You are on page 1of 182

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/320840328

Design of Transfer Slabs Using Strut-and-Tie Model

Thesis · June 2016


DOI: 10.13140/RG.2.2.31899.36647

CITATIONS READS

0 13,751

1 author:

Mohamed Ibrahim Metwally


Cairo University
5 PUBLICATIONS   2 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Assistant Lecturer at MC Academy for Engineering & Tech. View project

Ordinary Analysis vs. Staged Construction Analysis for High Rise Building. View project

All content following this page was uploaded by Mohamed Ibrahim Metwally on 03 November 2017.

The user has requested enhancement of the downloaded file.


Metwally, M. I., “Design of Transfer Slabs Using Strut-and-Tie Model”, M. Sc. Thesis
in Structural Engineering, Faculty of Engineering, El-Mansoura University, 2016

MANSOURA UNIVERSITY

FACULITY OF ENGINEERING

STRUCTURAL ENG. DEPARTMENT

Design of Transfer Slabs Using Strut-and-Tie


Model
A thesis presented by

Eng. Mohamed Ibrahim Metwally


B. Sc. Civil Engineering, Faculty of Engineering, Mansoura University
Demonstrator in Structural Engineering Department, Delta Higher
Institute for Engineering and Technology

Submitted in Partial Fulfillment for the


Requirements of the Degree of Master of Science
in

Structural Engineering
"Design of Reinforced Concrete Structures"

Under the Supervision of

Prof. Dr. Assoc. Prof. Dr.

Salah El-Din E. El-Metwally Mohamed El-Said El-Zoughiby


Professor of Concrete Structures Structural Engineering Department
Faculty of Engineering Faculty of Engineering
Mansoura University Mansoura University

2016
MANSOURA UNIVERSITY

FACULITY OF ENGINEERING

STRUCTURAL ENG. DEPARTMENT

Supervisors

Thesis Title:

Design of Transfer Slabs Using Strut-and-Tie Model


Researcher Name:

Eng. Mohamed Ibrahim Metwally


Scientific Degree:

M. Sc. in Engineering (Structural Engineering)


Supervisors:

Name Position Signature


Prof. Dr. Professor of Concrete Structures
Structural Eng. Dept.
1
Salah El-Din E. El-Metwally Faculty of Engineering
Mansoura University.
Assoc. Prof. Dr. Associate Professor
Structural Eng. Dept.
2
Mohamed El-Said El-Zoughiby Faculty of Engineering
Mansoura University.

Head of the Department Vice Dean for Postgraduate Dean of the Faculty
Studies and Researches

Prof. Dr. Ahmed M. Yousef Prof. Dr. Kassem Salah El-Alfy Prof. Dr. Mohamed I. El-Said
Zedan
MANSOURA UNIVERSITY

FACULITY OF ENGINEERING

STRUCTURAL ENG. DEPARTMENT

JUDGES

Thesis Title: Design of Transfer Slabs Using Strut-and-Tie Model

Researcher Name: Eng. Mohamed Ibrahim Metwally

Scientific Degree: M. Sc. in Engineering (Structural Engineering)


Supervisors:

Name Position Signature


Prof. Dr. Professor of Concrete Structures
Structural Eng. Dept.
1
Salah El-Din E. El-Metwally Faculty of Engineering
Mansoura University
Assoc. Prof. Dr. Associate Professor
Mohamed El-Said El- Structural Eng. Dept.
2
Faculty of Engineering
Zoughiby Mansoura University

Judges:

Name Position Signature


Prof. Dr. Professor of Concrete Structures
1 Faculty of Engineering
Yousef Ibrahim Aggag
Mansoura University
Prof. Dr. Professor of Concrete Structures
2 Faculty of Engineering
Salah El-Din E. El-Metwally
Mansoura University
Prof. Dr. Professor of Concrete Structures
3 Mataria Faculty of Engineering
Ahmed Moussa Abdel-Rahman
Helwan University
Assoc. Prof. Dr. Associate Professor
4 Faculty of Engineering
Mohamed El-Said El-Zoughiby
Mansoura University

Head of the Department Vice Dean for Postgraduate Dean of the Faculty
Studies and Researches

Prof. Dr. Ahmed M. Yousef Prof. Dr. Kassem Salah El-Alfy Prof. Dr. Mohamed I. El-Said
General Administration Library
Thesis Summary ( ) Library
Faculty: Engineering Dep. : Structural Engineering General No. :
Name: Mohamed Ibrahim Degree: Master of Science in Date:
Metwally Mohamed Structural Engineering
Thesis: Design of Transfer Slabs Using Strut-and-Tie Model
Summary
The need to create transfer structures has become an important and necessary
matter in many structures; for instance, high rise buildings, which could attribute to the
solution of the problems related to the different uses between the upper and lower areas
of the building. These transfer structures may be in the form of transfer beams, girders
or slabs. Transfer slab is considered one of the most common transfer elements which
could exist in high rise buildings. In order to draw a complete picture of this three-
dimensional element in this thesis, pile caps are considered as a simple transfer slab .

The geometry of pile cap doesn't provide engineers with a clear understanding of
its physical behavior which is considered as a complete disturbed region and there is a
very important volume of concrete that is subjected to low normal stresses and
significant shear stresses is called "inactive concrete", Strut-and-Tie model, on the other
hand, can provide this understanding and hence offer the possibility of improving the
design. This fact should be reflected in the strength assessment of the nodes and struts.

In this study, a three-dimensional nonlinear finite element analysis has been


conducted on pile caps that had been experimentally tested and the output results of the
cracking pattern, deflection, failure mode and stress distribution from stress trajectories
(that can't be obtained using the strut-and-tie model) are obtained and compared with
the experimental results in order to verify the validity of the FEM procedure. The
obtained results demonstrates the accuracy and power of the finite element method in
dealing with problem of pile caps.

The method of strut-and-tie model has been applied to the selected pile caps that
had been examined with aid of the finite element. The spatial nature of pile caps has
been reflected on the strength of struts and nodes of strut-and-tie model models. The
obtained results demonstrate the reliability of the strut-and-tie model method in
obtaining a lower bond estimate of the collapse load of pile caps.
Keywords: Reinforced concrete; Pile caps; Discrete model; Strut-and-Tie model;
Finite element method; Inactive concrete; Confinement; Cracking; Failure.
General Administration Library- Mansoura University- 60 Gomhoria Street,
Mansoura, Egypt. Box: 35516

URL: http:// www.mans.edu.eg EMAIL: mucentrlib@mans.edu.eg


ACKNOWLEDGEMENTS

In the name of Allah, praise is to Allah, prayer and peace is upon the Messenger
of Allah, the prophet Mohammed-peace is upon him-. Allah is the first and the last
who, always guided and aided me to bring-forth for the achievement of this work. All
praise and thanks are due to the Almighty Allah. There are many people whom I
have to acknowledge for their support, help and encouragement during the journey of
preparing this thesis. So, I will make an effort to give them their due here, and I
sincerely apologize for any inattention.

First and foremost, I wish to express my thanks and gratitude to my parents,


the ones who can never ever be thanked enough, for the overwhelming love and
kindliness they bestow upon me, and who have supported me financially as well as
morally and without whose proper guidance it would have been out of the question for
me to complete my higher education.

I would like to record my deepest gratitude to Prof. Dr. Salah El-Din E.


El- Metwally who, very kindly, and generously, devoted much of his time and
experience in helping, guiding, and advising me. Indeed, this work is the outcome of
his great continuous efforts and wide experience in the field of structural engineering.

Also, I am especially grateful and especially indebted to Assoc. Prof. Dr.


Mohamed El-Said El-Zoughiby for his constructive keen supervision, fruitful
criticism, continuous support and encouragement to complete this work, Above all and
the most needed, he provided me unflinching encouragement and support in various
ways. I am really indebted to him more than he know and O Allah give him much
better.

My special, profound and affectionate thanks, love, affectionate gratitude are


due to my wife, Om Khadija, who has been struggling with me, hand by hand, to
secure and shape a brighter future. Her harmony, support and looking after my Child
during my study all stand behind my success. At the same time, I would like to express
my love to „the beats of my heart,‟ my child, Khadija, who is the only source of
revelation to me, and her love and pure smiles that have made the hardship of this task
possible. My deep love and thanks are due to my brothers, sister and the entire family.

I
Acknowledgments

I do not forget to express my deep sense of gratitude to Dr. Ragab El-Shahawy,


Dr. Saad Mouharm, Eng. Hatem Khalifa and my faithful friends in Delta Higher
Institute of Engineering and Technology for their understanding, encouragement and
support. In addition, I am extremely grateful to Eng. Ahmed El-Nady and Eng.
Waleed El-Demerdash, who taught me how to model in the program uses in this work.
They are not only friends but also brothers.

I am immensely thankful for the sincere and brotherly friends in my country


who accompanied me and shared suffering with me during my study.

Finally, O Allah make this work purely for you alone, O Lord of the Worlds.

Mohamed Ibrahim Metwally


2016

II
ABSTRACT

The need to create transfer structures has become an important and necessary matter
in many structures; for instance, high rise buildings, which could attribute to the
solution of the problems related to the different uses between the upper and lower areas
of the building. These transfer structures may be in the form of transfer beams, girders
or slabs. Transfer slab is considered one of the most common transfer elements which
could exist in high rise buildings. In order to draw a complete picture of this three-
dimensional element in this thesis, pile caps are considered as a simple transfer slab.

The geometry of pile cap doesn't provide engineers with a clear understanding of
its physical behavior which is considered as a complete disturbed region and there is a
very important volume of concrete that is subjected to low normal stresses and
significant shear stresses is called "inactive concrete", Strut-and-Tie model, on the
other hand, can provide this understanding and hence offer the possibility of improving
the design. This fact should be reflected in the strength assessment of the nodes and
struts.

In this study, a three-dimensional nonlinear finite element analysis has been


conducted on pile caps that had been experimentally tested and the output results of the
cracking pattern, deflection, failure mode and stress distribution from stress trajectories
(that can't be obtained using the strut-and-tie model) are obtained and compared with
the experimental results in order to verify the validity of the FEM procedure. The
obtained results demonstrates the accuracy and power of the finite element method in
dealing with problem of pile caps.

The method of strut-and-tie model has been applied to the selected pile caps
that had been examined with aid of the finite element. The spatial nature of pile caps has
been reflected on the strength of struts and nodes of strut-and-tie model models. The
obtained results demonstrate the reliability of the strut-and-tie model method in
obtaining a lower bound estimate of the collapse load of pile caps.

III
CONTENTS
ACKNOWLEDGEMENTS________________________________________ I
ABSTRACT_____________________________________________________ III

CHAPTER 1: INTRODUCTION
1.1 Background _______________________________________________ 2
1.1.1 Transfer Slabs__________________________________________ 2
1.1.2 Design Practice_________________________________________ 3
1.2 Sectional Design Approach and Flow of Forces Approach___________ 4
1.3 Objectives and Scope________________________________________ 5
1.4 Thesis Structure____________________________________________ 6

CHAPTER 2: THE STRUT-AND-TIE MODEL METHOD


2.1 Introduction_______________________________________________ 8
2.2 Strut and Tie Design in Codes_________________________________ 10
2.3 Design Procedure via Strut-and-Tie Modeling____________________ 10
2.3.1 Development of a Strut-and-Tie Model______________________ 10
2.3.2 Components of Strut-and-Tie Model________________________ 13
2.4 Design of STM Elements ____________________________________ 13
2.4.1 Struts_________________________________________________ 13
2.4.1.1 Strut Inclination Angle _____________________________ 13
2.4.1.2 Types of Strut ____________________________________ 14
2.4.1.3 Strength of Strut__________________________________ 15
2.4.2 Tension Ties (or Ties) ___________________________________ 17
2.4.2.1 Development Length of Reinforcement________________ 18
2.4.2.2 Width of Tension Ties_____________________________ 18
2.4.3 Model Nodes__________________________________________ 20
2.4.3.1 Types of Nodes___________________________________ 20
2.4.3.2 Geometry of Nodal Zones__________________________ 21
2.4.3.3 Strength of Nodal Zones___________________________
24

CHAPTER 3: TRANSFER ELEMENTS


3.1 Introduction________________________________________________ 26
3.2 Definition_________________________________________________ 26

IV
Contents

3.3 Types_____________________________________________________ 27
3.3.1 Transfer Beams________________________________________ 27
3.3.2 Transfer Girders________________________________________ 28
3.3.3 Transfer Slabs__________________________________________ 29

CHAPTER 4: 3-D STRUT-AND-TIE MODEL APPLIED TO PILE CAPS


4.1 Introduction_______________________________________________ 34
4.2 Challenges of 2-D Strut-and-Tie Design Procedure ________________ 35
4.3 Three-Dimensional Strut-and-Tie Model_________________________ 37
4.3.1 Three-Dimensional Nodal Zones___________________________ 37
4.3.1.1 Geometry and Examples of Three-Dimensional Nodal
Zones__________________________________________ 38
4.3.1.2 Nodes With Overlapping Struts in the Same Quadrant___ 39
4.3.1.3 Three-Dimensional Nodal Zone Strength ______________ 40
4.3.2 Three-Dimensional Struts________________________________ 43
4.3.2.1 Strut Limitation Angle_____________________________ 44
4.3.3 Inactive Concrete in Three-Dimensional Structures ____________ 44
4.3.3.1 Effect of Inactive Concrete on Bearing Capacity of the
Structure _______________________________________ 45
4.3.3.2 Bearing Strength of Compressive Struts Confined by
Inactive Concrete_________________________________ 45
4.4 Comparison between 2D-STM and 3D-STM______________________
50

CHAPTER 5: 3-D NONLINEAR FINITE ELEMENT MODELING


5.1 Introduction_______________________________________________ 52
5.2 Finite Element Model by ANSYS _____________________________ 52
5.2.1 Element Types_________________________________________ 52
5.2.1.1 Concrete________________________________________ 52
5.2.1.2 Reinforcement____________________________________ 53
5.2.2 Material Modeling in ANSYS_____________________________ 53
5.2.2.1 Concrete________________________________________ 53
5.2.2.2 Steel Reinforcement _______________________________ 56
5.2.2.3 Bond between Concrete and Reinforcement____________ 56

V
Contents

5.2.3 Non-linear solution_____________________________________ 56


5.2.3.1 Time Stepping and Loading________________________ 56
5.2.3.2 Newton-Raphson Iterative Solution___________________ 57
5.3 Case Study (Experimental Test of Pile Caps of Adebar, Kuchma and
Collins) ___________________________________________________ 59
5.3.1 Geometry and Details of Reinforcement_____________________ 59
5.3.2 Results of Experimental Test______________________________ 62
5.3.2.1 Pile Cap A_______________________________________ 62
5.3.2.2 Pile Cap B_______________________________________ 63
5.3.2.3 Pile Cap C_______________________________________ 63
5.3.2.4 Pile Cap D and E_________________________________ 63
5.3.2.5 Pile Cap F_______________________________________ 64
5.3.2.6 Summary of Experimental Results ___________________ 64
5.4 Analysis of the Six Deep Pile Caps of Adebar, Kuchma, and Collins_____ 65
5.4.1 Pile Caps A, B, D, and E_________________________________ 65
5.4.1.1 Model Description and Material Properties_____________ 65
5.4.1.2 Meshing_________________________________________ 68
5.4.1.3 Loads and Boundary Conditions _____________________ 70
5.4.1.4 Finite Element Results_____________________________ 70
5.4.1.5 Comparison of the Results__________________________ 73
5.4.2 Pile Cap C____________________________________________ 74
5.4.2.1 Model Description and Material Properties_____________ 74
5.4.2.2 Meshing_________________________________________ 75
5.4.2.3 Loads and Boundary Conditions _____________________ 76
5.4.2.4 Finite Element Results_____________________________ 76
5.4.3 Pile Cap F_____________________________________________ 78
5.4.3.1 Model Description and Material Properties_____________ 78
5.4.3.2 Meshing_________________________________________ 79
5.4.3.3 Loads and Boundary Conditions______________________ 79
5.4.3.4 Finite Element Results_____________________________ 80
5.5 The Effect of Pile Cap Thickness on Pile Load Distribution_________ 81
5.6 Conclusions_______________________________________________
82

VI
Contents

CHAPTER 6: 3-D STRUT-AND-TIE MODELS


6.1 Introduction_______________________________________________ 84
6.2 Case Study 1 - Pile Cap A____________________________________ 84
6.3 Case Study 2 - Pile Cap B ____________________________________ 95
6.4 Case Study 3 - Pile Cap C ____________________________________ 107
6.5 Case Study 4 - Pile Cap D ____________________________________ 119
6.6 Case Study 5 - Pile Cap E ____________________________________ 129
6.7 Case Study 6 - Pile Cap F ____________________________________ 139
6.8 STM vs. Experimental Data _________________________________ 149
6.9 Conclusions_______________________________________________ 149

CHAPTER 7: SUMMARY AND CONCLUSIONS


7.1 Summary_________________________________________________ 151
7.2 Conclusions_______________________________________________
151

REFERENCES_________________________________________________ 153

VII
LIST OF TABLES
Page
Table 2.1 ACI 318M-14 Code values of coefficient for struts__________ 16
Table 2.2 ACI 318-14 Code values of coefficient for nodes____________ 24
Table 5.1 Material models for SOLID 65____________________________ 54
Table 5.2 Summary of reinforcement_______________________________ 61
Table 5.3 Results of concrete cylinder test____________________________ 61
Table 5.4 Proprieties of reinforcing steel 62
Table 5.5 Summary of pile cap test results____________________________ 64
Table 5.6 The main reinforcement and their effective depths using in
ANSYS_______________________________________________ 66
Table 5.7 Comparison of ultimate loads______________________________ 73
Table 6.1 Summary of calculation results of short direction of pile cap A___ 92
Table 6.2 Summary of calculation results of long direction of pile cap A___ 95
Table 6.3 Summary of calculation results of short direction of pile cap B___ 102
Table 6.4 Summary of calculation results of long direction of pile cap B___ 107
Table 6.5 Summary of calculation results of short direction of pile cap C___ 115
Table 6.6 Summary of calculation results of long direction of pile cap C___ 119
Table 6.7 Summary of calculation results of short direction of pile cap D___ 125
Table 6.8 Summary of calculation results of long direction of pile cap D___ 129
Table 6.9 Summary of calculation results of short direction of pile cap E___ 135
Table 6.10 Summary of calculation results of long direction of pile cap E___ 139
Table 6.11 Summary of calculation results of short direction of pile cap F___ 145
Table 6.12 Summary of calculation results of long direction of pile cap F___ 148
Table 6.13 The STM results compared with test results___________________ 149

VIII
LIST OF FIGURES
Page
Fig. 1.1 Combined structural system with transfer slab__________________ 2
Fig. 1.2 Pile cap as a transfer slab___________________________________ 3
Fig. 1.3 Examples of D-regions (ACI-318-14[5]) ______________________ 5
Fig. 2.1 Truss model used by Ritter, 1899[37] _________________________ 8
Fig. 2.2 B-regions and D-regions in several structural members___________ 9
Fig. 2.3 Stress trajectories in a B-region and near discontinuities (D-regions)_ 10
Fig. 2.4 Development of an STM ___________________________________ 11
Fig. 2.5 Flowchart illustrating STM steps ____________________________ 12
Fig. 2.6 Typical direct strut and tie model for a deep beam_______________ 13
Fig. 2.7 Angle recommendations in a deep beam with stirrups, for deviation
of concentrated forces and between struts and ties________________ 14
Fig. 2.8 The different strut shapes with examples in a beam_______________ 15
Fig. 2.9 Types of struts____________________________________________ 17
Fig. 2.10 Calculation of the development length at the nodal zone__________ 18
Fig. 2.11 The width of the tie used to determine the dimensions of the node 19
Fig. 2.12 Different types of nodes ____________________________________ 21
Fig. 2.13 Example of hydrostatic nodal zone____________________________ 22
Fig. 2.14 Example of non-hydrostatic nodal zone _______________________ 23
Fig. 3.1 Schematic diagram of transfer structure________________________ 26
Fig. 3.2 Transfer beam in beam-shear wall system______________________ 17
Fig. 3.3 Distribution of horizontal stress in beam-shear wall system________ 28
Fig. 3.4 Hotel cross section________________________________________ 29
Fig. 3.5 Stress trajectories in a transfer girder__________________________ 29
Fig. 3.6 Commencement of the superstructure construction on top of the
transfer slab_____________________________________________ 30
Fig. 3.7 Erection of a temporary platform using universal steel sections with
bracket support to the columns as work station for the transfer plate_ 31
Fig. 3.8 The final completed external appearance of the Olympian City
project__________________________________________________ 32
Fig. 3.9 Pile cap geometry_________________________________________ 32

IX
List of Figures

Fig. 4.1 Example of a 3-D strut-and-tie model and corresponding


reinforcement arrangement for a pile plinth_____________________ 34
Fig. 4.2 A combination of 2D strut-and-tie models______________________ 35
Fig. 4.3 Examples of nodal zones in transfer slabs______________________ 38
Fig. 4.4 Nodes with overlapping struts _______________________________ 39
Fig. 4.5 Triaxial compression tests Results, An application to find A2 in
stepped or sloped supports__________________________________ 41
Fig. 4.6 An application to find A_2 in stepped or sloped supports_________ 42
Fig. 4.7 Inclined 3-D strut in 2C2T-node over a pile located at the corner of a
pile cap_________________________________________________ 43
Fig. 4.8 Inactive concrete in pile cap________________________________ 44
Fig. 4.9 Development of strut-and-tie model for deep member or pile cap___ 46
Fig. 4.10 Maximum bearing stress to cause transverse splitting in biaxial stress
field____________________________________________________ 46
Fig. 4.11 Splitting in a concrete cylinder, tension develops in radial directions_ 47
Fig. 4.12 Analytical study of transverse tension in triaxial stress field_______ 48
Fig. 4.13 Analytical study of the ratio between stress at cracking, to concrete
strength ______________________________________________ 48
Fig. 4.14 Applied strut-and-tie model in two different planes______________ 50
Fig. 5.1 SOLID65 element geometry ________________________________ 52
Fig. 5.2 Discrete models for reinforcement ___________________________ 53
Fig. 5.3 LINK8-3D geometry _____________________________________ 53
Fig. 5.4 Stress-strain curve for concrete______________________________ 55
Fig. 5.5 Element connectivity; concrete solid and link elements ___________ 56
Fig. 5.6 Load steps, sub-steps, and time______________________________ 57
Fig. 5.7 Newton-Raphson iterative solution (2 load increments) __________ 58
Fig. 5.8 Arc-length method vs. Newton-Raphson method________________ 59
Fig. 5.9 Geometry and details of reinforcement of the six test specimens 60
Fig. 5.10 Final deformation pattern of pile cap A________________________ 62
Fig. 5.11 Appearance of pile cap B after testing_________________________ 63
Fig. 5.12 Sequence of relevant crack formation in pile cap F______________ 64
Fig. 5.13 Investigated quadrants of tested pile caps A, B, D and E__________ 65
Fig. 5.14 Pile cap model (volumes created in ANSYS) __________________ 66

X
List of Figures

Fig. 5.15 Reinforcement configurations of pile caps A, B, D and E_________ 67


Fig. 5.16 The overall meshing for pile caps A, B, D and E________________ 69
Fig. 5.17 Boundary conditions for a typical finite element model __________ 70
Fig. 5.18 Output of “ANSYS” Program for pile caps A, B, D and E________ 71
Fig. 5.19 Investigated quadrant of tested pile cap C_____________________ 74
Fig. 5.20 Pile Cap model (Volumes Created in ANSYS) _________________ 74
Fig. 5.21 Reinforcement configurations of pile cap C____________________ 75
Fig. 5.22 The overall meshing for pile cap C___________________________ 75
Fig. 5.23 Boundary conditions for a typical finite element model __________ 76
Fig. 5.24 Output of “ANSYS” Program for pile cap C___________________ 77
Fig. 5.25 Investigated quadrant of tested pile cap F______________________ 78
Fig. 5.26 Pile cap model (volumes created in ANSYS) __________________ 78
Fig. 5.27 Reinforcement configurations of pile cap F____________________ 79
Fig. 5.28 The overall meshing for pile cap F___________________________ 79
Fig. 5.29 Boundary conditions for a typical finite element model __________ 80
Fig. 5.30 Output of “ANSYS” Program for pile cap F___________________ 80
Fig. 5.31 Effect of pile cap rigidity on piles load________________________ 82
Fig. 6.1 Geometry of pile cap A____________________________________ 85
Fig. 6.2 Proposed 3D-STM of pile cap A_____________________________ 85
Fig. 6.3 STM for pile cap A in the short direction______________________ 86
Fig. 6.4 Nodes of pile cap A in the short direction_____________________ 87
Fig. 6.5 Amount of confinement for pile caps A, B, D and E_____________ 88
Fig. 6.6 STM for pile cap A in the long direction______________________ 92
Fig. 6.7 Nodes of pile cap A in the long direction_____________________ 93
Fig. 6.8 Geometry of pile cap B____________________________________ 96
Fig. 6.9 Proposed 3D-STM of pile caps B___________________________ 96
Fig. 6.10 STM for pile cap B in the short direction______________________ 97
Fig. 6.11 Nodes of pile cap B in the short direction_____________________ 98
Fig. 6.12 STM of pile cap B in the long direction______________________ 102
Fig. 6.13 Nodes of pile cap B in the long direction_____________________ 103
Fig. 6.14 Geometry of pile cap C____________________________________ 107
Fig. 6.15 Proposed 3D-STM of pile cap C____________________________ 108
Fig. 6.16 STM of pile cap C in top view______________________________ 108

XI
List of Figures

Fig. 6.17 STM of pile caps C in the short direction_____________________ 109


Fig. 6.18 Nodes of pile cap C in the short direction_____________________ 110
Fig. 6.19 Amount of confinement for pile cap C________________________ 111
Fig. 6.20 STM of pile caps C in the long direction______________________ 115
Fig. 6.21 Nodes of pile cap C in the long direction_____________________ 116
Fig. 6.22 Geometry of pile cap D____________________________________ 119
Fig. 6.23 Proposed 3D-STM of pile caps D and E________________________ 120
Fig. 6.24 STM of pile cap D in the short direction______________________ 121
Fig. 6.25 Nodes of pile cap D in the short direction_____________________ 121
Fig. 6.26 STM of pile cap D in the long direction______________________ 126
Fig. 6.27 Nodes of pile cap D in the long direction_____________________ 126
Fig. 6.28 Geometry of pile cap E____________________________________ 129
Fig. 6.29 STM of pile cap E in the short direction______________________ 130
Fig. 6.30 Nodes of pile cap E in the short direction_____________________ 131
Fig. 6.31 STM of pile cap E in the long direction______________________ 136
Fig. 6.32 Nodes of pile cap E in the long direction_______________________ 137
Fig. 6.33 Geometry of pile cap F____________________________________ 140
Fig. 6.34 Proposed 3D-STM of pile cap F_____________________________ 140
Fig. 6.35 Amount of confinement at column and pile nodal zones__________ 142

XII
XII
CHAPTER 1

INTRODUCTION
1.1 Background

1.1.1 Transfer Slabs


Transfer slabs have become one of the most practical methods in construction of High-
rise buildings, which could overcome the functionality problems of the buildings. A
podium structure in the lower zone in the building needs large spacing between the
supporting columns in order to use this produced open area for parking or shopping
malls - as an example; in contrast, in the upper zone no need for this large spacing
between columns, as it is often used as residential units. Fig. 1.1 illustrates the
combination system between the upper and lower zones using transfer slabs.

Fig. 1.1 Combined structural system with transfer slab.

2
Chapter 1 Introduction

Pile caps may be used to describe a reinforced concrete slab constructed on the
top of a group of foundation piles to spread the load they are to carry, as shown in Fig.
1.2. In this thesis, pile caps are considered as a simple example of transfer slabs.

Fig. 1.2 Pile cap as a transfer slab.

Because of their geometrical feature, pile caps do not obey the simplified
sectional approaches normally used to design most of reinforced structural elements.
Pile caps cannot rigorously be considered as beams (design using beam or deep beam
theory), neither as slabs (slab and flat slab design approach) and neither as walls (shear
wall or plate theory). A pile cap is an interface element between the superstructure and
the substructure, which is considered a three-dimensional element. The designer can
base the design of this structure on a study where the complete transfer of forces in the
pile cap is considered at once. For instance, this could be done based on the theory of
elasticity using fine mesh finite element analysis or another methodology, which
requires less calculation efforts; e.g., the strut-and-tie method.

1.1.2 Design Practice


A transfer slab (e.g., a pile cap) is a structural member whose function is to transfer load
from a column to a group of piles. Current design procedures for this structure don’t
supply engineers with a clear understanding of the physical behavior of these elements.
Strut-and-Tie model, on the other hand, can provide such understanding and hence
offer the possibility of improving current design practice.

3
Chapter 1 Introduction

Linear and nonlinear analyses illustrate that pile caps behave as three-
dimensional elements in which there is a complex variation in straining over the
dimensions of the discontinuity or disturbed region, in which compressive struts
develop between columns and piles. Of particular concern was that many pile caps that
were designed to fail in flexure have been reported to fail in the brittle mode of shear
(Chantelot and Mathern, 2010 [9]). For this reason, design procedure for pile caps
should not be based on a sectional design. On the other hand, the strut-and-tie modeling
can be applied to any structural component with any loading and support conditions.

The strut-and-tie model is a design procedure already implemented and


strongly recommended for the design of pile caps (i.e. transfer slabs) in, among others,
the European and the American building codes.

1.2 Sectional Design Approach and Flow of Forces Approach


In the design of reinforced concrete structures, a distinction can be made between B-
regions (standing for Bernoulli’s regions) and D-regions (standing for discontinuity-
regions) (Schlaich et al., 1987 [35]). In B-regions, the linear strain distribution of
flexure theory applies and thus a sectional analysis is appropriate to design these
regions. The sectional design methodologies are predicated on the traditional beam
theory, and hence are not appropriate for application to D-regions.

In D-regions, geometrical discontinuities or static discontinuities result in


disturbances and the plane sections assumption is not valid anymore. According to St.
Venant’s principle, the D-regions are assumed to extend to a characteristic distance h
away from the discontinuity, depending on the geometry as shown in Fig. 1.3. Thus, the
design of D-regions must proceed on a regional, rather than a sectional, basis. Like-
truss model or STM provides the means by which this goal can be accomplished
(Williams et al., 2012 [41]).

4
Chapter 1 Introduction

Fig. 1.3 Examples of D-regions, ACI-318-14 [5].

Unlike the sectional methods of design, the strut-and–tie method does not lend
itself to a cook book approach and therefore requires the application of engineering
judgment. The Strut-and-Tie Method (STM) can be practically used to design Transfer
Slabs; e.g., pile-caps, where the entire structure is classified as D-region. Although
there is no generally accepted procedure for the design of pile caps; many empirical
detailing rules are followed in practice, but these approaches vary significantly. The
main reason for these disparities is that most codes do not provide a design
methodology that provides a clear understanding of the strength and behavior of this
important structural element.

1.3 Objectives and Scope


The purpose of this study focuses on the design and the detailing of reinforced concrete
transfer slabs with pile caps considered as simple form of these elements by utilizing
both:

 a 3-D nonlinear finite element analysis using ANSYS 12.0- package computer
program and

 the Strut-and-Tie Models STM method.


The verification process of both the strut-and-tie results and the finite element model is
also achieved using experimental data available in literature.

5
Chapter 1 Introduction

1.4 Thesis Structure


This thesis contains the following chapters:

 Chapter 2 presents an introduction to 2D-STM including the development of the


strut-and-tie model, STM elements with an example of strategic placement of
nodes, struts and ties and their strengths.

 Chapter 3 gives a brief review on transfer elements which could be encountered


in practice.

 Chapter 4 illustrates the art of the three-dimensional strut-and-tie method


applied to pile caps (i.e. transfer slabs) including the geometry of three
dimensional nodal zones and their strengths.

 Chapter 5 presents a three-dimensional nonlinear finite element analysis of the


experimental specimens using ANSYS V12.0. The element types and material
models of both reinforcing steel and concrete are presented along with the
numerical model verification. A parametric study using ANSYS V12.0 finite
element package is presented.

 Chapter 6 shows comparisons of the output results of the finite element model
with the results from the proposed three dimensions strut-and-tie models.

 Chapter 7 summarizes the work carried out and the main conclusions drawn in
this thesis.

6
CHAPTER 2
THE STRUT-AND-TIE MODEL METHOD

2.1 Introduction
The analysis of shear in reinforced concrete members is based on the so called truss
analogy. Since the beginning of the 20th century, designers started to use regular truss
models in order to design structural concrete members by following the flow of forces.
These models have been used to handle regions with high shear force or torsional
moment, where the simple theories of flexure do not apply. An illustration of that is the
use of truss models for shear design by Ritter, 1899 [32], see Fig. 2.1.

Fig. 2.1 Truss model used by Ritter, 1899[32].

When a member is loaded, the crack pattern appears. This crack pattern helps to
describe how load is transferred in the member and then the equivalent truss model can
be identified. Ritter, 1899 [32] proposed the truss analogy in reinforced concrete
members for the calculation of shear reinforcement, which is the oldest and well known
example of reasoning by means of “strut-and-tie models”. However, the method is
also very well suited for the design calculation of compact structural members such as
foundation slabs (or blocks) supported by piles, corbels, deep beams, walls and
anchorage regions in prestressed members. The strut-and-tie method is essentially
founded on the publications of SCHLAICH and his colleagues and students.

Based on the distribution of stresses and strains within any cross-section of


concrete structure members, two types of regions could appear. One of these regions is
the Bernoulli region (B-region) and the other is the discontinuity regions (D-region). In
B-regions, the assumptions of Bernoulli‟s hypothesis of flexure is applicable (i.e., plane
sections before bending remain plane after bending) and have been successfully treated
using the the method of truss model while, in D-regions, Bernoulli‟s hypothesis does
not apply. Strut-and-Tie model is a powerful tool a rational method for understanding

8
Chapter 2 The Strut-and-Tie Model Method

these D-regions behavior as they are considered the most critical regions in concrete
structural elements (Chen and El-Metwally, 2011 [10]).

St. venant's principle suggests that the localized effect of the disturbance dies
out in about one member depth from the point of the disturbance. On this basis, D-
regions are assumed to extend one member depth each way from the discontinuity. Fig.
2.2 shows B- and D-regions in several structural members (MacGregor, 1997 [26]).

Fig. 2.2 B-regions and D-regions in several structural members [26].

Fig. 2.3 shows that in B-regions, the stress trajectories are smooth while in D-
regions, trajectories are not smooth as the intensities of the stresses decrease rapidly
near the origin of the discontinuity and such behavior the basic concept of B- and D-
regions of a structure.

9
Chapter 2 The Strut-and-Tie Model Method

Fig. 2.3 Stress trajectories in a B- and D-regions.

2.2 Strut-and-Tie Design in Codes


Nowadays, most of the major codes of practice allow the use of strut-and-tie models.
Since 1984, STM has been included in the Canadian Concrete Code as an alternative for
shear design in regions including statical or geometrical discontinuities. The design
according to stress fields using the strut-and-tie method became an alternative for the
structural analysis of discontinuity regions in the CEB-FIP Model Code 1990. Strut-
and-tie models were then introduced into the ACI Building Code in 2002. Then the
Appendix A “Strut-and-tie models” was created and different parts of the codes were
modified to allow the design with strut-and-tie models. In 2004, the strut-and-tie
method was introduced to Eurocode 2 (Chantelot and Mathern, 2010 [9]).

2.3 Design Procedure via Strut-and-Tie Modeling


2.3.1 Development of a Strut-and-Tie Model

In order to develop a suitable strut-and-tie model, STM, some techniques and rules
should be used such as the „load path method’ according to (Schlaich et al., 1987
[35]), „stress field approach‟ proposed by (Muttoni et al, 1997 [29]) or by linear
finite element analysis. These methods can help the designer in choosing a suitable
stress field. Fig. 2.4 illustrates the development of a strut-and-tie model by means of the
load path method.

10
Chapter 2 The Strut-and-Tie Model Method

(a) The region and (b) Load paths through (c) Corresponding
boundary load the region STM
Fig. 2.4 Development of a STM (Schlaich and Schäfer, 1991 [34]).

Fig. 2.5 illustrates one of several different design procedures that can be
followed for the design of a structural member using a STM method. Schlaich and
Schäfer, 1991[34] suggested that the best STMs are the models, when the tension ties
are short. In addition, Schlaich and Schäfer supposed that hyperid STMs could be
obtained by assembling two or more simple STMs within a D-region. while, Brown et
al. [8] illustrated that “statically determinates strut and tie model is the most preferable
one” since it is easy to calculate the member forces. Contrariwise, in statically
indeterminate systems, the true geometry of the struts are very difficult to be accurately
estimated.

11
Chapter 2 The Strut-and-Tie Model Method

Fig. 2.5 Flowchart illustrating STM steps (Brown et al., 2006 [8])

12
Chapter 2 The Strut-and-Tie Model Method

2.3.2 Components of Strut-and-Tie Model


The strut-and-tie model consists of:
Major compression diagonals (struts)
Tension ties (or ties)
Model nodes
Fig. 2.6 illustrates typical components of Strut-and-Tie Model using deep beam
example.

Fig. 2.6 Typical direct STM for a deep beam.

2.4 Design of STM Elements


2.4.1 Struts
Concrete struts carries the compressive forces in strut-and-tie models and they represent
the compressed concrete stress fields, often represented by dashed lines in the model.

2.4.1.1 Strut Inclination Angle

There are two kinds of problems related to the option of the struts inclination angle
when creating a strut-and-tie model. The first is the plastic redistribution and strain
compatibility problems between unstressed and stressed regions with the reason of an
inadequate deviation angle at concentrated forces. The second is the strain compatibility
problem because of the small angles between ties and struts. In a STM, the
recommendations on minimum angles to use differ between different authors and codes.
In FIB bulletin 3 cited in Engström, 2009 [17], the recommendations of Schäfer, 1999
[33] are given hereafter, using the notations of Fig. 2.7.

13
Chapter 2 The Strut-and-Tie Model Method

a) Deviation of concentrated loads


α ≈30° and α< 45° (2.1)
Additionally, the stresses under concentrated loads should be directly spread out when
entering the D-region.
b) Angles between struts and ties
θ1 ≈60° and θ1> 45° (2.2)
θ2 ≈45° and θ2> 30° (2.3)
θ3 ≈45° and θ3> 30° (2.4)

Fig. 2.7 Strut Inclination Angle.

But, the choice of 25° is considered between a strut and a tie joining at a node in
the ACI Building Code-318-14 [5].

2.4.1.2 Types of Strut


As discussed by Schlaich and Schäfer 1991 [34], Struts are generally divided in three
types, prismatic-, bottle- and fan-shaped struts, see Fig. 2.8. The prismatic-shaped strut
has a constant width. The bottle-shaped strut contracts or expands along the length and
in the fan-shaped strut a group of struts with different inclinations meet or disperse from
a node. The shape of a strut is highly dependent upon the force path from which the
strut arises and the reinforcement details of any reinforcement connected to the tie. The
capacity of the strut must be reduced, if the strut is subjected to unfavorable multi-axial
effects. On the other hand, if the strut is confined in concrete (i.e. multi-axial
compression exists), the capacity of the strut becomes greater.

14
Chapter 2 The Strut-and-Tie Model Method

Fig. 2.8 the different strut shapes with examples in a beam [9].

2.4.1.3 Strength of Strut


As the strut has no transverse reinforcement, the longitudinal cracks in the strut occurs
due to the spreading of compression forces, which gives rise to transverse tension, and
hence, the struts may fail after this cracking occurs. While, the strut will fail by
crushing, if sufficient transverse reinforcement is supplied.

According to ACI-318-14 [5], the nominal compressive strength of a strut


without longitudinal reinforcement shall be taken the smaller value of:

(2.5)
at the two ends of the strut, where is the cross-sectional area at one end of the strut,
and is the smaller of:

The effective compressive strength of the concrete in the strut.


The effective compressive strength of the concrete in the nodal zone.

where the effective compressive strength of the concrete in a strut shall be taken as

(2.6)

where is the effectiveness factor for concrete struts, takes into account the stress
conditions, strut geometry and the angle of cracking surrounding the strut. Table 2.1
shows the value of in this investigation.

15
Chapter 2 The Strut-and-Tie Model Method

Table 2.1 ACI 318M-14 Code values of coefficient for struts.

Strut condition

A strut with constant cross-section along its length (for example a strut 1.0
equivalent to the rectangular stress block in a compression zone in a
beam).

For struts located such that the width of the midsection of the strut is
larger than the width at the nodes (bottle-shaped struts):
a) With reinforcement normal to the center-line of the strut to resist
0.75
the transversal tensile force.
0.60λ
b) Without reinforcement normal to the center-line of the strut

For struts in tension members, or the tension flanges of members, for 0.40
example, two compression struts in a strut-and-tie model used to design
the longitudinal and transverse reinforcement of the tension flanges of
beams, box girders, and walls.

For all other cases applies to strut applications not included in above 0.60λ
cases (struts in a beam web compression field in the web of a beam where
parallel diagonal cracks are likely to divide the web into inclined struts,
and struts are likely to be crossed by cracks at an angle to the struts, Figs.
2.9a and 2.9b respectively.

where λ is a modification factor to account for the use of lightweight concrete. λ = 0.85
for sand-lightweight concrete and 0.75 for all-lightweight concrete and λ = 1.0 for
normal weight concrete.

16
Chapter 2 The Strut-and-Tie Model Method

(a) Struts in a beam web with inclined cracks parallel to struts.

(b) Struts crossed by skew cracks.


Fig. 2.9 Types of struts [5].

The design of struts shall be based on


(2.7)
In another form
( ) (2.8)

where is the largest factored force acting in a strut and obtained from the applicable
load combinations and the factor is 0.75 for ties, struts, and nodes.

2.4.2 Tension Ties (or Ties)


The tension tie is the second major component of a strut-and-tie model. The cross-
section of the tie along its length does not change. This element represents one or
several layers of steel in the same direction to be designed according to;
(2.9)

where is the cross section of area of steel, is the yield strength of steel and the
nominal strength of a tie shall be taken as

(2.10)

To make the design safe, special attention has still to be paid to the
anchorage in order to avoid failure of tension ties due to the lake of end anchorage. One
of the main advantages of the STM is that it static factorily indicates the need for
anchorage. Tension ties are represented as solid lines in strut and tie models.

17
Chapter 2 The Strut-and-Tie Model Method

2.4.2.1 Development Length of Reinforcement


Fig. 2.10 illustrates that the tension reinforcement in a tie should be developed with a
length equal to at the end of the tie. Hooks, mechanical devices and additional
confinement are working to reduce this development length of the tie reinforcement.

Fig. 2.10 Calculation of the development length at the nodal zone [5].

2.4.2.2 Width of Tension Ties

The effective tie width assumed in design can vary between the following limits,
depending on the distribution of the tie reinforcement:
In case of using one row of bars without sufficient development length beyond
the nodal zones (Fig. 2.11a):
(2.11a)
In case of using one row of bars and providing sufficient development length
beyond the nodal zones for a distance not less than , where is the concrete
cover (Fig. 2.11b):

(2.11b)

where is the bar diameter.


In case of using more than one row of bars and providing sufficient
development length beyond the nodal zones for a distance not less than ,

where is the concrete cover (Fig. 2.11c):


( ) (2.11c)
where is the number of bars and is the clear distance between bars.

18
Chapter 2 The Strut-and-Tie Model Method

In the three cases in Fig. 2.11, the development length according to the
Egyptian code [15] and ACI code [5] is equal to and , respectively. Where the
development length, begin at intersect the center of tie with extended nodal zone.

( ) (b)

(c) ( )

Fig. 2.11 The width of the tie used to determine the dimensions of the node [5].

The upper limit is established as the width corresponding to the width in a hydrostatic
nodal zone, calculated as

(2.12)

19
Chapter 2 The Strut-and-Tie Model Method

where is the applicable effective compressive strength of a nodal zone and is


computed from[5,15] as,

(2.13)

The stands for a cylinder concrete compressive strength and for a cube
concrete compressive strength, is the effectiveness factor for nodal zones, and is
the breadth of the beam. The width of the tie is to be determined to satisfy the
compressive stresses at nodes.

2.4.3 Model Nodes

2.4.3.1 Types of Nodes


Nodes represent the connections between struts and ties or the positions where the
stresses are redirected within the STM. There are two kinds of nodes which may exist in
a STM; “singular” (or “concentrated”) nodes and “smeared” (or “continuous”) nodes.
Smeared nodes do not need to be checked in a STM because the reinforcement is
properly developed until the extremities of the stress field. While, the singular nodes,
corresponds to the nodes close to the concentrated loads, reactions, .. etc. In the case of
deep beams or pile caps, singular nodes would be the nodes under the columns and over
the supports or piles (i.e., transfer structure); the other nodes being smeared over large
regions.
Singular nodes in STMs are the critical nodes because of the concentration of
stresses in the concrete at these nodal regions. Therefore, the check of these singular
nodes is very significant in STMs.

20
Chapter 2 The Strut-and-Tie Model Method

The concentrated nodes are divided into four major node types, CCC-, CCT-,
CTT- and TTT- nodes illustrated in Fig. 2.12;

C-C-C nodal zone bounded by compression struts only (hydrostatic node),


C-C-T nodal zone bounded by compression struts and one tension tie,
C-T-T nodal zone bounded by a compression strut and two tension ties, and
T-T-T nodal zone bounded by three tension ties.

Where C stands for compression force and T stands for tension force.

Fig. 2.12 Different types of nodes.

2.4.3.2 Geometry of Nodal Zones


The volume of concrete around a node “in which forces acting in different directions,
meet and balance” (Schäfer, 1999 [33]) is assumed to transfer strut-and-tie forces
through the node is called "Nodal Zone". As discussed by Brown et al. in 2006 [8], a
node can be detailed to be either hydrostatic or non-hydrostatic in theory. Hydrostatic
nodal zones were used at the beginning of the development of strut-and-tie models. The
faces of these nodal zones were perpendicular and proportional to the forces acting on
the node; see Fig. 2.13a. Therefore, no shear stresses were created at the node (Fig.
2.13e). However it is almost impossible to manage to have geometries assuring
hydrostatic nodes in a model. For this reason, all the major codes recognize non-
hydrostatic nodes nowadays (Fig. 2.14). For non-hydrostatic nodes, Schlaich et al.,
1987 [35] suggested that the ratio of maximum stress on a face of a node to the
minimum stress on another face of the node should be less than 2.0; otherwise, the non-
uniformity of stress distribution could make the check of the node unsafe.

21
Chapter 2 The Strut-and-Tie Model Method

Fig. 2.13 Example of hydrostatic nodal zone.

22
Chapter 2 The Strut-and-Tie Model Method

Fig. 2.14 Example of non-hydrostatic nodal zone.

The width of the inclined strut in Figs. 2.13 and 2.14 is defined by:

(2.14)

23
Chapter 2 The Strut-and-Tie Model Method

2.4.3.3 Strength of Nodal Zones


The compressive strength of concrete of the nodal zone depends on many factors
including the tensile straining from intersecting ties, confinement provided by
compressive reactions and confinement provided by transverse reinforcement.
The nominal compressive strength of a nodal zone, , shall be taken as;
(2.15)
where is the effective compressive strength of the concrete in the nodal zone and
is the smaller of:
The area of the face of the nodal zone on which acts, taken perpendicular
to the line of action of the strut force .
The area of a section through the nodal zone, taken perpendicular to the line
of action of the resultant force on the section.
the effective compressive strength of the concrete in a nodal zone can be obtained
from:
or (2.16)
where is the effectiveness factor of a nodal zone and it is assumed as given in Table
2.2 according to the ACI 318-14 code[5].

Table 2.2 ACI 318-14 Code values of coefficient for nodes.


Nodal zone
Compression-Compression-Compression, C-C-C 1.00
Compression-Compression-Tension, C-C-T 0.80
Two ties or more; C-T-T or T-T-T 0.60

The value of βn expresses how much disturbance of the node which could exist
because of the compatibility problems of the tension tie strains in the reinforcement and
the compression strains of concrete in the struts.

24
CHAPTER 3
TRANSFER ELEMENTS
3.1 Introduction
The use of transfer elements between the upper and lower zones of structures of a
high-rise building has become popular and sometimes even inevitable. The design of
reinforced concrete transfer elements, although routinely performed by structural design
engineers, is a very challenging task. One of the major characteristics of buildings with
transfer structures is that the spacing of vertical supporting elements above a transfer
structure (typical floor) is comparatively closer than below it (podium) for easy and
flexible architectural planning purposes. Transfer structures were usually idealized as
deep beams or thick slabs.

3.2 Definition
Transfer structures can be defined as either flexural or shear elements that transmit
heavy loads from columns or walls acting on its top and redistribute them to supporting
columns or walls, see Fig. 3.1.

Fig. 3.1 Schematic diagram of transfer structure.

The traditional design of these elements is similar to a pile cap or a beam.


Normally, the column supports strength or shear walls strength above these transfer
system are much lower than the flexural stiffness and strength of the transfer system
itself. Zhang et al. 2003[44] and 2005 [46] assumed that the deformations of these
transfer structures could not be taken into considerations. The concept of this rigid
diaphragm in any building with transfer structures were accepted in routine structural
analyses.

26
Chapter 3 Transfer Elements

3.3 Types
Transfer structures may be in the form of one of three elements;
i. Transfer Beams,
ii. Transfer Girders or
iii. Transfer Slabs.

3.3.1 Transfer Beams


Transfer beams are horizontal members, normally deep and large. They are used to
transfer loading from shear wall/columns of the upper structure to the lower frame
structure through shearing action by forming a diagonal cracks, since a diagonally
cracked deep beam behaves as a tied arch (Ley et al., 2006 [25], Zhang and Tan, 2007
[45]). The conventional plane section remaining plane approach is not applicable to the
analysis of deep beams (Kong, 1990 [23]). Besides, for beams without web
reinforcement, it has been shown that shear strength decreases as member size
increases. This is associated with the phenomenon of size effect (Tan and Cheng, 2006
[37], Tan et al., 2008 [38], Yang et al., 2007 [42]).

Fig. 3.2 illustrates three cases of transfer beams in shear wall system, which are
similar to deep beams when the wall is extending to columns. The depth of the interface
beams has to be kept much higher than the conventional beams, ranging from 1.0 m to
4.5 m.

Fig. 3.2 Transfer beam in beam-shear wall system.

When the depth of the beam is large enough, in beam-shear wall system, the
compression stress may appear in the upper part of it and compression zone is relatively
small, as shown in Fig. 3.3a. In contrary, when the beam depth is relatively small, Fig.
3.3b indicates that the beam is in full tension long the span owing to the interaction
between the wall and the beam (Dar, 2007 [13]).

27
Chapter 3 Transfer Elements

a) Large Transfer Beam b) Small Transfer Beam


Fig. 3.3 Horizontal stress distribution in beam-shear wall system.
The behavior of the transfer beam in beam-shear wall systems, as illustrated in
Fig. 3.3, does not look like the behavior of the ordinary beams, but due to the inter
action between the wall and the beam, the beam is in full tension or flexural-tension
along the span. So that, in beam-shear wall system, the transfer beam which carries a
shear wall should be considered as a flexural-tension member (Dar, 2007 [13]).

3.3.2 Transfer Girders


A transfer girder is not different a lot from a transfer beams conception. To illustrate
that, for the given hotel cross section, seen in Fig. 3.4 it is necessary to provide transfer
elements for discontinuous columns at the third floor level of the hotel (as an example),
in order to provide large meeting rooms within a podium structure beneath the tower
portion containing the guest rooms. Fig. 3.5 shows the stress trajectories in a transfer
girder.

28
Chapter 3 Transfer Elements

Fig. 3.4 Hotel cross section.

Fig. 3.5 Stress trajectories in a transfer girder.

3.3.3 Transfer Slabs


A transfer slab is a structural element sometimes encountered in high-rise buildings.
Building design often involves a podium structure that houses other functional spaces
such as a shopping mall or a large lift lobby which requires an unobstructed spatial
layout in order to give a more impressive view as illustrated in Fig. 3.6.

29
Chapter 3 Transfer Elements

Fig. 3.6 The construction of residential building (super structure) above


the transfer element (transfer slab) [33]

The averaged thickness of the transfer slab is around 2.5m to 3.5m; this makes
the construction of the transfer slab quite difficult since it is a very large and heavy in
weight, but to overcome these difficulties, Fig. 3.7, a very heavy-duty false-work
system is erected as support and work platform (Raymond, 1999 [31]). Fig. 3.8
illustrates the final completed external appearance of the Olympian City in Tai Kwok
project.

30
Chapter 3 Transfer Elements

Fig. 3.7 Erection of a temporary (false-work) platform.

31
Chapter 3 Transfer Elements

Fig. 3.8 Final Completed External Appearance of the Olympian City project
in Tai Kwok

The principle of transfer slab is very similar to a raft foundation suspended on piles.
In this thesis, pile caps are taken as a simple example of transfer slabs, Fig. 3.9.

Fig. 3.9 Pile cap geometry.

32
CHAPTER 4

3-D STRUT-AND-TIE MODEL APPLIED TO PILE CAPS


4.1 Introduction
It may not be adequate to employ 2-D models in the design of transfer slabs, such as
pile caps and wind power plant foundations which are subjected to load that result in
3-D stress fields. Schlaich et al., 1987 [35] stated that “If the state of stress is not
predominantly plane, as for example in the case with punching or concentrated loads,
three-dimensional strut-and-tie models should be used.” However, in most of the time,
details have not been provided on how to deal with three dimensional strut-and-tie
models, for instance concerning the verification of nodes. There are two different
approaches for the construction of a 3-D strut-and-tie model, by modeling in 3-D or by
combining 2-D models. A 3-D strut-and-tie model for a centric loaded pile cap is shown
in Fig. 4.1.

Fig. 4.1 Example of a 3-D strut-and-tie model and corresponding reinforcement


arrangement for a pile plinth (Engström, 2011 [18]).

Fig. 4.2 indicates how 2-D strut-and-tie models can be combined to solve spatial
structure, one model in the plane of the flanges and one model in the plane of the web.
For such a model each strut-and-tie model transfers the load in its own plane. The two
models are joined with common nodes. Most of the articles about 3-D strut-and-tie
models do not detail how to consider the intersection between struts and ties and how to
check the nodal zones.

34
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Fig. 4.2 A combination of 2-D strut-and-tie models (Engström, 2011 [18]).

4.2 Challenges of 2-D Strut-and-Tie Design Procedure


The standard procedure of the 2-D strut-and-tie design methodology can be found in
design, code provisions; e.g., ACI-318-14[5]; a conceptual flowchart is depicted in Fig.
2.5 for later comparison in this chapter. Generally speaking, this procedure is a trial-
and-error iterative design process based mainly on the experience of the designer. The
conventional 2-D Strut-and-Tie Design Procedure faces at least three major challenges
in design practices.

The first challenge is that it is a very difficult to create a suitable strut-and-tie


model of a structure, especially a three-dimensional model. In the literature, most strut-
and-tie models are developed by utilizing stress trajectories from finite element analysis
(Schlaich and Schäfer, 1991 [34] and Schlaich et al., 1987 [35]). Directions of each
strut and each tie are taken in accordance with those directions of principal compressive
and tensile stresses, respectively. This strategy, however, suffers from two problems.
 First, for a region with complex stress distribution, it is not a simple task to
create a corresponding strut-and-tie model by estimation.
 Second, for a three-dimensional region, it is not easy for designers to observe
the stress trajectories in the interior.

35
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

The second challenge is the problem concerned with stiffness determination of


members in a statically indeterminate strut-and-tie model. One way to handle the
statically indeterminate case is to use the so-called plastic truss method. However,
strain compatibility requirements and limited ductility in concrete must be taken into
account (Tjhin and Kuchma, 2002 [39]). Schlaich and Schäfer, 1991 [34] proposed
that a statically indeterminate STM could be divided into several statically determinate
ones, While Ali, 1997 in [4] suggested that each statically determinate truss model was
self-sustaining for external loading. However, it is not clear whether or not a statically
indeterminate truss model can always be divided. In addition, the assumption that each
statically determinate truss model takes the same external loading may not be
reasonable. Recently, Yun, 2000 [43] proposed an iterative approach to determine the
relative stiffness of statically indeterminate strut-and-tie members; however, its
convergence property may not be assured.

The third challenge is related to the problem of indirectly evaluating concrete


bearing capacity. The effective compressive strengths of the struts and nodes are first
obtained by looking up guidelines and codes (e.g., Schlaich et al., 1987 [35]; FIP 1996
[19]; MacGregor, 1997 [26]; ACI-318-14 [5]). The effective width of the strut is then
calculated from dividing the member force by the effective strength. The shapes and
dimensions of the nodes are determined after the widths of incoming struts are resolved.
Finally, the truss model with limited widths is evaluated to determine its suitability for
the structure. If the widths of struts are not sufficient, the selected strut-and-tie model
has to be changed. For 2-D STMs, values for effective compressive strengths of struts
and nodal zones are specified in codes and guidelines. However, for 3-D STMs, no
proven guidelines are yet available. (Leu et al., 2006 [24])

36
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

4.3 Three-Dimensional Strut-and-Tie Model


To facilitate dealing with 3-D strut-and-tie models, the ACI 318-14 [5] recommended
the following; “In a three-dimensional strut-and-tie model, the area of each face of
a nodal zone shall not be less than that given in A.5.1, and the shape of each face of the
nodal zones shall be similar to the shape of the projection of the end of the struts onto
the corresponding faces of the nodal zones”, but in the case of the design of transfer
slabs, which are elements with large dimensions in the three directions, the design using
strut-and-tie model is governed by the nodal zones at the columns position. These nodal
zones are subjected to complex three dimensional states of stress and using a method
based on 2-D analogies and other simplifications appear to be inadequate in this case.
Therefore, revisions were made to the ACI Building Code where Section 23.9.5 was
added to simplify the detailing of nodal zones in 3-D, by not requiring exact geometry
compatibility between the struts and the faces of the nodal zone. Pile caps are a good
example to illustrate the three-dimensional analysis of transfer slabs.

4.3.1 Three-Dimensional Nodal Zones


Chantelot and Mathern, 2010 [9] proposed a solution to improve the design of nodal
zones, when the struts and ties joining a node are not in the same plane. Defining
consistent nodal zones is the aim of the method, which fulfill static equilibrium and with
compatibility between the faces of the nodal zone and the cross-sectional areas of the
struts and ties meeting at the node. The improvements proposed are justified by the
importance of the check of the strength of nodal zones in the design by strut-and-tie
models, as expressed by Schäfer, 1999 [33]; “Poor detailing of singular node regions is
the most frequent cause for insufficient bearing capacity of reinforced concrete
members”.

The same denomination of nodes as in 2-D will be used, that is to say one C for
every strut reaching the node and one T for every tie. In the definition of the nodal
zones, the tensile stresses from the ties will be represented as compression acting on the
other side, like in the two-dimensional case.

37
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

4.3.1.1 Geometry and Examples of Three-Dimensional Nodal Zones


The geometry of the three-dimensional nodal zones shown in Fig. 4.3 is consistent. To
define a consistent nodal geometry, Chantelot and Mathern, 2010 [9] supposed a
three-dimensional method. This method consists in determining the shape of the
undefined struts, using the known or assumed corners of the nodal zone and the
direction vector parallel to the axis of the strut, so in order to define the remaining
struts, the following parameters of the nodal zones should be known:

 the dimensions of the supporting areas and loading areas,


 the height of the node, which is defined by the height of horizontal struts or the
height of influence areas of ties (as illustrated in Sec. 2.4.2.2)

(a) 5C-node under the column with four inclined struts

(b) 6C-node under the column for a 5-pile cap (three of the struts are shown)

Fig. 4.3 Examples of nodal zones in transfer slabs [9].

38
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

(c) 9C-node under a column for an 8-pile cap, combination of the two previous
nodes (two struts are shown at each level)
Fig. 4.3 Cont.

4.3.1.2 Nodes With Overlapping Struts in The Same Quadrant


For common singular nodes in a pile cap, an orthogonal basis can be defined at the node
by the vertical direction of the external load at the support, and the horizontal
orthogonal directions of the main reinforcement in the structure. In three-dimensions,
this orthogonal basis defines eight quadrants, of which four are located inside the
structure while, in two-dimensions, it defines four quadrants with two inside the
structure. If two struts meeting at the node are located in the same quadrant, the
definition of the nodal region becomes more complex. Some authors treated this
problem in the case of two-dimensional strut-and-tie models. The different methods
proposed to solve this issue are illustrated in Fig. 4.4 (Chantelot and Mathern, 2010
[9]).

(a) Problem description (b) Sol. of Schlaich [36] (c) Sol. of Clyde [12]

Fig. 4.4 Nodes with overlapping struts [9].

39
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

In his solution of the problem, Schlaich, 1990 [36] systematically combines two
adjacent struts, in order to obtain only simple nodes formed by the intersection of three
struts, as illustrated in Fig. 4.4b. The solution of Clyde, 2008 [12] is actually a variant
of the solution proposed by Schlaich, without the intermediate struts between the
triangular areas, as illustrated in Fig. 4.4c. Consequently, the main difference between
the two solutions is that in Schlaich’s solution the axes of all the struts intersect at the
same point, while in Clyde’s solution the resulting forces of the struts are not
concurrent. However, the moment equilibrium is still verified as all the triangle areas
(ABC, ACD and ADE in Figs. 4.4b and 4.4c), which are common to both methods, are
in equilibrium. This solution corresponds to what is usually done in practice, that is to
say, to divide the support area in proportion to the incoming forces in the inclined struts
and to divide the node in sub-nodes (Chantelot and Mathern, 2010 [9]).
Nevertheless, these methods which are applicable quite easily in two-dimensions
cannot be applied to three-dimensional cases, because the interface between the struts
cannot be defined as easily as in 2-D and thus the consistency of the nodal zones would
not be preserved. An alternative method had to be found to conduct this work, which is
suitable for three-dimensional nodal zones. The method proposed here consists in
checking the stress at the face of the node, from a hypothetical strut, resultant of all the
converging struts located in the same quadrant. The axis of the resultant strut
corresponds to the average between the directions of the converging struts, while the
force in the resultant strut is equal to the projections of the forces in the different struts
on this axis. Then, the polygonal area of the resultant strut, on which the force is
checked, is found as explained in the previous section (Chantelot and Mathern, 2010
[9]).

4.3.1.3 Three-Dimensional Nodal Zone Strength


The strength value of the concrete can be increased in the case of nodes subjected to
triaxial compression where, there must be only struts joining at these nodes. Wang et
al., 1987 [40] conducted some experimental results of tests on cubes loaded by triaxial
compression as shown in Fig. 4.5.

40
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Fig. 4.5 Triaxial compression tests results (Wang et al., 1987 [40]).

The chart shows that rather low transversal stresses produce an important
increase of the bearing strength. For instance, if the two transversal stresses are equal to
the uniaxial cube strength, the strength in the third direction is raised to approximately
five times the uniaxial strength. In addition, if the two transversal stresses are equal to
20% of the uniaxial cube strength, the strength in the third direction is in the order of
two times the uniaxial strength. The refinement of the nodal zone improves the three
dimensional state of stress of the nodal zone, by assuring higher and more homogeneous
stresses on all its faces.
In such cases, the strength of concrete under a triaxial state of stress is
; where is the maximum stress and is the lateral confining pressure
(or the other two perpendicular stresses) (Park and Paulay, 1975 [30]). If is
assumed any value as small as (i.e., ), the corresponding value of
, with accounting the size effect .
Several standards give recommendations for the triaxial compressive strength.
Eurocode 2 [18] gives the following upper limit for the concrete design strength for
triaxial compression, , which may be used if the transverse stresses are known and
bigger than ,

( ) (4.1)

With (recommended value by EC2 [16])

41
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

where is the characteristic value of concrete compressive strength at 28 days


and is the design value of concrete compressive strength.
In the Recommendations of FIP, 1999 [20], the following value is recommended:
(4.2)
While in the ACI 318-14 [5], the design bearing strength of concrete, shall not
exceed , except when the supporting surface area, is wider on all sides than
the loaded area, (Fig. 4.6) then the following value is recommended:

√ , where√ (4.3)

Fig. 4.6 An application to find in stepped or sloped supports [5].

The ACI 318-14 upper limit on√ ⁄ limits the maximum strength of a node
under a triaxial state of stress, , to .

42
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

4.3.2 Three-Dimensional Struts


Three-dimensional struts are compression members which are considered the resultants
of parallel or fan-shaped compressive stress fields like in two-dimensional strut-and-tie
model. In three-dimensional structures such as, transfer slabs (e.g. pile caps), three-
dimensional struts could appears in plane which differs from other planes of struts and
ties, Fig. 4.7, where a and b are the dimensions of support, in the x- and y-direction,
respectively, and u refers to the height of the node.

Fig. 4.7 Inclined 3-D strut in 2C2T-node over a pile located at


the corner of a pile cap.
The main difference between struts in 2-D STM and in 3-D STM that these
struts may be surrounded by large volumes of plain concrete far from stressed regions,
called "inactive concrete". Therefore, in Section 4.3.3 Inactive Concrete and its
influence in Three-Dimensional Structures are further discussed.

43
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

4.3.2.1 Strut Limitation Angle


As in 2-D strut-and-tie models, the limitation of 3-D strut angle has to be respected due
to the need for strain compatibility and the limited ductility of concrete. The rules in
(Section 2.4.1.1) applied for 2-D models can be adopted in 3-D models. On the one
hand, limitations impose for the spreading of a concentrated load, that the main inclined
struts should be located in a cone, whose axis follows the direction of the load and
makes an angle of about 30° with the generatrix. On the other hand, the angle between
struts and ties has to be above a certain limit. However, in 3-D, the limitations should
apply to the real angle between the tie and the strut, which is different from the angle
between the tie and the projection of the strut in the vertical plane of the tie. As in 2-D,
this angle should be around 60 degrees, and not less than 45 degrees. This angle can
easily be calculated by the following formula (Chantelot and Mathern, 2010 [9]).

⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗ ⃗⃗⃗⃗⃗⃗⃗⃗
(|⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗ | |⃗⃗⃗⃗⃗⃗⃗⃗ |) (4.4)

4.3.3 Inactive Concrete in Three-Dimensional Structures


As a characteristic geometry of transfer slabs, such as pile caps, very significant
volumes of concrete that are subjected to low stresses are called "inactive concrete". In
well-designed strut-and-tie models, these volumes are the ones that are far from any
strut or tie and confinement by these inactive concrete is a feature of great interest in the
model developed in this thesis work. This feature is clearly pointed out in a strut-and-tie
model, for example in Fig. 4.8. In reinforced concrete structure, when the element is
loaded, these important volumes of inactive concrete give rise to important internal
restraint.

Fig. 4.8 Inactive concrete in pile cap.

44
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

4.3.3.1 Effect of Inactive Concrete on Bearing Capacity of Structure


Internal restraint from inactive concrete makes positive and negative effect on the
bearing capacity of the structure. Compressive stresses induced due to this internal
restraint have a positive effect, this effect is called “inactive concrete confinement”; it
can be defined as the radial compression that develops around the compressive struts far
from nodal regions.
Confinement by inactive concrete reduces greatly the tendency of compressive
struts to develop transverse tensile stresses within the strut, hence increasing the
compressive capacity of these struts. However, inactive concrete confinement does not
have an important effect on the capacity of the nodal areas below the columns, which
already subjected to a triaxial state of stress due to the loading conditions.
Both triaxial compression due to the loading and to the confinement lead to the
choice of a failure criterion for concrete subjected to a multi-axial state of stress. In
qualitative terms, the compressive capacity of a concrete strut is enhanced by
compression in other directions and decreased by tension in transverse directions. On
the other hand, deformation limitations due to internal restraint from inactive concrete
have some negative effects on bearing capacity of the structure. Indeed, even for rather
small deformation or load levels, a highly internally restrained structure can develop
wide cracks that would deteriorate the bearing capacity of the structure. As a result,
important redistribution of forces cannot take place in the structure because they require
large deformations to develop. When the structure is loaded, the inactive concrete
restrains for instance and steel reinforcement can hardly develop its ultimate strength
because the need for deformation is not acceptable for the highly restrained structure
(Chantelot and Mathern, 2010 [9]).

4.3.3.2 Bearing Strength of Compressive Struts Confined by Inactive Concrete


In deep members such as pile caps, Adebar and Zhou [1,2] stated that compression
struts don't fail by crushing of concrete and the failure may occurs due to longitudinally
splitting of these struts due to the transverse tension caused by spreading of compressive
stresses. The spreading of the compression will cause biaxial or triaxial compression in
the nodal zone, but transverse tension near the mid-length of the strut, Fig. 4.9a. A
refined truss model, which includes a tension tie to resist the transverse tension, is
shown in Fig. 4.9b. The maximum bearing stresses are a good indicator of the
likelihood of a strut splitting failure.

45
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

(a) Linear elastic stress trajectories with (b) Refined truss model with
transverse tension due to spreading of concrete tension tie to resist
compression transverse tension

Fig. 4.9 Development of strut-and-tie model for deep member or pile cap.

In the case of plane stress, when the tension is resisted only in one direction field
as in walls or deep beams, the influence of the "amount of spreading" on the bearing
stress to cause transverse splitting is shown in Fig. 4.10 (Schlaich et al., 1987 [35]).
Based on the results in the Figure 4.10, Schlaich et al., 1987 deduced that the concrete
compressive stresses within a disturbed region can be considered safe if the maximum
bearing stress in all nodal zones is limited to .

Fig. 4.10 Maximum bearing stress to cause transverse splitting in


biaxial stress field.

In three-dimensional structures such as transfer slabs; e.g., pile caps, the


compressive stresses spread in two directions, reducing the transverse tension in any
one direction. For example, in the case of a double-punch test carried out by Chen and
Trumbauer, 1972 [11] and Adebar and Zhou, 1996 [2], as shown in Fig. 4.11, a crack
will form at the middle of the cylinder height first, reducing the capacity of the
specimen to carry additional compression thus leading to the cylinder failure.

46
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Fig. 4.11 Splitting in a concrete cylinder, tension develops in radial directions [9].

Nevertheless, in the case of double-punch, the compressive stresses spread in all


radial directions resulting in radial tension in the middle of the specimen also resisted in
all radial directions. Hence, in a double-punch test the risk of splitting is less decisive
than for the two-dimensional case because:
 The tension in the bottle shape is reduced due to the confinement provided by
surrounding concrete, see Section 4.3.3.1.
 The tensile stresses are reduced in each single direction, as the tension is resisted
in all radial directions. (Chantelot and Mathern, 2010 [9])

Adebar and Zhou, 1993 state that to better understand the transverse splitting
phenomenon, idealized compression struts were studied. Linear elastic finite elements
were used to determine the triaxial stresses at first cracking within cylinders of various
diameters and height subjected to concentric axial compression over a constant-
size circular bearing area of diameter , see Fig. 4.12a. The geometry of the problem
can be summarized in terms of two parameters, namely the ratio of the cylinder
diameter to the load diameter ⁄ , and the ratio of the cylinder height to the loaded
diameter ⁄ .

47
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Fig. 4.12 Analytical study of transverse tension in triaxial stress field [2].

Figs. 4.13a and 4.13b summarized the influence of ⁄ and ⁄ on the


bearing stress at first cracking, assuming the ratio of compressive strength to tensile
strength . When ⁄ , the cylinder is subjected to cylinder uniaxial
compression, while ⁄ , the compression stresses are spread out, creating
triaxial compression close to the loaded surface, and biaxial tension near the mid-height
of the cylinder, Fig. 4.13b.

a) Influence of confinement b) Influence of height

Fig. 4.13 Analytical study of the ratio between stresses at cracking, to concrete

strength, (Adebar and Zhou, 1993 [1]).

It is interesting to note that for large ⁄ and ⁄ values, increasing the


amount of confining concrete (i.e., increasing ⁄ and ⁄ further) does not
significantly change the internal stress flow and hence does not affect bearing stress
which causes first cracking beyond a certain limit.

48
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Adebar and Zhou, [1,2] proposed that when designing deep member (disturbed
regions) without sufficient reinforcement to insure redistribution after cracking, the
maximum bearing stress should be limited to

( ) (4.5)

(√( ⁄ ) ) (4.6)

(( ⁄ ) ) (4.7)

where the ratio ⁄ which represents the aspect ratio (height/width) of the compression
strut, should not be taken less than 1.0 (i.e., ⁄ ). The parameter accounts for
the amount of confinement, while the parameter accounts for the geometry of the
compression stress field.

The upper limits on and were set by Adebar and Zhou to guarantee an
upper limit of which corresponds approximately to the upper limit of bearing
strength given in the ACI-318-14 [5]. The intersection of confinement and geometry
(aspect ratio) was chosen to give a reasonably simple expression and yet correspond
well with the finite element predictions and the experimental results. The lower bearing
stress limit of is appropriate if there is no confinement (i.e., √( ⁄ ) ),
regardless of the height of the compression strut, as well as when the compression strut
is relatively short ( ⁄ 1), regardless of the amount of confinement. For actual
structural members the size effect has to be implemented leading to the following
design value

( )( ) (4.8)

The concrete bearing strength is actually proportional to the concrete tensile


strength. If the concrete compressive strength is significantly greater than 34.5 MPa,
Eq. (4.5), (which is similar in form to the ACI bearing stress limit) may become unsafe.
In that case a more appropriate limit for the bearing stress is given by

√ ( ) (4.9)

where is in MPa. Eq. (4.8) expresses the bearing strength enhancement in a form
similar to what was proposed by Hawkins, 1968 [22].

49
Chapter 4 3-D Strut-and-Tie Model Applied to Pile Caps

Thus, the design bearing strength of a strut in pile cap if the concrete compressive
strength is significantly greater than 34.5 MPa should be

( ) √( )( ) (4.10)

4.4 Comparison between 2-D STM and 3-D STM


With large three-dimensional reinforced concrete blocks, load paths for the concentrated
load may be considered in two directions; this is the approach adopted in Fig. 4.14,
where the strut-and-tie method is applied in two perpendicular planes. Unlike 2-D
STMs, 3-D STMs are required when the structure and loading are considerably spread
over all three dimensions, such as pile caps with two or more rows of piles.

Fig. 4.14 Applied strut-and-tie model in two different planes.

One of the major differences between 2-D STM and 3-D STM, the cross-
section area of the struts obtained by the 3-D model is always greater than or equal to
the cross-section area of the struts in the 2-D model, which allows in most of the cases a
higher lever arm, therefore reducing the force in the horizontal struts and ties, hence the
flexural reinforcement. Thus, the design which takes 3-D STM into considerations is
more economic than a design uses the 2-D analogy method and hence leads to a more
efficient design.

50
CHAPTER 5
3-D NONLINEAR FINITE ELEMENT MODELING

5.1 Introduction
For its power and versatility, the finite element method is utilized in this study in order
to trace the stress trajectories, displacements, strains, crack pattern at different stages of
loading and distribution of normal and shear stresses in concrete. The ANSYS 12.0
computer program is employed in the analysis of pile caps investigated in the current
study. Concrete is modeled using a 3D- reinforced concrete element named SOLID65
which is capable of crushing in compression and cracking in tension and reinforcement
is modeled using LINK 8-3D bar element.
This chapter discusses model development for the full-size pile caps. Element
types used in the models are first reviewed. The material model for concrete and
reinforcement and the input data (geometry, meshing, boundary conditions, and loads)
are then defined. Finally, Comparisons will be made between obtained results and the
experimental data.

5.2 Finite Element Model by ANSYS


5.2.1 Element Types
5.2.1.1 Concrete
Concrete is a quasi-brittle material and has different behavior in tension and
compression. A 3-D solid element, SOLID65, was used to model the concrete. This
element has eight nodes with three degrees of freedom (ux, uy and uz) at each node. This
element is capable of plastic deformation, cracking (in three orthogonal directions) and
crushing. The geometry, nodes locations and the coordinate system for this element
type are shown in Fig. 5.1.

Fig. 5.1 SOLID65 element geometry (ANSYS, 2009[7]).

52
Chapter 5 3-D Nonlinear Finite Element Modeling

SOLID65 is used for the modeling of solids with/without reinforcing bars


(rebar). In the work presented in this thesis, discrete modeling technique is used for
modeling the reinforcement as bar or beam elements that are connected to concrete
mesh nodes. Fig. 5.2 shows the discrete models for reinforcement.

Fig. 5.2 Discrete models for reinforcement (ANSYS, 2009 [7]).

5.2.1.2 Reinforcement
A LINK8-3D element is used to model steel reinforcement. The 3-D spar element is a
uniaxial tension-compression element with three degrees of freedom at each node:
translations (ux, uy and uz) in the x, y, and z-directions. The geometry, node locations, and
the coordinate system for this element are illustrated in Fig.5.3. This element is also
capable of plastic deformation.

Fig. 5.3 LINK8-3D geometry (ANSYS, 2009 [7]).

5.2.2 Material Modeling in ANSYS


5.2.2.1 Concrete
Material model number 1 refers to the SOLID65 element which requires linear
isotropic and multilinear isotropic material properties to properly model concrete. As
seen in Table 5.1, there are multiple parts of the material model for each element. The
multilinear isotropic material uses von Mises failure criterion along with the (Willam
and Warnke, 1975 [6]) model to define the failure of concrete.

53
Chapter 5 3-D Nonlinear Finite Element Modeling

Table 5.1 Material models for SOLID65

Material Model No. 1

Element Type SOLID65

Material Properties

Linear Isotropic
-Elasticity Modulus, EX, is equal to at point on the curve

-Poisson’s Ratio, PRXY, is equal to 0.20


Multilinear Isotropic
-Six coordinates are needed to represent the stress-strain curve for
concrete, Fig. 5.4.
Concrete
-Open Shear Transfer Coeff. 0.3
-Closed Shear Transfer Coeff. 1.0

-Uniaxial Cracking Stress The concrete tensile strength is


typically 8% -15% of the compressive
(Modules of rupture)
strength or equal to 0.62√ [5].
-1.0
-Uniaxial Crushing Stress This value means that the crushing
stress value is taken from the stress-
strain curve.
-Biaxial Crushing Stress 0
-Hydrostatic Pressure 0
-Hydro Biax Crush Stress 0
-Hydro Uniax Crush Stress 0
-Tensile Crack Factor 0

The value of shear transfer coefficient βt ranges from 0.0 to 1.0, with 0.0
representing a smooth crack (complete loss of shear transfer) and 1.0 representing a
rough crack (no loss of shear transfer) (ANSYS, 2009 [7]).

A nonlinear elasticity model was adopted for concrete. In this model the
compressive uniaxial stress–strain values for the concrete model are calculated using the
following equations with which are obtained the multilinear isotropic stress–strain curve
for the concrete (Desayi and Krishnan, 1964 [14]).

54
Chapter 5 3-D Nonlinear Finite Element Modeling

⁄ ⁄ ) ) MPa (5.1)
where
⁄ MPa (5.2)
and at first point,
⁄ ⁄ MPa (5.3)
where is the stress at any strain and is the strain at the cylinder compressive

strength

√ (MPa), (ACI 318-14 [5]) for NSC (5.4)


The uniaxial tensile cracking stress,

√ (MPa), (ACI 318-14 [5]) for NSC (5.5)

Fig. 5.4 shows the simplified compressive axial stress-strain relationship that
was used in this study. This simplified stress-strain curve concrete is constructed from
six points connected by straight lines. The curve starts at zero stress and strain. Point
No.1, at, is calculated for the stress-strain relationship of the concrete in the line arrange
to satisfy Hooke’s Law, Eq. (5.3). Point No.2, No.3, and No.4 are obtained from Eq.
(5.1). Point No.5 is at and , indicating crushing strain for unconfined concrete.
After Point No.5, an assumption was made of perfectly plastic behavior for concrete
(William et al., 1975 [6], and Meisam, 2009 [28]).

Fig. 5.4 Simplified multilinear isotropic stress-strain curve for concrete


in compression.

55
Chapter 5 3-D Nonlinear Finite Element Modeling

5.2.2.2 Steel Reinforcement


The steel reinforcement used for the finite element models is assumed to be an elastic
perfectly plastic material, identical in tension and compression. The bilinear elastic-
plastic stress- strain for steel reinforcement to be used with Link8 element is furnished
in two sets of data. Modulus of elasticity of 2*105 N/mm2 and Poisson’s ratio of 0.3 is
used to setup a linear isotropic model, which is for the elastic range. It is also
experienced that for tangent modulus a small value of 10 to 20 N/mm2 shall be used to
avoid loss of stability upon yielding. In the present study, yield stress (fy) of 479 N/mm2
and 486 N/mm2 with tangent modulus of 20 N/mm2.

5.2.2.3 Bond between Concrete and Reinforcement


In this study, perfect bond between concrete and reinforcement has been assumed. The
implementation of such condition is illustrated in Fig. 5.5. The perfect bond condition is
introduced by using the command "Merge Items" in ANSYS program which merges
separate entities that have the same location.

Fig. 5.5 Element connectivity; concrete solid and link elements.

5.2.3 Non-linear solution


The Solution Controls command dictates the use of a linear or non-linear solution for
the finite element model. In nonlinear analysis, the total load applied to a finite element
model is divided into load steps, a series of load increments. Before proceeding to the
next load increment, the stiffness matrix of the model is adjusted to reflect nonlinear
changes in structural stiffness for a completion of each incremental solution. In the
particular case considered in this thesis the analysis is small displacement and static.

5.2.3.1 Time Stepping and Loading


In order to reduce the computation time required in the analysis, ANSYS offers an
alternate automatic time step option by determining the size of load increments between
sub-steps (ANSYS, 2009 [7]). Due to this advantage, this option is selected. The

56
Chapter 5 3-D Nonlinear Finite Element Modeling

objective of using this kind of analysis is to obtain the descending branch of the stress-
strain curve of the pile cap specimens under axial loading. In automatic time stepping
(or automatic loading), the program calculates an optimum time step at the end of each
sub-step, based on the response of the structure or component to the applied loads.
There are two features of the automatic time stepping. The first feature concerns the
ability to estimate the next time step size based on the current and past analysis
conditions and make proper load adjustments. The second feature is referred to as the
time step bisection component. To make a proper closed relation between load-
deflection curves, the applied loads are divided into sub-steps up to failure, the sub-
steps are the points within a load step at which solutions are calculated, Fig. 5.6. They
are used in nonlinear static to apply the loads gradually so that an accurate solution can
be obtained. The load is applied incrementally by axial pressure at the top surface of the
column of pile cap model and is increased incrementally with an automatic increment.

Fig. 5.6 Load steps, sub-steps, and time.

5.2.3.2 Newton-Raphson Iterative Solution


Newton-Raphson equilibrium iterations provide convergence at the end of each load
increment within tolerance limits. Fig. 5.7 shows the use of the Newton- Raphson
approach in a single degree of freedom nonlinear analysis. The ANSYS program uses
Newton-Raphson equilibrium iterations for updating the model stiffness.

57
Chapter 5 3-D Nonlinear Finite Element Modeling

Fig. 5.7 Newton-Raphson iterative solution (2 load increments) (ANSYS, 2009 [7]).

Prior to each solution, the Newton-Raphson approach assesses the out-of-


balance load vector, which is the difference between the restoring forces (the loads
corresponding to the element stresses) and the applied loads. To satisfy the state of
equilibrium, the load vector exists beyond the equilibrium should be zero.
Subsequently, the program carries out a linear solution, using the out-of-balance loads,
and checks for convergence. If convergence criteria are not satisfied, the out-of-balance
load vector is reevaluated, the stiffness matrix is updated, and a new solution is attained.
This iterative procedure is repeated until the problem converges is achieved (William et
al., 1975 [6]). A number of convergence-enhancement and recovery features, such as
line search, automatic load stepping, and bisection, can be activated to help the problem
to converge. If convergence cannot be completed, then the program attempts to solve
with a smaller load increment.
In some nonlinear static analyses, using Newton-Raphson method alone, the
tangent stiffness matrix may become singular (or non-unique), causing severe
convergence difficulties. Such occurrences include nonlinear buckling analyses in
which the structure either collapses completely or "snaps through" to another stable
configuration. For such situations, another method, the arc-length method, can be
activating an alternative iteration scheme in order to help avoid bifurcation points and
track unloading. This iteration method is presented schematically in Fig. 5.8.

58
Chapter 5 3-D Nonlinear Finite Element Modeling

Fig. 5.8 Arc-length method vs. Newton-Raphson method.

5.3 Case Study (Experimental Test of Pile Caps of Adebar, Kuchma and Collins)
5.3.1 Geometry and Details of Reinforcement
Adebar, Kuchma, and Collins tested six full-scale pile caps (five four-pile caps and one
six-pile cap) in order to examine the suitability of three dimensional strut-and-tie
models for the design of pile caps. Pile caps A, B, D and E were all four-pile caps with
identical external dimensions while pile cap C was supported on six piles and pile cap F
was made identical to pile cap D except that four ’’corner‘‘ pieces of plain concrete
were omitted. All these pile caps had an overall depth of 600mm and were loaded
through 300mm square cast in place reinforced concrete columns and supported by
200mm diameter precast reinforced concrete piles embedded 100mm into the underside
of the pile cap. The details of the six test specimens are shown in Fig. 5.9 and the
reinforcement details are given in Table 5.2. The measured properties of the concrete
used in producing the pile caps, are given in Table 5.3, where the secant modulus
corresponds to a stress of , and the proprieties of the reinforcing bars are given
Table 5.4 (Adebar et al., 1990 [3]).

59
Chapter 5 3-D Nonlinear Finite Element Modeling

a) Plan Views

b) Cross section of pile caps B, D, and F

Fig. 5.9 Geometry and details of reinforcement of the six test specimens (Adebar et
al., 1990 [3]).

60
Chapter 5 3-D Nonlinear Finite Element Modeling

Table 5.2 Summary of reinforcement


Tension tie mark Reinforcement Area of Depth to Bar spacing,
(see Fig. 5.10) steel steel, mm2 centroid, mm mm
TA1 9-No. 10* 099 440 260
TA2 15-No. 10 1500 450 100
4-No. 10 340
TB1 4-No. 10 1200 390 70
4-No. 10 440
6-No. 10 350
TB2 8-No. 10 2200 400 45
8-No. 10 450
3-No. 10 340
TC2 5-No. 10 1100 390 45
3-No. 10 440
7-No. 10 350
TC3 7-No. 10 2100 400 45
7-No. 10 450
4-No. 15+ 330
TD1 4-No. 15 2400 380 70
4-No. 15 430
8-No. 15 350
TD2 8-No. 15 4800 400 45
8-No. 15 450
TE3 9-No. 10 900 495 210
TE4 5-No. 10 500 485 240
1-No. 10 100 250 ---
1-No. 10 100 325 ---
TE5
1-No. 15 200 400 ---
1-No. 15 200 470 ---
2
* No. 10 is a 11.3 mm diameter bar with a cross-sectional area of 100 mm .
+
No. 15 is a 16.0 mm diameter bar with a cross-sectional area of 200 mm2

Table 5.3 Results of concrete cylinder test


Tensile strength from
Cylinder compression test
Pile indirect tension test, MPa
Cap Secant Peak stress, Strain at Split Double
modulus*, MPa MPa peak stress cylinder punch
A 19400 24.8 0.0024 2.9 ---
B 19400 24.8 0.0024 2.9 ---
C 26000 27.1 0.0020 3.7 ---
D 28000 30.3 0.0020 2.2 2.0
E 31000 41.1 0.0022 2.7 2.5
F 28600 30.3 0.0020 2.2 2.0
* Secant modulus corresponds to a stress of

61
Chapter 5 3-D Nonlinear Finite Element Modeling

Table 5.4 Proprieties of reinforcing steel

Bar destination Nominal area, mm2 Yield stress, fy MPa


No. 10M 100 479
No. 15M 200 486

5.3.2 Results of Experimental Test


A universal testing frame was used to apply concentric compressive load onto the
column and the piles were supported on steel pedestals, which in turn were supported on
rubber pads. The load on the specimens was increased monolithically to failure,
stopping at approximately 10 deformation stages to take manual displacement readings
and to photograph crack development. All pile caps typically had very few cracks
before failure. The flexural cracks in pile caps A, B, D, and E extended diagonally
between the piles [3].

5.3.2.1 Pile cap A


Pile cap A was predicted to fail at a total load of 2138 kN by the design Eq. of the ACI
Building Code, with flexural being critical, while the specimen failed at only 83% of the
this predicted load, namely 1781 kN and the first cracking load was 1186 kN. The
flexural reinforcement yielded prior to failure. Fig. 5.10 illustrates the final deformation
patterns at top, bottom and all side surfaces at failure [3].

a) bottom surface and sides b) Top surface and sides

Fig. 5.10 Final deformation pattern of pile cap A [3].

62
Chapter 5 3-D Nonlinear Finite Element Modeling

5.3.2.2 Pile cap B


Pile cap B was similar to pile cap A except in the reinforcement. The majority of the
load was initially carried to the two piles closest to the column. After the tension tie in
the short direction yielded, the load distribution among the piles began to change. The
specimens failed before significant redistribution occurred. The pile cap resisted a
maximum load of 2189 kN and the tension reinforcement in the long direction did not
yield. Fig. 5.11 shows the appearance of pile cap B after testing [3].

Fig. 5.11 Appearance of pile cap B after testing [3].

5.3.2.3 Pile cap C


As in pile cap B, the vast majority of the load was resisted by the two piles closest to the
column, while the four outer piles resisted very small load. When the total column load
reached 2892kN, a shear failure occurred while the first cracking load occurred at a load
1780kN. The tension reinforcement between the center two piles reached yield [3].

5.3.2.4 Pile cap D and E


Pile caps D and E, which were similar to pile cap B except that they had twice as much
reinforcement, both failed prior to yielding of either tension tie. Once again the failure
surfaces resembled typical punching shear cones. Pile cap D failed at a column load of
3222kN, while Pile cap E failed at a column load of 2709kN. Pile cap E was stronger
than pile cap D because of the distribution reinforcement in the specimen and also the
higher concrete strength [3].

63
Chapter 5 3-D Nonlinear Finite Element Modeling

5.3.2.5 Pile cap F


Pile cap F was similar to pile cap D except for the four ’’missing corner ‘‘pieces of
plain concrete were omitted. It looked like two perpendicular deep beams intersecting at
mid-span. Failure occurred when the short beam failed in shear. The sequence of
relevant crack formation in the shorter span beam is illustrated in Fig. 5.12. The first
cracks to form were flexural cracks at mid-span, which propagate up the interface of the
two beams labeled 1 in Fig. 5.12. Then, a vertical crack, labeled 2, appeared directly
above the pile. At a pile load of 949 kN, a new diagonal crack suddenly appeared (crack
4). Analogous to a web-shear crack, this diagonal crack formed independently of any
previously existing cracks. With a small increase in load, the diagonal crack widened
considerably. None of the longitudinal reinforcement yielded. The specimen failed
when the total column load was 3026kN [3].

Fig. 5.12 Sequence of relevant crack formation in pile cap F [3].

5.3.2.6 Summary of Experimental Results


All pile caps experimental cracking and failure loads results are illustrates in table 5.5.

Table 5.5 Summary of pile cap test results


Experimental Results
Pile cap No. of Reinforcement
First Cracking Ultimate
specimen piles MPa layout
Load (kN) Load (kN)
A 4 24.8 Grid 1186 1781
B 4 24.8 Bunched 1679 2189
C 6 27.1 Bunched 1780 2892
D 4 30.3 Bunched 1122 3222
E 4 41.1 Bunched and grid 1228 4709
F 4 30.3 Bunched 650 3026

64
Chapter 5 3-D Nonlinear Finite Element Modeling

5.4 Analysis of the Six Deep Pile Caps of Adebar, Kuchma, and Collins
As a comparative study, a nonlinear finite element analysis has been performed for the
six full-scale reinforced concrete pile caps tested by Adebar et al., 1990 [3], one six-
pile cap and five four-pile caps. The conditions of symmetry have been utilized to
reduce the computation effort by considering one quarter of the pile cap only. The
complete details of the six test specimens are illustrated in Fig. 5.9 and Table 5.2.
Based on their geometry, these six deep pile caps are divided into three groups;
i. Pile Caps A, B, D, and E,
ii. Pile Cap C,
iii. Pile Cap F

5.4.1 Pile Caps A, B, D, and E


Pile Caps A, B, D, and E were all four-pile caps with identical external dimensions. The
complete details of these pile caps are shown in Fig. 5.9 and Table 5.2. Fig. 5.13
indicates the geometry of one of them showing the investigated quadrant used in
ANSYS model.

Fig. 5.13 Investigated quadrants of tested pile caps A, B, D and E.

5.4.1.1 Model Description and Material Properties


Pile cap, column, and piles were modeled as volumes. Since a quarter of the pile cap is
being modeled, the model is 1180mm, long and 850mm, wide. The pile caps all had an
overall depth of 600mm and loaded through 300mm square column. They were
supported by 177mm equivalent square piles embedded 100mm into the underside of
the pile cap. The combined volumes of the column, piles, and pile cap are shown in Fig.
5.14.

65
Chapter 5 3-D Nonlinear Finite Element Modeling

Fig. 5.14 Pile cap model (volumes created in ANSYS).

Link8 elements were used to create the reinforcement. Table 5.6 shows the main
reinforcement and their effective depths for these pile caps in long and short directions.

Fig. 5.15 illustrates the distribution of steel reinforcement of each pile cap.

Table 5.6 The main reinforcement and their effective depths used in ANSYS

Pile Reinforcement Steel Depth to Reinforcement


Cap Short Dir. Long Dir. centroid, mm layout

A 9-No. 10 15-No. 10 445 Grid


4-No. 10 6-No. 10 345
B 4-No. 10 8-No. 10 395 Bunched
4-No. 10 8-No. 10 445
4-No. 10 3-No. 10 7-No. 10 345
C 4-No. 10 5-No. 10 7-No. 10 395 Bunched
4-No. 10 3-No. 10 7-No. 10 445
4-No. 15 8-No. 15 340
D 4-No. 15 8-No. 15 390 Bunched
4-No. 15 8-No. 15 440
E shown in Fig 5.10 and Table 5.2
4-No. 15 8-No. 15 340
F 4-No. 15 8-No. 15 390 Bunched
4-No. 15 8-No. 15 440

66
Chapter 5 3-D Nonlinear Finite Element Modeling

a) pile cap A

b) pile cap B

c) pile cap D
Fig. 5.15 Reinforcement configurations of pile caps A, B, D and E.

67
Chapter 5 3-D Nonlinear Finite Element Modeling

d) pile cap E

Fig. 5.15 Cont.


5.4.1.2 Meshing
The pile caps A, B, D, and E are modeled using a nonlinear solid element SOLID65. To
obtain good results from the SOLID65 element, the use of (quadrilateral, triangle,
hexahedral, or tetrahedral) mesh are recommended. The choices of that mesh depend on
the geometry of the pile cap. Due to the irregularity of the pile cap model, triangular and
rectangular elements were created. The Volume Sweep command was used to mesh the
model of column and piles. This properly sets the width and length of their elements to
be consistent with the elements and nodes in the concrete portions of the model.

The meshing of the reinforcement is a special case compared to the volumes.


No mesh of the reinforcement is needed because individual elements were created in the
modeling through the nodes created by the mesh of the concrete volumes. By using
sharing nodes option in ANSYS, SOLID65 and Link8 elements can be interconnected
one to another forming a single solid model which is capable of simulating the actual
behavior of reinforced concrete pile cap. However, the necessary mesh attributes as
described above need to be set before each section of the reinforcement is created. The
overall meshing of the pile cap volumes is shown in Fig. 5.16 for each pile caps A, B
and E.

68
Chapter 5 3-D Nonlinear Finite Element Modeling

a) pile cap A

b) pile cap B

c) pile cap E
Fig. 5.16 The overall meshing for pile caps A, B and E.

69
Chapter 5 3-D Nonlinear Finite Element Modeling

5.4.1.3 Loads and Boundary Conditions


At a plane of symmetry, the displacement in the direction perpendicular to the plane
was held at zero (i.e., ux=0 and uz=0), while the bottom surface of the piles must be
constraint at Y-direction which uy=0. In order to avoid stress concentration problems, an
axial pressure was implemented over the entire top surface of the column model. Fig.
5.17 shows loading and boundary conditions for a typical finite element model.

Fig. 5.17 Boundary conditions for a typical finite element model.

5.4.1.4 Finite Element Results


The time at the end of the load step refers to the ending load per load step. The final
load applied from the finite element analysis, is the last load before the solution
diverges. The cracking pattern(s) in of the pile cap can be obtained using the
Crack/Crushing plot option in ANSYS. In order to view the first and final cracking
modes of each pile cap, Vector Mode plots must be turned on. ANSYS program records
a crack pattern at each applied load step. The initial (first) cracking of the pile caps A,
B, D and E in the FE model corresponds to a load of 758.9kN, 9.160kN, 1295kN and
2066kN in the pile cap models, respectively. It is observed that the first crack, Fig.
5.18c, occurs diagonally (Shear cracks) at about 41 percent of the ultimate load in the
region between closest two piles.
Increasing the load capacity of pile caps is a result of the increase in the concrete
strength, as noticed in pile caps B and D; their ultimate loads are 2273kN and 4703kN,
respectively. The higher compressive stresses exist at the position of the point loads and
supports, while a reduction in the compressive stresses occurs in the inclined struts
joining the point loads and supports especially at the closest piles because of the

70
Chapter 5 3-D Nonlinear Finite Element Modeling

diagonal cracks across the load path. Table 5.7 shows the finite element results for the
first cracking load and ultimate load and Fig. 5.19 shows the output of “ANSYS”
Program (Vector plots of principal stresses, Deformed shape, 1st cracks and Cracks
pattern at failure load) for the tested pile caps A, B, D and E.

i. pile cap A ii. pile cap E

a) Vector plots of principal stresses of pile caps A and E

b) Deformed shape of pile cap B.

c) 1st cracks for pile caps B and D


Fig. 5.18 Output of “ANSYS” program for pile caps A, B, D and E.

71
Chapter 5 3-D Nonlinear Finite Element Modeling

i. pile cap A

ii. pile cap B

iii. pile cap D iv. pile cap E


d) Cracks pattern at failure load for pile caps A, B, D, and E

Fig. 5.18 Cont.

72
Chapter 5 3-D Nonlinear Finite Element Modeling

5.4.1.5 Comparison of the Results


To study the accuracy of the nonlinear finite element approach, the obtained results are
compared with the test results of pile caps A, B, D and E from (Adebar, Kuchma and
Collins, 1990[3]). The recorded experimental cracking and ultimate failure loads and
the predicted cracking and failure loads for the tested pile caps calculated from the finite
element model are given in Table 5.7. The results in the table indicate that the
experimental and FEM results are in very good agreement. This demonstrates that the
nonlinear Finite element models of reinforced concrete pile caps, constructed in
ANSYS program using the dedicated concrete elements have accurately captured the
nonlinear response of these systems up to failure and can be considered a very good tool
in understanding the behavior of reinforced concrete transfer slabs such as pile cap.

Table 5.7 Comparison of ultimate loads

Experimental
Numerical Results
Results
1st Ultimate Max. Stresses 1st Ultimate
Pile
Cracking Failure Stresses (Strain) Cracking Failure
Cap Load, Load, in in Load, Load,
Conc. Steel rft.
FEM FEM near near Exp. Exp.
failure, failure,
(kN) (kN) (MPa) (MPa) (kN) (kN)
496
A 758.9 1744 16.2 1186 1781 0.44 0.98
(0.0043)
483
B 861.9 2273 18 1679 2189 0.38 1.04
(0.0033)
420
C 1100 2864 25.4 1780 2892 0.38 0.99
(0.0021)
260
D 1205 3163 31 1122 3222 0.38 0.98
(0.0013)
200
E 2066 4703 39 1228 4709 0.44 1.00
(0.001)
300
F 1377 2893 29.6 650 3026 0.48 0.96
(0.0015)

Average 0.42 0.99

73
Chapter 5 3-D Nonlinear Finite Element Modeling

5.4.2 Pile Cap C


Pile cap C was supported on six piles; the complete details of this pile cap are shown in
Fig. 5.9 and Table 5.2. Fig. 5.19 indicates the geometry of pile cap C showing the
investigated quadrant used in ANSYS model.

Fig. 5.19 Investigated quadrant of tested pile cap C.

5.4.2.1 Model Description and Material Properties


As described above in Sec. 5.4.1.1, the combined volumes of the pile cap C (Front, Top,
Side, and 3-D views) are shown in Fig. 5.20.

Fig. 5.20 Pile cap model (volumes created in ANSYS).

74
Chapter 5 3-D Nonlinear Finite Element Modeling

Link8 elements were used to create the reinforcement. Table 5.6 shows the
main reinforcement and their effective depths for pile cap C in long and short directions
and Fig. 5.21 illustrates the distribution of steel reinforcement of pile cap C.

Fig. 5.21 Reinforcement configurations of pile cap C.

5.4.2.2 Meshing
Due to the shape uniformity of pile cap C and unlike pile caps A, B, D and E, the
rectangular elements were created. The overall mesh of the concrete pile cap, column,
and support volumes is shown in Fig. 5.22 for pile cap C.

Fig. 5.22 The overall meshing for pile cap C.

75
Chapter 5 3-D Nonlinear Finite Element Modeling

5.4.2.3 Loads and Boundary Conditions


Following the same procedure in Sec. 5.4.1.3, the loading and boundary conditions are
considered here; Fig. 5.23 illustrates a typical finite element model.

Fig. 5.23 Boundary conditions for a typical finite element model.

5.4.2.4 Finite Element Results


The initial (first) cracking of the pile cap C in the FE model corresponds to a load of
1100kN, and it can be seen that the first crack occurs diagonally (Shear crack) at about
38 percent of the ultimate load in the region between closest two piles. The same
observations for high and low stresses in concrete in pile caps A, B, D and E, are
observed here too.
Table 5.7 shows the finite element results for the first cracking load and
ultimate load and Fig. 5.24 shows the output of “ANSYS” Program (Vector plots of
principal stresses, deformed shape, 1st cracks and Cracks pattern at failure load)
for tested pile cap C.

76
Chapter 5 3-D Nonlinear Finite Element Modeling

i. 3-D view ii. Side view


a) Vector plots of principal stresses of pile cap C

b) Deformed shape c) 1st cracks position

d) Cracks pattern at failure

Fig. 5.24 Output of “ANSYS” program for pile cap C.

77
Chapter 5 3-D Nonlinear Finite Element Modeling

From the results in Table 5.7 for pile cap C, it is obvious that the experimental
and FEM results are in very good agreement. Again, the nonlinear Finite element
models of reinforced concrete pile caps, constructed in ANSYS program using the
dedicated concrete elements have accurately captured the nonlinear response of these
systems up to failure.

5.4.3 Pile Cap F


Pile Cap F was similar to pile cap D, supported on four piles, except for the four
‘‘missing corners’’ of plain concrete. The complete details of this Pile cap are shown in
Fig. 5.9 and Table 5.2. Fig. 5.25 illustrates the geometry of pile cap F showing the
investigated quadrant used in ANSYS model.

Fig. 5.25 Investigated quadrant of tested pile cap F.


5.4.3.1 Model Description and Material Properties
Following the same procedure in Sec. 5.4.1.1, the combined volumes of pile cap F are
shown in Fig. 5.26.

Fig. 5.26 Pile cap model (volumes created in ANSYS).

78
Chapter 5 3-D Nonlinear Finite Element Modeling

Table 5.6 shows the main reinforcement and their effective depths for these pile
caps in long and short directions. Fig. 5.27 illustrates the distribution of steel
reinforcement of each pile cap.

Fig. 5.27 Reinforcement configurations of pile cap F.

5.4.3.2 Meshing
The overall mesh of the concrete pile cap, column, and support volumes is shown in
Fig. 5.28 for pile cap F.

Fig. 5.28 The overall meshing for pile cap F.

5.4.3.3 Loads and Boundary Conditions


As explained before in section 5.4.1.3, Fig. 5.29 shows the loading and boundary
conditions for a typical finite element model.

79
Chapter 5 3-D Nonlinear Finite Element Modeling

Fig. 5.29 Boundary conditions for a typical finite element model.

5.4.3.4 Finite Element Results


The initial (first) cracking of the pile cap F in the Finite Element model corresponds to a
load of 1377kN, and it can be seen that the first crack occurs diagonally (Shear crack)
at about 48 percent of the ultimate load in the region between closest two piles. The
same observations for high and low stresses in concrete in pile caps A, B, D and E, are
observed here too. Failure occurred in the short direction in shear.
Table 5.7 shows the finite element results for the first cracking load and ultimate load
and Fig. 5.30 shows the output of “ANSYS” Program (Vector plots of principal
stresses, deformed shape, 1st cracks and Cracks pattern at failure load) for tested
pile cap F. The results in the table indicate that the experimental and FEM results are in
good agreement. Again, the nonlinear Finite element models of reinforced concrete pile
caps, constructed in ANSYS program using the dedicated concrete elements have
accurately captured the nonlinear response of these systems up to failure.

a) Vector plots of principal stresses b) Deformed shape

Fig. 5.30 Output of “ANSYS” program for pile cap F.

80
Chapter 5 3-D Nonlinear Finite Element Modeling

c) 1st cracks for pile caps F

d) Cracks pattern at failure load for pile cap F


Fig. 5.30 Cont.

5.5 The Effect of Pile Cap Thickness on Pile Load Distribution


The designers of pile caps usually assumed a uniform distribution of loads among piles,
assuming that they are rigid. But Ghali, 1999 [21] deduced that the design of the pile
cap as a rigid footing is acceptable when the ratio ⁄ does not exceed 2.4 and if the

ratio ⁄ exceeds 2.4, a flexible behavior is expected and the support reactions have to
be determined by an analysis that reflects the behavior of the actual structure; where
is the distance between the column centerline and the center of the furthermost corner
pile and is the thickness of the pile cap.

81
Chapter 5 3-D Nonlinear Finite Element Modeling

A three-dimensional linear finite element analysis has been conducted on one of


the pile caps A, B, D, and E with different depths (0.6, 1.0, 1.40, 1.80 and 2.20 m) in
order to illustrate the effect of pile cap thickness on pile load distribution. Fig. 5.31
indicates that when the pile cap thickness increases the pile reactions on the closest two
piles decrease unlike the loads on the other piles.
80

70
% of Total Load

60

50

40

30

20

10

0
depth 0.60 m 1.0 m 1.4 m 1.8 m 2.2 m
Closest two piles 72.46 61 55.4 53 50
Other piles 27.54 39 44.6 47 50

Fig. 5.31 Effect of pile cap rigidity on piles load


5.6 Conclusions
From the results of the three dimensional nonlinear finite element analysis of the tested
pile caps, the following conclusions can be drawn:

 The finite element model (FEM) is able to simulate the behavior of reinforced
concrete transfer slabs such as pile cap.
 For the six pile caps, the first crack occurs in the constant shear region (shear-
span). It, in general, happens in the short direction of pile cap and it is a diagonal
crack. When increasing the loads, the cracking increases towards the supports.

 Confinement of plain concrete can increase the bearing capacity of the model, as
in the case of pile caps D and F.
 The cracking load and failure load predicted from the finite element models are
very close to measured values and the crack patterns at the ultimate loads
correspond well with the observed failure modes of the experimental
observations.
 The obtained straining actions of the pile caps are influenced by the rigidity of
the pile caps.

82
CHAPTER 6
3-D STRUT-AND-TIE MODELS
6.1 Introduction
The strut-and-tie model design method considers the flow of forces within pile-caps
rather than just the forces at one particular section. In the STM approach, the forces in
the pile caps are derived from idealized models. Concrete inside 3-D structures such as
pile caps is significantly more confined than concrete inside 2-D structures; e.g., beams,
walls, and slabs, and this is the basic difference between pile caps and other structure.
Therefore, using the same nominal compressive strength of concrete per as those
commonly used for 2-D structures may be too conservative for the well-confined
concrete in pile caps. Besides, the allowable contact stress between a pile and a pile cap
is greater than the values allowed in node design as per the ACI.

For 3-D analysis using the STM in which the load paths for the concentrated load
may be considered in two directions, especially in large three-dimensional reinforced
concrete blocks such as pile caps, the strut-and-tie method is applied in two directions.
In each direction, the vertical position of the top nodes located directly beneath the
column (and horizontal strut) must be determined. When the strength of nodal zones is
sufficient, it is assumed that the strength of concrete struts is also sufficient because of
the following reasons:

- The available area of concrete struts is larger than that of nodal zones.
- The concrete struts in pile caps are well confined inside the pile caps.

This chapter introduces the approach for how to predict the strength of pile caps
using 3-D STMs based on the flow of forces from elastic analysis and/or following the
load path method. The method is demonstrated by the application to six case studies;
pile caps A, B, C, D, E and F tested by Adebar et al., 1990 [3]. The STM results are
compared with both the experimental data and the finite element results.

6.2 Case Study 1 - Pile Cap A


Input data:
Pile cap size; overall depth, h=600 mm, number of piles = 4 piles, square column of
size 300mm, circular piles of diameter 200mm, with 100mm embedded into the
underside of the pile cap. Fig. 6.1 shows the geometry of pile cap A.

84
Chapter 6 3-D Strut-and-Tie Models

Fig. 6.1 Geometry of pile cap A.

Materials; the cylinder compressive strength of concrete, = 24.8MPa, the yield stress
of longitudinal steel, =479MPa, the area of main reinforcement, As (TA1) =9No.10
(900mm2), with depth to steel centroid= 440mm in the short direction and As (TA2)
=15No.10 (1500mm2), with depth to steel centroid= 450mm in the long direction. The
diameter of No. 10 is 11.3mm and the cross-sectional area is 100mm2.

Strut-and-Tie Modeling
In order to predict the failure load of pile cap A, Fig. 6.2 illustrates a simple 3-D strut-
and-tie model for pile cap A. The concentrated column load is transmitted directly to
the four supports (piles) by four inclined compression struts while, the horizontal
tension ties (longitudinal reinforcement) are required to prevent the piles from being
spread apart. Nodes represent the connections between struts and ties. The analysis will
be performed in both the short and long directions in the following sections.

Fig. 6.2 Proposed 3D-STM of pile cap A.

85
Chapter 6 3-D Strut-and-Tie Models

Analysis in short direction, Sec. 1-1

1. The internal lever arm, hs1:


As shown in Fig. 6.3, the term u1 (height of lower node ) can be computed from;
( ) , where is the number of steel layers, is the
longitudinal steel diameter, is the clear concrete cover, and is the clear distance
between bars or,
( )
( )

The top bearing area of the column is divided into four equal triangles; each one
corresponds to the load transferred to a pile. For simplification in the calculations the
area of each triangle is replaced with an equivalent rectangular area having the same
centroid of the triangle as shown in Fig. 6.1.

The width of horizontal strut, can be computed from Eq. (6.1).


(6.1)
, where is upper limit strength of the upper nodal zone C-
C-C due to confinement, was chosen to correspond approximately to the upper limit of
bearing strength given in the ACI-318-14[5].

Fig. 6.3 STM of pile cap A in the short direction.

86
Chapter 6 3-D Strut-and-Tie Models

Assume 7 bars of main reinforcement in the short direction are required to resist and
is assumed to be the equivalent width of the upper node ( ), then

thus, ( )
( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
The upper width of strut , , and the lower width, , as shown in Fig. 6.4 can
be calculated based on the dimension of column and piles as follows;

 The width of the strut at the upper node, , Fig. 6.4a;

 For the lower node, the width of the strut , , Fig. 6.4b;

a) Upper Node b) Lower Node

Fig. 6.4 Nodes of pile cap A in the short direction.

4. STM forces:
From equilibrium of the model joints, the following relations can be written:

87
Chapter 6 3-D Strut-and-Tie Models

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


The width of the diagonal strut changes linearly between the upper and lower nodes;
(uniformly tapered). The effective compression strength of this strut differs at its two
ends, due either to different nodal zone strengths at the two ends (Upper Node and
Lower Node), or to different bearing areas.

The compressive strength of concrete in this strut may be significantly enhanced


with confinement (Sec. 4.3.3.2). Therefore, the maximum bearing stress should be
limited to;

( )( ) (6.2)

The parameter accounts for the amount of confinement (Fig 6.5),


(√( ⁄ ) ) (6.3)

a) Column nodal zone b) Pile nodal zone

Fig. 6.5 Amount of confinement for pile caps A, B, D and E.

 At the column nodal zone (Upper Node);

√( ⁄ ) √ , this gives ( ) , take

88
Chapter 6 3-D Strut-and-Tie Models

 At the pile nodal zone (Lower Node);

√( ⁄ ) √ , this gives ( )

While the parameter accounts for the geometry of the compression stress field,

[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression


strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, the bearing area is column

cross sectional area divided by the number of piles; thus, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.97 ) and the node strength (1.7 ). Hence, the smallest value of (0.97 )
is governing the strut strength at the Upper Node.

89
Chapter 6 3-D Strut-and-Tie Models

Upon substituting in where is assumed to be the equivalent


upper node width ( );

[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.97 ) and the node strength (0.65 ). Hence, the smallest value of
(0.65 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

It is noted that there is a significant difference between the strength of strut at the
upper and lower nodes, and this can be attributed to the significant difference between
the strength of the upper nodal zone, 1.7 and the strut strength, 0.97 . In order to
reduce the gap between to the two values of strut strength, the height of upper node
'
should be estimated based on a strength lower than 1.7 f c .
Upon using a strength of the upper node ( ⁄ ) , and
redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the angle


. The width of strut at the upper node, and at the
lower node, . The pile force , the force in the
diagonal strut, .
 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

90
Chapter 6 3-D Strut-and-Tie Models

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.96 ) and the node strength (1.2 ). Hence, the smallest value of (0.96 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );

[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.96 ) and the node strength (0.64 ). Hence, the smallest value of
(0.64 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

In order to determine the predicted failure load in this direction, Table 6.1
summarizes the calculations for the critical struts and nodes in the short direction.

91
Chapter 6 3-D Strut-and-Tie Models

Table 6.1 Summary of calculation results of short direction of pile cap A

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN

335.3 335.3 yes


Struts
520.6 582.7 yes
Lower node,
397.5 634.9 yes
C-C-T

Hence, in this direction

Analysis in long direction, Sec. 2-2


Assume 11 bars of main reinforcement in the long direction are required to resist .
As performed for sec. 1-1

1. The internal lever arm, hs2:


As shown in Fig. 6.6, the term u2 (height of lower node )
( )

( )

Fig. 6.6 STM of pile cap A in the long direction.

2. Angle of inclined strut,


, from pile cap geometry.

92
Chapter 6 3-D Strut-and-Tie Models

3. Widths of struts:
 The width of the strut at the upper node, , Fig. 6.7a;

 For the lower node, the width of the strut , , Fig. 6.7b;

a) Upper Node b) Lower Node

Fig. 6.7 Nodes of pile cap A in the long direction.

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before in Fig. 6.5, for Upper Node and for Lower
Node. In order to obtain the parameter , the ratio ( ⁄ ) is calculated in the manner as
for the short direction. The strut vertical height . For the Upper
Node, the bearing area is , and for the Lower Node, .
Thus,

93
Chapter 6 3-D Strut-and-Tie Models

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.97 ) and the node strength (1.7 ). Hence, the smallest value of (0.97 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.97 ) and the node strength (0.65 ). Hence, the smallest value of
(0.65 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;

[ ]

Take the smallest value

94
Chapter 6 3-D Strut-and-Tie Models

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

In order to determine the predicted failure load in this direction, Table 6.2
summarizes the calculations for the critical struts and nodes in the long direction.

Table 6.2 Summary of calculation results of long direction of pile cap A

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN

526.9 526.9 yes


Struts
620.6 540.4 No, 87.1%
Lower node,
327.1 642.0 yes
C-C-T
Hence, in this direction

Finally, the capacity of pile cap A

( ) ( )

6.3 Case Study 2 - Pile Cap B


Input data:
Pile cap size; overall depth, h=600 mm, number of piles = 4 piles, square column of
size 300mm, circular piles of diameter 200mm, with 100 mm embedded into the
underside of the pile cap. Fig. 6.8 shows the geometry of pile cap B.

Materials; the cylinder compressive strength of concrete, =24.8 MPa, the yield stress
of longitudinal steel, =479MPa, the area of main reinforcement, As (TB1) = 12No.10
(1200mm2), with depth to steel centroid= 390mm in the short direction and As (TB2)
=22No.10 (2200mm2), with depth to steel centroid= 400mm in the long direction.

95
Chapter 6 3-D Strut-and-Tie Models

Fig. 6.8 Geometry of pile cap B.

Strut-and-Tie Modeling
Fig. 6.9 illustrates a simple three-dimensional strut-and-tie model for pile cap B. The
concentrated column load is transmitted directly to the four supports (piles) by four
inclined compression struts while, the horizontal tension ties (longitudinal
reinforcement) are required to prevent the piles from being spread apart. Nodes
represent the connections between struts and ties. The analysis will be performed in
both the short and long directions in the following sections.

Fig. 6.9 Proposed 3D-STM of pile cap B.

Analysis in short direction, Sec. 1-1

1. The internal lever arm, hs1:


As shown in Fig. 6.10, the term u1 (height of lower node ) can be computed from;

( ) , where is the number of steel layers, is the


longitudinal steel diameter, is the clear concrete cover, and is the clear distance
between bars or,

96
Chapter 6 3-D Strut-and-Tie Models

( )
( )

The width of horizontal strut, can be computed from Eq. (6.4).


(6.4)

Fig. 6.10 STM of pile cap B in the short direction.

, where is the strength of the upper nodal zone C-C-C


node and is the equivalent width of the upper node ( ), then

thus, ( )

( )

2. Angle of inclined strut,

, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.10 based on the dimension of column and piles as follows;

97
Chapter 6 3-D Strut-and-Tie Models

 The width of the strut at the upper node, , Fig. 6.11a;

 For the lower node, the width of the strut , , Fig. 6.11b;

a) Upper Node b) Lower Node

Fig. 6.11 Nodes of pile cap B in the short direction.

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before in Fig. 6.5, for Upper Node and for Lower
Node.

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes).

98
Chapter 6 3-D Strut-and-Tie Models

From trigonometry and for easiness in the calculations, the ratio ( ⁄ ) can expressed as
( ⁄ ), where is the strut vertical height and is the dimension of the bearing
area of the nodal zone connected with the strut, with equal to the square root of this
bearing area.

For strut , . For the upper node, the bearing area is column

cross sectional area divided by the number of piles; thus, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.86 ) and the node strength (1.7 ). Hence, the smallest value of (0.86 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

99
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.86 ) and the node strength (0.58 ). Hence, the smallest value of
(0.58 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

It is noted that there is a significant difference between the strength of strut at the
upper and lower nodes, and this can be attributed to the significant difference between
the strength of the upper nodal zone, 1.7 and the strut strength, 0.86 . In order to
reduce the gap between to the two values of strut strength, the height of upper node
'
should be estimated based on a strength lower than 1.7 f c .

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the


angle . The width of strut at the upper node, and at
the lower node, . The pile force , the force in the
diagonal strut, .
 At the column nodal zone (Upper Node);
( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);

( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

100
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.82 ) and the node strength (1.0 ). Hence, the smallest value of (0.82 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );

[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.82 ) and the node strength (0.56 ). Hence, the smallest value of
(0.56 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
The There is no need to carry out any checks for the upper node since the strength of
the node is greater or equal to the strength of the surrounding struts. The critical node is
the lower node, and as estimated before, . Based on this, the nominal
value of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following.

Since a value of ( ) was used in estimating the


width of strut at the lower node, a new value of the strut width should be calculated;

101
Chapter 6 3-D Strut-and-Tie Models

This value of leading to a new nominal value of the strut strength at this node;
[ ]

This last value of the nominal strength, the strut is still greater that the value at the
upper node; and therefore, the solution should proceed with no change for this strut.

In order to determine the predicted failure load in this direction, Table 6.3
summarizes the calculations for the critical struts and nodes in the short direction.

Table 6.3 Summary of calculation results of short direction of pile cap B

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN

574.8 574.8 yes


Struts
800.4 656.0 No, 82.0%
Lower node,
555.8 553.2 No, 99.5%
C-C-T
From the results in the table, the size of the upper nodal zone should be reduced to
82.0%. This finally gives a value of . It should be noted that the
reinforcing steel does not yield at this obtained pile load.

Analysis in long direction, Sec. 2-2

As performed for sec. 1-1


1. The internal lever arm, hs2:
As shown in Fig. 6.12, the term u2 (height of lower node )

Fig. 6.12 STM of pile cap B in the long direction.

102
Chapter 6 3-D Strut-and-Tie Models

( )

( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
 The width of the strut at the upper node, , Fig. 6.13a;

 For the lower node, the width of the strut , , Fig. 6.13b;

a) Upper Node b) Lower Node

Fig. 6.13 Nodes of pile cap B in the long direction.


4. STM forces:

5. Checking stress limits;


a. Struts:

103
Chapter 6 3-D Strut-and-Tie Models

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before in Fig. 6.5, for Upper Node and for Lower
Node. In order to obtain the parameter , the ratio ( ⁄ ) is calculated in the manner as
for the short direction. The strut vertical height . For the Upper
Node, the bearing area is , and for the Lower Node, .
Thus,
 At the column nodal zone (Upper Node);
( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);

( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;

( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.83 ) and the node strength (1.7 ). Hence, the smallest value of (0.83 )
is governing the strut strength at the Upper Node.
Upon substituting in where is assumed to be the equivalent
upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

104
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.83 ) and the node strength (0.56 ). Hence, the smallest value of
(0.56 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

In order to reduce the difference between to the two values of strut strength, the height
of upper node should be estimated based on a strength lower than .

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the angle


. The width of strut at the upper node, and at the
lower node, . The pile force , the force in the
diagonal strut, .

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

105
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.82 ) and the node strength (1.5 ). Hence, the smallest value of (0.82 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );

[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.82 ) and the node strength (0.56 ). Hence, the smallest value of
(0.56 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following;

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

Table 6.4 summarizes the calculations for the critical struts and nodes in the
long direction.

106
Chapter 6 3-D Strut-and-Tie Models

Table 6.4 Summary of calculation results of long direction of pile cap B

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
1053.8 1053.8 yes
Struts
1176.1 715.6 No, 60.9%
Lower node,
522.3 551.6 yes
C-C-T
From the results in the table, the size of the upper nodal zone should be reduced to
60.9%. This finally gives a value of . It should be noted that the
reinforcing steel does not yield at this obtained pile load.

Finally, the capacity of pile cap B

( ) ( )

6.4 Case Study 3 - Pile Cap C

Input data:
Pile cap size; overall depth, h=600 mm, number of piles = 6 piles, square column of
size 300mm, circular piles of diameter 200mm, with 100 mm embedded into the
underside of the pile cap. Fig. 6.14 shows the geometry of pile cap C.

Fig. 6.14 Geometry of pile cap C.

107
Chapter 6 3-D Strut-and-Tie Models

Materials; the cylinder compressive strength of concrete, = 27.1MPa, the yield stress
of longitudinal steel, =479MPa, the area of main reinforcement, As (TB1) =12No.10
(1200mm2), with depth to steel centroid= 390mm and As (TC2) =11No.10 (1100mm2),
with depth to steel centroid= 390mm in the short direction and As (TC3) =21No.10
(2100mm2), with depth to steel centroid= 400mm in the long direction.

Strut-and-Tie Modeling
Fig. 6.15 illustrates a simple three-dimensional strut-and-tie model of pile cap C. The
concentrated column load is transmitted directly to the six supports (piles) by six
inclined compression struts while, the horizontal tension ties (longitudinal
reinforcement) are required to prevent the piles from being spread apart.

Fig. 6.15 Proposed 3D-STM of pile cap C.

The analysis will be performed in both the short direction and diagonal strut,
direction (i.e. long direction) as follows. Fig. 6.16 illustrates the proposal of STM in top
view.

Fig. 6.16 STM of pile cap C in top view.

108
Chapter 6 3-D Strut-and-Tie Models

Analysis in short direction, Sec. 1-1

1. The internal lever arm, hs1:


As shown in Fig. 6.17, the term u1 (height of lower node ) can be computed from;
( ) , where is the number of steel layers, is the
longitudinal steel diameter, is the clear concrete cover, and is the clear distance
between bars or,

( )
( )

The width of horizontal strut, can be computed from;


, where is the strength of the upper nodal
zone C-C-C node and is the equivalent width of the upper node ( ),
then

thus, ( )

( )

Fig. 6.17 STM of pile cap C in the short direction.

109
Chapter 6 3-D Strut-and-Tie Models

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.18 based on the dimension of column and piles as follows;

a) Upper Node b) Lower Node

Fig. 6.18 Nodes of pile cap C in the short direction.

 The width of the strut at the upper node, , Fig. 6.18a;

 For the lower node, the width of the strut , , Fig. 6.18b;

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

110
Chapter 6 3-D Strut-and-Tie Models

ii. Diagonal strut,


The parameter accounts for the amount of confinement (Fig 6.19),
(√( ⁄ ) ) (6.5)

a) Column nodal zone b) Pile nodal zone

Fig. 6.19 Amount of confinement for pile cap C.

The parameter ;
a) At the column nodal zone

√( ⁄ ) √ , this gives

b) At the pile nodal zone

√( ⁄ ) √ , this gives

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, Fig. 6.17, the dimension of the

bearing area of the nodal zone connected with the strut, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

111
Chapter 6 3-D Strut-and-Tie Models

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.89 ) and the node strength (1.7 ). Hence, the smallest value of
(0.89 ) is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.89 ) and the node strength (0.63 ). Hence, the smallest value of
(0.63 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

In order to reduce the gap between to the two values of strut strength, the height of
'
upper node should be estimated based on a strength lower than 1.7 f c .

112
Chapter 6 3-D Strut-and-Tie Models

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the


angle . The width of strut at the upper node, and at
the lower node, . The pile force , the force in the
diagonal strut, .

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.85 ) and the node strength (1.0 ). Hence, the smallest value of (0.85 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

113
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.85 ) and the node strength (0.60 ). Hence, the smallest value of
(0.60 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following.

Since a value of ( ) was used in estimating the


width of strut at the lower node, a new value of the strut width should be calculated;
,

This value of leading to a new nominal value of the strut strength at this node;

[ ]

This last value of the nominal strength, the strut is still greater that the value at the
upper node; and therefore, the solution should proceed with no change for this strut.

In order to determine the predicted failure load in this direction, Table 6.5
summarizes the calculations for the critical struts and nodes in the short direction.

114
Chapter 6 3-D Strut-and-Tie Models

Table 6.5 Summary of calculation results of short direction of pile cap C

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
574.8 574.8 yes
Struts
805.9 713.6 No, 88.5%
Lower node,
563.1 653.9 yes
C-C-T
From the results in the table, the size of the upper nodal zone should be reduced to
88.5%. This finally gives a value of .

Analysis in long direction, Sec. 2-2

As performed for sec. 1-1


1. The internal lever arm, hs2:
As shown in Fig. 6.20, the term u2 (height of lower node )
( )

Fig. 6.20 STM of pile cap C in the long direction.

From the STM in top view of pile cap C, Fig. 6.16,

Also, , where is assumed to be the equivalent width of the


upper node ( ) as shown in Fig. 6.20, then

115
Chapter 6 3-D Strut-and-Tie Models

( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.21 based on the dimension of column and piles as follows;

a) Upper Node b) Lower Node


Fig. 6.21 Nodes of pile cap C in the long direction.

 The width of the strut at the upper node, , Fig. 6.21a;

 For the lower node, the width of the strut , , Fig. 6.21b;

4. STM forces:

√ √

5. Checking stress limits;


a. Struts:

116
Chapter 6 3-D Strut-and-Tie Models

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before in Fig. 6.19, for Upper Node and for Lower
Node.

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, Fig. 6.20, the dimension of

the bearing area of the nodal zone connected with the strut, √ ⁄
. For the lower node, the bearing area is pile area; hence, is equal to the
pile diameter, . Thus,

At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

117
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.89 ) and the node strength (1.7 ). Hence, the smallest value of (0.89 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-T-T) Node, 6. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.89 ) and the node strength (0.40 ). Hence, the smallest value of
(0.40 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following;

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

Table 6.6 summarizes the calculations for the critical struts and nodes in the
long direction.

118
Chapter 6 3-D Strut-and-Tie Models

Table 6.6 Summary of calculation results of long direction of pile cap C


Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
1135.3 1135.3 yes
Struts
1193.9 546.5 No, 45.8%
Lower node,
359.9 577.6 yes
C-T-T
From the results in the table, the size of the upper nodal zone should be reduced to
45.8%. This finally gives a value of .

Finally, the capacity of pile cap C

6.5 Case Study 4 - Pile Cap D


Input data:
Pile cap size; overall depth, h=600 mm, number of piles = 4 piles, square column of
size 300mm, circular piles of diameter 200mm, with 100 mm embedded into the
underside of the pile cap. Fig. 6.22 shows the geometry of pile cap D.

Fig. 6.22 Geometry of pile cap D.

Materials; the cylinder compressive strength of concrete, = 30.3MPa, the yield stress
of longitudinal steel, =486MPa, the area of main reinforcement, As (TD1) =12No.15
(2400mm2), with depth to steel centroid= 380mm in the short direction and As (TD2)
=24No.15 (4800mm2), with depth to steel centroid= 400mm in the long direction. The
diameter of No. 15 is 16mm and the cross-sectional area is 200mm2.

119
Chapter 6 3-D Strut-and-Tie Models

Strut-and-Tie Modeling
Fig. 6.23 illustrates a simple three-dimensional strut-and-tie model for pile caps D and
E. The concentrated column load is transmitted directly to the four supports (piles) by
four inclined compression struts while, the horizontal tension ties (longitudinal
reinforcement) are required to prevent the piles from being spread apart. Nodes
represent the connections between struts and ties. The analysis will be performed in
both the short and long directions in the following sections.

Fig. 6.23 Proposed 3D-STM of pile caps D and E.

Analysis in short direction, Sec. 1-1

1. The internal lever arm, hs1:


As shown in Fig. 6.24, the term u1 (height of lower node ) can be computed from

( )
where is the number of steel layers, is the longitudinal steel diameter, is the
clear concrete cover, and is the clear distance between bars or,

( )
( )

The width of horizontal strut, can be computed from Eq. (6.6)


(6.6)
, where is the strength of the upper nodal zone C-C-C
node and is the equivalent width of the upper node ( ), then

120
Chapter 6 3-D Strut-and-Tie Models

thus, ( )
( )

Fig. 6.24 STM of pile cap D in the short direction.

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as shown
in Fig. 6.25 based on the dimension of column and piles as follows;

a) Upper Node b) Lower Node

Fig. 6.25 Nodes of pile cap D in the short direction.

121
Chapter 6 3-D Strut-and-Tie Models

 The width of the strut at the upper node, , Fig. 6.25a;

 For the lower node, the width of the strut , , Fig. 6.25b;

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before, for Upper Node and for Lower Node.

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, the bearing area is column

cross sectional area divided by the number of piles; thus, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

122
Chapter 6 3-D Strut-and-Tie Models

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.81 ) and the node strength (1.7 ). Hence, the smallest value of (0.81 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.81 ) and the node strength (0.55 ). Hence, the smallest value of
(0.55 ) is governing the strut strength at the Lower Node.
Upon substituting in where is assumed to be the pile diameter,
( ) for lower node;
[ ]

It is noted that there is a significant difference between the strength of strut at the
upper and lower nodes, and this can be attributed to the significant difference between
the strength of the upper nodal zone, 1.7 and the strut strength, 0.81 . In order to
reduce the gap between to the two values of strut strength, the height of upper node
'
should be estimated based on a strength lower than 1.7 f c .

123
Chapter 6 3-D Strut-and-Tie Models

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the


angle . The width of strut at the upper node, and at
the lower node, . The pile force , the force in the
diagonal strut, .

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.78 ) and the node strength (1.3 ). Hence, the smallest value of (0.78 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

124
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.78 ) and the node strength (0.53 ). Hence, the smallest value of
(0.53 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following.

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

In order to determine the predicted failure load in this direction, Table 6.7
summarizes the calculations for the critical struts and nodes in the short direction.

Table 6.7 Summary of calculation results of short direction of pile cap D

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN

1166.4 1166.4 yes


Struts
1569.5 872.9 No, 55.6%
Lower node,
1046.9 643.9 No, 61.5%
C-C-T
This leads to a value of =

125
Chapter 6 3-D Strut-and-Tie Models

Analysis in long direction, Sec. 2-2

1. The internal lever arm, hs2:


As shown in Fig. 6.26, the term u2 (height of lower node )
( )

Fig. 6.26 STM of pile cap D in the long direction.

( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.27 based on the dimension of column and piles as follows;

a) Upper Node b) Lower Node


Fig. 6.27 Nodes of pile cap D in the long direction.

126
Chapter 6 3-D Strut-and-Tie Models

 The width of the strut at the upper node, , Fig. 6.27a;

 For the lower node, the width of the strut , , Fig. 6.27b;

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Diagonal strut,
As explained before in Fig. 6.5, for Upper Node and for Lower
Node. In order to obtain the parameter , the ratio ( ⁄ ) is calculated in the manner as
for the short direction. The strut vertical height . For the Upper
Node, the bearing area is , and for the Lower Node, .
Thus,

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

127
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.75 ) and the node strength (1.7 ). Hence, the smallest value of (0.75 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.75 ) and the node strength (0.52 ). Hence, the smallest value of
(0.52 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following;

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.
Table 6.8 summarizes the calculations for the critical struts and nodes in the
long direction.

128
Chapter 6 3-D Strut-and-Tie Models

Table 6.8 Summary of calculation results of long direction of pile cap D


Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
2332.8 2332.8 yes
Struts
2548.9 824.4 No, 32.3%
Lower node,
1027.1 624.4 No, 60.8%
C-C-T
This leads to a value of .

Finally, the capacity of pile cap D

( ) ( )

6.6 Case Study 5 - Pile Cap E


Input data:
Pile cap size; overall depth, h=600 mm, number of piles =6 piles, square column of size
300mm, circular piles of diameter 200mm, with 100 mm embedded into the underside
of the pile cap. Fig. 6.28 shows the geometry of pile cap E.

Fig. 6.28 Geometry of pile cap E.

Materials; the cylinder compressive strength of concrete, = 41.1MPa, the yield stress
of longitudinal steel, =486MPa, the area of main reinforcement, As (TD1) =12No.15
(2400mm2), with depth to steel centroid= 380mm in the short direction and As (TD2)
=24No.15 (4800mm2), with depth to steel centroid= 400mm in the long direction.

129
Chapter 6 3-D Strut-and-Tie Models

In addition, As (TE3) =9No.10 (900mm2), with depth to steel centroid= 495mm


in the short direction, also for As (TE4) =5No.10 (500mm2), with depth to steel
centroid= 485mm in long direction.

Analysis in short direction, Sec. 1-1


In short direction the steel reinforcement As (TD1) and part of As (TE3)
(5No.10) were taken into consideration when determining the predicted failure load, V1
of STM, with equivalent depth to centroid=400 mm.

As performed before in pile cap B, but when the concrete compressive


strength, is greater than 34.5 MPa, the maximum bearing stress should be limited to;

( ) √( )( ) (6.7)

1. The internal lever arm, hs1:


As shown in Fig. 6.29, the term u1 (height of lower node ) can be computed from
( )

where is the number of steel layers, is the longitudinal steel diameter, is the
clear concrete cover, and is the clear distance between bars or,

( )

Fig. 6.29 STM of pile cap E in the short direction.

130
Chapter 6 3-D Strut-and-Tie Models

( )

The width of horizontal strut, can be computed

, where is the strength of the upper nodal zone C-C-C node


and is the equivalent width of the upper node ( ), then

( )

It should be noted that only the distributed reinforcement with twice the column width
are considered.

thus, ( )
( )

2. Angle of inclined strut,

, from pile cap geometry.

3. Widths of struts:
The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.30 based on the dimension of column and piles as follows;

a) Upper Node b) Lower Node


Fig. 6.30 Nodes of pile cap E in the short direction.

 The width of the strut at the upper node, , Fig. 6.30a;

131
Chapter 6 3-D Strut-and-Tie Models

 For the lower node, the width of the strut , , Fig. 6.30b;

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Horizontal strut,
The width of strut was determined based on its designed strength; hence, there is
no need to carry out any further checks for this strut.

ii. Diagonal strut,


As explained before, for Upper Node and for Lower Node.

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, the bearing area is column

cross sectional area divided by the number of piles; thus, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);

( ⁄ ) ⁄ this gives [( ) ]

132
Chapter 6 3-D Strut-and-Tie Models

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( ) √( )( )

( ) √( ) ( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( ) √( ) ( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.80 ) and the node strength (1.7 ). Hence, the smallest value of (0.80 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.80 ) and the node strength (0.55 ). Hence, the smallest value of
(0.55 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

In order to reduce the gap between to the two values of strut strength, the height of
'
upper node should be estimated based on a strength lower than 1.7 f c .

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

133
Chapter 6 3-D Strut-and-Tie Models

The width of the horizontal strut , , and the


angle . The width of strut at the upper node, and at
the lower node, . The pile force , the force in the
diagonal strut, .
 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( ) √( )( )

( ) √( ) ( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( ) √( ) ( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.78 ) and the node strength (1.3 ). Hence, the smallest value of (0.78 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );

[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.78 ) and the node strength (0.54 ). Hence, the smallest value of
(0.54 ) is governing the strut strength at the Lower Node.

134
Chapter 6 3-D Strut-and-Tie Models

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following.

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

Table 6.9 summarizes the calculations for the critical struts and nodes in the short
direction.

Table 6.9 Summary of calculation results of short direction of pile cap E

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
1405.9 1405.9 yes
Struts
1964.4 1101.4 No, 56.1%
Lower node,
1371.8 882.5 No, 64.3%
C-C-T
This leads to a value of =

Analysis in long direction, Sec.2-2


In long direction the steel reinforcement As (TD2) and part of As (TE4) (3No.10)
were taken into consideration when determined the predicted failure load V2 of STM,
with equivalent depth to centroid=405 mm.

135
Chapter 6 3-D Strut-and-Tie Models

As performed for sec. 1-1


1. The internal lever arm, hs2:
As shown in Fig. 6.31, the term u2 (height of lower node )
( )

Fig. 6.31 STM of pile cap E in the long direction.

( )

It should be noted that only the distributed reinforcement with twice the column width
are considered.

( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:

The upper width of strut, , and the lower width can be calculated as
shown in Fig. 6.32 based on the dimension of column and piles as follows;

 The width of the strut at the upper node, , Fig. 6.32a;

 For the lower node, the width of the strut , , Fig. 6.32b;

136
Chapter 6 3-D Strut-and-Tie Models

a) Upper Node b) Lower Node


Fig. 6.32 Nodes of pile cap E in the long direction.
4. STM forces:

5. Checking stress limits;


a. Struts:

i. Diagonal strut,
As explained before in Fig. 6.5, for Upper Node and for Lower
Node. In order to obtain the parameter , the ratio ( ⁄ ) is calculated in the manner as
for the short direction. The strut vertical height . For the Upper
Node, the bearing area is , and for the Lower Node, .
Thus,

 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( ) √( )( )

( ) √( ) ( )

137
Chapter 6 3-D Strut-and-Tie Models

While, the maximum bearing stress of strut at the Lower Node should be limited to
( ) √( ) ( )

For the design strength of strut , take the average strength at the upper and lower
nodes;

( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.75 ) and the node strength (1.7 ). Hence, the smallest value of (0.75 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.75 ) and the node strength (0.52 ). Hence, the smallest value of
(0.52 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
The There is no need to carry out any checks for the upper node since the strength of
the node is greater or equal to the strength of the surrounding struts. The critical node is
the lower node, and as estimated before, . Based on this, the nominal
value of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following;

138
Chapter 6 3-D Strut-and-Tie Models

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

In order to determine the predicted failure load in this direction, Table 6.10
summarizes the calculations for the critical struts and nodes in the long direction.

Table 6.10 Summary of calculation results of long direction of pile cap E


Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN

2476.5 2476.5 yes


Struts
2746.7 1110.5 No, 40.4%
Lower node,
1188.0 860.1 No, 72.4%
C-C-T
This leads to a value of

Finally, the capacity of pile cap E

( ) ( )

6.7 Case Study 6 - Pile Cap F

Pile Caps D and F were identical except the four ‘‘missing corners’’ of plain concrete as
shown in Fig. 6.33. As observed from the experimental results and the FEM, The ACI
building Code predicts that pile cap D will have much higher strength than pile cap F,
because there is no confinement in pile cap F.

Input data:
Pile cap size; overall depth, h=600 mm, number of piles = 4 piles, square column of
size 300mm, circular piles of diameter 200 mm, with 100 mm embedded into the
underside of the pile cap.

139
Chapter 6 3-D Strut-and-Tie Models

Fig. 6.33 Geometry of pile cap F.

Materials; the cylinder compressive strength of concrete, = 30.3MPa, the yield stress
of longitudinal steel, =486MPa, the area of main reinforcement, As (TD1) =12No.15
(2400mm2), with depth to steel centroid= 380mm in the short direction and As (TD2)
=24No.15 (4800mm2), with depth to steel centroid= 400mm in the long direction. The
diameter of No. 15 is 16mm and the cross-sectional area is 200mm2.

Strut-and-Tie Modeling
Fig. 6.34 illustrates a simple three-dimensional strut-and-tie model for pile cap F. The
concentrated column load is transmitted directly to the four supports (piles) by four
inclined compression struts while, the horizontal tension ties (longitudinal
reinforcement) are required to prevent the piles from being spread apart. Nodes
represent the connections between struts and ties. The analysis will be performed in
both the short and long directions in the following sections.

Fig. 6.34 Proposed 3D-STM of pile cap F.

140
Chapter 6 3-D Strut-and-Tie Models

Analysis in short direction, Sec.1-1


As performed for Pile Cap D

1. The internal lever arm, hs1:


As shown in Fig. 6.24, the term u1 (height of lower node ) can be computed from;

( )
( )
The width of horizontal strut, can be computed from equation.

thus, ( )
( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
 The width of the strut at the upper node, , Fig. 6.25a;

 For the lower node, the width of the strut , , Fig. 6.25b;

4. STM forces:

5. Checking stress limits;


a. Struts:

i. Diagonal strut,
The compressive strength of concrete in this strut may be significantly enhanced with
confinement (Sec. 4.3.3.2). Therefore, the maximum bearing stress should be limited to;

( )( )

141
Chapter 6 3-D Strut-and-Tie Models

The parameter accounts for the amount of confinement (Fig 6.35),


(√( ⁄ ) )

 At the column nodal zone (Upper Node);

√( ⁄ ) √ , this gives ( )

 At the pile nodal zone (Lower Node);

√( ⁄ ) √ , this gives ( )

Fig. 6.35 Amount of confinement at column and pile nodal zones.

While the parameter accounts for the geometry of the compression stress field,
[( ⁄ ) ] , where ( ⁄ ) reflects the geometry of the compression
strut. (i.e., is assumed to be the strut length and is assumed to be the strut width at
the upper and lower nodes). From trigonometry and for easiness in the calculations, the
ratio ( ⁄ ) can expressed as ( ⁄ ), where is the strut vertical height and is
the dimension of the bearing area of the nodal zone connected with the strut, with
equal to the square root of this bearing area.

For strut , . For the upper node, the bearing area is column

cross sectional area divided by the number of piles; thus, √ ⁄ .


For the lower node, the bearing area is pile area; hence, is equal to the pile
diameter, . Thus,

142
Chapter 6 3-D Strut-and-Tie Models

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to

( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to

( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.57 ) and the node strength (1.7 ). Hence, the smallest value of (0.57 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.57 ) and the node strength (0.47 ). Hence, the smallest value of
(0.47 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

In order to reduce the gap between to the two values of strut strength, the height of
'
upper node should be estimated based on a strength lower than 1.7 f c .

143
Chapter 6 3-D Strut-and-Tie Models

Upon using a strength of the upper node ( ⁄ ) , and


redoing the calculations, the following has been obtained.

The width of the horizontal strut , , and the


angle . The width of strut at the upper node, and at
the lower node, . The pile force , the force in the
diagonal strut, .

 At the column nodal zone (Upper Node);

( ⁄ ) ⁄ this gives [( ) ]

 At the pile nodal zone (Lower Node);

( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.56 ) and the node strength (1.0 ). Hence, the smallest value of (0.56 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

144
Chapter 6 3-D Strut-and-Tie Models

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.56 ) and the node strength (0.45 ). Hence, the smallest value of
(0.45 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;
[ ]

Take the smallest value

b. Nodes:
There is no need to carry out any checks for the upper node since the strength of the
node is greater or equal to the strength of the surrounding struts. The critical node is the
lower node, and as estimated before, . Based on this, the nominal value
of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following.

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

Table 6.11 summarizes the calculations for the critical struts and nodes in the short
direction.

Table 6.11 Summary of calculation results of short direction of pile cap F

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
1166.4 1166.4 yes
Struts
1524.9 741.6 No, 48.6%
Lower node,
981.3 546.9 No, 55.7%
C-C-T
This leads to a value of .

145
Chapter 6 3-D Strut-and-Tie Models

Analysis in long direction, Sec.2-2

As performed for sec. 1-1

1. The internal lever arm, hs2:


As shown in Fig. 6.26, the term u2 (height of lower node )
( )

( )

2. Angle of inclined strut,


, from pile cap geometry.

3. Widths of struts:
 The width of the strut at the upper node, , Fig. 6.27a;

 For the lower node, the width of the strut , , Fig. 6.27b;

4. STM forces:

5. Checking stress limits;


a. Struts:
i. Diagonal strut,
With reference to Fig. 6.35, at Upper Node and at Lower Node.

In order to obtain the parameter , the ratio ( ⁄ ) is calculated in the manner as for the
short direction. The strut vertical height . For the Upper Node,
the bearing area is , and for the Lower Node, . Thus,

 At the column nodal zone (Upper Node);


( ⁄ ) ⁄ this gives [( ) ]

146
Chapter 6 3-D Strut-and-Tie Models

 At the pile nodal zone (Lower Node);


( ⁄ ) ⁄ this gives [( ) ]

Thus, the maximum bearing stress of strut at the Upper Node should be limited to
( )( )

While, the maximum bearing stress of strut at the Lower Node should be limited to
( )( )

For the design strength of strut , take the average strength at the upper and lower
nodes;
( )

The strength of strut at the upper node is controlled by the smaller of the strut
strength (0.56 ) and the node strength (1.7 ). Hence, the smallest value of (0.56 )
is governing the strut strength at the Upper Node.

Upon substituting in where is assumed to be the equivalent


upper node width ( );
[ ]

While the lower node is (C-C-T) Node, 8. The limiting compressive strength,

The strength of strut at the lower node is controlled by the smaller of the strut
strength (0.56 ) and the node strength (0.45 ). Hence, the smallest value of
(0.45 ) is governing the strut strength at the Lower Node.

Upon substituting in where is assumed to be the pile diameter,


( ) for lower node;

[ ]

Take the smallest value

147
Chapter 6 3-D Strut-and-Tie Models

b. Nodes:
The There is no need to carry out any checks for the upper node since the strength of
the node is greater or equal to the strength of the surrounding struts. The critical node is
the lower node, and as estimated before, . Based on this, the nominal
value of the pile reaction is

Since the reinforcement of the tie, , is arranged in multiple layers, the height of the
node should be checked against a maximum value as illustrated in the following;

Since a value of ( ) was used in estimating the


width of strut at the lower node, the solution should proceed.

Table 6.12 summarizes the critical struts and node calculations in the long
direction.

Table 6.12 Summary of calculation results of long direction of pile cap F

Actual Max.
Model Label Member Satisfaction
force, kN capacity, kN
2332.8 2332.8 yes
Struts
2548.9 723.7 No, 28.4%
Lower node,
1027.1 548.8 No, 53.4%
C-C-T
This leads to a value of

Finally, the capacity of pile cap F

( ) ( )

148
Chapter 6 3-D Strut-and-Tie Models

6.8 STM vs. Experimental Data


In order to examine the results from STM above, Table 6.13 summarizes the
comparison between the predicted capacities of each pile cap from STM analysis and
test results. From the table, it is obvious that the STM predictions represent good lower
bound estimates.

Table 6.13 The STM results compared with test results.

Exp. Data STM Data


Pile cap
, kN Failure Mode , kN Failure Mode

A 1781 flexural 1365.0 flexural 0.77


B 2189 flexural 1547.8 flexural 0.71
C 2892 flexural 1656.4 flexural 0.57
D 3222 shear 1827.8 shear 0.57
E 4709 shear 2499.2 shear 0.53
F 3026 shear 1537.2 shear 0.51
Average 0.61

6.9 Conclusions
In summary, this chapter introduces an approach for how to develop design models for
reinforced concrete pile caps using the strut-and-tie model method. From the obtained
results, the following can be concluded:

 STM approach gives mean values ranges from 0.51 to 0.77 of the experimental
ultimate loads for all chosen pile caps from literature.
 The proposed Strut-and-Tie approach is a powerful tool to predict the ultimate
strength, mode of failure and behavior of reinforced concrete transfer slabs such
as pile cap.
 Confinement of concrete in pile caps is essentially responsible for increasing
their capacity.
 The strength of nodes and struts in pile caps should be derived with
consideration to the spatial nature of the structural system in order to derive
good estimate of pile cap capacity.

149
CHAPTER 7

SUMMARY AND CONCLUSIONS

7.1 Summary
Transfer slab is considered one of the most common transfer elements which could
exist in high rise buildings. In this thesis, pile caps are considered to be as a simple
transfer slab. The Strut and Tie Model (STM) is an invaluable tool used for the design
of this disturbed region. In pile caps there is a very important volume of concrete that is
subjected to low stresses. Confinement by this inactive concrete increases the capacity
of these pile caps; therefore, 3-D STMs are required. In this thesis, a 3-D nonlinear
finite element analysis has been conducted in order to predict the ultimate capacity of
selected reinforced concrete pile caps using discrete reinforcement modeling technique.
Then, the Strut-and-Tie Models, STM, for all selected pile caps are suggested in two
and three dimensions based on the available experimental results of crack patterns,
modes of failure, and stress trajectors obtained from elastic finite element analysis. The
obtained 3-D-STM results are then compared with experimental test results.

7.2 Conclusions
Based on the work presented in this study, the following conclusions can be drawn:

 The 3-D nonlinear finite element analysis of reinforced concrete pile caps yields
accurate predictions of both the ultimate load and the complete response.

 The strut-and-tie model gives reasonable lower bound estimates of the load
carrying capacity of reinforced concrete pile caps.

 Unlike 2D strut-and-tie model, 3-D STMs are required when the structure and
loading are considerably spread over all three dimensions, such as pile caps.

 The strength of nodes and struts in pile caps should be derived with
consideration to the spatial nature of the structure system.

151
REFERENCES
1 Adebar, P., and Zhou, Z., 1993, “Bearing Strength of Compressive Struts
Confined by Plain Concrete,” ACI Structural Journal, V. 90, No. 5, September-
October 1993, pp. 534-541.

2 Adebar, P., and Zhou, Z., 1996, “Design of Deep Pile Caps By Strut-and-Tie
Models,” ACI Structural Journal, V. 93, No. 4, July-August 1996, pp.1-12.

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

4 Ali, M. A., 1997, “Automatic generation of truss models for the optimal design of
reinforced concrete structures,” PhD. thesis, Cornell Univ., Ithaca, N.Y.

5 American Concrete Institute, ACI 318, “Building Code Requirements for


Structural Concrete,” ACI 318-14, ACI Committee 318, 519 pp., Detroit, USA,
2014.

6 ANSYS User's Manual, Swanson Analysis Systems, Inc. 10, William, K. J. and
Warnke, E. D., 1975, “Constitutive model for the triaxial behavior of concrete,”
Proceedings of the International Association for Bridge and Structural Engineering.
7 ANSYS, 2009, ANSYS User’s Manual Revision 12.0, ANSYS, Inc., United States.
8 Brown, M. D., Sankovich, C. L., Bayrak, O., Jirsa, J. O., Breen, J. E., and
Wood, S. L., 2006, “Examination of the AASHTO LRFD Strut and Tie
Specifications,” Center for Transportation Research Report 0-4371-2, University of
Texas, Austin, April 2006, 330 pp.
9 Chantelot, G. and Mathern, A., 2010, “Strut-and-tie modeling of reinforced
concrete pile caps,” Master’s Thesis, Department of Structural Engineering,
Chalmers University of Technology, Publication no. 2010:51, Göteborg, Sweden,
51-53 pp.
10 Chen, W. F., and El-Metwally, S. E., 2011, “Understanding structural
engineering; from theory to practice,” Florida, CRC Press.
11 Chen, W. F., and Trumbauer, B. E., 1972, “Double Punch Test and Tensile
Strength of Concrete,” Journal of Materials, JMLSA, V. 7, No. 2, June 1972, pp.
148-154.

153
References

12 Clyde, D., 2008, “Affinity Properties of Truss Model Nodes,” Magazine of


Concrete Research, (Morley Symposium on Concrete Plasticity and its
Application), V. 60, No. 9, November 2008, pp. 651-655.
13 Dar, O. J., 2007, ‘‘Analysis and design of shear wall-transfer beam structure,”
boring pengeshan status thesis, University of Technology, Malaysia.
14 Desayi, P. and Krishnan, S., ‘‘Equation for the Stress-Strain Curve of Concrete,”
Journal of the American Concrete Institute, 61(3):345-350, March 1964.
15 Egyptian Code for the Design and Construction of Concrete Structure, ECP
203-2006, Ministry of Housing, Utilities and Urban Communities, National
Housing and Building Research Center, Cairo,2006.
16 EN 1992-1-1:2004, Euro code 2: Design of Concrete Structures, Part 1-1: General
rules and rules for buildings, December 2004.
17 Engström, B., 2009, “Design and Analysis of Deep Beams, Plates and other
Discontinuity Regions,” Chalmers University of Technology, Gothenburg, Sweden,
2009, 50 pp.
18 Engström, B., 2011, “Design and analysis of deep beams, plates and other
discontinuity regions,” Department of Structural Engineering, Chalmers University
of Technology, Göteborg, Sweden, 2011.
19 Fédération Internationale Du BétonFIB, 2002, “Design examples for the 1996
FIP recommendations - Practical design of structural concrete,” Technical Rep.,
International Federation for Structural Concrete fib, London.
20 Fédération Internationale Du BétonFIB; Fip, 1999, “Practical Design of
Structural Concrete,” fédération internationale de la précontrainte, September 1999,
113 pp.
21 Ghali, Mona K., 1999, “Effect of pile cap flexural rigidity on the behaviour of
bridge foundations,” Article, IABSE report, 80 (1999), p2410-246.
22 Hawkins, Neil M., “Bearing Strength of Concrete Loaded through Rigid Plates.”
Magazine of Concrete Research (London), V. 20, N. 62, Mar. 1968, pp. 31-40.
23 Kong, F. K., 1990, Reinforced Concrete Deep Beams, Van Nostrand Reinhold in
New York.
24 Leu, L. J., Huang, C. W., Chen, C. S., and Liao, Y. P., 2006 “Strut-and-Tie
Design Methodology for Three-Dimensional Reinforced Concrete Structures”
Journal of Structural Engineering 132(6):929-938, 2006.

154
References

25 Ley, T. M., Riding, K. A., Widianto, Bae, S. J. and Breen, J. E., 2007
“Experimental verification of strut-and-tie model design method”, ACI Struct. J.,
104(6), 749-755.

26 MacGregor, J. G., 1997, “Reinforced concrete mechanics and design,” 3rd Ed.,
Prentice-Hall, Upper Saddle River, N.J.
27 Martin, B. and Sanders, D., 2007, “Verification and Implementation of Strut-
and-Tie Model in LRFD Bridge Design Specifications,” 1-14 pp.

28 Meisam, S. G., 2009, “Analysis of FRP Strengthened Reinforced Concrete Beams


Using Energy Variation Method,” World Applied Sciences Journal 6 (1): 105- 111.
29 Muttoni, A., Schwartz, J., and Thürlimann, B., 1997, “Design of Concrete
Structures with Stress Fields,” Birkhäuser Verlag, Basel, Boston and Berlin,
Switzerland, 1997, 145 pp.
30 Park, R. and Paulay, T., 1975, ”Reinforced Concrete Structures,” John Wiley &
Sons, New York.
31 Raymond, W. M. Wong, 1999, ”Construction of Residential Buildings,
Developments and Trends in Methods and Technology, Hong Kong Housing
Development, Book 2, China Trend Building Press, 1999.

32 Ritter, W., 1899, Die Bauweise Hennebique, Schweizerische Bauzeitung, Bd,


XXXIII, No. 7, January 1899.
33 Schäfer, K.,1999, “Nodes,” Section 4.4.4 in Structural Concrete,Vol.2, fédération
internationale du béton (fib), Bulletin 2, Lausanne, Suisse, pp. 257-275.
34 Schlaich, J. and Schäfer, K, 1991, “Design and detailing of structural concrete
using strut-and-tie models,” Journal of the Structural Engineer, V. 69, No. 6, pp.
113–125.

35 Schlaich, J., Schäfer, K., and Jennewein, M., 1987, “Toward a Consistent Design
of Structural Concrete,” Journal of Prestressed Concrete Institute, V. 32, No. 3,
May 1987, 74-150.
36 Schlaich, M., and Anagnostou, G., 1990, “Stress Fields for Nodes of Strut-and-
Tie Models,” ASCE Journal of Structural Engineering, V. 116, No. 1, January
1990, pp. 13-23.
37 Tan, K. H. and Cheng, G. H., 2006, “Size effect on shear strength of deep beams:
Investigation with strut-and-tie model,” J. Str. Eng., 132(5), 673-685.

155
References

38 Tan, K. H., Cheng, G. H. and Zhang, N., 2008, “Experiment to mitigate size
effect on deep beams,” Mag. Concrete Res., 60(10), 709-723.

39 Tjhin, T. N., and Kuchma, D. A., 2002, “Computer-based tools for design by
strut-and-tie method: Advances and challenges,” ACI Structure. J., 995, 586–594.

40 Wang, C. Z., Guo, Z. H., and Zhang, X. Q., 1987, “Experimental Investigation of
Biaxial and Triaxial Compressive Concrete Strength,” ACI materials Journal,
March-April 1987, pp. 92-100.

41 Williams, C., Deschenes, D., and Bayrak, O., 2012, “Strut-and-Tie Model Design
Examples for Bridges,” Final Report 5-5253-01-1, Center for Transportation
Research, University of Texas, Austin, June 2012, 276 pp.

42 Yang, K. H., Chung, S. H. and Ashour, A. F., 2007, “Influence of section depth
on the structural behaviour of reinforced concrete continuous deep beams,” Mag.
Concrete Res., 59(8), 575-586.

43 Yun, Y. M., 2000, “Computer graphics for nonlinear strut-tie model approach,” J.
Computer, Civil Eng., 142, 127–133.

44 Zhang, L., Li, Y. and Wu, Q., 2003, “Seismic Response Analysis of Frame-
Supported Shear Wall Structure with High Transfer Story,” Industrial Construction
33(6), p24-27.

45 Zhang, N. and Tan, K. H., 2007, “Size effect in RC deep beams: Experimental
investigation and STEM verification,” Eng. Str., 29(12), 3241-3254.

46 Zhang, X., Zhou, Y. and He, J., 2005, “Seismic Design of the Short-Piered Shear
Wall Structure with High Transfer Floor,” Earthquake Resistant Engineering and
Retrofitting 27(4), p20-24.

156
‫يحخىيبث انشعبنت‬
‫ذرى‪ ْٛ‬اٌهساٌح ِٓ األت‪ٛ‬اب اٌراٌٍح‬
‫انببة األول‪:‬‬

‫ٌؽر‪ٛ‬ي ٘ما اٌثاب ػًٍ ِمكِح ػاِح ػٓ اٌّشىٍح ‪ٚ‬طهق اٌرصٍُّ اٌّفرٍفٗ ٌؽٍ‪ٙ‬ا ‪ٚ‬ذ‪ٛ‬ضٍػ أل٘كاف‬
‫اٌثؽس ‪ٚ‬وٍفٍح أعاو ٘مٖ األ٘كاف ِٓ ـالي ِؽر‪ٌٛ‬اخ اٌهساٌح‪.‬‬

‫انببة انزبَي‪:‬‬

‫ٌرٕا‪ٚ‬ي طهٌمح اٌرصٍُّ "أًَىرس انضبغظ وانشذاد"‪ ،‬وٍف ٌٕشؤ ‪ٚ‬اٌؼٕاصه اٌّى‪ٔٛ‬ح ٌألّٔ‪ٛ‬لض‬
‫‪ِٚ‬ما‪ِٚ‬ح وً ػٕصه ِٓ ٘مٖ اٌؼٕاصه‪.‬‬

‫انببة انزبنذ‪:‬‬

‫ٌرٕا‪ٚ‬ي ٘ما اٌثاب ِمكِح ػٓ ػٕاصه اٌرؽ‪ ًٌٛ‬اٌّفرٍفح اٌرى لك ذر‪ٛ‬اظك فً إٌّشآخ ‪ٚ‬األسثاب اٌرى‬
‫أقخ اًٌ ذ‪ٛ‬اظك٘ا ‪ٔٚ‬ثمج ِفرصهٖ ػٓ وً ٔ‪ٛ‬ع ‪ٚ‬ـص‪ٛ‬صا تالطح اٌرؽ‪ ًٌٛ‬اٌرً ً٘ ِؽً اٌكناسح‪،‬‬
‫‪ٌٚ‬ر‪ٛ‬ضٍػ اٌهإٌح أوصه‪ -‬ذُ اػرثان ٘اِاخ اٌف‪ٛ‬اوٌك وثالطح ذؽ‪ ًٌٛ‬تسٍطح فً ٘ما اٌثؽس‪.‬‬

‫انببة انشابع‪:‬‬

‫ٌرٕا‪ٚ‬ي ٘ما اٌثاب اٌرؽكٌاخ اٌرى ذ‪ٛ‬اظٗ طهٌمح اٌرصٍُّ اٌرمٍٍكٌح تاسرفكاَ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق‬
‫‪ٚ‬اٌكناساخ اٌساتمح ٌ‪ٙ‬مٖ اٌطهٌمح ٌرالئُ اٌؼٕاصه شالشٍح األتؼاق ٘اِاخ اٌف‪ٛ‬اوٌك (وثالطح ذؽ‪ًٌٛ‬‬
‫تسٍطح)‪ ،‬ؼٍس ذفرٍف ٘مٖ اٌؼٕاصه ػٓ غٍه٘ا ِٓ شٕائٍح األتؼاق فى ‪ٚ‬ظ‪ٛ‬ق اؼاطح ِٓ اٌفهسأٗ‬
‫تؼض‪ٙ‬ا ٌثؼض ذؼًّ ػًٍ وٌاقج ِما‪ِٚ‬ر‪ٙ‬ا‪.‬‬

‫انببة انخبيظ‪:‬‬

‫فً ٘ما اٌثاب ذُ ػًّ ّٔمظح ٌؼٍٕاخ ٘اِاخ ـ‪ٛ‬اوٌك ِفرثهج ِؼٍٍّا تاسرفكاَ تهٔاِط اٌرؽًٍٍ غٍه‬
‫اٌفطً (⑫‪ ،(ANSYS V‬شُ اٌرؤوك ِٓ صؽح ٘مٖ إٌّمظح تّمانٔح إٌرائط إٌظهٌح ِٓ اٌثهٔاِط‬
‫تإٌرائط اٌّؼٍٍّح ٌٍؼٍٕاخ اٌّفرثهج‪.‬‬

‫انببة انغبدط‪:‬‬

‫ٌرٕا‪ٚ‬ي ٘ما اٌثاب ِمانٔٗ تٍٓ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق شالشى األتؼاق ‪ٚ‬إٌّمظح ِٓ اٌرؽًٍٍ غٍه‬
‫اٌفطً تاسرفكاَ تهٔاِط (⑫‪.(ANSYS V‬‬

‫انببة انغببع‪:‬‬

‫فً ٘ما اٌثاب ذُ ػًّ ٍِفص ٌٍثؽس ‪ٚ‬إٌرائط اٌرى ذُ اٌر‪ٛ‬صً اٌٍ‪ٙ‬ا‪.‬‬

‫"فاللوه إجعل هذا العنل خالصا لوجوك الكزيه‪ ،‬وآخز دعواىا أن احلند هلل رب العاملني"‬
‫يهخض انشعبنت‬

‫بسه اهلل واحلند هلل والصالة والسالو علي رسول اهلل ‪ -‬سيدىا حمند وعلي آلى وصحبى وسله ‪-‬‬
‫ثه أما بعد‪...‬‬

‫فبْ اٌؽاظٗ اًٌ ػًّ يُشآث ححىيم أصثؽد أِه ٘اَ ‪ٚ‬ضه‪ٚ‬ني ‪ٚ‬ـاصح فً اٌّثأى ػاٌٍح‬
‫اإلنذفاع‪ ،‬ؼٍس ِٓ ـالٌ‪ٙ‬ا ٌس‪ ًٙ‬اٌرغٍة ػًٍ اٌّشاوً اٌّرؼٍمٗ تاـرالف اسرفكاَ اٌّساؼاخ قاـً‬
‫اٌّثٕى‪ٚ .‬لك ذى‪ِٕ ْٛ‬شآخ اٌرؽ‪ ًٌٛ‬ػًٍ شىً وّهج ذؽ‪ ًٌٛ‬أ‪ ٚ‬تالطح ذؽ‪ٚ .ًٌٛ‬ذؼرثه بالطت انخحىيم‬
‫‪ٚ‬اؼكج ِٓ أش‪ٙ‬ه اٌؼٕاصه اٌّ‪ٛ‬ظ‪ٛ‬قج تىصهج فً اٌّثأى اٌّهذفؼح‪ٌٚ .‬هسُ ص‪ٛ‬نٖ واٍِح ‪ٚ‬اضؽٗ ػٓ‬
‫٘مٖ اٌؼٕاصه شالشٍح األتؼاق‪ ،‬ذُ اػرثان هبيبث انخىاصيق وؤٔ‪ٙ‬ا تالطح ذؽ‪ ًٌٛ‬تسٍطح‪ٌ ،‬رس‪ًٍٙ‬‬
‫قناسر‪ٙ‬ا فً ٘ما اٌثؽس‪.‬‬

‫ذؼرثه هبيبث انخىاصيق ِٕطمٗ واٍِح اإلضطهاب‪ ،‬تسثة شىٍ‪ٙ‬ا غٍه إٌّرظُ ‪ٚ‬اؼر‪ٛ‬ائ‪ٙ‬ا‬
‫ػًٍ وٍّاخ وثٍهج ِٓ ؼعُ اٌفهسأح ت‪ٙ‬ا لاخ اظ‪ٙ‬اقاخ ػّ‪ٛ‬قٌح ِٕففضح تاإلضافح اًٌ اظ‪ٙ‬اقاخ‬
‫لص ػاٌٍح ٔسثٍا‪ِّ ،‬ا ٌعؼً سٍ‪ٛ‬و‪ٙ‬ا غٍه ِف‪ٌٍّ َٛٙ‬صُّ‪ٚ .‬تاٌراًٌ ذُ اسرفكاَ طهٌمح ذصٍُّ ظكٌكج‬
‫ذسًّ "أًَىرس انضبغظ وانشذاد" ٌٍرّىٓ اٌّصُّ ِٓ ف‪ ُٙ‬سٍ‪ٛ‬ن ٘مٖ اٌؼٕاصه شالشٍح األتؼاق‬
‫‪ٚ‬وٍفٍح ذصٍّّ‪ٙ‬ا تاسٍ‪ٛ‬ب ‪ٚ‬اضػ ‪ٚ‬شفاف‪.‬‬

‫فً ٘ما اٌثؽس‪ ،‬ذُ ػًّ ّٔمظٗ ٌؼكق سد ٘اِاخ ـ‪ٛ‬اوٌك ‪ِ -‬فرثهج ِؼٍٍّا‪ -‬تاسفكاَ‬
‫تهٔاِط اٌرؽًٍٍ غٍه اٌفطً ٌٍؼٕاصه شالشٍح األتؼاق (⑫‪ ِٕٗٚ (ANSYS V‬ذُ اسرفهاض إٌرائط‬
‫(أشىاي اٌشه‪ٚ‬ؾ ‪ٚ‬أشىاي اإلٔ‪ٍٙ‬ان ‪ٚ‬اٌرشىً ‪ٚ‬ذ‪ٛ‬وٌغ اظ‪ٙ‬اقاخ اٌشك ‪ٚ‬اٌضغظ ِٓ ِساناخ اإلظ‪ٙ‬اقاخ‬
‫‪ٚ‬غٍه٘ا) ‪ٚ‬اٌرى ال ٌّىٓ ذٕثئ٘ا ِٓ "أًَىرس انضبغظ وانشذاد" ‪ِٚ‬مانٔر‪ٙ‬ا تٕرائط اإلـرثاناخ‬
‫اٌّؼٍٍّح ٌٍرؽمك ِٓ صؽح إٌّمظح غٍه اٌفطٍح‪ٚ .‬تؼك لٌه ذُ الرهاغ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق شٕائً‬
‫‪ٚ‬شالشى األتؼاق ٌ‪ّٙ‬اخ اٌف‪ٛ‬اوٌك اٌّكن‪ٚ‬سح ‪ِٚ‬مانٔح ٘مٖ إٌرائط تإٌرائط اٌّؼٍٍّح‪.‬‬

‫ِٓ ـالي ٘ما اٌثؽس‪ ،‬أظ‪ٙ‬هخ ٔرائط اٌرؽًٍٍ ٌٍؼٕاصه شالشٍح األتؼاق تاسرفكاَ تهٔاِط‬
‫(⑫‪ (ANSYS V‬ذ‪ٛ‬افك ظٍك تؼك ِمانٔر‪ٙ‬ا تإٌرائط اٌّؼٍٍّح‪ ،‬ؼٍس ؼكز ذؤوٍك لٌه ِٓ ـالي‬
‫اٌفه‪ٚ‬لاخ اٌمٍٍٍح (‪ %1‬اًٌ ‪ )%10‬تٍٓ ٔرائط اٌرؽًٍٍ ‪ٚ‬إٌرائط اٌّؼٍٍّح‪ٚ .‬ومٌه ذُ اٌر‪ٛ‬صً اًٌ أْ‬
‫أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق رالري األبعبد أفضً ِٓ غٍهٖ (رُبئي األبعبد) ‪ٚ‬ـص‪ٛ‬صا ػٕك ذصٍُّ‬
‫ِٕاطك كبيهت اإلضطشاة ِصً ٘اِاخ اٌف‪ٛ‬اوٌك‪.‬‬
‫اإلداسة انعبيت نهًكخبت‬
‫يهخض انشعبنت ( ) ببنًكخبت‬
‫نلُ‪:‬‬ ‫انهُذعت انًذَيت‬ ‫لسُ‪:‬‬ ‫انهُذعت‬ ‫اٌىٍٍح‪:‬‬
‫اٌرانٌؿ‪:‬‬ ‫يبجغخيش انهُذعت‬ ‫اٌكنظح‬ ‫يحًذ إبشاهيى يخىني يحًذ‬ ‫االسُ‪:‬‬
‫(هُذعت إَشبئيت)‬ ‫اٌؼٍٍّح‪:‬‬
‫اسُ اٌهساٌح‪ :‬حظًيى بالطبث انخحىيم ببعخخذاو أًَىرس انضبغظ وانشذاد‬
‫اٌٍّفص‬
‫أصثؽد اٌؽاظٗ اًٌ ػًّ ِٕشآخ ذؽ‪ ًٌٛ‬أِه ٘اَ ‪ٚ‬ضه‪ٚ‬ني ‪ٚ‬ـاصح فً اٌّثأى ػاٌٍح اإلنذفاع‪ ،‬ؼٍس ِٓ ـالٌ‪ٙ‬ا ٌس‪ًٙ‬‬
‫اٌرغٍة ػًٍ اٌّشاوً اٌّرؼٍمٗ تاـرالف اسرفكاَ اٌّساؼاخ قاـً اٌّثٕى‪ٚ .‬لك ذى‪ِٕ ْٛ‬شآخ اٌرؽ‪ ًٌٛ‬ػًٍ شىً وّهٖ ذؽ‪ ًٌٛ‬أ‪ٚ‬‬
‫تالطح ذؽ‪ٚ .ًٌٛ‬ذؼرثه تالطح اٌرؽ‪ٚ ًٌٛ‬اؼكج ِٓ أش‪ٙ‬ه اٌؼٕاصه اٌّ‪ٛ‬ظ‪ٛ‬قج تىصهج فً اٌّثأى اٌّهذفؼح‪ٌٚ .‬هسُ ص‪ٛ‬نٖ واٍِح‬
‫‪ٚ‬اضؽٗ ػٓ ٘مٖ اٌؼٕاصه شالشٍح األتؼاق‪ ،‬ذُ اػرثان ٘اِاخ اٌف‪ٛ‬اوٌك وؤٔ‪ٙ‬ا تالطح ذؽ‪ ًٌٛ‬تسٍطح‪ٌ ،‬رس‪ ًٍٙ‬قناسر‪ٙ‬ا فً ٘ما‬
‫اٌثؽس‪.‬‬
‫ذؼرثه ٘اِاخ اٌف‪ٛ‬اوٌك ِٕطمٗ واٍِح اإلضطهاب‪ ،‬تسثة شىٍ‪ٙ‬ا اٌغٍه ِٕرظُ ‪ٚ‬اؼر‪ٛ‬ائ‪ٙ‬ا ػًٍ وٍّاخ وثٍهج ِٓ ؼعُ‬
‫اٌفهسأح ت‪ٙ‬ا لاخ اظ‪ٙ‬اقاخ ػّ‪ٛ‬قٌح ِٕففضح تاإلضافح اًٌ اظ‪ٙ‬اقاخ لص ػاٌٍح ٔسثٍا‪ِّ ،‬ا ٌعؼً سٍ‪ٛ‬و‪ٙ‬ا غٍه ِف‪ٌٍّ َٛٙ‬صُّ‪.‬‬
‫‪ٚ‬تاٌراًٌ ذُ اسرفكاَ طهٌمح ذصٍُّ ظكٌكج ذسًّ "أًَىرس انضبغظ وانشذاد" ٌٍرّىٓ اٌّصُّ ِٓ ف‪ ُٙ‬سٍ‪ٛ‬ن ٘مٖ اٌؼٕاصه شالشٍح‬
‫األتؼاق ‪ٚ‬وٍفٍح ذصٍّّ‪ٙ‬ا تاس‪ٛ‬ب ‪ٚ‬اضػ ‪ٚ‬شفاف‪.‬‬
‫فً ٘ما اٌثؽس‪ ،‬ذُ ػًّ ّٔمظٗ ٌؼكق سرح ٘اِاخ ـ‪ٛ‬اوٌك ‪ِ -‬فرثهٖ ِؼٍٍّا‪ -‬تاسفكاَ تهٔاِط اٌرؽًٍٍ اٌالـطً ٌٍؼٕاصه‬
‫شالشٍح األتؼاق (⑫‪ ِٕٗٚ )ANSYS V‬ذُ اسرفهاض إٌرائط (أشىاي اٌشه‪ٚ‬ؾ ‪ٚ‬أشىاي اإلٔ‪ٍٙ‬ان ‪ٚ‬اٌرشىً ‪ٚ‬ذ‪ٛ‬وٌغ اظ‪ٙ‬اقاخ اٌشك‬
‫‪ٚ‬اٌضغظ ِٓ ِساناخ اإلظ‪ٙ‬اقاخ ‪ٚ‬غٍه٘ا) ‪ٚ‬اٌرى ال ٌّىٓ ذٕثئ٘ا ِٓ "أًَىرس انضبغظ وانشذاد" ‪ِٚ‬مانٔر‪ٙ‬ا تٕرائط اإلـرثاناخ‬
‫اٌّؼٍٍّح ٌٍرؽمك ِٓ صؽح إٌّمظح االـطٍح‪ٚ .‬تؼك لٌه ذُ الرهاغ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق شٕائً ‪ٚ‬شالشى األتؼاق ٌ‪ّٙ‬اخ‬
‫اٌف‪ٛ‬اوٌك اٌّكن‪ٚ‬سٗ ‪ِٚ‬مانٔح ٘مٖ إٌرائط تإٌرائط اٌّؼٍٍّح‪.‬‬
‫ِٓ ـالي ٘ما اٌثؽس‪ ،‬أظ‪ٙ‬هخ ٔرائط اٌرؽًٍٍ اٌالـطً ٌٍؼٕاصه شالشٍح األتؼاق تاسرفكاَ تهٔاِط (⑫‪ )ANSYS V‬ذ‪ٛ‬افك ظٍك‬
‫تؼك ِمانٔر‪ٙ‬ا تإٌرائط اٌّؼٍٍّح‪ ،‬ؼٍس ؼكز ذؤوٍك لٌه ِٓ ـالي اٌفه‪ٚ‬لاخ اٌمٍٍٍح (‪ %1‬اًٌ ‪ )%10‬تٍٓ ٔرائط اٌرؽًٍٍ ‪ٚ‬إٌرائط‬
‫اٌّؼٍٍّح‪ٚ .‬ومٌه ذُ اٌر‪ٛ‬صً اًٌ أْ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق رالري األبعبد أفضً ِٓ غٍهٖ (رُبئي األبعبد) ‪ٚ‬ـص‪ٛ‬صا ػٕك‬
‫ذصٍُّ ِٕاطك واٍِح اإلضطهاب ِصً ٘اِاخ اٌف‪ٛ‬اوٌك‪.‬‬

‫وحقع انشعبنت في عبعت أبىاة عهي انُحى انخبني‪:‬‬


‫انببة األول‪ٌ :‬ؽر‪ٛ‬ي ٘ما اٌثاب ػًٍ ِمكِح ػاِح ػٓ اٌّشىٍح ‪ٚ‬طهق اٌرصٍُّ اٌّفرٍفٗ ٌؽٍ‪ٙ‬ا ‪ٚ‬ذ‪ٛ‬ضٍػ أل٘كاف اٌثؽس ‪ٚ‬وٍفٍح‬
‫أعاو ٘مٖ األ٘كاف ِٓ ـالي ِؽر‪ٌٛ‬اخ اٌهساٌح‪.‬‬
‫انببة انزبَي‪ٌ :‬رٕا‪ٚ‬ي طهٌمح اٌرصٍُّ "أًَىرس انضبغظ وانشذاد"‪ ،‬وٍف ٌٕشؤ ‪ٚ‬اٌؼٕاصه اٌّى‪ٔٛ‬ح ٌألّٔ‪ٛ‬لض ‪ِٚ‬ما‪ِٚ‬ح وً ػٕصه‬
‫ِٓ ٘مٖ اٌؼٕاصه‪.‬‬
‫انببة انزبنذ‪ٌ :‬رٕا‪ٚ‬ي ٘ما اٌثاب ِمكِح ػٓ ػٕاصه اٌرؽ‪ ًٌٛ‬اٌّفرٍفح اٌرى لك ذر‪ٛ‬اظك فً إٌّشآخ ‪ٚ‬األسثاب اٌرى أقخ اًٌ‬
‫ذ‪ٛ‬اظك٘ا ‪ٔٚ‬ثمج ِفرصهٖ ػٓ وً ٔ‪ٛ‬ع ‪ٚ‬ـص‪ٛ‬صا تالطح اٌرؽ‪ ًٌٛ‬اٌرً ً٘ ِؽً اٌكناسح‪ٌٚ ،‬ر‪ٛ‬ضٍػ اٌهإٌح أوصه‪ -‬ذُ اػرثان‬
‫٘اِاخ اٌف‪ٛ‬اوٌك وثالطح ذؽ‪ ًٌٛ‬تسٍطح فً ٘ما اٌثؽس‪.‬‬
‫انببة انشابع‪ٌ :‬رٕا‪ٚ‬ي ٘ما اٌثاب اٌرؽكٌاخ اٌرى ذ‪ٛ‬اظٗ طهٌمح اٌرصٍُّ اٌرمٍٍكٌح تاسرفكاَ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق ‪ٚ‬اٌكناساخ‬
‫اٌساتمح ٌ‪ٙ‬مٖ اٌطهٌمح ٌرالئُ اٌؼٕاصه شالشٍح األتؼاق ٘اِاخ اٌف‪ٛ‬اوٌك (وثالطح ذؽ‪ ًٌٛ‬تسٍطح)‪ ،‬ؼٍس ذفرٍف ٘مج اٌؼٕاصه ػٓ‬
‫غٍه٘ا ِٓ شٕائٍح األتؼاق فى ‪ٚ‬ظ‪ٛ‬ق اؼاطح ِٓ اٌفهسأٗ تؼض‪ٙ‬ا ٌثؼض ذؼًّ ػًٍ وٌاقج ِما‪ِٚ‬ر‪ٙ‬ا‪.‬‬
‫انببة انخبيظ‪ :‬فً ٘ما اٌثاب ذُ ػًّ ّٔمظح ٌؼٍٕاخ ٘اِاخ ـ‪ٛ‬اوٌك ِفرثهج ِؼٍٍّا تاسرفكاَ تهٔاِط اٌرؽًٍٍ اٌالـطً‬
‫⑫‪ ،ANSYS V‬شُ اٌرؤوك ِٓ صؽح ٘مٖ إٌّمظح تّمانٔح إٌرائط إٌظهٌح ِٓ اٌثهٔاِط تإٌرائط اٌّؼٍٍّح ٌٍؼٍٕاخ اٌّفرثهج‪.‬‬
‫انببة انغبدط‪ٌ :‬رٕا‪ٚ‬ي ٘ما اٌثاب ِمانٔٗ تٍٓ أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق شالشى األتؼاق ‪ٚ‬إٌّمظح ِٓ اٌرؽًٍٍ اٌالـطً تاسرفكاَ‬
‫تهٔاِط ‪.ANSYS V12.0‬‬
‫انببة انغببع‪ :‬فً ٘ما اٌثاب ذُ ػًّ ٍِفص ٌٍثؽس ‪ٚ‬إٌرائط اٌرى ذُ اٌر‪ٛ‬صً اٌٍ‪ٙ‬ا‪.‬‬
‫كهًبث دالنيت‪:‬‬
‫ِٕشآخ ذؽ‪٘ -ًٌٛ‬اِاخ اٌف‪ٛ‬اوٌك‪ -‬اإلضطهاب‪ -‬أّٔ‪ٛ‬لض اٌضاغظ ‪ٚ‬اٌشكاق‪ -‬االؼاطح‪ -‬اٌرؽًٍٍ اٌالـطً‬
‫ص‪.‬ة‪35516:‬‬ ‫يكخبت اإلداسة انعبيت‪ -‬انًُظىسة‪ 60 -‬شبسع انجًهىسيت‪ -‬انًُظىسة ‪ -‬يظش‬

‫‪URL: http:// www.mans.edu.eg‬‬ ‫‪EMAIL: mucentrlib@mans.edu.eg‬‬


‫جبيعت انًُظىسة‬
‫كهيــــــت انهُذعــــــــت‬
‫قغـــى انهُذعــت اإلَشـــبئيت‬
‫إقشاس‬
‫ٌمه اٌثاؼس‪ /‬يحًذ إبشاهيى يخىني يحًذ تاإلٌرىاَ تم‪ٛ‬أٍٓ ظاِؼح إٌّص‪ٛ‬نج ‪ٚ‬أٔظّر‪ٙ‬ا ‪ٚ‬ذؼٍٍّاذ‪ٙ‬ا‬
‫‪ٚ‬لهاناذ‪ٙ‬ا اٌسانٌح اٌّفؼ‪ٛ‬ي اٌّرؼٍمح تبػكاق نسائً اٌّاظسرٍه ػٕكِا لاَ تبػكاق اٌهساٌح اٌؼٍٍّح ذؽد‬
‫ػٕ‪ٛ‬اْ‪:‬‬

‫"حظًيى بالطبث انخحىيم ببعخخذاو أًَىرس انضبغظ وانشذاد"‬

‫‪ٌٚ‬عٕح اإلشهاف‪-:‬‬
‫‪ -1‬أ‪.‬د‪ .‬طالح انذيٍ انغعيذ انًخىني‬
‫‪ -2‬أ‪.‬و‪.‬د‪ .‬يحًذ انغعيذ انضغيبي‬

‫وؤؼك ِرطٍثاخ ًٍٔ قنظح اٌّاظسرٍه فً اٌ‪ٕٙ‬كسح اإلٔشائٍح ‪.‬‬

‫واإلقشاس تؽكاشح ِ‪ٛ‬ض‪ٛ‬ع اٌكناسح اٌثؽصٍح ‪ٚ‬أٔٗ ٌُ ٌسثك ذٕا‪ٚ‬ي اٌّ‪ٛ‬ض‪ٛ‬ع ‪ٚ‬اٌؼٕ‪ٛ‬اْ اٌثؽصً تص‪ٛ‬نذٗ‬
‫إٌ‪ٙ‬ائٍح اٌىاٍِح أ‪ٔ ٚ‬شهٖ ساتما فً أي نسائً أ‪ ٚ‬أطانٌػ أ‪ ٚ‬ورة أ‪ ٚ‬أتؽاز أ‪ ٚ‬أي ِٕش‪ٛ‬ناخ ػٍٍّح تّا‬
‫ٌٕسعُ ِغ األِأح اٌؼٍٍّح اٌّرؼانف ػٍٍ‪ٙ‬ا فً وراتح اٌهسائً ‪ٚ‬األطانٌػ ‪ٚ‬لث‪ٛ‬ي ػكق (‪ )1‬تؽس ٌٍٕشه‬
‫تؼٕ‪ٛ‬اْ‪:‬‬

‫"‪"Strut-and-Tie Model for Strength Assessment of Pile Caps‬‬

‫"حقذيش يقبويت هبيبث انخىاصيق ببعخخذاو طشيقت أًَىرس انضبغظ وانشذاد"‬

‫فً ِعٍح ػٍٍّح ِرفصصح فى ِعاي اٌ‪ٕٙ‬كسح ‪-:ً٘ٚ‬‬


‫يجهت انبحىد انهُذعيت بكهيت انهُذعت ببنًطشيت‪ -‬جبيعت حهىاٌ‬
‫بًجهذ سقى (‪ )150‬نشهش يىَيى نعبو ‪2016‬‬
‫‪ٚ‬أْ اٌثؽس إٌّش‪ٛ‬ن ِسرفهض ِٓ اٌهساٌح اٌّمو‪ٛ‬نج تؼاٌٍٗ ‪ٚ‬أْ أسّاء ظٍّغ اٌساقج اٌّشهفٍٓ ِ‪ٛ‬ظ‪ٛ‬قج‬
‫ػًٍ اٌثؽس‪.‬‬
‫وهزا إقشاس يُى بزنك‬
‫اٌّمه تّا فٍٗ‬
‫و‪ .‬يحًذ إبشاهيى يخىني‬
‫جبيعت انًُظىسة‬
‫كهيــــــت انهُذعــــــــت‬
‫قغـــى انهُذعــت اإلَشـــبئيت‬

‫نجُت انًُبقشت وانحكى‬


‫"حظًيى بالطبث انخحىيم بإعخخذاو أًَىرس انضبغظ وانشذاد"‬ ‫عُىاٌ انشعبنت‪:‬‬

‫يحًذ إبشاهيى يخىني يحًذ إبشاهيى‬ ‫اعى انبـــبحذ‪:‬‬

‫انذسجت انعهًيت انًطهىة انحظىل عهيهب‪ :‬دسجت انًبجغخيش في انهُذعت اإلَشبئيت‬

‫ححج إششاف‪:‬‬

‫انخىقيع‬ ‫انىظيفت‬ ‫اإلعى‬ ‫و‬


‫أسرال تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر دكخىس‬
‫‪1‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫طالح انذيٍ انغعيذ انًخىنى‬
‫أسرال ِساػك تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر يغبعذ دكخىس‬
‫‪2‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫يحًذ انغعيذ انضغيبى‬

‫نجُت انًُبقشت وانحكى‪:‬‬


‫انخىقيع‬ ‫انىظيفت‬ ‫اإلعى‬ ‫و‬
‫أسرال تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر دكخىس‬
‫‪1‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫يىعف إبشاهيى عجبس‬
‫أسرال تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر دكخىس‬
‫‪2‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫طالح انذيٍ انغعيذ انًخىنى‬
‫أسرال تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر دكخىس‬
‫‪3‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح تاٌّطهٌح‪ -‬ظاِؼح ؼٍ‪ٛ‬اْ‬ ‫أحًذ يىعي عبذ انشحًٍ‬
‫أسرال ِساػك تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر يغبعذ دكخىس‬
‫‪4‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫يحًذ انغعيذ انضغيبى‬

‫عًيذ انكهيت‬ ‫وكيم انكهيت نهذساعبث انعهيب وانبحىد‬ ‫سئيظ انقغى‬

‫أ‪.‬ق ِؽّك اتهاٍُ٘ اٌسؼٍك‬ ‫أ‪.‬ق لاسُ صالغ األٌفى‬ ‫أ‪.‬ق أؼّك ِؽّ‪ٛ‬ق ٌ‪ٛ‬سف‬
‫جبيعت انًُظىسة‬
‫كهيــــــت انهُذعــــــــت‬
‫قغـــى انهُذعــت اإلَشـــبئيت‬

‫انًششفــــىٌ‬
‫عُىاٌ انشعبنت‪:‬‬

‫"حظًيى بالطبث انخحىيم ببعخخذاو أًَىرس انضبغظ وانشذاد"‬


‫اعى انبـــبحذ‪:‬‬

‫يحًذ إبشاهيى يخىني يحًذ إبشاهيى‬


‫انذسجت انعهًيت انًطهىة انحظىل عهيهب‪:‬‬

‫دسجت انًبجغخيش في انهُذعت اإلَشبئيت‬


‫ححج إششاف‪:‬‬

‫انخىقيع‬ ‫انىظيفت‬ ‫اإلعى‬ ‫و‬


‫أسرال تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر دكخىس‬
‫‪1‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫طالح انذيٍ انغعيذ انًخىنى‬

‫أسرال ِساػك تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أعخبر يغبعذ دكخىس‬


‫‪2‬‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫يحًذ انغعيذ انضغيبى‬
‫جبيعت انًُظىسة‬
‫كهيــــــت انهُذعــــــــت‬
‫قغـــى انهُذعــت اإلَشـــبئيت‬

‫حظًيى بالطبث انخحىيم ببعخخذاو أًَىرس انضبغظ وانشذاد‬

‫نساٌح ػٍٍّح ِمكِح ِٓ‬

‫انًهُذط‪ /‬يحًذ إبشاهيى يخىني يحًذ إبشاهيى‬


‫تىاٌ‪ٛ‬نٌ‪ٛ‬ي اٌ‪ٕٙ‬كسح اٌّكٍٔح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬
‫ِؼٍك تمسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‪ِ -‬ؼ‪ٙ‬ك اٌكٌرا اٌؼاًٌ ٌٍ‪ٕٙ‬كسح ‪ٚ‬اٌرىٕ‪ٌٛٛ‬ظٍا‬

‫حىطئت نهحظىل عهي دسجت انًبجغخيش‬


‫فً‬

‫انهُذعت اإلَشبئيت‬
‫"حظًيى انًُشآث انخشعبَيت انًغهحت"‬

‫ذؽد اشهاف‬

‫انذكخــىس‬ ‫األعخبر انذكخىس‬

‫يحًذ انغعيـذ انــضغــيبي‬ ‫طالح انذيٍ انغعيذ انًخىني‬


‫أسرال ِساػك‪ -‬لسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬ ‫أسرال اٌفهسأح اٌّسٍؽح‪ -‬لسُ اٌ‪ٕٙ‬كسح اإلٔشائٍح‬
‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬ ‫وٍٍح اٌ‪ٕٙ‬كسح‪ -‬ظاِؼح إٌّص‪ٛ‬نج‬

‫‪2016‬‬

‫‪View publication stats‬‬

You might also like