 MPa concrete.

microstrain
(a) representative stress-strain curves

35 MPa

A--

50

100

150

tirne (days)
(b) average shrinkage strains
Figure 2.3 Concrete Properties

200

250

air content (% )

I
r

average 28-day sumgth @Pa)

6.5

2.3

36.1

72.8

2.18

3.08

age at testing (days)


--

secant modulus (GPa)

characteristic peak suain ( d m )

average

37.6

average

3.2

modulus of rupture at testing

average

4.6

6.7

(MW

std. dev.

0.3

0.3

compressive strength at testing

79.2

(MW
splitting tensile strength at testing

5.2

(MW

-. -

..

Table 2.3 Concrete properties

2.2.2

Reinforcing Steel
Steel reinforcement consisted of No. 10, No. 15 and No.25 epoxy-coated deforrned bars

with a specified grade of 400 MPa. A minimum of 3 tensile samples were testeci for each bar
size to determine their mechanical properties. Table 2.4 summarizes the average values and the
standard deviation of the yield and ultimate stresses and strains at strain hardening, strains at the
dtimate stress and the rupture strains. Figure 2.4 shows typical stress-straincurves for the three
different bar sizes. The modulus of elasticity for d l reinforcing steel has been taken as 200 GPa

for the purpose of both design and anaiysis.

0.02

0.06

0.04

0.08

0.1

0.12

0.14

0.12

0.14

0.16

strain (mm/mrn)

(a) No. 10 bars

strain (mm/mrn)

(b) No. 15 bars

--

I UY

0.02

0.04

0.06

0.08

0.1

0.16

strain (mm/mm)

(c) No. 25 bars

Figure 2.4 Typical stress-strain responses of reinforcing bars

average

1
1

std. dev.

R O P ~

441.0

1
1

2.7

No. 10

419.0

1
1

1.0

1.12

OS4

No. 15

No. 25
468.5

7.1

strain at strain

average

0.77

hardenhg ( %)

std. dev.

0.04

0.02

0.06

ultimate stress

average

707.5

688-4

768.6

std. dev.

2.2

2.9

L.2

strain at ultimate stress

average

13.4

12.0

10.2

(w)

std. dev.

0.2

0.5

O. 1

rupture strain,

average

14.1

over 200 mm ( % )

std. dev.

16.0
1

0.6

13.5
1

1.1

0.6

Table 2.4 Reinforcing steel properties

2.3

Test Setup and Instrumentation


The pier caps were installeci upsidedown under the 11,000 kN capacity MTS universai

testing machine (see Fig. 2.5). Figure 2.6 shows the loading arrangement at the top of the stub
column for each specimen. Load was transferred through the 559 mm diameter bottom platen
of the MTS's spherical seat, which in turn loaded two plates. These plates had a total thickness

of 127 mm in order to ensure sufficient spreading of the load. The 51 mm thick bottom plate
rneasured 660 x 660 mm, and was seated with plaster to the top of the 750 x 750 mm stub
column. The size of this bouom plate was chosen such that the load wodd be transrnitted to the
vertical column bars without loading the concrete cover.
The cap beams were simply supported on the laboratory strong floor. Figure 2.6 shows
the bearing details used for specimens CAPN and CAPH. Two 20 mm thick by 152 mm wide
by 600 mm long bearing plates were seated with a plaster mortar compound on the bottorn of

CAPN. The bearing plates for specimen CAPH had a width of 76 mm, that is, one-half that
provided for CAPN due to the higher concrete compressive strength of specimzn CAPH. These
600 mm long bearing plates were centred across the 750 mm wide cap beams such that they did
not bear on the cover concrete. The centre of the bearings was Iocated 375 mm from the end
faces of the cap beams (see Fig . 2.6). The bearing plates rested on a rocker, with a radius of

Figure 2.5 Specirnen CAPN under the MTS testing machine

Notes;
dimensions in mm
I, = 152 mm for CAPN
I, = 76 mm for CAPH

Figure 2.6 Different bearing details of specimens CAPN and CAPH

250 mm, which in tum rested on two 152 mm diamerer rollers sandwiched between two 76 mm
thick steel plates.
Three Linear Voltage Differential Transducers (LVDT's) or extensometers were installecl
to m u r e vertical displacements of the specimen at the supports and at mid-span (see Fig. 2.7).
The centre deflection reporteci is taken as the deflection from the LVDT at midspan (CV),minus
the average of the LVDT's at ends A and B (AV and BV), in order to remove rnovements at the

In addition, extensometers were used to determine average strains in the test


specimens. These LVDT's were attached to threaded rods which, in turn, were glued in holes
drilled 50 mm into the concrete. Ten extensometers were positioned, as shown in Fig. 2.7, at
the level of the centroid of the tension steel, between the centres of the bearhgs. A second Iine
of eight extensometers was located at nid-height of the cap beam. Additional extensometers were
positioned, as shown if Fig. 2.7, to fonn 260 mm rosettes within the shear spans of the cap
beam. The purpose of these rosettes was to determine principal strains and their directions in the
diagonal compressive stnits.
supports.

Figure 2.8 shows the locations of the twenty-two electricd resistance strain gauges which
were glued to the reinforcing bars prior to casting. Twelve gauges were Iocated on the bonom
layer of the main tension reinforcement, 6 on the centre and 6 on the outerrnost bar. These
gauges were positioned at the start of the hooks, at the inner edges of the bearing plates, and at
locations aligned with the colurnn faces (see Fig. 2.8). An additional ten gauges were glued to
the h o ~ o n t a and
i vertical distributed bars in the shear spans of the cap beams as shown.

2.4

Testing Procedure
The loading was displacement controlled at a rate of approximateiy 0.004 mrnkec.

Throughout the testing, load, displacement and strain readings were recorded at intervals of 25

khi or 0.1 mm, whichever came first. During the early stages of a test, load application was
haited, to create major load stages, at increments of approximately 500 kN. At these load stages,
widths of cracks crossing the horizontal lines shown in Fig. 2.8 were measured using a crack
comparator, and the crack patterns were sketched and photographed. Afier yielding of the main
tension reinforcement, load stages were taken at increments of 1 to 2 mm of the mid-span vertical
deflection. The loading of the specimens continueci until a significant decrease in load-carrying
capacity was observed.

Note:
dimensions in mm

Figure 2.7 LVDT locations

crack widths

measured on

figure 2.8 Strain gauge locations and crack width lines of measurernent

CHAPTER 3

EXPERIMENTAL RESULTS

This chapter describes the experimentai results of each specimen. Appendix A gives
more details of the rneasurements taken.

3.1

Load-Deflection Responses
The response is described in terms of the applied shear on each shear span, which is

taken as one-haif of the total load applied to the top of the column. Figures 3.l(a) and (b) show
the shear vs. centre deflection responses for specirnens CAPN and CAPH, respectively.

3.1.1

Specimen CAPN

First cracking of CAPN occurred in ends A and B simultaneousiy at a shear of 430 W.


These flexural cracks which occurred on the bottom face of the cap barn, in line with the colurnn
faces, resulted in a slight decrease in member stiffness (see Fig. 3.l(a)). At a shear of 870 kN,
a major diagonal crack fonned on each end, causing a slight drop in load and a reduction of the
stif-fhess. First yielding of the tension tie occurred at a shear of less than 2310 IcN near the
location of the first flexural crack of end A. First yielding of the crack control reinforcement
occurred at a shear of 1350 kN for end A and 1740 kN for end B. Generai yielding of the
specimen occurred at a shear of 25 10 kN and a centre deflection of 4.33 mm. A maximum shear
of 2920 kN (that is, a total applied load of 5840 kN) was reached at a deflection of 14.78 mm.
FaiIure occurred by diagonal cnishing of the concrete at the re-entrant corner between the column
and end A of the cap beam followed by slippage dong the diagonal crack which f o d between
the re-entrant corner and the support of end A. This resulted in a drop of about 30% of the loadcarrying capacity as shown in Fig. 3.l(a)

ig and shear slip

first blding of end B +tributed steel

7-

lding of end A distibuted steel

centre deflection (mm)


(a) Specimen CAPN

first diagon l cracking


I

'first ffexural citacking


5

10

15

20

centre deflection (mm)

(b) Specimen CAPH


Figure 3.1 Load-deflection responses of specimens

First cracking of CAPH occurred in end A at a shear of 490 W, and resuIted in a


signifiant decrease in rnember stiffness (see Fig. 3.l(b)). The first crack in end B did not
develop until a shear of 570 IcN was reached. These f i e d cracks were aligned with the
colurnn faces and propagated from the bottom face of the cap beam. At a shear of 890 k.N, a
major diagonal crack formed on end A which caused considerable Ioad dropoff and a reduction
in stiffness. First yielding of the tension tie occurred at a shear of 2120 kN near the Iocation of
the first flexural cracks of end A. First yielding of the crack conuol reinforcement occurred at
a shear of 1360 kN for end A and 1910 kN for end B. Generai yielding of the specirnen took
place at a shear of 2620 kN and a centre deflection of 4.06 mm. A maximum shear of 3000 kN,
or a total applied Ioad of 6 0 kN, was reached at a deflection of 9.36 mm. Failure occurred
by shear slippage dong the diagonal crack which formed between the re-entrant corner and the
support of end A, with only rninor signs of crushing near the re-entrant corner. This resulted
in a drop of approxirnately 38% of the load-carrying capacity as s h o w in Fig. 3.l(b)

LVDT's were used to deterrnine the average strains over the gauge lengths provided. and
the main gauges, glued to the reinforcing bars, were used to determine local strains in the bars.

3.2.1

Specimen CAPN
Figure 3.2shows the variation of shear vs. horizontal strains rneasured in the bottom bars

of the main tension tie. The development of strains is shown at different Iocations of the main
tension tie for both ends A and B. with solid lines used to identify readings from gauges placed
on the outermost bar and dashed lines used to identifjr those readings from the innennost bar.
From gauges A5 and A6, as well as B3 and B4, it is ckar that the outermost bars are strained
approxirnately the sarne amount as the uinermost bars. Gauge B5 did not work during testing.
Gauges A5 and A6, as well as gauge 86 (see Fig. 3.2(e) and (f)), located close to where the first
flexural cracks f o d , clearly indicate the change from pre-cracking to post-cracking stiffness
at a shear of 430 kN, corresponding to first flexural cracking. A11 of the strain measurements
on end A were Iost after a shear of 1590 khi was reached due to a rnalfunction in the data
acquisition system. Up to this shear level of 1590 kN, end A experienced slightly greater strains
than end B. Gauge B6 indicates that first yielding of the tension tie in end B occurs at a shear

A1 (outer bar)
A2 (inner bar)

B l (outer bar)
82 (inner bar)

---

2.000

4,000

6,000

8.

micmstrain
vield

A3 (outer bar)
A4 (inner bar)

- --

micmstrain

rnicmsttain

yield

vield

---

500 s'
1

2.000

4,000

microstrain

6.000

2,000

86 (inner bar)

4,000

6.000

8,

microstrain

Figure 3.2 Strains in bottom bar of CAPN tension tie. determined from strain gauges

of 23 10 kN,that is, somewhat less than the shear corresponding to general yielding of 25 10 W.
Extrapolation of readings from gauges AS and A6 indicates that first yielding of the tension tie
in end A occwed at a shear of about 2260 W. The readings from gauges B3 and B4 (see Fig.
3.2(d)), located at the imer edge of the bearing in end B, indicate that the strains at this location
were close to the yield strain at maximum shear level. At general yield of the specimen (2510
kN), these strains had only reached about 65% of their yield suain. Strains in the bars at the

start of the main tension tie hook (BI and B2) remained well below yield throughout the loading
(see Fig. 3.2(b)).
Figure 3.3 shows the applied shear vs. the measured strains in the distributed
reinforcement. Gauges located on the distributed reinforcement in end A were Iost after a shear
of 1590 kN. Gauges A10 and Al 1, as well as BI0 and Bl 1, glued to the vertical legs of the
closed hoops, experienced significant tende strains afier the first major diagonal cracking at a
shear of 870 kN. As can be seen from Fig. 3.3(a) and (b), gauges A10 and B10, which were
glued to the outer hoop legs, experienced larger suains than gauges Al 1 and B11, attached to the
inner hoop legs. Gauges A7 through Ag, and B7 and B8 were placed outside the region where
major diagonal cracks formed, and therefore experienced very Iittle straining (see Fig. 3.3(c) and
(W.
Figure 3.4 shows the strains determined from the sets of LVDT's placed dong a line
corresponding to the mid-height of the cap beam, and at the level of the centroid of the tension
tie. The average strains determined from these LVDT readings are plotted for four different load
stages: at a loading corresponding to Ml-service plus impact, at general yielding, at maximum
shear, and after failure. Also shown in Fig. 3.4 are the yield strains of the uniformly distributed
steel and the tension tie reinforcement. Figure 3.4(a) shows that some regions of the tension tie
had reached yield at a load corresponding to full-service plus impact, while the uniformiy

distributed crack control reinforcement had a maximum suain of 76% of its yield strain. At
generai yield (see Fig. 3.4(b)), the average strains exceeded the strain-hardening strain of 5.4
millistrain in three regions of the main tension tie steel. in addition, yielding of the distributed
reinforcernent at mid-height of the section occurred. At maximum shear (see Fig . 3.4(c)) there
is a noticeable difference between ends A and B, with strains in end A reaching a maximum
strain of 2.83% in the main tension tie reinforcement in linc with the c o l m face (this average
strain corresponds to a steel stress of about 600 ma). The maximum strain achieved in the
stronger end (end B) was 1.46%- As can be obsewed in Fig. 3.4(d), the strains generally
decreased due to the d r o p f f in load after failure, with the exception of the regions where a
major diagonal crack formed between the re-entrant corner and the support in end A.

microstrain

(b) vertical strain,

(a) horizontal strain, E,

microstrain

micmstrain

(c) maximum principal strain, E,

(d) minimum principal strain, E,

O
micmstrain

(e) shear strain, y,

10

20

30

40 50 60
deg rees

70

(f) principal angle, 8,

Figure 3.5 Responses of CAPN-A rosettes A6 and A7

80

90

microstrain

(a) horizontal strain, E,

5,000

10,000

15.000

(b) vertical strain, 5

20,000

microstrain

(c) maximum principal strain, E ,

500

1
O

(d) minimum principal strain, E~

.
5.000

10.000

15,000

20.000

mianstrain

( e )shear strain, y,

(f) principal angle, 8,

Figure 3.6 Responses of CAPN-6 rosettes B6 and 87

Figures 3-5 and 3-6 show the shear vs. horizontal strain, vertical strain, principd strains,
shear strain and angle of minimum principal strain detennined fiom the rosettes in ends A and
B, respectively. The strains detennined from the rosettes were very small until a crack formed
within the gauge lengths of the LVDT's. The first diagonal cracks, which fonned in ends A and

B at a shear of 870 kN, resulted in the development of significant principal tende strains and
shear strains. These are the same cracks that caused the slight drop off in load as shown in Fig.
3.1(a). The shear vs. horizontal suain,

E,,

and the shear vs. vertical strain,

responses are

described in Fig. 3.5(a) and (b), and Fig. 3.6(a) and (b), for ends A and B, respectively, Both
the horizontal and vertical strains in the region of the cap beam close to the column face
experienced a sudden increase in strains upon first diagonal cracking. In end A, the horizontal
strains are slightly larger than the vertid strains, with yielding of the u n i f o d y distributed steel

in the horizontal and vertical directions taking place at a shear of 1350 kN and 1500 kN,
respectively. In end B, the strains were Iower than in end A due to the larger arnount of
uniformiy distributed reinforcement in end B. This reinforcement, in end B, yielded in the
horizontal and vertical directions at shears of 1740 kN and 1780 kN, respectively. Both the
principal tensile strain,

and the shear strain, y,,

plots indicate that significant yielding took

place, resulting in very Iarge principal tensile strains and shear strains, particularly in end A.

The principal tensile strain was greater than 2%. resulting in very large cracks, and hence slip
dong the crack interface occurred at shears greater than 2700 W. In end B, the strains were
somewhat lower than those experienced in end A, with a maximum tende strain of 1.28% and
a maximum shear strain of 0.78%. For both ends A and B, the angle, Oz, corresponding to the
minimum principal strain was close to 45" from the horizontal until significant yielding took
place, which resulted in the angle becoming steeper.

Figure 3.7 shows the variation of shear vs . horizontal strains rneasured in the bottom bars
of the main tension tie. Solid lines are used to identiQ readings from gauges placed on the
outermost bar and dashed lines are used to identiQ those readings from the innermost bar. It can
be seen fiom gauges A3 through A6, and B3 through B6, that the strains in the outermost bars
are approxirnately the sarne as those in the imemost bars. Gauges A5 and A6, Iocated close to
where the first flexural crack fonned, clearly indicate the change from pre-cracking to postcracking stifiess at a shear of 490 kN, corresponding to first flexurai cracking (see Fig. 3.7(e)).
Readings fiom gauges B5 and B6 indicate that the first flexural crack in end B did not forrn until
a shear of 570 kN was reached (see Fig. 3.7(9). Strains in the tension tie in cantilever end A

vield

soo

-1

A? (outer bar)
A 2 (inner bar)

500

-- -

2,000

-81 (outer bar)


- - - 82 (inner bar)
t

4,000

6,000

8.

microstrain

500

k;-

-t1

~ i ( o u t ebar)
r
A4 (inner bar)

2.000

4,000

- --

6,000

8,000

microstrain
yieid

3,500
3,000

1,500 .-

A6 ( m e r bar)

2.000

4,000

(outer bar)
- - - 85
86 (inner bar)

- --

6.000

8,000

microstrain

Figure 3.7 Strains in bottom bar of CAPH tension tie, detenined from strain gauges

were slightly greater than those in end B. Gauges A5 and A6 indicate that first yielding of the
tension tie occurred at a shear of about 2120 kN, chat is, somewhat less than the shear
corresponding to general yielding (2620 W. At general yielding. the strains fiom gauges A3
and A4, located at the inner edge of the bearing in end A, were at 85% of their yield strain.
while strauis from gauges 83 and 84 had only reached 76% of their yield strain (see Fig. 3.7(c)
and (d)). The readings from gauges A3 and A4 indicate that, at the maximum shear, the strains
at this location were slightly greater than the yield suain, while strains from gauges B3 and B4
remained just below yield. Strains in the bars at the start of the main tension tie hooks (Al, A2,
B1 and B2) remained well below yield throughout the loading (see Fig. 3.7(a) and (b)).
Figure 3.8 exhibits the applied shear vs. the measured strains in the distributed
reinforcement. Gauges A10 and A l 1, expenenced significant vertical tensile strains after the first
major diagonal cracking occurred at a shear of 890 W. As can be seen fiorn Fig. 3.8(a), gauge
A10 which was glued to the outer hoop leg, expenenced large; strains than gauge A l 1, located
on the inner hoop leg. Gauges B 10 and B 11, experienced significant tensile strains at a shear of

L625 kN when a major inclined crack propagated up towards the re-entrant corner of end B (see
Fig. 3.8(b)).

Gauges A8, and B7 through B9 experienced very small strains as they were

positioned outside the region where major diagonal cracks formed (see Fig. 3,8(c) and (d)).
Figure 3.9 shows the strains determincd fiom the lines of LVDT's positioned at midheight of the cap beam, and at the level of the centroid of the tension tie. The LVDT readings
indicate that the tension tie reinforcement in end A had reached yield at a location in line with
the colurnn face at full service loading plus impact (see Fig. 3.9(a)). At this load level, the crack
control reinforcement rernained below yield. Figure 3.9(b) shows that strain hardening of the
tension tie occurred at the same location where first yielding was measured and yielding of the
distributeci reinforcement at rnid-height of the section was measured at generai yield of the
specimen. Strains in end A are considerably greater than those of end B at this stage. At the
extremities of the rnid-height Iine of measurernent, the strains were close to zero throughout
loading until a diagonal crack formed through the outer gauge length at failure. At maximum
shear, mid-height strains in end A are roughly twice those in end B (see Fig. 3.9(c)). Strains at
the level of the tension tie are largest at the locations of the major flexural cracks. The maximum
strains in ends A and B at the maximum shear were 2.62%. and 1.9496, respectively. The postfailure strains decreased due to the drop in load, with the exception of the extreme left reading
at mid-height due to the formation of the diagonai crack which caused failure (see Fig. 3.9(d)).

Figures 3.10 and 3.1 1 show the shear vs. horizontal strain. vertical strain. p ~ c i p a l
strains, shear strain and angle of minimum principal strain detennined from the rosettes in ends

5,000

10,000

15.000

micmstrain

(b) vertical strain. E,

(a) horizontal strain, E,

"

5,000

10,000

15,000

"

5,000

10,000

15,000

microstrain

microstrain

(c) maximum principal strain, E,

(d) minimum principal strain, E,

5.000

10.000

15.000

10 20

30

40

50

60

70

micmstrain

degrees

(e) shear strain, y,

(f) principal angle, 8,

Figure 3.10 Responses of CAPH-A rosettes A6 and A7

80

90

5,000

10.000

15,000

microstrain

(a) horizontal strain, E,

(b) vertical strain, E,

microstrain

(c) maximum principal strain, E,

5,000

10,000

(d) minimum principal strain, E,

15.000

microstrain

degrees

(e) shear strain, y,

(f) principal angle, 8,

Figure 3.11 Responses of CAPH-B rosettes B6 and B7

A and B, respectively. The strains detennined from each rosette were very srnall until a crack

f o d within the rosette. Significant principal tensile strains and shear strains developed at a
shear of 890 W, which corresponds to the formation of the first diagonal crack in end A. This
is the same crack that caused the slight drop-off in load as shown in Fig . 3.l(b). The rosettes
in end B did not register significant strains until the formation of the first diagonal crack in that
end at a shear of 1040 W. The shear vs. horizontal strain and the shear vs. vertical strain
responses are described in Fig. 3.10(a) and (b), and Fig . 3.1 1(a) and @), for ends A and B,
respectively. Both the horizontal and vertical strains in the cantilever ends near the colurnn face
experienced sudden increases upon first diagonal cracking. In end A, the horizontal strains were
somewhat larger than the vertical strains, with yielding of the uniformly distnbuted steel in the
horizontal and vertical directions taking place at shears of 1360 kN and 1610 kN, respectively.

In end B, the strains were lower than in end A due to the larger arnount of uniformly distributeci
reinforcement in end B. The horizontal distributed reinforcement in end B yielded at a shear of
1910 kN, while the vertical distributed reinforcement did not yield until a shear of 2730 W.
Both the principai tensile strain and the shear strain plots show that significant yielding had taken
piace resulting in very large principal tensile strains and shear strains, particularly in end A.
As can be seen from Fig. 3.10, significant strains were rneasured in CAPH-A rosette A7

when the shear reached the diagonal cracking shear of 890 kN. A maximum tensile strain of
1.4% and a maximum shear strain of 0.7% were reached during the test. Afier this maximum

strain was reached, failure occurred by shear slippage as shown in Fig. 3.15. Significant strains
developed in rosette B7 of CAPH-B at shears greater than 1040 kN (see Fig. 3.1 1). This
corresponds to the formation of diagonal cracks through this rosette. A maximum [ensile strain
of 0.65% and a maximum shear strain of 0.69% were reached in end B.

The angles of minimum principal strain for both ends were considerably steeper (see Fig.
3.10(9 and 3.1 l(f)) than those reached in specimen CAPN (see Fig. 3.5(f) and 3.6(f)).

3.3

Development of Cracking
Since one of the main objectives of this research programme is to examine the influence

of different arnounts of uniformly distributed reinforcement on crack control at service load


levels, particular care was taken to measure crack widths at al1 load stages. In order to have a
consistent means of measuring crack widths, they were rneasured at locations where the cracks
crossed two horizontal lines, one at middepth of the cap beam and the other at the level of the
cenuoid of the tension tie (see Fig. 2.8).

3.3.1

Specimen CAPN
Figures 3.12(a) and (b) illustrate the change in cracking pattern and crack widths which

were measured over the full service load range for the cap beam. First cracking of specimen
CAPN occurred at a shear of 430 kN with the formation of two short flexurai cracks, near the

bottom of the specimen, in Iine with the column faces (see Fig. 3.12(a)). As the shear was
increased to 460 kN,the load corresponding to the self-weight of the superstructure, the cracks
extended slightly but their widths had not changed significantly. Figure 3.12(b) shows the crack
pattern at a shear of 1120 kN, which corresponds closely to NI service plus impact loading on
the superstructure. This crack pattern had essentially developed at a shear of 870 kN, and as the
load was increased to 1120 kN, the cracks grew in width alone. It can be seen from Fig . 3.12(b)
that ends A and B had maximum diagonal crack widths of 0.20 and 0-25 mm, respectively, even
though end B had a larger amount of unifonnly disuibmed reinforcement (p = 0.003). than end
A (p = 0.00 18). At this load Ievel the maximum flexural crack width was 0.20 mm. It is noted
that the maximum diagonal crack width and the maximum flexural crack width were about the
same, and both are within acceptable limits for this member containhg epoxy-coated bars
As the shear was increased beyond 1120 kN, it was observed that cracks had formed at
nearly every hoop location dong the bottom of the bearn. Figure 3.12(c) shows the crack pattern
at a shear of 25 10 kN, corresponding to general yietding. The diagonal cracks had a maximum
width of 1.00 and 0.60 mm in ends A and B, respectively. For this loading case, well above the
service load range, the higher percentage of uniformly distributed reinforcement in end B
provided better crack control. Minor cnishing at both re-entrant corners was observed at load
levels slightly higher than general yield.
The crack pattern at maximum shear is shown in Fig. 3.12(d). In end A, a new major
diagonal crack formed with a width of 2.20 mm, white two existing diagonai cracks, also had
widths greater than 2.00 mm. Two of these cracks delineate the "bulging" of the newly formed
strut in end A (see Fig. 3.12(d)).
At failure, a new 3 .Omm wide diagonal crack opened suddeniy, delineating the strut
between the re-entrant corner and the bearing in end A, as shown in Fig. 3.13. This was
followed immediately by major crushing near the re-entrant corner of end A (see Fig. 3.13).

Figure 3.13 CAPN after failure

Figures 3.14(a) and (b) illusuate the change in cracking pattern and crack w i d h which
were rneasured for the HPC specimen CAPH over its MI service load range. First cracking of
specimen CAPH occurred in end A at a shear of 490 kN,slightly higher than 460 W, the load
corresponding to the superstructure dead load. This flexural crack propagated from the bottom
of the specimen, in line with the column face, over a distance of approximately one-half metre
(see Fig. 3.14(a)). This behaviour is typical of HPC as large amounts of energy are released
upon initial cracking due to the elevated tensile strength of the concrete. At a shear of 570 kN,
the first flexural crack in end B had fonned in line with the column face. Figure 3.14(b) shows
the crack pattern at a shear of 1120 kN, which corresponds closely to full service load plus
impact Ioading on the superstructure. This crack pattern had essentially developed at a shear of
890 kN, with the only change taking place, as the Ioad was increased to 1120 kN, being the
widening of the cracks. It can be seen from Fig. 3.14(b) that end A had a maximum diagonal
crack width of 0.45 mm, which is greater than permissible Iimts. The maximum diagonal crack
width in end B was only 0.25 mm, indicating that the higher percentage of unifonnly distributed
reinforcement in this end provided sufficient crack control. The maximum flexurai crack width
was 0.25 mm at this load levei, and splitting cracks could be observed dong the main tension tie.
As the shear was increased beyond 1120 kN, nearly every hoop location dong the bottom

of the beam had attracted a crack. Figure 3.14(c) shows the crack pattern at a shear of 2620 kN,
the load corresponding to general yielding. The diagonal cracks had a maximum width of 1.25
and 0.60 mm in ends A and B, respectively.
The crack pattern at maximum shear is s h o w in Fig. 3.14(d). In end A, a new major
diagonal crack forrned with a width of 1.25 mm between the re-entrant corner and the support

of end A. Just before failure occurred, minor cnishing at both re-entrant corners was observed,
and a horizontal crack at the top of tile cap beam directly under the column formed.

Figure 3-15 shows the crack pattern of specimen CAPH after failure. Failure was causeci
by relative shear slip of 8.0 mm dong the newly formed diagonal crack in end A. A minor
arnount of crushing aiso occurred near the top of this crack.

Figure 3.15 CAPH after failure

CHAPTER 4

COMPARISONS AND ANALYSES OF RESULTS

4.1

Cornparison of Responses of Normal- and High-Strength Concrete


Specimens
Loaddeformation responses of the two specimens are compared in Fig. 4.1.

First

flexurd cracking occurred at a shear of 430 kN for the normal-strength concrete specimen,
CAPN, and 490 kN for the high-strength concrete specimen, CAPH. The first diagonal cracking

in CAPN occurred at a shear of 870 kN,while the first inclineci crack CAPH occurred at a shear
of 890 W. While the high-strength concrete had about twice the compressive strength of the
normal-strength concrete. it is somewhat surprishg that CAPH had only rnarginaily higher
cracking loads. It is important to realize that at the tirne of testing, the shrinkage strains in
specimens CAPN and CAPH were about 320 and 420 microsuain, respectively (see Fig .2.3@)).
The higher expected restrained shrinkage stresses in specimen CAPH would cause a larger
reduction in the cracking load than that of specimen CAPN.
Specimen CAPN reached general yield at a shear of 2510 kN and exhibited a
displacement ductility, defined as the ultimate deflection divided by the yield deflection (Au/$),
of 3.4. Specimen CAPH exhibited a slightly stiffer response than specimen CAPN and yielded
at a shear of 2630 W. However, this specimen was considerably less ductile than CAPN,
achieving an ultimate deflection of only 2.3 times 4. The failure of the nonnal-strength concrete
specimen was caused by c m h i n g at the re-entrant corner of end A, followed by shear slip (see
Fig. 3.13). The high-performance concrete pier cap failed by shear slip dong a diagonal crack
extending from the re-entrant corner to the support of end A (see Fig. 3.15).
Steel strains in cantilever ends A of both specimens were generally slightly greater than
those of ends B. The strains measured on the tension tie reinforcement at the inside edges of the
bearing pads were significantly lower than those measured in line with the column faces. This
curtaihnent of stresses in the main tension tie is described in section 1.7.1. The steel stresses

measured at the inner edge of the bearing in end A of each specimen was typically 69 % of those
measured in Lne with the column faces after signifiant cracking had occurred. In end B of each
specimen, the stresses at the inner edge of the bearing were approximately 63 96, representative

_..I

1O

CAPN

shear slio

15

20

25

centre defiection (mm)

Figure 4.1 Cornparison of loaddeflection responses of specirnens

of the higher ratio of vertical distributed reinforcement. Most of the strain gauges located on the
distributed reinforcement did not register large strains as they were located just outside the region
of significant diagonal cracking.

By comparing the horizontal strains at mid-height for specimens CAPN and CAPH (see
Fig . 3.4(a) and Fig . 3.9(a)), at a load level correspondhg to full service plus impact loading, the
following observations can be made:
1.

End A of each specimen, which contained a distributed reinforcement ratio of 0.0018,


had a maximum horizontal strain in the cantilever portion of 1.6 millistrain.

.3

End 0 of each specirnen, contaking a distributed reuiforcement ratio of 0.003, had a

maximum horizontal strain of 1.1 millistrain for CAPN and 1.2 millistrain for CAPH.
At higher load levels the differences between these horizontal strains at mid-height of
ends A and B for both specimens becarne more significant.
The rosettes of specirnen CAPN indicate that after cracking and up to a load of about
2700 kN the principal tensile and shear strains in ends A and B were virtudly the same (see Fig.

3.5 and 3.6). At loads higher than 2700 kN, general yielding of the reinforcement resulted in
very large principal tensile and shear strains in end A. The angle of minimum principal strain
deterrnined frorn rosettes A7 and B7 were roughly 45

O.

Principal tensile strains determuied from

rosette A7 of the high-performance concrete specirnen, CAPH, were considerably greater than
those determined from rosette 87, while the shear strains were virtually the sarne in the two ends
(see Fig. 3.10 and 3.1 1). The angle of minimum principai strain detennined from rosette B7 was
slightly steeper than that of A7.
Figure 4.2 compares the flexural crack widths measured at the level of the tension tie in
the normai- and high-strength concrete pier cap specimens. Crack widths are slightly higher in
the high-strength concrete specirnen due in large part to the greater release of energy upon initial
cracking. Figure 4.3 compares the diagonal crack widths measured at mid-height of the norrnaland high-strength concrete specimens. It is clear that crack widths are considerably larger in end
A of the high-strength concrete specimen than in the normal-strength concrete specimen, while

the crack widths in end B of each specimen are roughly the same.
Figures 4.4(a) and (b) compare the maximum diagonal crack widths in ends A and B of
specimens CAPN and CAPH, respectively. There was nor a significant difference between the
crack widths of the two ends of CAPN under upper serviceability conditions (refer to Fig.

3.12(b)), and they were al1 smaller than required by code limits. However, at higher load levels,
the extra reinforcement in end B caused a moderate improvement in crack control over end A.

manimum crack width (mm)

maximum crack width (mm)

(a) Maximum crack widths


3,500

3,000

---

N d -

,
/

CAPN-A

500

CAPH-A

---

sum of crack widths (mm)

sum of crack widths (mm)

(b) Surn of crack widths


Figure 4.3 Diagonal crack widths measured at mid-height of specimens

0.5

1.O

1.5

2.0

maximum diagonal crack width (mm)


(a) Normal-strength concrete specirnen,CAPN

0.5

1.O

1.5

2.0

maximum diagonal crack width (mm)


(b) High-strength concrete specimen, CAPH

Figure 4.4 Influence of distnbuted reinforcement ratio on crack control

Altematively, a significant difference in crack control performance of the two ends of CAPH
could be observed at full service plus impact Ioading. A diagonal crack in end A had already
opened up to 0.45 mm, while the cracks in end B were well controlled (refer to Fig. 3.14@)).
It is interesthg to note that the high-strength concrete specirnen has a post-cracking
stifmess which is only slightly higher than that of the normal-strength concrete specimen (see Fig.
4.1). This result is sornewhat surprishg since the tensile strength of the high-strength concrete
is about 1.6 times the tensile strength of the no&-strength

concrete (see Table 2.3). The

reduced tension stiffening observed in the high-strength concrete specimen may be due to the
greater tendency for bond splitting cracks as c m be obsewed by comparing the crack patterns
of CAPN and CAPH (see Fig. 3.13 and 3.15). A

' '

'

et al. (1993) have postulated that

high-suength concrete, due to its higher compressive strength, develops higher , more localized
bond stresses. This, together with the fact that the bearing capacity of the concrete is related to
fcl whereas the tensile strength is related to
results in bond splining of the concrete before

K.

a uniform bond stress c m be achieved. Abrisharni et al. (1995) conciuded that the presence of

epoxy-coating on reinforcement results in fewer flexural cracks, larger crack widfhs, more
splitting cracks and decreased ductility. In addition, Abrisharni and Mitchell (1996) concluded
that bond splitting cracks reduce the tension stiffening in the concrete and that after significant
deformations in the postcracking range, the tension stiffening of hi&-strength concrete specimens
approaches that of normal-strength concrete specimens. The effects associated with the use of
high-strength concrete in combination with the use of epoxy-coated reinforcing bars has resulted
in a reduced tension stiffening, fewer flexural and diagonal cracks, bond splitting cracks and a
lower ductility.

A number of different types of predictions were carrieci out to determine the response of

the specimens tested.

Although pIane-sections analysis is not applicable for prdicting the responses of disturbed
regions, it is of interest to compare the predicted cracking loads, using this method, with the
rneasured cracking loads.
The predicted shears corresponding to first flexural cracking, at rnid-span of the

specimens, from plane-sections analyses were 462 kN and 661 kN for specimens CAPN and
CAPH. respectively. These values are significantly higher than those measured during testing,

that is, 430 kN for CAPN and 490 kN for CAPH. The key reasons why these predictions are
unconservative are that the strain distribution in disturbed regions is significantly non-linear and
that concrete shrinkage strains were not given any consideration.

The cornputer prograrn

RESPONSE (Collins and Mitchell 1991) was used to cary out the sarne predictions including the
effect of concrete shrinkage strains. The shears corresponding to first flexural cracking including
shrinkage strains were predicted to be 304 kN and 476 kN for specimens CAPN and CAPH,
respectively, which are conservative predictions.

4.2.2

Simple Strut-and-Tie Models


Simple stnit-and-tie rnodels were developed to establish preliminary predictions of the pier

cap yield strengths (see Fig. 4.5 and 4.6). Both specimens were governed by yielding of the
main tende tie. The main tension tie, which consists of 10 No. 25 bars with a yield stress of
468

Mh, has a yield force of 2340 kN. It is assumed that the lines of action of the diagonal

stmts intersect the lines of action of the compressive resultants in the column (Le., at the quarter

points of the column). From equilibriurn, the shear which corresponds to yielding of the main
tension tie is 1995 kN for specirnen CAPN and 2050 kN for specimen CAPH (see Fig. 4.5 and
4.6). In these predictions, the materid reduction factors were taken as 1.0. These models are
simple and no consideration is given to any strength enhancement provided by the distributeci
reinforcement, particularly the horizontal bars. The predictions are therefore conservative.

Figure 4.5 Simple strut & tie modei for specimen CAPN

Figure 4.6 Simple strut 8 tie model for specirnen CAPH

4.2.3

Refined Sm-and-Tie Modeis


The r e f d strut-and-tie modeis s h o w in Fig. 4.7 and 4.8 account for the contribution

of the crack controI reinforcement to the strength of the specimens. In addition to the main
tensile tie at the bottom of the specimens, additional ties are provided to represent the horizontal
and vertical distributed steel. These secondary tension ties are positioned at the centroids of the
effective horizontal and vertical unifonnly distributed reinforcement. The refined strut-and-tie
models shown in Fig. 4.7(a) and 4.8(a) model the response of each specnen as though they were
reinforced throughout with a reinforcement ratio of 0.0018 for the distributed steel. These
details, which were used in ends A of both test specimens, are rnodelled by a secondary
horizontal tension rie (4 legs of No. 15 bars) with a yield force of (800 mm2)(419 MPa) = N O kN
and vertical tension ties (3 sets of 4 legged No. 10 hoops) at each end with yieid forces of (1200
mrn2)(441 MPa) =Yi0 kN. The predictions obtained from Fig 4.7(a) and 4.8(a) are representative

of the weaker side of each specimen and hence should be used when comparing with the actual
strengths. The refined strut-and-tie models s h o w in Fig. 4.7(b) and 4.8@) model a distributed
reinforcement ratio of 0.003, which is the same reinforcement ratio contained in ends B of
specimens CAPN and CAPH. The yield forces of the horizontai and vertical tension ties in Fig.
4.7(b) and 4.8(b) were calculated to be 670 kN and 880 kN, respectively. The predictions
obtained frorn Fig. 4.7(b)and 4.8(b) are presented in order to demonstrate the how the strengths
would increase if the Iarger amount of uniformly distributed reinforcement (Le., a ratio of 0.003)
were present throughout the specimens.
The changing inclinations of the main diagonal struts in Fig. 4.7 and 4.8 are induced by
the presence of the vertical and horizontal distributed reinforcement. The tensile forces result
in discrete angular changes at the nodes where the secondary tension ties intersect the struts. The
resulting arching action provides steeper struts above the supports, ultirnately resulting in higher
strengths. If more distributed steel were present, then the arching action would be even more
pronounceci (see Fig. 4.7(b) and 4.8 (b)).
This more detailed strut-and-tie model gives a better representation of the flow of
compressive stresses. The modelling of the flow of compressive stresses from the column into
the cap bearn results in higher localized compressive stresses near the re-entrant corners and
secondary struts which represent the fanning compressive stresses anchored by the vertical
uniformly distributeci reinforcement. A comparison of Fig. 4.7 with 4.8 illustrates that the struts
for the high-strength concrete specimen are considerably smaller than those in the normal-strength
concrete specimen. This effect gives a slight increase in the capacity for the high-strength
concrete specimen.

(a) End A details

V = 2540 kN
(b) End B details
Figure 4.7 Refined strut & tie models for specimen CAPN

(a) End A details

(b) End R details

Figure 4.8 Refined strut & tie models for specirnen CAPH

Table 4.1 compares the predictions made with the simple strut-and-tie mode1 and the
refmed suut-and-tie mode1 with the measured values of total load applied to the pier caps at
general yield. in rnaking these predictions, it was assumed that both ends of the cap beam were
reinforced with the smaller amount o f uniforiniy distributed reinforcement, that is consistent with
end A, since end A will give a lower predicted load. It is apparent that the refined strut-ad-tie
models give excellent predictions of the load at general yielding. Accounting for the uniformly
distributed reinforcement can significantly increase the predicted yield load. while giving slightly
conservative predictions. It m u t be pointed out that the acniai faiIure loads are somewhat higher
than the general yielding loads due to strain hardening in the reinforcement. The predictions

made with the stmt-and-tie models neglected the effects of strain hardening.

1 Measured 1
Specimen

Load at
Yield

Odv)

Simple Strut-and-Tie

Re-

Stru-and-Tie

Predicted

Yield

Measured
Predided

Predicted

Measured

Yield
OrN)

CAPN

5020

3990

1.26

4820

1.04

CAPH

5240

4100

1.28

5 120

1.O2

Table 4.1 Conparison of strut-and-tie predictions with measured loads at generai yielding

4.2.4

Non-Linear Finite Element Analysis Using Program FIELDS

Figure 4.9 compares the measured loaddeflection responses with the predicted responses
obtained by using the non-linear finite element prograrn FIELDS (Cook 1987, Cook and Mitchell
1988) for specimens CAPN and CAPH. In predicting the responses the cracking stress was
adjusted to account for the size effect of these NI-scaie specirnens. Using a cracking stress of
0.33

fifor these specirnens which experience signifiant diagonal cracking within the cantilever

portions of the cap beams, and assuming that the cracking stress is inversely proportional to the
fourth root of the size, then the cracking stress for these 1 1 0 mm deep members compared to
the 150 mm deep control specirnens would be:

10

15

20

25

20

25

centre deflection (mm)


(a) Specirnen CAPN

10

15

centre deflection (mm)


(b) Specimen CAPH
Figure 4.9 Predicted and measured loaddeflection responses of specimens

CAPN and CAPH


agree very well with the rneasured responses up to loads of 2490 kN and 2670 kN,respectively.
As can be seen from Fig. 4.9 the predicted responses for spechens

At these load levels, the non-linear analysis predicts local crushing at the re-entrant corner. The
load deflection responses up to cracking and from cracking up to the points where local cnishing
is predicted, agree reasonably well with the measured response. As can be seen from comparing
Fig . 4.9(a) and (b), the predictions over-estimate the tension-stiffening ,particularly for the highstrength concrete specimen (see discussion in Section 4- 1).
In order to reduce the sensitivity due to local cnishing, the elements at the re-entrant
corner were softened by specifying a compressive stress-strain curve having a peak strain equal
to 1.5 times the cylinder peak strain. The frnite element analysis gives an accurate prediction of
yielding, however since the analysis relies on a tangent stiffness model, it was unable to converge
after local crushing was predicted.
Figure 4.10 shows the deflected s~hapesof specimens CAPN and CAPH at the predicted
maximum load levels. It is apparent from this figure that the deformations are not symmetrical
about the centrelines of the pier caps due to the fact that end B of each specimen contains a
greater arnount of uniformly distributed horizontal and vertical reinforcement.
Figures 4.11 through 4.16 show the predicted strains and concrete stresses for specimens
CAPN and CAPH at three different load levels. At the lower service load level, that is a total

applied load of 920 kN, for both the normal-strength and high-strength concrete specimens the
stresses are nearly elastic with only minor cracking predicted for specimen CAPN. in addition,
no distinct compressive strut action is apparent at this load level (see Fig. 4.11 and 4.12).
Figures 4.13 and 4.14 show the predicted strains and stresses at the upper service load level
corresponding to a total applied load of 2280 kN. Significant principal tensile strains are
predicted in both spec'mens at this load level. It is apparent that Iarger principal tensile strains
occur in end A than in end B due to the smaller amount of uniformiy distributed reinforcement.
Figures 4.15 and 4.16 show the predicted strains and stresses at the maximum predicted
load IeveIs. It is apparent that the principal tensile strains are Iarger for the diagonal cracks than
for the flexural cracks. By observing the flow of compressive stresses it is apparent that more
direct compressive strut action is taking place close to failure. Some bulging of the compressive
stmts between the coIumn and the reaction bearhgs is apparent. The high-strength coricrete
specimen CAPH exhibits struts having smaller widths and higher compressive stresses. The nonlinear finite-element analysis predicts a 7% higher ultirnate strength for CAPH than for CAPN.
The predicted strains in the tension tie for the high-strength concrete specimen are higher than
those predicted for the normal-strength concrete specimen-

)1

,
I

displacement scale:
5.00 mm

: - ----

---.-

V = 2490 kN

(a) Specimen CAPN

<

1
I

displacernent scale:
5.00 mm

! ! !

V = 2670 kN

(b) Specimen CAPH


Figure 4.1 0 Predictions of deflected shapes of specirnens at maximum
predicted loads

(a) Principal strains

stress scale:
5 MPa

(b) Stresses in concrete

Figure 4.11 Predicted strains and stresses in specimen CAPN at a load of 920 kN

(a) Principal strains

stress scale:
5 MPa

(b) Stresses in concrete


Figure 4.12 Predicted strains and stresses in specimen CAPH at a load of 920 kN

(a) Principal strains

stress scale:
5 MPa

(b) Stresses in concrete

Figure 4.1 3 Predicted strains and stresses in specimen CAPN at a load of 2280 kN

(a) Principal strains

stress scale:
5 MPa

(b) Stresses in concrete

Figure 4.14 Predicted strains and stresses in specimer! CAPH at a load of 2280 kN

(a) Principal strains

stress scale:
20 MPa

(b) Stresses in concrete

Figure 4.1 5 f redicted strains and stresses in specimen CAPN at a load of 4980 kN

strain scaie:

(a) Principal strains

stress scale:
20 MPa

(b) Stresses in concrete


Figure 4.1 6 Predicted strains and stresses in specirnen CAPH at a Ioad of 5340 kN

Figure 4.17 shows the development of stress in the main tension ties of specimens CAPN
and CAPH. Stresses are plotted at applied shears of 460 kN and 1140 k;N (the service load range
bounds), 2000 kN and at the maximum load predicted by the finite element analyses, The
predictions for both specimens indicate a significant stress drop-off within the region containhg
the confinement reinforcement provided by the column ties. The measured strains are typically
somewhat higher than the predicted strains. It m u t be pointed out that the predicted strains
shown in the figure are "average" strains and therefore would be Iess than the strains measured
at or near cracks. The non-linear finite element anaiyses provided excellent predictions of the
strains in the main tension tie at the inner edge of the bearing. The presence of uniformly
distributed reinforcement results in a drop-off in stress from the location of maximum moment
towards the inner edge of the bearing . As can be seen from the measured and predicted stresses
the tension tie force decreases for locations close to the bearing due to the contribution of the

vertical uniformly distributed steel. The larger arnount of vertical reinforcement in end B results

in reduced force dernands on the tension tie at the imer edge of the bearings. (see Fig . 4-17).
Figure 4.18 compares the predicted stresses in the main tension tie using finite element
analysis and refined strut-and-tie modelling with the stress computed from the measured steel
strains at maximum predicted loads. It can be seen that the refined strut-and-tie model gives
reasonable predictions for these stresses.
Table 4.2 compares the predictions made using the r e f d strut-and-tie model and the
predictions from the non-linear fuiite-element analysis with the measured values of total load
applied to the pier caps at general yield. Although the non-linear finite element analysis gives
slightly better predictions, the refined strut-and-tie mode1 compares exceptionally well with this
more sophisticated approach. It mus: be pointed out however that the non-linear finite element
analysis is capable of predicting strains and crack widths at service load 1eveIs.

s-u
CAPN

Yield

5020

Predided

Me&

Predicted

Table 4.2 Cornparison of refmed strut-and-tie predictions and non-linear finite element
predictions with measured loads at generai yielding

(a) Specimen CAPN

rneasured

(b) Specirnen CAPH


Figure 4.17 Predictions of stress development in main tension ties

measured

refined sM-and-tie

(a) Specirnen CAPN

measured

-------

finite element

-refined
- - -stnit-and-tie
------

(b)Specimen CAPH
Figure 4.d8 Predictions of stress development in main tension ties at general yield

Estimates of Crack Widths

4.3

Tables 4.3 and 4.4 compare the measured principal tensile strains and crack widths with
those predicted using the results from the non-linear f i t e element analyses. The crack width
predictions were made for both flexural and diagonal cracks. The expecte flexural crack widths,
w , were determined fiom:

where:

td

= maximum predicted principal tensile strain at the level of the niain tension tie,

Sm

average crack spacing predicted fkom CEB expression (see Section 1.9).

The expected diagonal crack widths were determineci h m :

where: e#

maximum predicted principal tensiie strain at mid-height of pier cap.

The predicted average spacing, s,,

of the diagonal cracks is deterrnined from (Collins and

Mitchell 1991):

where,s

and s,, are the crack spacings indicative of the crack control characteristics of the

horizontal and vertical distributed reinforcement. respectively. For simplicity sm and s.,

were

taken as the spacings of reinforcement in the two directions and the angle of principal
compression, 8, was assumed to be 45 O . In addition, the predicted crack widths were multiplied
by a factor of 1.2 to account for the influence of epoxy coating on the reinforcement (Abrishami

et al. 1995).
As can be seen fkom Table 4.3, the flexural crack widths predicted using non-linear finite

element analyses compare very well with the measured maximum crack widths.
The predicted widths of diagonal cracks can Vary considerably frorn the crack widths
observed (see Table 4.4). One concern is that when applying normal procedures to the high
strength concrete specimen, the principal tensile strain and the crack width rnay be
underestimated. This rnay be due to the larger energy released when cracks forrn in high-strength
concrete members, which can Iead to the formation of longer and larger cracks. In addition this

Load
OrN)

Specimen

cm

~%~mmi

Wpndiacd

(id,

(mm)

(mm)

%e=Wtd

920

O. 169

0.29 1

0.05

0.05

2280

0.805

1.134

0.24

0.20

CAPN-B

920

O. 107

O. 144

0.03

0.05

s, = 2 4 8 m

2280

0.704

1.000

0.2 1

O. 15

CAPH-A

920

0.037

0.064

s, =248mm

2280

0.704

1.O87

0.2 1

0.25

CAPH-B

920

0.037

0.067

CAPN-A

s,

= X8mm

2280

0.562

1.068

0.17

0.20

Table 4.3 Comparison of predicted and measwed crack widths and principai tensile strains
in the main tension tie

Load

Specime~~

'marumi

EPreti~

wmeas~t~

wprcdicttd

(lm

(107

920

O .O25

2280

2.174

2-459

0.55

0.30

CAPN-B

920

0.030

sd=124mm

2280

O. 985

1.953

0.15

0.25

CAPH-A

920

0.019

sme = 2 10mm

2280

1.638

2.203

0.4 1

0.45

CAPH-B

920

0.019

CAPN-A
S~=~~OII-IIII

124mm

2280

0.585

1.096

(-1

(mm)

0.09

0.25

Table 4.4 Comparison O predicted and measured diagonal crack widths and principal
tens ile strains at rnid-height

phenomenon may be due to the fact that the tension stiffening in high-strength concrete members
tends to approach that of normal-strength concrete members after signifiant cracking has
developed (see Section 4.1).

CONCLUSIONS

The conclusions arising from this research project are surnmarized as follows:
1.

A reinforcement ratio for the unifomily distributeci steel of 0.002 was sufficient to control

cracking over the depth of the normal-strength concrete pier cap specimen. This amount
of reinforcement is required in the 1994 CSA Standard for controiling cracking in
disturbed regions. Side A of the normal-strength concrete specimen contained a
reinforcement ratio of 0.0018. and exhibited adequate crack control at service load Ievels.
2,

A unifonnly distributed steel reinforcement ratio of 0.003 was necessary to adequately

control cracking over the depth of the high-strength concrete pier cap specimen at service
load levels. Fcr the high-strength concrete specirnen, initial cracks tended to be longer
and wider than those observed in the normd-strength concrete specirnen.
3.

The high-strength concrete specirnen had a slightly higher strength dian the normaistrength concrete specimen due to the sinaller compressive struts in the high-strength
concrete pier cap, leading to a slightty larger effective depth. The high-strength concrete
pier cap specimen exhibited a 32% Iower ductility than the normal-strength concrete
specimen.

4.

Both the normai- and high-strength concrete specimen exhibited cracking Ioads which
were influenced by the large size of the specimens and by the restrained shrinkage
stresses. The cracking load of the high-strength concrete specimen was only slightly
higher than that of the normal-strength concrete specirnen due to the higher shnnkage
strains experienced in the high-strength concrete.

5.

Simple strut-and-tie models provided conservativeestimates of the strength of the pier cap
specimens.

6.

Refined strut-and-tie models which simulate the effect of the horizontal and vertical
distributed reinforcement, provided better estimates of the general yielding load of the
specimens than the simple strut-and-tie model. In the refined strut-and-tie model, the
inclusion of the horizontai tension tie representing the unifonnly distributed horizontal

reinforcement significantiy increases the strength prediction. The vertical tension ties
representing the uniformly distributeci vertical reinforcement reduces the required force
in the main tension tie near the support bearings.

7.

The predictions using non-linear fiaite elernent analyses gave accurate predictions of the
variation of stress in the main tension tie and provideci a means of assessing the principal
tensile strains and crack widths at service Ioad levels.

8.

Reasonably accurate predictions of flexural crack widths were made by applying the usual

crack spacing assurnptions to the principal tensile strains obtained from the non-linear
finite elernent analyses.
9.

More research is requireci to accurately predict the inclined crack widths in very large
disturbed regions and to properly account for the influence of high-strength concrete on
inclined crack w idths .

REFERENCES
AC1 Committee 3 18 (1995), "Building Code Requirements for Structural Concrete (AC1
3 18-95)", American Concrete Institute, Detroit, 1995.

Abrishami, H. H., Cook, W. D. and Mitchell, D. (1995). "Muence of Epoxy-Coated


Reinforcement on Response of Normal and High-Strength Concrete Beams " , ACI Smctural
Journal, Vol. 92, No. 2, March-April 1995, pp. 157-166.
Abrishami, H.

H. and Mitchell, D. (1996), "influence of Splitting Cracks on Tension

Stiffening", ACI Structural Journal, Vol. 93, No. 6, Nov.-Dec. 1996, pp. 703-710.
Al-Soufi, S. (1990), "The Response of Reinforced Concrete Bndge Pier Caps", Masters
thesis, McGill University, MontreaI, 1990. 134 pp.
Azizinamini, A., Stark, M., Roller, J. J. and Ghosh, S. K. (1993), "Bond Performance
of Reinforcing Bars", ACI Structural Journal, Vol. 90, No. 5, Sept.-Oct. i396, pp. 554-58 L .

CSA Committee A23 -3 (1994). "Design of Concrete Structures (A23.3-94)", Canadian


Standards Association, Rexdale, ON, 1994.
"Canadian Highway Bridge Design Code (CHBDC)" , Canadian Standards Association,

Rexdale, ON, 1997.


Collins, M. P. and Mitchell, D. (1980), "Shear and Torsion Design of Prestressed and
Non-Prestressed Concrete Beams " , Journal of the Prestressed Concrete h t i t u t e , Vol. 25, No.

5, Sept.-Oct. 1980, pp. 32-100.


Collins, M. P. and Mitchell, D. (1985), "EvaIuating Existing Bridge Structures using the
Modified Compression Field Theory", AC1 Special Symposium Vol. SP-88 Strength Evaluation

of Msting Concrete Bridges, Amencan Concrete Institute, Detroit, 1985, pp. lC9- 141.
Collins, M. P. and Mitchell, D. (1986). "A Rational Approach to Shear Design - The
1984 Canadian Code Provisions", ACI Journal, Vol, 83, No. 6, Nov.-Dec. 1986, pp. 925-933.
Collins, M. P. and Mitchell, D. (199 l), "Prestressed Concrete Structures", Prentice-Hall
Inc., EngIewood Cliffs, NJ, 1991, 766 pp.
Collins, M. P., Mitchell, D. and MacGregor, J. G. (1993), "Structural Design
Considerations for High-Strength Concreten, Concrete Incemational, Vol. 15, No. 5, May 1993,
pp. 27-34.

Collins, M. P. and Porasz, A. (1989), "Shear Design for High Strength Concreten, CEB

Bul!drin d'lnfonnation, No. 193, Dec. 1989, pp. 77-83.


Comit Euro-international du Bton (1990). "CEB-FIP Mode1 Code (MC 90)". Thomas
Telford Services Ltd., London, 1993.
Cook, W. D. (1987)- "Smdies of Reinforced Concrete Regions Near Discontinuities",
PhD thesis. McGill University, Montreal, 1987. 153 pp.
Cook, W. D. and Mitchell, D. (1988). "Studies of Disturbed Regions near Discontinuities
in Reinforced Concrete Members",ACI Smtctural Journal, VOL 85, No. 2, March-April 1988,
pp. 206-2 16.
Franz, G. and Niedenhoff, H. (1963). "The Reinforcement of Brackets and Short Deep
Beam", Cernent and Concrete Association, Library Translation No. 6 1.114, London, 1964.
Gergely, P. and Lutz, L. A. (1968), "Maximum Crack Width in Reinforced Concrete
Flexural Members", AC1 Special Symposium Vol. SP-20 Causes, Mechanisms, and Control of
Cracking in Concrete, American Concrete Institute, Detroit, 1968, pp. 87- 117.
Hognestad, E. (1957), "Confirmation of inelastic Stress Distribution in Concrete",
Proceedings of the Amen'can Society of Civil Engineers, Vol. 83, No. ST2, March 1957, pp.
1189-1 to 1189-17.
Kani, M. W., Huggins, M. W. and Wittkopp, R. R. (1979). "Kani on Shear in
Reinforced Concrete", Department of Civil Engineering, University of Toronto, Toronto, 1979,
225 pp.
Kriz, L. B. and Raths, C. H. (1965), "Connections in Precast Concrete Structures

Strength of Corbels", Journal of the Prestressed Concrete Institue, Vol. 10, No. 1, Feb . 1965,

pp. 16-47.
Leonhardt, F. and Wdther, R. ( 1966). "Wandatiger Trager (Wall-like Bearns) ",
Deutscher AusschussfLir Stahlbeton, Bulletin No. 178, Wilhelm Ernst und Sohn, Berlin, 1966,
159 pp.
MacGregor, J. G. ( 1997), " Reinforced Concrete: Mecfianics and Design", Prentice-Hall
Inc., Upper Saddle River, NJ, 1997, 939 pp.
Marti, P. (1985), "Basic Tools of Reinforced Concrete Beam Design", AC1 Jourrial, Vol.
82, No. 1, Jan.-Feb. 1985, pp. 46-56.

Mast, R. F. (1968), "Auxilliary Reinforcement in Concrete Connections", Proceedings


of the American Society of Civil Engineers, Vol. 94, No. ST6,June 1968, pp. 1485- 15W.
Mattock, A. H . ( W 6 ) , "Design Proposais for Reinforced Concrete Corbels", J o u m l
of the Prestressed Concrete Institute, Vol. 2 1, No. 3, May-iune 1976, pp. 2-26.

Mattock, A. H.,Chen, K. C. and Soongswang, K. (1976), "The Behviour of Reinforced


Concrete Corbels", l o u d of the Prestressed Concrete Imtitute, Vol. 2 1, No. 2, March-April
1976, pp. 52-77.
Morsch, E. (1909), "Concrete-Steel Construction (Der Eisenbetonbau)", translation of
3rd German edition by E. P. Goodrich, McGraw-Hill Book Co., New York, 1909, 368 pp.
Park. R. and Paulay , T. ( l975), "Reinforced Concrete Structures" , Wiley-Interscience,
New York, 1975, 769 pp.
Popovics, S. (1973), ''A Numericai Approach to the Complete Stress-Strain Curve of
Concrete". Cernent and Concrete Research, Vol. 3, No. 5, May 1973, pp. 583-599.
Ramirez, J. A. and Breen, J. E. (1991), "Evaluation of a Modified Tniss-Mode1
Approach for Beams in Shear", ACI Srmctural Journal, Vol. 88, No. 5, Sept.-Oct. 1991, pp.
562-57 1.
Ritter, W. ( l899), "Die Bauweise Hennebique (The Hennebique Design Method)",
Schweizerische Bauzeitung , Vol. 33, No. 7, Feb- :899, pp. 59-6 1.
Rogowsky, D. M. and MacGregor, J. G. (1986), "Design of Reinforced Concrete Deep
Beams", Concrete International, Vol- 8, No. 8, August 1986, pp. 49-58.
Rogowsiq, D. M-,MacGregor, J. G. and Ong, S. Y. (1986). "Tests of Reinforced

Concrete Deep Beams" ACI Structural Journal, Vol. 83, No. 4, July-August 1986, pp. 6 14-623.
Schlaich, J. and Schafer, K. ( l984), "Konstruieren im Stahibetonbau (Reinforced
Concrete Construction)" , Beron-Kalender I984, W ilhelm Ernst und Sohn, Berlin, 1984, pp. 7871004.

Schlaich, J., Schafer, K. and iemewein, M. (1987). "Toward a Consistent Design of


Structurd Concrete" , Journal of the Prestressed Concrete Imtitute, Vol. 32, No. 3, May-June

1987, pp. 74-150.

Thorenfeldt, E., Tomaszewicz, A. and Jensen, J. J. ( 1987). " Mechanical Properties of


High-Strength Concrete and Application in Design", Proceedings of the Symposium Utilimtion

of High-Strength Concrete, Tapir, Trondheim, 1987, pp . 149- 159.


?'hriiman, B., Marti, P., Pralong, J., Ritz, P. and Zimrnerli, B. (1983), "Anwendung
der Plastizitaetstheorie auf Stahlbeton (Application of the Theory cf Plasticity to Reinforced
Concrete)", Institue of Structural Engineering, ETH Ziirich, Germany, 1983, 252 pp.
Vecchio, F. J. and Collins, M. P. (1986), "The Modified Compression Field Theory for
Reinforceci Concrete Elernents Subjected to Shear", ACI Journal, Vol. 83, No. 2, March-April
1986, pp. 219-231.
Walraven, J. C. (1981), "Fundamentai Analysis of Aggregate Interlock", Proceedings
of the Amencan Sociery of Civil Engineers, Vol. 107, No. ST 11, Nov. 1981. pp. 2245-2270.

APPENDIX

EXPERIMENTAL DATA

This appendix presents a surnmary of the experirnental data recorded for the two pier cap
specimens. The data presented includes applied shear, LVDT readings, and strains from the
electrical resistance strain gauges.

instrumentation.

Refet to Fig. 2.7 and 2.8 for descriptions of the

Table A.l Readings h m vertical LVDTs used to determine the deflection of


specimen CAPN

Table A.2 Readings from LVDTs located at the level of the main tension tie in
specimen CAPN-A

Table A.3 Readings from LVDTs located at the level of the main tension tie in
specimen CAPN-B

Table A.4 Readings from LVDTs Iocated at rnid-height of specimen CAPN-A

Table A S Readings from LVDTs located at mid-height of specimen CAPN-B

TabIe A.6 Readings from LVDT rosettes located in end A of specimen CAPN

Table A.7 Readings frorn LVDT rosettes located in end B of specimen CAPN

Shear

Al

A2

A3

A4

AS

A6

Wv

(106,

(109

(109

(103

(109

(103

246 .O

-2

-4

44

46

427.5

-2

-6

210

254

615.5

-2

-8

14

496

520

864.5

-4

-6

36

66

8 12

834

1123.0

10

382

540

1052

Il06

1502.5

44

26

986

1446

1526

NIA

1753.5

NIA

NIA

NIA

NIA

NIA

NIA

1998.5

NIA

NIA

NIA

NIA

NIA

NIA

225 1.5

N/A

N/A

NIA

N/A

NIA

NIA

2505 .O

NIA

NIA

NIA

NIA

NIA

NIA

2665.5

NIA

NIA

NIA

N/A

NIA

NIA

272 1.5

NIA

NIA

NIA

N/A

NIA

NIA

2786.O

NIA

NIA

N/A

NIA

NIA

NIA

2833 .O

NIA

NIA

NIA

N/A

NIA

NIA

2895.5

NIA

NIA

NIA

NIA

NIA

NIA

29 12.5

NIA

NIA

NIA

NIA

NIA

NIA

2104.0

NIA

NIA

NIA

NIA

N/A

NIA
L

Table A.8 Strains from strain gauges located in end A of specirnen CAPN

Shear

A7

AS

A9

A10

Al1

(kW

(106,

(106,

t103

(109

(106,

246 .O

- 14

- 12

427.5

-24

-20

18

1O

615.5

-34

-32

32

-2

-2
-

864.5

-34

-2 8

20

14

18

1123.0

-60

-38

14

2 12

58

1502.5

-80

-46

22

338

124

1753.5

NIA

NIA

NIA

NIA

NIA

1998.5

NIA

NIA

NIA

NIA

NIA

225 1.5

NIA

NIA

NIA

NIA

N/A

2505.0

NIA

NIA

NIA

NIA

N/A

2665 -5

NIA

NIA

NIA

NIA

NIA

272 1.5

NIA

NIA

NIA

NIA

NIA

2786 .O

NIA

NIA

NIA

NIA

NIA

2833 .O

NIA

NIA

NIA

NIA

NIA

2895.5

NIA

NIA

NIA

NIA

N/A

2912.5

NIA

N/A

NIA

NIA

N/A

2104.0

NIA

NIA

NIA

NIA

NIA

Table A.8 (Cont 'd) Strains from strain gauges located in end A o f specimen CAPN

Shear

B1

B2

(10~)

(10'3

B3
(10~)

B4

B5

(106,

(10'9

M
(106,

NIA

246.0

-2

-4

NIA

42

427 -5

-4

-6

10

NIA

100

615.5

-4

- 10

14

NIA

440

864.5

-8

-10

24

42

NIA

722

1123.0

- 10

-8

186

178

NIA

970

1502.5

10

642

554

NIA

1368

1753-5

24

14

872

758

NIA

1620

1998.5

48

22

1140

968

NtA

1878

225 1.5

110

38

1462

1234

NIA

2 180

2505 .O

164

52

1612

1476

NIA

2506

2665 -5

232

72

1828

1664

NIA

2740

272 1.5

274

90

1882

!734

NIA

2776

2786.0

302

100

1950

1794

NIA

3 138

2833.0

388

108

2168

1902

NIA

5698

2895. 5

446

122

2220

1992

NIA

3054

29 12.5

466

128

2266

2054

NIA

2998

2 104.0

454

130

1802

1774

NIA

2638

Table A.9 Strains from strain gauges located in end B of specimen CAPN

Shear

B7

Ba

OrN)

(109

(109

B9
(106,

BI0

BI1

(109

W6)

NIA

246.0

-12

- 12

NIA

-6

427.5

-20

-24

NIA

- 10

-14

-28

-36

NIA

-28

-32

864.5

-22

-40

NIA

12

26

1123.0

-34

-54

NIA

410

148

1502.5

-64

-80

NIA

710

338

1753.5

-82

- 100

NIA

882

458

1998.5

-100

-120

NIA

1080

598

225 1.5

-1 12

- 142

NIA

1376

796

2505 .O

-1 14

-164

NIA

1564

956

2665.5

-94

-180

NIA

1730

1138

272 1-5

-82

-192

NIA

1818

1268

2786 .O

-70

-200

NIA

1802

1292

2833 .O

340

-24

NIA

1378

1230

2895.5

6 12

54

NIA

1304

1260

2912.5

668

80

NIA

1244

1162

2104.0

636

106

NIA

1060

1016

6 15.5
-

Table A.9 (Cont 'd) Strains from strain gauges located in end B o f specimen CAPN

Table A.10 Readings from vertical LVDTs used to detennine the deflection of
specimen CAPH

Table A.11 Readings from LVDTs located at the level of the main tension tie in
specirnen CAPH-A

Shear

B5

B4

B3

(mm)

(mm)

(mm)

B2
(-1

B1
(mm)

250.0

-0.003

-0.006

495.5

-0.003

-0.007

571-5

-0.003

0.003

0.020

0.006

745.5

-0.014

0.3 15

0.013

0.013

0.003

890.5

0.028

0.419

0.333

-0.034

0.000

1124.0

0.087

0.45 1

0.499

0.1 15

-0.0 12

1326-5

O. 140

0.498

0.552

0-211

-0.009

1611.O

O. 189

0.582

0.600

O. 027

O. 166

1870.5

0.217

0.655

0.743

0.286

0.29 1

2122.0

0.24 1

O. 745

0.852

0.3 13

0.350

2391.O

0.259

0.855

0.982

0.320

0.423

2618.0

0.280

0.960

1.134

0.333

0.500

277 1-5

0.297

1.O59

1.426

0.299

0.555

2844.0

0.304

1-405

1.871

4.055

0.61 1

2902.5

0.304

1.783

2.525

-0.34 1

0.65 1

2960-5

0.217

2.402

3.760

-1 .O63

0.688

2996 .O

O. 185

2.763

4.500

- 1-539

0.718

2792. O

O. 157

2.763

4.528

-1 -676

0.72 1

2910.0

O-154

2.800

4.665

-1,717

0.728

1842.O

0.080

2.501

4.123

- 1-676

0.632

Table A.12 Readings from LVDTs located at the level of the main tension tie in
specimen CAPH-B

Table A.13 Readings From LVDTs located at mid-height of specimen CAPH-A

Table A.14 Readings fiom LVDTs located at mid-height of spechen CAPH-B

Table A.15 Readings frorn LVDT rosettes located in end A of specimen CAPH

Table A.16 Readings fiom LVDT rosettes located in end B of specimen CAPH

Table A.17 Strains from strain gauges Iocated in end A of specimen CAPH

Shear

A7

A8

A9

A10

(106,

(106,

(106,

(104

Al!
(104

NIA

NIA

250.0

NIA

-6

NIA

495 -5

NIA

-20

NIA

-2

-4

57 1.5

NIA

-16

NIA

-2

-6

745.5

NIA

-24

NIA

-26

-26

890.5

NIA

-6

NIA

50

1124.0

NIA

- 16

NIA

224

80

1326.5

NIA

-26

NIA

340

140

1611.O

NIA

-38

NIA

460

220

1870.5

NIA

-46

NIA

628

308

2 122.0

NIA

-54

NIA

744

376

239 1.O

NIA

-54

NIA

1050

4 14

2618.0

NIA

-56

NIA

1216

444

277 1.5

NIA

-64

NIA

1332

494

2844.O

NIA

-66

NIA

1418

568

2902 -5

NIA

-74

NIA

1464

744

2960.5

NIA

-66

NIA

1512

1010

2996.0

NIA

-60

NIA

1608

1088

2792 .O

NIA

34

NIA

48

-1 180

2910.0

NIA

148

NIA

-26

- 12

1842.0

NIA

272

NIA

206

rn

--

Table A.17 (Cont'd)Strains From main gauges Iocated in end A of specimen CAPH

Shear

BI

(106,

B2
(10~)

B3

B4

B5

Bd

(106,

(107

(m

(103

250.0

-4

30

34

495 -5

-2

-6

56

64

571.5

-2

-8

10

70

82

745.5

-2

- 12

14

688

644

890.5

-2

-6

12

32

874

782

1124.0

-2

-8

60

72

1090

988

1326.5

-2

-6

f 44

200

1282

1152

161f .O

16

806

702

1556

1510

1870.5

34

1164

1132

1832

1820

2 122.0

46

12

1340

1362

2170

2098

239 1.O

52

12

1554

1630

2536

2394

26 18.0

72

18

1748

1846

2884

2606

277 1.5

58

22

1892

1982

3248

2670

2844.0

116

24

1976

2040

4316

2762

2902.5

132

26

2046

2092

6818

2976

2960.5

144

28

2122

2140

6196

3554

2996. O

158

28

2190

2172

3794

5204

2792.O

160

30

2162

2146

3680

5 174

2910.0

168

30

2170

2154

2846

4458

1842.0

154

30

1714

1674

2348

3806

--

1
i

Table A.18 Strains from strain gauges located in end B of spechen CAPH

Table A.18 (Cont'd) Strains fiom strain gauges located in end B of specirnen CAPH

IMAGE NALUATION
TEST TARGET (QA-3)

APPLIED 4 IMAGE. lnc

--

1653 East Main Street


NY 14609
71
Fax 7161288-5989

Rctiester.
USA
--- Phocre:
W482-W0O
---

: +- - ,

Anda mungkin juga menyukai