Never Uploaded PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 280

SHEAR DESIGN OF PILE CAPS AND OTHER MEMBERS WITHOUT

TRANSVERSE REINFORCEMENT
By
Zongyu Zhou
B.Eng. Tongji University 1982, M.Eng. Tongji University 1987

A ThESIS SUBMITTED IN PARTIAL FULFILLMENT OP

THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in
THE FACULTY OF GRADUATE STUDIES

DEPARTMENT OF CIVIL ENGINEERING

We accept this thesis as conforming


to the required standard

THE UNIVERSITY OF BRITISH COLUMBIA

January 1994

© Zongyu Zhou, 1994


In presenting this thesis in partial fulfilment of the requirements for an advanced
degree at the University of British Columbia, I agree that the Library shall make it
freely available for reference and study. I further agree that permission for extensive
copying of this thesis for scholarly purposes may be granted by the head of my
department or by his or her representatives. It is understood that copying or
publication of this thesis for financial gain shall not be allowed without my written
permission.

(Signature)

Department of V L
The University of British Columbia
Vancouver, Canada

Date

DE-6 (2/88)
Abstract

This thesis deals with the shear design of structural concrete members without trans
verse reinforcement. The three major parts of this study are the transverse splitting
of compression struts confined by plain concrete, the development of a rational design
procedure for deep pile caps, as well as a general study of the shear transfer mechanisms
of concrete beams.
Three-dimensional compression struts that are unreinforced and confined by plain
concrete, as occur in deep pile caps, were studied both analytically and experimentally.
Based on the study results, bearing stress limits are proposed to prevent compression
struts from transverse splitting. The maximum bearing stress depends on the amount of
confinement, as well as the aspect ratio (height to width) of the compression strut.
The proposed bearing stress limit was incorporated into a strut-and-tie model to
develop a rational design procedure for deep pile caps. Two methods are proposed. The
first method is a direct extension of two-dimensional strut-and-tie models used for deep
beams. The second method is presented in a more traditional form in which “flexural
design” and “shear design” are separated. The shear design is accomplished by limiting
the bearing stress at the columns and the piles. The first method is more appropriate for
analysis, while the second method is more appropriate for design. The rationality and
accuracy of the proposed methods are demonstrated by the comparison with previous
test results.
In the final part of this study, the influence of bond between concrete and longitudinal
reinforcement upon the load transfer mechanism of both deep members and slender
members without stirrups are investigated. An interpretation of an important shear

11
failure mechanism is presented.

111
Table of Contents

Abstract ii

List of Tables viii

List of Figures ix

Acknowledgement xvi

1 Introduction 1
1.1 Background 1
1.2 Objectives and Outlines of Thesis 4

2 Analytical Study of Transverse Splitting 8


2.1 Finite Element Modelling 8
2.2 Internal Stress Distributions within Cylinders 9
2.3 Definitions of Geometrical Parameters and First Cracking 10
2.4 Location of First Cracking 10
2.5 Bearing Stress at First Cracking 11
2.6 Summary 12

3 Experimental Study of Transverse Splitting 17


3.1 Description of Test Program . . 17
3.2 Specimen Preparation 17
3.3 Instrumentation and Data Acquisition 18

iv
3.4 Test Observations 19
3.4.1 Internal Cracking 19
3.4.2 Influence of Concrete Confinement 20
3.4.3 Failure Mode • • • . 20

4 Bearing Strength of Concrete Compression Struts 29


4.1 Comparison of Measured and Predicted Cracking Loads • • . . 29
4.2 Comparative Study of Ultimate Bearing Strength 30
4.2.1 Previous Studies of Bearing Strength • . . . 30
4.2.2 Influence of Concrete Compressive Strength 32
4.2.3 Influence of Size Effect • . . . 33
4.2.4 Influence of Loading Geometry 34
4.2.5 Summary 34
4.3 Design Recommendations for Concrete Compression Struts • • . . 35

5 Shear Design of Deep Pile Caps 45


5.1 Introduction 45
5.2 ACI Code Approach for Pile Cap Design 47
5.2.1 Code Procedure for Shear Design 47
5.2.2 Inadequacies of ACI Code Procedure for Shear Design of Deep Pile
Caps 49
5.2.3 Comparison of Different ACI Code Editions 50
5.3 CRSI Approach for Shear Design of Deep Pile Caps 51
5.4 Strut-and-Tie Model Approach 53
5.4.1 CSA Approach for Deep Beams 53
5.4.2 Proposed Design Method (1) 54
5.4.3 Proposed Design Method (2) 57

V
5.5 Summary of Previous Pile Cap Tests 58
5.6 Comparative Study of Different Design Procedures • . . . 63
5.6.1 Comparison of Design Methods 63
5.6.2 Comparison of Prediction Results 64
5.7 Conclusions 66

6 Shear Failure of Beams Without Stirrups 99


6.1 Introduction 99
6.2 Brief Review of the Literature 100
6.2.1 Transition from Deep Beam to Slender Beam 101
6.2.2 Behaviour of Slender Beams 102
6.3 One Interpretation of Shear Failure of Beams without Stirrups 105
6.4 Bond Effect in Longitudinally Reinforced Concrete Beams . 108
6.4.1 Bond Influence in Uncracked Beams 109
6.4.2 Bond Influence in Cracked Beams 112
6.5 Shear Displacements along Cracks 113
6.6 Load Transfer Mechanism 114
6.7 Interpretation of Some Beam Test Results 116
6.8 Conclusions 117

7 Bond Splitting Failure 143


7.1 Introduction 143
7.2 General Bond Action 144
7.3 Previous Experimental Studies 146
7.4 Previous Studies of Ultimate Splitting Failure 147
7.5 Previous Studies of Splitting Initiation 149
7.6 Proposed Design Equation for Bond Splitting Initiation 151

vi
7.7 Some Comments and Conclusions . 153

8 Brief Summary and Further Research 168

Bibliography 173

Appendices 183

A Measured Bearing Stresses and PUNDIT Readings 183

B ACI Code and CRSI Handbook Predictions 193

C Predictions from Proposed Design Method (1) 237

D Predictions from Proposed Design Method (2) 250

VII
List of Tables

3.1 Summary of experimental results 22


3.1 (cont’d) Summary of experimental results 23

5.1 Summary of pile cap test results 68


5.1 (con’t) Summary of pile cap test results 69
5.2 Summary of ACT Building Code and CRSI Handbook predictions (flexure
and bearing) 70
5.2 (con’t) Summary of ACT Building Code and CRST Handbook predictions
(flexure and bearing) 71
5.3 Summary of ACT Building Code and CRSI Handbook predictions (shear). 72
5.3 (con’t) Summary of ACT Building Code and CRST Handbook predictions
(shear) 73
5.4 Comparison of ACT Code and CRSI Handbook predictions: ratio of mea
sured capacity to predicted capacity and failure mode 74
5.4 (con’t) Comparison of ACT Code and CRST Handbook predictions: ratio
of measured capacity to predicted capacity and failure mode 75
5.5 Comparison of proposed design methods with experimental results. . . . 76
5.5 (con’t) Comparison of proposed design methods with experimental results. 77

viii
List of Figures

1.1 Strut-and-tie model for a deep beam or pile cap: (a) the idealized load-
resisting truss; (b) linear elastic stress trajectories with transverse tension
due to spreading of compression and (c) refined truss model with concrete
tension tie to resist transverse tension 6
1.2 Maximum bearing stress to cause transverse splitting in a biaxial stress
field, from Schlaich et al.[3} 7

2.1 Structural modelling of an idealized compression strut: (a) geometry; (b)


finite element mesh; (c) isometric view showing stresses produced by ax
ially symmetric loading and (d) an axially symmetric finite element of
rectangular cross section. Figs.(c) and (d) are from Ref.[15] 13
2.2 Typical internal stress distributions along the central Z axis within the
cylinders: (a) tall cylinder and (b) short cylinder 14
2.3 Influence of D/d on the bearing stress to cause cracking for the case of
H/d=2 15
2.4 Analytical study of the bearing stress at first cracking: (a) influence of
confinement and (b) influence of height 16

3.1 Schematic of test set-up 24


3.2 Photograph of test set-up 25
3.3 Typical relationships of load vs. PUNDIT readings (transit time incre
ments) for the case of H/d = 2 26
3.4 Typical relationships of load vs. axial deformation for the case of H/d = 2 27

ix
3.5 Photograph of the typical failure mode of a concrete cylinder 28

4.1 Comparison of experimental results and analytical predictions: influence


of confinement for various heights 37
4.2 Comparison of experimental results and analytical predictions: influence
of height for various degrees of confinement 38
4.3 Comparison of ACT Building Code bearing strength prediction with ex
perimental results: (a) present investigation and (b) previous tests. . . 39
4.4 Comparison of ACT Building Code bearing strength with Hawkins’s sug
gested equation for various concrete strengths 40
4.5 Influence of specimen height on ultimate bearing strength: (a) the force
flow in a single-punch test; (b) the force flow in a double-punch test and
(c) the force flow in a tall double-punch test 41
4.6 Influence of height on accuracy of ACT Building Code bearing strength
prediction of double-punch tests 42
4.7 Comparison of suggested design equation with the analytical results. . . 43
4.8 Comparison of predictions from suggested design equation with experi
mental results 44

5.1 ACT Code specified critical sections for flexure and shear investigation of
pile cap 78
5.2 A deep two-pile cap 79
5.3 A deep three-pile cap 80
5.4 Comparison of two-way punching shear calculations: (a) the cap with
square column; (b) the cap with circle column 81

x
5.5 CR51 approach for shear design of deep pile caps: (a) allowable shear
stress, v, for one-way shear while w/d < 1.0; (b) allowable shear stress,
v, for two-way shear while wfd < 0.5, from Ref.[36) 82
5.6 Flow chart for ACT and CRST design procedures for pile caps 83
5.7 Strut-and-tie model for deep beam, from Ref.[39] 84
5.8 Loading geometry of compression struts with linearly varying cross section 85
5.9 Various layouts of main reinforcing bars used by Blévot and Frémy, Ref.[32] 86
5.10 Various anchorage lengths used by Clarke, Ref.[44] 87
5.11 Comparison of one-way shear design methods for two-pile caps: (a) plan
view of pile cap; (b) to (d) influence of pile cap depth on column load for
various pile cap widths 88
5.11 (cont’d) Comparison of one-way shear design methods for two-pile caps:
(a) plan view of pile cap; (b) to (d) influence of pile cap depth on column
load for various pile cap widths 89
5.12 Comparison of two-way shear design methods for a typical four-pile cap:
(a) plan view of pile cap; (b) influence of pile cap depth on column load 90
5.13 Comparison of ACT ‘77 predictions with experimental results 91
5.14 Comparison of ACT ‘83 predictions with experimental results 92
5.15 Comparison of ACT ‘[11.8] predictions with experimental results, clause
11.8 considered 93
5.16 Comparison of CRST predictions with experimental results 94
5.17 Comparison of proposed method (1) predictions with experimental results
of all specimens in Table 5.1 95
5.18 Comparison of proposed method (1) predictions with experimental results
of specimens with bunched reinforcement in Table 5.1 96

xi
5.19 Comparison of proposed method (2) predictions with experimental results
of all specimens in Table 5.1 97
5.20 Relationship of measured ultimate bearing stress and confinement on col
umn zones of pile caps in Table 5.1 98

6.1 Distribution of transverse compressive stress for various shear span ratios,
from Mau and Hsu, Ref. [52] 120
6.2 Load near the support: transition from deep beam to slender beam, from
Schlaich et al., right side simple models; left side refined models, Ref.[3]. 121
6.3 Predictions of shear strength versus a/d ratio for tests reported by Kani[13J,
from Collins and Mitchell, Ref.[39]. 122
6.4 Truss model developed by Adebar, Ref.[54J 123
6.5 Truss model developed by Reineck, Ref.[55j 123
6.6 Truss models developed by Al-Nahlawi and Wight, Ref.[53]. 124
6.7 Structural model developed by Muttoni and Schwartz, Ref.[56]. 124
6.8 Structural model developed by Kotsovos, Ref.[57] • • • . 125
6.9 Crack pattern of a beam tested by Kani, Ref.[13] • . • . 125
6.10 Geometric relationship of a crack at reinforcement level 126
6.11 A simply supported and central loaded beam 127
6.12 Linear elastic finite element analysis: modelling of (a) perfect bond and
(b) no bond case 128
6.13 Internal force flows in uncracked beams: (a) perfect bond case and (b) no
bond case 129
6.14 Modelling of bond effect upon transverse splitting of compression struts
in deep beam 130

XII
6.15 Bond influence upon transverse tensile stresses of compression struts in
deep beam 131
6.16 Internal stress distributions of cracked beams due to pure beam action,
from Ref.[39J 132
6.17 Shear stress distributions of cracked beams due to combined beam action
and arch action: (a) modelling of combined beam and arch actions; (b)
tension force in longitudinal reinforcement; (c) shear stress distribution at
section n — n 133
6.18 Shear displacement along cracks in shear span: (a) vertical cracks and (b)
inclined cracks 134
6.19 Shear transfer at cracks by aggregate interlock 135
6.20 A beam with inclined cracks in shear span: (a) strut force due to arch ac
tion; (b) one assumed tension force in reinforcement; (c) another assumed
tension force in reinforcement; (d) measured concrete strain and tension
force in reinforcement from Ref.[71] 136
6.21 Measured force in bar, bond stress and crack locations, from Ref.[59]. . 137
6.22 A load transfer mechanism just before the occurrence of splitting along
longitudinal reinforcement: (a) truss model and (b) tension force in rein
forcement 138
6.23 Design details of beams tested by Kotsovos: (a) a/d = 1.5 and (b) a/d>
2.5, from Ref.[73j 139
6.24 Load-deflection curves of beams shown in Figure 6.5: (a) a/d = 1.5; (b)
aid = 3.3 and (c) a/d = 4.4, from Kotsovos, Ref.[73J 140
6.25 Beam with internal stirrups tested by Chana, from Ref.[58] 141
6.26 Beam with external stirrups tested by Kim et al., from Ref.[74] 141
6.27 Beams tested by Muttoni et al., from Ref.[56j 142

xlii
6.28 Beams by Kuttab et al., from Ref.[75] . 142

7.1 Bond stress—slip relationship, from Gambarova et al., Ref.[80] 155


7.2 Pullout tests: (a) concentric pullout-test specimen; (b) commonly used ec
centric pullout-test specimen and (c) eccentric pullout-test specimen used
by Ferguson, Ref.[60] 156
7.3 The University of Texas beam tests: (a) possible tension force distribution
along the top test bar after the occurrence of inclined crack; (b) side view
of specimen; (c) plain view of top reinforcement; and (d) moment diagram,
adapted from Ferguson and Thompson, Ref.[61] 157
7.4 Stub-beam or cantilever bond specimen, from Gergely, Ref.[77] 158
7.5 Mechanism representation for bond, from Tepfers, Ref.[84] 159
7.6 Bond Splitting Failure Patterns, from Oranguri et al., Ref.[82] 160
7.7 Analysis of Bond Splitting Stresses: (a) bursting and bond stresses; (b)
uncracked elastic state; (c) uncracked plastic state and (d) partly cracked
elastic state. Adapted from Tepfers, Ref.[84} 161
7.8 Comparison of test and prediction for bond splitting initiation, from Tepfers,
Ref.[84] 162
7.9 Comparison of Equation 7.7 with test data on cracking loads carried out
by Kemp and Wilhelm Ref.[66] 163
7.10 The relationship of bond splitting strength and ld/d proposed by Teng and
Ye Ref.[86] for concentric pullout tests 164
7.11 Comparison of Equation 7.2 arid 7.10 with test results from Kemp and
Wilhelm Ref. [66] 165
7.12 Comparison of Equation 7.10 with test results from Tepfers Ref.[84]. . . . 166
7.13 Relationship of bond strength versus ld/db, presented by Eqs. 7.2 and 7.10. 167

xiv
8.1 Load transfer mechanisms of structural concrete members without trans
verse reinforcement: (a) very deep members (aid < 1) as well as more
slender members with no bond between concrete and reinforcement; (b)
slender members with normal bond between concrete and reinforcement. 172

xv
Acknowledgement

The author would like to express his sincerest gratitude to his supervisor Dr. Perry
Adebar for his constant guidance, criticism, and encouragement throughout the course
of research work and in the preparation of this thesis.
Financial support to undertake this study was provided by the Natural Sciences and
Engineering Research Council of Canada.
Finally, this thesis is dedicated to the author’s family for their understanding and
support, and especially to his wife, Jialing, and son, George, for their patience through
the duration of this study.

xvi
Chapter 1

Introduction

1.1 Background

Plain concrete is strong and reasonably ductile in compression, but is weak and brittle
in tension. Thus reinforcing steel bars are usually provided in concrete members to resist
any tension force. In a reinforced concrete beam subjected to bending, for example,
the flexural tension force is resisted by longitudinal reinforcing bars. For more complex
loading conditions, such as combined shear and bending, structural concrete members
are usually reinforced with both longitudinal and transverse reinforcement.
Structural concrete members reinforced with both longitudinal and transverse rein
forcement can be designed using simple truss models or in the more general case strut-
and-tie models [1-3]. It has been suggested by Schlaich et al. [3] that structures be
subdivided into B-regions and D-regions for the purpose of design. B-regions are where
the Bernoulli hypothesis of plane strain distribution is reasonable (B standing for beam
or Bernoulli). Internal stresses in B-regions can be derived from the sectional forces with
out considering the details of how the forces are applied. Truss models for B-regions with
stirrups involve compression and tension chords to model the flexural stresses, tension
ties to model the stirrups and diagonal compression struts to model the cracked concrete
web. Truss models that have been developed for the shear design of beams with stirrups
are rational and work well [4-8].
In D-regions (D standing for discontinuity or disturbance) the strain distribution is

1
Chapter 1. Introduction 2

significantly nonlinear, and the details of how the forces are applied are very important.
The more general form of truss models, strut-and-tie models can be used for both B- and
D-regions. The theoretical justification for strut-and-tie (truss) models comes from the
theory of plasticity (limit analysis) [1]. Strut-and-tie models fall within the domain of
lower-bound limit analysis. As strut-and-tie models allow designers to freely choose the
load path, a sufficient amount of uniformly distributed reinforcement must be provided to
distribute cracks, thereby ensuring adequate ductility (plastic deformation) for internal
redistribution of stresses so that the assumed strut and tie system can eventually develop.
Unfortunately, it is not always practical to provide the reinforcement needed for crack
control and ductility. Traditionally many structural concrete members are designed and
built without transverse reinforcement. In North America, both the ACT Building Code
[9] and the Canadian Concrete Code [10] allow the following structural concrete members
to be designed without stirrups: slabs and footings; concrete joists; beams with a total
depth less than 10 in. (254 mm); wide beams having a total depth less than the width
of the web; and tee beams having a total depth less than 2.5 times the thickness of the
flange.
Many of the current design procedures for members without stirrups are less than
satisfactory. For example, the design procedures currently specified in the ACT Building
Code and the Canadian Concrete Code for deep pile caps (footings) are inappropriate.
The ACT Building Code procedure for the shear design of footings supported on piles
(pile caps) is the same sectional approach used for footings supported on soil and for
slabs. The procedure involves determining the section thickness which gives a concrete
contribution V greater than the shear force applied on the code defined critical section.
While this approach is reasonable for slender footings supported on numerous pile, it is
not appropriate for deep pile caps which are entirely a D-region. On the other hand, ap
plying strut-and-tie (truss) models to deep pile caps, which are usually without transverse
Chapter 1. Introduction 3

reinforcement and have very limited ductility, is also questionable.


The internal force flow within a deep pile cap can be represented by the simple
strut-and-tie model shown in Figure 1.1(a). The column load is transmitted directly to
the piles by compression struts. In order to prevent the piles from being spread apart,
tension ties must be provided. The results of an experimental study on pile caps [11] has
demonstrated that such a simple strut-and-tie model is a better model for deep pile caps
than the basis of the traditional design procedures for pile caps.
A prismatic compression strut is a highly idealized representation of the stress state
between a column and pile in a deep pile cap. In reality, the compressive stresses in
the compression strut spread out and transverse tensile stresses are created due to strain
compatibility [see Figure 1.1(b)]. A refined strut-and-tie model for the compression strut,
which includes a tension tie to model the transverse tension, is shown in Figure 1.1(c).
As the compression struts which transfer load in deep pile caps are usually unreinforced,
the tension tie must be provided by concrete tensile strength which is only about a tenth
to a fifteenth of the concrete compressive strength. It is obvious that the failure of
unreinforced compression struts could be initiated by transverse splitting.
Figure 1.2 from Schlaich et al. [3] shows the influence of the “amount of spreading”
on the bearing stress to cause transverse splitting for a plane stress field (appropriate for
the case of a deep beam). Based on Figure 1.2 Schlaich et al. suggest that the concrete
compressive stresses within an entire unreinforced D-region can be considered safe if the
maximum bearing stress in all nodal zones is limited to O.6f (or in unusual cases O.4f)

[3]. However what bearing stress limit is appropriate for compression struts confined by
surrounding plain concrete, such as occurs in deep pile caps, is not known. This hinders
the application of strut-and-tie models for the design of deep pile caps.
While some truss models have been developed for slender members without transverse
reinforcement (i.e., B-regions), there is still no generally accepted interpretation of the
Chapter 1. Introduction 4

shear failure mechanism for such members. Experimental investigations have shown that
the bond between concrete and longitudinal reinforcement has a large influence upon
the shear capacity of reinforced concrete beams without transverse reinforcement [12-14].
Truss models that have been developed for members without transverse reinforcement
do not adequately account for the influence of bond, particularly bond splitting failures
at the base of the critical diagonal crack.

1.2 Objectives and Outlines of Thesis

This thesis deals with the shear design of structural concrete members without stirrups,
and includes three major parts. The first part is a study of the transverse splitting of
compression struts confined by plain concrete as occurs in deep pile caps. The second
part involves the development of a rational design procedure for deep pile caps, which
makes use of the bearing stress limit developed in the first part.
In deep pile caps the column load is transmitted to the piles by direct compression
struts. In more slender members a direct compression strut cannot form and the load
transfer mechanism is influenced by the bond between concrete and reinforcing bars.
A more general study of the load transfer mechanisms in members without transverse
reinforcement is undertaken in the third part of this thesis.
In Chapter 2 three-dimensional compression struts confined by plain concrete are
studied analytically to investigate the internal stress distributions. The objective is to
determine the bearing stress at which first cracking occurs due to the transverse tension.
The influence of various geometrical parameters upon the bearing stress at first cracking
is presented.
In Chapter 3 the transverse splitting of compression struts is investigated experimen
tally. The results from tests on plain concrete cylinders of various sizes loaded over a
Chapter 1. Introduction 5

constant bearing area are presented.


Design recommendations for the bearing stress limit of unreinforced concrete com
pression struts based on a combination of the analytical and experimental results are
proposed in Chapter 4. Comparisons are also made with previous studies for ultimate
bearing strength.
In Chapter 5, the proposed bearing stress limit of compression struts is incorporated
into a simple strut-and-tie model to develop a design procedure for deep pile caps. Two
design procedures are proposed. The first design method is appropriate for a detailed
analysis, while the second design method is more suitable for design. A comparison of
predictions from the proposed procedures, as well as the ACT Code and CRSI Handbook
procedures are carried out for available pile cap test results.
In Chapter 6, a general study of the load transfer mechanisms in members without
stirrups is carried out. Recently developed truss models are reviewed. The shear failure
mechanism of beams without stirrups is studied from the compatibility of the critical
diagonal cracking propagation. The effect of bond between concrete and reinforcing bars
upon shear resistance mechanism are also studied. Based on the study in this chapter,
some interesting conclusions are drawn. Finally in this chapter, experimental results
from beam shear tests are interpreted.
As a natural extension to Chapter 6, the bond splitting failure along reinforcing bars
is investigated in Chapter 7. Based on previous studies on bond, an empirical equation is
proposed to predict the tension force causing bond splitting in beams without transverse
reinforcemeilt.
Finally, some major conclusions from this study and a few recommendations for fur
ther investigations are summarized in Chapter 8.
Chapter 1. Introduction 6

V V

comPressio/,.,.N

,
I/ /
J. N/ Ti
nocal zone 1
tie
Lension
T = V I tan e
D=V/sine

(a)

(b)

///

(c)

Figure 1.1: Strut-and-tie model for a deep beam or pile cap: (a) the idealized
load-resisting truss; (b) linear elastic stress trajectories with transverse tension due to
spreading of compression and (c) refined truss model with concrete tension tie to resist
transverse tension.
Chapter 1. Introduction

fb

Failure
f,
C

Transverse Cracking

0
1 2 3 4 5 6 7 8
b
9
a

1. b

Figure 1.2: Maximum bearing stress to cause transverse splitting in a biaxial stress field.
from Schlaich et al.[3J.
Chapter 2

Analytical Study of Transverse Splitting

In order to better understand the transverse splitting phenomenon in deep pile caps and
develop an appropriate strength criterion for plain concrete compression struts, idealized
three-dimensional compression struts are studied analytically. The study focuses on the
initial cracking within struts rather than the post-cracking stage, which is a much more
complicated nonlinear problem.

2.1 Finite Element Modelling

The compression struts are idealized as cylinders of exterior diameter D, and height H,
subjected to concentric axial compression over a constant size circular bearing area of
diameter d, both on top and bottom. See Figure 2.1(a). Linear elastic finite elements
were used to study the internal triaxial stress distributions, and to determine when and
where first cracking occurs within the cylinders.
For the problem at hand, the geometry, material properties and external loading are
all axis-symmetric so that the problem is mathematically two-dimensional. The static
displacements and stresses are independent of angle 0, circumferential displacement v
is zero and material points have only u(radial) and w(axial) displacement components.
Non-zero stress components are O, o, o,. and Tzr. The stress o is a principal stress
and o, r and Tzr are components of the two other principal stresses. In the analysis,
axis-symmetric 4-node quadrilateral elements were used [15]. There are two integration
points in each one of radial, axial and circumferential directions. Stresses at the central

8
Chapter 2. Analytical Study of Transverse Splitting 9

point of each element were calculated. See Figure 2.1(c) and (d).
The finite element mesh used is shown in Figure 2.1(b), which also shows the geom
etry, grid layout and boundary conditions. Due to symmetry of the problem and the
characteristic of the elements used, only a quarter of the cylinder is shown. The radial
displacement u = 0 was prescribed at all nodes that lie on the Z axis and the rigid
body motion along Z direction was restrained by prescribing w = 0 at the mid-height
of the symmetrical cylinder. But no restrictions were imposed for vertical displacement
on the Z axis and horizontal displacement along the radial direction with the exception
of one location at the corner. By adding more elements either horizontally or vertically
(i.e., increasing D or H), the influence of various loading geometry on the internal stress
distributions was investigated.

2.2 Internal Stress Distributions within Cylinders

Based on the numerical analysis results the internal stress distributions along the central
Z axis are shown qualitatively in Figure 2.2. The vertical stress component r has
its maximum value beneath the loading plate and decreases gradually with increasing
distance from the loading plate. The stress o is distributed uniformly on the transverse
cross section at a certain distance away from the loading plate. The stress components 0
and ,. on the Z axis are equal and may be negative (compressive), or positive (tensile),
as well as zero at locations where the stress cr becomes uniformly distributed.
The internal stress distribution within the cylinder can be divided into three zones
[see Figure 2.2(a)]. A zone of triaxial compression exists immediately below the loading
plate. Further down from the loading plate is a zone with horizontal biaxial tension and
vertical uniaxial compression. At the mid-height of relatively tall cylinders is a zone of
vertical uniaxial compression. In short cylinders, the third zone (of uniaxial compression)
Cha
p
0ter 2. Analytical Study of Transverse Splitting 10

does not exist. See Figure 2.2(b).

2.3 Definitions of Geometrical Parameters and First Cracking

The geometry of the problem can be summarized in terms of two parameters, namely
the ratio of the cylinder diameter to the loaded area diameter, D/d, and the ratio of the
cylinder height to the loaded area diameter, H/d.
The stress field within a cylinder can be expressed in terms of the four stress compo
nents oo, o, o,. and re,. or by three principal stresses and corresponding principal stress
directions. The stress O is always a principal stress. It was assumed that cracking will
occur when the maximum principal tensile stress reaches the concrete tensile strength.
The location of the maximum principal tensile stress defines where the first cracking oc
curs. The influence of the other principal stress components upon the cracking strength
was neglected for simplicity.

2.4 Location of First Cracking

After the internal stress distributions within the cylinders were examined, it was found
that the maximum tensile stress would occur either on Z axis (at the centre of the
cylinder) or near the exterior surface of the cylinder depending on the geometry. Hence,
the location of first cracking could be located either in the inside of the cylinder or near
the outside of the cylinder.
The influence of loading geometry upon the location of first cracking is shown in
Figure 2.3, where the predicted bearing stress at cracking versus the ratio of Did are
summarized for the case of H/ d = 2. There are two curves in the figure: one indicating
cracking inside, the other cracking outside. It can be seen that for small ratios of Did,
first cracking occurs on the surface, while for D/d > 1.5 the first cracking occurs at the
Chapter 2. Analytical Study of Transverse Splitting 11

centre of the cylinder.


As cracking outside occurs only when there is very little confinement, which is not of
much practical interest, oniy cracking inside was considered below.

2.5 Bearing Stress at First Cracking

Figure 2.4 summarizes the analytical results regarding the the influence of D/ d ratio and
H/d ratio upon the bearing stress at first cracking (inside). In order to plot the bearing
stress as a function of the compressive strength of concrete, the concrete tensile strength

f’ was assumed to one-fifteenth of the concrete compressive strength f (i.e., f’ = f/l5).


Figure 2.4(a) shows the relationship between bearing stress at first cracking and D/d
for specific H/d values. For small ratios of D/d (i.e., close to 1.0), the bearing stress
at first cracking is independent of height and tends to be very large. In this range
the uniaxial compression is more critical than the transverse tension. In the range of
D/d > 1.5 2.0, increasing the amount of confinement (i.e., increasing D/d further)
will increase the bearing stress at first cracking, but not very much for small and medium
H/ d values. The significant transverse tension is introduced due to the spreading-out of
compressive force. The transverse tensile stresses will reach concrete tensile strength at
relatively low external load in this range.
Figure 2.4(b) shows the relationship between bearing stress at first cracking and H/d
for specific D/ d values. Figure 2.4(b) illustrates a similar trend to Figure 2.4(a). In
the range of H/ d < 1, the bearing stress at first cracking is independent of confinement
and tends to be very large. The localized uniaxial compressive stress field between two
loading plates is dominant within the compression strut. For H/d > 1.5, increasing
height will increase the bearing stress at first cracking. But increasing height further
(i.e., Hid > 4 ‘-‘ 6) does not have significant influence upon the bearing stress at first
Chapter 2. Analytical Study of Transverse Splitting 12

cracking.
It can be seen that the minimum bearing stress at first cracking is approximately
O.6f at D/d = 2 and H/d = 1.5. More discussions about aspect ratio (H/d) upon
the bearing stresses at first cracking will be given in the comparative study of ultimate
bearing strength in Chapter 4.

2.6 Summary

The numerical investigation indicates that the loading geometry has a significant influence
upon the internal stress distribution within a confined compression strut. See Figure 2.2.
Qualitatively, the response of an idealized compression strut (cylinder) to external load
can be described as follows. For a compression strut without confinement (D/d = 1),
the strut is uniaxial stressed. The compressive stress is constant over the height and
is equal to the bearing stress. However, if there is confinement outside the loaded area
(D/d> 1), the additional concrete will be mobilized to sustain the external load, and the
axial compressive stress along the central axis of the strut will decrease continually with
increasing distance away from the loading plate. As expected, higher bearing stresses may
be possible due to the confinement. On the other hand, as a result of strain compatibility

the transverse tensile stresses are introduced. Because concrete tensile strength f’ is
about a tenth to a fifteenth of concrete compressive strength f, internal cracking of
the concrete compression strut (cylinder) may occur due to the weak tensile strength at
relatively low compression stresses.
Chapter 2. Analytical Study of Transverse Splitting 13

d/2

TT

H
H12

J
1c

D12

(a) (b)

z,w

4 3x’
\;
Axis-
—r,u
2

(C) (d)

Figure 2.1: Structural modelling of an idealized compression strut: (a) geometry; (b)
finite element mesh; (c) isometric view showing stresses produced by axially symmetric
loading and (d) an axially symmetric finite element of rectangular cross section. Figs.(c)
and (d) are from Ref.[15].
Chapter 2. Analytical Study of Transverse Splitting 14

5.r
Fy

2 — -

bd H/2
fb

Tension Compression
H (a)

I
I
D e.J

ZS(L
Tension Compression

(b)

Figure 2.2: Typical internal stress distributions along the central Z axis within the
cylinders: (a) tall cylinder and (b) short cylinder.
Chapter 2. Analytical Study of Transverse Splitting 15

f /
3- H/d=2
/
fc’/ft’=15
2
Cracking Outside

Cracking Inside

I I
0
1 2 3 D
d

Figure 2.3: Influence of D/d on the bearing stress to cause cracking for the case of
H/d=2.
Chapter 2. Analytical Study of Transverse Splitting 16

H/d = 10
3 H/d—8
HId =6

H/d =4

H/d =3

H/d—2

0
(a)
B
D

Did =7
D/d—6
D/d =5

D/d—4
2
D/d =3

D/d =2
I

0
(b)
12
H
d
Figure 2.4: Analytical study of the bearing stress at first cracking: (a) influence of
confinement and (b) influence of height.
Chapter 3

Experimental Study of Transverse Splitting

In order to confirm the transverse splitting phenomenon within compression struts, and
to investigate the influence of H/d and D/d ratios upon the bearing stress at first cracking
experimentally, a series of experiments were conducted on plain concrete cylinders.

3.1 Description of Test Program

Table 3.1 summarizes the properties of the 60 different specimens which were tested.
The diameter of the concrete cylinders varied from 6 inches (152 mm) up to 24 inches
(610 mm), while the diameter of the circular bearing plate was constant at 6 inches (152
mm). The height of the cylinders varied from 9 inches (230 mm) up to 36 inches (914
mm). The variation of geometrical parameters covers the range of 1 < Did 4 and
1.5 Hid 6, which is believed to be the practical range. Note that D6-l2 in Table
3.1 denotes a 6 inch (152 mm) diameter cylinder which is 12 inch (305 mm) high.

3.2 Specimen Preparation

The specimens were constructed from two batches of concrete. Both batches were sup
plied by a local ready-mix supplier (25 MPa was specified) and had crushed stone as
aggregate with a maximum size of 3/4 inch (19 mm). The specimens were cast in sono
tube forms with plywood bases, which were stripped after about 7 days. The specimens
continued to cure in the laboratory for an additional 50 to 60 days before being tested.

17
Chapter 3. Experimental Study of Transverse Splitting 18

The specimens were always covered by plastic sheets during curing time. As it turned
out, Batch A concrete had a cylinder compressive strength of 30 MPa (4350 psi), while
Batch B concrete had a compressive strength of only 20 MPa (2900 psi) at testing.

3.3 Instrumentation and Data Acquisition

The test set-up used is shown in Figure 3.1 and 3.2. The specimen was placed on the
testing table of a 400 kip capacity Baldwin testing machine and loaded through 6 inch di
ameter, half-inch thick steel discs. Special care was taken to ensure the specimen was not
loaded eccentrically. Two Linear Variable Differential Transducer (LVDT) displacement
transducers were mounted on two steel bars which were connected to the top loading
(steel) disc. Two transducers (one transmitter and one receiver) of a Portable Ultrasonic
Non-Destructive Digital Indicating Tester (PUNDIT) [16] were put at the height where
the maximum tensile stresses were expected based on the analytical results. Two small
wood pieces were glued to the side face of the specimen to act as supports for the two
transducers. An elastic rope was used to tighten the transducers to the specimen, and
at same time to give as little transverse constraint to the specimen as possible.
The axial deformations of specimens were measured by the two LVDTs. The axial
deformation versus the applied loads were plotted with the help of an X- Y plotter. The
measured axial deformation included the deformations of the steel discs, plaster and the
concrete cylinder so the deformation of the steel discs and plaster had to be eliminated
before obtaining the true axial deformation of a concrete cylinder.
An important piece of equipment used in the experimental investigation was the
PUNDIT [16]. The PUNDIT measures the travelling time of an ultrasonic pulse whose
velocity is proportional to the density of the material. The transit time is displayed
by three ‘in-line’ numerical indicator tubes. The transit time was recorded by hand at
Chapter 3. Experimental Study of Transverse Splitting 19

predetermined load values. If the last digit either remained at a constant value or would
oscillate between two adjacent-values with a bias for one of the values, this value was
recorded. If the last digit oscillated between two adjacent-values evenly, the mean of the
two values was recorded. In this way, an accuracy of 0.05 microsecond was possible.

3.4 Test Observations

Load was applied at an approximate rate of 2.75 MPa every minute, continuously until
failure. Major test observations are summarized below.

3.4.1 Internal Cracking

The internal cracking within concrete cylinders could not been seen by eye so that the
PUNDIT was used to detect it. By observing the change in ultrasonic pulse transit time,
the internal cracking was determined indirectly.
While load was being increased, the transit time of ultrasonic pulse passing through
the specimens (i.e., from transmitter to receiver) also increased. The rate of change
of the ultrasonic pulse travel time varied at different loading stages. The transit time
increased very little or remained almost stable at low load levels, and then increased
steadily at medium load levels. Finally, the transit time increased very quickly just
before failure. The quick increase of transit time was always a sign that failure would

soon occur. Typical load PUNDIT reading (transit time increments) relationships for
specimens with H/d = 2 are shown in Figure 3.3.
The relationship between load and PUNDIT reading (transit time increment) is ini
tially linear. As the velocity of ultrasonic pulses travelling in a solid material depends on
the density and elastic properties of the material, the transit time depends on both the
path length and the quality of concrete. In the linear stage, the increase in transmit time
Chapter 3. Experimental Study of Transverse Splitting 20

is believed to be a result of the transverse deformation of cylinders due to the Poisson’s


ratio effect (i.e., the small increase in path length). In the post-linear stage, both the
quality of the concrete and the transverse deformation of concrete are believed to have
an influence upon the transmit time increment. First cracking is believed to be indicated
by the onset of nonlinearity. The non-linearity in the PUNDIT readings for 6 inch (152
mm) specimens is believed to be due to the micro-cracking that occurs when concrete is
subjected to uniaxial compression.

3.4.2 Influence of Concrete Confinement

It is obvious that the confinement has an influence on the structural behaviour of concrete
cylinders subjected local compression on both the top and bottom surfaces. Typical
bearing stress versus axial deformation relationships for specimens with H/ d = 2 are
shown in Figure 3.4, in which the deformations coming from the steel discs and plaster
have been eliminated and the applied loads have been converted to bearing stresses.
It can be seen that increased confinement will make the axial stiffness of the spec
imens larger and allow higher ultimate bearing strengths to be achieved. In contrast,
confinement has a minor influence on the bearing stress at first cracking. See both Figs.
3.3 and 3.4.

3.4.3 Failure Mode

In general, the failure of specimens occurred suddenly and with a loud noise. Typically,
three or four radial cracks split the specimen into approximately equal segments. Con
crete “cones” usually formed in the triaxial compressed zone below the loading plates.
Figure 3.5 shows a typical concrete cylinder after failure.
The 6 inch (152 mm) diameter and 8 inch (203 mm) diameter specimens were less
brittle than other specimens. See Figure 3.4. In the 8 inch (203 mm) specimens surface
Chapter 3. Experimental Study of Transverse Splitting 21

cracks were observed before failure. That is, the specimens resisted additional load after
exterior cracks were observed. Note that the 8 in (203 mm) specimens were predicted to
crack on the outside before the inside. See Figure 2.3. The 6 inch (152 mm) specimens
were standard cylinders that failed in compression.
ei

CD C])
I : I--i I--i I_ , : I— : :‘ : i— .• .• i— ,— — — ;‘ ,— I. oo 00 00 00 00 ) CD

C C C I--s I--i I.- I.-. I—i I--i C C C.3 CZ C3 I— I I ç C.C c t’.) t.) t\) t\) t t t’.D s

CAZ L’3 I— C L3 I— C ‘3 I C t’ I—i CD


*

C)
o

CD

‘CD I
—‘ I-s
t’. t3 ISZ t’. t’.Z I t\1 t.. t’. I- .) k1 I— I—i t’3 t t. L2
I
- t.Z 00 00 t’Z : t c
I

‘.— I-”
c-I
I-s
F
CI)

I-s
CD
,

. —.‘

CI)
.3 C C’ I— —1 ) . C I—i C CT © © —4 © 00 00 —1 t’Z C —4 CD

O:s

tN3
Chapter 3. Experimental Study of Transverse Splitting 23

Table 3.1: (cont’d) Summary of experimental results.

Specimen Concrete Bearing Stress Max. Bearing


Designation* Batch at Cracking Stress
(MPa) (MPa)
D14-9-1 B 15.9 32.5
D14-9-2 B 14.6 31.7
014-9-3 B 17.1 33.2
014-18-1 B 26.8 48.1
014-18-2 B 25.6 50.0
D14-18-3 B 24.4 48.8
D14-36-1 B 34.2 58.1
014-36-2 B 28.1 53.2
D14-36-3 B 29.3 51.7
018-9-1 B 26.8 44.4
018-9-2 B 19.5 42.2
D18-9-3 B 20.7 44.4
D18-12-1 A 26.8 40.2
018-12-2 A 26.8 43.4
D18-12-3 A 29.3 42.0
D18-12-4 A 29.3 41.7
D18-18-1 B 34.2 61.0
018-18-2 B 31.7 55.6
018-18-3 B 29.3 56.9
D18-36-1 B 37.8 71.5
D18-36-2 B 37.8 68.3
018-36-3 B 39.1 68.3
024-12-1 A 32.9 59.8
D24-12-2 A 35.4 65.5
024-12-3 A 34.1 62.0
*D612 denotes 0 = 6 in; H = 12 in
1 MPa = 145 psi
1 in = 25.4 mm
Chapter 3. Experimental Study of Transverse Splitting 24

d=6 in.
Displacement
r1
ansducer

Plaster

Specimen H=Varies
UNDIT PUNDIT
receiver

Steel disc

D =varies

Figure 3.1: Schematic of test set-up.


Chapter 3. Experimental Study of Transverse Splitting 25

Figure 3.2: Photograph of test set-up.


Chapter 3. Experimental Study of Transverse Splitting 26

80

70 • Cracking

Cl) 60-
D24-12--
G)
-4-i
C,)
50 -

D18-12

40
- / D12-12

30 -
D8-12

D6-12
20 -

io4
1
(: I
0 I

0 1 2 3 4 5 6 7
PUNDIT Transmit Time Increment
(microseconds)

Figure 3.3: Typical relationships of load vs. PUNDIT readings (transit time increments)
for the case of H/d = 2.
Chapter 3. Experimental Study of Transverse Splitting 27

80

60- D24-12
U)

Cl) 50 -

Dl 8-12

ctI 40 -

D12-12 D8-12

30 -

D6-12
20 -

10 -

0 I I I
0 0.5 1 1.5

Displacement (mm)

Figure 3.4: Typical relationships of load vs. axial deformation for the case of H/d = 2.
Chapter 3. Experimental Study of Transverse Splitting 28

I
Figure 3.5: Photograph of the typical failure mode of a concrete cylinder.
Chapter 4

Bearing Strength of Concrete Compression Struts

4.1 Comparison of Measured and Predicted Cracking Loads

Figure 4.1 and 4.2 compare the measured cracking loads with the linear elastic predic
tions. Figure 4.1 shows the influence of confinement (D/d) for various height, while
Figure 4.2 shows the influence of height (H/d) for various degrees of confinement. The
tensile strength used in the prediction,f = f/13, was chosen to give the best fit. Con
sidering that only one parameter (f/f) is adjusted, there is a very good correlation
between the analytical prediction and the experimental results regarding the influence of

Did and H/d on the bearing stress to cause first cracking.


It should be pointed out that the large deviation of some experimental data points
from analytical prediction in the case of H/cl = 2 of Figure 4.1 is not surprising. As
discussed in Section 2.4 (Figure 2.3) the analytical results predict that there is a range
of first cracking inside as well as first cracking outside. The ratio of D/d = 1.33 for
the 8 inch (203 mm) diameter and 12 inch (305 mm) height specimens lies in the range
of first cracking outside. Test observation also confirmed that some external cracking
occurred before the specimens failed. The early external cracking is believed to be the
main reason for this deviation from the analytical prediction. It should be remembered
that the predictions shown in Figure 4.1 represent that of first cracking inside.

29
Chapter 4. Bearing Strength of Concrete Compression Struts 30

4.2 Comparative Study of Ultimate Bearing Strength

While this study is concerned primarily with the bearing stress to cause initial transverse
cracking, the test results also give valuable information for ultimate bearing strength. A
comparison of the bearing stress which caused failure in the present tests and the bearing
strength predicted by the ACT Building Code [9] was therefore carried out. According to
the ACT Building Code the maximum bearing strength of concrete is 0.85f, except when
the supporting surface area A
2 is wider on all sides than the loaded area A
, the bearing
1
strength is multiplied by ,
1
/
2
A A but not more than 2. Note that D/d = i.JA
/
2 Ai.
Surprisingly it was found that the ACT approach is unconservative in the range 1.5 <
D/d < 3.5. See Figure 4.3(a). Thus, it seems that further study on ultimate bearing
strength is warranted.

4.2.1 Previous Studies of Bearing Strength

As bearing strength is important in the design of many concrete structures, a great


number of investigations have been done on the bearing strength of concrete.
It is generally believed by most investigators that the principal variables influencing
the concrete bearing strength are:

1. The geometry of loaded area and specimen;

2. the nature of the supporting bed under the specimen;

3. the supported area of specimen;

4. the concrete strength, and

5. the specimen size.

The main conclusions based on the previous experimental work are that:
Chapter 4. Bearing Strength of Concrete Compression Struts 31

1. The failure is due to the punching out of an inverted cone of concrete beneath the
loaded area and the radial pressures exerted by this cone split the specimen;

2. The ratio of bearing strength over concrete compressive strength increases contin
uously for increasing confinement, ie. increasing the ratio of D/d;

3. There are cracking loads and failure loads which are sometimes different and depend
upon the size of bearing plates and the height of specimens;

4. The larger the loading plate size (smaller D/d) and smaller the relative height of
specimen, the greater is the difference between failure load and external cracking
load;

5. Specimens supported on a compressible bed give less bearing strength than for
similar specimens on a rigid support medium;

6. Specimens subjected to localized forces from opposite ends exhibit lower bearing
strength compared to localized compression from one end only;

7. The higher the compressive strength of concrete, the lower the ratio of ultimate
bearing strength to concrete compressive strength;

8. The ratio of bearing strength over concrete compressive strength decreases with the
increase in specimen size,

Conclusions (3)—(8) are mainly from Niyogi’s tests [17].


However, there are different opinions regarding the influence of specimen height. Some
authors considered that bearing strength over concrete compressive strength fb/f’ is
independent of specimen height [18-26]. But Niyogi [17] concluded that for double punch
tests the ratio f&/f increases with the increase in the specimen height and supporting
Chapter 4. Bearing Strength of Concrete Compression Struts 32

area. For single punch tests the bearing strength decreases with increasing specimen
height in some ranges of geometrical parameters, and in the reverse (the bearing strength
decreases with decreasing specimen height) in other ranges of geometrical parameters.
Double-punch tests involves the specimens being loaded symmetrically using bearing
plates on both the top and bottom surfaces. In single-punch tests the base of the specimen
is set directly on the lower platen of the testing machine and the upper platen is used to
load the bearing plate.
Figure 4.3(b) compares the ACT Building Code bearing strength prediction with pre
vious tests conducted by Shelson [19], Au and Baird [20], Douglas and Trahair [23],
Middendorf [24], Hawkins [26], Chen and Trumbauer [27], as well as Niyogi [17]. As
shown in Figure 4.3(a) for the present investigation, this figure also indicates that the
ACT prediction is unconservative for quite a number of data points. After the unconser
vative data points were carefully examined, it was found that the major reasons for the
lower ratios of bearing strength to concrete compressive strength f are the influence of
concrete compressive strength itself, size effect and loading geometry. These are discussed
in detail below.

4.2.2 Influence of Concrete Compressive Strength

The Commentary to ACT 318-89 Section 10.15 makes reference to Hawkins [26] who
suggested the following expression for the bearing strength of concrete

fb = f+1
J1(/A
50
/
2 A — 1) (4.1)

where f is in psi units. That is, the enhancement in bearing strength due to confining
concrete is proportional to the tensile strength of concrete. In the ACT Building Code this
equation has been simplified so that the bearing strength enhancement is proportional to
Chapter 4. Bearing Strength of Concrete Compression Struts 33

the concrete compressive strength. Figure 4.4 compares the ACT approach with Equation

4.1 for different concrete strengths. Note that when the concrete compressive strength is
high than 5,000 psi (34.5 MPa), the original expression suggested by Hawkins [26] gives
lower bearing strengths than the ACT Building Code. The low strength reduction factor
of 0.70 used in the ACT Building Code for bearing on concrete partly compensates for
this over-simplification.

4.2.3 Influence of Size Effect

A second factor which contributes to lower bearing strengths in Figure 4.3 is size effect.
The ACT Building Code approach is based on tests of relatively small specimens, while
data points shown in Figure 4.3(a) are from large specimens. Since bearing failures
involve fracture of concrete or crack propagation process due to indirect tension, there is
a significant size effect involved.
Based on the result of 42 double-punch tests, in which the specimens had the geo
metrical proportions D = H = 4d, Marti [28] concluded that the size effect was in good
agreement with Baant’s (nonlinear fracture mechanics) size effect relation [29, 30]. Thus
the bearing strength of concrete is proportional to

fo cc (4.2)
)Da

where D is the maximum aggregate size and .X is an empirical constant, which was found
to be 38.0 and 68.5 in Marti’s two series of tests [28]. For the geometrically identical
specimens that were made with the same concrete, f = 3425 psi (23.6 MPa), and ranged
in size from 8 to 128 times the maximum aggregate size (i.e., the size varied by a factor
of 16), the bearing strength decreased by a factor of 1.6 from the smallest specimen to
the largest specimen.
Chapter 4. Bearing Strength of Concrete Compression Struts 34

4.2.4 Influence of Loading Geometry

The most important factor to contribute to the actual bearing strength being lower than
the ACT Building Code prediction is the geometry of loading.
The majority of the test results, upon which the ACT Building Code procedure is
based, are from single-punch tests. In the present investigation, as well as some previ
ous studies, the specimens were loaded by double-punch method. In single punch tests
transverse tensile stresses are created at one end only and the lower platen of the testing
machine restrains the expansion of the specimen. In double-punch tests the compression
at both ends of the specimen results in higher tensile stresses at the mid-height of the
specimen as shown in Figure 4.5(b) and the restraint is eliminated. When the specimens
are relatively tall [Figure 4.5(c)], the double-punch tests produce similar results as the
single-punch tests except for the restraint issue [Figure 4.5(a)]. Figure 4.6 illustrates that
the accuracy of the ACT bearing strength depends on the H/d ratio for double-punch
tests.
In addition to the influence of specimen height discussed above, the data points in
Figure 4.3 also show that the effect of confinement is not as large as the ACT procedure
predicts in the range of small D/d ratios, where quite a number of test results either are
unsafe or have a very low safety margin. Tt is interesting to compare Figures 4.3 and
4.6 with Figure 2.4. Most lower bearing strength data points are located in the range
1 < H/d < 3 and 1.5 < D/d < 3.5, exactly where the analytical results indicate the
least bearing stress to cause first cracking.

4.2.5 Summary

On the basis of present investigation, some conclusions which supplement the conclusions
of previous studies (Section 4.2.1) are that:
Chapter 4. Bearing Strength of Concrete Compression Struts 35

1. Bearing failure is due to transverse splitting, provided that the supporting surface
area is wider on all sides than the loaded area;

2. The bearing stress at external visible cracking could be less than or equal to the
bearing stress at failure depending upon whether first cracking occurs outside or
inside;

3. Bearing strength depends, to a major extent, upon the loading geometry, which
includes the loading method, surrounding confinement (D/d) as well as specimen
heights (H/d);

4. Due to either higher concrete compressive strength or size effect or loading geome
try, the bearing strength according to the ACT Code can be unconservative.

4.3 Design Recommendations for Concrete Compression Struts

Based on the results of the analytical and experimental studies presented in Chapter 2
and 3, it is suggested that when designing deep members (disturbed regions) without
sufficient distributed reinforcement to ensure redistribution after initial cracking, the
maximum bearing stress should be limited to

fb O.6f(1 + 2a/3) (4.3)

where
= O.33(/ — 1) 1.0

0.33( — 1) 1.0

The ratio h/b, which represents the aspect ratio (height/width) of the compression
strut, should not be taken less than 1.0 (i.e., 3 0). The parameter a accounts for
Chapter 4. Bearing Strength of Concrete Compression Struts 36

the amount of confinement, while the parameter j3 accounts for the geometry of the
compression stress field.
The lower bearing stress limit of O.6f is appropriate if there is no confinement (i.e.
2 /A
/A 1 1) regardless of the height of the compression strut, as well as when the com
pressive strut is relatively short (i.e. h/b 1) regardless of the amount of confinement.
The upper limit of Equation 4.3, O.6f x 3 = l.8f, is chosen to correspond approximately
to the upper limit of bearing strength given in the ACT Building Code. The interaction
of confinement and geometry (aspect ratio) is chosen so as to give a reasonably simple
expression and yet correspond well with the finite element predictions and the exper
imental results. Figure 4.7 compares Equation 4.3 with the finite element prediction,
while Figure 4.8 compares predictions from Equation 4.3 with the experimental results.
As mentioned previously, concrete bearing strength is actually proportional to the
concrete tensile strength. If the concrete compressive strength is significantly greater
than 5,000 psi (34.5 MPa), a more appropriate limit for the bearing stress is

fb 0.6f + 6aç8fl (4.4)

where f in MPa units. If psi units are used, the 6cr6 in Equation 4.4 should be replaced
by 72a/3.
Chapter 4. Bearing Strength of Concrete Compression Struts 37

80
Experiment&:
60 -
• Cracking H/d=1.5
Cl) o Failure
U)
40
Cl)
20
I Predicted Cracking
• ft’=1.52MPa
Q) 0
2 3 4

-
4
s
80

H/d=2.0

40

20 Predicted Cracking
ft’=2.3OMPa
(Concrete Batch A)
01 2 3 4
80

60- o H/d=3
0

Predicted Cracking
ft’=1.52 MPa

01 2 3 4

80
8
H/d=6
60 -

40
Predicted Cracking
20 -
ft’=1.52MPa

0
2 3
Did

Figure 4.1: Comparison of experimental results and analytical predictions: influence of


confinement for various heights.
Chapter 4. Bearing Strength of Concrete Compression Struts 38

80
Experimental:
60 • Cracking
-
D/d=1.67
Cl) o Failure
Cl)

40 -

Cl)

20-
edicted Cracking
ft=1.52MPa
00 1 2 3 4 5 6 7 8

80

60 - D/d=2.0

0
6 0
40 -

20
- “%-“edied Cracking
1.52 MPa

1 2 3 4 5 6 7 8

80

D/d =3.0 8
60-
0

40-
8

- \_JedictedCracking
20 1.52 MPa

1 2 3 4 5 6 7 8
H/d

Figure 4.2: Comparison of experimental results and analytical predictions: influence of


height for various degrees of confinement.
Chapter 4. Bearing Strength of Concrete Compression Struts 39

5-
f,C
4-
.

3- .
I
II
$ S
2- $

11--,
ACt Code
(a)
01 2 3I 4 5

5
f,C
4

(b)
0
1 2 3
1
/
2
A A

Figure 4.3: Comparison of ACT Building Code bearing strength prediction with experi
mental results: (a) present investigation and (b) previous tests.
Chapter 4. Bearing Strength of Concrete Compression Struts 40

f3
-‘b
/
A

B
C
2
D

- ACI code
S
S
S
S

S
S
S
S
F
•1 S
I ——
F

A = 2,500 psi (17 PAPa)


B = 5,000 psi (34.5 PAPa)
C = 7,500 psi (51.7 PAPa)
D = 10,000 psI (69 MPa)

0
1 2 3

Figure 4.4: Comparison of ACT Building Code bearing strength with Hawkins’s suggested
equation for various concrete strengths.
Chapter 4. Bearing Strength of Concrete Compression Struts 41

II

(a) (b)

(c)

Figure 4.5: Influence of specimen height on ultimate bearing strength: (a) the force flow
in a single-punch test; (b) the force flow in a double-punch test and (c) the force flow in
a tall double-punch test.
Chapter 4. Bearing Strength of Concrete Compression Struts 42

b(exp)

b(AcI)

2-

I I I I
0
0 1 2 3 4 5 6 7 8 9 10

H/d

Figure 4.6: Influence of height on accuracy of ACT Building Code bearing strength pre
diction of double-punch tests.
Chapter 4. Bearing Strength of Concrete Compression Struts 43

f;
2 -j H/d =4

H/d=3

Hid = 1

o I I I
4
1 2 3

JA
1 A
D/d=
/
2

— —.

2 ,.-13Id4_-—
-

D/cI= 1

I I I
0
0 1 2 3 4
H/d

Figure 4.7: Comparison of suggested design equation with the analytical results.
Chapter 4. Bearing Strength of Concrete Compression Struts 44

(exp)

(pred)

• • .
.
I
: •

:
. .
:
.

.
• . .
.4
I S
I

01 2 3 4 5

1
J
4 A
A
D/d=
/
2

Figure 4.8: Comparison of predictions from suggested design equation with experimental
results.
Chapter 5

Shear Design of Deep Pile Caps

5.1 Introduction

This chapter deals with shear design of deep pile caps. The main objective is to de
velop a rational design procedure for deep pile caps based on the design recommendation
proposed in Chapter 4 for concrete compression struts.
The methods currently used in the design of reinforcement for deep pile caps are
mainly classified as two categories. Some designers assume that a cap acts as a beam
spanning between the piles and design the reinforcement on the basis of simple bending
theory [9]. Others assume that the cap acts as a truss, the column load is transmitted
to the piles through axial struts and the tension force required for equilibrium is taken
by reinforcement [3], 32]. A literature survey for previous theoretical and experimental
studies on pile caps can be found in Ref.[33].
Bending theory and truss analogy may lead to similar amounts of steel, but suggest
different arrangements in plan: the former leads to a uniform grid whereas the latter

favours bands of steel running between the piles. In addition, the two methods have
different anchorage requirements. Nominal anchorage is required beyond the centre-lines
of the piles for bending theory. Truss analogy, on the other hand, requires that full
anchorage lengths are provided beyond the centre-lines of the piles.
A sectional method is usually used for the shear design of deep pile caps. The shear
stresses on the code defined critical sections are limited to the shear strength contributed

45
Chapter 5. Shear Design of Deep Pile Caps 46

by concrete. However a previous experimental study [11] revealed that a sectional method
is not appropriate for the shear design of deep pile caps. Further investigations on the
use of sectional methods for the design of deep pile caps are carried out in this chapter.
This chapter is composed of six parts. In the first part, the ACT Code approach
for design of pile caps is summarized. Some practical design examples are presented,
which show that the ACT Code approach is inadequate for shear design of deep pile
caps. The definition of deep pile caps is also given in this part. The background of the
CR51 (Concrete Reinforcing Steel Institute) approach for shear design of deep pile caps
is then discussed. The CRST suggested allowable concrete shear stresses for critical beam
sections across the width of the pile cap at the face of the column and for two-way slab
punching shear on the periphery (faces) of the column are summarized.
Thirdly, the Canadian Code (CSA A23.3-M84) approach of using strut-and-tie mod
els to design deep members is summarized. However, there is actually no shear design
procedure for deep pile caps in the Canadian concrete code, i.e., there is no appropriate
strength criterion for compression struts confined by plain concrete. By applying the de
sign equation developed in Chapter 4 for the bearing stress limit of concrete compression
struts, two design methods are proposed for the shear design of deep pile caps.
Previous experimental work for pile caps and some important conclusions suggested
by various investigators are summarized. A comparative study is carried out in the fifth
part of this chapter. The predictions using various design procedures (ACT Code, CRSI
and proposed methods) are compared with the previous experimental results. Finally,
some conclusions are drawn at the end of this chapter.
Chapter 5. Shear Design of Deep Pile Caps 47

5.2 ACT Code Approach for Pile Cap Design

The ACT Building Code (ACT 318) [9, 34-35] uses a sectional force approach for the
design of pile caps regardless of the depth. The procedure involves three separate steps:
(1) shear design, in which the depth of the slab or pile cap is chosen so that the concrete
contribution to shear resistance is greater than the shear applied on the code defined
critical section; (2) fiexural design, in which the usual procedures for beams are used to
determine the required amount of longitudinal reinforcement at the critical section for
flexure, and; (3) a check that the bearing stresses on the column and piles do not exceed
the bearing strength.
The procedure to check bearing strength is to satisfy that bearing on concrete at
contact surface (e.g., the top of the pile cap under the column, the bottom of the pile
cap above the pile) does not exceed the code bearing strength (ACT 318-89: 15.8.1.1) [9].
This check is to ensure that the concrete does not crush at these locations.
For moment and shear calculations the pile force may be considered concentrated at
the centre of the pile. Flexural design is straight forward. The critical moment is equal
to the pile reaction times distance from pile centre to face of a concrete column (Clause
15.4.2). For example, the moment at section m — m in Figure 5.1 is equal to the product
of the reactions from three left side piles and the distance s. The required amount of
longitudinal reinforcement is provided to resist this moment, similar to the procedure
used for the fiexural design of slabs or beams.

5.2.1 Code Procedure for Shear Design

Shear design involves checking one-way beam shear and two-way punching shear (see
Figure 5.1). For one way beam shear, the critical section is measured at a distance d from
the face of supported member (column) (Clauses 15.5.2 and 11.12.1.1 of ACI 318-89). For
Chapter 5. Shear Design of Deep Pile Caps 48

two way punching shear, the critical section is taken at a distance d/2 out from perimeter
of the column, or the pile (Clause 11.12.1.2), where d is effective height measured from
extreme compression fibre to centroid of longitudinal tension reinforcement. Any pile
whose centre is located d/2 or more outside the critical section shall be considered as
producing shear on that section, where d is diameter of pile. In Figure 5.1 the shear on
the section n-n is introduced by the reactions from three right side piles. All eight side
piles produce shear on the critical section for two-way shear at d/2 from the perimeter
of the column. Reaction from any pile whose centre is located d/2 or more inside the
critical section shall be considered as producing no shear on that section (Clause 15.5.3).
Figure 5.2 shows a two pile cap, for which the pile produces no shear on the section
n-n. For intermediate positions of pile centre, the pile reaction is assumed as distributed
linearly across the pile diameter d in the direction shear is accumulated (Clause 15.5.3.3).
For pile caps where the pile reactions are located either under the column or outside
the critical sections (d from the face of the column for one-way shear or d/2 for two-way
shear calculation), the code procedure for shear design is straight forward. It is proposed
herein that such a pile cap, as shown in Figure 5.1, can be defined as a slender pile cap.
However, a great number of pile caps used in practice are not slender pile caps. These pile
caps are herein called ‘deep’ pile caps. The effective height d of deep pile caps is equal
to or greater than the distance from the centre line of the closet pile to the face of the
supported column. The ACT Code approach becomes less transparent and inadequate
for the shear design of deep pile caps. This inadequacy is illustrated in the following
examples.
Chapter 5. Shear Design of Deep Pile Caps 49

5.2.2 Inadequacies of ACI Code Procedure for Shear Design of Deep Pile
Caps

Figure 5.2 shows a deep two-pile cap, for which the piles are located within the code
defined critical section from the face of the column (i.e., within the distance d from the
column face). There are no code provisions specifically for the shear design of such a
pile cap. Hence ACT 318-89 Commentary Section 15.5.3 says: “when piles are located
inside the critical sections d or d/2 from face of column, analysis for shear in deep flexural
members in accordance with Section 11.8 needs to be considered.” In Clause 11.8 some
guidance for deep beams is given. A critical section is established at 0.5 a from face of
support, where a is shear span and is defined as a distance from a concentrated load to
face of support. It becomes hard to define what is the support (the pile or the column?).
Based on the guidance for deep beams, it could be considered that piles are supports and
column load is concentrated load. In Figure 5.2, the critical section j-j is defined by this
interpretation. The code equations 11-29 or 11-30 are applicable and thus shear strength
on that section can be calculated. However, if the cap depth is increased, this defined
critical section could eventually cut through the column so that again no code provisions
can be applied.
Figure 5.3 shows a deep three-pile cap. If the piles are still considered as supports,
it is difficult to determine the shear span a (i.e., where is the concentrated load ?). It
should also be noted that there are no code provisions whatsoever applicable to perform
punching shear calculations when the three piles are located within the square critical
section shown in the figure.
Figure 5.4 shows another example that demonstrates the inadequacy of the ACT Code
approach. Two four-pile caps are identical except that one has a square column and the
other has a circular colunm. The two columns have the same areas. For the pile cap
Chapter 5. Shear Design of Deep Pile Caps 50

that has the square column and four circular piles, as shown in Figure 5.4(a), the square
critical section d/2 away from the perimeter of the column just has its corners at the
centres of piles. The ACT Code assumption that the pile reaction is distributed linearly
across the depth, d (pile diameter) (Clause 15.5.3.3), can be utilized for punching shear
calculation. If the code is followed literally, half of all pile reactions will introduce shear
on the critical section. On the other hand, if the square column is replaced by a circular
column with the same area and correspondingly the critical section is circular rather than
square, [see Figure 5.4(b)], the pile reactions are now totally located outside the critical
section so that the critical section is subjected to the shear from the full value of pile
reactions. Consequently, the prediction for load capacity of the pile cap with the square
column is as much as almost two times that of the pile cap with the circle column, based
on the two way punching shear calculation. The sensitivity of the predicted capacity to
the shape of the colunm is a demonstration that a sectional approach is not appropriate
for deep pile caps.

5.2.3 Comparison of Different ACI Code Editions

In the above section the 1989 ACT Code [9] procedures for shear design of pile caps have
been summarized, and it has also been pointed out that these procedures are transparent
and directly applicable to slender pile caps. From the 1977 edition to the 1989 edition,
the ACT Code have made some changes in shear design clauses which do not affect the
shear design of slender pile caps, but do affect the shear design of deep pile caps. These
changes are summarized below.
In ACT 318-77 [34], Clause 11.1.3.1 stated that for nonprestressed members, sections
located less than a distance d from face of support may be designed for the same shear
as that computed at a distance d. In the accompanying commentary to ACT 318-77, a
statement is made that if the shear at sections between the support and a distance d
Chapter 5. Shear Design of Deep Pile Caps 51

differs radically from the shear at distance ci, the shear at the face of the support should
be used.
In ACT 318-83 [35], the contents of the cormnentary have been incorporated into the
formal body of the code, which means that the requirement that the shear capacity at
the face of the support should be evaluated is imposed when there is a concentrated load
between the face of support and the code defined critical section.
From ACT 318-83 to ACT 318-89 no additional changes have been made to Clause
11.1.3. In comparison with the previous ACT Code editions, more explanations about
Clause 11.1.3 have been added to the Commentary (ACT 318R-89). Something new in
the ACT 318R-89, which is relevant to this study, is that the shear at the face of the
column should be investigated for footings supported on piles when the pile reaction is
located within a distance d from the face of the column.
The ACT Building Code procedures for two-way shear have not been modified recently.
The critical section remains at d/2 from the perimeter of the column regardless whether
there is a concentrated load applied within the critical section.

5.3 CRSI Approach for Shear Design of Deep Pile Caps

As there are actually no specific procedures in the ACT Code for the shear design of deep
pile caps [36], formulas are presented in the CR51 Handbook [36, 37] for allowable shear
on concrete as a function of the ratio w/d, where w is the horizontal distance from face of
column to centre of pile reaction and ci is effective depth of pile cap. Based on recognition
of considerable reserve shear strength observed in deep beam tests [38], allowable shear
strength on concrete is assumed to increase rapidly from ‘diagonal tension’ to ‘pure shear’
as the distance w varies from d to 0 for one-way shear and from d/2 to 0 for two-way
punching shear. In addition, the critical sections for shear calculation of deep pile caps
Chapter 5. Shear Design of Deep Pile Caps 52

are specified.
For one-way shear of deep pile caps, the critical section is at the face of the column.
The formula for shear strength calculation, which is a modification of the ACT Equation
11-30, is

V = (--)(3.5 —
2.5)(1.9iTh+ 2500p )bd 10Jbd (5.1)

where (w/d) < 1.0, 1.0 > M/Vd> 0 and f is in psi unit. In the CRSI handbook [37]
the formula is further reduced for the case of p 0.002 and f’ = 3000psi to

V = (-)(3.5 —

2.5#)(1.9.Th + O.l/)bd 1O/bd (5.2)

which gives when the product of the last bracketed expressions = 2’J and
(d/w) 1. Figure 5.5(a) summarizes the one-way shear calculations for deep pile caps.
The CRSI critical section for two-way shear in deep pile caps is defined at the perime
ter of the column face. The formula for the allowable shear stress at the critical section,
which is a modification of the ACT Equation 11-36, is

V. = (-)(1 + )(4/) <32./jb


d
0 (5.3)

where for a square column of dimension c, b


0 equals 4 x c. When d/2w = 1, the CRST
formula leads to the same expression as the ACT approach where the critical section is
at d/2 from the column perimeter (b
0 equals 4 x (c + d)). Figure 5.5(b) summarizes the
CRSI two-way shear calculations for deep pile caps.
The flow chart summarizing ACT Code and CRSI approaches is presented in Figures
5.6.
Chapter 5. Shear Design of Deep Pile Caps 53

5.4 Strut-and-Tie Model Approach

An alternative approach which can overcome the drawbacks of ACI approach for the
shear design of deep pile caps is strut-and-tie models which consider the complete flow of
forces within a pile cap rather than the forces at any one particular section. The internal
force flow indicates that the vertical column load is transmitted to the piles by inclined
compression struts and in order to prevent the piles from spreading apart, tension ties
(reinforcement) must be provided. Assuming reinforcing steel bars are properly anchored
in the nodal zone, it is believed that a “shear failure” of a deep pile cap will occur when
a concrete compression strut fails prior to yielding of tension ties. It should be noted
that truss models previously proposed for the design of pile caps, as mentioned in the
introduction of this chapter, have been used only for “flexural design”.

5.4.1 CSA Approach for Deep Beams

Strut-and-tie models have been adopted in the Canadian Concrete Code (CSA A23.3-
M84 [10]) for the design of deep beams [39]. The internal flow of forces in a simply
supported deep beam can be approximated by the truss model shown in Figure 5.7.
The truss is composed of several components. The zones of unidirectional compressive
stress in the concrete are modelled as compression struts, while tension ties are used to
model the principal reinforcement. The regions of concrete subjected to multi-directional
stresses, where the struts and the ties meet (the joints of the truss) are represented by
‘nodal zones’.
The Canadian Code [10] requires that the concrete compressive stresses in the nodal
zones not exceed the following limits (Section 11.4.7.5):

(a) 0.85f in nodal zones bounded by compressive struts and bearing areas;
Chapter 5. Shear Design of Deep Pile Caps 54

(b) O.75f in nodal zones anchoring only one tension tie; and

(c) 0.60f in nodal zones anchoring tension ties in more than one direction,

The tension tie reinforcement is distributed and anchored over an effective area of
concrete at least equal to the tensile tie force divided by the stress limits given above
(Sections 11.4.7.4 and 11.4.7.6).
The key part of this approach is the establishment of stress limit for the compression
struts based on the work by Vecchio and Collins [40]. The code requires that the concrete
compressive stress f2 in the struts shall not exceed f2ma (Section 11.4.7.3), i.e.

f2 f2rnax (5.4)
= 0.8 +170c
1

where f2ma shall not exceed f unless the concrete is triaxial confined and e is determined
by considering the strain conditions of the concrete and the reinforcement in the vicinity
of the strut.
This approach is rational for deep beams, however extending this approach to deep pile
caps seems questionable. In deep pile caps, compression struts are usually unreinforced
and confined by surrounding plain concrete. Thus the Canadian Code has no provisions
specifically for the shear design of deep pile caps. In the Canadian Concrete Design
Handbook [41], Suter and Fenton have ignored shear calculations for deep pile caps and
incorporated a more conservative flexural model to predict deep pile cap behaviour.

5.4.2 Proposed Design Method (1)

The stress limit for compression struts in the Canadian Code was established by testing
reinforced concrete panels [40], which were subjected to in-plane stresses. The stress
limit is appropriate for planar reinforced concrete members and is not applicable to three
dimensional pile caps where compression struts are confined by plain concrete.
Chapter 5. Shear Design of Deep Pile Caps 55

In Chapter 4 the design equations for the bearing stress limit of compression struts
confined by surrounding plain concrete were proposed, based on the analytical and ex
perimental investigation. For convenience, the equations are again written here:

= 0.6f(1 + 2afl) (5.5)

where

a = 0.33(J4- 1 1.0

= O.33( — 1) 1.0

If the concrete compressive strength is significantly greater than 5,000 psi (34.5 MPa),
the limit for the bearing stress is

fb 0.6f + 6a/3/ (5.6)

where f is in MPa units. If psi units are used, the 6cq3 should be replaced by 72a.
These equations are directly applicable to the idealized compression struts as studied
in Chapter 2 to Chapter 4. See Figure 5.8(a). However, the geometry of compression
struts in deep pile caps may be different from that of the idealized struts. Some factors
which may affect the determination of geometrical ratios of D/d and h/b have to be taken
into consideration.
One situation that could occur in design of deep pile caps is shown in Figure 5.8(b)
and (c). In addition to different loading areas on two ends, the compression strut does
not have a constant cross-section along its length. Hence new ratios of both h/b and D/b
are needed in using the design equations. Herein a weighted average method is proposed
to calculate the new geometrical ratios. These are
Chapter 5. Shear Design of Deep Pile Caps 56

D — /b + 2
1
(D
2
) /b
(D
)
5 7
b — D
+
1
/
2 b
D
h — (h/b + (h/b
2
)
1 )
2
58
b — +h/b
1
h/b
2
where b
, D
1 , b
1 2 and D
2 are loading geometry of the compression struts. See Figure 5.8.
The allowable bearing stress limit of the compression strut in Figure 5.8(a) is greater
than that in Figure 5.8(b). And again the allowable bearing stress limit of the one in
Figure 5.8(b) is less than that in Figure 5.8(c).
It should be noted that the predicted bearing stress fb for each compression strut is
on the end with smaller loaded area. As the force in a strut is constant, the bearing
stress on the other end can be determined from equilibrium.
By incorporating the developed bearing stress limit for compression struts confined by
plain concrete into a simple three-dimensional strut-and-tie model, one design procedure
is proposed.
As strut-and-tie models suggested in the Canadian Concrete Code (CSA A23.3-M84)
[39] for deep beam design, the first design procedure proposed here for deep pile caps does
not split the procedure into “flexural design” and “shear design,” which are traditionally
used for the design of slender members. The procedure involves several steps to develop an
equilibrium force system. The initial pile cap dimension can be chosen based on previous
design experience or with the help of published design aids. The effective depth d of a
pile cap can be determined from the concrete cover requirement and then the location
of reinforcement can also be determined. The strut-and-tie model is drawn and the
geometric parameters of compression struts are then calculated. The bearing stress limit

on the nodal zones and the corresponding force in the compression struts can be calculated
by using Equations 5.5 or 5.6, 5.7 and 5.8. The horizontal components of the inclined
compression strut forces are resisted by providing properly anchored reinforcement. The
Chapter 5. Shear Design of Deep Pile Caps 57

sum of vertical components of the inclined compression strut forces is the calculated load
capacity of the pile cap. Prediction examples can be found in Appendix C.

5.4.3 Proposed Design Method (2)

The second design method is a simplified procedure of the first design method and is
presented in a more traditional way with a separated “flexural design” and “shear design”.
In the second design method, the calculation of the bearing stress limit is based on
some simple parameters such as the effective depth of the pile cap and the dimensions
of the column and the pile rather than more complicated geometric parameters of the
compression strut. As a result, this simplified procedure is more easy to apply in the
design of deep pile caps.
The traditional truss analogy principle is used for “flexural design”. The fiexural
capacity depends strongly on the inclination of the compression strut, which is defined
by the location of the nodal zones. The truss model used in this procedure is that the
lower nodal zones are located at the center of the piles at the level of the longitudinal
reinforcement, while the upper nodal zone is assumed to be at the top surface of the pile
cap at the column quarter point.
The “shear design” of a deep pile cap involves applying Equation 5.5 or 5.6 to limit
the concrete stresses in compression struts and nodal zones to ensure the tension tie
reinforcement yields prior to any significant diagonal cracking in the plain concrete com
pression struts. The ratio A
1
/
2 A is identical to that used in the ACT Code to calculate
bearing strength. To calculate the maximum bearing stress for the nodal zone below a
column, where two or more compression struts meet, the aspect ratio of the compression
strut can be approximated as

(5.9)
Chapter 5. Shear Design of Deep Pile Caps 58

where d is the effective depth of the pile cap and c is the dimension of a square column.
For a round column, the diameter may be used in place of c. To calculate the maximum
bearing stress for a nodal zone above a pile, where only one compression strut is anchored,
the aspect ratio of the compression strut can be approximated as

(5.10)

where 4, is the diameter of a round pile.


A general shear design procedure for deep pile caps can be accomplished as follows.
First, the initial pile cap depth is chosen using the ACT Code one-way and two-way shear
design procedures. In the case of one-way shear, the critical section should be taken at
d from the column face, and any pile force within the critical section should be ignored
(i.e., the ACT procedure prior to 1983). Secondly, the nodal zone bearing stress should
be checked using above described procedure. If necessary, the pile cap depth may be
increased (/3 increased), or the pile cap dimensions may be increased in order to increase
the confinement of the nodal zones (increase a), or else the bearing stresses may need to
be reduced by increasing the column or pile dimensions. Prediction examples showing
this method can be found in Appendix D.

5.5 Summary of Previous Pile Cap Tests

In order to evaluate the various pile cap design procedures quantitatively, these pro
cedures will be used to predict previously tested pile caps. In this section previous
experimental work on pile caps is summarized.
Hobbs and Stein (1957) [42] tested 70 one-third scale models of two-pile caps to verify
their design method, which is based on analytically determined bending stress distribu
tions on vertical planes through two-pile caps. The specimens had various amounts of
Chapter 5. Shear Design of Deep Pile Caps 59

either straight or curved reinforcing bars, which were anchored by a number of different
methods. During the test, the first cracks consistently appeared at or close to the pile cap
mid-span. Later, diagonal cracks, which originated at the top of the piles, propagated to
the loading plate and were usually followed directly by failure.
Deutsch and Walker (1.963) [43] tested four two-pile caps. The pile caps were designed
with the same centre to centre pile spacing 42 inches (1067 mm), the same width 15 inches
(381 mm), the same stub pile dimensions 10 x 10 in. (254 x 254 mm) and the same stub
column dimensions 6.5 x 6.5 in. (165 x 165mm). The depth of the pile caps and amount of
reinforcing steel were varied in an attempt to investigate the structural action of pile caps
and to compare different design methods (sectional force methods and the truss analogy
with compression struts inclined at 35 and 45 degrees). Owing to the limitation of the
loading capacity of the testing apparatus, only two pile caps (No.3 and No.4) were tested
to failure, which were designed on the truss analogy with compression struts inclined at
45 and 35 degrees respectively. None of the pile caps were actually thought to have failed
in shear. All pile caps behaved similarly with one main crack forming at the vertical
centre line, extending to within one inch from the top of the pile cap at failure.
Blévot and Frémy (1967) [32] tested two series of pile caps. First series specimens
included 51 four-pile caps, 37 three-pile caps and 6 two-pile caps and were all of about
half scale size. Second series specimens were of full scale size and had 8 four-pile caps, 8
three-pile caps and 6 two-pile caps. The main objectives was to determine the influence
of different reinforcing steel layouts and to verify their proposed truss models, which were
used to design reinforcement. The height of the pile caps and the layouts of reinforcing
steel were varied in the study. Figure 5.9 shows the five different reinforcing steel layouts
investigated in the four-pile caps of first series specimens. These layouts can be named
bunched square, bunched diagonal, bunched hybrid, and grid. The three-pile caps of the
first series specimens had similar layouts to those in the four-pile caps. In the second
Chapter 5. Shear Design of Deep Pile Caps 60

series specimens, three-pile caps and four-pile caps had two kinds of layouts, bunched
hybrid, and bunched square or triangle plus grid. Layouts in the two-pile caps of both
series specimens are similar to that in simply supported deep beams.
Clarke (1973) [44] tested 15 four-pile caps, with the objective to compare the strength
of caps with values predicted by various design methods. Moreover the effect of different
steel layouts and different anchorage lengths upon the behaviour of pile caps at service
loads and ultimate loads were also investigated. The specimens included two types of
pile caps, namely type A and type B. A pile diameter of 200 mm, a column dimension
of 200 x 200 mm and the total depth of 450 mm were used throughout, giving plan
dimensions of 950 mm square for the type A caps and 750 mm square for the type
B. Three different steel layouts (grid, bunched square and bunched diagonal) and four
different anchorage lengths (nil, nominal, full and full-plus-bob) were considered. See
Figure 5.10. Most pile caps failed in shear after reinforcement yielded. Four pile caps
were thought to have failed in flexure.
Clarke and Taylor (1974) [45] also tested a number of eight-pile caps at 1:4, 1:15 and
1:38 scale to investigate the influence of pile stiffness upon distribution of the column
load in pile caps. The uncracked pile caps were found to distribute pile loads as predicted
by an elastic solution. When the concrete cracked, load carried by the outer piles was
considerably decreased.
In order to confirm their proposed shear design approach for deep pile caps, Sabnis
and Gogate (1980, 1984) [46, 47] tested nine four-pile caps, all at a one-fifth scale. The
first three specimens, one of which was of plain concrete, served to finalize the test set-up.
The remaining six specimens were used to study the effect of the amount of uniformly
placed reinforcement upon the shear capacity of deep pile caps. All the tested pile caps
had the same dimensions: 3 inch (76 mm) diameter piles and columns, 13 inch (330 mm)
square caps and 6 inch (152 mm) total depth of caps. The reinforcement ratio was varied
Chapter 5. Shear Design of Deep Pile Caps 61

between 0.0014 and 0.0079 for the main six specimens. In order to prevent anchorage
failure, the reinforcement was hooked and further extended vertically the full depth of
the pile cap. Punching shear was thought to be the predominant failure mechanism for
all the specimens tested.
Adebar et al. (1990) [11] tested six full scale pile caps in an attempt to investigate
the applicability of strut-and-tie models (truss models) to the design of pile caps and to
evaluate the validity of the current ACT design procedures for pile caps. Of six tested
pile caps, four caps were of diamond shapes, one cap of rectangular shape and one cap
of cruciform shape. The pile caps all had an overall depth of 600 mm and were loaded
through 300 mm square cast-in-place reinforced concrete columns. They were supported
by 200 mm diameter precast reinforced concrete piles. The layouts of reinforcing steel
used were bunched bars, grid, and bunched plus grid. Full straight anchorage lengths
were provided for reinforcing bars passing over the piles. Of six pile caps, four are believed
to have failed in shear while two are thought to have failed in flexure.
From above mentioned experimental investigations for pile caps, some important con
clusions obtained by the investigators are summarized in the following:

1. Truss models are more appropriate for deep pile caps than simple bending theory.
Blévot and Frémy [32] confirmed their truss models by carrying out a comprehensive
series of tests. Best results were obtained with the imaginary struts inclined at
between 45° and 55° to the horizontal. Clarke [44] concluded that the truss analogy
was an adequate method of analyzing a four-pile cap in order to ascertain its flexural
capacity and to determine the required amount of tensile reinforcement. Adebar et
al. [11] concluded that the ACT Building Code (sectional method) failed to capture
the trend of the experimental results.

2. Layouts of reinforcing steel have a great effect on the behaviour of deep pile caps
Chapter 5. Shear Design of Deep Pile Caps 62

under service loading and on the ultimate capacity of deep pile caps. The most
desirable behaviour was obtained by the layouts of the bunched hybrid or bunched
steel plus a relative light grid of steel. Blévot and Frémy [32] observed that the
bunched steel gave approximately 20% higher strength than the same weight of steel
spread out in a grid pattern. Clarke [44] concluded, from his four-pile cap tests,
that bunching the steel in the form of a square increases the load by about 14%,
while bunching it along the diagonals only increases the strength by a negligible
amount. It was reported by both Blévot and Clarke that the crack control was very
much improved by using bunched hybrid or bunched hybrid plus grid, instead of
using bunched diagonal only.

3. Only nominal anchorage of reinforcing steel is required. Deutsch and Walker [43]
showed analytically and confirmed by their tests that only nominal anchorage was
required past the edges of the piles. Blévot and Frémy [32] recommended the use of
standard hooks as regular requirement for anchorage based on their experimental
results. Clarke [44] concluded from his four pile cap tests that caps with bunched
square steel with nil anchorage was the most efficient if the efficiency is defined as
the load carried by the cap divided by the total weight of steel used. The reason
that the anchorage requirement can be reduced was, suggested by Clarke, that
reinforcing bars bunched over piles are subjected to high lateral bearing stresses
which help to lock the bars into place. It was also shown in Clarke’s tests that the
full anchorage resulted in a 30% increase in load capacities. Clarke thought that
this substantial increase was mostly due to the vertical portion of the reinforcement
providing reinforcing across shear cracks. These vertical portions of the reinforcing
bars acted as stirrups. It should be noted that the reinforcing bars in the specimens
tested by Sabnis and Gogate [47] were also full anchorage with the bars extended
Chapter 5. Shear Design of Deep Pile Caps 63

vertically the full depth of the pile cap.

5.6 Comparative Study of Different Design Procedures

5.6.1 Comparison of Design Methods

To compare the one-way shear design procedures, Figure 5.11 summarizes the relationship
between the maximum column load and the width b and depth d of a two-pile cap. When
the width of the pile cap is the same as the column width (b = c), the pile cap is essentially
a deep beam. See Figure 5.11(b). When the width of the pile cap is increased, larger
shear forces can be resisted by the increased concrete area at the critical section, and the
limiting column load due to bearing strength is increased as a result of confinement, see
Figure 5.11(c) and (d).
Three different ACT Code predictions for one-way shear are given in Figure 5.11. The
least conservative prediction, entitled “ACI ‘77,” is what designers of pile caps would have
used prior to the 1983 edition of the ACT Building Code (any pile within d of the column
face is assumed to produce no shear on the critical section). The “ACT ‘83” prediction, in
which the critical section is at the face of the column, gives very conservative predictions.
The predicted column load based on one-way shear calculation at the critical section half
way (a/2) between the face of the column and the centre of the closest pile, “ACT [11.8],”
gives an intermediate result. The CRST method, in which the critical section is also at
the face of the column, is much less conservative than “ACT ‘83” due to an enhanced
concrete contribution, but is more conservative than “ACT ‘77.”
Note that all methods predict that as the pile cap becomes very deep, the maximum
column load is limited by bearing strength. When the pile cap is twice as wide as the
column (b 2c) the ACT Code predicts that the confinement is sufficient so that the
bearing strength has reached the upper limit of 2 x O.85f.
Chapter 5. Shear Design of Deep Pile Caps 64

Figure 5.12 compares the influence of pile cap depth on two-way shear strength predic
tions for a typical four-pile cap. Although the CRSI Handbook method gives a concrete
contribution which is significantly larger for deep pile caps, the maximum column load
is always smaller than the ACT method. This is because in the ACT method the critical
section is at d/2 from the column face and any pile which intercepts the critical section
is assumed to transmit part of the load directly to the column. For example, if a pile
is centred on the critical section, only half of the pile reaction must be resisted by the
critical section.

5.6.2 Comparison of Prediction Results

Table 5.1 summarizes the 48 specimens which were chosen for the comparative study.
These include two pile caps tested by Deutsch and Walker [43], eighteen pile caps tested
by Blévot and Frémy [32] in their second series, fourteen pile caps tested by Clarke [44],
eight pile caps tested by Sabnis and Gogate [47] as well as six pile caps tested by Adebar
et at. [11]. Specimens not considered in the comparative study include the small wide
beam models tested by Hobbs and Stein, the small-scale specimens (first series) tested
by Blévot and Frémy, and the one specimen tested by Clarke and two specimens tested
by Deutsch and Walker which did not fail.
Table 5.2 and Table 5.3 give the ACT Code and CRST predictions. The predicted
flexural strengths and bearing strengths are summarized in Table 5.2, while the predicted
shear capacities are summarized in Table 5.3.
In the case of one-way shear, three different predictions are given from the ACT
Building Code: the 1977 edition of the ACT Building Code (critical section at d from
the column face); the 1.983 ACT Building Code (critical section at the column face); and
the special provisions for deep flexural members (Clause 11.8, critical section at half way
between the column face and the closest pile centre). Tn the case of two-way shear, ACT
Chapter 5. Shear Design of Deep Pile Caps 65

Code procedures involve critical section at d/2 from the column face and the pile face.
Details of the predictions can be found in Appendix B.
Table 5.4 presents the ratio of measured pile cap capacity to predicted capacity for
the three ACT Code predictions as well as the CRSI prediction. As mentioned before,
actually there are no shear design procedures for deep pile caps in ACT ‘77, therefore most
pile caps are predicted to fail in flexure. If Clause 11.8 is considered, the load capacity
of some pile caps is limited by shear (especially specimens tested by Blévot and Frémy).
Tf the 1983 ACT Building Code is applied, it can be seen that the load capacity of almost
all pile caps is controlled by shear due to the lower predicted shear capacities.
It is very interesting to examine the CRST predictions. The original objective of the
CRSI equations is to reflect the considerable reserve strength observed in the tests of deep
beams. Thus the CRSI equations generally give higher shear capacities. This results in
the change of predicted failure modes for most of the pile caps to flexural failure as in
the ACT ‘77 predictions. The comparisons are illustrated graphically in Figure 5.13 to
5.16.
Table 5.5 summarizes the predictions from the proposed design methods and compares
the predictions with the experimental results.
The comparisons of measured pile cap capacities and the predictions from the pro
posed design method (1) are shown in Figure 5.17 and 5.18. The predictions for specimens
listed in Table 5.1 are shown in Figure 5.17, while the predictions for the pile caps with
bunched reinforcement are shown in Figure 5.18. The comparison of measured test results
with the predictions from the proposed design method (2) is shown in Figure 5.19.
Chapter 5. Shear Design of Deep Pile Caps 66

5.7 Conclusions

Some conclusions can be arrived at from the study in this chapter. The ACT Code
design procedures (especially shear design) are not adequate for deep pile caps. The ACT

Code 1977 edition gives scattered and not conservative predictions. The ACT Code 1983
edition, in contrast with 1977 edition, gives overly conservative predictions, obviously
due to the equation used for shear calculation on the critical section of the face of the
column which underestimates the shear capacity of deep pile caps. Though this problem
has been realized in the development of CRST shear design method, the problem is not
totally solved. In addition, the ACT Code design procedures or CRSI procedures are
sometimes very complex and tedious, involving many calculations in order to meet the
code requirement. For example, predicting the three-pile cap specimens tested by Blévot
is difficult.
The ACT Code flexural strength predictions are unconservative for deep pile caps.
These fiexural strength procedures are meant for lightly reinforced beams which are able
to undergo extensive fiexural deformations (increased curvatures) after the reinforcement
yields. As the curvature increases, the fiexural compression stresses concentrate near the
compression face of the member. Deep pile caps are too brittle to undergo such deforma
tions, therefore, assuming the fiexural compression is concentrated near the compression
face is inappropriate. Assuming the flexural compression is uniform across the entire
pile cap, which strain measurements have shown to be incorrect [11], leads to a further
overprediction of the flexural capacity.
In Figure 5.20, the bearing capacities on column zones for previously tested pile caps
are illustrated in terms of the amount of confinement at column zones. The dashed line
shown in the figure demonstrates a similar trend as explored analytically and experi
mentally on the double punch loaded concrete cylinders. This indirectly further confirms
Chapter 5. Shear Design of Deep Pile Caps 67

that the transverse splitting phenomenon is predominant for the shear failures of deep
pile caps.
Both proposed design methods herein capture the physical behaviour of deep pile
caps, and produce conservative and reasonably accurate predictions. The proposed de
sign method (1) is a direct extension of the two-dimensional strut-and-tie model of the
Canadian concrete code. The proposed design method (2) is a simplified method which
treats “fiexural design” and “shear design” separately. This is the concept traditionally
accepted by current engineering practice.
Chapter 5. Shear Design of Deep Pile Caps 68

Table 5.1: Summary of pile cap test results.

Number d Pile Column f Failure


Specimen of Size Size Reinf. Load
Piles (mm) (mm) (mm) (MPa) Layouts (kN)
Blévot & Frémy
2N1 2 495 350 sq. 350 sq. 23.1 bunched 2059
2Nlb 2 495 350 sq. 350 sq. 43.2 bunched 3187
2N2 2 703 350 sq. 350 sq. 27.3 bunched 2942
2N2b 2 698 350 sq. 350 sq. 44.6 bunched 5100
2N3 2 894 350 sq. 350 sq. 32.1 bunched 4413
2N3b 2 892 350 sq. 350 sq. 46.1 bunched 5884
3N1 3 447 350 sq. 450 sq. 44.7 bunched 4119
3Nl 3 486 350 sq. 450 sq. 44.5 bunched 4904
3N3 3 702 350 sq. 450 sq. 45.4 bunched 6080
3N3 3 736 350 sq. 450 sq. 40.1 bunched 6669
4N1 4 674 350 sq. 500 sq. 36.5 b.& g. 6865
4N1 4 681 350 sq. 500 sq. 40.0 b.& g. 6571
4N2 4 660 350 sq. 500 sq. 36.4 bunched 6453
4N2b 4 670 350 sq. 500 sq. 33.5 bunched 7247
4N3 4 925 350 sq. 500 sq. 33.5 b.& g. 6375
4N3b 4 931 350 sq. 500 sq. 48.3 b.& g. 8826
4N4 4 920 350 sq. 500 sq. 34.7 bunched 7385
4N4b 4 926 350 sq. 500 sq. 41.5 bunched 8581
Deutsch & Walker
No.3 2 533 254 sq. 165 sq. 23.8 bunched 596
No.4 2 373 254 sq. 165 sq. 23.6 bunched 289
Chapter 5. Shear Design of Deep Pile Caps 69

Table 5.1: (con’t) Summary of pile cap test results.

Number d Pile Column f’ Failure


Specimen of Size Size Reinf. Load
Piles (mm) (mm) (mm) (MPa) Layouts (kN)
Clarke
Al 4 400 200 rd. 200 sq. 20.9 grid 1110
A2 4 400 200 rd. 200 sq. 27.5 bunched 1420
A3 4 400 200 rd. 200 sq. 31.1 bunched 1340
A4 4 400 200 rd. 200 sq. 20.9 grid 1230
A5 4 400 200 rd. 200 sq. 26.9 bunched 1400
A6 4 400 200 rd. 200 sq. 26.0 bunched 1230
A7 4 400 200 rd. 200 sq. 24.2 grid 1640
A8 4 400 200 rd. 200 sq. 27.5 bunched 1510
A9 4 400 200 rd. 200 sq. 26.8 grid 1450
AlO 4 400 200 rd. 200 sq. 18.2 grid 1520
All 4 400 200 rd. 200 sq. 17.4 grid 1640
A12 4 400 200 rd. 200 sq. 25.3 grid 1640
Bl 4 400 200 rd. 200 sq. 26.9 grid 2080
B3 4 400 200 rd. 200 sq. 36.3 grid 1770
Sabnis & Gogate
SS1 4 111 76 rd. 76 rd. 31.3 grid 250
SS2 4 112 76 rd. 76 rd. 31.3 grid 245
SS3 4 111 76 rd. 76 rd. 31.3 grid 248
SS4 4 112 76 rd. 76 rd. 31.3 grid 226
SS5 4 109 76 rd. 76 rd. 41.0 grid 264
SS6 4 109 76 rd. 76 rd. 41.0 grid 280
5G2 4 117 76 rd. 76 rd. 17.9 grid 173
SG3 4 117 76 rd. 76 rd. 17.9 grid 177
Adebar, Kuchma and Collins
A 4 445 200 rd. 300 sq. 24.8 grid 1781
B 4 397 200 rd. 300 sq. 24.8 bunched 2189
C 6 395 200 rd. 300 sq. 27.1 bunched 2892
D 4 390 200 rd. 300 sq. 30.3 bunched 3222
E 4 410 200 rd. 300 sq. 41.1 bk g. 4709
F 4 390 200 rd. 300 sq. 30.3 bunched 3026
Chapter 5. Shear Design of Deep Pile Caps 70

Table 5.2: Summary of ACT Building Code and CRSI Handbook predictions (flexure and
bearing).

Specimen Flexural Bearing Capacity


Name Capacity Column Pile
(kN) (kN) (kN)
2N1 2197 2746 5492
2Nlb 3756 5135 10270
2N2 3429 3244 6488
2N2b 5553 5310 10620
2N3 5413 3816 7632
2N3b 7259 5485 10970
3N1 3825 15378 23860
3Nlb 5290 15311 23757
3N3 6014 15631 24255
3N3b 7982 13808 21424
4N1 7969 15525 26083
4Nlb 8171 17005 28570
4N2 7812 15464 25977
4N2b 8546 14234 23913
4N3 8283 14234 23913
4N3b 10788 20549 34521
4N4 9864 14734 24753
4N4b 10866 17631 29621
No.3 512 1100 3907
No.4 270 1093 3881
Chapter 5. Shear Design of Deep Pile Caps 71

Table 5.2: (con’t) Summary of ACT Building Code and CRSI Handbook predictions
(flexure and bearing).

Specimen Flexural Bearing Capacity


Name Capacity Column Pile
(kN) (kN) (kN)
Al 1258 1421 3908
A2 1266 1870 5140
A3 1256 2115 5812
A4 1258 1421 3908
A5 1265 1829 5028
A6 1252 1768 4860
A7 1262 1646 4524
A8 1266 1870 5140
A9 1264 1822 5008
AlO 1252 1238 3404
All 1252 1183 3252
A12 1262 1720 4728
Bi 2022 1829 5028
B3 1528 2468 6784
SS1 133 242 808
SS2 116 242 808
553 194 242 808
SS4 157 242 808
SS5 317 318 1060
SS6 455 318 1060
SG2 302 139 463
SG3 628 139 463
A 2256 3794 5296
B 2790 3794 5296
C 4008 4146 8682
D 5646 4636 6472
E 7428 6288 8780
F 5324 3091 6472
Chapter 5. Shear Design of Deep Pile Caps 72

Table 5.3: Summary of ACT Building Code and CRSI Handbook predictions (shear).

Specimen One-Way Shear (kN) Two-Way Shear (kN)


Name ACT CRSI Column Pile
1977 1983 [11.8] ACT CRSI
2N1 1053a 316 947 775 c c c
2Nlb 1438a 432 1296 898 c c c
2N2 b 488 1463 2438 c c c
2N2b b 616 1846 2634 c c c
2N3 b 673 2018 3364 c c c
2N3b b 804 2414 4023 c c c
3N1 2130a 1583a 4490 2130 3722a 6594 3789
3Nlb 2707a 1715a 4737 2639 4397a 8062 4157
3N3 b 2504a 7496 9321 b 21249 b
3N3b b 2468a 7385 9871 b 22297 b
4Nl b 2436 7308 12004 11831a d b
4Nlb b 2575 7702 12114 12824a d b
4N2 b 2377 7130 11319 10959a d b
4N2b b 2318 6955 10672 11103a d b
4N3 b 3201 9603 16005 56759a 13338 b
4N3b b 3858 11286 19292 70248a 16111 b
4N4 b 3236 9695 16178 54102a 13425 b
4N4b b 3566 10437 17831 63794a 14870 b
No.3 1881a 329 925 558 c c c
No.4 285 230 504 d c c c
Chapter 5. Shear Design of Deep Pile Caps 73

Table 5.3: (con’t) Summary of ACT Building Code and CRSJ Handbook predictions
(shear).

Specimen One-Way Shear (kN) Two-Way Shear (kN)


Name ACI CRSI Column Pile
1977 1983 [11.8] ACT CRSI
Al b 578 1646 2714 2916a 1458 1996
A2 b 662 1848 3078 3344a 1672 2288
A3 b 704 1934 3246 3556a 1778 2432
A4 b 578 1646 2714 2916a 1458 1996
A5 b 654 1830 3046 3308a 1654 2264
A6 b 644 1790 2988 3252a 1626 2224
A7 b 620 1750 2902 3138a 1569 2148
A8 b 662 1848 3078 3344a 1672 2288
A9 b 654 1828 3042 3302a 1651 2260
AlO b 538 1554 2550 2722a 1361 1860
All b 526 1526 2498 2660a 1330 1820
A12 b 634 1784 2962 3208a 1604 2196
Bl b 516 2066a 2584 b 3308 b
B3 b 600 2344a 3002 b 3843 b
SS1 b 69 186 254 122 d 228
SS2 b 68 178 249 122 d 228
SS3 b 68 181 250 121 d 226
SS4 b 71 190 258 122 d 228
SS5 b 84 229 296 134 d 251
SS6 b 89 229 315 134 d 251
SG2 b 65 164 261 101 d 185
SG3 b 84 164 273 101 d 185
A 3248 2400 6056 6352 2360a d 6250
B 3408 2084 5306 4278 1839 d 2764
C 6304 1820 4938 3734 1899 d 2989
D 3770 2432 6348 4726 1968 d 3107
E 4478 3076 8140 7078 2475 d 3970
F 1604 572 1224 1618 c c c
a = Increased capacity since piles partially within critical section;
b = Infinite capacity since piles totally within critical section;
c = Procedure not applicable;
d = CRSI prediction not applicable use ACT.
-
Chapter 5. Shear Design of Deep Pile Caps 74

Table 5.4: Comparison of ACT Code and CRSI Handbook predictions: ratio of measured
capacity to predicted capacity and failure mode.

Specimen Reported
Name ACT ‘77 ACT ‘83 ACT [11.8] CRSI Failure
Mode
2N1 1.96 s
1 6.52 s 2.17 i 2.66 s
1 s
2Nlb 2.22 s
1 7.38 i 2.46 s
1 3.55 s
1 s
2N2 0.91 b 6.03 s 2.01 s
1 1.21 i s
2N2b 0.96 b 8.28 i 2.76 Sj 1.94 i S
2N3 1.16 b 6.56 .5 2.19 s
1 1.31 i S
2N3b 1.07 b 7.32 i 2.44 s 1.46 s
1 s
3N1 1.93 s 2.60 s 1.11 2 1.93 s
1 s
3Nlb 1.81 s 2.86 s 1.04 i 1.86 si s
3N3 1.01 f 2.43 s 1.01 f 1.01 f s
3N3b 0.84 f 2.70 s 0.90 s 0.84 f s
4Nl 0.86 f 2.82 1
s 0.94 s 0.86 f s
4Nlb 0.80 f 2.55 s, 0.85 i 0.80 f s
4N2 0.83 f 2.71 L
8 0.91 i 0.83 f S
4N2b 0.85 f 3.13 s
1 1.04 si 0.85 f s
4N3 0.77 f 1.99 s 0.77 f 0.77 f s
4N3b 0.82 f 2.29 s 0.82 f 0.82 f s
4N4 0.75 f 2.28 s 0.76 5j. 0.75 f 5
4N4b 0.79 f 2.41 i 0.82 s 0.79 f s
No.3 1.16 f 1.81 s 1.16 f 1.16 f f
No.4 1.07 f 1.26 1
s 1.07 f 1.07 f f
Chapter 5. Shear Design of Deep Pile Caps 75

Table 5.4: (con’t) Comparison of ACT Code and CRSI Handbook predictions: ratio of
measured capacity to predicted capacity and failure mode.

Specimen Reported
Name ACT ‘77 ACT ‘83 ACT [11.8] CRSI Failure
Mode
Al 0.88 f 1.92 s
1 0.88 f 0.88 f s
A2 1.12 f 2.15 s
1 1.12 f 1.12 f s
A3 1.07 f 1.90 s
1 1.07 f 1.07 f s
A4 0.98 f 2.13 s 0.98 f 0.98 f s
A5 1.11 f 2.14 s 1.11 f 1.11 f s
A6 0.98 f 1.91 s
1 0.98 f 0.98 f s
A7 1.30 f 2.65 s 1.30 f 1.30 f s
A8 1.19 f 2.28 s
1 1.19 f 1.19 f s
A9 1.15 f 2.31 s 1.15 f 1.15 f s
AlO 1.23 b 2.83 s 1.23 b 1.23 b f
All 1.39 b 3.12 i 1.39 b 1.39 b f
A12 1.30 f 2.59 s 1.30 f 1.30 f f
Bl 1.14 b 4.03 s
1 1.14 b 1.14 b s
B3 1.16 f 2.95 s
1 1.16 f 1.16 f f
SS1 2.05 2 3.62 i 2.05 s
2 2.05 2 S
SS2 2.11 f 3.60 i 2.11 f 2.11 f s
SS3 2.05 s
2 3.65 s 2.05 2 2.05 s 2 s
SS4 1.85 2 3.18 s
1 1.85 s
2 1.85 s2 s
SS5 1.97 2 3.14 i 1.97 2 1.97 2 S
SS6 2.09 s
2 3.15 s
1 2.09 2 2.09 s 2 s
SG2 1.71 2 2.66 s
1 1.71 2 1.71 2 S
SG3 1.75 2 2.11 s
1 1.75 2 1.75 2
A 0.79 f 0.79 f 0.79 f 0.79 f f
B 1.19 2 1.19 2 1.19 s
2 1.19 s2 s
C 1.52 2 1.59 s
1 1.52 1.52 2 S
D 1.64 2 1.64 s
2 1.64 2 1.64 2 S
B 1.90 2 1.90 s
2 1.90 2 1.90 2 S
F 1.89 s 5.29 si 2.47 s 1.87 i s

f=flexure; b=column bearing; s


=one-way shear; 2
1 s
= two-way shear; s=shear.
Chapter 5. Shear Design of Deep Pile Caps 76

Table 5.5: Comparison of proposed design methods with experimental results.

Specimen Predicted Predicted (2) Experimental Exp. Exp.


Name (1) Flexure Shear Pred.(1) Pred.(2)
(kN) (kN) (kN) (kN)
2N1 1663 2128 1053a 2059 1.24 1.96 s
2Nlb 2921 3570 1438a 3187 1.09 2.22 s
2N2 2581 3109 2148 2942 1.14 1.37 s
2N2b 4186 5051 3504 5100 1.22 1.46 s
2N3 3893 4835 2552 4413 1.13 1.73 s
2N3b 5317 6443 3622 5884 1.11 1.62 s
3N1 3207 3256 2l30a 4119 1.28 1.93 s
3Nlb 4107 4531 2707a 4904 1.19 1.81 s
3N3 5060 5070 7493 6080 1.20 1.20 f
3N3b 6170 6767 6869 6669 1.08 0.99 f
4N1 5025 6041 9037 6865 1.37 1.14 f
4Nlb 5398 6178 9790 6571 1.22 1.06 f
4N2 4876 5933 8868 6453 1.32 1.09 f
4N2b 4864 6512 8393 7247 1.49 1.11 f
4N3 6041 6208 10604 6375 1.06 1.03 f
4N3b 7018 7010 13993 8826 1.26 1.26 f
4N4 6704 7414 10825 7385 1.10 1.00 f
4N4b 7704 8150 12450 8581 1.11 1.05 f
No.3 467 480 732 596 1.28 1.24 f
No.4 250 254 285 289 1.16 1.14 f
Chapter 5. Shear Design of Deep Pile Caps 77

Table 5.5: (con’t) Comparison of proposed design methods with experimental results.

Predicted
Specimen Predicted (2) Experimental Exp. Exp.
Name (1) Flexure Shear Pred.(l) Pred.(2)
(kN) (kN) (kN) (kN)
Al 792 1030 1420 1110 1.40 1.08 f
A2 928 1031 1716 1420 1.53 1.38 f
A3 992 1020 1868 1340 1.35 1.31 f
A4 792 1030 1420 1230 1.55 1.19 f
A5 916 1031 1688 1400 1.53 1.36 f
A6 900 1020 1648 1230 1.37 1.21 f
A7 868 1030 1572 1640 1.89 1.59 f
A8 928 1031 1716 1510 1.63 1.47 f
A9 916 1030 1684 1450 1.59 1.41 f
AlO 696 1030 1296 1520 2.18 1.48 f
All 668 1030 1256 1640 2.46 1.59 f
A12 888 1030 1620 1640 1.85 1.59 f
Bl 1008 1374 1592 2080 2.06 1.51 f
B3 1028 1030 1972 1770 1.72 1.72 f
SS1 98 97 122a 250 2.55 2.58 f
SS2 84 85 122a 245 2.92 2.88 f
SS3 116 144 121a 248 2.14 2.05 s
SS4 112 116 122a 226 2.02 1.95 f
SS5 149 238 134a 264 1.77 1.97 s
SS6 149 347 134a 280 1.88 2.09 s
SG2 71 231 lOla 173 2.44 1.71 s
SG3 71 543 lOla 177 2.49 1.75 s
A 1448 1445 1915 1781 1.23 1.23 f
B 1659 1662 1696 2189 1.32 1.32 f
C 1499 1502 1684 2892 1.93 1.93 f
D 2320 3454 1968a 3222 1.39 1.64 s
E 3505 5084 2475a 4709 1.34 1.90 s
F 1774 3472 1303 3026 1.71 2.32 s
ACI ‘77 prediction critical; s=snear critical; f=flexure critical
Chapter 5. Shear Design of Deep Pile Caps 78

Flexural Calculation

s
/ d

h L
m

I I / I I
I I j I I I I
\ / \ J \
Two Way

0n Way Shear
Punching Shear

/ /\ \ I
! ‘ I I
I Ijl I I I I

/ H

d/2 - I
/
dP) I I I I

L
m n

. . .

Figure 5.1: ACT Code specified critical sections for fiexure and shear investigation of pile
cap.
Chapter 5. Shear Design of Deep Pile Caps 79

Deep Beam Behaviour d


H
N___ One Way Shear
,__/
I
dpi I
)%_ “—

a/2

Figure 5.2: A deep two-pile cap


Chapter 5. Shear Design of Deep Pile Caps 80

-- One WayShear?
— ,--‘ I
/
/

r———--—-—-——-7—I

Two Way
d/2
Punching Shear
-
÷

I’
/ /
—,
I—’

Figure 5.3: A deep three-pile cap


Chapter 5. Shear Design of Deep Pile Caps 81

/ I

I
/ I
\ I /
4— _,

Two Way
Punching Shear _Hd12m__

(a)

/
/ /

N’ /

Two Way
Punching Shear

(b)

Figure 5.4: Comparison of two-way punching shear calculations: (a) the cap with square
column; (b) the cap with circle column.
Chapter 5. Shear Design of Deep File Caps 82

0.3 0.1

2
)[a.z.s
0
.4
U.

(a)
0 0.Z 0.4 0. 0.8
w-/d

‘ I

:
3 ——

C,.
— —

‘_‘
1
•_Ic

rut
‘1

IC
k

k
}c)
Ii

‘JcI÷d/cZ
s
{(I+d1c_..I

1
4
r
I -_w. (b)
U 0.1 0.Z 03 0.4 0.5 O.

Figure 5.5: CRSI approach for shear design of deep pile caps: (a) allowable shear stress,
v, for one-way shear while w/d < 1.0; (b) allowable shear stress, v, for two-way shear
while w/d < 0.5, from Ref.[36}.
Chapter 5. Shear Design of Deep Pile Caps 83

C )
Given:
Column size
Pile group
Pile capacity
fy

pile cap?

No
Yes
(ACI 77)
Have no appropriate design
Select d: procedure for shear
One-way shear
(11.11.1.1, 11.1, 11.2, 11.3and 11.5)
Two-way shear Select d:(ACI 83)
(11.11.1.2, 11.11.2 and 15.5) One-way shear (11.1—3)
Minimum footing depth Have no appropriate design
(15.7) procedure for two-way shear

Select d: (CRSI)
One-way shear (CRSI)
Two-way shear (CRSI)
Find As:
Moment calculation
(15.4.1 and 15.4.2)
Detailing
(15.4.3, 15.4.4 and 12.2)

Compute cap thickness h:


(7.7.1)

Shear check:
(for Individual pile)
Perimeter shear
(11.11.1.2 arid 11.11.2)
Beam shear
(11.11.1.1 and 11.3,1.1)

Bearing check:
(10.16)

Figure 5.6: Flow chart for ACT and CRSI design procedures for pile caps
CD
cD(D
CD Cf) <

CD ‘C) CD
D
C-) ( CD
Cl) — —
C—
-4. f-Th :3-a CD
T1
H CD . D
0 ‘0 0
U) D 0.
C’, Q)
9- 0
-a —i, N
0
0 0 D
0 CD
CD C)
CD
C’,

p
* -. co
QCD 01
0(n C)
CD
0-
D

CD 01
00

3c 0 0
0
CD m
0) D
Cl) ci
0 CD0
0
:3 0)’
I CD 10 CD
CD
Chapter 5. Shear Design of Deep Pile Caps 85

1
D

(a) (b) (c)

Figure 5.8: Loading geometry of compression struts with linearly varying cross section.
Chapter 5. Shear Design of Deep Pile Caps 86

(a) (b)

N
:,

/ N

(c) (d)

(e)

Figure 5.9: Various layouts of main reinforcing bars used by Blévot and Frémy, Ref.[32j.
Chapter 5. Shear Design of Deep Pile Caps 87

(1) Nil (2) Nominal

L I
‘p

(3) Full (4) Full-plus-bob

Figure 5.10: Various anchorage lengths used by Clarke, Ref. [44].


Chapter 5. Shear Design of Deep Pile Caps 88

36” ‘I

d=1O”
,-—

/ I i

(a)
‘i c=1O”

b=c=1O
(ki p)
°
40
300 lAd ‘77 CRSL—
/ Proposed (2)
200 -

-
ACI [11.8]
100

(b) ——ZC 183


0
12 14 16 18 20 22 24 26 28 30
d(in.)

Figure 5.11: Comparison of one-way shear design methods for two-pile caps: (a) plan
view of pile cap; (b) to (d) influence of pile cap depth on column load for various pile
cap widths.
Chapter 5. Shear Design of Deep Pile Caps 89

b=2c=20”
p 700
(kip) 600
500
400
300
200
100
(c) 01

d (in.)

b=4c=40” Proposed (2)


700

(kip) 600
500
400
300
200
100
(d) 0
12 14 16 18 20
d (in.)

Figure 5.11: (cont’d) Comparison of one-way shear design methods for two-pile caps: (a)
plan view of pile cap; (b) to (d) influence of pile cap depth on column load for various
pile cap widths.
Chapter 5. Shear Design of Deep Pile Caps 90

/dp8 15”

/
36”

I
—-. f.. 14’ I

(a) 15”

15” 36” 15”


—. .-4 —

P 1800
1600
(kip)
1400
1200
1000
800
600
400
200

(b) 12 16 20 24 28 32 36 -

d (in.)

Figure 5.12: Comparison of two-way shear design methods for a typical four-pile cap: (a)
plan view of pile cap; (b) influence of pile cap depth on column load.
Chapter 5. Shear Design of Deep Pile Caps 91

4
Pexp
ACI ‘77
I I
pred.
3
0 Flexure
One-way shear
A
Two-way shear

2
A

0
*
1 U
0 .0_000
0 0 0
Mean 1.31
CCV 34.8%

2000 4000 6000 8000 10000 12000


Pexp (kN)

Figure 5.13: Comparison of ACT ‘77 predictions with experimental results.


Chapter 5. Shear Design of Deep Pile Caps 92

4
Pexp ACI ‘83
Ppred.
• One-way shear

:
- Two-way shear
:

2 A

1
C
Mean 3.09
C0V55.1%

2000 4000 6000 8000 10000 12000

Pexp (kN)

Figure 5.14: Comparison of ACI ‘83 predictions with experimental results.


Chapter 5. Shear Design of Deep Pile Caps 93

Pexp
ACI [11.8]
Ppred
-

• One-way shear
A Two-way shear

2
A

A
*

,C*A
00 A
1 U

0 10

Mean 1.42
CCV 38.9%

2000 4000 6000 8000 10000 12000

Pexp (kN)

Figure 5.15: Comparison of ACT ‘[11.8] predictions with experimental results, clause 11.8
considered.
Chapter 5. Shear Design of Deep Pile Caps 94

4
Pex
CRSI
pred.
Flexure
• One-way shear
A
Two-way shear
Bearing
.
I •

• •
I
A
• A
.

1 0
0 0&00
0 0
Mean 1.40
CCV 40.7%

0 2000 4000 6000 8000 10000 12000


Pexp. (kN)

Figure 5.16: Comparison of CRSI predictions with experimental results.


Chapter 5. Shear Design of Deep Pile Caps 95

4
Pexp.
Ppred. Proposed (1)
3.

. .

2
• $

• .
•.
.
••, •
• • •
• • .
•• •• • •
• •• • • •• •

Mean 1.57
CCV 29.6%
I
0
0 2000 4000 6000 8000 10000 12000

Pexp (kN)

Figure 5.17: Comparison of proposed method (1) predictions with experimental results
of all specimens in Table 5.1.
Chapter 5. Shear Design of Deep Pile Caps 96

4
Pexp
ppred. Proposed (1)
3 (bunched reinforcement)

2
.
.
.
.
.
.
.
•. .
.

Mean 1.32
CCV 16.5%
0
0 2000 4000 6000 8000 10000 12000
Pexp (kN)

Figure 5.18: Comparison of proposed method (1) predictions with experimental results
of specimens with bunched reinforcement in Table 5.1.
Chapter 5. Shear Design of Deep Pile Caps 97

4
Pexp.
Proposed (2)
I
Ppred
3 o
° “Flexure”
0
t
“Shear
L

.
.
S
2 .
• 0
0
• o .
0.
0
o 0 o
0

1 00 0

Mean 1.55
• CCV 27.8%
I
0 I

o 2000 4000 6000 8000 10000 12000

Pexp. (kN)

Figure 5.19: Comparison of proposed method (2) predictions with experimental results
of all specimens in Table 5.1
Chapter 5. Shear Design of Deep Pile Caps 98

3
f b • Adebar et a
— °
elévot and Frémy
.c ‘ • Clarke
I Deutsch and Walker
Sabns and Gogate
8
2
.

.
..

0
D
0
0 —
0

0
C
C

2 3 4 5

Figure 5.20: Relationship of measured ultimate bearing stress and confinement on column
zones of pile caps in Table 5.1.
Chapter 6

Shear Failure of Beams Without Stirrups

6.1 Introduction

In the previous chapters the discussion focused on the shear resistance of deep pile caps.
The model that was proposed assumes that the load is transmitted from the column to
the piles by direct compression struts, and that the tensile stresses which cause cracking
of the compression struts is due only to the transverse spreading of compression stresses.
As pile caps become more slender, other mechanism will be involved in creating tensile
stresses which may lead to a shear (diagonal tension) failure. Note that this issue was
accounted for in the proposed design method (Section 5.4) by suggesting that the ACT
Building Code (empirical) method be used for slender pile caps.
In order for this thesis to be a relatively comprehensive treatment of pile caps and
other members without transverse reinforcement, it is necessary to consider the mecha
nisms involved in the shear failure of more slender members. Towards that end, a study
was made of the shear resisting mechanisms in slender beams. While this study is really
a separate topic, it is closely related to the first part of this thesis (transverse splitting
in deep members). In fact, as will be shown in Chapter 8, the truss model which is
associated with the transverse splitting mechanism in deep members is in fact very sim
ilar to the truss model associated with what is believed to be an important shear failure
mechanism in slender beams without transverse reinforcement.
It should be pointed out however, that unlike the first part of this thesis, which was a

99
Chapter 6. Shear Failure of Beams Without Stirrups 100

complete study of transverse splitting (it resulted in the development of a new improved
design method for pile caps), this study on shear in slender beams is much more of a
pilot study. In this and the next chapter considerable information is given about the
shear resisting mechanisms in slender beams, but additional research is needed before
the concepts presented here can be implemented into a shear design procedure which can
be used by practising engineers.
This chapter is divided into seven parts. A brief review of the literature is presented
and some important conclusions about the current knowledge of shear failures in beams
without web reinforcement are summarized first. Then, an interpretation of an important
shear failure mechanism is presented based on the deformation compatibility of critical
diagonal crack propagation. Next, bond influence upon internal stress distributions of
both uncracked and cracked beams are investigated by using linear elastic finite elements
and the concepts of arch action and beam action. This is followed by a study on shear
displacements along vertical cracks and inclined cracks in beams, and the presentation
of a load transfer mechanism based on studies in this chapter and the experimental
measurements of previous work. As an application of the understanding obtained in this
study, the test results carried out by various researchers are explained in the sixth part.
Finally, the conclusions arrived at are summarized.

6.2 Brief Review of the Literature

State-of-the-art summaries of the shear resistance of structural concrete members without


transverse reinforcement have been reported previously in Ref. [38, 48-51]. The literature
review presented here focuses on more recently developed models for shear resistance of
slender beams and understanding the transition from deep beams to slender beams.
Chapter 6. Shear Failure of Beams Without Stirrups 101

6.2.1 Transition from Deep Beam to Slender Beam

The distribution of the transverse compression stresses within the shear span estimated
by Mau and Hsu [52] is shown in Figure 6.1. When a/h = 0, transverse compression stress
is maximum at the line of actions. While a/h = 0.25 and 0.5, the maximum values of
transverse compression stress still occur at the centre of the line in. connection with load
point and support point, but the magnitudes decrease with increasing a. While a/h = 1,
the distribution of transverse stress shows the characteristics of two humps. These two-
humps become more distinct and the transverse compressive stresses approach zero at
the centre of the shear span when a/h = 2. Hence, it is seen that direct compressive
stress field between the load and the support is formed when a/h < 1.

Following the internal force flow, Schlaich et al. [3] gave a description of the transition
from deep beams to slender beams using truss models for uncracked beams. See Figure
6.2. If a single load is applied at a distance a < h near the support, the load is carried
directly to the support by a compressive stress field as simulated by a simple compression

strut. A further examination indicates that the transverse tensile stresses will be intro
duced in the compression strut as shown by the refined truss model and as discussed in
Chapter 1. With increasing shear span a, the compressive member C
1 joins the part of
the tensile force T
1 and simultaneously the transverse tensile ties blend into the vertical
ties of the truss model, which are now needed for hanging up the shear forces in slender
beam. Thus the beam having a < h, where the load can be transmitted directly to the
support, is treated as a deep beam.
Collins and Mitchell [39] compared the shear strengths of a series of simply supported
beams tested by Kani [13] with the predicted capacities from both sectional and strut
and-tie analyses for assumed cracked beams. In Kani’s tests the shear span to depth
ratio a/d varied from 1 to 7 and no web reinforcement was provided. The comparison
Chapter 6. Shear Failure of Beams Without Stirrups 102

indicated that the shear resistance is governed by strut-and-tie action, which assumes
the loads are carried to the supports with direct compression struts. The failures are
governed by a strut-and-tie model (crushing of the compression strut) at a/d less than
about 2.5, but are governed by a sectional model for aid greater than 2.5. See Figure
6.3. Their conclusions suggest that beams with aid < 2.5 can be treated as deep beams.
The truss models developed by Al-Nahlawi and Wight [53] illustrate that when a/d>
1, the shear failure of the beams without transverse reinforcement is characterized by the
failure of concrete tension ties. Their work suggested that the beams with a/d> 1 could
be considered as slender beams. In the next section more details of Al-Nahlawi and
Wight’s truss models will be presented.

6.2.2 Behaviour of Slender Beams

Based on the modified compression field theory which suggests that concrete tensile stress
is present in cracked reinforced concrete members, a simple truss model with concrete
tension ties was developed by Adebar [54] for a diagonally cracked beam which is simply
supported, subjected to point loads and reinforced with only longitudinal reinforcement.
See Figure 6.4. The inclination of compression struts in the web is equal to half of the
inclination of the uniformly spaced diagonal cracks, and concrete tension ties are perpen
dicular to the compression struts. The shear capacity of the beam depends primarily on
the crack width, which is strongly influenced by the crack inclination.
A mechanicaI model was developed by Reineck [55] based on a tooth model con
sidering the shear carrying actions of friction along cracks, dowel force of longitudinal
reinforcement, cantilevering action of tooth from the compression zone and shear force
component in the compression chord. After a detailed analysis of the shear carrying
actions involved, a stress field for a beam without transverse reinforcement was proposed
and a truss model representing this stress field was developed. See Figure 6.5. This truss
Chapter 6. Shear Failure of Beams Without Stirrups 103

model is similar to the one given by Adebar [54] except that the inclination of uniformly
spaced diagonal cracks are assumed to be 60 degrees. The failure is characterized by a
crack further propagating into the compression zone and breaking off the tooth, and is
defined by the critical slip along the inclined crack which reaches a critical crack width.
Based on an idea that a truss model must include a concrete tension member that
will fail in tension before yielding of the flexural reinforcement, Al-Nahiawi and Wight
[53] developed the truss models shown in Figure 6.6. The truss models have a 45-deg
compression strut originating at the node under the applied load and a 35-deg compres
sion strut originating at the support. The shear capacity of beams without stirrups is
controlled by the failure of a concrete tension tie when its maximum tensile stress reaches
the tensile strength of concrete subjected to transverse compression. They consider the
failure of a concrete tensile tie as corresponding to the unchecked propagation of an in
clined crack. One of conclusions from their models is that for beams with a/d> 1, shear
failure modes are similar and indicated by the failure of concrete tensile ties.
Muttoni and Schwartz [56] developed a structural model for the loading stage when
a typical crack pattern has formed. See Figure 6.7. They indicated that the form of this
critical crack pattern is accompanied by a collapse of three shear carrying actions, which

are cantilever, interlocking and dowelling actions, and then the direct transfer of the load
to the support also becomes impossible because of the wider crack. Their model shows
that the compression strut turns in the central region of the beam and acts together with
a concrete tie. Failure occurs either when the tensile strength in the region D is reached
or when the strength in zone E, which is subjected to both tension and compression, is
exceeded.
After carrying out tests designed to test the validity of current design concepts,
Kotsovos [57] has concluded that aggregate interlock and dowel action make a negli
gible contribution to the load-carrying capacity and the strength of compressive zones
Chapter 6. Shear Failure of Beams Without Stirrups 104

increases due to triaxial stress/strain states, thus making a significant contribution to


shear capacity. In an attempt to summarize the experimental information, the concept
of the ‘compressive force path’ has been developed by Kotsovos. The concept considers
that the load-carrying capacity of beams without transverse reinforcement is associated
with the strength of uncracked concrete in the region of the paths along which compres
sive forces are transmitted to the supports. See Figure 6.8 (it should be noted that the
original model presented by Kotsovos in this figure does not satisfy equilibrium). The
shear failure is believed to be related to the development of tensile stresses mainly in the
region of the path. The tensile stresses may be developed due to changes in the path
direction, the varying intensity of compression stress field along the path and bond fail
ure at the level of the tension reinforcement between two consecutive fiexural or inclined
cracks, etc. More discussion about Kotsovos’ experimental work is given in Section 6.7.
In summary, it is generally accepted that point loads can be transmitted directly
to the support in beams either uncracked or cracked while a < d. However, there are
different opinions about load transfer mechanisms for cracked beams when 1 <a/d < 2.5.
Secondly, all the investigators, who tried to give rational interpretations of shear fail
ure mechanism of reinforced concrete beams without transverse reinforcement, believe
that the tensile strength of concrete plays an important role in the shear resistance for
both uncracked beams and cracked beams, which is illustrated by various truss mod
els composed of tension ties. However the interpretations of where and how the tensile
strength of concrete is mobilized for shear resistance are different. Adebar [54], Reineck

[55], A1-Nahlawi and Wight [53] considered that concrete tensile strength is needed for
hanging up the shear force in the lower part of beams. Muttoni and Schwartz [56] con
sidered that concrete tensile strength is needed in the upper part of beams to help the
compression strut to deviate from the critical diagonal crack. In Kotsovos’ compressive
force path concept [57], concrete tensile strength mainly contributes to the change of
Chapter 6. Shear Failure of Beams Without Stirrups 105

the compressive force path direction. Thus it is evident that there is still no univer
sally accepted shear failure mechanism for reinforced concrete beams without transverse
reinforcement.

6.3 One Interpretation of Shear Failure of Beams without Stirrups

In order to investigate the shear failure mechanism, the behaviour of a slender, longi
tudinally reinforced concrete beam simply supported and subjected to two symmetrical
point loads is examined herein. See Figure 6.9. The vertical flexural cracks first form in
the pure bending region and then the additional cracks form in the shear spans between
the concentrated loads and supports. The vertical cracks in the pure bending region keep
propagating upward vertically with increasing load. However the flexural cracks formed
in the shear spans, after extending vertically to longitudinal reinforcement level, will
become slightly inclined toward the load. Traditionally it is believed that these cracks
become inclined because of shear. As the loading is further increased, a characteristic
inclined crack, the so-called critical diagonal tension crack, forms. Its propagation into
the compression zone of the beam near the section of maximum moment is usually rel
atively sudden, splitting the beam into two pieces and causing collapse. Above shear
failure observations are well known and quite general.
As reviewed briefly in Section 6.2, the various interpretations for shear failure mech
anism given by previous studies are generally based on force transfer concept, which
emphasizes the various actions involved in shear carrying mechanism and on the satis
faction of equilibrium conditions. On the contrary, an alternate interpretation of shear
failure mechanism is based on the deformation compatibility of a critical diagonal crack.
It is very reasonable to choose the crack opening of the critical diagonal crack at
the reinforcement level as an important geometric parameter to keep track of the crack
Chapter 6. Shear Failure of Beams Without Stirrups 106

propagation. See Figure 6.10. If the crack opens horizontally only, the crack propa
gates vertically, which is the case in the pure bending region. If the crack opening has
both horizontal and vertical components, the crack propagation will deviate from verti
cal direction and become inclined. This is the situation in the shear span range. The
orientation of stable crack propagation can be approximately determined by the ratio
of horizontal and vertical opening components. Figure 6.10 shows this geometric rela
tionship. Test observations indicate that there are several inclined cracks in the shear
span, all of which have the potential to become the critical diagonal crack. The shear
failure observations show that only one of the inclined cracks, the critical diagonal crack,
penetrates into the compression zone and extends horizontally to load point before the
beam collapses. This means that of all inclined cracks, the critical diagonal crack that
distinguishes itself from other inclined cracks must have a significant vertical opening
component, which is geometrically compatible with this crack propagation, at the rein
forcement level or the other locations along the crack. Then it is logical to infer that the
formation of a horizontal crack, which introduces the significant vertical opening compo
nent of the critical diagonal crack, is a critical stage in the shear failure of a reinforced
concrete beam without transverse reinforcement.

It is of interest that more information can be found for the phenomenon described
above. The shear failure descriptions given in ACT 426 Committee report [38] are:

Beams may exhibit a number of different modes of shear failure, the most
common of which is the crushing or shearing of the compression flange over
the inclined crack which is often accompanied or initiated by splitting along
the tension reinforcement.

It can be seen from this description that there are some uncertainties of which one occurs
first in this statement, i.e., the crushing of the compression flange or splitting along the
Chapter 6. Shear Failure of Beams Without Stirrups 107

tension reinforcement.
Fortunately, there is valuable experimental evidence that can help to clarify this
point. The evidence comes from the experimental investigation carried out by Chana
[58]. Chana used a high speed tape recorder in conjunction with electrical demountable
strain transducers to continuously monitor crack widths at critical locations while the
tested beams were approaching failure. His test results clearly show that the beam shear
failure was preceded by splitting along the longitudinal reinforcing bars.
Another factor which is of help to understand this mechanism is the crack control
efficiency of longitudinal reinforcement. The major function of longitudinal reinforcement
in concrete beams is to compensate for the weakness of concrete low tensile strength and
provide enough tensile strength to have high concrete compressive strength fully utilized.
This function is well obtained in the pure bending region of the beam. After the vertical
cracks have occurred in this region, their propagation are most efficiently controlled by
the longitudinal reinforcement, which is at right angle to them. The crack opening at
the reinforcement level is proportional to the deformation of the reinforcement if there
is proper bond between concrete and reinforcement. Under a same load level a larger
percentage of reinforcing bars results in smaller stresses in the bars (or smaller crack
opening). On the other hand, the force system in the pure bending region also favours
checking crack propagation. If more reinforcement is arranged at the bottom of the beam,
more compressive force (or a larger compression zone) can be developed at the top of the
beam. This compressive force effectively restrains the crack from penetrating vertically
into the compression zone.
However in the shear spans of the beam, the cracks are diagonal rather than vertical.
The reinforcement is skew to the cracks. In order to check the propagation of the diag
onal cracks, not only axial action but also dowel action of the reinforcement should be
mobilized. The crack control of longitudinal reinforcement to inclined cracks is much less
Chapter 6. Shear Failure of Beams Without Stirrups 108

efficient than to vertical cracks. The flatter the inclined cracks are, the less the crack con
trol efficiency of longitudinal reinforcement. The crack control efficiency of longitudinal
reinforcement is least when the cracks are horizontal.
From the above discussion, an intuitive and rational conclusion about shear failure
mechanism of slender, longitudinal reinforced concrete beams is that splitting along the
longitudinal reinforcement, which produces a considerable vertical opening component to
the critical diagonal crack, is the immediate cause of shear failure. In general, a distinct
characteristic of slender beams without transverse reinforcement is lack of the ability
to control horizontal crack propagation and the brittleness of shear failure results from
the occurrence of either flat inclined or horizontal cracks. Should the defect have been
overcome by any mechanism, the shear resistance behaviour will be greatly improved.
Traditionally, the transverse reinforcement in conjunction with longitudinal reinforcement
is provided for shear design of structural concrete members.

6.4 Bond Effect in Longitudinally Reinforced Concrete Beams

It was concluded above that the horizontal splitting along longitudinal reinforcement,
which precedes the propagation of a critical diagonal crack into compression zone, indi
cates shear failure of a beam without transverse reinforcement. In this section the load
carrying mechanism before the occurrence of this horizontal splitting and the cause of
the horizontal splitting are investigated. The influence of bond between concrete and
reinforcing bars upon the load transfer mechanism is focused on as so much information
from previous studies [13, 59-67] indicates that the bond has a significant effect, which
was either ignored or considered to a less extent in previous studies of shear failure of

beams without transverse reinforcement.


Kani [13] tested a series of point loaded and simply supported beams without web
Chapter 6. Shear Failure of Beams Without Stirrups 109

reinforcement to investigate the influence of various bond qualities on the shear failure
mechanism. By introducing an intermediate layer of a vermiculite-cement mix with
different elastic properties and strengths between the reinforcement and the concrete in
the shear span, the varying bond quality (varying average ultimate bond stress) was
obtained. The tests showed that “the better the bond, the lower is the load capacity of
the beam.” As the extreme case, the no bond beam presented flexural failure with no
single crack occurring in the shear spans. Though the test results did not tell the whole
story about the shear failure mechanism of beams without transverse reinforcement, they
did reveal the fact that bond plays an important role in load sustaining mechanism.
The second example showing the interaction of bond and shear failure is the bond
tests done by Ferguson et al. [61, 62] and Kemp et al. [66]. Ferguson et al. tested beams
to investigate development length of reinforcing bars in bond. They found that there
were always the combinations of diagonal tension failure and bond splitting failure for the
narrow beams that were tested. Kemp et al. [64, 66] observed that the shear crack formed
only after a longitudinal bond crack and bond slip had occurred for the cantilever-type
bond specimens tested (more discussions on these tests are given in the next chapter).
The third example is the experimental work by Mains [59]. His measurement of bond
stresses along reinforcing bars in beams indicated that the distribution was not uniform
and that there was a significant localization after the occurrence of the critical diagonal
crack. This evidence conflicts with the basic assumption adopted by many authors that
the stress variation of longitudinal reinforcement in cracked beams has a similar shape
as the moment diagram.

6.4.1 Bond Influence in Uncracked Beams

In order to investigate the effect of bond on the load transfer mechanism of reinforced
concrete beams without web reinforcement, a simply supported and centrally loaded
Chapter 6. Shear Failure of Beams Without Stirrups 110

beam was studied to find the internal stress distributions in the uncracked concrete.
See Figure 6.11. Based on the work of Schlaich et al. [3], this beam can be divided
into B-regions and D-regions, which are also shown in the figure. The prediction using
classical beam theory, which assumes that cross sections remain plane during deformation
(Bernouli assumption), is only valid for B-regions and not valid for D-regions so that the
linear elastic finite element method was used to investigate the effect of bond on the
internal stress distributions. Nine node Lagrange quadratic plane elements were used in
analysis. The biaxial stresses in x and y directions and principal stresses were calculated
at each node of the element.
First, perfect bond and no bond cases for slender beams were studied. The beam
was assumed to already have a single flexural crack formed at mid-span. Because of the

symmetry of the problem, only half of the beam was modeled. See Figure 6.12. In Figure
6.12(a) the force T modelling the tension force in the reinforcement is applied at the the
front of the beam, which is the simulation of the case with perfect bond between the
longitudinal reinforcement and concrete. In Figure 6.12(b), the corresponding force T
is applied at the end of the beam to represent the no bond case. Final analysis results,
showing the internal stress distributions, are also given in Figure 6.12.
The shear stresses, in both the perfect bond case and the no bond case, give similar
distributions in the B-region. The shear stresses are confined within a certain part along
the cross sections in the D-regions and distributed uniformly in the B-regions. The shear
stress distributions demonstrate the shear transfer mechanism in the uricracked beams,
but do not indicate the load transfer mechanism. The fundamental difference between
these two internal stress distributions is that the whole cross sections are subjected to
compression for the no bond case whereas the cross sections are subjected to tension on
the bottom and compression on the top for the perfect bond case.
The load transfer mechanism can be found by looking at the internal force flows in
Chapter 6. Shear Failure of Beams Without Stirrups 111

the beam. Figure 6.13 shows two principal stress trajectories for perfect bond and no
bond situations respectively. In Figure 6.13(b), the load is transferred to the support
directly by a compression stress field (compression strut). The internal force flow is
consistent with the externally applied loads. The resultant of forces at the support and
the resultant of forces at the upper compression zone act along the same line but in
opposite directions. In Figure 6.13(a), the force system is different from that in Figure
6.13(b). As a requirement of static equilibrium, the action line of the resultant of forces
at the upper compression zone will still go through the point where the support reaction
and the tension force in the reinforcement intersect. However the force flow in the beam
is curved rather than straight so that the tensile strength is mobilized to make the force
flow change direction. Though the force flow in Figure 6.13(a) is different from that in
Figure 6.13(b), this force flow is consistent with the corresponding external force system.
As there is no horizontal force applied at the bottom corner of the beam, the force
flow originating from the support is almost vertical. Then the additional internal force
(tension force) is required to make the force flow from the upper load point to the support
possible.
It has been mentioned before that the load can be directly transferred to the support
in deep beams (a/d < 1). The transverse splitting of compression struts dominates shear
failures of deep beams. In this section the influence of various bond stress distributions
upon the transverse splitting of compression struts is also investigated. Various bond
stress distributions are shown in Figure 6.14. The analysis results are summarized in
Figure 6.15. The bond influence on the transverse tension of the compression strut is
concentrated at the lower end, where the transverse tensile stresses are amplified due to
bond. For the no bond case, the transverse tensile stress is almost uniformly distributed
along the inclined compression strut. The shear stress distributions along the cross
sections for various bond cases are similar to those in Figure 6.12.
Chapter 6. Shear Failure of Beams Without Stirrups 112

6.4.2 Bond Influence in Cracked Beams

In order to investigate the bond effect upon the load transfer mechanism of cracked
beams, the traditional concepts of arch action and beam action [68] are used in a sim
ple equilibrium analysis procedure. It is believed that the bond effect can be better
understood by using these simple concepts.
For a cracked beam in the shear span, the relationship between external moment and
internal moment of resistance, at a distance x from the support, can be approximated as
[68]

M = Tjd (6.1)

where T = T(x) =tensile force resultant acting at the centroid of longitudinal reinforce
ment, j = j(x) =variable coefficient and d =effective depth measured from the extreme
compression fibre to the centroid of longitudinal reinforcement. The shear force may be
expressed as V = dM/dx. Hence, by means of Equation 6.1 one obtains

V = jd + T- (6.2)

where V is the constant shear force acting through the shear span. The first term at
the right-hand side of Equation 6.2 is usually termed the “beam action”, which reflects
the change of the force in reinforcement, and the second term is called the “arch ac
tion”, which represents the inclination of the internal thrust force. The simultaneous
occurrence of both actions require the corresponding internal stress distribution and the
compatibility of the deformation within the beam.
A half beam with idealized flexural cracks is shown in Figure 6.16. The shear resis
tance modes of the varying tension force in reinforcement with the constant lever arm

and the constant tension force in reinforcement with the varying lever arm are the two
Chapter 6. Shear Failure of Beams Without Stirrups 113

extreme cases, which could be termed the “pure beam action” and the “pure arch ac
tion” respectively. The internal stress distributions for the “pure beam action” are well
known and first described by Mörsch [69]. See Figure 6.16. The shear stress is calcu
lated by horizontal equilibrium of a piece of beam element and is found to be uniformly
distributed along the cross section. However, the shear stress for the “pure arch action”
is concentrated as shown in Figure 6.12(b). The shear stresses combine with horizontal
compression stresses to produce a resultant which travels straight from the load point to
the support.
The internal shear stress distribution, for combined “beam action” and “arch action,”
is somewhat different from that of either one action. The same half beam, as shown in
Figure 6.16, is again given in Figure 6.17, and a plot of the tension force in the longitudinal
reinforcement is also given. It is assumed that the combined beam action and arch action
exists in the range of from the load point to section m — m with the resultant thrust force
shown in Figure 6.17(a). Correspondingly the variation of the force in the longitudinal
reinforcement is shown in Figure 6.17(b). Then the shear stress distribution along section
n. — n., coming from the contributions of both actions, can be determined. One part of
shear stresses is calculated by equilibrium consideration, in which the shear stress is
related to the bond stress between the concrete and the reinforcement, as shown in
Figure 6.16. Another part can be determined from the vertical component of inclined
thrust force. Summing up these two parts should be equal to external applied shear force.
See Figure 6.17(c). Equation 6.2 is the mathematical expression of Figure 6.17.

6.5 Shear Displacements along Cracks

After the shear stress distributions along the idealized vertical cracks have been examined
above, the corresponding shear displacement along the cracks is investigated in this
Chapter 6. Shear Failure of Beams Without Stirrups 114

section. It is generally believed that there are two kinds of deformations in cracked
beams, which could introduce the relevant shear displacements along the cracks [70].
One is the flexural rotation of the compression zone and another is the bending within a
concrete tooth caused by bond force LIT.
For the idealized vertical flexural cracks in the shear span, the compatibility of de
formations illustrates that only the bending of the concrete teeth causes the shear dis
placement along the cracks and thus develop uniform shear stress distributions along the
cracks, which is consistent with the assumed beam action. See Figure 6.18(a). The rota
tion of the compression zone makes the vertical crack opening and does not introduce the
shear displacement along the crack. This latter deformation for idealized vertical cracks
is only consistent with the beam elements subjected to pure bending moment, where no
bond force and no shear stresses will be developed.
However, the existence of the shear force in the shear span makes the cracks inclined
rather than vertical so that the both deformations have contributions to the shear dis
placement. See Figure 6.18(b). The shear displacement will still be uniform along the
inclined crack, if only the bending of the concrete teeth is considered. The shear dis
placement due to the rotation of the compression zone will not be uniform, with the
largest shear displacement occurring at the mouth of the crack. One of the important
conclusions about the shear displacement along the inclined cracks is that the flatter the
cracks are, the more shear displacement will be introduced along the cracks for the same
compression zone deformation.

6.6 Load Transfer Mechanism

Based on the previous detailed analysis, an insight into the load transfer mechanism can
be obtained by examining the shear failure process of a simply supported beam subjected
Chapter 6. Shear Failure of Beams Without Stirrups 115

to the point loading.


When the beam is loaded to a certain stage, the obvious aggregate interlock at the
critical inclined crack is mobilized with the company of the occurrence of considerable
shear displacement along the crack due to the increasing load and the reduced compres
sion zone (i.e., the increasing bending of the tooth and the increasing rotation of the
compression zone). This can be related to the experimental observation that only the
critical crack keeps opening and the widths of other cracks remain almost unchanged
under increasing loading. In addition, the shear displacement along the crack is always
accompanied by the crack opening for shear transfer by aggregate interlock action due to
the rough contacting surfaces. See Figure 6.19. At this stage, it can be considered that
both beam action and arch action already exist for shear resistance, based on the previous
analysis. With the loading being further increased, in addition to the shear stresses, the
compression stresses along the crack will also be introduced as a result of the longitudinal
reinforcement’s constraint to the crack opening. These shear stresses and compression
stresses acting on the crack are one of the direct reasons for the occurrence of the second
diagonal crack, which is commonly observed in shear failures of longitudinal reinforced
concrete beams. See Figure 6.20(a). As the loading is approaching this stage, just before
the occurrence of the second diagonal crack, the full arch action can be approximately
assumed in the range from the load point to the critical crack. See Figure 6.20. This
assumption is justified by measurements of reinforcement strain and concrete strain in
the shear span [59, 71, 72], an example of which is also shown in Figure 6.20(d).
If full arch action is assumed in the range from the load point to the critical crack,
and full beam action with a reduced lever arm is assumed in the range from the critical
crack to the support, the variation of tension force in reinforcement is shown in Figure
6.20(b). Because the magnitude of bond stress is proportional to the gradient of the
tension force in reinforcement, larger bond stresses will be developed near the support.
Chapter 6. Shear Failure of Beams Without Stirrups 116

If the variation of the tension force is further assumed as a solid line shown in Figure
6.20(c), the bond stress will be amplified locally in the zone near the critical crack. This
assumption can be fully confirmed by Mains’ measurements [59]. See Figure 6.21. It can
be concluded that the second diagonal crack or even splitting along reinforcement partly
results from this high bond stress.
A possible load transfer model is shown in Figure 6.22. The full arch action is assumed
in the range from the load point to the cross section n — n where the critical crack occurs.
Then the resultant compression force on the cross section n — n, F, should be as shown
in the figure. Because the compression force flow N originating from the support is not
along the same line as F, the tension force T is required for equilibrium. The F. has
two components, one meeting with the force N and another going downward to intersect
the longitudinal reinforcement.

6.7 Interpretation of Some Beam Test Results

With the help of the model presented above, it is possible to give an alternative interpre
tation of the experimental information presented by Kotsovos in Ref.[57, 73] and other
researchers in Ref.[56, 58, 74, 75].

Kotsovos [73] tested a series of simply supported beams with various arrangements of
shear reinforcement, and subjected to two-point loading with various shear span to depth
ratios (aid). See Figure 6.23. The main test results are given in Figure 6.24 which shows
the load-deflection curves of the beams tested. Series C and D beams were found to
have a load-carrying capacity significantly higher than that of series A beams which had
no shear reinforcement throughout their span. Series D beams, in all cases, exhibited a
ductile behaviour, which is indicative of a flexural mode of failure, and their load-carrying
capacity was higher than that of series A beams by an amount varying from 40 to 100%
Chapter 6. Shear Failure of Beams Without Stirrups 117

depending on a/d. In addition, near the peak load the inclined crack of series D beams
had a width in excess of 2 mm. Kotsovos considers that the test results are conflict with
the concept of shear capacity of critical sections, the view that aggregate interlock makes
a significant contribution to shear resistance, and even the truss analogy concept.
Chana [58] tested beams with traditional internal stirrups, but locally arranged as
shown in Figure 6.25. Kim et al. [74] tested beams with external stirrups. See Figure
6.26. The prestressed external stirrups were provided at the outer third sections of beams.
Muttoni [56] and Kuttab [75] tested beams with the arrangements of reinforcement shown
in Figure 6.27 and 6.28 respectively. All beams failed in ductile modes and reached failure
loads which are almost the full flexure capacities of the beams. The shear capacity of
these beams was about double the capacity of similar beams without the nonconventional
transverse reinforcement.
The test results from Kotsovos or the other researchers are very interesting, but not
that surprising. The common point of these tests was that the beams were provided
with a mechanism that gave effective constraint to the development and propagation of
flat (or horizontal) cracks in the critical zones, making the critical diagonal crack more
stable. In reality, the arrangements of reinforcement used by these researchers can be
considered as various modifications of the conventional form in which the stirrups are
placed throughout the whole beam and anchored in the compression zone.

6.8 Conclusions

Based on the study in this chapter, some conclusions are arrived at:

1. A distinct characteristic of slender beams without transverse reinforcement is the


lack of ability to control horizontal crack propagation and the brittleness of shear
failure results from the occurrence of either flat inclined cracks or horizontal cracks.
Chapter 6. Shear Failure of Beams Without Stirrups 118

2. The shear stresses in both the perfect bond case and the no bond case have similar
distributions along the cross sections in the B region of the beam. This suggests
that simply focusing on on the shear stress distributions is not adequate for the
investigation of shear failure mechanism. How a load transfers is much more im
portant than shear stress distributions.

3. As the entire cross section of a beam is subjected to compression stress and the
shear stress in the no bond case, no inclined cracks will be developed in the shear
span. This was illustrated in the experimental work by Kani [13].

4. Bond has an influence upon the transverse splitting of compression struts of very
deep beams (a/d < 1); however this influence can be ignored because the deteri
orated bond in deep beams with inclined cracks will reduce this influence signifi
cantly.

5. Generally, the shear stress distributions over the cross section of cracked beams
are not uniform. The localized shear stress distributions can be considered as the
direct result of arch action or vice versa.

6. The ideal direct load transfer mechanism or “pure arch action” can only be realized
in the no bond condition, regardless of the shear span ratio a/d. The form of the
direct compression strut between the point load and the support depends on the
resultant forces acting on the both ends of the strut. These resultant forces on
two ends of the strut should act along the same line and in opposite directions.
Consequently, the inclination of the strut originating from the support cannot be
arbitrarily chosen when the truss models are developed.

7. In addition to dowel action, which must be introduced due to the shear displacement
along the crack, the locally amplified bond stress at the zone near the critical crack
Chapter 6. Shear Failure of Beams Without Stirrups 119

contributes to the occurrence of the second diagonal crack and the splitting parallel
to the longitudinal reinforcement. This bond problem has also been recognized by
other researchers [76] but from a different point of view.
Chapter 6. Shear Failure of Beams Without Stirrups 120

h/4
— S.

‘I’
—-S.

— Distribution
/ I’ /

....7—-—j--------4.F
: /7
ot Transverse
Compression
hCT> 1 I,
I hi g h/2 , h/2 ._._— Isostatic
I i I i / Compressive
I I I I /


I ‘ I’
‘1’
5---,
‘ /
/
i

-r

F
- Curve

(a) a/h = 0 (b) a/h = 0.25 Cc) a/h = 0.5

a=h a=2h
— .-
, — .. ,
,
, ,
, 7
7

Z’’
,••
,
I
F
F
‘ /
.. —
‘- —

I
Cd) a/h 1 Ce) a/h = 2

Figure 6.1: Distribution of transverse compressive stress for various shear span ratios.
from Mau and Hsu, Ref.[52].
Chapter 6. Shear Failure of Beams Without Stirrups 121

F”
j
.‘

h
IS,,,
Lrc

f ‘I

, -*k--
/

/
I , \
, ,
I

j, z-v.

Figure 6.2: Load near the support: transition from deep beam to slender beam, from
Schlaich et al., right side simple models; left side refined models, Ref.[3].
Chapter 6. Shear Failure of Beams Without Stirrups 122

• 6 x 6 x 1 in, (152 x 152 25 mm) plate


69 6 *9 x 2 in, (152 x 229 x 51 mm) plate
o 25 -

o 6 3*0.38 in. (152* 76* 95mm)plate

•67 V V
ia
0.20

}m)

V 0.15 l 3940 psi (272 MPa)


bdf max. aag.- 3/4 in. (19mm)
d21.2in. (538mm)
b6.1in. (155mm)

2 (2277 mm
As 3.53 in )
2
0.10 -
• 72 f.=53.9 ksi (372 MPa)

081

65
.76
U.V.J 0
71 063 066

strut and tie model sectionaI model

I
0
0 1 2 3 4 5 6 7
aid

Figure 6.3: Predictions of shear strength versus a/d ratio for tests reported by Kani[13j,
from Collins and Mitchell, Ref.[39}.
Chapter 6. Shear Failure of Beams Without Stirrups 123

vJ
I
(a) Geometry and Loading

r iiV

1
d,

V4j
2
f J
Ib) Truss Mode!

Figure 6.4: Truss model developed by Adebar, Ref.[54}.

F

/ \Nc —
/
I’- _\_.
— 1L
fv

Figure 6.5: Truss model developed by Reineck, Ref.{55}.


Chapter 6. Shear Failure of Beams Without Stirrups 124

h
T

* *
O 450 450

Figure 6.6: Truss models developed by Al-Nahiawi and Wight, Ref.{53j.

Figure 6.7: Structural model developed by Muttoni and Schwartz, Ref.[56J.


Chapter 6. Shear Failure of Beams Without Stirrups 125

‘4

I I

Figure 6.8: Structural model developed by Kotsovos, Ref.[57J.

24j
Z/f/)f

Figure 6.9: Crack pattern of a beam tested by Kani, Ref.[13}.


Chapter 6. Shear Failure of Beams Without Stirrups 126

HHA

Figure 6.10: Geometric relationship of a crack at reinforcement level.


Chapter 6. Shear Failure of Beams Without Stirrups 127

Figure 6.11: A simply supported and central loaded beam.


ij
o b-’.

r-- - -.

-- -

o
cfIfjIL1LIw
CD CD

CD

C’,

C,
7
IUI[EW
z
0
I-.
c-l cD
CD
CD -S
0
Cl)
- .
cD
(I)
-uvi11J
-. .III
Cl)
0
Cl)
-‘
0 ±:
C
0 11]y- I
rrrIrJnrrr1a ZQfffl
I
-1
1IIP .111
I-’
L’3
I
Chapter 6. Shear Failure of Beams Without Stirrups 129

— — — — —
— — — -, ‘-

— — .— .- — — — — — — — — /
— — C. — — — C C c_ — —

1
— —
/ / ,/
—,
/
— —
/ 1 ,) / /
— — — —
K

.‘ /

/
7//il

- ///‘ / /7/ /
/ / / / 1’! / / / i’, / 7 1 /

---
I 7 / I I
----------
I I I I I I I
------
I I I I -

(a)

-----

- — — — — -, — -- —, — - H
— — — — — — — — — — — —
— — .- .- ... . _ _ —

, / — .-. — — — — — — ,, — — / /
- — t.t. c .. —

- / — 2 — — — •_
— — _• — / /
./ / — — —. — — — — — — — ,, /
/ ; /
- — — — C C — — — —

— .- .- — — — — — — ; ./.

(b)

Figure 6.13: Internal force flows in uncracked beams: (a) perfect bond case and (b) no
bond case.
Chapter 6. Shear Failure of Beams Without Stirrups 130

-1
-

- - - - —

———----—--q
-i - - - -
-

- - - - — — - — — - - — — — —

—- -
———- -

(a) (b)
- - ————--

-
————- - -——--
—1
- - - -

-
-
———- - - -——-
- - -—

- - -

LU LU
(c) (d)

Figure 6.14: Modelling of bond effect upon transverse splitting of compression struts in
deep beam.
Chapter 6. Shear Failure of Beams Without Stirrups 131

1T

ff/
/ 9.

/\/

\____

(a)
(b)
(c)
(d)

Figure 6.15: Bond influence upon transverse tensile stresses of compression struts in deep
beam.
Chapter 6. Shear Failure of Beams Without Stirrups 132

2 b

T 2

-
1
;5z:--- C+AC
jd b id
1
—I—-
___
T—’ T+AT T
— a

T+ AT
1
Ax

longitudinal equilibrium longitudinal shear


stresses stresses

Figure 6.16: Internal stress distributions of cracked beams due to pure beam action, from
Ref. [39].
Chapter 6. Shear Failure of Beams Without Stirrups 133

m m

(a)

pure arch action

(b)

(c) beam arch combined

Figure 6.17: Shear stress distributions of cracked beams due to combined beam action
and arch action: (a) modelling of combined beam and arch actions; (b) tension force in
longitudinal reinforcement; (c) shear stress distribution at section n n.

Chapter 6. Shear Failure of Beams Without Stirrups 134

//
7 7 /77

(a) (b)

Figure 6.18: Shear displacement along cracks in shear span: (a) vertical cracks and (b)
inclined cracks.
Chapter 6. Shear Failure of Beams Without Stirrups 135

V+dV

VtOV

V+dV

Figure 6.19: Shear transfer at cracks by aggregate interlock.


Chapter 6. Shear Failure of Beams Without Stirrups 136

second C

(a) lv n

jT
(b)

(C)
Cracked
Concrete

4.

rr.

NJ
Cracked
Steel

Figure 6.20: A beam with inclined cracks in shear span: (a) strut force due to arch
action; (b) one assumed tension force in reinforcement; (c) another assumed tension
force in reinforcement; (d) measured concrete strain and tension force in reinforcement
from Ref.[71].
Chapter 6. Shear Failure of Beams Without Stirrups 137

/
Measured \
steel tension
Calculated
teéI tension
Critical flexural I
crack section 4

(a)

•l

IRLI&ILP\JeL/
I.

I I

(b)

Figure 6.21: Measured force in bar, bond stress and crack locations, from Ref.[59].
Chapter 6. Shear Failure of Beams Without Stirrups 138

-4

(a)

(b)

Figure 6.22: A load transfer mechanism just before the occurrence of splitting along
longitudinal reinforcement: (a) truss model and (b) tension force in reinforcement.
JCD
CD O

c..

cJ-
CD
CD
a,
C a,
TI
CD II.
i ii • It
I II II
• I
Ii
I
1• I•
I II 1W I I
o
I-. I yb
III I,- I
I I II IC I
U’ I ‘I I
CD 0 W
0 0 UI
I 01W Iw
o I—.
C’) II..
0
I I
CD
II
i I
Iiii
cJ
I I II I
I I I W
CDI
i ‘ I 0

(ID
0
Cl) D C) W 0 0’
o
C
Cl)

•1____ ‘. ‘Tn
I. I!0
90 —i
I-’-
iooI 100
H
WII
’L
0 a
iiI

V
Chapter 6. Shear Failure of Beams Without Stirrups 140

a) a)
1

2 2

x
ci) —
— -\ZE
0 0S
C
- 0 06
0

00? 002-
0 /c5:— 0
• I I I I
c 5 10 15 20 25 2 1. 6 8 10 l2
.2 0

deflection mm -
deflection mm -

(a) (b)

ci) 12
I-

2 B

D
x

0
•0
Ca
.2
V
a)

0.2
0
-

0
V 0
0 2 L 5 8 10 12
.2
central deflection mm-

(c)

Figure 6.24: Load-deflection curves of beams shown in Figure 6.5: (a) a/d = 1.5; (b)
a/d = 3.3 and (c) a/d = 4.4, from Kotsovos, Ref.[73].
Chapter 6. Shear Failure of Beams Without Stirrups 141

35

30 Linknos
321

25 lW
Lnk 2/ Lnk I

20
Failure of span without
links at 9SkN
/ -

15
/k3

10.
1/ Failure load
=158kN

0 20 40 60 80 100 120 140 160


TOTAL SHEAR FORCE - kN

Figure 6.25: Beam with internal stirrups tested by Chana, from Ref.{58].

L
.
4 J —

L
1 25

-—r- f S I I PS

IO7 I .iLIPS 24.1 *IPS

II 4.
I
NUT
SECT ION A—A

Figure 6.26: Beam with external stirrups tested by Kim et al., from Ref.[74j.
Chapter 6. Shear Failure of Beams Without Stirrups 142

S ‘ —— — ,/.
/

(b)

Figure 6.27: Beams tested by Muttoni et aL, from Ref.[56].

Path of Compressive Force


/
n-i 11

Path of Compressive Force



I aid = 2
[I_ujH1_rH

.1
I I I I I I I aid = 3.6

Figure 6.28: Beams by Kuttab et al., from Ref.[75].


Chapter 7

Bond Splitting Failure

7.1 Introduction

The main objective of this chapter is to deal with the question of what causes the horizon
tal splitting along longitudinal reinforcement in a beam without transverse reinforcement.
It is generally believed that splitting of concrete along reinforcing bars in a beam occurs
primarily due to the combined effect of wedging action of bar deformations (bond) and
dowel action of reinforcement. Ferguson et al. [61, 62j focused on bond splitting. They
noticed that the existence of a diagonal crack resulted in lower bond strengths. Based on
an extensive experimental study, Cergely [77] concluded that the dowel force is the most
important factor producing splitting in beams without stirrups, and that the dowel effect
overshadows the pure bond effect in most situations, especially in beams with small bar
spacing. Jimenez et al. [78] found that bond strength and dowel capacity are independent
of each other. Kemp et al. [66] observed from their experimental work that there is a
weak interaction between dowel force and bond resistance until approximately 80 percent
of the pure dowel capacity is reached at which time bond capacity decreases very rapidly.
They suggested a 20 percent reduction in design ultimate bond to include the effect of
dowel action.
The conclusions arrived at in Chapter 6 about the load transfer mechanism prior to the
occurrence of splitting, are consistent with above mentioned experimental observations.
Dowel action and severity of bond stresses (i.e., amplification and localization) will both

143
Chapter 7. Bond Splitting Failure 144

be introduced due to the occurrence of the critical diagonal crack. Consequently, dowel
action and bond action both will make contributions to splitting along reinforcement. To
date no satisfactory quantitative analysis results are available to describe the interaction
of these two actions (bond and dowel). Traditionally bond and shear have been dealt
with separately so that previous investigations on bond splitting were mostly concerned
with ultimate bond splitting failure. However, the conclusions arrived at in Chapter
6 make it evident that the occurrence of horizontal splitting along reinforcement in a
critical zone indicates shear failure of beams without transverse reinforcement. Then it
seems that the initiation of bond splitting could be the more appropriate criterion than
the ultimate bond splitting strength for the study of shear capacity of beams without
transverse reinforcement.
In this chapter, the attention is focused on bond splitting with an objective to develop
a tentative design criterion for the initiation of longitudinal bond splitting. The chapter
is divided into three parts. In the first part, previous studies are briefly reviewed. This
includes general bond actions, experimental studies, as well as bond splitting strength
(ultimate and initiation). A proposed design equation for bond splitting initiation is
presented in the second part. Finally, some comments are given about Ferguson’s [61, 62]
experimental work and a number of conclusions are drawn.

7.2 General Bond Action

Studies of bonding forces for plain reinforcing bars and deformed bars by Lutz and
Gergely [79] showed that bond for plain bars is made up of three components: (1) chemical
adhesion, (2) friction and (3) mechanical interaction between concrete and steel. When
plain bars without surface deformations are used, bond depends mainly upon chemical
adhesion, and after slip, upon friction. There is also some negligible mechanical action
Chapter 7. Bond Splitting Failure 145

due to the roughness of the bar surface. With use of deformed bars, the main reliance is
changed to bearing of lugs on concrete and to shear strength of concrete sections between
lugs.
Because slip of deformed bars can occur in two ways namely: (1) the lugs can split
the concrete by wedging action and (2) the hugs can crush the concrete, two types of
bond failures can occur. If the surrounding concrete resistance is moderate, as it is
for ordinary concrete cover, the lugs of large steel bars can split the concrete without
crushing it. With small bars or with large cover over the bars, the lugs will shear the
concrete and pull out without splitting the concrete.
Bond between concrete and a deformed reinforcing bar that is subjected to a pull-out
force as well as is with a moderate concrete cover, can be characterized by the relationship
between averaged bond stress along the embedment length and slip at the loaded end,
with four different stages as shown in Figure 7.1 [80]. In Stage 1. (small values of the
bond stress), bond is assured by chemical adhesion, and no bar slip occurs. In stage 2
(larger bond stress values), the chemical adhesion breaks down and bonding is assured
by bearing action or wedging action of the bar lugs. In Stage 3 (still larger values of
bond stress), the first longitudinal cracks form as a result of the increasing wedge action
of lugs and more bond stress can be sustained by the interlock between concrete and
reinforcement lugs. Once the longitudinal cracks break out through the whole cover,
failure occurs abruptly in this stage if no transverse reinforcement is provided. If enough
transverse reinforcement is provided, the confinement exerted by the reinforcement would
allow the bond stress to reach a larger value in spite of concrete splitting. See Stage 4 in
Figure 7.1.
For bond related shear failures of structural concrete members without shear rein
forcement, the confinement provided by transverse reinforcement is not available. Hence
only Stage 3 described above is relevant to the problem to be dealt with. The beginning
Chapter 7. Bond Splitting Failure 146

and end of Stage 3 indicate splitting initiation and ultimate splitting failure. Both split
ting initiation and splitting failure are investigated in the following. But it is believed
that the bond related shear failure is more relevant to splitting initiation than to ultimate
splitting failure, as dowel action, which can aggravate the bond splitting problem, will
always be introduced due to the occurrence of diagonal cracks in beams.

7.3 Previous Experimental Studies

In the history of bond research, the problem most considered has been bar development
length d,
1 which is the embedment length necessary to assure that a bar can be stressed
to its yield strength without failing in bond.
In order to determine bond stresses or development length, a variety of test methods
have been used. Typically there are ordinary pull-out tests (also denoted as concentric
pull-out test), eccentric pull-out tests, full-beam tests and semi-beam tests (often called
stub-beam or cantilever tests).
In Figure 7.2 ordinary pull-out test specimens and eccentric pull-out test specimens
commonly used by researchers are shown. It was believed by Ferguson et al. [60] that the
ordinary pull-out test is not entirely realistic as the measure of bond strength in beams,
because it carries no shear on the splitting plane. In a beam, a short 1x length of beam, as
shown in Figure 6.16, transfers the change in bar tension by bond stress into a horizontal
shearing stress, a considerable part of which acts on the section through the level of bars.
Also the loaded end of a concentric pull-out specimen is in compression and is restrained
due to friction forces. Hence the concentric pull-out test gives higher bond strengths than
those expected in beams, especially where splitting is an important factor. The eccentric
pull-out test [Figure 7.2(b) and (c)] greatly improves the disadvantages mentioned for

the ordinary pull-out test and keeps the advantage of simplicity.


Chapter 7. Bond Splitting Failure 147

Full-beam tests are considered most reliable because the influences of both transverse
shear stresses and flexural tension cracks are included. One kind of full-beam test used at
the University of Texas [61, 62] is shown in Figure 7.3. In order to eliminate the influence
of support reaction upon bond strength, the development length of the bar is placed in
a negative moment region, where the L” shown in the figure is the development length
investigated. However the full-beam tests are very expensive test procedures and difficult
to generalize for different variables. Semi-beam specimens (or stub cantilever specimens),
have been utilized at Cornell University [77] and West Virginia University [66], as well
as in Japan [81]. See Figure 7.4. This type of specimen still provides a realistic strain
gradient through the depth of the specimen as in beams, but the cost is much lower than
that of full-beam specimen tests. In addition, it is quite versatile since ratios of bond,
shear and flexure can be easily varied from one specimen to another.
In summary, in order to get correct information about bond splitting in beams without
transverse reinforcement, full-beam test procedures are most preferred. Semi-beam tests
and eccentric pull-out tests are satisfactory substitutes for full-beam tests. The test
results from semi-beam specimens and eccentric pull-out specimens can be considered
applicable to real beams.

7.4 Previous Studies of Ultimate Splitting Failure

The generally accepted bond mechanism, when wedging action has been mobilized, is
that reinforcing bar force is transferred to surrounding concrete by inclined compressive
forces radiating out from lugs on reinforcing bars and making an angle with the bar
axis. The inclined compressive forces can be decomposed into radial and tangential
components. The radial components (bursting forces) are balanced by circumferential
tensile (ring) stress in the surrounding concrete. For the case of concern to this study,
Chapter 7. Bond Splitting Failure 148

longitudinal cracks (splitting) appear when the tensile rings, which are usually weakest
in the thinnest concrete cover protecting the reinforcement, are stressed to the tensile
strength. See Figure 7.5. Several factors can affect bond splitting strength. These are
concrete cover or clear spacing of bars, concrete tensile strength (which is related to the
compressive strength f), embedment length and bar diameter etc.
Orangun et al. [82] proposed an approach for determining bond strength or develop
ment length, that included all the variables mentioned above. The approach is based on a
physical bond model. The radial forces, generated between the lugs and the surrounding
concrete, can be regarded as water pressure acting against a thick—walled cylinder with
an inner diameter equal to the bar diameter and a thickness c that is the smaller of the
clear bottom or side cover cb or 1/2 the clear spacing c
8 between adjacent bars. See Figure
7.6. The capacity of the cylinder depends on the tensile strength of the concrete. With
cb greater than 3
c
/ 2, a horizontal split develops at the level of the bars and is termed a
“side split failure.” With 3
c
/ 2 greater than cb, a “face-and-side split failure” forms with
longitudinal cracking through the cover followed by splitting through the plane of the
bars. When c/2 is much greater than cb, a “V-notch failure” forms with longitudinal
splitting followed by inclined cracks that separate a V-shaped segment of cover from the
member. From the results of 62 beam tests, an equation of the bond stress was developed
by using a nonlinear regression analysis. That is

u 3.23c 53db
= 1.22 + + (7.1)
db
in which u is the ultimate bond strength; c is the smaller of c
8 or c; c is the smaller of
one half of clear spacing or side cover; cb is the concrete cover; d
6 is the bar diameter;
and id is the development length or splice length (all units are in psi and inches).
This equation was further modified by rounding the coefficients to obtain a somewhat
Chapter 7. Bond Splitting Failure 149

more conservative value for u, denoted as uj

(7.2)
d, d
1

Orangun et al. [82] compared the bond stresses calculated by Eq. 7.2 to test results
obtained from a total of nine studies (over 500 tests) of splice and development strength
for bars not confined by transverse reinforcement. The predicted strengths gave a close
match with the test results.
According to the recent work by Darwin et al. [83], the dimensionless equations Eqs.
7.1 and 7.2 for bond strength can better be expressed as the tension force in reinforcement
in terms of the same variables. They believed that bond force provides a better measure
of member response than bond stress, since bond strength can be considered as being a
structural property rather than a material property. The recommended equation is

= 3ld(c + 0.4db) + 200Ab (7.3)

where Ab is bar area and f is steel stress at ultimate bond strength.

7.5 Previous Studies of Splitting Initiation

It should be noted that the previous study results presented above are interesting but
only related to ultimate bond splitting strength. Equations 7.1 and 7.2 are valuable
for bond development design if shear and bond problems are dealt with separately. As
a matter of fact, shear and bond are always co-existing and interactive, therefore the
initiation of bond splitting is most concerned in this study. Unfortunately little work
has been carried out on bond splitting initiation in comparison with the work done on
ultimate bond splitting failure, especially experimental investigations.
Chapter 7. Bond Splitting Failure 150

Using the thick-walled pipe analogy, Tepfers [84] performed an analysis of bond crack
ing for the concrete in uncracked elastic state, plastic state and partly cracked elastic
state. See Figure 7.7. By setting the known maximum tensile hoop stress equal to ma
terial tensile strength and approximating the angle of inclined compression forces as 45
degrees, expressions to predict splitting bond stress were developed for above mentioned
different states. These expressions are:

i 12 u12
-r) )
+ 1)2 + ()2
7
ft — (

(7.5)
db

= (0.3 + 0.6-) (7.6)

where Equations 7.4, 7.5 and 7.6 are for the uncracked elastic state, uncracked plastic
state and partly cracked elastic state respectively, and tt is the bond stress indicating
crack initiation, c is the thickness of cover, db is the diameter of a deformed reinforcing
bar and f is concrete tensile strength.
Figure 7.8 compares Tepfers’ [84] predictions with his eccentric pull-out test results
and other test results [85]. It can be seen that the plastic prediction gives the upper
bound solution, while the partly cracked elastic prediction gives a lower bound to the
results.
Kemp and Wilhelm [66] conducted a linear regression analysis of their experimental
data from semi-beam tests (see Figure 7.4) and suggested the following equation:

= (2.64 + 2.37-) (7.7)


Chapter 7. Bond Splitting Failure 151

where c is concrete cover and db is diameter of test bars, all units in psi and inches. A
comparison of Equation 7.7 and test data on cracking loads is shown in Figure 7.9.

7.6 Proposed Design Equation for Bond Splitting Initiation

There is a significant difference between the equations for bond cracking and ultimate
bond splitting strength. The embedment length d
1 of reinforcing bars is not included in
the Equations 7.4—7.7 for initial bond cracking in comparison with the Equations 7.1 and
7.2 for ultimate bond strength. The uncertainties about the validity of Equations 7.4—7.7
to predict bond splitting initiation strength in real beams exist due to lack of enough
test data. The ld/d ratios of the specimens used by Kemp et al. and Tepfers were 11.34
and 3.13 respectively. Because the bond stresses at which first concrete crack was visible
were averaged along the embedment lengths in the studies of Kemp et al. and Tepfers,
none of the equations developed by them can be readily applied to different beams with
different beam sizes.
Teng and Ye [86] have carried out a series of concentric pull-out tests to study bond
and slip relationships for deformed reinforcing bars. Both ultimate bond splitting loads
and bond splitting initiation loads were recorded during testing. Based on the statistical
regression analysis of test data, the following design equations were developed:

= (1.106 + 1.3)- (7.8)

= (1.162 + 1.8O2)- (7.9)

where u, is bond cracking stresses and u is ultimate bond splitting strengths, all units
in kg/cm
2 and cm. These two Equations 7.8 and 7.9 are shown graphically in Figure
7.10. It can be seen that the ultimate bond strength and cracking bond strength have
Chapter 7. Bond Splitting Failure 152

a similar trend. They both are a function of concrete cover, bar size, concrete strength
as well as embedment length. Although the two equations cannot be directly applied to
beams, as discussed in Section 7.2, the evidence revealed by the two equations can be
used as a basic assumption to develop an empirical equation for the prediction of bond
splitting initiation strength in beams.
The following equation is proposed for the bond splitting initiation strength

——=1.2+3--+15--. (7.10)
d
1

where all units of stress are in psi. This equation is actually a simple modification of
Equation 7.2, whose accuracy of predictions for ultimate bond splitting strength was
recently confirmed by Darwin et al. [83]. Equations 7.2 and 7.10, as well as the experi
mental results measured by Kemp and Wilhelm [66], are shown in Figure 7.11. Equation
7.10 results from Equation 7.2 by shifting a parallel displacement downwards for the
tested idid ratio. The comparison of Equation 7.10 and experimental results measured
by Tepfers [84] is shown in Figure 7.12. The close agreement of the predictions with test
results is illustrated. Figure 7.13 demonstrates Equations 7.2 and 7.10 for the c/d& of
0.5.
Equations 7.10 can also be modified to express bar force at cracking normalized with
respect to as recommended by Darwin [83]. That equation is

= 1
7
3
d (c
r + 0.4db) + 6OAb (7.11)

where Ab is bar area and 3


f is steel stress at cracking.
Chapter 7. Bond Splitting Failure 153

7.7 Some Comments and Conclusions

From Figure 7.13, it can be seen that the bond splitting initiation strength is much lower
than ultimate bond splitting strength. The combination of this lower bond cracking
strength and dowel action can lead to lower shear strengths as well as lower ultimate
bond splitting strengths.
The University of Texas beam [61] is shown in Figure 7.3 with a potential diagonal
crack within development length L”. In experimental studies carried out by Ferguson et
al. [61, 62], diagonal cracks always developed in narrow beams. These diagonal cracks
hastened the bond splitting process, particularly with longer L” values, and led to lower
bond strengths (splitting along whole development length). Sometimes, diagonal cracks
dominated failures in shear with the ends of the bars still fully bonded near the inflection
point. In both cases, the occurrence of diagonal cracks led to lower shear capacities.
Based on the study in Chapter 6, after the occurrence of a diagonal crack the variation
of tension force in a reinforcing bar may be shown by the solid line in Figure 6.20(b). As
discussed before, dowel action will be introduced and bond stress will be locally amplified
due to the development of a diagonal crack. Based on the study in this chapter, it can
be seen that the bond problem is further aggravated by the reduced development length
<L” by referring to Equation 7.3 and lower allowable tension force in reinforcement by
the comparison of Equation 7.3 and Equation 7.11. See Figure 7.3. Then lower shear
capacity or lower bond splitting strength can be expected. When wider beams are used
for testing, the unfavourable factors for bond splitting strength will be eliminated with
the disappearance of a critical diagonal crack. The higher shear capacities and the higher
bond splitting strengths can be expected.
Another issue is about whether dowel force or bond force dominates splitting along
reinforcing bars in beams without transverse reinforcement. If a single bar is put in a wide
Chapter 7. Bond Splitting Failure 154

beam (small dowel force and large dowel capacity), bond splitting could be prominent as
the low steel percentage and the large steel stresses cause splitting on the bottom face,
Ref. [62]. If a couple of bars are put in the same dimension beam with small spacing in one
layer (large dowel force and little dowel capacity), dowel action could dominate splitting
as the splitting occurs on the sides under comparable load, Ref.[77]. Both situations,
experimentally indicated, lead to lower shear capacities.
The shear resistance mechanism and the bond strengths between concrete and longi
tudinal reinforcing bars are very complicated in a reinforced beam without stirrups. The
interaction of shear and bond seems much more complicated. Another example showing
this complexity is Hall’s discussion [67] of the experimental work carried out by Baant
and Kazemi [87] for the investigation of size effect on diagonal shear failure of beams
without stirrups. Hall considered that the observed results are due to the location and
distribution of the reinforcement, which implies bond failure.
In summary, the study in this chapter has led to a tentative equation for bond splitting
initiation and an improved understanding of the interaction of shear and bond in a beam
without transverse reinforcement. Further work on this topic is necessary, however, the
problem is very complicated and many factors are involved. It is not possible to complete
this topic as part of the present study. Suggestions for further study are given in the
next chapter.
Chapter 7. Bond Splitting Failure 155

Cl)
C’)
ci)
Cl)
-o
C
0
co
inadequate confinement

Bar Slip

Figure 7.1: Bond stress—slip relationship, from Gambarova et aL, Ref.[80].


Chapter 7. Bond Splitting Failure 156

Reaction
Each Side

Reacrion

(a)

(c)
(b)

Figure 7.2: Pullout tests: (a) concentric pullout-test specimen; (b) commonly used ec
centric pullout-test specimen and (c) eccentric pullout-test specimen used by Ferguson,
Ref. [60].
Chapter 7. Bond Splitting Failure 157

(a)

iF
IIIIIIIII I I I Potential inclined crack
1
(b)

L” P.’.

(c)

(d)

Figure 7.3: The University of Texas beam tests: (a) possible tension force distribution
along the top test bar after the occurrence of inclined crack; (b) side view of specimen;
(c) plain view of top reinforcement; and (d) moment diagram, adapted from Ferguson
and Thompson, Ref.[61].
CD

(I)

CJ(
cc
Chapter 7. Bond Splitting Failure 159

ZzrzE
Figure 7.5: Mechanism representation for bond, from Tepfers, Ref.[84].
Chapter 7. Bond Splitting Failure 160

Failure plane

C
>
6 C/2. C=C,/2

Side split failure


I? Just before failure
C
>z
I
I
- -
-

At failure I
I

V-Notch failure
LzFace-and-side split failure
>Cb
C,/
>
2 C,/2>Cb

Figure 7.6: Bond Splitting Failure Patterns, from Orangun at aL, Ref.[82].
Chapter 7. Bond Splitting Failure 161

1
a / Or+dr

(a) (b)

edb/2

(c) (d)

Figure 7.7: Analysis of Bond Splitting Stresses: (a) bursting and bond stresses; (b)
uncracked elastic state; (c) uncracked plastic state and (d) partly cracked elastic state.
Adapted from Tepfers, Ref.[84j.
Chapter 7. Bond Splitting Failure 162

6 .-—--________

0
5—
0
0
0

/ •

4:
/ . ..
Icbc

0c
0
00

£ 0
I Ct o •
e
0.%
•. •
2 -

x °

x
- elastic stage (equabon 5)
1

X • ordinary concrete

0 Iightweght concrete

I I I
0 1 2 3 4 5 6

c/d

Figure 7.8: Comparison of test and prediction for bond splitting initiation, from Tepfers,
Ref. [84].
Chapter 7. Bond Splitting Failure 163

15
uc

00 05
4 5

Figure 7.9: Comparison of Equation 7.7 with test data on cracking loads carried out by
Kemp and Wilhelm Ref. [66].
Chapter 7. Bond Splitting Failure 164

fl 6 Ultimate Strength (Eq. 7.9)


— — —
— Cracking Strength (Eq. 7.8)

2=20
db
2- c
—=1.5
db

• db

db
I
0
0 5 10 15 20
Id
db

Figure 7.10: The relationship of bond splitting strength and ld/d proposed by Teng and
Ye Ref.[86] for concentric pullout tests.
Chapter 7. Bond Splitting Failure 165

_!_ . Cracking
o Ultimate (
15

- Equation 7.2
10

:osFE7b0T

Figure 7.11: Comparison of Equation 7.2 and 7.10 with test results from Kemp and
Wilhelm Ref.[66].
Chapter 7. Bond Splitting Failure 166

uc
7r • Cracking

15

. .

Equation 7.10

(---=3.13)

I I
0
0 0.5 1 1.5 2
C
db

Figure 7.12: Comparison of Equation 7.10 with test results from Tepfers Ref.[84j.
Chapter 7. Bond Splitting Failure 167

U
25
1
fl
Ultimate Strength

Cracking Strength
20 -

15 -

Equation 7.2

10 -

=0.5
db
5-

Equation 7.10
I I I I
0
0 5 10 15 20

db

Figure 7.13: Relationship of bond strength versus ld/db, presented by Eqs. 7.2 and 7.10.
Chapter 8

Brief Summary and Further Research

This study, on the shear design of structural concrete members without transverse rein
forcement, includes three main topics: transverse splitting of compression struts, devel
opment of a rational design procedure for deep pile caps, as well as a more general study
of the load transfer mechanism in concrete beams without stirrups.
For deep pile caps, the compression strut that transmits column load to a pile is
usually unreinforced and confined by surrounding plain concrete. The compression in a
strut will spread out thereby introducing transverse tension near mid-height of the strut
due to strain compatibility. As there is no reinforcement provided to resist this transverse
tension, the concrete tensile strength must be mobilized. A refined truss model includes
a concrete tension tie to model the transverse tension. It is believed that the brittle shear
failure of deep pile caps is initiated by internal cracking due to this transverse tension.
That is, the shear failure of a deep pile cap results from transverse splitting rather than
crushing of compression struts.
In this study, the compression struts have been idealized as concrete cylinders of
various diameter D, and height H, subjected to concentric axial compression over a
constant size circular bearing area of diameter d. Linear elastic finite elements were used
to determine the triaxial stresses at first cracking within cylinders (Chapter 2). The
numerical study has indicated that the bearing stress at first cracking within cylinders
depends on the amount of confinement, the aspect ratio (height/width), as well as the
ratio of concrete compressive strength to concrete tensile strength. In order to confirm the

168
Chapter 8. Brief Summary and Further Researth 169

transverse splitting phenomenon within compression struts a series of experiments were


conducted on large size concrete cylinders (Chapter 3). A good correlation was found
between the analytical prediction and the experimental results regarding the influence of
D/d and H/d on the bearing stress to cause first cracking.
Based on the results of the analytical and experimental studies, a bearing stress limit
was proposed in terms of the amount of confinement and the aspect ratio (height/width)
of the compression strut, as well as concrete strength (Chapter 4). The proposed bearing
stress limit is given for the maximum nodal zone bearing stress to prevent diagonal tension
(shear) failures in deep pile caps with unreinforced compression struts. In contrast,
the ACT Building Code bearing stress limit is intended to prevent crushing of concrete
in nodal zones and does not preclude a shear failure due to transverse splitting of a
compression strut.
By incorporating the proposed bearing stress limit into a strut-and-tie model that
emphasizes internal force flow rather than the “shear stress” on any prescribed section,
two rational design methods for deep pile caps were proposed (Chapter 5). The two
methods are similar except for the details of how the bearing stress limit is applied.
The first design method is a direct extension of the two dimensional strut-and-tie model
for deep beams (CSA approach). The procedure involves defining an equilibrium force
system in a deep pile cap. The forces in the compression struts are calculated from the
proposed bearing stress limit. The horizontal components of the compression strut forces
must be equilibrated by tension forces in the provided reinforcement. The sum of the
vertical components of the compression strut forces gives the designed load capacity of
the deep pile cap.
The second design method is presented in a more traditional way, in which “flexural
design” and “shear design” are separated. A truss model is used for “flexural design”
(i.e., to calculate required longitudinal reinforcement). “Shear design” is accomplished
Chapter 8. Brief Summary and Further Research 170

by limiting the maximum bearing stress (between pile and cap or column and cap) below
the proposed bearing stress limit. Although a similar force flow is implied in the second
method as in the first, the details of the strut geometry are not needed in the more
simplified second method.

Comparison of predictions from the proposed design method with predictions from
ACT Code procedure and CRST Handbook procedure for 48 previously tested pile cap
specimens demonstrated that the proposed method is more rational and more accurate
than what is presently used to design deep pile caps (Chapter 5). Tn deep pile caps
the shear stress is concentrated in zones (compression struts) between the column and
piles, and is not uniform over the height making it difficult to calculate a meaningful
average shear stress. The sectional methods of the ACT Code and CRST Handbook are
not appropriate for the shear design of deep pile caps. For example, the one-way shear
design provisions of the 1983 ACT Building Code (and subsequent edition) are excessively
conservative for deep pile caps.
It was found that the traditional ACT Building Code flexural design procedures are
unconservative for deep pile caps. These flexural strength procedures are meant for lightly
reinforced beams which are able to undergo extensive flexural deformations (increased
curvature) after the reinforcement yields. Deep pile caps are large blocks of plain concrete
which cannot undergo significant flexural deformations without triggering a brittle shear
failure.
In the third part of this study, the load transfer mechanisms of beams, which unlike
pile caps transmit the load in one direction, without stirrups were investigated. Addi
tional considerations herein are the crack propagation of discrete diagonal cracks and the
influence of bond between concrete and longitudinal reinforcing bars.
Based on a study of the compatibility of the displacements of a critical inclined crack,
an interpretation of an important shear failure mechanism of slender beams without
Chapter 8. Brief Summary and Further Research 171

stirrups is presented (Chapter 6). The interpretation suggests it is the occurrence of


either very flat or horizontal cracks near the critical inclined crack that is most indicative
of shear failure in beams. In order to get a better understanding of this interpretation,
the influence of bond upon the load carrying mechanism of the beam was studied.
In this study, two mechanisms of shear resistance in structural concrete members
without stirrups have been identified. In a very deep member (a/d < 1) or somewhat
more slender member with no bond between concrete and reinforcement, loads are trans
mitted directly to supports by compression struts [see Figure 8.1(a)]. The shear failure
of such members is characterized by transverse splitting of compression struts. For more
slender members with normal bond between concrete and longitudinal reinforcement, the
load transfer mechanism is as shown in Figure 8.1(b). A concrete tension tie is needed to
transfer load to the support. In this case, the capacity of the concrete tension tie relies
upon the bond strength between concrete and reinforcement.
An empirical equation for the strength of bond splitting initiation has been devel
oped based on previous experimental results (Chapter 7). However, the shear resistance
mechanisms in slender beams are very complicated, involving both shear and bond. A
design procedure which can be used by practising engineers cannot be developed in this
pilot study which is only a small part of this thesis. Additional concentrated research
is necessary. For example, more analytical and experimental research is required on the
bond splitting initiation in beams without stirrups. The interaction of dowel action and
bond splitting to cause cracking along reinforcement should be studied further both ex
perimentally and analytically with the objective to develop a quantitative relationship
of this interaction. Based on a better understanding of the splitting phenomenon along
longitudinal reinforcing bars, a rational design procedure for beams without stirrups can
be developed, in which bond splitting is avoided by limiting the maximum tension force
in longitudinal reinforcement.
Chapter 8. Brief Summary and Further Research 172

H
**
F__I \

(a)
$1

(b)

Figure 8.1: Load transfer mechanisms of structural concrete members without transverse
reinforcement: (a) very deep members (a/d < 1) as well as more slender members with
no bond between concrete and reinforcement; (b) slender members with normal bond
between concrete and reinforcement.
Bibliography

[1] Marti, P., “Basic Tools of Reinforced Concrete Beam Design,” ACT Journal,
Proceedings, V. 82, No. 1, Jan.-Feb. 1985, pp. 46-56.

[2] Collins, Michael P., and Mitchell, Denis, “Rational Approach to Shear Design-
The 1984 Canadian Code Provisions,” ACT Journal, Proceedings, V. 83, No. 6,
Nov.-Dec. 1986, pp. 925-933.

[3] Schlaich, J., Schafer, K., and Jennewein, M., “Towards a Consistent Design of
Structural Concrete,” PCI Journal, V. 32, No. 3, May-June, 1987, pp. 74-150.

[4] Ritter, W., “Die Bauweise Hennebique,” Schweizerische Bauzeitung (Zurich), V.


33, No. 7, Feb. 1899, pp. 59-61.

[5] MacGregor, James, “Dimensioning and Detailing,” IABSE Colloquium, “Struc


tural Concrete,” Stuttgart, April 1991, pp. 391-409.

[6] Ramirez, J.A., and Breen, J.E., “Evaluation of a Modified Truss-Model Approach
for Beams in Shear,” ACI Structural Journal, V. 88, No. 5, September-October
1991, pp. 562-571.

[7] Reineck, K.H., and Hardjasaputra, H., “Consideration of Strains in Shear Design of
Reinforced Concrete and Prestressed Concrete Beams,” (in German), Bauingenieur,
Vol. 65, February 1990, pp. 73-82.

[8] Ramirez, J.A., “Truss Model Approaches for Shear Design in Beams,” State-of
the-Art Report, ASCE-ACT Committee 445 Shear and Torsion,

173
Bibliography 174

[9] ACT Committee 318, “Building Code Requirements for Reinforced Concrete and
Commentary (ACT 318-89/ACT 318 R-89),” American Concrete Institute, Detroit,
1989, 353 pp.

[10] “Design of Concrete Structures for Buildings,” (CAN3 A23.3-M84), Canadian Stan
dards Association, Rexdale, 1984, 281 pp.

[11] Adebar, P., Kuchma, D., and Collins, M.P., “Strut-and-Tie Models for the Design
of Pile Caps: An Experimental Study,” ACI Structural Journal, V. 87, No. 1,
Jan.-Feb. 1990, pp. 81-92.

[12] ACT-ASCE Committee 326, “Shear and Diagonal Tension,” ACT Journal,
proceedings V. 59, Jan. and Feb., 1962, pp. 1-30, 277-333, and 353-395.

[13] Kani, M.W., Huggins, M.W., and Wittkopp, R.R., “Kani on Shear in Reinforced
Concrete,” Department of Civil Engineering, University of Toronto, 1979, 225 pp.

[14] Ferguson, P.M., Breen, J.E., and Jirsa, J.O., “Reinforced Concrete Fundamena
tals,” 5th Edition, John Wiley & Sons, New York, 1988, 746 pp.

[15] Cook, RD., Malkus, D.S., and Plesha, M.E., “Concepts and Applications of Finite
Element Analysis,” Third Edition, John Wiley & Sons, New York, 1989, 630 pp.

[16] “PUNDIT Manual for Use with the Portable Ultrasonic Non-Destructive Digital
Indicating Tester,” C.N.S. Instruments Ltd, 61-63 Holmes Road, London, NW5,
42 pp.

[17] Niyogi, Sanat K., “Bearing Strength of Concrete - Geometric Variations,” Journal
of Structural Division, Proceedings, ASCE, V. 99, No. ST7, July 1973, pp. 1471-
1489.
Bibliography 175

[18] Meyerhof, G.G., “The bearing capacity of concrete and rock,” Magazine of Concrete
Research, Vol. 4, No. 12, April 1953, pp. 107-116.

[19] Shelson, W., “Bearing capacity of concrete,” ACT Journal, Proceedings Vol. 54, No.

5, Nov. 1957, pp. 405-414.

[20] Au, T., and Baird, D.L., “Bearing capacity of concrete blocks,” ACT Journal, Pro
ceedings Vol. 56, No. 9, March 1960, pp. 869-879.

[21] Hawkins, N.M., “Discussion of reference 6,” ACT Journal, Proceedings Vol. 57, No.
2, Sept. 1960, pp. 1469-1479.

[22] Campbell-Allen, D., “Discussion of reference 2,” ACT Journal, Proceedings Vol. 57,
No. 12, June 1958, pp. 1185-1187.

[23] Douglas, D.J., and Trahair, N.S., “An examination of the stresses in the anchor
age zone of a post-tensioned prestressed concrete beam,” Magazine of Concrete

Research, Vol. 12, No. 34, March 1960, pp. 9-18.

[24] Middendorf, K.H., “Practical aspects of end zone bearing of post-tensioning ten
dons,” PCI Journal, Vol. 8, No. 4, August 1963, pp. 57-62.

[25] Zielinski, J., and Rowe, R.E., “An investigation of the stress distribution in the an
chorage zone of post-tensioned concrete members,” London, Cement and Concrete
Association, Sept. 1960, Research Report 9, 32 pp.

[26] Hawkins, N.M., “Bearing Strength of Concrete Loaded through Rigid Plates,” Mag
azine of Concrete Research (London), V. 20, No. 62, March 1968, pp. 31-40.

[27] Chen, W.F., and Trumbauer, WE., “Double-Punch Test for Tensile Strength of
Concrete,” Journal of Materials, V. 7, No. 2, June 1972, pp. 148-154.
Bibliography 176

[28] Marti, Peter, “Size Effect in Double-Punch Tests on Concrete Cylinders,” ACI
Materials Journal, V. 86, No. 6, Nov.-Dec. 1989, PP. 597-601.

[29] Baant, Z.P., “Size Effect in Blunt Rracture: Concrete Rock and Metal,” Journal
of Engineering Mechanics, ASCE, V. 110, No. 4, April 1984, pp. 518-535.

[30] Baant, Z.P., Kim, J.-K., and Pfeiffer, P.A., “Nonlinear Fracture Properties from
Size Effect Tests,” Journal of Structural Engineering, ASCE, V. 112, No. 2, Feb.
1986, pp. 289-307.

[31] Yan, H.T., “Bloom base allowable in the design of pile caps,” Civil Engineering
and Public Works Review, Vol. 49, No. 575, May 1954, pp. 493-495. No. 576, June
1954, pp. 622-623.

[32] Blévot, J., and Frémy, R., “Semelles sur pieux,” Annales, Institut Technique du
Bâtiment et des Travaux Publics(Paris), Vol. 20, No. 230, February 1967, pp. 223-
295.

[33] Adebar, P., “The Behaviour of Pile Caps: An Experimental Investigation,” MASc
thesis, Department of Civil Engineering, University of Toronto, April 1989, 137 pp.

[34] ACT Committee 318, “Building Code Requirements for Reinforced Concrete (ACT
318-77),” American Concrete Institute, Detroit, 1977, 102 pp.

[35] ACT Committee 318, “Building Code Requirements for Reinforced Concrete (ACT
318-83),” American Concrete Institute, Detroit, 1983, 111 pp.

[36] Rice, P.F., and Hoffman, E.S., “Pile Caps — Theory, Code, and Practice Gaps,”
Structural Bulletin No.2, Concrete Reinforcing Steel Institute, Chicago, Feb. 1978,
14 Pp.
Bibliography 177

[37] CRSI Handbook, Concrete Reinforcing Steel Institute, Chicago, 1984, 800 pp.

[38] ACI-ASCE Committee 426, “Shear Strength of Reinforced Concrete Members,”


proceedings, ASCE, V. 99, ST6, June 1973, pp. 1091-1188.

[39] Collins, M.P., and Mitchell, D., “Prestressed Concrete Structures,” Prentice Hall,
Englewood Cliffs, 1990, 500 pp.

[40] Vecchio, F.J., and Collins, M.P., “The Response of Reinforced Concrete to In-
Plane Shear and Normal Stresses,” University of Toronto, Department of Civil
Engineering, Publication No. 82-03, March 1982.

[41] “Concrete Design Handbook,” Canadian Portland Cement Association, 1985.

[42] Hobbs, N.B., and Stein, P., “An Investigation into the Stress Distribution in Pile
Caps with Some Notes on Design,” Proceedings of the Institution of Civil Engineers,
Vol. 7, July 1957, pp. 599-628.

[43] Deutsch, G.P., and Walker, D.N.O., “Pile Caps,” Fourth Year Civil Engineering
Research Project, University of Melbourne, 1963, 75 pp.

[44] Clarke, J.L., “Behaviour and Design of Pile Caps with Four Piles,” Cement and
Concrete Association, London, Report No. 42.489, November 1973, 19 pp.

[45] Clarke, J.L., and Taylor, H.P.J., “Model Tests to Determine the Influence of Sup
port Stiffness Upon the Distribution of Pile Loads on an Eight-Pile Cap,” Magazine
of Concrete Research, Vol. 26, No. 86, March 1974, pp. 39-46.

[46] Gogate, A.B., and Sabnis, G.M., “Design of Thick Pile Caps,” ACI Journal, pro
ceedings, Vol. 77, No. 1, Jan.-Feb. 1980, pp. 18-22.
Bibliography 178

[47] Sabnis, G.M., and Gogate, A.B., “Investigation of Thick Slab (Pile Cap) Be
haviour,” ACT Journal, Proceedings, Vol. 81, No. 1, Jan.-Feb. 1984, pp. 35-39.

[48] ACT-AS CE Committee 326, “Shear and Diagonal Tension,” ACT Journal,
proceedings V. 59, Jan. and Feb., 1962, pp. 1-30, 277-333, and 353-395.

[49] Bresler, B., and MacGregor, J.G., “Review of Concrete Beams Failing in Shear,”
proceedings, ASCE, V. 93, ST1, Feb. 1967, pp. 343-372.

[50] Wairaven, J.C., “Shear in Elements without Shear Reinforcement,” Bulletin


d’Information 146, Comité Euro-Tnternational du Béton, Paris, Jan. 1982, pp. 9-41.

[51] Reineck, K.-H., “Models for the Design of Reinforced and Prestressed Concrete
Members,” Bulletin d’Information 146, Comité Euro-International du Béton, Paris,
Jan. 1982, pp. 43-96.

[52] Mau, S.T., and Hsu, T.T.C., “A formula for the shear strength of deep beams,”
ACI Structural Journal, V. 86, No. 5, 1989, pp.

[53] Al-Nahiawi, K.A., Wight, J.K., “Beam Analysis Using Concrete Tensile Strength
in Truss Models,” ACT Structural Journal, May-June 1992, pp. 284-289.

[54] Adebar, P., Discussion of Paper by Peter Marti, “Design of Concrete Slabs for
Transverse Shear,” ACI Structural Journal, September-October 1993.

[55] Reineck, K.-H., “Ultimate Shear Force of Structural Concrete Members without
Transverse Reinforcement Derived from a Mechanical Model,” ACT Structural Jour
nal, V. 88, No. 5, Sept.-Oct. 1991, pp. 592-602.

[56] Muttoni, A., and Schwartz, J., “Behaviour of Beams and Punching in Slabs without
Shear Reinforcement,” TABSE Colloquium Stuttgart 1991, IABSE Report, V. 62,
Bibliography 179

Zurich, 1991, pp. 703-708.

[57] Kotsovos, M.D., “Compressive Force Path Concept: Basis for Reinforced Concrete
Ultimate Limit State Design,” ACI Structural Journal, V. 85, No. 1, Jan.-Feb.
1988, pp. 68-75.

[58] Chana, P.S., “Investigation of the Mechanism of Shear Failure of Reinforced Con
crete Beams,” Magazine of Concrete Research (Wexham Springs), V. 39, Dec. 1987,

pp. 196-204.

[59] Mains, R.M., “Measurement of the Distribution of Tensile and Bond Stresses Along
Reinforcing Bars,” ACT Journal, Proceedings V. 48, No. 3, November 1951, pp. 225-
252.

[60] Ferguson, P.M., Turrin, R.D., and Thompson, J.N., “Minimum Bar Spacing as a
Function of Bond and Shear Strength,” ACT Journal, June 1954, pp. 869-887.

[61] Ferguson P.R., and Thompson J.N., “Development Length of High Strength Re
inforcing Bars in Bond,” ACT Journal, Proceedings V. 59, No. 7, July 1962, pp.
887-922.

[62] Ferguson P.R., and Thompson J.N., “Development Length for Large High Strength
Reinforcing Bars,” ACT Journal, Proceedings V. 62, No.1, Jan. 1965, .
94
-
71
pp.

[63] ACT Committee 408, “Bond Stress—The State of the Art,” ACT Journal,
Proceedings V.63, No.11, November 1966, pp. 1161-1188.

[64] ACT Committee 408, “Opportunities in Bond Research,” ACT Journal, proceedings
V. 67, No. 11, November 1970, pp. 857-867.
Bibliography 180

[65] Ferguson, P.M., “Small Bar Spacing or Cover—A Bond Problem for the Designer,”
ACT Journal, September 1977, pp. 435-439.

[66] Kemp, E.L., and Wilhelm, W.J., “Investigation of the Parameters Influencing Bond
Cracking,” ACT Journal, January 1979, pp. 47-71.

[67] Hall, R.T., discussion of “Size Effect on Diagonal Shear Failure of Beams without
Stirrups,” by Bazant, Z.P., and Kazemi, M.T., ACI Structural Journal, March-
April 1992, pp. 211-212.

[68] Park, P., and Paulay, T., “Reinforced Concrete Structures,” John Wiley & Sons,
New York, 1975.

[69] Morsch, E., “Concrete Steel Construction,” English Translation E.P. Goodrich,
McGraw-Hill, New York, from 3rd edn of Der Eiseribetonbau (1st edition 1902).

[70] Fenwick, R.C., and Paulay, T., “Mechanisms of Shear Resistance of Concrete
Beams,” JournaloftheStructuralDivision, ASCE, V. 94, ST1O, Oct. 1968, pp.
2325-2350.

[71] dePaiva, H.A.R., and Siess, C.P., “Strength and Behaviour of Deep Beams in
Shear,” Journal of the Structural Division, proceedings of the ASCE, ST5, Oc
tober 1965, pp. 19-41.

[72] Watstein, D., and Mathey, R.G., “Strains in Beams Having Diagonal Cracks,” ACT
Journal, proceedings V. 55, December 1958, pp. 717-728.

[73] Kotsovos, M.D., “Strength and Behaviour of Deep Beams,” Reinforced Concrete
Deep Beams, F.K. Kong, ed., Blackie and Son, London/Van Nostrand Reinhold,
New York, 1990, pp. 21-59.
Bibliography 181

[74] Kim, W., Prak, B.K., and White, R.N., “Fracture Analysis of Shear Cracking
in Reinforced Concrete Beams,” Proceedings of the International Conference on
Structural Engineering and Computation, 25-28 April 1990, Beijing, China.

[75] Kuttab, A.S., and Haldane, D., “Detailing for Shear with the Compressive Force
Path Concept,” TABSE Colloquium Stuttgart 1991, IABSE Report, V. 62, Zurich,
1991, pp. 661-666.

[76] Kim, W., and White, R.N., “Initiation of Shear Cracking in Reinforced Concrete
Beams with No Web Reinforcement,” ACT Structural Journal, V. 88, No. 3, May-
June 1991, pp. 301-308.

[77] Cergely, Peter, “Splitting Cracks Along the Main Reinforcement in Concrete Mem
bers,” Report to Bureau of Public Roads, U.S. Department of Transportation,
Cornell University, April 1969.

[78] Jimenez, R., White, R.N., and Gergely, P., “Bond and Dowel Capacities of Rein
forced Concrete,” ACT Journal, January 1979, pp. 73-92.

[79] Lutz, L. and Gergeley, P., “Mechanics of bond and slip of deformed bars in con
crete,” Journal of the American Concrete Institute, Proceedings Vol. 64, No.11,
November 1967, pp. 711- 721.

[80] Gambarova, P.G., Rosati, G.P., and Zasso, B., “Steel-to-concrete bond after con
crete splitting: test results,” Materials and Structures, 1989, 22, pp.
.
4
-
357

[81] Morita, S., and Fujii, S., “Bond capacity of deformed bars due to splitting of
surrounding concrete,” Bond in Concrete, edited by P. Bartos (Applied Science,
London, 1982), pp. 331-341.
Bibliography 182

[82] Orangun, C. 0., Jirsa, J. 0., and Breen, J. E., “A Reevaluation of Test Data on
Development Length and Splices,” ACT Journal, Proceedings V. 74, No. 3, March
1977, pp. 114- 122.

[83] Darwin, D., McCabe, S.L., Idun, E.K., and Schoenekase, S.P., “Development
Length Criteria: Bars Not Confined by Transverse Reinforcement,” ACT Struc
tural Journal, V. 89, No. 6, November-December 1992, pp. 709-720.

[84] Tepfers, Ralejs, “Cracking of concrete cover along anchored deformed reinforcing
bars,” Magazine of Concrete Research, Vol. 31, No. 106, March 1979, pp. 3-12.

[85] Tilantera, T., and Rechardt, T., “Bond of reinforcement in light-weight aggregate
concrete,” Otaniemi, Helsinki University of Technology, Division of Structural En
gineering, 1977. Publication 17. pp. 1-36.

[86] Teng, Z.M., and Ye, Z.M., “An experimental study of bond and slip relationship of
deformed reinforcing bars,” Department of Civil Engineering, Tsinghua University,
Beijing, China, 1984.

[87] Baant, Z. P., and Kazemi, M. T., “Size Effect on Diagonal Shear Failure of Beams
without Stirrups,” ACI Structural Journal, May-June 1991, pp. 268-276.
Appendix A

Measured Bearing Stresses and PUNDIT Readings

D6-12-3 D6-12-4
60 60
-50 50
Cl) 40
Cl)
a)
30 30



20
C
( 15.9
10
I2

1 I 2 3- 4 00

D6-1 2-5 D8-12-1


60 60
50 50

40 40 .

r
30 30
..
24.4
20 20

10 10
r______
00 2 3 4 5 0 2 3 4 5

Transmit Time Increments (microseconds)

183
Appendix A. Measured Bearing Stresses and PUNDIT Readings 184

D8-1 2-2 D8-1 2-3


c 60 60

-5o 50
Cl)
040 40
C)

( ,zz r
•1-
(I) 30

C 20

1o 10

00 1 2 3 4 5 00 3 4 5

D8-12-4 D8-1 2-5


60 60
50 50
40 40
30
20
10

00•
r 1’ 2 3 4 5
30
20
10

Dl 0-9-1 Dl 0-9-2
60 1:IJ————

50 50
40 40

30 n
20 20
10 4 10 / 13.4

00 2 4 5 Oo 1 2 3 4 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 185

010-9-3 D10-18-1
60 60
0
50 50
Cl,
Cl, 40 40
G)

C,) 30 30
.I

C 20

ci) 10
20
10
fZ.3
00 15 00 1 2 3 4 5

D10-18-2 D10-18-3
60 60
50 50
40 40
30
20
10

00
( 4 5
30
20
10

00

Dl 0-36-1 Dl 0-36-2
60 60
50 50
40 40
30 30 .......................................................................

20
10

00
/c4T
2 3 4 5
20
10 4 1
19.5

2 3
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 186

010-36-3 Dl 2-9-1
60 60
0
50 50
U) 40 40
U)
ci)
30 ,u
C) 20 20
10 4 1%
I I
I 12.2
ci
m 0
) 12 3 4 5 Cl 2 34 5

Dl 2-9-2 Dl 2-9-3
60 60
50- 50
40
30

4 001
6 23 4 5
D12-12-1 D12-12-2
60 60
50 50 —-.-—

40
30
20
10

Transmit Time Increments (microseconds)


2 3 4 - 5
Appendix A. Measured Bearing Stresses and PUNDIT Readings 187

012-12-3 D12-12-4
60
0
-50
Cl)
cn4O
a)
4-..
C’)
c,)
C

ci)

2 3
D12-12-5 D12-12-6
60 60
50 50
40 40
30 /1 23.2
30
20 20
10 10
oc ) 1 3 4
012-18-1 D12-18-2
60 bU

50 50 —. —

40 4 7/
30 30/

24.4
20 122.0 20--
10 0•
n_________—
00 2 3 4 5 I 2 3 4I 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 188

D12-18-3 Dl 2-36-1
Cu
0

Cl)
Cl)
G)
C’)
cz)
C
I-
Cu
CL)

Dl 2-36-2 Dl 2-36-3
60
50
40
30
20
10

D14-9-1 Dl 4-9-2
60 60

50 50

40 40

19
30 30
20 20
10 10

00 2 3 4 5 1 2 3 4 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 189

Dl 4-9-3 D14-18-1
60
a
50 50
C’)
C,) 40
ci
30 30
/7’”26.8
c,) 20
C

o 10
0
) 2 3 4 5
D14-18-2 D14-18-3
60
50
40
30
/‘6
20
10

34 5

z 1-
r0
Dl 4-36-1 Dl 4-36-2
60

40
7 ——-

30
28.1
20
10

.10 2 3 4 5 00 1 2 3 4 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 190

Dl 4-36-3 Dl 8-9-1
Ca 60
a
-50 50
U)
U)
a)
30
c,)
C 4 1
29.3

2 3 4 5
40
30
20
10
U
01
26.8

2 34 5
Dl 8-9-2 Dl 8-9-2
60— 60—

z——
50 -

40
30
20
20.7
10
n
3 4 S ‘0 1 2 3 4
D18-12-1 D18-12-2
60—
50
40
30
26.8
20 -

0 -

n____
2 3 4 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 191

D18-12-3 D18-12-4
Cu 60 60
0
50 50
U)
U) 40 40
a)
Ci) 30 30
C 20 20

C) 10 10
m
.10 I 2 3 4 5 00 12Th4 5
D18-18-1 D18-18-2
60 60
50 50
40
30
20
10

00 .K: 3
40
30
20
10

00
4
‘ 1 2 3 4 5
D18-18-3 Dl 8-36-1
60 80
50 70

40
30
20
10
29.3
60
50
40
30
20
10
4ZJ
2 3 4 5 00 2 3 4 5
Transmit Time Increments (microseconds)
Appendix A. Measured Bearing Stresses and PUNDIT Readings 192

Dl 8-36-2 Dl 8-36-3
80

c3) 30
C
2O
1o
Oü Kz 2
D24-1 2-1
4

60
50
D24-1 2-2

40
30
20
10

00 2 3 4 5
D24-1 2-3
60
50
40
30
20
10

Transmit Time Increments (microseconds)


Appendix B

ACT Code and CRSI Handbook Predictions

?er/. I Coo. jyo),. c-f 2c’ J.

ecre’€. le& To-hC- .i


r&L — a o-
O o1 P-L

o- -
CQ c Cc’: Zê’(’.
4

€_ot or—- cc—


J± +
O—vj — Eedt-+ec& co&m. -&4e-

O€—v-&, SheCvrL3) —
CIL4. ,
2
c’c’ O Q -&‘JC

CdV&’. o-
,Cp

U20 c’ +L ?kC/3. ts) 1r


.

- he---i L4) —

cJc,}-v c’v i _-C4LaL o* -cC_ Of ‘“ “.

-- —‘-
h€cy — . a—vJu

-i -

Th.e_j evy-C’2) —
LcL’k , +o— €-.,--

C r- C —frs —-

cR —r---- -p

j93
Appendix B. ACI Code and CRSI Handbook Predictions 194

—t —- -
vv

JJ

OAVtA

() ——

Coi’-. ;OQ nc

o’’r. 14.’;L.

-
_w.&

cit:.

1=’4

Ori€—’j c4—
‘1J ZJf, cpa) —
J t4f’D
1
aij =

<a.zS?J LH?’.)
(oo3 ...
o2J’
r= c3_2. £IJ (?-)

= (3—24 )C o43

= (35—..5 )(o o3 4

= -t-z5’-) to c)

—25 ) --
7744
=

C-) C3_2 )(°.°S (‘e?E)

1)1:. rs) jj ( i L1?’)

(}‘3)
=

-
Appendix B. ACI Code and CRSI Handbook Predictions 195

ftL&J4 <4±:
az-c, 4cM_
t Lt t z3 /cA

S
1

J1 4
-c3-L3I49.c —
L_J
too 42. = 112 L)

= 2 -rz z) 27 LkU)

_-vj jLI)

-- 103’ = o34J,’3 xos-4’f.S = 1.t t)


L = = 3.7t)
1caQ 2’iT4 -
— = le3 L4)
one—-JJ øi—
w
1 j skeL7.J:
.5 4s1’.1
23
yo4J (t)
.-___---1 I
L_
‘H-’ )4
Q
4 ==- 2
V ’32.2.Lt) =

beo-’- C3;
2l.25 122
10= /O32M

O—k)O. &le.&ri3) (35—Z o.43) z.f 2.

244

.aLt) >

3 Lt) 1c
:
4 ’ iN = 7&.6 it-) 47 L& )
Q—iQ heOX L.4): cñi’• 4’-):
3/4;Q.€?
“%CQQ
0 •-“
(35 —25cQ. =
2OC)

vc, = - .35 ( a.c39j -- t71 Q.O244 __7)4oe 49.; = 274i (4


J
3 )
= )
t
3q.L . z.togS c4o4’,5’ O.L’f) h€c srr& L2-):

\J3.5C-t) ZY7it) 7L’k$) 2 ‘•t-)



= 2 )=
Appendix B. ACI Code and CHSI Handbook Predictions i96

}€tL ç4c4 fL)

A=43.21e. 3O4 -44’?.az’ (:)


Yc=
-F= °/ = V. = 73 ct)

= = Ct)

c= 2? = 33 c-c) =-
Qk)

O—v)- SeocL-) )0$(3)

Vt: = 0. 04 t4ts49.5 =
t)
2 p /Lk0)(4’5) = °2o3, =
o Ct) (-2. ,43= z.+3 . 25

44Ct) 4
4J) ‘/= 24’( 5e3q, -i75.qo2o3)44
>

= zv= ia L&)=

QP€ eL4): beojn ii)

40 3Z 3(-)= 3L&)
O.O203. = c4
%‘
(3—2.-’) eo’r 4it2).

VC. 3(o.MJ — z.CCt)


= 4tt) . 44’? •=HO.l Ct) ca. = = 04r1.1 Ct) a2Yjo c&’_)
v=4u)

i(LAL —‘,it kec)


A=42’crnt) tjl31 =7Q3cA r’e

-;=Z7 Y/L

4.
\—
74;3
‘4=
743xOZ/
rtL
= i74 ct)

X)= 14 ck) 4Lt


2iL 4 L)

che3vc 3)
= ).OcZ
473
42
1
,

Yz9o-t .OIr2 75.!,Q7- 2.


Appendix B. ACI Code and CRSI Handbook Predictions 197

cr. — J
4
( fr
42.5
7
t
0
/_ = .gç g?c
7
,2 33Ut)=3z44 c)

(3—25o.. )= k’ex’i c-he-, c)


÷i75.sIai72-)4ø73 33Oä)
;:;
Ut) > =

V tt4.3Lt) c.= 2z4.3,(1iJ,)


3
‘j.

2Z

fL .1.:
4
ev+ G€—)Rj k’L)
nst
On.,—io C2-):

Jo.4 —
)=I2o3tt-r)
4
Vc 4
o.53o4-J L3
3.4 it)
= 422
G
3
3tt)
/

-)‘= 53e+) — 1=b1.tt) ikt-’)

Ofl€LJ ‘e’.’’C)

43.l/3)o.O14 p =aoi4, Ya= t


U= 041

C3_1)(3) 2.737 2. (3. — 2.5LO)’ t

= 2. (aof oc4cL3 V .3g(o 3J4;t; -I-


-

j
31.Lt) >.
= I4..3Ct)

‘4=4iit) %4343Lb?, Q= 2V zsi’4(4)


Ltoi-j s4€.+kL

IL..tc,L = Q.-
)L?:SZ= rJ,)
30
1Ut)

= cO-=

= = Z2-7 ‘ = 31/
(c4 t173(t-m) ii’7.3 o>4i,= z’ L)

Co- 2LI z.a)


Appendix B. ACI Code and CRSI Handbook Predictions 198

r€—’i)
,oi: P= OL93, -= oz4-

Or—”Wj c-L2..): (3.5 — 2.424.)= 2. >2.5

+S4os-.4= 34.3 tt (odj ..t

•>.


= Lt) 93(kJ ‘/=azLt, kCD2205.th.2.L,

one— skerc4) ectr’r t


4
<.r•’ )

?.0L3,
-‘-°4 Q..Q5_ix33Pct)
4
tsJC 3g (.)
(3—2.5?L44) — 2.3 bem.r’j .1j-iL2)

23(O3’ t$10.°I3)4ot4 = 3?J et)

= 24 Lt..) 2 j7 lLQL 7.2C)= %32L4e-)


1
r

V71.5(.) ,

@ne—’ seor-)
=4O.24c4 $q=4q, dSI.tqirt t
= 4Jo o—’J-J chea,LZ)
)#é )= 7, v= o53c4f 4o.2 =ALO 1t)

— = 370.1 Ct-) = Ye...

t
0
W = 2)sipe t74Lt) 7251 ckiJ) = 2IJp Szt)= 4-t-?..)

ofl€—i. arie—w e1Vr (4..)


4,
4 ‘O, 2J ‘4cr2 -‘=
(3.—2.5Q.24-)’ 2. )Z.5 =

> xS’2 -
z.5Ct) )

V, = I Ct) zV = z44.’Z (t) V 205iCt), NccZY4I0.2L4O23(J)

‘ie- s4-H (..)


5 -*41lo35 53It)= SLJ-)

Ci-)

Q.3tt) -
Z(t’J
Appendix B. ACI Code and CRSI Handbook Predictions 199

3U

As = 2.7cr.\
,
1 297 V/cm , 45ci.

3 co334j.7t
1 92
f
—2-

-A = 44)
l2.LO 419qX2
.ç. ,g
1
_

L__. i-tt L_J



=
a c?&-7 I
)(-t1t)

lo I 37.S •
=
l.)L45I3’

I3°° 1*)

29/-

sf
3
1 M= 3.13 =43.
— -
A 4’7.°’ e’?
r IO.S297O+32252O —

47 /
= 1’°t’’ .f-& 2529x45
47’lx57€
= 49
S (II4T1+
7
5 —

) = 32 t-)
= .ZO/
q
3 = I4i. c-t)
LJ

=4c4’i

A.
“ $.252c
•ttj= 1
fGq

tE4 4
9
(L4 )= 3Ct)
= —
I = /j
2
!.3)O -=

22.Lt)

= 3L3O.O qocL= 3Z5(L)


Appendix B. ACI Code and CRSI Handbook Predictions 200

°
I
Vt = S3a.J s%4474o5 (;t)
= =

b —:
43

=
= I47.t)


3I 3x2.4 tt)z13oL4i)

Oi1.—J

vco3L441 =
(t)

12 —&;

30 Ct), \/

3Lk)
Zt4L+’6,

— 3 rJjI4 =
eUc3):

= — O.o1S

= s—zo.)= 2.’L
_4
— z(o31fi -+ 75l
:

IS.75 .75

P +3/( i ‘447) =

v=( 3S (-t)

y=’.2a) , D2.L±) >

(t +Ci&, iork
Appendix B. AOl Code and CRSI Handbook Predictions 201

o€—w k’r (4):

b
4T’S/çi.r I G, hO

ri ae—j’-” 72kt)

—b icì. icrL,irore)

= 3 zo k;

C C
tjxi n-i)

4 (4 + 447) 35 C.ft

v=iJ;- x3 3• L)
f 37.
-
.zi4?t V =

=
V=
1 tt) 372Ju-i)
=

1jo—,Ja eCC2.)

CW-, = 44
Z.14/
= =

v’ = o,o4-G
,72.4 Ct) . 4jj ?x4+!7J7.2tt)

V= 72.+C) 4 4)

shei’u.)
M’i
— --7.4€ =.

i.°’7j-
a-
1
3.i-Lt) 37 (-kiJ)

1
X2

-i4- (z)

L9

4e 3 tq-33p L+) z3 &)


Appendix B. ACI Code and CRSI Handbook Predictions 202

1 :2ct
A 1

44 , - =43. c/
‘---7 4OZ+!%3 ‘,
444 i 3t
— —---


S4-.

f44n

A= A -t ‘

—43(4—--)— 3L,Lt—) -
-4 3Lt)

23)

L), z-i -
g4-x Z3C) f7t)

= Lt) zc (:&eJ)

-—M he—c’)

vc=’*_

,
*JL

V
— .3
Appendix B. AOl Code and CHSI Handbook Predictions 203

e—c h!G-)

P 432/(
o&4.I’) 24=Q3 ( —25XO4) z.32.5

2.(O3Jç II.O t) r
t
4 )

‘4i6T.Q’i) -‘/c

t—c
lAu/= (3w— 2.’:2—’ 3.2,-aS
••••)

‘4 .; (o3 -7cgLYøz44 -)ot


=32.o t) >- goxs2. z4-)

VL21%,+Lt), , Jz2cMt)

bri- C i jrec) q =3Mb1.o = 43L±=4737 &ki


4
t’

One—’Jej heo’L4-)

= )
y —
.(o3J --t7$ “c’L —37t) .

O77J fD 31 (3_2so7’fl 5S’-Z.5


4)°’4+
v (° 5°31 - 75.I i.’j Lt) c
c—c. l2.I4/=o2 4o.z/,=o4b, L3—2.5)=2.2.
f3
L2
(asJ +jT d443X4b)2ZI.71t) > 2 J4t I Lt)

v 4q4t) i-p’c V
6 Z7L)

-t’o—’i he)

‘4 = .r2.ijTh *-4 = 411 .o t)

/\
J
/
I \
— =

/r
L_J -

t -

t = ,22.l(±)

S2
243
Appendix B. ACT Code and CR51 Handbook Predictions 204

O-’. ‘—r ()
7I---4 = t
7 cM,
l2c5 . ‘4 J_ lk.t) =

= 1
J 4qL4) =

F1T JfT’. 4S()

=2 .L43S3’? o’Ct)

= = 2.?t &)

fl€-L &h-ø+K


= 73.m, 7 7t
ZPc
=
3
$ 2 ,
S3a

‘2.43o COO = 3 cl . 44.3

A.qs& -
= z37

= z
4a
7
t1r4.73 zS3 (7o’z t--) /4 =to44ct)
e
7

A
=
1 I7 coO°= cA6 -= 2q13 l/ =3 A= j4p9 cc’2& , ‘z4-c

A4+4oco °=4m, -1’ 33k =W7


ot3Z+331
t80 X5S
=

O.O23 (70.0— J ‘17.7


94(.3

23?( 71 — 0 ‘j3c2€37 Ct’) , JpIe= xlO/


14 .%
4
3 tt)
= .‘T3

3iS’ 3zD44 3
.21t)=
1
£ ,l4L-kL)

C /A )
Oi— WCM SieuC2):

0—: V 304 ‘e,S(t)

b— , V= 3’44C3 xt4-x-70 a.o(t)-

c—cr i1S5.tLt)

= 3w =2)2O44f.N)
Appendix B. ACI Code and CRSI Handbook Predictions 205

O—vJQ S’.)

Q +$/ rJ’OD, tJAj) z3A-’


c
2 3 , (3—2T5?25

— ) , )

V=4Lt) ,

C-—c
2.5
-j
v)

V = 71 =zbI.ILt) > oQ2 4Z.Ct)

V=432.!) ZJp€ = — 32L2.Lt)

Ct+4-L, orc&) aç4S = 744Ct) 74C-kJ)


&e—vJc çhCOV((+)

J— 4’,7g/T

2’L, p_4.

2E(5oj .jt75 _ _)3c7O.2 3,Lt) 4


=

_44 O
P /
o
7
r
4 ) °“‘4+,
-- i.i(o o3jZ- (cl ‘44 )M.-7o . Ct) .

jr 039
C—; 1A)/
P /LIsos 7Y2)
= 4
3 ,

V 25(o 3J4 I75 ij


9
3
&
4772IPtt)

V. = 9z.3t) , = 4z
.
5 t)

— 3 3 3l = O.4-Ct 3 = 2! (k)

“-): /A)

LA)/ 9
2S4/ a
O
7 I

n
0 4çJ 34.)

SCL’) C/A )

t’e&i r Z)

4.3>&31 34f.••
°‘‘ iP_i€ =a.7
= 3It)
C7-
Appendix B. AOl Code and CRSI Handbook Predictions 206

-kLr&,L

==2cAt, lSe A2)c3C0S3/4J


7
d. d,74.t
=
±=4°’? A=
.
5 -

$02 43U -z
C3.C z7.3t.

3±3o=

=‘74, A7e _ 4-

34’ 4-31
34c! )4-it &73. — i)= CVI) .
37
IlO.7?d/

tz7-))

= J
1
3 L3
7
3z $I%(t)- ‘7LL4d-)

OflP_—OjerC4)

O —VAJ t/L2.)

oI-JiSZ73’’ o+.z(t)= J€
V= WP&.
k—i,:

C—C: O4j
3 I’T
3 2i(t)
1
,
ct)

= = z5 Lt) 1

—vc .hoL3):
)

‘/2.(°34 c,ct73 z4Lt) -> x13’ 21-3;

‘4=. 2.GA)=

Q—w
-— -U)

= 42 2)
Appendix B. ACI Code and CRSI Handbook Predictions 207

‘! (øi -i-’, .s;c’o4.a 2.j3(r3. 53

‘‘c 3L1)= 1? )
335.S.) -k)
51,e3J<J) C

1
t—P’ ecrC2)

iJ/ 24/3 o

= o—- o= 4o.t; .- a+fox73= 2Z7t)

V c-) 4-L V z 73.,c.;= i2ZLkF)

()
)(43;2=
(243 )(4? 140s Ott) t€ i 7
z (t)

4IJ 1:

f .3c72.s/#
I I I
14p7 cm’ 4= 23 ,o
4
s/c A

L L.
2
,
7 g

+4Z7bZX.Lg+ i4o7.x Z3 )•‘ 5


9
4 ,cZ7b

-:1-- = 272(47.4-

L1 LtJ 1—. = 4iZI = .-f)’ 3(&1)

e71—uXAJ c-eci)

E1 o—)o’’-j heo-r2):

L_1 i__I v= o54-f xi’7.4- 244.C±)

= Z.4W = Z43 (ki))

Ofle—Jck1 h€o-’(3)

P= 32

TT Z rr V
I
)
7
>2.5

OOO5)I(W7.4

j .G I 120 l° = 4z.3 L) > .s ;J ,ø7.4= 372. Ct)


Appendix B. ACI Code and CHSI Handbook Predictions 208

= 3(7,Ct)= aiJL :I?e= 74a*)=73o


fl1—W S€t2C(4-)

/6ci-= O52
=5 (3.’5—Z5)OS) 22

V. — 2,p03J ÷tiod74

= ,IZ.O(t) ‘.

‘Jt=V2PC±;)
I
2 J., cot 44J.1 I24(k04{)

V 4-(sV 7.4) 94
= 2O4Lt) 3( (-I&)
-o---,JO.. 1ie?’-Cz)

= c wi- acot)

t,eOrift

— S S3.I
2
ofZY3 L5ZtC441)

etr4C):

—-: QZ 32
35=Lt)

= 4t€’

fLLVL
= 4o.ocm 44’O ‘8/ =e-f=4 2 t-F=5L7 d
4443a &B5÷924.Sit7?cbb .

— ,
1
4z.4t 4’ = 4•
44455 ( —
1k t-), 2Lw =

‘-=
44LL= 2()= 7!(&I)

c€o’) C/Pt)

fl€—VJ. A’L-):

Z Øox’.
VL=
-
4 j’31.3Lt), ztJJe = = Z24C+) 7kS)
ce—-O- t’xL):
7/6 Zb, L322) 2.5 )‘2,5
/,a)

/= a.5IS j— Y3Q .. 34-o C&)


Appendix B. ACT Code and CRSI Handbook Predictions 209

= ct) = = g5.4t)
7
‘ = 7732(&)

OAi0j sie(C4)

= = (3—25XO22.’

V 223(o.5° &ai= 6i
kCt)
7 2. Ij4 v.1= 4t)

= 67.1 ) = = = l2352CU z4Lk)

4= = v= 37.)= z24&t.)

4o-IMi-): ( /e&
3 4
rio

_ioj erLS):

Ci); ?axfn
1

1t= saxo5= I
3+°t)
7 c4os3 r(23tt)

= 4= 2.I3.Z.Lt)2S57 ()

+t
- 2795 S/- c L,Cm’, /k 27 QS4 Z
.7C?t’f? ‘3
7

‘Th /em? s -c”=


7&G2x28b (it)
= L.)L19o37I ) i3.4
pZ2,65

=,•a±), ‘Iot

oJ-U):
/A)

one—t)Jc’ zH

VL = 53o4j o = i’Z z(tr) ziJr V. , cQ 2iJpe z+z.s1-ct 239j k.43)

ofl€.—-j ieL3)

= AAvh&a L7/
4 27 (39.—2,5)L027) 2.3’2.5

1O?(b5. 411.1-ct)> .5 jxjZ0Xc.9 = 3C35 c-t)


{
vc.= 5(±) = J = ‘}27.Oct) 73C L-JJ)

ore—vja OfC4)

= .53, (3S_2.5)2.

V 791 )iox6. = 77.l tk).

Vc7JCt) 24L, 42Ct)lLJJ)


Appendix B. ACI Code and CHSI Handbook Predictions 210

—oj €&r)
‘4 ô&
J
7j 62’f.’ZL), Wt= = tlJ.c5(t)= 1ocC&t)

_t14 /s.’i = o53 ‘ , nci a

G—U ceckr(3):

i,u—irw +‘4-, 1e+Cz):

= izt37 x &= l5
6.(t)
7
= 4o= 4$Jpt 2L4.)Z17CkJ)

4a
r-L
4
f{
k3oc, -F 4’so = c-*t Az d’4
-cc=-34- /c4.42, A= +S.11cm> ..f —

x4 +ibi4 2.t--W)
4l((7

= 52.O/ =439 Ct)) Icøt 7i4C4. (j.)

one—- heov)
one — v’JL ¶4i€OYC?)

o 3ô4j a7= fl2 Lt), zN,,t= ‘/. = 44 = 23b.4 Ct) 23) (-?iJ)

erC3)

=4
/rp °‘-°. /(/a)
i75/,

‘4= 2o5D5 + ,i3Ct)> 34(±)

= 4-.1)= 1tJR. ) 4JpL ‘7.2 C1)

or —icu.j 4ec’cL4);
=o.i, 3-23-) z
21(O5Of 4-
1 t)L )(J=9c41Lt)< 2.!5i9 9o(t)

0
V 441() o44 a,t)=jaZ)

-tuio—w&

‘4=tC&77 bI4Lt),

b-wj ?heo.rL7-)

eA” ¶d7€+k(t) 4€X)

.JZL
QQ
0
‘J
LO’t = 2M-3.4-’tt= zc1 r3 -&ii)
Appendix B. ACI Code and CRSI Handbook Predictions 211

iqi-:

i= 0/cmj lj;’ z’i/-,


c-= qsi f 4l5 = oCW

(3! 6t l3,OX23lsO —

—z 7 “Ltt
27o
= 2W (2S —
) = 47. LtW)
2L
7
2j =-
147. 1Oz/ ‘4’223Lt) , 44L±) 2S3(k3)

o—uX 4A.rC)

_vo&Lj e.r(.Z):
VO.3C4S L72 = 132 G) , Y = Zt’JpJ(l2.) ‘JeL = ç(±)32 4JJ

Oflf—WO- €,A’():

ojG.,
/cv..A) =
= C3._2.5)cIV37,2.5

‘4= 2S (
qj. t7! oQo3’ --)!ox25
-
0 o.: t) -
,c(€o25 = !:t)

V4-.tt) = 4\7’.2i) Q3(kFi)


cfl’L-uJWJ hXC4);

= —i z.O3J4.t5 +!72is°37 —)o25 !ze7 Lt) > O(72= &(*)

V = J2= 4

—o—W.J hetML):

J4
V t.
7 -t-kD92-5= o3ft), 774.i)= (-JJ)

-+1*30 — JCA

V o4-(i-i 4X 923D(±) .
4-Jj 2Ox25=2o!.I (t)
JL-=
1

ke t) .j /f)

J,L 45l,4Lt)

— c,L 24.3g.4ct) = a3’I3 (.J)


Appendix B. AOl Code and CHSI Handbook Predictions 212

PS
43

-çL 4
Skv

k=2.L’-, -c=4s4 /- =‘935 AI2.Gt 44/ =cW

k3,I4’ ,;+Z /(A.t-, J ‘43 Ac. øC


r (Zo+S4+ p24x414o)+ 3 I
tj

Ar 94 °(‘1_ rI ‘-
43) 192. t4)

= = .ot-), = 4
J-j= 1IgCt-)= p7 (-?dJ)
c’heo..rci): (/F)

OhE_—v) t74XL2)

v= j tV- = i%.’7 Lt)


2 , Jt= v’4 — ?j3.4(t)’ tkii)

w&j

c’= —,

= + I752 ‘-2. ) Loj2. = 75.4-W) 5. I L)

‘4= S7.4Lt) = L 1150.8(.t) —

one_—W’- €-A4-

= (32So.3)2.’ 2.S

v -= ‘ji ot,28 = 1371.9 ct) cga t et)

V’3.&Lt) = 4JL= jq
.Lt)
7 L9Z1Z(iJ)

o—J& S1eosA’O:

i.o1o7j4oi.1z)Xc12. 24,4Lt) 7 .OCt) 9OZ4(4tJ)


“4=
—bo — &tj ie12)
LA)/cA ‘92 Q3

vc _c.4-<j -÷ - 4xD)92= k2.8 34-17.0(t)

il1-Z. L) =Y 4-ZCt,= (QJJ)

-_J8vXC3): C’/A)

e”s 4
e M1Ii C7-)
I’Jc.Q = OS 4I3xo 2D1S.3(t) O 4A3 X? 2kO(t)

= oL(-kN) = = 320oCt)342 I (&)


Appendix B. AOl Code and CHSI Handbook Predictions 213

441+
hec-tk

Pcs T.24 Z4 c )= 23cS’=27.7LC,


72I4/eA

y=
J
1

2J.jt ô/= tzLt) , L’ 4.”


(‘/A)

V=omo+Jtoc1;.L1= iS3t&) , k12 =vI 3O.)


2I’(k-I)
3

1
t’ne—w 5I1er(3):

/j,J) = =9) L3S-2SX9)3O3>2,

= 2.(3gJ + I
5.oc44
7 )fQX 4j (t) > 443 ±)

v= , —

oie—v Me—rc4):

y __L o3J +i75flx&ôt4 j)o[ =jZ3


,Lt)
7

N/S4CtE) = —
‘f’7&)

( 5)6.7Lt) io2 (*±J)

o— ewrC)
W/a 3/o

Vt. = (- e.4f4- 3tt)


tt. Vc

r(3): /A)

bexr 1-tJ seMHC,)

t(35. 2t)
1 3 ,
Lt)

= 1473+ (k.i) JJ.


1
4
9 ‘25Z’ t)=A-
73 L&t)
t
Appendix B. ACI Code and CRSI Handbook Predictions 214

‘44*

1
-€tt i
9
cm€’

-fk-1/, =T4c
92.w ,
134 /CA%j
1
344’4
4r
k3 ‘
—---) = 3.1 -)
93O1-= ;50Ct) , co
1 44( 1tO3pC)=)O/,CkiJ)
e—Lc’j e-LU (/A)

—j ç1iec’-U-)

y=oo1-Jv’z.&=’ J.t) = 2-V 33.1t) 3,bc&i)

,‘he-rt3)

6
O.OZ , 3.S-25XI)=3p3 25

I o2.L, L = 32 L1) LR92. =


= -

a. L
‘4 3jLt), kco’1cI42L) lo437L-ki)

o—iioj iec+)

= Z (PD1 PJ I’(5.& 021g) IC/J ) J.4L4)7 ‘—fl

V o9.ILt) , J€ = 2-V jI2Lt)= 73I(-&)

o—u ectv- LU:

I2,3t) lJ,L — Z.
= 4:9 Li)=’ J)

&rL2):

‘4 = o,3o4 — (I4 — T4-x T= lb3Lt) < 48i 23Z3Lt)

‘6 13Lt) kcoL = II.3L) l’f7O (JJ)


—je--W ie—[3); ‘-VA)
he’f eM-rL2)

=çI3)7q7.gIt)

=
7’ -) R= )
2
O
4 .)
tt)9&2-(4
Appendix B. ACI Code and CRSI Handbook Predictions 215

o.3
LLtJ S:

AS2?.i 4iol,1:, = 2I ,fc A4 p


4.%cYD

= 574
= 1I,2= 5i2()

4eoii)

iL= a 1-2L9 p = iiCfr)

rii
LZ1 U
2t”d
II
V =2js1
kci;

— l’Lr 2V 740kr 32()

an—i 4,e&fC3):
‘)7/O4Z

(35_2,XO12) = 4z.5

-)
.

2Z?r25C1J)

LJ
D (fi5JSQSi
a—vJ° eaA-c4-):

47#I_ = !7/2 -S) C3-Q.X°25) = 375

i jg J z?eq37 - ) iX1)
oq—w cheorci . O ?(52i#i k’

i4
t
0 = 2V 54 =

rjr1 eu?r’ -iji-)

JcoQ
=
-I z45 O= 2.

Lr L.iJ) = 3’?O’It4JJ)

1o.1-
+ku-t the p+ti:
=a.92, =4,oD1z ‘=)47”, i
t
1 =
3Z/=
)=
Appendix B. AOl Code and ORSI Handbook Predictions 216

ene—Aj -heacL)

= z5. .4 jc -LV — —

øfl€— 4e2)

=
, 23okkJ)

—‘ h’3)

GZ/ 5?(4vT) J

u +2 a4Z &7P
=zJj3.44’(4&)

on€_w&i ietrL4) _ +

beoir ç-c) eoMYUj c.4--eL7)

Q
0
ki &c(2 X34Q2 )L LS= 24 -r =g X34 \O .43,,3 ep

— 1clt 21J 3kt)

A1&A4

A )L3i0 =5.4mr, jMoH1’si., =4oon) 4ai


200 () = 5r)LG/
)= )2.7(—M.)
2ipJ =19 k) =4)=t2S8c4)

( J7p)

v= ‘=z9 =zV= 7
ØkJ)

ofle—WR ie-r(’)

‘°°

FI±T 7VfxoDZi
f /
L 3S9
2V

Q—)O 1eo4)

y=29o/ =o L3,5_1.)21S
L_F_H
V= -

M-3(&JJ
I 600 IL7 = 9)4O(O
Appendix B. ACI Code and CRSI Handbook Predictions 217

Wco=

‘€r-i) V - 33Z (2o+4o)4ov= I4sQJJ, zv


—+-ja-4 ce(I)
‘1% •5.. V =

G—1JO1 h-eAr(3): = ‘v3322J ki4i kt= 44=?q6(kM)


‘9’: 17eoJin $ +kL2)

A & A
fLQsr.L = - 410 , Iz66&iJ)

,33 &t-1

C N/n)

exO) VOIJ5X944VJJ NcoL2V = 2


( jJ)

OQ-’AJ ?4L3) z.87E,2S

‘4= 25(57jE5 - 1
2
7 4,OPclZl 4ct,= 21--bJ . a-J = S

‘424-J QVlC*)

ane— -keoJ(4) v z.zc(I57Sf i724 >OOZ.1 97?c4OD j53’ ki)

4
9VJ,
\I3

—o—-. SoL): y=. Q33fl z4’o


= = 2Y 3344 c-ku)

—)---Jo heoc). .lAJ/ =• V= 1j2-4&, LVc,= tCft3)

-Jo—1)1c.Jj 4At3). S2I mi’-j V 33zzg2.1 c72 J, 0


Q = 4’4 z( J)

A
A= 77SPrftf4Ih1&, 4’,-f3’-1 r
c
4 1Jpl€
o4
4iJ
)4= 774l’’— ‘2 z-{ 12O

are—’ €O(U):

= vj= 7o4c1)

OflPJ€X(3): ?
/
7 ct,4’a) = O2-5 (c-zsxo.zs)=2.75 ‘25

V 2.i978JIt )95D?4O W7 * < 4fS3j D.x4o O5é -k


= qc = 1 = ‘ Sk -k)

o—Jaj ch4 ‘4= J - I7z4


Z5 (o I
7 2O ) ox4oO= -a3t) oj 7(frjj)

‘/i3C) ‘c324(,’-kJ
Appendix B. ACI Code and CRSI Handbook Predictions 218

— *‘eR’( ) \= 3
‘3322j- 240D x4or I7f
J, ‘icQ 2-V 5(J)

J/=o., v= -
7 ii, 4
i
0
c
f lI
S

—ha— V = 332zSfl zc4o-= aiJ= JJ 4tJ 243zL-kJ)

Jir L7)
bert k C’):
S+-€AC
+
1

t Q(2;’-3LI 2C2I5C4&)
0
k x c&

z
73413
M= 7g4)L4o (40o ccz) — ‘j= 35 J

coQVfc

one—wjseLtr3): = 2.5(PI57Ji + )5Ox4OD= 04

11= ‘z-V, =
e
3 ’(IJ)

one —ioj e’X(4) = 225(o.IJ * 7.2cDPe2 )so4oo=1z3 &t )J


QJIQ=3
L
7

I’JcoQ zV. = (4))

O—AWv ‘WLI) c,&=2V33O&k)J)

H-oio—woj c’1Z). L4i7=o., =V=54t)

3 —W ‘c3): V.= Q.332ZJ 2I4’’ +—F) 4(


,e.CLFin 4
i
1(J):
ske°
’ mip

a z,c-2L,i .2c’D= 2kJ)

, {jZ6pK?ci.
2
A17Smst ,

1r t2i-tt), ZkiL{

(/A )

0—Wj e4rc2) b5x4OO322

o— 1xi3)’ “4 = 2.5( 2 4J’ z,s4y= % jJ

Vt95& co2V7(42J-)

3—o.%j SeiAL4): 17.z4 o)95rs4o= 4L4 g3oJ -ki’)

= J cøt’ 2-Vt -

—-v.o—vi&q c*XC) Q32Z Zxi4 1


L48’zi) ct ZV 3QZL&)
Appendix B. ACI Code and CRSI Handbook Predictions 219

w—i- se&rc2);
J/ =oE y= 2CJ)

— -f.-L’) (t), 14t= 4


tJJ zztiC&k1)
\4
a -h’en -F+. ().

Wco 0g 22&L2aO’ øZ-X2Oø 215(k)


1 4I(R)

?
1
A
j+{\; 7rn 24Zl4? ,
— lZ2(’LtJ)

4o’ —__--—-—-)=124.1J--nt) ]{ - (3 (k’i)


o1€—W heci.

€-r(2); Nct= -V
24LO2
7 )5Dx4oo yt 43J zcV p L4i)
‘4=
= (JJ)

Vt = _J
2 t3Jj*
5

= = 2Y LO2.L)
e%C): Jco.t 2V 3
i31J)

O—IA3AJ sheoca) ‘V’)/o.s,

•-IiWQ- ]Cj ecaAL); 332zj2C2\)t’O 537 Lk.i), 4lJLc z4


U).

ZG= 131 tCJ -z4(J)

s4Te1L1-; _ p&
9 22 , 4t IZ,4LIJ)
75-)(43
) $J-), z’1= 12’O/,, 32 kiJ

(J/)

kli( C2). 4-LktD

v= z;(7SS 71 o’z1 ) rx4o q 4*) q3 ,cjc14pO=78P r-)

1’c 2-Vc

v= Iz I&-(kJ)

V2.iL),

o-t.1C’Aj se-rL) ,5 L(-ki), = 333Z L-k)

—-VJ e4-ct) i/ —O5)


Appendix B. ACI Code and CRSI Handbook Predictions 220

-; V= J \g=32 OQS(*JJJ)
84
%.

H
9iLt)
€i -

c s2O 2iJJtJ) oS ±!!x2,2 4x2oo l22(1Ji) 4pI 5D(4J)

l€w h+; A 1
y
7
-y -cL=\e.2 c4pt 125Z LJi)
78;x4I
My 74ioC4— 1 ZJ?& 52
I22L41—ff) 4i)

O,-WQ’M
-J
EOJL) -

-
u-• /A)
29L+)J), c.,L= 53C-))

o€—c .he&rc3): V = 2S Co.I7i l2O.21 )9 o4o= 77’7(-) .

Wct=2V;= 4&)
flf—vJc- 1eL1-4C4 ‘4= -2 =}75L4). 3J 4oo 134L&bJ)
cQ LkI-)

*iO—iOA.i .1.eLkYL1) i’4= 25Va(-kJ’))

-Yio- icj /= O. QV I3’ kiJ

V3t2-3 j 4 &)=+J, = b’ (-kg)

e’-’ fr+t) 12)

QO.W7Z2Y,2X2Ot )23(k1J) A1LO ø cJ 4)Jiie 4O4 (-)

MI
9
5 m, 7?
7s4IO L4oo —
)= I ()J-fli-) zJt 25 .XI°= zJ £-kiJ)
O-Wo 4iecx-(e9 C
1bI45X4 23(k)) cot - Z( (PJ)

(3)- /. 2
; (oi’37i4 - --°i -)9ox4= 73L-b). 33x4vqkJ)
cZ’ z)4 — ia’ ki’i)

OWOj i€/L4) = Lzç . o3o5 )(cW)(4po= L3i6LJ)

= i= ck)
-+Wo—W’ eo.tL1) 4c YcZ’ C’)
—kJo—’?J-4 2
*E

-ecxrC3) I 2O?&)
Appendix B. ACI Code and CHSI Handbook Predictions 221

bead +flq+1L) rL2)

1cot = 0,S5’c22 fl,3&fr1) xL74xX23(.i) l C= 4.=32Lk1J)

z.31PoL Jt= 4$JQ ,2L+-)


75x4l 0
H 7x4ko(40o— 116.21’kii-P’9) l22X10/

, 4CkJ)

(,3) 27S +.rX0D0ZI o,g3j Z4ao =zL ki.3


9
3s.

\/2’12-(k) lct iVC 7


g-(&1i)

o—iaj “4= C’57i - q4x=iIs)’ 3o5x4oo’’15tkJJ)


,

MOV’O V=°33flj 1
24odx=/,c14(J2P) bica — — 2O (ki))

—-io—Mcj shea.rL2): W/=oc,

ck€o-L3: &) = ztc1L cJ)


to_V’.Oj
r “4 = 32aj
1
-
eo”n c-høi- L2)

ozs!s4 x2t= fl3zLkJ) I1 = 4 l-i


1Q=
4
4ce =o.x ZxZ.3 2
OJ= 14Q))
r-- L+j.J)

-çj=z,.9 HP
1
44o
tr s-410C400— )“ 101.
$2. )(75•D
9
l.
f
t

CD /
3
2Jhl1so
l
0 1-kJJ,
ç’Ø (’
2
t”4Jp 2O22.,&)

oe—9Aw”
i4ct2Y iL L4’&)
(H-) (H-)
one—A3t
- 4c 950S02) = O 125 (35_25XI2) )-Z
;ga)=;;
-

\ •/ \ •1 Po ‘4=. 2(ol73j
24,coZ1)0M0O
7

31 (fJ’)> 9 974o=
o4’3fj L41SJ)

I
i
I
-40O r
1
V’ L2JJ), Zp2o ?v 033(k1J), =4Ck-i)
Appendix B. ACI Code and CRSI Handbook Predictions 222

o—u 4) (3S- z?02.7’2.


=

25 7
(o
S J -i-i7z4-oo2I
= lct 2V,’ 25.4L4&)

—-uo-- -L): C/A)

-4,—W- ct2XCZ-) i/ =O2 eo

V —
—i-- &&i i —i- x S
7
3 333g (-ji)
- O25

= V;c&) .

-3o_vjcod(L2):
/A)

fmioi41LI): i1\(2
9

).Ico-Q = ,cZOt 2.CF) L21 2O (27(b) =42N)

A4”, co4L 12S)

2Mp$(
‘i 9
,c.
3
5 D —

Ofl€—”3Oj Shec) ( /A)


ic,Q 2V./ UJ)
on L2):
ot€-wcj 7 ‘/c1.
c>)
7 O1é
=

v,= 25cI7w?
- xOO , -—
oI25 )T O-37L> e.4’3 )c4

V = 7’ ki4), ZIL - -
— 49t— a344.&iJ)
h€L1-c(#) _L) r 4tr
Gflj v=5 (5
sJ -
7 ,4x°”L’ 3,

Ye IDI (kiJ), c24 3cn’2(kI) =

±cio—vJaj SiecuU) (t’/A)

*ecu) V 575Jgoox4oz
= 5I24(R)
vG
ax(3) (7A)

)J= oS .32,z1SL-kJJ), x33 =ib% tkiJ) Jce474-LkI)


2
xo
Appendix B. ACI Code and CHSI Handbook Predictions 223

5EZ—3

- 4E3 A12, c

—- )=-si& i-

41 P

o.O77
0
/ I ?-
L7b

‘779 97b(4.34—
7 3)23O /.4i’..
4

20 =
2
3 Z I3k-1p, t4ae2Lpkp= n,tLJ)

-t -
L43 ‘.

/ \
12


.4c t ) —
. a3233(5+

— 25
A r14’I077O (44 _--) = 44 3 I-Iit

4 B”

+= DT’& k =- 427.
=z S—
7
z?63Q(4Z )=8Q7 /4-, ZPJ$ 9o’35 èZp, I7hd)
1
47L2k

-fD, A49’737i6’j 5ItL,


7
=4Z -P-= 2’o 1’
“‘7i i243 (+.27V— ib -;. . 2iJJ = =i.i -lJp
O2Z-k L-k)

- 2t’’ ?) As — o.33 &1,


4 joi- çpt,

24’2 47
M °.33 4S J) H

-=z&’s- A.=°i’7
,
1 jOCO’ ?

= J 4.lz’k =
z3 L&)
Appendix B. ACI Code and CRSI Handbook Predictions 224

kvf&) :
/A)

.ii V,4.3T&= .c, p = 3.IiZ/


4375 )
I 35 13x47S344.c i1.
V=(I .j

— L)

= l’7)
5
4
94
Yd/H=43 f= or1r7/( 4.E;y: ‘.‘I4

12 1’. <3. J ,x434-= 3’-7 ik.

3t Vc) Q
0
Jc 441 3
V/M=43$/2=I.74I

y=i_ -j- ao,7x i.74)3 (‘7Ot, fl. - 3.5S 374 I


zL’v, L1I’Jr1.= i4&’()
$4.: ‘4/= 47
43q4. 4/3?(4)a.O2.5

‘=
-- ’) 4i4 793B
7
3 j x34-34c=

I1L I.9 7IL4)


sc /’1
4
/K ‘

= :142., I’ 3.5J J5

MJ=V 4La= ‘ftJ)

= = /xi..zY
737 °1
OZZ iL 35J’?( I3X47’ I6
zk= 4, ILC.Q 44Q zock =

5:!_ /o,x4,z) =
q34 I. ‘

2t4k= \4, 4 LQ47 4= 5(-kN)

‘Y= .3c-) p=
)(42S) OOiZ

V= +z- oiz.) 3?c462 93 3i I2’

4a-•= =
Appendix B. ACT Code and CRSI Handbook Predictions 225

3flij S4€OA”3):

4375/. = 3, (3, —
= 2.9

V 25(9j, +2ODOZI 3YT) ia35= zo


S fl,.
9 I3xb3875= l,

= ‘4,, = t=
4;J_ () = t4 c’kt)
25 =352) (3 25/)
4c/’, =
= 29

2c,36 ii,. ‘- 34sq4 4


Z3o ‘.

2Q \/, Q 1
t
4 r)= 1Jg()
Vd/ 43/ = (.3,5 — 278 ‘25’

2
(
5 ç ÷2v ko)97x34)13 (145 2a344 iI, (-
I3L13b5 2Z931 li’.

2J V, UCG.9
1
r
444
40.’7 r

25
43q4-5,/ (35 2/) Z.7’T > 25

V. =25( 1ij4s-k 2.)3430’3 b. . x3k.3’?+523ag ‘.


1
i
2 /-

1= Vc 4k=4z.S4v’ 1cbJ)
4’3T5/j ,s _2s/
3 ‘2 S

2(rr7 -7 x3)(4Z= ‘2572.1 I1.

2-(4tVc 2S2Ll , 4$L? 5I. -‘r (J)

,ss. vd/ 4
7 /,zs=.2 2.5/.) = 2.y “zS’
\4 =
.c
2 j+2vX3.Z
7)(34z1I5=
c74r( ‘ G .f 34275 Z5’Z( ,

2=V7’272) ( .1c49 51.4 1c. ‘r2. (i)

S92: J/.j= .b
4 )= 2,72,S
t3c’2
/
5

V xE5’c3.7) 3x4625 z’z2- i. ‘‘J (3 I

c’ 35 , = 4- -‘ L4
v/= 71
,
2
4bzc%
.=”3

3Z5’7 I. > J(4i2= 3’5 17.

2i’ V j )1cQ’ I1-(k)


Appendix B. ACI Code and CRSI Handbook Predictions 226

Seoc4);

1J/ = 4375
2S, o57 i.° (—2’7)= 2.OS

v.= -x2o9ic.78Z4 2.3.b1r

, 4t=

SSZ- )

= = Z.O 0c 3414 =

ZWk=\4,
j3 25/, 3—l,7) 2.Dg
=
7 .i .
‘/e’ 2pS77oo= 2) 4
-f’ < e.2k:r

J= 44L= c2*fl25O(ktI)
5’- 4
1J/ 4314
z.,/ 517 LO, (- 2s)-=2DS

y,=- c ‘21.O
1
2p3,r Lf’ t-4-

= L-.5o5=- z.O

\/ *ae
2
_L 333z <oJ 142’(5 42. &?

vs., i*t 4L L4?’ 2’bLi4)


‘L
ç<-

= 5 42..9
3-’ <tGjj’4 &,
‘I L)
= =e1E4 O) 2.4)2.I5

34-
1
T =

tt= 4=&’r -k’J)


21
=
= 25/4= --G (,5—25O5k=

Vc30.’7
f
4 ?
Appendix B. ACI Code and CRSI Handbook Predictions 227

two—QvJ e.cct)

SSI. r( 3+ 3S7) 23.Z , V 4j4.4 23.2 37=’ 27.4-k

= V =z7.4”=
t6
ii

— zai., Vc=4 23.2s-+3q4 7544

=i(43) Z3J .
,

IJcQVt 2l(&)J)

SS+: 11(- 4s4.5)= 232 1*.. \4. 4


j .i 7-Z3,2t 4s94-
V — -

ES- 22. ‘kr


=

<.s- b(34275)= ‘ •J ‘4= 3

S(2 - S-i12) 240 jr., v

toX=• V’22Jt? o)

o o+ -a-bt)
a-JJ hc(a) (‘/ >

—ie—WCj e%L):

.
(4.•375) = t) 4
V 4
g
3 t(5E

+149)t° Ij V 4i3 434S 4z.:)


SS2: ba
442L’= 4-V -I’k.? 2lJ)

-- (3- 4.’5)O.t V =4J (-kzp)=

-J= 4V= *-‘P 2Z6 ke-)


S2 2Z ‘kt)

=s (3-4275)i, V4i o?27=4 ()


S-S;

4 =4V 5&’A-’ =sik


SG2SGi3: {3t4.625)= W°J y 4J t.ox4i o.4’k

= — 4-’/ — =
Appendix B. ACI Code and CRSI Handbook Predictions 228

4-k () eir1g 4
-itz

S)S2= 3= -: scs2=Ss=-

WGQ 1-536
L-kT)= 242(4W) r-
4
Ss5 SS,:

57J,(k)
= 7i.(-ki)= 313 (-kiJ) =

cGrz= SS: Sz=


4
t3S2 2__fop (:&p)
= = I3 c-ki) 4)J.pLL
= r)
4
4- ‘-‘ c&)

A:
I6O
i€wk:
$

44nn’, -L43-W. 47L1 141’o.

H=9ox4J7(44o—
234D)
ki —-

‘11-’.S) 7%.3 = z4.L1oJ)


0
)_o
j630 k a=o)z4r 4r9

r()= = 3J,. (.çJ3)

41I)= ‘-‘/b3 21kF))


d
iof__ 2(222561i)

©
ne-W’- k€O.((-I)

576 ÷(—. )Ciz4) I2-O ‘‘

7ojOPOZ, —-=
25
0
‘4 = (I51iJj -t-i724-x25) 2’4V4-S7 (*s)
Tt -r tiu-*S)

\4= 417c_ = fl3](-kN)

1coZ 2( I3-
-4-)
1 324(-kJ)
Appendix B. ACI Code and CHST Handbook Predictions 229

z6O o€-Jj

Sot

b’ 234Drrn)-) p -

V( o.i7-H7z+ )Z3&V4O =

9 Z’o44c o3C4}-1)

cj tp: r1,-crr’ -1-f e4)


heorC3)
o_j-
4

3q6 ‘/=

Wc
ane—Joj $ke-):

i@t

= .3 --
—j) p
(_34)52
—=

‘°‘ 2Z €-k-)
Q4L3j 2344O2p c4)
ØI
4
t -dXO1 eA1)

‘4= z)20 =C-kN)


Lt- 2qO- zI2O) )
4
boE(

eccL4-)

ei:

iS

Yo vZ-

c= ca-)= l,3(-&)

-z = z f(1ox+)4-*- 22 ---o =
2
oj
44
V= °-33J* I31& 2J4L-kI)

-
= 4 c> L,t-o), 3LJ)

= . a. 5- -. -
cyk
Appendix B. ACT Code and CHSI Handbook Predictions 230

+_ WOj &FQ3);

= 2(4W(I- ) --= is

= 33 4-?q2J33 ‘j- =
2 37 ji)

co-t LZ: L-ft -fr- #- 24-7 -

c- 2(3Z)j b2vt&1)

dt-. 3Ornfl, zP

U&-n., ,22q’
= —
) = 22k
7 (L)
,
2
A=’l ---4’j ‘W
1Z )(49
22.C49j (4o45—
z4S
41i 1)- 6&

‘2 7÷6) Z7’o(k)
3b
N0h-R((2): f
3 OP 3 --=
Ot2/
v
7(W) Q2O7J 23O?(3’TOj33Z (kJ)

ti\

— zL)-Y v24L+-)-)

“-‘Q): iot -(—:)vz1.= I.zTzmnl /az2i445)°r7-

Mt
V= (ai
g
7 ,LI’N) 737(k3)

çl.ot .irect Ph2oz;I) °‘

= l/= 33, C3.W—2,3S) =

=2.(3 7’j - 7z4xi3 )23t3o=qi )o.4*3. sq&= z244(kJ.)


1-eon;

oe -c c4-) sI,ot €tü: =3373°i1) I 35 2. OT7)

\/

I4-
t)’
Appendix B. AOl Code and CRSI Handbook Prediciions 231

Jo—J0’J

V03322ç[ A B171ss9 kJ) 4cQY = kiJ

3G—W l’et,%c7): (LJ/ = > a.3)


=

*i)o t 4L hfs AArec1’ti 0


b o--r(oo ÷)= 15O

c4 =2(4J 2(3)7+(kJ)

on rectL-
V 32ZJ2 x 2()

= &)j
4
2ll

c& f.1Z =

C:
2600

I t f

)
o
.

0 0 ii- = o4-
o NCQZ)(l0*&Z004fi5 tkiJ)
crj

Q 0
Jk( b= z-4tjJ 4cZ 4oo CjJ)
0—AcIj

aom’n j,
=°‘ =oooZ

ML4.

‘Ic (P7iV7 4-)’7ci4o= 4


j-2

(iii)

= IVL =

I24 -
Appendix B. AOl Code and CELSI Handbook Predictions 232

oe_w&ç
V/ç
cket1: Z6°)i 0
3Q/ ,3

V (O7 JoN) ° O
3
ZOc’i =3(-i4)

—tDAJ cJe&c (3):

2omn
sko —
, ?04_ /‘ =
V = 2.(O.1578J 41’z4xo.,34-x Z. )2&t)(39O = = zC’ LkJ)

c’L iV 43k)

Ofl€—j A&f)

çhrr az: j .a L35_25,r77)

V=.)(qIa= )7 L4jJ) <


,Ic..f= J
0
4
9
vV3
)
-f

-W SoW): bA z ft3’?o o)4+C4--3OV)CJ = oooo ht


2

\4=c22jxqq(kI4) jq&kt
ko—oj c11e(2): = nc )
4
GCøi

Ol’J0J heic3:
— OO--2lOO?(4O
—co-r1i€r.:O =&1,m’*., br(2b)4n
= fleo-2l
2.303
V,= q3 2j74?. 37 = 127c-bi) J Ir..LL), 4)o7—- = 2.o )

.20i2+ — 2SV)7L4Qt)

c2) m’t Vc 3322] L3< W71 Lki3) ‘l€iv


ce- j+
1
,
4

lZx3O+2W0x4
3b.4Y)ftt ,=--b.4-)= 3•w1.
7
;g
l2OD-i2’

= (-- 4-)c)y3fl7fli< 17L mn1 v=43zJz7 ,74 i,9a)

& o--

2’3om, = 3omM -f,= 33.4&) t,i?o.


4vc4.
)=43.z,o(&4-m) cu)= 43)44(3)

33(¼)
2i+
1
Appendix B. ACI Code and CRSI Handbook Predictions 233

Ofle—VJJ *e&r () 4ret.,v’:


0
2
4q7440)&3

4o/ 1’74

\4=(oi 78!+L7 Fx0o3xI 74 2s7Jc1Oo


220
‘Jot ‘Ic — 4’tRs) 3 I20 L4()’)

-i-;) 4p)
7
r
370

ot reetoi b = 23O
=
-- =
3S0
= Q7

7q374oo27 )z3iG3 -33z(bJ) o.z9or1J $23.)(3B0


1- J)
Lo oAre-$ V = 3Z =

tcQ 2134 + 33t) =

Ofl—1XAJ ,IqeL-rL3) 23[.I, P .Y


I’ M ==ZS3
j-o
2.>2.S

V= Zi57J -z4 27 3) z3I I $3o=zI72L q4?33J9L23IlX3&,= zd&)

ree4-vt V= 2JZ ---‘ O2(&)J)


22- O

Wt zCtc3Z+2(Z)=’

heC) i:7 LS—2.%,


)
1 =l.3
*rt rt.-:

V .27s S3Z

L-kM) — 4)
1017
Lc. cret-Dv’. v= i7 74d,1kII)

czt C’( 744) 47)

-+?Jo—1.d €QvTt))
!bi = iC
o%v my
2
\/o.3Z2j
o
33.c1o7i,p bC) )4L= V.

(1)/ A%so.71>0s-j 0

I24
hLG A30&AJ 1L3): Shor& rt; Z)+60 ÷r1)o —

V 322..j3G? I2’4 c3S0 = LQ3

2?(O3-+ ZI0€ —
2205
3107 LkiJ) co-h-Q
4oo 5&
9
pn Ar€1Zc’i-.; b= 2X4Ot

\/o
3
3 2ZJ *1Z32J s.44’c eI (4L1)
210
Z’!(1 101 r
)J9tk)
Appendix B. ACI Code and CRSI Handbook Predictions 234

S1JC Short ritt;on: 33on’n 4?


z4oo’43-i- o,c4’7q
4
3

4#.i (411 4ML&4’9, k41,!,

tar, 3O (Y’Y’t


-

53X’ S
9i”O, J 99= is177 1J)
— I

t=2-L2I37-N97)= 74zSL-ki.i)

frl€—WC i1eoc.4-u). an AIr€t..: 2i 57L,t f(4o?ci2l)


-
1 4e7A/
VAd/ 21

Vc(0.’78iTh --.24.xooIeo ) 2.! -4’7.1 =kR7L&lJ)2qoJ4 x/238. b4o79 = t


3Q43
ritot Y=7 --=f1t)

— 4+7S&ki))

$hort O1I(€et Zøt = ‘?73

v(o.Is7JI
1724 3
o
z
2
._L)
i
44,,(
4 /P &&JJ) <Q29O7J .234’x4-J I )o (*MJ
3 oreet
Lon -- = 474i)

2_= 2(Jo7)a7Ji)

O—WO *€cLF(3); sret 1?=23Iø.)mn, ‘ 1i4

’94)22
2
C3

‘4 Z5(°.7kW +7.z 274)2316i ll344-(*j


4
rtEoY ‘4= 2.o3- =

).JZ(J2Jo7±ZS3) l’fo (k)1)

oe—J Iheut4): horétw. /444t/ i.37

4
N = 1.37 247(&l) <

lpnecto- ’flbZ1)
3
VZ47

= 2C2.4-37 -r iloz.) = 9070 L4ü)

*0—’j 4eocc 2 [ 441)4 300 -4o7q)4J =


\4=33Zl4i.I Xl llTZS7 = L=
0
c V = 24-&)

±0AIC’j JLZ). (-7I = 3/ 7S).o5 tpt)


Appendix B. ACI Code and CHSI Handbook Predictions 235

2O4.
*(AJO_j 1..c-(3) 1’o•-t redt: 2A- tr -iT3- 9.w’
= L3(d7&k)

Ne 2(I37(1-1 )=qc€i)
3 -Q
35
lJ3rL ,)rea = z4b+ir(w
0
b Z I G ‘I23t,2v)%M

Vt = 3322J I23.2.L4o7. = IO7+(,JJ)

= 2o94(. I ) = 69e c-?J)

I 6O
+Ie1S+(:
reeefl: Aieomn 3BOl1Wt, -=44-1?

olE )377z (-kil-4v1)


4
2
17 3a.3 x4L
=

A 4emi
=
5
4OO
4gooSb (UJ-+)
Mr 44,(4— —

1s7
1\iL) = 5
b’!.% = II53(h)

t’4caQ ZCI 1153) =

Gn-VJQLj gLl)

,)=q33
4
,X
40
/(
j 4
ireci’ b=4”’,
- .74.
I4’ 3O
V (ol
S
7 23 &iJ

-> oo4CO 25& i)

NY2&L)) 1.J 22S&(1+

ort ree+t
4=
l.4A
-= i.sz
V
g.÷l
(l
z
7 4 jxt.2)4ot)L3S
‘4=
2.k3Lk)J)

Vt = 195 I ckw)
I’icZ 2L--95)’ 57zLk)
Appendix B. ACI Code and CRSI Handbook Predictions 236

one—ij& 4hecA (3) odrL; = 4oc’mn, —23,C3_2


>
3 2)25

\425(o.I573ji-l- 24 253)4,Bo....
L4)
re: l
t
v=’
coL= q54)2-tk)

SOr-tE recfii d/w3°/, S2)

jt I9 L)J) <

C-kF4)

1n4r’ect,1:

zcs) I3C-kt)

be.&r1 4 jq4- beoc c-fr-ik ez)

AL = o52 X 2Sx 3T+kJ)


2 A: =

NL4M32’ =
&.
-1ct= O.)L2Xi7. :

= 4,324 zC ()
! ]c.L

E C,: tO,ZIz20 7C-ki)


5
l44

Ncd i44J7 =
£S2 L
ii)
4
! (bi) =

qSz3X2ottIui)

L
0
‘J 4c b47Z ck)

4ki
L=
0
lL1c 4)(25

J??.L= 53xvisek
4-4-
2
7 kJ)
Appendix C

Predictions from Proposed’ Design Method (1)

1. etZ x.m?i€

I ° I t I ,= 4O.2 c, 4t, 3/2

i E1 I
I
I
I
ml I I I I
=

I L_1 4- I I L__J
I I
I
the eJt

-i=.
& -rt- ‘net

—t
r- r’
tL
0(2
I

e=c( ç
)
7 =3a2

j oi(

Q2=

‘= —
e—c=
600

—J’7.+ cf
= 4o3
S’vt2

237
Appendix C. Predictions from Proposed Design Method (1) 238

-tue eii ‘1t cj -tt c esi?e. +mct ok


Z ?- i.34 Ø3
0 = 30.27

— 5347 rA -—---
.77

- -r Qf -*k co 6€ s-h-d cd c0lrrn’L 0n

= ‘2?( Z. 6
COSS.15

h 53.b7 40
4-
r3 -.

= 22O

±kQ. W€r

9 VOL ‘

h — Z.I3’-+ r(L
7
T
T — 2.l.77

—n\_ OJ9fl.9 i-u

033(J43_l) 4
OCI.ii--—t)=Lo4

(S = Q.3L137— )=

o ccl x1rt. 2elL’e

Fb’ 0kf(I = i-r 20 0.32) o.af


(j€rw,’t)
CoS°
-c- .

‘_ ±. = o.s4f 52
— 2 cs4b.3° CS3’1 -

0A)2 caci toa. Qt


ve-co-t e-.d- of L’eI. -frut fm
= 9
C0s46.S
I75 ? 35 = — 333 x440= ij Lt)

o-%zL cao- o IicLn- -coc


4- . =
.2 — C) ct)
Appendix C. Predictions from Proposed Design Method (1) 239

coL -LF V 2q7,Lt)- 2L1 -kt)

2.
kkt)= Scmt. c 4S ,
A ,
2
ZCW 23/2

2,4-L,cmt, 1
’ 2J°

53.2.CMt b /= .7S,

/=i92 (j4vt43
.o47. =o.z bzffo.7f

Fv S4Stt) , F ii.aut) <-

— — lb.L’U L,W3(1)

1l11: C
= Z7.tt) ‘9ZI(&)

!I=.3efrt, A= ,
2
25ck
4
7
2.3’ ,
3
q=24.3 o<=3°, f—iZ.5°

b 32,OeM /=2.o4, (J)—’.*3


= (J)
t
1 .i4
o= 1S3S3, b °ft, f O.iYfc’
F= ‘3C)
ia= ‘z.F=
?4O.2lC 47/z,
rl,ft,
4.2J .O’, I 23.9i, 23.7

4= = 33.7GeAt, h/= 97 .

:—= ZIML,
2
b 4t/,=z.3. (f) =
(=L,lz-3, c=oo47, f3c-
E= Zr3.+Ct), F lqo,ct) =
v i4—z,.%tt) 4k

}.{= CAt, 4zcmt, 227/1, 3327 1r/.


I7.T. ‘L b3.,

4= 1
= 3.O2t7i ,
t.)j1fS
Appendix C. Predictions from Proposed Design Method (1) 240

2j4,
3
J-= tx=oc47,
t°-7
F I,1t) F= 4I.oL1)<

= 37.ott)= 3S3(-bi)
• f4
c
7 /
t L.ZCA?t 9= l5’ x— = —.

= 7b.+e- , =
S2.19cA’., /b’23, (J.)=4.3
)yzLs J=ts / =o.o47 =o42,f =51f
F =1-7t.1() E =
= 2f = = 37 CO)
311 -f emt, 4+7em., A-= =z7 /c4.t2, -
.

= I .4’2t , Q = 37.9 — 4°, 5.S O(- = = 4 .

1=°•- b,z3.3ct, -k/E,=24 (J)=.7


=j.9lct, (i*)22

k/8) J4=ZP (‘O3,3 ob f’O.22fc’

Fv loq.o’± i4Lt)= )t4h-’7ai4’JLt2

‘icQ 3F7• 7’)


)1= 5’c.’t, c4I,I ott, 4C.øt?, 443 /i -
i-’
O- 9.imt, &= 7.1 0
33c, qi° ci 19.r, q= 3V°
k=,bjøt, b=iz.1em ,
,

(J)=22
=scA-i
2
b

/=7°, !-=.o f3f


fv= 3.t) 1’?3Lt) 4.2(44= 794Lt)

ISO(t =4ss, =zS 4s/

= --O = =i7 )
-CdQ ‘= =
Appendix C. Predictions from Proposed Design Method (1) 241

=0c_1) 3h,cM,
OI3e, =4’M°, c=,3a°, 7l0, zZ71’, l3.o
= Z3 , h/b 3.b ,
l7I

= z7.’, h/=.zc ,

h/b = 330, Z,a2 O3


= Q7p
=
2
f =

,=Z0.7Lt)

3F,

= 7,n c&= L4ce, Ac= Z7b f;7z.’/c

S, O( 7.IJ CI 33.G, X i’7.7


h=/e-’, b=43M h/,=L5

= 1 (j3
h/ , =
1
=a.7 Oo3 ’ f=
—F o7b-f
2 f
Pj33LL)
= 4Fv iZ.41t) o25kFi)
Ac=*.ie 4S/
1= &= 41S
,
6 X 32’ a2S, c’22.7

h=l7e4 =4j,et , ,

i23, c])3
h/tvJ4,
r1 =-? , = =L79h, f=o4
= I37 = 2-Lt) ‘- i54()
kl. =
3cg kt*)

4z; 7cøi, cL= •qe+tt, p= fL=37 S/


CL=23.C41&, O(ZO.0°, O(-334° , ==i75

k= 4jmt,
1 ‘/b7
b h/b LM-,

J I4i çSf

I ‘L+. 3 Ct) = 4o Ct) I-u, 54a L)


4P = 4-7.2it)= 47’k1’))
Appendix C. Predictions from Proposed Design Method (1) 242

4: 9.oew. 3
c
4 -

0.. = — 3L,.S, = —
Of’- 4-3.4, _

ko4r?4.t, b=4z4cwi K%=t52 (J)ii


h/= 3)
S3c’$t. (j)= 3o
/b=L.44- J*=2.1 ) j=Oi4-h f=o.7I. ..

= Fh=.Lt’ -L= b.bLt)

JeZ 4Fv 4%.0C±)4LAC-ktJ)

43: let, 925O’).-, 43CAI 1


2,7cJ%q’- .
34-/c4

&= 3.3° o(=’•9° LX’1°,

(J). Lrp

b= ‘•
/b= 3.°4, (j)=3.
,
X=ki’7, =o33 , -,=l.o2$L’, t
=1f
— I41) Fk J4.9Ct) (4()

i’4P iLi)iO4L-kiJ)

IOOCA) cA 14$1, As’ 2727, -4bZ

& 0, +20 7.7x4b2oL)

F k4-s$ lz9.c-tz= I
gtt)
7

I1c- 4F 7iUz) 7aL-kJ)

4k j=IOOeM., 1
cJ.I.9e4fl
47:3ekM
-
2 ,

=ji’ ,=i7.9°, q=.z,.O° oizh4°, 42°


‘‘ 9334cmt. = 4g.57ct, j-)1= i7I
= -oent.
(J)
0.’3, ..Q1lfL’ -‘=o,aF
Pk= .°)= =l3tt)

4F, b361t)704)

=2flt As ,
1
toe.4& /c/t

j ob2 4=2o.7 22°

k -A-tcn, k/= ()=


%=2:17,
Appendix C. Predictions from Proposed Design Method (1) 243

24-3 (j-)= 2, =A13 °IJ

F= I&4-Lt), f= Lt) <Lt)


7
5O.

kQ 4Fv = Ct)z fflQ4d’J)

2wt. , 4!DJ
121fl.,

&I.32i.
44ø, O(2-I.°, 1 Zj°
O&!Ll°


1/,= ro, (}&)
c=O3L’7 = -‘ =o[Llfc’
= 25Lk) , F
NL = Ioo L41’) = 4t7 (-k1)

= I ., 4I?
Q= O2ifrt, = 37.0° , , = q’= 33,’2.°, o=3S C? 41
-, b= cj)=i
,
k/ qi,

7,
5
O=o2 (=.o -=‘fL

(-r) F 37.3(p) ‘ 37.7


‘21 = t.(p)= zt,(kii)

Al &A4: = 4wt = ,k 4iv 4WMi’tc, -f c9


‘ )4-it, Mis, O. 2r,k, 4 22., = t.c’, Q
=4ic2m%, =2o,7rAm)
1
b 03
h/

) ,

, oL=o)7,4-, 44f:

1
4R =z&€i)

A2 & A = 4 mM, 4o A= - = 4-to t4?’ -cL’ = 27. M

= 7- e=.’ o(= 2.&°, I7.’) O(29o, j0


1
L

=449,5bwt /=2t

15 .z mn’t, ,
1
h/ 21)

, -=,3’L =764-, -=t.07f ,


Appendix C. Predictions from Proposed Design Method (1) 244

p = = 227, ((r)) ‘1 2V77(J)

ls1= 4F— SckiJ)

)—4ionia =4’viflt 4Io4P,

O=-477, X

1 2° i97 1’32.

k= 45 ,
= z ,
)=
‘‘‘
2o 39 nI P1
1
2
/
1
r b (j=
319 )J.E=33
=
4
h +‘!1U, t,zb- ,
P= ‘24-iJ) F=.7Lk) -=D-.3L’kI’J)
coL 4Fv TfZ L)

or0Mi
5
l1=4 OmY0t, 1fm, -4o?-’W, f24’fr1
L,t O
3
I?i
O4°,
h 44mj b= Za4Sr11t, J-2 t7
=
1 ,
— •?z= L,Qra1fl) h/zy, (J4

,j-=3.32. X0’b4 4c4, -Ii,= c.36o-j

Fv -j) Fk= 7,z()27.7(k)

4F qit4ci”)

H=4°, 4oo1W1, A1T 21GM?c.

= 9 (,vnt, 44-9° o = i-L°, q = ia’, i34l,

h44.Z0t =2I.Mt’ h/k= 2,fl , (J)=


k/
,
1 = z7, )= 4’
k/= 42 3Z, = 0.744, 49 -= io4f’, - = o3’7f
= zVC4&) i-

= -F = ‘o’(-kIJ)
A7 1 d=4ooit, As -4io-1P, .L Z4’ç2MP
450 rrtt
430 (4° S
4
OVLW(, U c?—l97

-h= 443.Un, ,= tilo.44flt .,


h/
.
2 L} Lj)
=
1 h

(j)=4o

=S- -R’=sf’
Fv= 27 (L), F=7 1
)2277(kO s]coL
F
4 / 8(&tJ)
Appendix C. Predictions from Proposed Design Method (1) 245

*=4oiwl, d=4omo 2
A4 c=41oHP, f2SMPa
= ot, cp= az-’ 32., 12
c=
k — 446 P = ) .3 ,

= (j)l - c
k/= Z.7’(,
k
, (7&+, 49 , -i.øE, FO3

F= , F= 227.7LeJJ) 4F cc4J)
H=.4om1rt a= 4cxD’niifl, A5n. ‘4-io HF, -=1ZHF
ci= zon’w, =4o.°, -•‘

k 4OQfl1fl
1 I2= 24P h/= t9q)
(J) 7c
=
1
=$32..wm, /= ,
4o

=ou)cf, f=of =o46f’

1e 4kJ) , F 2.3Lk)277&k))), k(1&)

Aii )(=4omnt, c=4o-mit, f=-myvj, =4-i&-P -yc:= 4i1€


4z3.I0, o=S1-.8’, =—t°
2
q?
O=4• O(,=2b,

h = 4o8.o1ii b (j) =
z43tn1M) ‘/,=I6S, (J)=4-
h/b = !, = 3.3Z, O( 764 = O27 ?.

=
1
R lVlLkrJ) )

kiL 4Pv =4.qw1 = jJ)

At2: 1 =4niw Ach’ f-


4omi1
&7fln) 2&°, 13J° 37.7°

k= 44.4jPint = II3I, k,
,
1 = .u ,

€,.511%t,
7
= /=2z, (f)?4-o
33332
h/ c=.7
= zzC4iJ) , , =2cL?,L-ct)) j- zzY].7(-o)
4
F
iL4Fv4-Z2t’ gc-ki)

H=4omm, =4oo,rn1.,

1= 2.CO1M-, Ø =4%,
1
c% -5°
1/,=
1:= 39.7Sn1fl = vzjo’
1
b
Appendix C. Predictions from Proposed Design Method (1) 246

‘2nt )a3.7c
/bi37 ,
+°9°-, =33c:

F,v V52L-kIJ) , = 78 8-Io)

jt= 4=iooS(kJ)
-(=4r3fl1M, 1
t.=4ooYnnt 4ç= 3331 3b3pU’

= ‘Z P° i-L= A- 333.2 440 = 3k


= 4J=.
02.S (4I)

£SI j= d=47t. A 4 ‘r3°? 43’Y-


cu=o , AoPS4a43=.It2)

P= Si)ct4t.2

C4)
• s 45 4., GOI f 97b7 $€
a=o 4 tj
l- DOt)(9,tJW1 = .31-kp)
R’-LO= t&4I.g°= 47t2)
)= 4ck)
-.kF=’ .4d
1

‘r1=.in, 436iv1., AOP7OS4) Z3OD1S, f’43


o= ‘ Q35.S°, c=47Se, q= 67°. c°

4, , h/ i.’v
h= .
,

, (J)
=
2 °
.t-z, = , x= 0764., (3 o39 , 39 4-fL:) fc-z4f

IS— b-?) ,
9r)
t’i= F’)= ck
S4: -= t.. c As°9 ro99?, L:=436F
_4,9D q,= c= 32°

# t., I = 4.9 t, /= 3;- 147

b (j4
I=3 OO76+, -= a4 ,f’ow-
F=Lk) ,
Appendix C. Predictions from Proposed Design Method (1) 247

1t=, c&= A-.= O2lI 4-= 72O p-


,
1
k= z.i4-’t., 3 = 331
a = 4’T.°, 4= 7.9’. c. D
0
f=
= =
,
h/b, t.- .

,
2
h/=
()=+o

/= L7 j= i c=o.7M-, 3’’ 2-8, -f=ciS4-f(’, ,“cz4fj

R
4-Pg = 33.& (ktp)’ (-ktJ)
i-= = 4Z75 A=o,o13 2 9z43c
4o
B= 331 =

15lat., =
1
b Cj-)=irj
21
k/. (J=4o

1>_ i.73 3.31, O(=O.7LA-, o.S+ f=4-’

= , F zSkj’) c

ki = 4Pj L&p) 14-9 C4J)

-=
1
k=4-snt. = j sz -r
LQ.3’Z6 43=?z.5.2-°, q=i, I71 =—2.°

1’ 393I

b= -71 fl•) h/,=

/= I.75, =332, 1 (3= 437 , -=o.S3-


cc=o.7”r -= o.-g+’

F 4(kF) = Yt-kp) < r L-&)

JcJL 4P’-° 1Ick)J)


= 71

t’ 44ovr&, ,
2
Ont =471P .
0
=zr
a.= 0, •= = Asfj = qo 47q =

= 4-3,.I4’1.6° a7c4vJ3

JcoI=

1tZc
t39Qr, 4’i9-,
Appendix C. Predictions from Proposed Design Method (1) 248

o.= a , —
vz,vx47’

F 57
(
7
= ii.J)

1
= zq7 )= IQi’J)

;ht iret-o;

-)= pii, ct= f: 4r7 f7 frP

. -(,i°, A’c 2o41 5’4L?&)


= ko-.t& =
.$q7(I-i.

..hoi.tr a€Z-
H=Ynnt ô3On1M-. Z1O)1VJ -
=
1 H?R, fo.siPo
i7O) a
11
a=iO’’, O=S7.2°i

b 3I2.O 1
h/
h/.. cJ)=3S7
h/bh92..J r3i4i ,

H4t42J1) 1 h io4+1kJ) 4
j
= Tq4(
T
2x —i- Z3’Lo Ck,)
Skoit dreetvi

1
-f=Ut’irnn 1
-1riWt 14? ,
M?.

j=4Zk°, c=41.T°, q 0
g c4°, %= •
h=4S.G4n11vt, b,zb.3rn1., ,

z.nwt, 1
/=L4-I, 7
(J)=3.s
3
,j= P ,-.°4fL’
O3i -

F1o(&t)< sc
(’kJ)
7
± -)=&t)

t=rIYtn, k=9Oflltt A Z4’Z 3-j?


1
&—)fl O• —25,

p)2 ,
Appendix C. Predictions from Proposed Design Method (1) 249

J4 (Q242, 7-PL’ c
= 0tL&) Fk=
WL vb44(i -- 2O) 1774 4eJ-L.)
Appendix D

Predictions from Proposed Design Method (2)

1 Freet 2tfl

A&’ C4t’,
1 -= 1.’s /t

35
t f44o/c fr4
.
9 5t.1t

“ ffL •‘;

.4
L_1
-&=-:;-=
coV= zAPj±o4
= 2 .4O.Zl (44 - =34)
=
‘ ka.r fr’+k”
C.oLan: c=33(-L)047

-o4o-tiqo =
3
0b03J444 275.3

1iLRL%) 275.3 )C 3i
= 337.2 Lt) = 3337 L-)
= o.047 33(—)o.(S7
o4 x4f-ij -+ x o37 = 5/c_i.2
I irr.j 600
= 26.6

Ofl.-4)0j ,
ieo..fi)
4 )4-33 () AiieJx .

IeLt3= isf33(f&)

‘a.

250
Appendix D. Predictions from Proposed Design Method (2) 251

=223/,

= = z4.Z52323. .Z /‘7tt) 2I2 ki)

=-4o.2its,
5
A = 4gg/- =4 ‘=

P-
0
4 2A-ci’8
=i’ .,
3t
L
7 t)c ()
2’Z A4O2Ie43, 473s/ctl, A1

, 4.25&, d= .4tiM - 5L.2CA4

= 2 z7 43 L±) = 43 kIJ)

A 4køt

=
0
j zs41 7Lt) =

4 the.- I c4e)= 4 + =

c.c. =441c), of vttkMk Crcl—= o.Sciii


c-e- t=
=7sr3/eA.,t.L 447ewt, +68jT

— 32±) 325 (-k14)

N;
= x445 - 4”(*)= 1- €-

A=43 27/4, =7O2, J ,

= z.S39 59 Ct) oO (ki;


=6w
! A= 4.ld tj
34o4-3L,Lt) Hlj
4
caL”

= 7”’ /‘- = Wf.+


-

coL 4At47’ t:;= o(j

J
Appendix D. Predictions from Proposed Design Method (2) 252

=3y-, k

-ic2. 43.
7
t Z3 x 4Lt) 3(’k)

bZS:
4 90c4
0
1 Lh 7’
42 ,4t) =

43t =43.o3, 1
o
7
=a
S/2
‘2S ,
KL= — -) Lr;2c.g iJ)

)•‘‘

4M4: A47c Z91/

44F732j4 —-j !7c±)= ri4i4Lk1J)

/‘ntj -t
ct= 4c3 =

f=
co) 2I.2’7

No.
=
7 C-)
A1’P4: 4iofr1? , -b

eat = 1o3aL4)

A2AS: As=t’’i fj 4toMVt ,=4cDri =33’,

toSi(k)

=° Mo 4X ni, pii

I =A4=A7=i=MO AW’AZ:
=

AA1;

h1V, j=(2oo-)JZi2.lmfk.

17
4
4
44oJ)
44o—

A
=
Appendix D. Predictions from Proposed Design Method (2) 253

= : = T2 ?- ‘ - i,

= 4. og444- 7.. 72430 2! .S 14’) =


‘ 7 (&J)
ft’, 9=
43Q4
L

4=)23
1’
2) l44Cd-)

A— Th 4.4t 4 , 4A

-1t= ?‘-°‘T s90’77o j.!t’) kJ)

= o.2fl , =7zr 4i
4-Lo2flx92os.
-[ 3.1(r)=3t)
4c5.)
A33-?,
39’3)L9ZOj Y
aç J 42il-, 141i
4.!iz-
MoL 4?cz2 ,6c’ocD(

sqS: Ao.-’ -L4l4,


4o54-,o’eD 4cfl 122Lr)E4t.k1J)

Ot3ThJ .ç= 41 fri?, 6{=44o’, _37m --


440 534
2. X9G) ‘ 49 = !O Z L
- )
4 r.(L)
4
l
!012..,L = 433 tl€ii)

c=1O1I-rk33I4i-C)

J395mM, ‘=----

2acz 1
Ic

2GOThM.., jOT1”, h=37m’) =_

fl
3=
-. 32 ck1)
Appendix D. Predictions from Proposed Design Method (2) 254

di
ZA1
32O
=OYt)

4°f 31]r5 234C&, ZrlJL)


= 23C.4 LOO1JJ)

eCtO

/=3,vrrn, 242441P 4-ltii 3flmrt’)

330 = t’! LkI’), zJ


Jtt
1 — 3

3DJ S3

A=Z4’, 3,-I?€&, 375P.) ±0A

L2OL41, 4LkJJ), 2,L) 234

L= 24i
0
I’ic i)OS LJ)

seJw ca
=23/i,
CO1fl: ,
)=°“

IJ_ 33(iY0P47 ,

DO4 O3j
-f= o6x25S t CILI) ki’1)

&2.355 Pft47 J23 N,nz) 35O.9tY344otk)

ane—ij he-orLI) NCOU3*

lo3COi)
=MoJ/nt, 45At

, (37= oi37

-S, =obx44
1 -\3 04 .bO3o •2’7:3 f/ 37.2;o7(kt)

o,Arc oi37 4-o


- =
- 2Jb. WZL)= z. 3x2 31t)4)

n—w&j netC()

gt= ef= 1433(&N)


Appendix D. Predictions from Proposed Design Method (2) 255

L= 27g d
=o33(--_-1)= o /,,;-)
c6c3oT, =33(----l)o3,;

— 31. 5=222.7L) 21 “)_ = Z7I X5= 429 (;t)= (ê)


nei
22 - = 4
O3 —I’7’7 =a.b4*o C4f’ 7
j+5=.
o=o”47, —l) 323, 323j z71z i/
oc)
= 3z ct)o1o ziZ7. o eki)
7
4,t)

O)1— JAj sh4) nt i’ti irt f ,2J 3I5Zj- L-&J4)

= 3a
7 . 8.4 e
Ø=bO47) ‘Q33(- t)>i.ô O327
212.4

7
4
ct—.O (-n.— 1)=o.i3, i-3 k,( W =295
/‘-

N c> 22.4 3’ zo,2.L) 22(&J), oi.°ti) = 4ii (k))


Ofl-JOj ShQRC(i) ( ptLCLtbL)
, N.t 1.L= z52 Ckt)
4t7 k/

=p4’7, H)11, q /2

; 2ZI.C Lt) 7oi4L.)


cottI)
N7)=

ofl€-M he’O o- apt,b)

C Zoitc Qf Iun= o.t)


f’= 45 l/ 44]e, .cLftJ-E

33(Ii)O’2cl, o4S-1- P’ 4!b2

Ot o.33(-- —l).2) =33(4_4)0Q4.3 1’ )(OX4O43T-. flr?

f Crt.& 3’1.541)

= 3.5Lt) = 22(kJ))
7
3.,x4-. N(L2) =
3x

O€—l-4 kc(i): C.L(3)O(ki’J), —‘J SJ’cLt)- NcL4)72Z(klJ)

HC I.o1&)i)
Appendix D. Predictions from Proposed Design Method (2) 256

= c4I, 33(— ) O3O 6s43. c4Oj435323

3(-E—--D=op7, - 4S3. xoo7Io,zz’


Ui)= 4-7 eiJ)
1 —2774 ,?35 o2&(t) OD7cLJ)

O—’ r’ -4j—Jo-4 eas’); 4.(4it)

tL 27O7’)

3N3; =7om
=4i )=o2, -fb4’3- .O52j 44,3= 3773
O.234, (- —I )oz, f=obs.43 + ?

ictLu’ 3773 kS 7+.° Lt)74


3(),
7 L2) 3O2.b’3 II ±) (fJ)

.cheo--u . €o4’t) 41o

4L = nnn t— t74 1-ki.i)

4o = fl3,e

(3=o33(-—)=ob2 OJ,4O &‘-j i/’.

OzL3.3 33{l o25, - fthX4Oi -PP XOZ3,X 2T z7!.4

= 7004 t)SLkIi)) 97X2 (-)


A)?)3J he’-rc.’)

CrCO =

e*-c o4 voJt y
-= 3z.c = 4-ent

O33(1 )Q49,
,
33_)oZ3, 33(-——1)’0233, _ =o)L37254 z3233 j3

1
t
7
33’
AtFi) 4z4-3)L3I-k±)
‘rt) N/ , o—’w e3rC) tJc.Lq.) 1k)

g , ;f3’/:

= 3 rTt; 2& ‘O)


r
23
Appendix D. Predictions from Proposed Design Method (2) 257

)—)L s-e’-. ( /) 224

= J) kM)

42:

= j) ?33(4 —)= .44- , --


3L7
, 373 22 ‘5922x =2’23
3

2
tb

= = = 432.3

(-ki)

= 3b’(k’)
bS
112

09 G33( )O# 3 &34i \ 5°4r’Th3 -‘z


0Z3t,, O.3(2- —O23 023IZ32Yr =g-
j
34z= ;.i= ±22. 35= O9)

tWO—i)) c9(4)= O(J)

x=o.’’5S, 1
3 3t— )a.732 , x5 j= 4• /t

tZ3b, —I) 44- —


2
-fb 3+IW- P> 23 443 I244 /e

424
l’icoLU ‘.tt’ 1 4ct)
S

5
1
c 5 ‘ie
, tobo4(kN)

4’: c9a.Seh
1
c=oZ 33() 9
4
Tb°’ ‘.0 =
I7Ø7 /t

=O,23, 33(—l)=O.4k5E, =o4-t tlP 23o44f


j/
= I33()) = 4X-4OS 3” ’ 5tt)=j353f-)
T
Ibb

3OV’1 €r1C): cc.2(i) 7o24C4zJ =


L ,

-o353.4- 5
r
45
o7 44 J/2

CO23b, °.33 — lj=.’g, f =b 354-p 243g}. ‘z.4.2


ccLL) 43 i22LLt) f75
(&‘))

1
4-zLkN) := r)
Appendix D. Predictions from Proposed Design Method (2) 258

44: 2 42 F/f

5L

= 32=3(-j_1)=a444-, -,&4z-- — 24j

o9,5= 1
t ?) 4s24-.9 3= 44ct)=

W{’)- c-tt4) 374(), —

-4p-k,

fl3>ç}>.)
3
O =
0ft°33f XIbXR363 z’23 ?$Z

ij- (&) 32(ki’i)


7
1 z 4okr)
= 21(kIJ)
&rt) 4c-fcy I31 L-kN) = 732(kI1)

F=3+2.p

ci= .4;z
1 ) 34
3kZ234? g
3
‘47
o•3! —i)Q 2) p$
NCØ-Lr 3ø’] I3.t
c
4 p) 9z(-*jJ) , 2 -G’ 321k) vzLkt)

Se4-(I) cett 2t-k)J), Nc= ftJcJ.

A1&A1- -9=aQM?&, c=4oOmrt


ct nie,r e vi€,,f
o33(—>t,
(,= o33(— L)O.3S (

=e•33( )2+3) -t)= 033

‘= —j-(, $ 3SJi = 3.5 M?& b 2 G33o 4.3 M1’

= 355I2O’ I’f 20 (4 t’ ct.) 4 c *3 K2= bo kij


-—ijc shec(I)- —o- SeeL3): tq’ (-ku)
= ii= I4’ZoC$)
&A -

(13S,

=24.5 o33, -f=-o.x27-i- a24So33=p&


rfr,()4) 24k)

jo—Jc Sr(): 44Lkk1) . hAJc sbexc3) 2ZS(NJ

mJz= 7:(-kt)
Appendix D. Predictions from Proposed Design Method (2) 259

HP.

Oi.°, , -

.) =a,7c3I.L± 24G7933Jñ —

4&, d—
2 coL(ij = ZhLk)

+J1”- , iO)’--i Sie(3) = Z432CkN)

j,,= a.33S, -,= ob-- &3j’2

= o. 1
f )c24.9 £ z4xo:3j .9 MI
icLLLY 422 ibSG4J) , 4cL2 I.4-2= (t)

-kt*io-tj Seut3) Ns 2?44CdJ)

= ‘rf1 =

=‘2&OM?i, 43’-,

o(i-°, X2-4j Jn 4i.2 I4?


24Ø, ‘
+ xo2A 33jzh. = 11

= 4i zøo= -S (-ki.3), = 4s i3.l x 2OI5 Z2Y

kUU) AS4.) V2±4}J) *30-WOj 5kcL2xt3). 22/ (e)

CcJ lkL?N)

L
t + =
=oz4z.-i- 3’.3 -i?o.

035, rab4Z + oa433J


320 (ZC&4)) eJj) 4&9 ZJJj
tL)=

—-- r): tL4) 3t3(k)} j_3VJ z4Lk)

=z-, =4e
=I0,=38, =4z1M?
=a.24 =O33 ,c2,.S-+ 2433J
-‘
=1 V?t’

/ 4- TX2 233’L&)

.32(k •O-0-’ ((1.) ‘-AcJc)


Appendix D. Predictions from Proposed Design Method (2) 260

= 7r = 4g4
A1O: -= 4Xi

= IQ, =oS2+

(O3, 4= o.2-t b

Nc) 32.4)c2’’ i29t&) , ‘4c’(1) x4xL


2
2fr ,3+Ck)
GAL
4 Sb_tLr(); )=ar22J.i)
4
L -) Sk€cx)

= = (4v)
All: -4}-, 4t

G()0) -f,= 74--i-


=o2J
S
1 , (3= o.33, -f ou7 4-i- b$O.Z4 o3jE4-j2. 4?c

2 (-1) 2 4-c= 7I (-k)

Jt) c 4)Zki, e-): aC-ki)

Th1- = (-kN)

Al2; ‘z.Sl?&, =4etmrk

-L-°,
—= GX2,-S3SE 4o. fri.

2=3) )(Z3 t 7b
‘ic-&L) 2DO lzoL&), — 17 —L2= zzt(ki

-); NcJ4 -+UjO—&A3’ -r) e)= 2


)

(kJ)
20
l2
11

c=zTM =4on

‘ =ck33 33(’)33’, =cb2ti - =


7
xooco38J .s
L33Zl)) =

= qzci)

= (S , , 3--(C 3’33= 44?

“33, 4f7 HP
2

- =‘
Appendix D. Predictions from Proposed Design Method (2) 261

434y , o22

4.;r5
51 11
43gf 1
aj,35

= -,- PZ/.X. otpst) =é43+ 7ZX2Z(°I53j4 2t)


32=
oI 4Io(k) 2(kL))J Nc-b

o— ,hetct). lC-(4)= 22.LkJJ) O0V€ tfL): 1c.t1> ZZ()

434 -?

= L?) 2 P)

WLI) = 4-i. I tkp) 3(kJ). 4-x3 gJ.t-kL-) -

—tuo-ij heof = raz ., -o—-w he’—t3

‘i S Qi’22(tJJ)

33I)O0 (?
a

1T7 ‘‘ (
2
2J )
b

_eArL1) 4c.Jk)
0 )2I&kJ) , te-W0j eL4-3) CI) Z2, L&J)

Th= ZI4ei.))

‘: c( .0 , O(t

4.1i.,

+?Z$ 1] SC-) 5V +2 2lFI4.5 = 374 LP)


N L€fl) — =4tZLk?)2I t
)
4 ctzy4 ks’74I IoS(k)=

ShW1): QI.I)
4 JCAJ&J.5h1X(3) CL5)2I()

SGz= s3= -=
.2_,c7 p • y =,o, x= o-z

= 4o& LP) q7 j= o7 Pit)


7

40X 2 (4p) -‘kN) .IckIz) rø 4Y7


-, 42k)=

€‘-ft
l’5)
c.-=
Appendix D. Predictions from Proposed Design Method (2) 262

oc dr3o; Lt+,

=
o4&, = —2S+kW
2)
,
27(?)

x2’== 3Lj()) iCO.e.L) = 34.(V± I’15 k)

1Jc) 23 (-kM)

Oe—ix eJ) MZL)= 32’fLk), —4A €X’) .J.j4) 26(kN)

hearc3) (-kJ)

= =

2 ,o(=’ 31O
= O43Z z÷432j 27.C M?)

= o, = 23L1P&)
= )zo t4vJ), L) a- )=
a X ceQcz) (I t 2Z.(4t)J)
7
IZ

j
3
[ i, 3kJL), -±wo-J&j shxi ) iLL+)

— — 2,4(4UJ)

i’v1l (&I3)

SbGrt t: 2j M?m.. 9e nt
cO33
I
2 )> o3(i)=o4z. f= Z7.1-i-xt42jj= LPR)
sq -
°,

34-I c-ku), 34- (i-+ lg4-ck)

= 2.I A2( f2ø= >. = A-° (I -‘- )= za (kiJ)


Ofl—t’J 6h’erU): jy= 3o4jJ) ——j- ie.tr 9kJ)
‘wo—t3J shzrL3) 3-±)=

b1fl $JL

3-IP,
= 03Li)3ii2,

c(= 1°, =o33(-j_l )=o2,

= ;4,(-k), 4-C-)

WD= 7(ki), c)T(—


Appendix D. Predictions from Proposed Design Method (2) 263

—) fLf:) c )37/€?J)
3
( , u—WM Sare).. c4c4-) L1OJJ
70
r
5
cL) k
= )11I”
{ Mc’k
r
1 (kIJ)

/=oc44, (iP)

&z= t — 3° (1#)
= 4-I 3â= (4z)J) I -+- —‘ 27sqck)

= 39.0 ) 2 ) *20 2339 c) NkL2-) -÷ = 347 (-k4)

hEzVf(j) 4c.-t) -il)O_1Jj IAau.

-to—iuj &1eaJ L3) ‘-co.L) 9cD ks))

‘J--t ii’fJ z4L-ki’)

rtt\ 03M’& O1Yt

40
= 33(_t) o, i)=0441, 33’1
3%.
3_)2T) f3o.3- jj=’zI.4 (4&)

Ncts •7s’= C-k3I),


7
82 coW 37(I + I3ok1)

IJtc)
4 = 2I42ZD r34(-k), M( +
5L-kJJ)
7
q

&r€—IAIcj ec3rLI): -cC3) I104.(&J)

L
0 iI4= ic-ki-

You might also like